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 519 毫秒
1.
2.
Cell surface receptors of the integrin family are pivotal to cell adhesion and migration. The activation state of heterodimeric αβ integrins is correlated to the association state of the single-pass α and β transmembrane domains. The association of integrin αIIbβ3 transmembrane domains, resulting in an inactive receptor, is characterized by the asymmetric arrangement of a straight (αIIb) and tilted (β3) helix relative to the membrane in congruence to the dissociated structures. This allows for a continuous association interface centered on helix-helix glycine-packing and an unusual αIIb(GFF) structural motif that packs the conserved Phe-Phe residues against the β3 transmembrane helix, enabling αIIb(D723)β3(R995) electrostatic interactions. The transmembrane complex is further stabilized by the inactive ectodomain, thereby coupling its association state to the ectodomain conformation. In combination with recently determined structures of an inactive integrin ectodomain and an activating talin/β complex that overlap with the αβ transmembrane complex, a comprehensive picture of integrin bi-directional transmembrane signaling has emerged.Key words: cell adhesion, membrane protein, integrin, platelet, transmembrane complex, transmembrane signalingThe communication of biological signals across the plasma membrane is fundamental to cellular function. The ubiquitous family of integrin adhesion receptors exhibits the unusual ability to convey signals bi-directionally (outside-in and inside-out signaling), thereby controlling cell adhesion, migration and differentiation.15 Integrins are Type I heterodimeric receptors that consist of large extracellular domains (>700 residues), single-pass transmembrane (TM) domains, and mostly short cytosolic tails (<70 residues). The activation state of heterodimeric integrins is correlated to the association state of the TM domains of their α and β subunits.610 TM dissociation initiated from the outside results in the transmittal of a signal into the cell, whereas dissociation originating on the inside results in activation of the integrin to bind ligands such as extracellular matrix proteins. The elucidation of the role of the TM domains in integrin-mediated adhesion and signaling has been the subject of extensive research efforts, perhaps commencing with the demonstration that the highly conserved GFFKR sequence motif of α subunits (Fig. 1), which closely follows the first charged residue on the intracellular face, αIIb(K989), constrains the receptor to a default low affinity state.11 Despite these efforts, an understanding of this sequence motif had not been reached until such time as the structure of the αIIb TM segment was determined.12 In combination with the structure of the β3 TM segment13 and available mutagenesis data,6,9,10,14,15 this has allowed the first correct prediction of the overall association of an integrin αβ TM complex.12 The predicted association was subsequently confirmed by the αIIbβ3 complex structure determined in phospholipid bicelles,16 as well as by the report of a similar structure based on molecular modeling using disulfide-based structural constraints.17 In addition to the structures of the dissociated and associated αβ TM domains, their membrane embedding was defined12,13,16,18,19 and it was experimentally recognized that, in the context of the native receptor, the TM complex is stabilized by the inactive, resting ectodomain.16 These advances in integrin membrane structural biology are complemented by the recent structures of a resting integrin ectodomain and an activating talin/β cytosolic tail complex that overlap with the αβ TM complex,20,21 allowing detailed insight into integrin bi-directional TM signaling.Open in a separate windowFigure 1Amino acid sequence of integrin αIIb and β3 transmembrane segments and flanking regions. Membrane-embedded residues12,13,16,18,19 are enclosed by a gray box. Residues 991–995 constitute the highly conserved GFFKR sequence motif of integrin α subunits.  相似文献   

3.
Previous studies on the activity of the rice Gα promoter using a β-Glucuronidase (GUS) reporter construct indicated that Gα expression was highest in developing organs and changed in a developmental stage-dependent manner. In this paper, GUS activity derived from the rice Gα promoter was analyzed in seeds and developing leaves. In seeds, GUS activity was detected in the aleurone layer, embryo, endosperm and scutellar epithelium. In developing leaves, the activity was detected in the mesophyll tissues, phloem and xylem of the leaf sheath and in the mesophyll tissue of the leaf blade. The activity in the aleurone layer and scutellar epithelium suggests that the Gα subunit may be involved in gibberellin signaling. The activity in the mesophyll tissues of the leaf blade suggests that the Gα subunit may be related to the intensity of disease resistance. The pattern of the activity in the developing leaf also indicates that the expression of Gα follows a developmental profile at the tissue level.Key words: expression pattern, Gα subunit, GUS staining pattern, heterotrimeric G protein, riceThe rice mutant d1 is deficient in the heterotrimeric G protein α subunit (Gα). Recently it was found that the dwarfism phenotype of d1 is due to a reduction in cell numbers.1 This discovery has led to new questions regarding how rice Gα regulates cell number, and which other signaling molecules are involved in this process in various tissues and at different development stages. Studies of d1 suggest that rice Gα participates in both gibberellin signaling24 and brassinosteroid signaling.58 Promoter studies using the β-Glucuronidase (GUS) reporter indicate that Gα expression is highest in developing organs.1 In this paper, we report on the expression pattern of a Gα promoter::GUS construct in seeds and developing leaves of rice.  相似文献   

4.
In this study, we report the functional characterization of heterotrimeric G-proteins from a nonvascular plant, the moss Physcomitrella patens. In plants, G-proteins have been characterized from only a few angiosperms to date, where their involvement has been shown during regulation of multiple signaling and developmental pathways affecting overall plant fitness. In addition to its unparalleled evolutionary position in the plant lineages, the P. patens genome also codes for a unique assortment of G-protein components, which includes two copies of and genes, but no canonical . Instead, a single gene encoding an extra-large Gα (XLG) protein exists in the P. patens genome. Here, we demonstrate that in P. patens the canonical Gα is biochemically and functionally replaced by an XLG protein, which works in the same genetic pathway as one of the Gβ proteins to control its development. Furthermore, the specific G-protein subunits in P. patens are essential for its life cycle completion. Deletion of the genomic locus of PpXLG or PpGβ2 results in smaller, slower growing gametophores. Normal reproductive structures develop on these gametophores, but they are unable to form any sporophyte, the only diploid stage in the moss life cycle. Finally, the mutant phenotypes of ΔPpXLG and ΔPpGβ2 can be complemented by the homologous genes from Arabidopsis, AtXLG2 and AtAGB1, respectively, suggesting an overall conservation of their function throughout the plant evolution.In all known eukaryotes, cellular signaling involves heterotrimeric GTP-binding proteins (G-proteins), which consist of Gα, Gβ, and Gγ subunits (Cabrera-Vera et al., 2003). According to the established paradigm, when Gα is GDP-bound, it forms a trimeric complex with the Gβγ dimer and remains associated with a G-protein coupled receptor. Signal perception by the receptor facilitates GDP to GTP exchange on Gα. GTP-Gα dissociates from the Gβγ dimer, and both these entities can transduce the signal by interacting with different effectors. The duration of the active state is determined by the intrinsic GTPase activity of Gα, which hydrolyzes bound GTP into GDP and inorganic phosphate (Pi), followed by the reassociation of the inactive, trimeric complex (Siderovski and Willard, 2005).In plants, G-protein signaling has been studied in only a few angiosperms to date at the functional level, although the proteins exist in the entire plant lineage (Hackenberg and Pandey, 2014; Urano and Jones, 2014; Hackenberg et al., 2016). Interestingly, while the overall biochemistry of the individual G-protein components and the interactions between them are conserved between plant and metazoan systems, deviations from the established norm are also obvious. For example, the repertoire of canonical G-proteins is significantly limited in plants; the human genome codes for 23 Gα, 5 Gβ, and 12 Gγ proteins, whereas most plant genomes, including those of basal plants, typically encode 1 canonical Gα, 1 Gβ, and three to five Gγ proteins (Urano and Jones, 2014). The only exceptions are some polyploid species, such as soybean, which have retained most of the duplicated G-protein genes (Bisht et al., 2011; Choudhury et al., 2011). Moreover, even in plants that possess only a single canonical Gα and Gβ protein, for example Arabidopsis (Arabidopsis thaliana) and rice, the phenotypes of plants lacking either one or both proteins are relatively subtle. The mutant plants exhibit multiple developmental and signaling defects but are able to complete the life cycle without any major consequences. These observations have questioned the significance of G-protein mediated signaling pathways in plants.Interestingly, plants also possess certain unique variants of the classical G-protein components such as the type III Cys-rich Gγ proteins and extra-large GTP-binding (XLG) proteins, which add to the diversity and expanse of the G-protein signaling networks (Roy Choudhury et al., 2011; Chakravorty et al., 2015; Maruta et al., 2015). The XLG proteins are almost twice the size of typical Gα proteins, with the C-terminal region that codes for Gα-like domain and an extended N-terminal region without any distinctive features. Plant XLGs are encoded by entirely independent genes and therefore are different from the mammalian extra-long versions of Gα proteins such as XLαs and XXLαs, which are expressed due to the use of alternate exons (Abramowitz et al., 2004). Three to five copies of XLG proteins are present in the genome of most angiosperms. At the functional level, the XLG proteins have been characterized only from Arabidopsis, to date, where recent studies suggest that the proteins compete with canonical Gα for binding with the Gβγ dimers and may form functional trimeric complexes (Chakravorty et al., 2015; Maruta et al., 2015). The XLG and Gβγ mutants of Arabidopsis seem to function in the same pathways during the regulation of a subset of plant responses, for example primary root length and its regulation by abscisic acid (ABA); the root waving and skewing responses; sensitivity to Glc, salt, and tunicamycin; and sensitivity to certain bacterial and fungal pathogens (Ding et al., 2008; Pandey et al., 2008; Chakravorty et al., 2015; Maruta et al., 2015). However, many of the phenotypes of Arabidopsis Gα and Gβγ mutants are also distinct from that of the xlg triple mutants. For example, compared to the wild-type plants, the canonical G-protein mutants exhibit altered response to gibberellic acid, brassinosteroids, and auxin and show changes in leaf shape, branching, flowering time, and stomatal densities (Ullah et al., 2003; Chen et al., 2004; Pandey et al., 2006; Zhang et al., 2008; Nilson and Assmann, 2010). The xlg triple mutants behave similarly to wild-type plants in all these aspects of development and signaling. Moreover, whether the XLG proteins are authentic GTP-binding and -hydrolyzing proteins and the extent to which they directly participate in G-protein-mediated signaling pathways remains confounding (Chakravorty et al., 2015; Maruta et al., 2015). Even in plants with a limited number of G-protein subunits such as Arabidopsis, one Gα and three XLGs potentially compete for a single Gβ protein, and the analysis of null mutants is not straightforward, that is, it is not possible to delineate whether the phenotypes seen in the Gα null mutants are truly due to the lack of Gα and/or because of an altered stoichiometry or availability of Gβ for the XLG proteins.As a bryophyte, Physcomitrella patens occupies a unique position in the evolutionary history of plants. It lacks vasculature but exhibits alteration between generations, which is dominated by a gametophytic (haploid) phase and a short sporophytic (diploid) phase (Cove et al., 2009). Many of the pathways related to hormone signaling, stress responses, and development are conserved between angiosperms and P. patens (Cove et al., 2009; Sun, 2011; Komatsu et al., 2013; Yasumura et al., 2015). It is also an intriguing example in the context of the G-protein signaling, because its fully sequenced genome does not encode a canonical Gα gene, although genes coding for the Gβ and Gγ proteins exist. A single gene for a potential XLG homolog also exists in the P. patens genome. This unique assortment of proteins predicts several alternative scenarios for G-protein signaling in P. patens. For example, the P. patens Gβγ proteins might be nonfunctional due to the loss of canonical Gα and are left in the genome as evolutionary artifacts. Alternatively, the Gβγ proteins of P. patens might maintain functionality regardless of the existence of a canonical Gα protein in pathways not regulated via classic G-protein signaling modes. Finally, a more likely scenario could be that the potential XLG protein can substitute for the Gα function in P. patens.To explore these possibilities and understand better the conserved and unique mechanisms of G-protein signaling pathways in plants and their significance, we examined the role of G-protein subunits in P. patens. We provide unambiguous evidence for the genetic coupling of XLG and Gβ proteins in controlling P. patens development. In contrast to all other plant species analyzed to date, where G-proteins are not essential for growth and survival, the XLG or one of the Gβ proteins is required for the sporophyte formation and life cycle completion in P. patens. Furthermore, one of the Arabidopsis XLG proteins, XLG2, and the canonical Gβ protein AGB1 can functionally complement the P. patens mutant phenotypes. These data provide new insights in the evolutionary breadth and the spectrum of signaling pathways regulated by G-proteins in plants.  相似文献   

5.
The heterotrimeric G protein α subunit (Gα) is targeted to the cytoplasmic face of the plasma membrane through reversible lipid palmitoylation and relays signals from G-protein-coupled receptors (GPCRs) to its effectors. By screening 23 DHHC motif (Asp-His-His-Cys) palmitoyl acyl-transferases, we identified DHHC3 and DHHC7 as Gα palmitoylating enzymes. DHHC3 and DHHC7 robustly palmitoylated Gαq, Gαs, and Gαi2 in HEK293T cells. Knockdown of DHHC3 and DHHC7 decreased Gαq/11 palmitoylation and relocalized it from the plasma membrane into the cytoplasm. Photoconversion analysis revealed that Gαq rapidly shuttles between the plasma membrane and the Golgi apparatus, where DHHC3 specifically localizes. Fluorescence recovery after photobleaching studies showed that DHHC3 and DHHC7 are necessary for this continuous Gαq shuttling. Furthermore, DHHC3 and DHHC7 knockdown blocked the α1A-adrenergic receptor/Gαq/11-mediated signaling pathway. Together, our findings revealed that DHHC3 and DHHC7 regulate GPCR-mediated signal transduction by controlling Gα localization to the plasma membrane.G-protein-coupled receptors (GPCRs) form the largest family of cell surface receptors, consisting of more than 700 members in humans. GPCRs respond to a variety of extracellular signals, including hormones and neurotransmitters, and are involved in various physiologic processes, such as smooth muscle contraction and synaptic transmission (20, 25). Heterotrimeric G proteins, composed of α, β, and γ subunits, transduce signals from GPCRs to their effectors and play a central role in the GPCR signaling pathway (13, 21, 24, 32). Although the Gα subunit seems to localize stably at the cytosolic face of the plasma membrane (PM), a recent report suggested that Gαo, a Gα isoform, shuttles rapidly between the PM and intracellular membranes (2). The PM targeting of Gα requires both interaction with the Gβγ complex and subsequent lipid palmitoylation of Gα (22). Thus, palmitoylation of Gα is a critical determinant of membrane targeting of the heterotrimer Gαβγ.Protein palmitoylation is a common posttranslational modification with lipid palmitate and regulates protein trafficking and function (7, 18). Gα is a classic and representative palmitoyl substrate (19, 38), and recent studies revealed that protein palmitoylation modifies virtually almost all the components of G-protein signaling, including GPCRs, Gα subunits, several members of the RGS (regulators of G-protein signaling) family of GTPase-activating proteins, GPCR kinase GRK6, and some small GTPases (7, 33). This common lipid modification plays an important role in compartmentalizing G-protein signaling to the specific microdomain, such as membrane caveolae and lipid raft (26). The palmitoyl thioester bond is relatively labile, and palmitates on substrates turn over rapidly, allowing proteins to shuttle between the cytoplasm/intracellular organelles and the PM (2, 3, 27). For example, binding of isoproterenol to the β-adrenergic receptor markedly accelerates the depalmitoylation of the associated Gαs, shifting Gαs to the cytoplasm (37). This receptor activation-induced depalmitoylation was also observed in a major postsynaptic PSD-95 scaffold, which anchors the AMPA (alpha-amino-3-hydroxy-5-methyl-isoxazole-4-propionic acid)-type glutamate receptor at the excitatory postsynapse through stargazin (6). On glutamate receptor activation, accelerated depalmitoylation of PSD-95 dissociates PSD-95 from postsynaptic sites and causes AMPA receptor endocytosis (6). Thus, palmitate turnover on Gαs and PSD-95 is accelerated by receptor activation, contributing to downregulation of the signaling pathway. However, the enzymes that add palmitate to proteins (palmitoyl-acyl transferases [PATs]) and those that cleave the thioester bond (palmitoyl-protein thioesterases) were long elusive.Recent genetic studies in Saccharomyces cerevisiae identified Erf2/Erf4 (1, 40) and Akr1 (29) as PATs for yeast Ras and yeast casein kinase 2, respectively. Erf2 and Akr1 have four- to six-pass transmembrane domains and share a common domain, referred to as a DHHC domain, a cysteine-rich domain with a conserved Asp-His-His-Cys signature motif. Because the DHHC domain is essential for the PAT activity, we isolated 23 mammalian DHHC domain-containing proteins (DHHC proteins) and developed a systematic screening method to identify the specific enzyme-substrate pairs (11, 12): DHHC2, -3, -7, and -15 for PSD-95 (11); DHHC21 for endothelial NO synthase (10); and DHHC3 and -7 for GABAA receptor γ2 subunit (9). Several other groups also reported that DHHC9 with GCP16 mediates palmitoylation toward H- and N-Ras (36) and that DHHC17, also known as HIP14, palmitoylates several neuronal proteins: huntingtin (14), SNAP-25, and CSP (14, 23, 35). However, the existence of PATs for Gα has been controversial because spontaneous palmitoylation of Gα could occur in vitro (4).In this study, we screened the 23 DHHC clones to examine which DHHC proteins can palmitoylate Gα. We found that DHHC3 and -7 specifically and robustly palmitoylate Gα at the Golgi apparatus. Inhibition of DHHC3 and -7 reduces Gαq/11 palmitoylation levels and delocalizes it from the PM to the cytoplasm in HeLa cells and primary hippocampal neurons. Also, DHHC3 and -7 are necessary for the continuous Gαq shuttling between the Golgi apparatus and the PM. Finally, blocking DHHC3 and -7 inhibits the α1A-adrenergic receptor [α1A-AR]/Gαq-mediated signaling pathway, indicating that DHHC3 and -7 play an essential role in GPCR signaling by regulating Gα localization.  相似文献   

6.
7.
AGS3, a receptor-independent activator of G-protein signaling, is involved in unexpected functional diversity for G-protein signaling systems. AGS3 has seven tetratricopeptide (TPR) motifs upstream of four G-protein regulatory (GPR) motifs that serve as docking sites for Giα-GDP. The positioning of AGS3 within the cell and the intramolecular dynamics between different domains of the proteins are likely key determinants of their ability to influence G-protein signaling. We report that AGS3 enters into the aggresome pathway and that distribution of the protein is regulated by the AGS3 binding partners Giα and mammalian Inscuteable (mInsc). Giα rescues AGS3 from the aggresome, whereas mInsc augments the aggresome-like distribution of AGS3. The distribution of AGS3 to the aggresome is dependent upon the TPR domain, and it is accelerated by disruption of the TPR organizational structure or introduction of a nonsynonymous single-nucleotide polymorphism. These data present AGS3, G-proteins, and mInsc as candidate proteins involved in regulating cellular stress associated with protein-processing pathologies.The discovery of AGS3 (GPSM1) and related accessory proteins revealed unexpected functional diversity for G-protein signaling systems (8, 36). AGS3 is involved in a number of different cellular activities, including asymmetric cell division during neuronal development (30), neuronal plasticity and addiction (9, 10, 12, 38, 39), autophagy (27), membrane protein trafficking (17), cardiovascular function (7), and metabolism (7). AGS3 is a multidomain protein consisting of seven tetratricopeptide repeats (TPR) in the amino-terminal portion of the protein and four G-protein regulatory (GPR) motifs in the carboxyl region of the protein. Each of the GPR motifs binds and stabilizes the GDP-bound conformation of Gα (Giα, Gtα, and Gi/oα), essentially behaving as a guanine nucleotide dissociation inhibitor. As such, AGS3 may be complexed with up to four Gα and function as an alternative binding partner for Gα independently of the classical heterotrimeric Gαβγ. Despite the clearly demonstrated function of AGS3 and the related protein LGN (GPSM2 or AGS5) in various model organisms and a fairly solid, basic biochemical understanding of the interaction of a GPR motif with Gα, the signals that operate “upstream” and/or “downstream” of AGS3 or an AGS3-Gi/oα complex are not well defined.AGS3 and other GPR proteins may regulate G-protein signaling directly by influencing the interaction of Gα with Gβγ or another Gα binding partner. In addition, a portion of Gα in the cell is complexed with GPR proteins to various degrees, and this interaction is regulated. Ric-8A interacts with an AGS3-Giα complex in a manner somewhat analogous to the interaction of a G-protein-coupled receptor with heterotrimeric Gαβγ, promoting nucleotide exchange and the apparent dissociation of AGS3 and Giα-GDP (37). The specific impact of AGS3 and other GPR proteins on signaling events is likely dependent upon where the individual protein is positioned within the cell and the nature of intra- and intermolecular interactions that influence the interaction of the GPR motif with Gi/oα.The TPR domain of AGS3 is an important determinant of its positioning within the cell through its interaction with specific binding partners (1, 8, 28, 36). As part of a broader effort to address the fundamental questions of AGS3 “positioning” and control of G-protein interaction, we focused upon the roles of individual TPR domains. Endogenous and ectopically expressed wild-type AGS3 is nonhomogeneously distributed in the cytoplasm, with obvious punctate structures, and it may be present at the cell periphery. Disruption of the TPR organizational structure by targeted amino acid substitutions or introduction of a nonsynonymous single-nucleotide polymorphism redistributes AGS3 to punctate structures throughout the cytoplasm that are similar in appearance to the preaggresomal assemblies or aggregates observed in neurodegenerative diseases. Upon cellular stress, both wild-type and TPR-modified AGS3 migrate, in a microtubule-dependent manner, to a perinuclear aggresome. The distribution of AGS3 to the aggresome is dependent upon the TPR domain, and it is differentially regulated by Giα and mammalian Inscuteable (mInsc), which bind to the GPR and TPR domains, respectively, of AGS3. These data present AGS3 and G-proteins as candidate proteins involved in regulating cellular stress associated with protein-processing pathologies and suggest that this involvement can be manipulated to therapeutic advantage.  相似文献   

8.
9.
Protein kinase C (PKC)-ε, a component of the serine/threo-nine PKC family, has been shown to influence the survival and differentiation pathways of normal hematopoietic cells. Here, we have modulated the activity of PKC-ε with specific small molecule activator or inhibitor peptides. PKC-ε inhibitor and activator peptides showed modest effects on HL-60 maturation when added alone, but PKC-ε activator peptide significantly counteracted the pro-maturative activity of tumor necrosis factor (TNF)-α towards the monocytic/macrophagic lineage, as evaluated in terms of CD14 surface expression and morphological analyses. Moreover, while PKC-ε inhibitor peptide showed a reproducible increase of TNF-related apoptosis inducing ligand (TRAIL)-induced apoptosis, PKC-ε activator peptide potently counteracted the pro-apoptotic activity of TRAIL. Taken together, the anti-maturative and anti-apoptotic activities of PKC-ε envision a potentially important proleukemic role of this PKC family member.Key words: acute myeloid leukemia, surface antigens, HL-60 cells, apoptosis, maturation.Activation of all protein kinase C (PKC) family of serine and threonine isoenzymes is associated with binding to the negatively charged phospholipids, phosphatidylserine, while different PKC isozymes have varying sensitivities to Ca2+ and lipid-derived second messengers such as diacylglycerol (Gonelli et al., 2009). Upon activation, PKC isozymes translocate from the soluble to the particulate cell fraction, including cell membrane, nucleus and mitochondria (Gonelli et al., 2009). PKC primary sequence can be broadly separated into two domains: the N-terminal regulatory domain and the conserved C-terminal catalytic domain.The regulatory domain of PKC is composed of the C1 and C2 domains that mediate PKC interactions with second messengers, phospholipids, as well as inter and intramolecular protein-protein interactions. Differences in the order and number of copies of signaling domains, as well as sequence differences that affect binding affinities, result in the distinct activity of each PKC isozyme (Gonelli et al., 2009).In recent years, a series of peptides derived from PKC have been shown to modulate its activity by interfering with critical protein-protein interactions within PKC and between PKC and PKC-binding proteins (Brandman et al., 2007, Souroujon and Mochly-Rosen, 1998). Focusing on PKC-ε isozyme and using a rational approach, one C2-derived peptide that acts as an isozyme-selective activator (Dorn et al., 1999) and another that acts as a selective inhibitor (Johnson et al., 1996) of PKC-ε, have been identified.These findings are particularly interesting since besides being involved in the physiology of normal cardiac (Braun and Mochly-Rosen, 2003, Johnson et al., 1996, Li et al., 2006), hematopoietic (Gobbi et al., 2009, Mirandola et al., 2006, Racke et al., 2001), and neuronal (Borgatti et al., 1996) cell models, mounting experimental evidences have linked altered PKC-ε functions to solid tumor development (Okhrimenko et al., 2005, Gillespie et al., 2005, Lu et al., 2006). Therefore, taking advantage of the recent availability of small molecule peptides able to activate or inhibit specifically PKC-ε by disrupting protein/protein interactions (Dorn et al., 1999, Johnson et al., 1996), which open important therapeutic perspectives, we have investigated the effects of both PKC-ε activator and PKC-ε inhibitor peptides on the maturation and survival of leukemic cells, using as a model system the HL-60 myeloblastic leukemia cell line, which can be induced to undergo terminal differentiation or apoptotic cell death by a variety of chemical and biological agents (Breitman et al., 1980, Zauli et al., 1996).  相似文献   

10.
11.
Heterotrimeric G proteins, consisting of Gα, Gβ, and Gγ subunits, are a conserved signal transduction mechanism in eukaryotes. However, G protein subunit numbers in diploid plant genomes are greatly reduced as compared with animals and do not correlate with the diversity of functions and phenotypes in which heterotrimeric G proteins have been implicated. In addition to GPA1, the sole canonical Arabidopsis (Arabidopsis thaliana) Gα subunit, Arabidopsis has three related proteins: the extra-large GTP-binding proteins XLG1, XLG2, and XLG3. We demonstrate that the XLGs can bind Gβγ dimers (AGB1 plus a Gγ subunit: AGG1, AGG2, or AGG3) with differing specificity in yeast (Saccharomyces cerevisiae) three-hybrid assays. Our in silico structural analysis shows that XLG3 aligns closely to the crystal structure of GPA1, and XLG3 also competes with GPA1 for Gβγ binding in yeast. We observed interaction of the XLGs with all three Gβγ dimers at the plasma membrane in planta by bimolecular fluorescence complementation. Bioinformatic and localization studies identified and confirmed nuclear localization signals in XLG2 and XLG3 and a nuclear export signal in XLG3, which may facilitate intracellular shuttling. We found that tunicamycin, salt, and glucose hypersensitivity and increased stomatal density are agb1-specific phenotypes that are not observed in gpa1 mutants but are recapitulated in xlg mutants. Thus, XLG-Gβγ heterotrimers provide additional signaling modalities for tuning plant G protein responses and increase the repertoire of G protein heterotrimer combinations from three to 12. The potential for signal partitioning and competition between the XLGs and GPA1 is a new paradigm for plant-specific cell signaling.The classical heterotrimeric G protein consists of a GDP/GTP-binding Gα subunit with GTPase activity bound to an obligate dimer formed by Gβ and Gγ subunits. In the signaling paradigm largely elucidated from mammalian systems, the plasma membrane-associated heterotrimer contains Gα in its GDP-bound form. Upon receiving a molecular signal, typically transduced by a transmembrane protein (e.g. a G protein-coupled receptor), Gα exchanges GDP for GTP and dissociates from the Gβγ dimer. Both Gα and Gβγ interact with intracellular effectors to initiate downstream signaling cascades. The intrinsic GTPase activity of Gα restores Gα to the GDP-bound form, which binds Gβγ, thereby reconstituting the heterotrimer (McCudden et al., 2005; Oldham and Hamm, 2008).Signal transduction through a heterotrimeric G protein complex is an evolutionarily conserved eukaryotic mechanism common to metazoa and plants, although there are distinct differences in the functional intricacies between the evolutionary branches (Jones et al., 2011a, 2011b; Bradford et al., 2013). The numbers of each subunit encoded within genomes, and therefore the potential for combinatorial complexity within the heterotrimer, is one of the most striking differences between plants and animals. For example, the human genome encodes 23 Gα (encoded by 16 genes), five Gβ, and 12 Gγ subunits (Hurowitz et al., 2000; McCudden et al., 2005; Birnbaumer, 2007). The Arabidopsis (Arabidopsis thaliana) genome, however, only encodes one canonical Gα (GPA1; Ma et al., 1990), one Gβ (AGB1; Weiss et al., 1994), and three Gγ (AGG1, AGG2, and AGG3) subunits (Mason and Botella, 2000, 2001; Chakravorty et al., 2011), while the rice (Oryza sativa) genome encodes one Gα (Ishikawa et al., 1995), one Gβ (Ishikawa et al., 1996), and either four or five Gγ subunits (Kato et al., 2004; Chakravorty et al., 2011; Botella, 2012). As expected, genomes of polyploid plants have more copies due to genome duplication, with the soybean (Glycine max) genome encoding four Gα, four Gβ (Bisht et al., 2011), and 10 Gγ subunits (Choudhury et al., 2011). However, Arabidopsis heterotrimeric G proteins have been implicated in a surprisingly large number of phenotypes, which is seemingly contradictory given the relative scarcity of subunits. Arabidopsis G proteins have been implicated in cell division (Ullah et al., 2001; Chen et al., 2006) and morphological development in various tissues, including hypocotyls (Ullah et al., 2001, 2003), roots (Ullah et al., 2003; Chen et al., 2006; Li et al., 2012), leaves (Lease et al., 2001; Ullah et al., 2001), inflorescences (Ullah et al., 2003), and flowers and siliques (Lease et al., 2001), as well as in pathogen responses (Llorente et al., 2005; Trusov et al., 2006; Cheng et al., 2015), regulation of stomatal movement (Wang et al., 2001; Coursol et al., 2003; Fan et al., 2008) and development (Zhang et al., 2008; Nilson and Assmann, 2010), cell wall composition (Delgado-Cerezo et al., 2012), responses to various light stimuli (Warpeha et al., 2007; Botto et al., 2009), responses to multiple abiotic stimuli (Huang et al., 2006; Pandey et al., 2006; Trusov et al., 2007; Zhang et al., 2008; Colaneri et al., 2014), responses to various hormones during germination (Ullah et al., 2002), and postgermination development (Ullah et al., 2002; Pandey et al., 2006; Trusov et al., 2007). Since the Gγ subunit appeared to be the only subunit that provides diversity in heterotrimer composition in Arabidopsis, it was proposed that all functional specificity in heterotrimeric G protein signaling was provided by the Gγ subunit (Trusov et al., 2007; Chakravorty et al., 2011; Thung et al., 2012, 2013). This allowed for only three heterotrimer combinations to account for the wide range of G protein-associated phenotypes.In addition to the above typical G protein subunits, the plant kingdom contains a conserved protein family of extra-large GTP-binding proteins (XLGs). XLGs differ from typical Gα subunits in that they possess a long N-terminal extension of unknown function, but they are similar in that they all have a typical C-terminal Gα-like region, with five semiconserved G-box (G1–G5) motifs. The XLGs also possess the two sequence features that differentiate heterotrimeric G protein Gα subunits from monomeric G proteins: a helical region between the G1 and G2 motifs and an Asp/Glu-rich loop between the G3 and G4 motifs (Lee and Assmann, 1999; Ding et al., 2008; Heo et al., 2012). The Arabidopsis XLG family comprises XLG1, XLG2, and XLG3, and all three have demonstrated GTP-binding and GTPase activities, although they differ from GPA1 in exhibiting a much slower rate of GTP hydrolysis, with a Ca2+ cofactor requirement instead of an Mg2+ requirement, as for canonical Gα proteins (Heo et al., 2012). All three Arabidopsis XLGs were observed to be nuclear localized (Ding et al., 2008). Although much less is known about XLGs than canonical Gα subunits, XLG2 positively regulates resistance to the bacterial pathogen Pseudomonas syringae and was immunoprecipitated with AGB1 from tissue infected with P. syringae (Zhu et al., 2009). xlg3 mutants, like agb1 mutants, are impaired in root-waving and root-skewing responses (Pandey et al., 2008). During the preparation of this report, Maruta et al. (2015) further investigated XLG2, particularly focusing on the link between XLG2 and Gβγ in pathogen responses. Based on symptom progression in xlg mutants, they found that XLG2 is a positive regulator of resistance to both bacterial and fungal pathogens, with a minor contribution from XLG3 in resistance to Fusarium oxysporum. XLG2 and XLG3 are also positive regulators of reactive oxygen species (ROS) production in response to pathogen-associated molecular pattern elicitors. The resistance and pathogen-associated molecular pattern-induced ROS phenotypes of the agg1 agg2 and xlg2 xlg3 double mutants were not additive in an agg1 agg2 xlg2 xlg3 quadruple mutant, indicating that these two XLGs and the two Gγ subunits function in the same, rather than parallel, pathways. Unfortunately, the close proximity of XLG2 and AGB1 on chromosome 4 precluded the generation of an agb1 xlg2 double mutant; therefore, direct genetic evidence of XLG2 and AGB1 interaction is still lacking, but physical interactions between XLG2 and the Gβγ dimers were shown by yeast (Saccharomyces cerevisiae) three-hybrid and bimolecular fluorescence complementation (BiFC) assays (Maruta et al., 2015). Localization of all three XLGs was also reexamined, indicating that XLGs are capable of localizing to the plasma membrane in addition to the nucleus (Maruta et al., 2015).Interestingly, several other plant G protein-related phenotypes, in addition to pathogen resistance, have been observed only in Gβ and Gγ mutants, with opposite phenotypes observed in Gα (gpa1) mutants. Traditionally, the observation of opposite phenotypes in Gα versus Gβγ mutants in plants and other organisms has mechanistically been attributed to signaling mediated by free Gβγ, which increases in abundance in the absence of Gα. However, an intriguing alternative is that XLG proteins fulfill a Gα-like role in forming heterotrimeric complexes with Gβγ and function in non-GPA1-based G protein signaling processes. If XLGs function like Gα subunits, the corresponding increase in subunit diversity could potentially account for the diversity of G protein phenotypes. In light of this possibility, we assessed the heterotrimerization potential of all possible XLG and Gβγ dimer combinations, XLG localization and its regulation by Gβγ, and the effect of xlg mutation on selected known phenotypes associated with heterotrimeric G proteins. Our results provide compelling evidence for the formation of XLG-Gβγ heterotrimers and reveal that plant G protein signaling is substantially more complex than previously thought.  相似文献   

12.
Jenny M. Woof 《MABS-AUSTIN》2012,4(3):291-293
Fc receptors and their interaction with antibodies will be a major theme at the forthcoming FASEB Science Research Conference on Immunoreceptors to be held in Snowmass this July (details available at www.faseb.org/src/home.aspx, follow the tabs for Immunoreceptors). Since its inception in the mid 1980s, this meeting series has maintained a focus on Fc receptors, and this year’s meeting will be no exception.From a therapeutic viewpoint, there is much to be gained from a detailed understanding of the biology of effector molecules such as Fc receptors and complement. Indeed, knowledge of the interaction of IgG with such molecules has been central to the development of improved mAbs with altered functions and transformed half-lives, tailored for particular therapeutic applications. Examples include mAbs designed to maximise complement recruitment1 or to enhance Fc receptor engagement and triggering of ADCC,2-5 or conversely, variants engineered to be unable to engage complement6 or Fc receptors.7 Glycoengineering of IgG Fc offers an alternative means to modify effector function capabilities,8 while development of IgG mutants that display extended or altered serum half-lives has been driven through exhaustive analysis of the interaction with FcRn.9,10Despite the appreciable advances that have been made in unravelling the various facets of Fc receptor biology, new information pertinent to mAb engineering continues to emerge. A flavour of some of these new advances will be given below. They span novel receptors and receptor roles, structure-function relationships, the molecular architecture of signaling complexes, the influence of the membrane lipid environment and scaffolding interactions, isotype considerations, through to technical innovations likely to inform the field.Remarkably, new receptors that have previously eluded characterization are now being described. These include the IgM receptor, which evidence indicates is a molecule also known as TOPO/Fas apoptotic inhibitory molecule 3 whose gene lies close to other known immunoglobulin receptors on chromosome 1,11 and a receptor for IgD recently documented on basophils.12 Moreover, we are seeing an appreciation of new roles for existing Fc receptors. An example is the demonstration in a transgenic study that human FcγRIIa can trigger active and passive anaphylaxis and airway inflammation. Moreover, human mast cells, monocytes and neutrophils were shown to produce anaphylactogenic mediators when FcγRIIA was engaged.13 Hence IgG may contribute to allergic and anaphylactic reactions in humans by engaging FcγRIIa.Exciting new structural information on Fc receptors and their ligands is emerging. An important example is the solving of the X-ray crystal structure for human FcγRI.14 While the structural information supports a ligand binding mode similar to those of FcγRII or FcγRIII, the FG-loop in domain 2 of FcγRI with its conserved one-residue deletion appears critical for high affinity IgG binding. A second example concerns the high responder/low responder (HR/LR) polymorphisms of FcγRIIa, which are linked to susceptibility to infections, autoimmune diseases, and the efficacy of therapeutic Abs. New insights into these differences have been provided by the recent solving of the structure for the complex of the HR allele with IgG Fc.15 Third, understanding of the human IgE-FcεRI interaction has moved forward significantly through the solving of the X-ray crystal structure of the complex of FcεRI and the entire Fc region of IgE (comprising domains Cε2, Cε3 and Cε4).16 In a final example, the structural basis for the improved efficacy of nonfucosylated mAbs has been investigated.17 The X-ray crystal structure of the complex between nonfucosylated IgG Fc and a soluble form of FcγRIIIa carrying two N-linked glycans showed that one of two receptor glycans interacts with nonfucosylated Fc to stabilize the complex. It is proposed that when the Fc glycan is fucosylated this interaction is inhibited due to steric hindrance and, together with the negative effects of Fc fucosylation on the dynamics of the receptor binding site, this provides a rationale for the improved ADCC displayed by nonfucosylated IgG.A question of interest is precisely how Fc receptors bound to antibody ligands organize themselves within signaling complexes in the cell membrane. Some intriguing clues to this conundrum of molecular architecture are now surfacing. In mast cells, FcεRI molecules loaded with IgE form a synapse when presented with antigen that is mobile within a lipid bilayer, via coalescence into large cholesterol-rich clusters.18 Of particular relevance to the therapeutic setting, clustering of receptors into immune synapses is also seen with FcγR. For instance, during in vivo ADCC mediated by tumor-specific mAb, clustering of FcγR, actin and phosphotyrosines has been noted at contact zones between tumor cells and macrophages or neutrophils.19 The theme of the influence of the membrane lipid domain environment on Fc receptor function is taken up elsewhere. It has been shown, for example, that serine phosphorylation of FcγRI influences membrane mobility and function. The cytoplasmic tail of FcγRI interacts with protein 4.1G,20 and it is proposed that this is mediated via a phosphoserine-dependent mechanism critical for localization of the receptor to lipid rafts.21 With regard to FcγRIIa, a major role for lipid rafts in the regulation of IgG binding to FcγRIIa has been revealed.22 Notably, exclusion of FcγRIIa from lipid raft membrane microdomains is able to suppress IgG binding in myeloid cells.Increased knowledge of the capabilities of Fc receptors specific for other antibody classes is opening up new options for therapy. For example, IgA antibodies may offer a highly useful and efficacious alternative approach of particular relevance to treatment at mucosal sites. Human IgA mAbs have been demonstrated to mediate efficient tumor cell killing23,24 and to have the capability to control certain infectious diseases.25,26 The detailed understanding of functional sites in IgA that has resulted from numerous mutagenesis studies,27 coupled with improved ways to produce and isolate recombinant IgA mAbs28 should facilitate developments toward therapeutics based on this immunoglobulin class. Similarly, recent studies indicate that IgE may serve as an alternative to the classic IgG backbone for therapeutic antibodies.29Finally, technical innovations seem poised to further inform the field and advances are arriving or may be anticipated from techniques such as solution nuclear magnetic resonance (NMR) spectroscopy,30 cryo-electron tomography,31 single particle tracking,32 and ultrasensitive force techniques such as adhesion frequency assays.33,34Interest in Fc receptors continues unabated, and the contribution that the field can make to mAb development and optimisation is unquestionable. The FASEB SRC on Immunoreceptors will serve as a forum for discourse on the above issues and much more, providing invaluable information and networking opportunities for all those interested in ways to maximise the efficacy of mAbs and mAb-based reagents. Registration is open until 24 June 2012.  相似文献   

13.
Eukaryotic DNA polymerase η (Polη) confers ultraviolet (UV) resistance by catalyzing translesion synthesis (TLS) past UV photoproducts. Polη has been studied extensively in budding yeast and mammalian cells, where its interaction with monoubiquitylated proliferating cell nuclear antigen (PCNA) is necessary for its biological activity. Recently, in collaboration with other investigators, our laboratory demonstrated that Arabidopsis thaliana Polη is required for UV resistance in plants. Furthermore, the purified enzyme can perform TLS opposite a cyclobutane pyrimidine dimer and interacts with PCNA. Intriguingly, the biological activity of Polη in a heterologous yeast assay depends on co-expression with Arabidopsis PCNA2 and Polη sequences implicated in binding PCNA or ubiquitin. We suggest that interaction of Arabidopsis Polη with ubiquitylated PCNA2 is required for TLS past UV photoproducts by Polη.Key words: polymerase η, proliferating cell nuclear antigen, translesion synthesis, ubiquitin, Arabidopsis thaliana, ultraviolet radiationUltraviolet (UV)-induced pyrimidine dimers can block the progression of DNA replication forks potentially disrupting the replication machinery and resulting in cell death. For this reason, cells have evolved non-essential, low fidelity DNA polymerases (Pols) capable of copying damaged templates,1,2 a process termed translesion DNA synthesis (TLS). In budding yeast, TLS past UV photoproducts is catalyzed by Polη and Polζ (composed of the Rev3 catalytic and Rev7 accessory subunits), but also involves the Rev1 protein in an as yet undetermined role linked to Polζ.1,3,4 Yeast and human Polη replicates cyclobutane pyrimidine dimers (CPDs), in particular thymine-thymine (TT) CPDs, in a relatively error-free manner whereas Polζ is essential for UV mutagenesis implicating it in error-prone TLS.1,4,5Both UV resistance due to TLS and the polymerases responsible have been well-studied in yeast and mammalian cells over the past decade. Only more recently has evidence emerged that TLS may also contribute to UV resistance in plants. Arabidopsis thaliana POLH, REV1, REV3 and REV7 encode homologs of Polη, Rev1, Rev3 and Rev7, respectively.610 T-DNA insertions in POLH, REV1 or REV3 sensitise root growth to acute UV doses,68,10 and these mutations, as well as inactivation of REV7, increase the sensitivity of whole plants to longer term UV treatment.6,8 Interestingly, polh rev3 double mutants show an additive increase in UV sensitivity over that observed for polh and rev3 single mutants,6,10 potentially pointing to differences in the UV photoproducts bypassed by the two polymerases. That the enhanced UV sensitivity of the mutants may reflect a TLS deficiency is suggested by the finding that purified Arabidopsis Polη catalyzes primer extension and TLS past a TT CPD in vitro.6For TLS to occur, Polη must gain access to the replication machinery arrested at a UV photoproduct. It does so in yeast and mammalian cells by interacting with proliferating cell nuclear antigen (PCNA), the eukaryotic sliding clamp required for processive DNA replication.1,3,11, DNA damage or stalling of the replicative polymerase triggers monoubiquitylation of PCNA at lysine 164 by a complex of the E2 ubiquitin conjugase Rad6 and the E3 ubiquitin ligase Rad18.1,3,11,12 This modification increases the affinity of Polη for PCNA, with which it interacts via a single PCNA interacting peptide (PIP) box and a single ubiquitin-binding zinc finger (UBZ) domain.1,3In contrast to its yeast and mammalian counterparts, Polη from Arabidopsis and Oryza sativa (rice) has two PIP boxes and lacks a UBZ.6,9,10 Instead the two polymerases each possess two ubiquitin-binding motifs (UBMs) similar to those present in the Arabidopsis Rev1 protein and a vertebrate TLS polymerase, Pol., for which there is no homolog in Arabidopsis.6,13 Considerable differences in the sequences flanking the UBMs in Polη and Rev1 argue that Polη did not acquire its UBMs from Rev1, and so, although perhaps unique to plant Polη, their origin remains a mystery.The presence of PCNA- and ubiquitin-binding sequences in plant Polη hint that it may operate in TLS in a manner similar to that for Polη from yeast or mammalian cells. Indeed, three lines of evidence6 lead us to suggest that the Polη PIP boxes and UBMs likely function in binding ubiquitylated PCNA and this interaction is probably required for TLS past UV photoproducts by Arabidopsis Polη. First, Arabidopsis Polη interacts physically and in yeast two-hybrid assays with Arabidopsis PCNA1 and PCNA2. Second, expression in yeast of Arabidopsis cDNAs encoding Polη and PCNA2, but not PCNA1, fully complements the UV sensitivity conferred by elimination of yeast Polη. In vitro mutagenesis suggests the inability of Polη plus PCNA1 to restore UV resistance is due to a lysine at position 201 in PCNA1 but not PCNA2. In the three-dimensional structure of PCNA, amino acid 201 lies adjacent to lysine-164, the residue that is ubiquitylated in yeast and human PCNA. Thus, one possibility is that lysine-201 in PCNA1 prevents complementation of UV sensitivity by inhibiting ubiquitylation of lysine-164. Third, altering presumed critical residues in either of the two PIP boxes or UBM2 in Arabidopsis Polη also prevents restoration of UV resistance in Polη-deficient yeast cells.Several important parts of the puzzle remain to be solved. In particular, the ubiquitylation of plant PCNA has yet to be demonstrated, and the identity of the proteins that might monoubiquitylate plant PCNA is uncertain. Although Arabidopsis Rad6 homologs can ubiquitylate target proteins in vitro, there is no evidence that Arabidopsis PCNA1 or PCNA2 is a substrate, and Arabidopsis lacks a Rad18 homolog.14,15 Finally, if PCNA is ubiquitylated in planta, does this occur at lysine-164 in response to DNA damage or replication fork stalling, is the interaction of Polη with PCNA stimulated by this modification, and is an enhanced interaction mediated by the Polη UBMs?  相似文献   

14.
15.
Stomata are formed by pairs of surrounding guard cells and perform important roles in photosynthesis, transpiration and innate immunity of terrestrial plants. Ionic solutes in the cytosol of guard cells are important for cell turgor and volume change. Consequently, trans-membrane flux of ions such as K+, Cl, and malate2 through K+ channels and anion channels of guard cells are a direct driving force for turgor change, while the opening of calcium permeable channels can serve as a trigger of cytosolic free calcium concentration elevations or oscillations, which play second messenger roles. In plants, heterotrimeric G proteins have fewer members than in animals, but they are well investigated and found to regulate these channels and to play fundamental roles in guard cell function. This mini-review focuses on the recent understanding of G-protein regulation of ion channels on the plasma membrane of guard cells and their participation in stomatal movements.Key words: guard cell, heterotrimeric G protein, ion channel, arabidopsis thaliana, stomata, plasma membrane, patch clampHeterotrimeric G proteins, composed of Gα, Gβ and Gγ subunits, are key elements of cellular signal transduction networks. In plant species, fewer members of G proteins are present than in animals. For example, only one Gα subunit (GPA1), one Gβ subunit (AGB1) and two Gγ subunits (AGG1 and AGG2) are reported in Arabidopsis while 23 Gα, 5 Gβ and 12 Gγ subunits have been identified in human.1 All three kinds of subunits are expressed in guard cells. Ubiquitous expression of GPA1 throughout plant was ascertained by northern and promoter::GUS analyses and RT-PCR results also indicate guard cell expression.24 AGB1 is ubiquitously expressed throughout the plant and its promoter::GUS transgenic lines show strong expression in guard cells.57 For Gγ subunits, RNA blots show AGG1 and AGG2 expression throughout the plant, however, reporter gene analysis shows guard cell expression of AGG2 but not AGG1.79 The guard cell expression of G protein subunits implies the function of G protein in guard cell signaling and stomatal movement regulation.Stomata are microscopic pores in the epidermis of terrestrial plants, which serve as the mouths of plants for gas change since through them CO2 enters leaves for photosynthesis and water vapor is lost as transpiration.1013 In addition, stomatal movements induced by pathogen and pathogen/microbe-associated molecular patterns (PAMPs or MAMPs) are a component of the plant innate immunity system.1416 Biotic and abiotic stresses (e.g. water deficiency, cold, pathogens) and their induced phytohormone changes (e.g. abscisic acid [ABA], ethylene) have been widely investigated in stomatal movement regulation, and stomatal apertures are directly regulated by volume change of the surrounding guard cell pairs. The accumulation/release of ionic solutes through ion channels on the guard-cell plasma membrane together with malate production/metabolism induces water influx/efflux driving increase/decrease of cell turgor and volume which co-operates with the radial reinforcement of the guard cell walls to widen/shrink stomatal aperture.10,17 Given that mature guard cells lack plasmodesmata with neighboring cells, all ion uptake and efflux must pass through ion channels and ion transporters on the plasma membrane.In Arabidopsis guard cells, the model cell type for cell signaling of the model plant species, all three kinds of ion channels (K+ channels, anion channels and Ca2+-permeable channels) have been investigated and found to be regulated by heterotrimeric G proteins.10,17 Their ion channel activities can be measured in intact guard cells, guard cell protoplasts, or cell membrane patches using the patch clamp technique.15,18,19 Patch clamping can be used to measure ion fluxes in whole cells or even through a single ion channel.20,21 The patch clamp technique under the whole-cell recording configuration can measure the currents through hyperpolarization-activated inward K+ channels which account for K+ accumulation during stomatal opening, and the depolarizationactivated outward K+ channels which, together with R-type and S-type anion channels, mediate solute removal during stomatal closure. Besides these ionic fluxes which directly elicit changes in turgor, Ca2+-permeable channels which participate in Ca2+ signaling are also regulated by G proteins. For better visualization of the currents through K+, anion and Ca2+permeable channels, real current traces and their idealized current/voltage relationships are indicated in Figure 1. The G-protein regulation of inward and outward K+ channels, S-type anion channels, and Ca2+-permeable channels and their significance for stomatal movements will be discussed below, and the genes encoding them which have been explored up to now also will be discussed.Open in a separate windowFigure 1Current traces and idealized current/voltage relationships of wild type guard cell plasma membrane ion channels involved in G-protein regulation (A–C), ABA inhibition of whole-cell inward K+ currents. (A) indicates inward K+ currents of wild type guard cell protoplasts in response to hyperpolarizing voltages under control conditions [Scale bar is shown in (B)]; (B) indicates inward K+ currents of wild type guard cell protoplasts with ABA treatment; (C) indicates the idealized current/voltage relationship of inward K+ currents for control (gray) and ABA treatments (black). (D–F), ABA activation of slow anion currents. (D) indicates anion currents of wild type under control condition and (E) shows current after ABA treatment; (F) indicates the idealized current/voltage relationship of anion currents for control (gray) and ABA treatments (black). (G–I), ABA activation of currents through Ca2+-permeable channels. (G) indicates currents through Ca2+-permeable channels of wild type under control condition and (H) shows current after ABA treatments; (I) indicates the idealized current/voltage relationship of currents through Ca2+-permeable channels for control (gray) and ABA treatments (black).  相似文献   

16.
In animals, heterotrimeric G proteins, comprising Gα, Gβ, and Gγ subunits, are molecular switches whose function tightly depends on Gα and Gβγ interaction. Intriguingly, in Arabidopsis (Arabidopsis thaliana), multiple defense responses involve Gβγ, but not Gα. We report here that the Gβγ dimer directly partners with extra-large G proteins (XLGs) to mediate plant immunity. Arabidopsis mutants deficient in XLGs, Gβ, and Gγ are similarly compromised in several pathogen defense responses, including disease development and production of reactive oxygen species. Genetic analysis of double, triple, and quadruple mutants confirmed that XLGs and Gβγ functionally interact in the same defense signaling pathways. In addition, mutations in XLG2 suppressed the seedling lethal and cell death phenotypes of BRASSINOSTEROID INSENSITIVE1-associated receptor kinase1-interacting receptor-like kinase1 mutants in an identical way as reported for Arabidopsis Gβ-deficient mutants. Yeast (Saccharomyces cerevisiae) three-hybrid and bimolecular fluorescent complementation assays revealed that XLG2 physically interacts with all three possible Gβγ dimers at the plasma membrane. Phylogenetic analysis indicated a close relationship between XLGs and plant Gα subunits, placing the divergence point at the dawn of land plant evolution. Based on these findings, we conclude that XLGs form functional complexes with Gβγ dimers, although the mechanism of action of these complexes, including activation/deactivation, must be radically different form the one used by the canonical Gα subunit and are not likely to share the same receptors. Accordingly, XLGs expand the repertoire of heterotrimeric G proteins in plants and reveal a higher level of diversity in heterotrimeric G protein signaling.Heterotrimeric GTP-binding proteins (G proteins), classically consisting of Gα, Gβ, and Gγ subunits, are essential signal transduction elements in most eukaryotes. In animals and fungi, ligand perception by G protein-coupled receptors leads to replacement of GDP with GTP in Gα, triggering activation of the heterotrimer (Li et al., 2007; Oldham and Hamm, 2008). Upon activation, GTP-bound Gα and Gβγ are released and interact with downstream effectors, thereby transmitting signals to multiple intracellular signaling cascades. Signaling terminates when the intrinsic GTPase activity of Gα hydrolyzes GTP to GDP and the inactive heterotrimer reforms at the receptor. The large diversity of mammalian Gα subunits confers specificity to the multiple signaling pathways mediated by G proteins (Wettschureck and Offermanns, 2005). Five distinct classes of Gα have been described in animals (Gαi, Gαq, Gαs, Gα12 and Gαv), with orthologs found in evolutionarily primitive organisms such as sponges (Oka et al., 2009). Humans possess four classes of Gα involving 23 functional isoforms encoded by 16 genes (McCudden et al., 2005), while only a single prototypical Gα is usually found per plant genome (Urano et al., 2013). Multiple copies of Gα are present in some species with recently duplicated genomes, such as soybean (Glycine max) with four Gα genes (Blanc and Wolfe, 2004; Bisht et al., 2011). In the model plant Arabidopsis (Arabidopsis thaliana), a prototypical Gα subunit (GPA1) is involved in a number of important processes, including cell proliferation (Ullah et al., 2001), inhibition of inward K+ channels and activation of anion channels in guard cells by mediating the abscisic acid pathway (Wang et al., 2001; Coursol et al., 2003), blue light responses (Warpeha et al., 2006, 2007), and germination and postgermination development (Chen et al., 2006; Pandey et al., 2006).It is well established that heterotrimeric G proteins play a fundamental role in plant innate immunity. In Arabidopsis, two different Gβγ dimers (Gβγ1 and Gβγ2) are generally considered to be the predominant elements in G protein defense signaling against a variety of fungal pathogens (Llorente et al., 2005; Trusov et al., 2006, 2007, 2009; Delgado-Cerezo et al., 2012; Torres et al., 2013). By contrast, these studies attributed a small or no role to Gα, because mutants deficient in Gα displayed only slightly increased resistance against the fungal pathogens (Llorente et al., 2005; Trusov et al., 2006; Torres et al., 2013). The Gβγ-mediated signaling also contributes to defense against a model bacterial pathogen Pseudomonas syringae, by participating in programmed cell death (PCD) and inducing reactive oxygen species (ROS) production in response to at least three pathogen-associated molecular patterns (PAMPs; Ishikawa, 2009; Liu et al., 2013; Torres et al., 2013). Gα is not involved in PCD or PAMP-triggered ROS production (Liu et al., 2013; Torres et al., 2013). Nonetheless, Arabidopsis Gα plays a positive role in defense against P. syringae, probably by mediating stomatal function and hence physically restricting bacterial entry to the leaf interior (Zhang et al., 2008; Zeng and He, 2010; Lee et al., 2013). Given the small contribution from Gα, the involvement of heterotrimeric G proteins in Arabidopsis resistance could be explained in two ways: either the Gβγ dimer acts independently from Gα, raising a question of how is it activated upon a pathogen attack, or Gα is replaced by another protein for heterotrimer formation.The Arabidopsis genome contains at least three genes encoding Gα-like proteins that have been classified as extra-large G proteins (XLGs; Lee and Assmann, 1999; Ding et al., 2008). XLGs comprise two structurally distinct regions. The C-terminal region is similar to the canonical Gα, containing the conserved helical and GTPase domains, while the N-terminal region is a stretch of approximately 400 amino acids including a putative nuclear localization signal (Ding et al., 2008). GTP binding and hydrolysis were confirmed for all three XLG proteins, although their enzymatic activities are very slow and require Ca2+ as a cofactor, whereas canonical Gα utilizes Mg2+ (Heo et al., 2012). Several other features differentiate XLGs from Gα subunits. Comparative analysis of XLG1 and Gα at the DNA level showed that the genes are organized in seven and 13 exons, respectively, without common splicing sites (Lee and Assmann, 1999). XLGs have been reported to localize to the nucleus (Ding et al., 2008). Analysis of knockout mutants revealed a nuclear function for XLG2, as it physically interacts with the Related To Vernalization1 (RTV1) protein, enhancing the DNA binding activity of RTV1 to floral integrator gene promoters and resulting in flowering initiation (Heo et al., 2012). Therefore, it appears that XLGs may act independently of G protein signaling. On the other hand, functional similarities between XLGs and the Arabidopsis Gβ subunit (AGB1) were also discovered. For instance, XLG3- and Gβ-deficient mutants were similarly impaired in root gravitropic responses (Pandey et al., 2008). Knockout of all three XLG genes caused increased root length, similarly to the Gβ-deficient mutant (Ding et al., 2008). Furthermore, as observed in Gβ-deficient mutants, xlg2 mutants displayed increased susceptibility to P. syringae, indicating a role in plant defense (Zhu et al., 2009). Nevertheless, a genetic analysis of the possible functional interaction between XLGs and Gβ has not been established.In this report, we performed in-depth genetic analyses to test the functional interaction between the three XLGs and Gβγ dimers during defense-related responses in Arabidopsis. We also examined physical interaction between XLG2 and the Gβγ dimers using yeast (Saccharomyces cerevisiae) three-hybrid (Y3H) and bimolecular fluorescent complementation (BiFC) assays. Our findings indicate that XLGs function as direct partners of Gβγ dimers in plant defense signaling. To estimate relatedness of XLGs and Gα proteins, we carried out a phylogenetic analysis. Based on our findings, we conclude that plant XLG proteins most probably originated from a canonical Gα subunit and retained prototypical interaction with Gβγ dimers. They function together with Gβγ in a number of processes including plant defense, although they most probably evolved activation/deactivation mechanisms very different from those of a prototypical Gα.  相似文献   

17.
In model organisms, resistance to inhibitors of cholinesterase 8 (Ric-8), a G protein α (Gα) subunit guanine nucleotide exchange factor (GEF), functions to orient mitotic spindles during asymmetric cell divisions; however, whether Ric-8A has any role in mammalian cell division is unknown. We show here that Ric-8A and Gαi function to orient the metaphase mitotic spindle of mammalian adherent cells. During mitosis, Ric-8A localized at the cell cortex, spindle poles, centromeres, central spindle, and midbody. Pertussis toxin proved to be a useful tool in these studies since it blocked the binding of Ric-8A to Gαi, thus preventing its GEF activity for Gαi. Linking Ric-8A signaling to mammalian cell division, treatment of cells with pertussis toxin, reduction of Ric-8A expression, or decreased Gαi expression similarly affected metaphase cells. Each treatment impaired the localization of LGN (GSPM2), NuMA (microtubule binding nuclear mitotic apparatus protein), and dynein at the metaphase cell cortex and disturbed integrin-dependent mitotic spindle orientation. Live cell imaging of HeLa cells expressing green fluorescent protein-tubulin also revealed that reduced Ric-8A expression prolonged mitosis, caused occasional mitotic arrest, and decreased mitotic spindle movements. These data indicate that Ric-8A signaling leads to assembly of a cortical signaling complex that functions to orient the mitotic spindle.The cortical capture of astral microtubules is essential to generate the forces needed for mitotic spindle positioning for both symmetric and asymmetric cell divisions (23, 29). Failure to either capture astral microtubules or the inappropriate application of pulling forces adversely affects mitotic spindle orientation, and can impede embryogenesis and alter cell fate decisions. Studies examining mitotic spindle orientation in Drosophila embryonic and larval neuroblasts have identified two critical pathways, the Gα/Pins/Mud pathway and the Pins/Dlg/Khc73 pathway (29). The heterotrimeric G-protein α subunit (Gα), Pins (Partner-of-Inscuteable), and Mud (Mushroom body defect) are members of an evolutionarily conserved noncanonical G-protein signaling pathway, which form a tripartite protein complex linked to the apical Par complex by the adapter protein Inscuteable (29, 37). Reducing the level of Gαi, Pins, or Mud prevents neuroblast mitotic spindle alignment. A second spindle orientation pathway involves Pins, the tumor suppressor Discs large (Dlg) and the microtubule plus-end-directed kinesin heavy chain 73 (Khc73). Khc73 binds Dlg and coimmunoprecipitates with Pins. Khc73 localized to astral microtubules can induce Pins-Dlg cortical polarity (27).In canonical G-protein signaling pathways, the binding of ligand to a seven-transmembrane receptor triggers a heterotrimeric G-protein α subunit (Gα) to exchange GTP for GDP, resulting in the dissociation of the Gα subunit from its associated Gβγ heterodimer (12, 20). This exposes interactive sites in the Gα and Gβγ subunits, allowing their binding to and activation of downstream effectors. Since Gα subunits possess an intrinsic GTPase activity, GTP hydrolysis leads to the reassembly of heterotrimeric G protein causing signaling to cease. In noncanonical G-protein signaling the seven-transmembrane receptor is replaced by an intracellular guanine nucleotide exchange factor, such as Ric-8 (37). In studies in Drosophila and Caenorhabditis elegans Ric-8 has been shown to positively regulate Gαi activity and is essential for asymmetric cell divisions (1, 2, 5, 8, 11, 36). Although initially characterized as a guanine nucleotide exchange factor (GEF) for isolated Gαsubunits, more recent biochemical studies have shown that Ric-8A (the mammalian equivalent of Ric-8) also acts on a complex of GDP-Gαi, the mammalian Pins homolog LGN, and NuMA (nuclear mitotic apparatus protein; the mammalian equivalent of Mud) catalytically releasing GTP-Gαi and causing liberation of NuMA from LGN (30, 31). Ric-8A can also catalyze guanine nucleotide exchange on Gαi1 bound to the GPR/GoLoco exchange inhibitor AGS3, a paralog of LGN (33). During mitosis the N-terminal portion of LGN binds NuMA and the C-terminal domain binds GDP-Gαi and the trimolecular complex localizes to the cell cortex, where the dynamic release of NuMA from LGN may regulate aster microtubule pulling during cell division (3, 9, 10, 22).In the present study we examined the role of Ric-8A in mitotic spindle orientation in adherent cells and in polarized MDCK cells. In nonpolarized adherent cells cell such as HeLa, integrin mediated cell-substrate adhesion orients the mitotic spindle parallel to the substratum, and thereby both daughter cells remain attached. This requires the actin cytoskeleton, astral microtubules, the microtubule plus end tracking protein EB1, myosin X, cdc42, LIM kinase 1, and phosphatidylinositol(3,4,5)-triphosphate (PIP3) (13, 18, 32, 34, 35). PIP3 may direct dynein/dynactin-dependent pulling forces on the spindle midcortex to orient the mitotic spindle (34). In polarized cells such as Madin-Darby canine kidney (MDCK) cells, the mitotic spindle is constrained by the topology of the cell and cortical cues provided by adherens junctions (24). In contrast to HeLa cells these cues are insensitive to phosphatidylinositol 3-kinase (PI3K) inhibition, which blocks the generation of PIP3 (34). We found that inhibiting either Ric-8A or Gαi expression impairs the orientation of the metaphase mitotic spindle in HeLa cells and pertussis toxin, which blocks Ric-8A triggered nucleotide exchange, disrupts the normal mitotic spindle alignment of both HeLa and MDCK cells. Impairment of Ric-8A expression or function inhibits the localization of Gαi1, LGN, NuMA, and dynein to the metaphase cortex opposite the spindle poles.  相似文献   

18.
Upon ligand binding, G-protein-coupled receptors (GPCRs) impart the signal to heterotrimeric G proteins composed of α, β, and γ subunits, leading to dissociation of the Gα subunit from the Gβγ subunit. While the Gα subunit is imperative for downstream signaling, the Gβγ subunit, in its own right, mediates a variety of cellular responses such as GPCR desensitization via recruiting GRK to the plasma membrane and AKT stimulation. Here we report a mode of spatial regulation of the Gβγ subunit through alteration in subcellular compartmentation. RKTG (Raf kinase trapping to Golgi apparatus) is a newly characterized membrane protein specifically localized at the Golgi apparatus. The N terminus of RKTG interacts with Gβ and tethers Gβγ to the Golgi apparatus. Overexpression of RKTG impedes the interaction of Gβγ with GRK2, abrogates the ligand-induced change of subcellular distribution of GRK2, reduces isoproterenol-stimulated phosphorylation of the β2-adrenergic receptor (β2AR), and alters β2AR desensitization. In addition, RKTG inhibits Gβγ- and ligand-mediated AKT phosphorylation that is enhanced in cells with downregulation of RKTG. Silencing of RKTG also alters GRK2 internalization and compromises ligand-induced Gβ translocation to the Golgi apparatus. Taken together, our results reveal that RKTG can modulate GPCR signaling through sequestering Gβγ to the Golgi apparatus and thereby attenuating the functions of Gβγ.Heterotrimeric G proteins are composed of distinct Gα, β, and γ subunits which relay extracellular signals from heptahelical G-protein-coupled receptors (GPCRs) to downstream effectors (16, 25, 30). Gα binds Gβγ when Gα is bound with GDP but dissociates from Gβγ after GDP is replaced with GTP upon activation of GPCRs by extracellular ligand (25). Under physiologic conditions, the Gβ and Gγ subunits form a dimer in which the two subunits are not separable (10, 30). Although Gα is the primary protein that transmits the signal of GPCRs to specific intracellular effectors, such as adenylyl cyclase and phospholipase C, emerging evidence has indicated that Gβγ is able to regulate GPCR signaling through interacting with GPCRs, the Gα subunit, and downstream effectors (30). Predominantly, Gβγ is able to directly interact with and affect the functions of a variety of membrane and intracellular effectors, such as ion channels, adenylyl cyclase, G-protein-coupled receptor kinases (GRKs), and phosphatidylinositol 3-kinase (PI3K) (30). The current model of Gβγ-mediated signaling restricts it mostly to the plasma membrane (PM) (30). In the case of membrane-bound effectors, such as adenylyl cyclases or GIRK channels, Gβγ regulates the activities of these transmembrane proteins through conformational alteration. In the case of cytosolic proteins such as PLCβ2 or GRK2, whose substrates are localized to PM, Gβγ regulates their activity by recruiting the proteins to PM. The activity of Gβγ is primarily regulated by GPCR and Gα, in which GPCR activation leads to conformational changes of Gα. Such change causes replacement of Gα-bound GDP with GTP and release of Gβγ from the heterotrimeric G proteins. The activity of Gβγ could also be regulated by interacting with cytosolic proteins such as RACK1 (7). However, how Gβγ-mediated signaling is regulated in a spatial manner via subcellular compartmentation is largely unknown.GRK2 is a member of a family of GRKs that can phosphorylate the agonist-occupied GPCRs (4). Specific phosphorylation of activated receptors is associated with a decreased responsiveness of GPCR to prolonged stimulation by the agonist, also known as desensitization (15, 26). Gβγ regulates the activities of GRK2 and GRK3 toward several GPCRs (9). In cooperation with phosphatidylinositol 4,5-bisphosphate, Gβγ binds to the pleckstrin homology (PH) domain of GRK2 and recruits GRK2 to PM, in which it phosphorylates activated GPCRs (18, 30). The crystallographic structure of GRK2 in complex with Gβ1γ2 has been solved (20, 32). On the other hand, AKT is an intracellular target of PI3K and plays a critical role in cell growth, proliferation, and survival. It has been reported that Gβγ could activate AKT in a PI3K-dependent fashion (5), and Gβγ could mediate AKT activation at endosomes (13). Recent data also indicate that the p110β subunit of PI3K signals downstream of GPCR, and the AKT activation mediated by p110β is G protein dependent (14, 17).PAQR3 is a member of the progestin and adipoQ receptor (PAQR) family, and the members of this family are predicted to have seven transmembrane domains similar to GPCRs (31). Recently, we demonstrated that PAQR3 is localized at the Golgi apparatus and is involved in the spatial regulation of Raf kinase, whereby this protein was named Raf kinase trapping to Golgi apparatus (RKTG) (12). Biochemical analysis of RKTG suggested that its N terminus is localized on the cytoplasmic side of the Golgi membrane (21). Using the N terminus of RKTG to screen a Saccharomyces cerevisiae two-hybrid library, we determined that RKTG is able to interact with Gβ, and detailed analyses indicate that RKTG is a spatial regulator of Gβγ signaling.  相似文献   

19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号