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 421 毫秒
1.
A Boolean network is a model used to study the interactions between different genes in genetic regulatory networks. In this paper, we present several algorithms using gene ordering and feedback vertex sets to identify singleton attractors and small attractors in Boolean networks. We analyze the average case time complexities of some of the proposed algorithms. For instance, it is shown that the outdegree-based ordering algorithm for finding singleton attractors works in time for , which is much faster than the naive time algorithm, where is the number of genes and is the maximum indegree. We performed extensive computational experiments on these algorithms, which resulted in good agreement with theoretical results. In contrast, we give a simple and complete proof for showing that finding an attractor with the shortest period is NP-hard.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32]  相似文献   

2.
Our previous work using a melanoma progression model composed of melanocytic cells (melanocytes, primary and metastatic melanoma samples) demonstrated various deregulated genes, including a few known lncRNAs. Further analysis was conducted to discover novel lncRNAs associated with melanoma, and candidates were prioritized for their potential association with invasiveness or other metastasis‐related processes. In this sense, we found the intergenic lncRNA U73166 (ENSG00000230454) and decided to explore its effects in melanoma. For that, we silenced the lncRNA U73166 expression using shRNAs in a melanoma cell line. Next, we experimentally investigated its functions and found that migration and invasion had significantly decreased in knockdown cells, indicating an essential association of lncRNA U73166 for cancer processes. Additionally, using naïve and vemurafenib‐resistant cell lines and data from a patient before and after resistance, we found that vemurafenib‐resistant samples had a higher expression of lncRNA U73166. Also, we retrieved data from the literature that indicates lncRNA U73166 may act as a mediator of RNA processing and cell invasion, probably inducing a more aggressive phenotype. Therefore, our results suggest a relevant role of lncRNA U73166 in metastasis development. We also pointed herein the lncRNA U73166 as a new possible biomarker or target to help overcome clinical vemurafenib resistance.  相似文献   

3.
Mathematical tools developed in the context of Shannon information theory were used to analyze the meaning of the BLOSUM score, which was split into three components termed as the BLOSUM spectrum (or BLOSpectrum). These relate respectively to the sequence convergence (the stochastic similarity of the two protein sequences), to the background frequency divergence (typicality of the amino acid probability distribution in each sequence), and to the target frequency divergence (compliance of the amino acid variations between the two sequences to the protein model implicit in the BLOCKS database). This treatment sharpens the protein sequence comparison, providing a rationale for the biological significance of the obtained score, and helps to identify weakly related sequences. Moreover, the BLOSpectrum can guide the choice of the most appropriate scoring matrix, tailoring it to the evolutionary divergence associated with the two sequences, or indicate if a compositionally adjusted matrix could perform better.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29]  相似文献   

4.
Plasmodesmata (Pd) are membranous channels that serve as a major conduit for cell-to-cell communication in plants. The Pd-associated β-1,3-glucanase (BG_pap) and CALLOSE BINDING PROTEIN1 (PDCB1) were identified as key regulators of Pd conductivity. Both are predicted glycosylphosphatidylinositol-anchored proteins (GPI-APs) carrying a conserved GPI modification signal. However, the subcellular targeting mechanism of these proteins is unknown, particularly in the context of other GPI-APs not associated with Pd. Here, we conducted a comparative analysis of the subcellular targeting of the two Pd-resident and two unrelated non-Pd GPI-APs in Arabidopsis (Arabidopsis thaliana). We show that GPI modification is necessary and sufficient for delivering both BG_pap and PDCB1 to Pd. Moreover, the GPI modification signal from both Pd- and non-Pd GPI-APs is able to target a reporter protein to Pd, likely to plasma membrane microdomains enriched at Pd. As such, the GPI modification serves as a primary Pd sorting signal in plant cells. Interestingly, the ectodomain, a region that carries the functional domain in GPI-APs, in Pd-resident proteins further enhances Pd accumulation. However, in non-Pd GPI-APs, the ectodomain overrides the Pd targeting function of the GPI signal and determines a specific GPI-dependent non-Pd localization of these proteins at the plasma membrane and cell wall. Domain-swap analysis showed that the non-Pd localization is also dominant over the Pd-enhancing function mediated by a Pd ectodomain. In conclusion, our results indicate that segregation between Pd- and non-Pd GPI-APs occurs prior to Pd targeting, providing, to our knowledge, the first evidence of the mechanism of GPI-AP sorting in plants.Plant cells are interconnected with cross-wall membranous channels called plasmodesmata (Pd). Recent studies have shown that the region of the plasma membrane (PM) lining the Pd channel is a specialized membrane microdomain whose lipid and protein composition differs from the rest of the PM (Tilsner et al., 2011, 2016; Bayer et al., 2014; González-Solís et al., 2014; Grison et al., 2015). In a similar manner, the cell wall domain surrounding the Pd channel is specialized and, unlike the rest of the cell wall, is devoid of cellulose, rich in pectin, and contains callose (an insoluble β-1,3-glucan; Zavaliev et al., 2011; Knox and Benitez-Alfonso, 2014). In response to physiological signals, callose can be transiently deposited and degraded at Pd, which provides a mechanism for controlling the Pd aperture in diverse developmental and stress-related processes (Zavaliev et al., 2011). Control of Pd functioning is mediated by proteins that are specifically targeted to Pd. Plasmodesmal proteins localized to the PM domain of Pd can be integral transmembrane proteins, such as Pd-localized proteins (Thomas et al., 2008), the receptor kinase ARABIDOPSIS CRINKLY4 (Stahl et al., 2013), and callose synthases (Vatén et al., 2011). Alternatively, Pd proteins can associate with the membrane through a lipid modification like myristoylation (e.g. remorins; Raffaele et al., 2009) or be attached by a glycosylphosphatidylinositol (GPI) anchor (e.g. Pd-associated β-1,3-glucanases [BG_pap]; Levy et al., 2007; Rinne et al., 2011; Benitez-Alfonso et al., 2013), Pd-associated callose-binding proteins (PDCBs; Simpson et al., 2009), and LYSIN MOTIF DOMAIN-CONTAINING PROTEIN2 (LYM2; Faulkner et al., 2013).Among the known Pd proteins involved in Pd-specific callose degradation is BG_pap, a cell wall enzyme carrying a glycosyl hydrolase family 17 (GH17) module as its functional domain (Levy et al., 2007). Another group of proteins controlling callose dynamics at Pd are PDCBs that harbor a callose-binding domain termed carbohydrate-binding module 43 (CBM43), implicated in stabilizing callose at Pd (Simpson et al., 2009). Some β-1,3-glucanases may combine the two callose-modifying activities by harboring both GH17 and CBM43 functional domains, and several such proteins were shown to localize to Pd (Rinne et al., 2011; Benitez-Alfonso et al., 2013; Gaudioso-Pedraza and Benitez-Alfonso, 2014). A distinct feature of BG_pap and PDCBs is that both are predicted glycosylphosphatidylinositol-anchored proteins (GPI-APs). The GPI anchor is a form of posttranslational modification common to many cell surface proteins in all eukaryotes. GPI-APs are covalently attached to the outer leaflet of the PM through the GPI anchor. The basic structure of the anchor consists of ethanolamine phosphate, followed by a glycan chain of three Man residues and glucosamine, followed by phosphatidylinositol lipid moiety (EtNP-6Manα1-2Manα1-6Manα1-4GlcNα1-6myoinositol-1-P-lipid; Ferguson et al., 2009). All predicted GPI-APs carry an N-terminal secretion signal peptide (SP) similar to other secreted proteins. Distinctly, GPI-APs also carry a structurally conserved 25- to 30-residue C-terminal GPI attachment signal, which typically begins with a small amino acid (e.g. Ala, Asn, Asp, Cys, Gly, or Ser) termed omega, followed by a spacer region of five to 10 polar residues, and ending with a transmembrane segment of 15 to 20 hydrophobic residues (Ferguson et al., 2009). The entire region between the N-terminal and the C-terminal signals of a GPI-AP is termed the ectodomain and carries the protein’s functional domain(s). The GPI modification process takes place in the lumenal face of the endoplasmic reticulum (ER) in a cotranslational manner. Upon translocation into the ER, a GPI-AP is stabilized in the ER membrane by its C-terminal signal, which is concurrently cleaved after the omega amino acid, and a preassembled GPI anchor is covalently attached to the C terminus of the omega amino acid. After attachment to a protein, the GPI anchor undergoes a series of modifications (remodeling), both at the glucan chain and at the lipid moiety. Such remodeling is crucial for the sorting of GPI-APs in the secretory pathway and the subsequent lateral heterogeneity at the PM (Kinoshita, 2015). In particular, the addition of saturated fatty acid chains to the lipid moiety of the anchor leads to the enriched accumulation of GPI-APs in the PM microdomains, also termed lipid rafts (Muñiz and Zurzolo, 2014). In Arabidopsis (Arabidopsis thaliana), GPI modification has been predicted for 210 proteins of diverse functions at the PM or the cell wall or both (Borner et al., 2002). Despite extensive research on the GPI modification pathway and the function of GPI-APs in mammalian and yeast cells, such knowledge in plant systems is scarce. In particular, despite an emerging role of GPI-APs in the regulation of the cell wall domain of Pd, their subcellular targeting and compartmentalization mechanism have not been studied. In addition, it is not known how the targeting mechanism of Pd-resident GPI-APs is different from that of other classes of GPI-APs, which are not localized to Pd.In this study, we investigated the subcellular targeting mechanism of Pd-associated callose-modifying GPI-APs, BG_pap and PDCB1, and compared it with that of two unrelated non-Pd GPI-APs, ARABINOGALACTAN PROTEIN4 (AGP4) and LIPID TRANSFER PROTEIN1 (LTPG1). Using sequential fluorescent labeling of protein domains, we found that the C-terminal GPI modification signal present in both Pd- and non-Pd GPI-APs can function as a primary signal in targeting proteins to the Pd-enriched PM domain. Moreover, we show that while the GPI signal is sufficient for Pd targeting, the ectodomains in BG_pap and PDCB1 further enhance their accumulation at Pd. In contrast, the ectodomains in non-Pd GPI-APs mediate exclusion of the proteins from the Pd-enriched targeting pathway. The Pd exclusion effect was found to be dominant over the Pd-targeting function of the GPI signal and the Pd-enhancing function of the Pd ectodomain, and it possibly occurs prior to PM localization. Our findings thus uncover a novel Pd-targeting signal and provide, to our knowledge, the first evidence of the cellular mechanism that regulates the sorting of GPI-APs in plants.  相似文献   

5.
The purpose of this table is to provide the community with a citable record of publications of ongoing genome sequencing projects that have led to a publication in the scientific literature. While our goal is to make the list complete, there is no guarantee that we may have omitted one or more publications appearing in this time frame. Readers and authors who wish to have publications added to subsequent versions of this list are invited to provide the bibliographic data for such references to the SIGS editorial office.

Phylum Euryarchaeota

Phylum Crenarchaeota

Phylum Deinococcus-Thermus

Phylum Proteobacteria

Phylum Tenericutes

Phylum Firmicutes

Phylum Actinobacteria

Non-Bacterial genomes

  相似文献   

6.
Klinfelter syndrome was first described in adult males with gynecomastia, azoospermia and hypergonadotropic hypogonadism. Children with the 47, XXY karyotype demonstrate few clinical findings so Klinefelter syndrome is rarely diagnosed until adult life. Besides children who have been diagnosed during prenatal genetic testing, in infancy a male with 47, XXY (or variants: 46, XY-47, XXY; 48, XXXY; 48, XXYY, 49, XXXXY) may be found while undergoing evaluation of micropenis, hypospadias, cryptorchidism or facial anomalies. The older child may present with learning disabilities, behavior disorders or tall stature. At the time of puberty, the clinical picture includes small testes, gynecomastia and an eunuchoid habitus. Early diagnosis of Klinefelter syndrome must be performed since it has been demonstrated that early treatment with androgens may ameliorate many aspects of the clinical symptoms and attenuate or prevent behavioral and psychiatric disorders associated with 47, XXY males.  相似文献   

7.
The inference of gene regulatory networks is a key issue for genomic signal processing. This paper addresses the inference of probabilistic Boolean networks (PBNs) from observed temporal sequences of network states. Since a PBN is composed of a finite number of Boolean networks, a basic observation is that the characteristics of a single Boolean network without perturbation may be determined by its pairwise transitions. Because the network function is fixed and there are no perturbations, a given state will always be followed by a unique state at the succeeding time point. Thus, a transition counting matrix compiled over a data sequence will be sparse and contain only one entry per line. If the network also has perturbations, with small perturbation probability, then the transition counting matrix would have some insignificant nonzero entries replacing some (or all) of the zeros. If a data sequence is sufficiently long to adequately populate the matrix, then determination of the functions and inputs underlying the model is straightforward. The difficulty comes when the transition counting matrix consists of data derived from more than one Boolean network. We address the PBN inference procedure in several steps: (1) separate the data sequence into "pure" subsequences corresponding to constituent Boolean networks; (2) given a subsequence, infer a Boolean network; and (3) infer the probabilities of perturbation, the probability of there being a switch between constituent Boolean networks, and the selection probabilities governing which network is to be selected given a switch. Capturing the full dynamic behavior of probabilistic Boolean networks, be they binary or multivalued, will require the use of temporal data, and a great deal of it. This should not be surprising given the complexity of the model and the number of parameters, both transitional and static, that must be estimated. In addition to providing an inference algorithm, this paper demonstrates that the data requirement is much smaller if one does not wish to infer the switching, perturbation, and selection probabilities, and that constituent-network connectivity can be discovered with decent accuracy for relatively small time-course sequences.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31]  相似文献   

8.
9.
10.
11.
The purpose of this table is to provide the community with a citable record of publications of ongoing genome sequencing projects that have led to a publication in the scientific literature. While our goal is to make the list complete, there is no guarantee that we may have omitted one or more publications appearing in this time frame. Readers and authors who wish to have publications added to subsequent versions of this list are invited to provide the bibliographic data for such references to the SIGS editorial office.

Non-Bacterial genomes

  相似文献   

12.
Cyclopropane fatty acids (CPAs) are desirable as renewable chemical feedstocks for the production of paints, plastics, and lubricants. Toward our goal of creating a CPA-accumulating crop, we expressed nine higher plant cyclopropane synthase (CPS) enzymes in the seeds of fad2fae1 Arabidopsis (Arabidopsis thaliana) and observed accumulation of less than 1% CPA. Surprisingly, expression of the Escherichia coli CPS gene resulted in the accumulation of up to 9.1% CPA in the seed. Coexpression of a Sterculia foetida lysophosphatidic acid acyltransferase (SfLPAT) increases CPA accumulation up to 35% in individual T1 seeds. However, seeds with more than 9% CPA exhibit wrinkled seed morphology and reduced size and oil accumulation. Seeds with more than 11% CPA exhibit strongly decreased seed germination and establishment, and no seeds with CPA more than 15% germinated. That previous reports suggest that plant CPS prefers the stereospecific numbering (sn)-1 position whereas E. coli CPS acts on sn-2 of phospholipids prompted us to investigate the preferred positions of CPS on phosphatidylcholine (PC) and triacylglycerol. Unexpectedly, in planta, E. coli CPS acts primarily on the sn-1 position of PC; coexpression of SfLPAT results in the incorporation of CPA at the sn-2 position of lysophosphatidic acid. This enables a cycle that enriches CPA at both sn-1 and sn-2 positions of PC and results in increased accumulation of CPA. These data provide proof of principle that CPA can accumulate to high levels in transgenic seeds and sets the stage for the identification of factors that will facilitate the movement of CPA from PC into triacylglycerol to produce viable seeds with additional CPA accumulation.Modified fatty acids (mFAs; sometimes referred to as unusual fatty acids) obtained from plants play important roles in industrial applications as lubricants, protective coatings, plastics, inks, cosmetics, etc. The hundreds of potential industrial uses of mFAs have led to considerable interest in exploring their production in transgenic crop plants. mFAs are produced by a limited number of species, and the transfer of genes encoding mFA-producing enzymes from source plants to heterologous hosts has generally resulted in only modest accumulation, usually less than 20% of the desired mFA in transgenic seed (Napier, 2007) compared with levels found in the natural source. For example, ricinoleic acid accounts for more than 90% of the fatty acid of castor bean (Ricinus communis) seeds, and tung (Aleuites fordii) seeds accumulate more than 80% α-eleostearic acid (Thelen and Ohlrogge, 2002; Drexler et al., 2003). In order to elevate the content of mFAs in the engineered plants to that found in the native plant, it is necessary to (1) optimize the synthesis of mFA (Mekhedov et al., 2001), (2) minimize its degradation (Eccleston and Ohlrogge, 1998), and (3) optimize its incorporation into triacylglycerol (TAG; Bafor et al., 1990; Bates and Browse, 2011; van Erp et al., 2011).Cyclic fatty acids (CFAs) are desirable for numerous industrial applications. The strained bond angles of the carbocyclic ring contribute to their unique chemistry and physical properties, and hydrogenation of CFAs results in ring opening to produce methyl-branched fatty acids. Branched chain fatty acids are ideally suited for the oleochemical industry as feedstocks for the production of lubricants, plastics, paints, dyes, and coatings (Carlsson et al., 2011). Cyclopropane fatty acids (CPAs) have been found in certain gymnosperms, Malvales, Litchi spp., and other Sapindales species. They accumulate to as much as 40% in seeds of Litchi chinensis (Vickery, 1980; Gaydou et al., 1993). Sterculia foetida accumulates the desaturated CFA (i.e. cyclopropene fatty acid) to more than 60% of its seed oil (Bohannon and Kleiman, 1978; Pasha and Ahmad, 1992). The first step in its synthesis is the formation of the CPA by the cyclopropane synthase (CPS) enzyme, which transfers a methyl group to C9 of the oleoyl-phospholipid followed by cyclization to form the cyclopropane ring (Grogan and Cronan, 1997; Bao et al., 2002, 2003). None of the known natural sources of CPA are suitable for its commercial production. Therefore, it would be desirable to create an oilseed crop plant that accumulates high levels of CPA by heterologously expressing CPS in seeds. However, to date, heterologous expression of plant cyclopropane synthase genes has led to only approximately 1.0% CPA in the transgenic seeds (Yu et al., 2011).Two pathways for the biosynthesis of TAG exist in plants (Bates and Browse, 2012; Fig. 1). The de novo biosynthesis from glycerol-3-phosphate and acyl-CoA occurs via the Kennedy pathway and includes three acyltransferases: glycerol-2-phosphate acyltransferase, acyl-CoA:lysophosphatidic acid acyltransferase (LPAT), and acyl-CoA:diacylglycerol acyltransferase (DGAT; Kennedy, 1961). Alternatively, acyl-CoAs can be redirected from phosphatidylcholine (PC) via the action of a phospholipase C, choline phosphotransferase, phosphatidylcholine:diacylglycerol cholinephosphotransferase (PDCT; Hu et al., 2012; Lu et al., 2009), or phospholipid:diacylglycerol acyltransferase (PDAT; Dahlqvist et al., 2000). An acyl group can be released from PC to generate lysophosphatidylcholine (LPC) by the back reaction of acyl-CoA:LPC acyltransferase (Stymne and Stobart, 1984; Wang et al., 2012) or a phospholipase A/acyl-CoA synthase (Chen et al., 2011).Open in a separate windowFigure 1.Schematic representation of the plant TAG biosynthesis network. Acyl editing can provide PC-modified fatty acids for de novo diacylglycerol/TAG synthesis. ACS, acyl-CoA synthase; CPT, CDP-choline:diacylglycerol choline phosphotransferase; G3P, glycerol-3-phosphate; GPAT, acyl-CoA:glycerol-3-phosphate acyltransferase; LPC acyltransferase, acyl-CoA:LPC acyltransferase; mFAS, modified fatty acid synthase (in this work, mFAS is CPS); PAP, phosphatidic acid phosphatase; PLA, phospholipase A; PLC, phospholipase C.LPAT is a pivotal enzyme controlling the metabolic flow of lysophosphatidic acid (LPA) into different phosphatidic acids (PAs) in diverse tissues. Membrane-associated LPAT activities, identified in bacteria, yeast, plant, and animal cells, catalyze the transfer of acyl groups from acyl-CoA to LPA to synthesize PA. In plants and other organisms, LPAT activities have been identified in the endoplasmic reticulum (Kim et al., 2005), plasma membrane (Bursten et al., 1991), and mitochondria (Zborowski and Wojtczak, 1969). In higher plants, endoplasmic reticulum-localized LPAT plays an essential role transferring fatty acid from CoA esters to the sn-2 position of LPA in the synthesis of PA, a key intermediate in the biosynthesis of membrane phospholipids and storage lipids in developing seeds (Maisonneuve et al., 2010). LPAT from developing seeds of flax (Linum usitatissimum), rape (Brassica napus), and castor bean preferentially incorporate oleoyl-CoA, weakly incorporate cyclopropane acyl-CoA, and were unable to incorporate methyl-branched acyl-CoA when presented with an equimolar mix of these potential substrates (Nlandu Mputu et al., 2009). Thus, LPAT activity from agronomic plants constitutes a potential bottleneck for the incorporation of branched chain acyl-CoA into PA. In this work, we investigate the utility of an LPAT from a cyclopropanoid-syntheizing plant, S. foetida, with respect to its ability to enhance CPA accumulation. In our efforts to enhance CPA accumulation in transgenic plants, we screened CPS genes from diverse sources and identified Escherichia coli CPS (EcCPS) as an effective enzyme for the production of CPA in plants. However, EcCPS is reported to prefer the sn-2 position of E. coli phospholipid (Hildebrand and Law, 1964), and the data presented here show that its expression primarily leads to the accumulation of CPA at the stereospecific numbering (sn)-1 position. Moreover, coexpression of S. foetida lysophosphatidic acid acyltransferase (SfLPAT) results in the incorporation of CPA at the sn-2 position of LPA. Thus, coexpression of EcCPS and SfLPAT enables a cycle that enriches the accumulation of CPA at both sn-1 and sn-2 positions of PC and increases the accumulation of CPA. This work underscores the utility of coexpressing an acyltransferase from mFA-accumulating species with mFA-synthesizing enzymes to help mitigate bottlenecks in mFA TAG synthesis.  相似文献   

13.
The plant root is the first organ to encounter salinity stress, but the effect of salinity on root system architecture (RSA) remains elusive. Both the reduction in main root (MR) elongation and the redistribution of the root mass between MRs and lateral roots (LRs) are likely to play crucial roles in water extraction efficiency and ion exclusion. To establish which RSA parameters are responsive to salt stress, we performed a detailed time course experiment in which Arabidopsis (Arabidopsis thaliana) seedlings were grown on agar plates under different salt stress conditions. We captured RSA dynamics with quadratic growth functions (root-fit) and summarized the salt-induced differences in RSA dynamics in three growth parameters: MR elongation, average LR elongation, and increase in number of LRs. In the ecotype Columbia-0 accession of Arabidopsis, salt stress affected MR elongation more severely than LR elongation and an increase in LRs, leading to a significantly altered RSA. By quantifying RSA dynamics of 31 different Arabidopsis accessions in control and mild salt stress conditions, different strategies for regulation of MR and LR meristems and root branching were revealed. Different RSA strategies partially correlated with natural variation in abscisic acid sensitivity and different Na+/K+ ratios in shoots of seedlings grown under mild salt stress. Applying root-fit to describe the dynamics of RSA allowed us to uncover the natural diversity in root morphology and cluster it into four response types that otherwise would have been overlooked.Salt stress is known to affect plant growth and productivity as a result of its osmotic and ionic stress components. Osmotic stress imposed by salinity is thought to act in the early stages of the response, by reducing cell expansion in growing tissues and causing stomatal closure to minimize water loss. The build-up of ions in photosynthetic tissues leads to toxicity in the later stages of salinity stress and can be reduced by limiting sodium transport into the shoot tissue and compartmentalization of sodium ions into the root stele and vacuoles (Munns and Tester, 2008). The effect of salt stress on plant development was studied in terms of ion accumulation, plant survival, and signaling (Munns et al., 2012; Hasegawa, 2013; Pierik and Testerink, 2014). Most studies focus on traits in the aboveground tissues, because minimizing salt accumulation in leaf tissue is crucial for plant survival and its productivity. This approach has led to the discovery of many genes underlying salinity tolerance (Munns and Tester, 2008; Munns et al., 2012; Hasegawa, 2013; Maathuis, 2014). Another way to estimate salinity stress tolerance is by studying the rate of main root (MR) elongation of seedlings transferred to medium supplemented with high salt concentration. This is how Salt Overly Sensitive mutants were identified, being a classical example of genes involved in salt stress signaling and tolerance (Hasegawa, 2013; Maathuis, 2014). The success of this approach is to be explained by the important role that the root plays in salinity tolerance. Roots not only provide anchorage and ensure water and nutrient uptake, but also act as a sensory system, integrating changes in nutrient availability, water content, and salinity to adjust root morphology to exploit available resources to the maximum capacity (Galvan-Ampudia et al., 2013; Gruber et al., 2013). Understanding the significance of environmental modifications of root system architecture (RSA) for plant productivity is one of the major challenges of modern agriculture (de Dorlodot et al., 2007; Den Herder et al., 2010; Pierik and Testerink, 2014).The RSA of dicotyledonous plants consists of an embryonically derived MR and lateral roots (LRs) that originate from xylem pole pericycle cells of the MR, or from LRs in the case of higher-order LRs. Root growth and branching is mainly guided through the antagonistic action of two plant hormones: auxin and cytokinins (Petricka et al., 2012). Under environmental stress conditions, the synthesis of abscisic acid (ABA), ethylene, and brassinosteroids is known to be induced and to modulate the growth of MRs and LRs (Achard et al., 2006; Osmont et al., 2007; Achard and Genschik, 2009; Duan et al., 2013; Geng et al., 2013). In general, lower concentrations of salt were observed to slightly induce MR and LR elongation, whereas higher concentrations resulted in decreased growth of both MRs and LRs (Wang et al., 2009; Zolla et al., 2010). The reduction of growth is a result of the inhibition of cell cycle progression and a reduction in root apical meristem size (West et al., 2004). However, conflicting results were presented for the effect of salinity on lateral root density (LRD; Wang et al., 2009; Zolla et al., 2010; Galvan-Ampudia and Testerink, 2011). Some studies suggest that mild salinity enhances LR initiation or emergence events, thereby affecting patterning, whereas other studies imply that salinity arrests LR development. The origin of those contradictory observations could be attributable to studying LR initiation and density at single time points, rather than observing the dynamics of LR development, because LR formation changes as a function of root growth rate (De Smet et al., 2012). The dynamics of LR growth and development were characterized previously for the MR region formed before the salt stress exposure, identifying the importance of ABA in early growth arrest of postemerged LRs in response to salt stress (Duan et al., 2013). The effect of salt on LR emergence and initiation was found to differ for MR regions formed prior and subsequent to salinity exposure (Duan et al., 2013), consistent with LR patterning being determined at the root tip (Moreno-Risueno et al., 2010). Yet the effect of salt stress on the reprogramming of the entire RSA on a longer timescale remains elusive.Natural variation in Arabidopsis (Arabidopsis thaliana) is a great source for dissecting the genetic components underlying phenotypic diversity (Trontin et al., 2011; Weigel, 2012). Genes underlying phenotypic plasticity of RSA to environmental stimuli were also found to have high allelic variation leading to differences in root development between different Arabidopsis accessions (Rosas et al., 2013). Supposedly, genes responsible for phenotypic plasticity of the root morphology to different environmental conditions are under strong selection for adaptation to local environments. Various populations of Arabidopsis accessions were used to study natural variation in ion accumulation and salinity tolerance (Rus et al., 2006; Jha et al., 2010; Katori et al., 2010; Roy et al., 2013). In addition, a number of studies focusing on the natural variation in RSA have been published, identifying quantitative trait loci and allelic variation for genes involved in RSA development under control conditions (Mouchel et al., 2004; Meijón et al., 2014) and nutrient-deficient conditions (Chevalier et al., 2003; Gujas et al., 2012; Gifford et al., 2013; Kellermeier et al., 2013; Rosas et al., 2013). Exploring natural variation not only expands the knowledge of genes and molecular mechanisms underlying biological processes, but also provides insight on how plants adapt to challenging environmental conditions (Weigel, 2012) and whether the mechanisms are evolutionarily conserved. The early growth arrest of newly emerged LRs upon exposure to salt stress was observed to be conserved among the most commonly used Arabidopsis accessions Columbia-0 (Col-0), Landsberg erecta, and Wassilewskija (Ws; Duan et al., 2013). By studying salt stress responses of the entire RSA and a wider natural variation in root responses to stress, one could identify new morphological traits that are under environmental selection and possibly contribute to stress tolerance.In this work, we not only identify the RSA components that are responsive to salt stress, but we also describe the natural variation in dynamics of salt-induced changes leading to redistribution of root mass and different root morphology. The growth dynamics of MRs and LRs under different salt stress conditions were described by fitting a set of quadratic growth functions (root-fit) to individual RSA components. Studying salt-induced changes in RSA dynamics of 31 Arabidopsis accessions revealed four major strategies conserved among the accessions. Those four strategies were due to differences in salt stress sensitivity of individual RSA components (i.e. growth rates of MRs and LRs, and increases in the number of emerged LRs). This diversity in root morphology responses caused by salt stress was observed to be partially associated with differences in ABA, but not ethylene sensitivity. In addition, we observed that a number of accessions exhibiting a relatively strong inhibition of LR elongation showed a smaller increase in the Na+/K+ ratio in shoot tissue after exposure to salt stress. Our results imply that different RSA strategies identified in this study reflect diverse adaptations to different soil conditions and thus might contribute to efficient water extraction and ion compartmentalization in their native environments.  相似文献   

14.
15.
Most proteins produced in the endoplasmic reticulum (ER) of eukaryotic cells fold via disulfide formation (oxidative folding). Oxidative folding is catalyzed by protein disulfide isomerase (PDI) and PDI-related ER protein thiol disulfide oxidoreductases (ER oxidoreductases). In yeast and mammals, ER oxidoreductin-1s (Ero1s) supply oxidizing equivalent to the active centers of PDI. In this study, we expressed recombinant soybean Ero1 (GmERO1a) and found that GmERO1a oxidized multiple soybean ER oxidoreductases, in contrast to mammalian Ero1s having a high specificity for PDI. One of these ER oxidoreductases, GmPDIM, associated in vivo and in vitro with GmPDIL-2, was unable to be oxidized by GmERO1a. We therefore pursued the possible cooperative oxidative folding by GmPDIM, GmERO1a, and GmPDIL-2 in vitro and found that GmPDIL-2 synergistically accelerated oxidative refolding. In this process, GmERO1a preferentially oxidized the active center in the a′ domain among the a, a′, and b domains of GmPDIM. A disulfide bond introduced into the active center of the a′ domain of GmPDIM was shown to be transferred to the active center of the a domain of GmPDIM and the a domain of GmPDIM directly oxidized the active centers of both the a or a′ domain of GmPDIL-2. Therefore, we propose that the relay of an oxidizing equivalent from one ER oxidoreductase to another may play an essential role in cooperative oxidative folding by multiple ER oxidoreductases in plants.In eukaryotes, many secretory and membrane proteins fold via disulfide bond formation in the endoplasmic reticulum (ER). Seed storage proteins of major crops, such as wheat, corn, rice, and beans, which are important protein sources for humans and domestic animals, are synthesized in the ER of the endosperm or cotyledon. A number of seed storage proteins fold by the formation of intramolecular disulfide bonds (oxidative folding) and are transported to and accumulate in protein bodies (Kermode and Bewley, 1999; Jolliffe et al., 2005). In contrast to normally folded proteins, misfolded and unfolded proteins are retained in the ER and degraded by an ER-associated degradation or vacuolar system (Smith et al., 2011; Pu and Bassham, 2013). Therefore, quick and efficient oxidative folding of nascent seed storage proteins is needed for their accumulation in protein bodies.During this process, protein disulfide isomerase (PDI; EC 5.3.4.1) and other ER protein thiol disulfide oxidoreductases (ER oxidoreductases) are thought to catalyze the formation and isomerization of disulfide bonds in nascent proteins (Hatahet and Ruddock, 2009; Feige and Hendershot, 2011; Lu and Holmgren, 2014). After phylogenetic analysis of the Arabidopsis genome, 10 classes of ER oxidoreductases (classes I–X) were identified (Houston et al., 2005). Among them, class I ER oxidoreductase, a plant PDI ortholog, has been studied in a wide variety of plants. Class I ER oxidoreductases have two catalytically active domains a and a′, containing active centers composed of Cys-Gly-His-Cys and two catalytically inactive domains b and b′. An Arabidopsis ortholog of class I ER oxidoreductases is required for proper seed development and regulates the timing of programmed cell death by chaperoning and inhibiting Cys proteases (Andème Ondzighi et al., 2008). OaPDI, a PDI from Oldenlandia affinis, a coffee family (Rubiaceae) plant, is involved in the folding of knotted circular proteins (Gruber et al., 2007). The rice ortholog (PDIL1-1) was suggested to be involved in the maturation of the major seed storage protein glutelin (Takemoto et al., 2002). Furthermore, rice PDIL1-1 plays a role in regulatory activities for various proteins that are essential for the synthesis of grain components as determined by analysis of a T-DNA insertion mutant (Satoh-Cruz et al., 2010).The oxidative refolding ability of class I ER oxidoreductases was confirmed in recombinant soybean (GmPDIL-1) and wheat proteins produced by an Escherichia coli expression system established from cDNAs (Kamauchi et al., 2008; Kimura et al., 2015).Class II and III ER oxidoreductases have an a–b–b′–a′ domain structure. Class II ER oxidoreductases have an acidic amino acid-rich sequence in the N-terminal region ahead of the a domain. Recombinant soybean (GmPDIL-2) and wheat class II ER oxidoreductases have oxidative refolding activities similar to that of class I (Kamauchi et al., 2008; Kimura et al., 2015). Class III ER oxidoreductases contain the nonclassical redox-center Cys-X-X-Ser/Cys motifs, as opposed to the more traditional CGHC sequence, in the a and a′ domains. Recombinant soybean (GmPDIL-3) and wheat proteins lack oxidative refolding activity in vitro (Iwasaki et al., 2009; Kimura et al., 2015). Class IV ER oxidoreductases are unique to plants and have an a–a′–ERp29 domain structure, which is homologous to the C-terminal domain of mammalian ERp29 (Demmer et al., 1997).Recombinant soybean class IV ER oxidoreductases (GmPDIS-1 and GmPDIS-2) and wheat class IV ER oxidoreductase possess an oxidative refolding activity that is weaker than that of classes I and II (Wadahama et al., 2007; Kimura et al., 2015). Class V ER oxidoreductases are plant orthologs of mammalian P5 and have an a–a′–b domain structure. A rice class V ER oxidoreductase, consisting of PDIL2 and PDIL3, plays an important role in the accumulation of the seed storage protein Cys-rich 10-kD prolamin (crP10; Onda et al., 2011). Recombinant soybean class V ER oxidoreductase, GmPDIM and wheat class V ER oxidoreductase possess an oxidative refolding activity similar to that of class IV (Wadahama et al., 2008; Kimura et al., 2015). In the soybean, GmPDIL-1, GmPDIL-2, GmPDIM, GmPDIS-1, and GmPDIS-2 were found to associate transiently with a seed storage precursor protein, proglycinin, in the ER of the cotyledon by coimmunoprecipitation experiments, suggesting that multiple ER oxidoreductases are involved in the folding of the nascent proglycinin.The disulfide bond in the active center of ER oxidoreductases is reduced as a result of catalyzing disulfide bond formation in an unfolded protein. The reduced active center of PDI was discovered to be oxidized again by ER oxidoreductin-1 (Ero1p) in yeast (Frand and Kaiser, 1998; Pollard et al., 1998). Ero1p orthologs are present universally in eukaryotes. Yeast and flies have a single copy of the ERO1 gene, which is essential for survival (Frand and Kaiser, 1998; Pollard et al., 1998; Tien et al., 2008). Mammals have two genes encoding Ero1-α (Cabibbo et al., 2000) and Ero1-β (Pagani et al., 2000) that function as major disulfide donors to nascent proteins in the ER, but are not critical for survival (Zito et al., 2010). Domain a of yeast PDI is the most favored substrate of yeast Ero1p (Vitu et al., 2010), whereas a′ of human PDI is specifically oxidized by human Ero1-α (Chambers et al., 2010) and Ero1-β (Wang et al., 2011). Electrons from Cys residues of the active centers of PDI are transferred to oxygen by Ero1 (Tu and Weissman, 2004; Sevier and Kaiser, 2008). The reaction mechanisms of yeast Ero1p and human Ero1s have been intensively investigated; their regulation by PDI has been extensively studied as well (Tavender and Bulleid, 2010; Araki and Inaba, 2012; Benham et al., 2013; Ramming et al., 2015). Only rice Ero1 (OsERO1) has been identified as a plant ortholog of Ero1p (Onda et al., 2009). OsERO1 is necessary for disulfide bond formation in rice endosperm. The formation of native disulfide bonds in the major seed storage protein proglutelin was demonstrated to depend upon OsERO1 by RNAi knockdown experiments. However, no plant protein thiol disulfide oxidoreductases that are oxidized by a plant Ero1 ortholog have been identified to date.In this study, we show that multiple soybean ER oxidoreductases can be activated by a soybean Ero1 ortholog (GmERO1a). In addition, we propose a synergistic mechanism by which GmPDIM and GmPDIL-2 cooperatively fold unfolded proteins using oxidizing equivalents provided by GmERO1 in vitro.  相似文献   

16.
17.
The essential cell division protein FtsL is a substrate of the intramembrane protease RasP. Using heterologous coexpression experiments, we show here that the division protein DivIC stabilizes FtsL against RasP cleavage. Degradation seems to be initiated upon accessibility of a cytosolic substrate recognition motif.Cell division in bacteria is a highly regulated process (1). The division site selection as well as assembly and disassembly of the divisome have to be strictly controlled (1, 4). Although the spatial control of the divisome is relatively well understood (2, 4, 14, 17), mechanisms governing the temporal control of division are still mainly elusive. Regulatory proteolysis was thought to be a potential modulatory mechanism (8, 9). The highly unstable division protein FtsL was shown to be rate limiting for division and would make an ideal candidate for a regulatory factor in the timing of bacterial cell division (7, 9). In Bacillus subtilis, FtsL is an essential protein of the membrane part of the divisome (5, 7, 8). It is necessary for the assembly of the membrane-spanning division proteins, and a knockout is lethal (8, 9, 12). We have previously reported that FtsL is a substrate of the intramembrane protease RasP (5).These findings raised the question of whether RasP can regulate cell division by cleaving FtsL from the division complex. In order to mimic the situation in which FtsL is bound to at least one of its interaction partners, we used a heterologous coexpression system in which we synthesized FtsL and DivIC. It has been reported before that DivIC and FtsL are intimate binding partners in various organisms (6, 9, 15, 21, 22, 26) and that FtsL and DivIC (together with DivIB) can form complexes even in the absence of the other divisome components (6, 21). We therefore asked whether RasP is able to cleave FtsL in the presence of its major interaction partner DivIC, which would argue for the possibility that RasP could cleave FtsL within a mature divisome. In contrast, if interaction with DivIC could stabilize FtsL against RasP cleavage, this result would bring such a model into question. An alternative option for the role of RasP might be the removal of FtsL from the membrane. It has been shown that divisome disassembly and prevention of reassembly are crucial to prevent minicell formation close to the new cell poles (3, 16).  相似文献   

18.
Neuropeptides induce signal transduction across the plasma membrane by acting through cell-surface receptors. The dynorphins, endogenous ligands for opioid receptors, are an exception; they also produce non-receptor-mediated effects causing pain and neurodegeneration. To understand non-receptor mechanism(s), we examined interactions of dynorphins with plasma membrane. Using fluorescence correlation spectroscopy and patch-clamp electrophysiology, we demonstrate that dynorphins accumulate in the membrane and induce a continuum of transient increases in ionic conductance. This phenomenon is consistent with stochastic formation of giant (~2.7 nm estimated diameter) unstructured non-ion-selective membrane pores. The potency of dynorphins to porate the plasma membrane correlates with their pathogenic effects in cellular and animal models. Membrane poration by dynorphins may represent a mechanism of pathological signal transduction. Persistent neuronal excitation by this mechanism may lead to profound neuropathological alterations, including neurodegeneration and cell death.Neuropeptides are the largest and most diverse family of neurotransmitters. They are released from axon terminals and dendrites, diffuse to pre- or postsynaptic neuronal structures and activate membrane G-protein-coupled receptors. Prodynorphin (PDYN)-derived opioid peptides including dynorphin A (Dyn A), dynorphin B (Dyn B) and big dynorphin (Big Dyn) consisting of Dyn A and Dyn B are endogenous ligands for the κ-opioid receptor. Acting through this receptor, dynorphins regulate processing of pain and emotions, memory acquisition and modulate reward induced by addictive substances.1, 2, 3, 4 Furthermore, dynorphins may produce robust cellular and behavioral effects that are not mediated through opioid receptors.5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29 As evident from pharmacological, morphological, genetic and human neuropathological studies, these effects are generally pathological, including cell death, neurodegeneration, neurological dysfunctions and chronic pain. Big Dyn is the most active pathogenic peptide, which is about 10- to 100-fold more potent than Dyn A, whereas Dyn B does not produce non-opioid effects.16, 17, 22, 25 Big Dyn enhances activity of acid-sensing ion channel-1a (ASIC1a) and potentiates ASIC1a-mediated cell death in nanomolar concentrations30, 31 and, when administered intrathecally, induces characteristic nociceptive behavior at femtomolar doses.17, 22 Inhibition of endogenous Big Dyn degradation results in pathological pain, whereas prodynorphin (Pdyn) knockout mice do not maintain neuropathic pain.22, 32 Big Dyn differs from its constituents Dyn A and Dyn B in its unique pattern of non-opioid memory-enhancing, locomotor- and anxiolytic-like effects.25Pathological role of dynorphins is emphasized by the identification of PDYN missense mutations that cause profound neurodegeneration in the human brain underlying the SCA23 (spinocerebellar ataxia type 23), a very rare dominantly inherited neurodegenerative disorder.27, 33 Most PDYN mutations are located in the Big Dyn domain, demonstrating its critical role in neurodegeneration. PDYN mutations result in marked elevation in dynorphin levels and increase in its pathogenic non-opioid activity.27, 34 Dominant-negative pathogenic effects of dynorphins are not produced through opioid receptors.ASIC1a, glutamate NMDA (N-methyl-d-aspartate) and AMPA (α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid)/kainate ion channels, and melanocortin and bradykinin B2 receptors have all been implicated as non-opioid dynorphin targets.5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 30, 31, 35, 36 Multiplicity of these targets and their association with the cellular membrane suggest that their activation is a secondary event triggered by a primary interaction of dynorphins with the membrane. Dynorphins are among the most basic neuropeptides.37, 38 The basic nature is also a general property of anti-microbial peptides (AMPs) and amyloid peptides that act by inducing membrane perturbations, altering membrane curvature and causing pore formation that disrupts membrane-associated processes including ion fluxes across the membrane.39 The similarity between dynorphins and these two peptide groups in overall charge and size suggests a similar mode of their interactions with membranes.In this study, we dissect the interactions of dynorphins with the cell membrane, the primary event in their non-receptor actions. Using fluorescence imaging, correlation spectroscopy and patch-clamp techniques, we demonstrate that dynorphin peptides accumulate in the plasma membrane in live cells and cause a profound transient increase in cell membrane conductance. Membrane poration by endogenous neuropeptides may represent a novel mechanism of signal transduction in the brain. This mechanism may underlie effects of dynorphins under pathological conditions including chronic pain and tissue injury.  相似文献   

19.
Immunogold localization revealed that OmcS, a cytochrome that is required for Fe(III) oxide reduction by Geobacter sulfurreducens, was localized along the pili. The apparent spacing between OmcS molecules suggests that OmcS facilitates electron transfer from pili to Fe(III) oxides rather than promoting electron conduction along the length of the pili.There are multiple competing/complementary models for extracellular electron transfer in Fe(III)- and electrode-reducing microorganisms (8, 18, 20, 44). Which mechanisms prevail in different microorganisms or environmental conditions may greatly influence which microorganisms compete most successfully in sedimentary environments or on the surfaces of electrodes and can impact practical decisions on the best strategies to promote Fe(III) reduction for bioremediation applications (18, 19) or to enhance the power output of microbial fuel cells (18, 21).The three most commonly considered mechanisms for electron transfer to extracellular electron acceptors are (i) direct contact between redox-active proteins on the outer surfaces of the cells and the electron acceptor, (ii) electron transfer via soluble electron shuttling molecules, and (iii) the conduction of electrons along pili or other filamentous structures. Evidence for the first mechanism includes the necessity for direct cell-Fe(III) oxide contact in Geobacter species (34) and the finding that intensively studied Fe(III)- and electrode-reducing microorganisms, such as Geobacter sulfurreducens and Shewanella oneidensis MR-1, display redox-active proteins on their outer cell surfaces that could have access to extracellular electron acceptors (1, 2, 12, 15, 27, 28, 31-33). Deletion of the genes for these proteins often inhibits Fe(III) reduction (1, 4, 7, 15, 17, 28, 40) and electron transfer to electrodes (5, 7, 11, 33). In some instances, these proteins have been purified and shown to have the capacity to reduce Fe(III) and other potential electron acceptors in vitro (10, 13, 29, 38, 42, 43, 48, 49).Evidence for the second mechanism includes the ability of some microorganisms to reduce Fe(III) that they cannot directly contact, which can be associated with the accumulation of soluble substances that can promote electron shuttling (17, 22, 26, 35, 36, 47). In microbial fuel cell studies, an abundance of planktonic cells and/or the loss of current-producing capacity when the medium is replaced is consistent with the presence of an electron shuttle (3, 14, 26). Furthermore, a soluble electron shuttle is the most likely explanation for the electrochemical signatures of some microorganisms growing on an electrode surface (26, 46).Evidence for the third mechanism is more circumstantial (19). Filaments that have conductive properties have been identified in Shewanella (7) and Geobacter (41) species. To date, conductance has been measured only across the diameter of the filaments, not along the length. The evidence that the conductive filaments were involved in extracellular electron transfer in Shewanella was the finding that deletion of the genes for the c-type cytochromes OmcA and MtrC, which are necessary for extracellular electron transfer, resulted in nonconductive filaments, suggesting that the cytochromes were associated with the filaments (7). However, subsequent studies specifically designed to localize these cytochromes revealed that, although the cytochromes were extracellular, they were attached to the cells or in the exopolymeric matrix and not aligned along the pili (24, 25, 30, 40, 43). Subsequent reviews of electron transfer to Fe(III) in Shewanella oneidensis (44, 45) appear to have dropped the nanowire concept and focused on the first and second mechanisms.Geobacter sulfurreducens has a number of c-type cytochromes (15, 28) and multicopper proteins (12, 27) that have been demonstrated or proposed to be on the outer cell surface and are essential for extracellular electron transfer. Immunolocalization and proteolysis studies demonstrated that the cytochrome OmcB, which is essential for optimal Fe(III) reduction (15) and highly expressed during growth on electrodes (33), is embedded in the outer membrane (39), whereas the multicopper protein OmpB, which is also required for Fe(III) oxide reduction (27), is exposed on the outer cell surface (39).OmcS is one of the most abundant cytochromes that can readily be sheared from the outer surfaces of G. sulfurreducens cells (28). It is essential for the reduction of Fe(III) oxide (28) and for electron transfer to electrodes under some conditions (11). Therefore, the localization of this important protein was further investigated.  相似文献   

20.
The multifunctional movement protein (MP) of Tomato mosaic tobamovirus (ToMV) is involved in viral cell-to-cell movement, symptom development, and resistance gene recognition. However, it remains to be elucidated how ToMV MP plays such diverse roles in plants. Here, we show that ToMV MP interacts with the Rubisco small subunit (RbCS) of Nicotiana benthamiana in vitro and in vivo. In susceptible N. benthamiana plants, silencing of NbRbCS enabled ToMV to induce necrosis in inoculated leaves, thus enhancing virus local infectivity. However, the development of systemic viral symptoms was delayed. In transgenic N. benthamiana plants harboring Tobacco mosaic virus resistance-22 (Tm-22), which mediates extreme resistance to ToMV, silencing of NbRbCS compromised Tm-22-dependent resistance. ToMV was able to establish efficient local infection but was not able to move systemically. These findings suggest that NbRbCS plays a vital role in tobamovirus movement and plant antiviral defenses.Plant viruses use at least one movement protein (MP) to facilitate viral spread between plant cells via plasmodesmata (PD; Lucas and Gilbertson, 1994; Ghoshroy et al., 1997). Among viral MPs, the MP of tobamoviruses, such as Tobacco mosaic virus (TMV) and its close relative Tomato mosaic virus (ToMV), is the best characterized. TMV MP specifically accumulates in PD and modifies the plasmodesmatal size exclusion limit in mature source leaves or tissues (Wolf et al., 1989; Deom et al., 1990; Ding et al., 1992). TMV MP and viral genomic RNA form a mobile ribonucleoprotein complex that is essential for cell-to-cell movement of viral infection (Watanabe et al., 1984; Deom et al., 1987; Citovsky et al., 1990, 1992; Kiselyova et al., 2001; Kawakami et al., 2004; Waigmann et al., 2007). TMV MP also enhances intercellular RNA silencing (Vogler et al., 2008) and affects viral symptom development, host range, and host susceptibility to virus (Dardick et al., 2000; Bazzini et al., 2007). Furthermore, ToMV MP is identified as an avirulence factor that is recognized by tomato (Solanum lycopersicum) resistance proteins Tobacco mosaic virus resistance-2 (Tm-2) and Tm-22 (Meshi et al., 1989; Lanfermeijer et al., 2004). Indeed, tomato Tm-22 confers extreme resistance against TMV and ToMV in tomato plants and even in heterologous tobacco (Nicotiana tabacum) plants (Lanfermeijer et al., 2003, 2004).To date, several host factors that interact with TMV MP have been identified. These TMV MP-binding host factors include cell wall-associated proteins such as pectin methylesterase (Chen et al., 2000), calreticulin (Meshi et al., 1989), ANK1 (Ueki et al., 2010), and the cellular DnaJ-like protein MPIP1 (Shimizu et al., 2009). Many cytoskeletal components such as actin filaments (McLean et al., 1995), microtubules (Heinlein et al., 1995), and the microtubule-associated proteins MPB2C (Kragler et al., 2003) and EB1a (Brandner et al., 2008) also interact with TMV MP. Most of these factors are involved in TMV cell-to-cell movement.Rubisco catalyzes the first step of CO2 assimilation in photosynthesis and photorespiration. The Rubisco holoenzyme is a heteropolymer consisting of eight large subunits (RbCLs) and eight small subunits (RbCSs). RbCL was reported to interact with the coat protein of Potato virus Y (Feki et al., 2005). Both RbCS and RbCL were reported to interact with the P3 proteins encoded by several potyviruses, including Shallot yellow stripe virus, Onion yellow dwarf virus, Soybean mosaic virus, and Turnip mosaic virus (Lin et al., 2011). Proteomic analysis of the plant-virus interactome revealed that RbCS participates in the formation of virus complexes of Rice yellow mottle virus (Brizard et al., 2006). However, the biological function of Rubisco in viral infection remains unknown.In this study, we show that RbCS plays an essential role in virus movement, host susceptibility, and Tm-22-mediated extreme resistance in the ToMV-host plant interaction.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

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