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 62 毫秒
1.
2.
The Bamboo mosaic virus (BaMV) is a positive-sense, single-stranded RNA virus. Previously, we identified that the chloroplast phosphoglycerate kinase (chl-PGK) from Nicotiana benthamiana is one of the viral RNA binding proteins involved in the BaMV infection cycle. Because chl-PGK is transported to the chloroplast, we hypothesized that chl-PGK might be involved in viral RNA localization in the chloroplasts. To test this hypothesis, we constructed two green fluorescent protein (GFP)-fused mislocalized PGK mutants, the transit peptide deletion mutant (NO TRANSIT PEPTIDE [NOTP]-PGK-GFP) and the nucleus location mutant (nuclear location signal [NLS]-PGK-GFP). Using confocal microscopy, we demonstrated that NOTP-PGK-GFP and NLS-PGK-GFP are localized in the cytoplasm and nucleus, respectively, in N. benthamiana plants. When NOTP-PGK-GFP and NLS-PGK-GFP are transiently expressed, we observed a reduction in BaMV coat protein accumulation to 47% and 27% that of the wild-type PGK-GFP, respectively. To localize viral RNA in infected cells, we employed the interaction of NLS-GFP-MS2 (phage MS2 coat protein) with the modified BaMV RNA containing the MS2 coat protein binding sequence. Using confocal microscopy, we observed that BaMV viral RNA localizes to chloroplasts. Furthermore, elongation factor1a fused with the transit peptide derived from chl-PGK or with a Rubisco small subunit can partially restore BaMV accumulation in NbPGK1-knockdown plants by helping BaMV target chloroplasts.Bamboo mosaic virus (BaMV) is a single-stranded, positive-sense RNA virus. The genomic RNA of BaMV contains five open reading frames (ORFs) and is 6,366 nucleotides in length with a 5′ cap and a 3′ poly(A) tail (Lin et al., 1994; Yang et al., 1997). ORF1 encodes a 155-kD replicase comprised of a capping enzyme domain that exhibits S-adenosylmethionine-dependent guanylyltransferase activity (Li et al., 2001a; Huang et al., 2004), a helicase-like domain with RNA 5′-triphosphatase activity (Li et al., 2001b), and an RNA-dependent RNA polymerase domain (Li et al., 1998; Cheng et al., 2001). The three overlapping ORFs (i.e. ORF2, ORF3, and ORF4) are known as the triple gene block. They encode for proteins involved in viral movement (Lin et al., 2004, 2006; Vijaya Palani et al., 2006). ORF5 encodes the viral capsid protein (CP), required for virion assembly and viral movement (Cruz et al., 1998).The genomes of positive-strand RNA viruses are templates for both translation and replication. Viral replication complexes are likely to be assembled using host factors to synthesize the minus-strand RNA and then the plus-strand progeny RNA. Recent studies have shown that host factors play important roles in assembling the viral RNA replication complex, selecting and recruiting viral replication templates, activating the complex for RNA synthesis, and other steps (Ahlquist et al., 2003; Patarroyo et al., 2012). The translation and the minus-strand RNA synthesis of poliovirus are regulated by host poly(C) and poly(A) binding proteins and viral polymerase precursor 3CD (Waggoner and Sarnow, 1998; Herold and Andino, 2001; Walter et al., 2002). A number of host genes required for Brome mosaic virus replication have been identified systemically by the yeast (Saccharomyces cerevisiae) genetic approach (Ishikawa et al., 1997; Kushner et al., 2003; Mas et al., 2006; Gancarz et al., 2011). The same approach was used to identify the host factors involved in the replication of Tomato bushy stunt virus (TBSV; Panavas et al., 2005; Li et al., 2009b). A heat shock protein90 homolog (Huang et al., 2012) and the Nicotiana benthamiana glutathione transferase U4 (NbGSTU4; Chen et al., 2013), were identified to interact with the 3′ untranslated region (UTR) of BaMV RNA and enhanced the minus-strand RNA synthesis at the early replication step. The Ser/Thr kinase-like protein localized on cell membrane facilitates the BaMV intercellular movement (Cheng et al., 2013).Previously, we have identified two host proteins (i.e. p51 and p43) interacting specifically with the 3′ UTRs of BaMV by using electrophoretic mobility shift assay (EMSA) and the UV cross-linking competition technique. The results of liquid chromatography-tandem mass spectrometry (LC-MS/MS) and BLAST indicate that the protein sequences of p43 and p51 match the chloroplast phosphoglycerate kinase (chl-PGK) and elongation factor1a (EF1a) of Nicotiana benthamiana, respectively (Lin et al., 2007). Phosphoglycerate kinase is an ATP-generating enzyme that acts in the glycolytic, gluconeogenic, and photosynthetic pathways (Banks et al., 1979; McHarg et al., 1999). chl-PGK is encoded in the nucleus and translated to produce a 50-kD precursor protein and is then processed into mature 43 kD in the chloroplast. In a knockdown experiment through virus-induced gene silencing, the reduction of PGK decreased the accumulation of BaMV coat protein (Lin et al., 2007).Eukaryotic EF1a has been shown to play a role in binding to the tRNA-like structure and upstream pseudoknot in the 3′ UTR of Tobacco mosaic virus to regulate the gene expression and viral replication (Pathak et al., 2008). EF1a has also been involved in the recruitment of viral RNA and has facilitated the replicase complex assembly of TBSV (Pogany et al., 2008). The 3′ UTR of BaMV cannot only bind its replicase but also the EF1a and has been proposed to regulate viral RNA replication (Lin et al., 2007).In this study, we transiently expressed two mislocalized PGK mutants to study the possible functions of chl-PGK that is involved in viral RNA replication. In addition, we used confocal microscopy to investigate the localization of BaMV RNA. Finally, we provided evidence that the down-regulation of BaMV accumulation in PGK-knockdown plants can be restored by the expression of the BaMV RNA binding protein EF1a that is fused to a chloroplast transit peptide.  相似文献   

3.
4.
5.
6.
7.
Diacylglycerol (DAG) is an intermediate in metabolism of both triacylglycerols and membrane lipids. Probing the steady-state pools of DAG and understanding how they contribute to the synthesis of different lipids is important when designing plants with altered lipid metabolism. However, traditional methods of assaying DAG pools are difficult, because its abundance is low and because fractionation of subcellular membranes affects DAG pools. To manipulate and probe DAG pools in an in vivo context, we generated multiple stable transgenic lines of Arabidopsis (Arabidopsis thaliana) that target an Escherichia coli DAG kinase (DAGK) to each leaflet of each chloroplast envelope membrane. E. coli DAGK is small, self inserts into membranes, and has catalytic activity on only one side of each membrane. By comparing whole-tissue lipid profiles between our lines, we show that each line has an individual pattern of DAG, phosphatidic acid, phosphatidylcholine, and triacylglycerol steady-state levels, which supports an individual function of DAG in each membrane leaflet. Furthermore, conversion of DAG in the leaflets facing the chloroplast intermembrane space by DAGK impairs plant growth. As a result of DAGK presence in the outer leaflet of the outer envelope membrane, phosphatidic acid accumulation is not observed, likely because it is either converted into other lipids or removed to other membranes. Finally, we use the outer envelope-targeted DAGK line as a tool to probe the accessibility of DAG generated in response to osmotic stress.Diacylglycerol (DAG) is a central metabolite in plant lipid metabolism. Its glycerol backbone is modified with two acyl chains. If a third acyl chain is added, triacylglycerol (TAG) is formed, whereas if a head group is added, it is converted into polar lipids such as a galactolipid. In green tissues, the majority of DAG is used as an intermediate in galactolipid synthesis, because the extensive thylakoid membrane system consists of approximately 85% galactolipids (Block et al., 1983). Although under normal conditions the galactolipids are exclusively chloroplastic, in Arabidopsis (Arabidopsis thaliana), the DAG used to make galactolipids is derived from assembly pathways in both the chloroplast and the endoplasmic reticulum (ER; Benning, 2009). In both pathways, the bulk of the fatty acids are synthesized in the chloroplast stroma (Browse et al., 1986) in the following order of abundance: 18:1, 16:0, and 18:0 (Wallis and Browse, 2002).In the chloroplast pathway, these fatty acids are directly attached to a glycerol-3-P, generating first lyso-phosphatidic acid (l-PtdOH) and then phosphatidic acid (PtdOH) in the inner leaflet of the chloroplast inner envelope (Fig. 1; Frentzen et al., 1983). The acyltransferases involved are specific to the extent that the sn-2 position of the glycerol backbone predominantly receives a 16:0 fatty acid. PtdOH is then used directly for phosphatidylglycerol (PtdGro) synthesis (Babiychuk et al., 2003) or converted to DAG by a PtdOH phosphatase (Joyard and Douce, 1977). The PtdOH phosphatase activity is known to be associated with the inner envelope, though which leaflet is obscured by the fact that DAG can efficiently flip across membranes (Hamilton et al., 1991) and the actual enzyme has not been unambiguously identified and located (Nakamura et al., 2007). However, the leaflet associations of two other enzymes that use DAG in the inner envelope have been established. MGD1, which uses DAG to synthesize the most abundant galactolipid, monogalactosyldiacylglycerol (MGDG), is on the outer leaflet of the inner envelope membrane (Xu et al., 2005), while SQD2, which uses DAG to generate the less abundant sulfolipid, sulfoquinovosyldiacylglycerol (SQDG), is located on the inner leaflet of the inner envelope membrane (Tietje and Heinz, 1998). Also associated with the inner envelope membrane are a number of fatty acid desaturases, including FAD4, FAD5, FAD6, FAD7, and FAD8 (Joyard et al., 2010). Two of these are specific, generating lipids with signature desaturations: FAD4 desaturates only the 16:0 fatty acid of PtdGro, giving plastidic PtdGro a distinct 16:1 Δ3 trans moiety (Browse et al., 1985; Gao et al., 2009), and FAD5 desaturates primarily the 16:0 fatty acid of MGDG, producing 16:1 Δ7 cis (Kunst et al., 1989). The remaining desaturases are less specific, with little preference for head group or acyl tail. They further desaturate 16:1 or 18:1 in the cis conformation to 16:2 or 18:2 (FAD6; Browse et al., 1989) and on to 16:3 or 18:3 (FAD7 and FAD8; Wallis and Browse, 2002). The combined actions of these FADs result in the highly desaturated fatty acid profiles seen for most chloroplast lipids.Open in a separate windowFigure 1.Overview of DAG pools in the chloroplast envelope membranes. Processes that are known to have activity feeding into or withdrawing from DAG pools in the chloroplast envelope membranes are shown. Enzymes are indicated, and their substrates and products are connected with black arrows. However, for space reasons, not all reactants are shown. Membrane leaflets are indicated, and enzymes with known membrane topology are displayed correctly, while those without known topology are displayed in the center of the appropriate membrane. The acyl group preferred by each l-PtdOH acyltransferase is given in parentheses. Proposed processes transporting lipids from the ER to the chloroplast are shown with dashed arrows. Enzymes are as follows: 1, ATS1; 2, ATS2; 3, lipid phosphate phosphatase γ; 4, MGD1; 5, SQD2; 6, cytosolic phospholipases; 7, MGD2 or MGD3; 8, SFR2; 9, acyl-CoA:glycerol-3-P acyltransferase; 10, l-PtdOH acyltransferase; 11, PtdOH phosphatase; 12, cytidine diphosphate-choline:DAG cholinephosphotransferase; 13, TGD4; and 14, TGD1, TGD 2, TGD3 lipid transport complex. OE, Chloroplast outer envelope membrane; IE, chloroplast inner envelope membrane; ACP, acyl carrier protein. [See online article for color version of this figure.]In unstressed plants, DAG seems to be used primarily in the inner chloroplast envelope. However, several conditions are known to cause extensive DAG use in the chloroplast outer envelope. During phosphate deprivation, MGD2 and MDG3 synthesize MGDG from DAG on the outer leaflet of the outer envelope membrane (Kobayashi et al., 2009). The DAG backbones are probably supplied from the phosphatidylcholine (PtdCho) pool by phospholipase activity, which was shown to be simultaneously up-regulated (Andersson et al., 2004; Nakamura et al., 2005). DAG is also generated during freezing stress by a galactolipid:galactolipid galactosyltransferase named Sensitive to FReezing2 (SFR2). This enzyme transfers the galactosyl head group of MGDG onto another MGDG, giving rise to digalactosyldiacylglycerol (DGDG) and DAG (Moellering et al., 2010). The DAG is subsequently sequestered into a lipid droplet by formation of TAG by an as yet unidentified enzyme.In the ER pathway, fatty acids synthesized in the chloroplast stroma are exported through a still poorly defined mechanism to the ER and activated to acyl-CoAs. Acyltransferases sequentially catalyze formation of l-PtdOH and PtdOH from glycerol-3-P and acyl-CoAs. Again, the acyltransferase working on the sn-2 position of the glycerol backbone is specific, but unlike the chloroplast isoform, it prefers an 18:1 carbon fatty acid (Frentzen et al., 1983). Newly generated PtdOH can be converted to PtdGro or phosphatidyl inositol (PtdIns) (Collin et al., 1999) or hydrolyzed to DAG (Shimojima et al., 2009). DAG can then be further metabolized to TAG and PtdCho. PtdCho acyl groups (18:1/18:1 and 18:1/16:0) are desaturated sequentially by desaturases FAD2 (Okuley et al., 1994) and FAD3 (Browse et al., 1993). These desaturases prefer PtdCho as substrate. The acyl chains modified on PtdCho are transferred to other ER lipids, including DAG, as a result of continual acyl editing of the PtdCho pool (Bates et al., 2012). Furthermore, PtdOH and many of the other extraplastidic phospholipids can be converted to DAG by action of phospholipases (Shimojima et al., 2009). These have as yet partially defined roles in response to stress or recycling of membrane lipids (Testerink and Munnik, 2005).Glycerolipid precursors generated by de novo synthesis, acyl editing, and possibly stress conditions in the ER are transported to the chloroplast by a mechanism that is likely to involve at least two putative lipid transporters: trigalactosyldiacylglycerol4 (TGD4) in the chloroplast outer envelope membrane and the TGD1, TGD2, and TGD3 complex in the inner envelope membrane (Wang and Benning, 2012). The actual lipid species transported remains unclear, but PtdCho, lyso-phosphatidylcholine, PtdOH, and DAG have been discussed in the literature (Andersson and Dörmann, 2009). The DAG moieties are then fully incorporated into all plastidic lipids except PtdGro, presumably using the same pathways that metabolize plastidic DAG, described above. Because of the preference of chloroplast and ER sn-2 acyltransferases for 16 or 18 carbon fatty acids, respectively, the origin of the DAG moieties can be distinguished by positional analysis of the acyl groups on the glycerol backbone (Roughan and Slack, 1982). In Arabidopsis, the chloroplast and ER lipid synthesis pathways contribute nearly equally to mature chloroplast lipids (Browse et al., 1986; Mongrand et al., 1998). Thus, the DAG pools described so far in the chloroplast inner and outer envelope membranes are each of dual origin.A challenge for the analysis of the different DAG pools is that this compound is not a bilayer-forming lipid and thus does not accumulate stably to high levels. Furthermore, during any lengthy fractionation procedure, its levels can be expected to alter, as DAG-modifying enzymes exist in multiple membranes. Moreover, because DAG is quickly metabolized and may have efficient transport systems (Dong et al., 2012), it is difficult to confirm whether metabolizing enzymes are accessing the same or separate DAG pools.To probe different DAG pools of chloroplast membranes in vivo, we have generated a series of stable transgenic Arabidopsis lines in which we target an Escherichia coli DAG kinase (DAGK) to each leaflet of the chloroplast envelope membranes. The basic utility of this approach was previously shown by targeting a DAGK to the chloroplast in tobacco (Nicotiana tabacum) using a single construct fusing the bacterial protein to the Rubisco small subunit N-terminal peptide (Fritz et al., 2007). Here, we present a full phenotypic analysis of these lines, determining which chloroplast membranes show steady-state alterations of DAG and PtdOH levels predicted by ectopic DAGK activity. We further determine the accessibility of DAG pools generated on the outer leaflet of the chloroplast outer envelope membrane during osmotic stress. Having this system established in Arabidopsis will allow characterization of DAG pools in multiple lipid mutant lines.  相似文献   

8.
Chloroplast division is performed by the constriction of envelope membranes at the division site. Although constriction of a ring-like protein complex has been shown to be involved in chloroplast division, it remains unknown how membrane lipids participate in the process. Here, we show that phosphoinositides with unknown function in envelope membranes are involved in the regulation of chloroplast division in Arabidopsis thaliana. PLASTID DIVISION1 (PDV1) and PDV2 proteins interacted specifically with phosphatidylinositol 4-phosphate (PI4P). Inhibition of phosphatidylinositol 4-kinase (PI4K) decreased the level of PI4P in chloroplasts and accelerated chloroplast division. Knockout of PI4Kβ2 expression or downregulation of PI4Kα1 expression resulted in decreased levels of PI4P in chloroplasts and increased chloroplast numbers. PI4Kα1 is the main contributor to PI4P synthesis in chloroplasts, and the effect of PI4K inhibition was largely abolished in the pdv1 mutant. Overexpression of DYNAMIN-RELATED PROTEIN5B (DRP5B), another component of the chloroplast division machinery, which is recruited to chloroplasts by PDV1 and PDV2, enhanced the effect of PI4K inhibition, whereas overexpression of PDV1 and PDV2 had additive effects. The amount of DRP5B that associated with chloroplasts increased upon PI4K inhibition. These findings suggest that PI4P is a regulator of chloroplast division in a PDV1- and DRP5B-dependent manner.  相似文献   

9.
10.
11.
Protein amino (N) termini are prone to modifications and are major determinants of protein stability in bacteria, eukaryotes, and perhaps also in chloroplasts. Most chloroplast proteins undergo N-terminal maturation, but this is poorly understood due to insufficient experimental information. Consequently, N termini of mature chloroplast proteins cannot be accurately predicted. This motivated an extensive characterization of chloroplast protein N termini in Arabidopsis (Arabidopsis thaliana) using terminal amine isotopic labeling of substrates and mass spectrometry, generating nearly 14,000 tandem mass spectrometry spectra matching to protein N termini. Many nucleus-encoded plastid proteins accumulated with two or three different N termini; we evaluated the significance of these different proteoforms. Alanine, valine, threonine (often in N-α-acetylated form), and serine were by far the most observed N-terminal residues, even after normalization for their frequency in the plastid proteome, while other residues were absent or highly underrepresented. Plastid-encoded proteins showed a comparable distribution of N-terminal residues, but with a higher frequency of methionine. Infrequent residues (e.g. isoleucine, arginine, cysteine, proline, aspartate, and glutamate) were observed for several abundant proteins (e.g. heat shock proteins 70 and 90, Rubisco large subunit, and ferredoxin-glutamate synthase), likely reflecting functional regulation through their N termini. In contrast, the thylakoid lumenal proteome showed a wide diversity of N-terminal residues, including those typically associated with instability (aspartate, glutamate, leucine, and phenylalanine). We propose that, after cleavage of the chloroplast transit peptide by stromal processing peptidase, additional processing by unidentified peptidases occurs to avoid unstable or otherwise unfavorable N-terminal residues. The possibility of a chloroplast N-end rule is discussed.Following synthesis, most proteins undergo various N-terminal (Nt) protein modifications, including removal of the Nt Met and signal peptide, N-terminal α-acetylation (NAA), ubiquitination, and acylations. These Nt modifications play an important role in the regulation of cellular functions. The N terminus of proteins has also been shown to be a major determinant of protein stability in bacteria (Varshavsky, 2011), eukaryotes (Graciet et al., 2009), mitochondria, and perhaps in plastids/chloroplasts (Apel et al., 2010; Nishimura et al., 2013; van Wijk, 2015). The role of the N terminus in protein stability is conceptualized in the N-end rule, which states that certain amino acids, when exposed at the N terminus of a protein, act as triggers for degradation (Bachmair et al., 1986; Dougan et al., 2012; Tasaki et al., 2012; Gibbs et al., 2014).Most of the approximately 3,000 plastid proteins are nucleus encoded (n-encoded) and are targeted to the plastid through an Nt chloroplast transit peptide (cTP). After import, the cTP is cleaved by the stromal processing peptidase (SPP; Richter and Lamppa, 1998; Trösch and Jarvis, 2011). The consensus site of cTP cleavage by SPP is only loosely defined, and the rules, mechanisms, and enzymes for possible subsequent processing, stabilization, and other posttranslational modifications (PTMs) are not well characterized (for discussion, see van Wijk, 2015). The exact N terminus is unknown for many chloroplast proteins and cannot be accurately predicted, because SPP specificity is not sufficiently understood (Emanuelsson et al., 2000; Zybailov et al., 2008) and probably also because additional Nt processing occurs for many chloroplast proteins (Fig. 1A). The approximately 85 plastid-encoded (p-encoded) proteins typically first undergo cotranslational Nt deformylation, followed by N-terminal Met excision (NME; Giglione et al., 2009; Fig. 1B); both these PTMs are required for normal plastid/chloroplast development and protein stability (Dirk et al., 2001, 2002; Giglione et al., 2003; Meinnel et al., 2006). Both n-encoded and p-encoded proteins can undergo NAA inside the plastid (Zybailov et al., 2008; Fig. 1). Postulated functions of NAA in eukaryotes include the mediation of protein location, assembly, and stability (Jones and O’Connor, 2011; Starheim et al., 2012; Hoshiyasu et al., 2013; Xu et al., 2015), thereby affecting a variety of processes, including drought tolerance in Arabidopsis (Arabidopsis thaliana; Linster et al., 2015).Open in a separate windowFigure 1.Conceptual illustration of Nt maturation of n-encoded and p-encoded proteins. Ac, Acetylated; MAP, Met amino peptidase; NAT, N-acetyltransferase; N-term, N-terminal; PDF, peptide deformylase. A, Nt maturation of n-encoded plastid proteins including removal of cTP by SPP and potential subsequent Nt modifications. B, Nt maturation of p-encoded proteins. *, The removal depends on the penultimate residue, generally following the N-terminal Met Excision (NME) rule; **, N-terminal acetylation typically occurs only for selected residues; “Results”).Typical proteomics work flows generally yield only partial coverage of protein sequences, and it is often difficult to know which peptides represent the true N termini (Nti) or C termini. Systematic identification of Nti or C termini requires specific labeling and enrichment strategies, such as combined fractional diagonal chromatography, developed by Gevaert and colleagues (Staes et al., 2011), and terminal amine isotopic labeling of substrates (TAILS), developed by the group of Overall (Kleifeld et al., 2011; Lange and Overall, 2013). These strategies allow the identification of different Nt proteoforms and were recently also applied to plants (Tsiatsiani et al., 2013; Carrie et al., 2015; Kohler et al., 2015; Zhang et al., 2015) and diatoms (Huesgen et al., 2013). We previously reported on Nti of chloroplast proteins based on tandem mass spectrometry (MS/MS) analysis, but because no Nt enrichment/labeling technique was used, only those that underwent NAA could be considered bona fide Nti (Zybailov et al., 2008). Nt Edman degradation sequencing was systematically carried out for thylakoid lumen proteins (Peltier et al., 2000, 2002) but not for stromal proteins or chloroplast membrane proteins with their Nti exposed to the stroma. The Nti of thylakoid lumen proteins are mostly generated by lumenal peptidases (Hsu et al., 2011; Midorikawa et al., 2014), and the thylakoid lumen contains a different set of peptidases than the stroma; hence, rules for Nt maturation and stability are likely different than those for stroma-exposed proteins.The objective of this study was to systematically determine the Nti of stroma-exposed chloroplast proteins of Arabidopsis (the N-terminome) and to provide a baseline for understanding Nt protein maturation and protein stability in the chloroplast stroma. To that end, we applied the TAILS technique and determined the Nti of approximately 250 chloroplast proteins by mass spectrometry (MS). We observed that many n-encoded plastid proteins accumulated with two or even three different Nt residues, in many cases both with and without NAA. The extent of accumulation of different Nt proteoforms is surprising and will be discussed. The p-encoded proteins generally showed very similar Nt residues as compared with the n-encoded proteins, with the exception of Met. Our data show that small, apolar, or hydroxylated residues (Ala, Val, Ser, and Thr) are the most frequent Nt residues of stromal proteins, whereas other residues are strictly avoided or are only present for very specific proteins likely to aid in their function. Chloroplast protein degradation products were also detected, with enrichment for peptides generated by cleavage between Arg and Thr residues. We present testable hypotheses for understanding Nt processing and maturation, stability, and a possible N-end rule in chloroplast stroma.  相似文献   

12.
13.
MicroRNAs (miRNAs) guide RNA-induced silencing complexes to target RNAs based on miRNA-target complementarity. Using a dual-luciferase based sensor system in Nicotiana benthamiana, we quantitatively assessed the relationship between miRNA-target complementarity and silencing efficacy measured at both the RNA and protein levels, using several conserved miRNAs and their known target sites from Arabidopsis thaliana. We found that naturally occurring sites have variable efficacies attributable to their complementarity patterns. We also observed that sites with a few mismatches to the miRNA 3′ regions, which are common in plants, are often equally effective and sometimes more effective than perfectly matched sites. By contrast, mismatches to the miRNA 5′ regions strongly reduce or eliminate repression efficacy but are nonetheless present in several natural sites, suggesting that in some cases, suboptimal miRNA efficacies are either tolerated or perhaps selected for. Central mismatches fully abolished repression efficacy in our system, but such sites then became effective miRNA target mimics. Complementarity patterns that are functional in animals (seed sites, 3′-supplementary sites, and centered sites) did not reliably confer repression, regardless of context (3′-untranslated region or open reading frame) or measurement type (RNA or protein levels). Overall, these data provide a robust and empirical foundation for understanding, predicting, and designing functional miRNA target sites in plants.  相似文献   

14.
15.
16.
17.
18.
19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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