首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
3.
4.
5.
Polar auxin transport (PAT), which is controlled precisely by both auxin efflux and influx facilitators and mediated by the cell trafficking system, modulates organogenesis, development and root gravitropism. ADP-ribosylation factor (ARF)-GTPase protein is catalyzed to switch to the GTP-bound type by a guanine nucleotide exchange factor (GEF) and promoted for hybridization to the GDP-bound type by a GTPase-activating protein (GAP). Previous studies showed that auxin efflux facilitators such as PIN1 are regulated by GNOM, an ARF-GEF, in Arabidopsis. In the November issue of The Plant Journal, we reported that the auxin influx facilitator AUX1 was regulated by ARF-GAP via the vesicle trafficking system.1 In this addendum, we report that overexpression of OsAGAP leads to enhanced root gravitropism and propose a new model of PAT regulation: a loop mechanism between ARF-GAP and GEF mediated by vesicle trafficking to regulate PAT at influx and efflux facilitators, thus controlling root development in plants.Key Words: ADP-ribosylation factor (ARF), ARF-GAP, ARF-GEF, auxin, GNOM, polar transport of auxinPolar auxin transport (PAT) is a unique process in plants. It results in alteration of auxin level, which controls organogenesis and development and a series of physiological processes, such as vascular differentiation, apical dominance, and tropic growth.2 Genetic and physiological studies identified that PAT depends on efflux facilitators such as PIN family proteins and influx facilitators such as AUX1 in Arabidopsis.Eight PIN family proteins, AtPIN1 to AtPIN8, exist in Arabidopsis. AtPIN1 is located at the basal side of the plasma membrane in vascular tissues but is weak in cortical tissues, which supports the hypothesis of chemical pervasion.3 AtPIN2 is localized at the apical side of epidermal cells and basally in cortical cells.1,4 GNOM, an ARF GEF, modulates the localization of PIN1 and vesicle trafficking and affects root development.5,6 The PIN auxin-efflux facilitator network controls root growth and patterning in Arabidopsis.4 As well, asymmetric localization of AUX1 occurs in the root cells of Arabidopsis plants,7 and overexpression of OsAGAP interferes with localization of AUX1.1 Our data support that ARF-GAP mediates auxin influx and auxin-dependent root growth and patterning, which involves vesicle trafficking.1 Here we show that OsAGAP overexpression leads to enhanced gravitropic response in transgenic rice plants. We propose a model whereby ARF GTPase is a molecular switch to control PAT and root growth and development.Overexpression of OsAGAP led to reduced growth in primary or adventitious roots of rice as compared with wild-type rice.1 Gravitropism assay revealed transgenic rice overxpressing OsAGAP with a faster response to gravity than the wild type during 24-h treatment. However, 1-naphthyl acetic acid (NAA) treatment promoted the gravitropic response of the wild type, with no difference in response between the OsAGAP transgenic plants and the wild type plants (Fig. 1). The phenotype of enhanced gravitropic response in the transgenic plants was similar to that in the mutants atmdr1-100 and atmdr1-100/atpgp1-100 related to Arabidopsis ABC (ATP-binding cassette) transporter and defective in PAT.8 The physiological data, as well as data on localization of auxin transport facilitators, support ARF-GAP modulating PAT via regulating the location of the auxin influx facilitator AUX1.1 So the alteration in gravitropic response in the OsAGAP transgenic plants was explained by a defect in PAT.Open in a separate windowFigure 1Gravitropism of OsAGAP overexpressing transgenic rice roots and response to 1-naphthyl acetic acid (NAA). (A) Gravitropism phenotype of wild type (WT) and OsAGAP overexpressing roots at 6 hr gravi-stimulation (top panel) and 0 hr as a treatment control (bottom panel). (B) Time course of gravitropic response in transgenic roots. (C and D) results correspond to those in (A and B), except for treatment with NAA (5 × 10−7 M).The polarity of auxin transport is controlled by the asymmetric distribution of auxin transport proteins, efflux facilitators and influx carriers. ARF GTPase is a key member in vesicle trafficking system and modulates cell polarity and PAT in plants. Thus, ARF-GDP or GTP bound with GEF or GAP determines the ARF function on auxin efflux facilitators (such as PIN1) or influx ones (such as AUX1).ARF1, targeting ROP2 and PIN2, affects epidermal cell polarity.9 GNOM is involved in the regulation of PIN1 asymmetric localization in cells and its related function in organogenesis and development.6 Although VAN3, an ARF-GAP in Arabidopsis, is located in a subpopulation of the trans-Golgi transport network (TGN), which is involved in leaf vascular network formation, it does not affect PAT.10 OsAGAP possesses an ARF GTPase-activating function in rice.11 Specifically, our evidence supports that ARF-GAP bound with ARF-GTP modulates PAT and gravitropism via AUX1, mediated by vesicle trafficking, including the Golgi stack.1Therefore, we propose a loop mechanism between ARF-GAP and GEF mediated by the vascular trafficking system in regulating PAT at influx and efflux facilitators, which controls root development and gravitropism in plants (Fig. 2). Here we emphasize that ARF-GEF catalyzes a conversion of ARF-bound GDP to GTP, which is necessary for the efficient delivery of the vesicle to the target membrane.12 An opposite process of ARF-bound GDP to GTP is promoted by ARF-GTPase-activating protein via binding. A loop status of ARF-GTP and ARF-GDP bound with their appurtenances controls different auxin facilitators and regulates root development and gravitropism.Open in a separate windowFigure 2Model for ARF GTPase as a molecular switch for the polar auxin transport mediated by the vesicle traffic system.  相似文献   

6.
7.
8.
Auxin is a phytohormone essential for plant development. Due to the high redundancy in auxin biosynthesis, the role of auxin biosynthesis in embryogenesis and seedling development, vascular and flower development, shade avoidance and ethylene response were revealed only recently. We previously reported that a vitamin B6 biosynthesis mutant pdx1 exhibits a short-root phenotype with reduced meristematic zone and short mature cells. By reciprocal grafting, we now have found that the pdx1 short root is caused by a root locally generated signal. The mutant root tips are defective in callus induction and have reduced DR5::GUS activity, but maintain relatively normal auxin response. Genetic analysis indicates that pdx1 mutant could suppress the root hair and root growth phenotypes of the auxin overproduction mutant yucca on medium supplemented with tryptophan (Trp), suggesting that the conversion from Trp to auxin is impaired in pdx1 roots. Here we present data showing that pdx1 mutant is more tolerant to 5-methyl anthranilate, an analogue of the Trp biosynthetic intermediate anthranilate, demonstrating that pdx1 is also defective in the conversion from anthranilate to auxin precursor tryptophan. Our data suggest that locally synthesized auxin may play an important role in the postembryonic root growth.Key words: auxin synthesis, root, PLP, PDX1The plant hormone auxin modulates many aspects of growth and development including cell division and cell expansion, leaf initiation, root development, embryo and fruit development, pattern formation, tropism, apical dominance and vascular tissue differentiation.13 Indole-3-acetic acid (IAA) is the major naturally occurring auxin. IAA can be synthesized in cotyledons, leaves and roots, with young developing leaves having the highest capacity.4,5Auxin most often acts in tissues or cells remote from its synthetic sites, and thus depends on non-polar phloem transport as well as a highly regulated intercellular polar transport system for its distribution.2The importance of local auxin biosynthesis in plant growth and development has been masked by observations that impaired long-distance auxin transport can result in severe growth or developmental defects.3,6 Furthermore, a few mutants with reduced free IAA contents display phenotypes similar to those caused by impaired long-distance auxin transport. These phenotypes include defective vascular tissues and flower development, short primary roots and reduced apical dominance, or impaired shade avoidance and ethylene response.715 Since these phenotypes most often could not be rescued by exogenous auxin application, it is difficult to attribute such defects to altered local auxin biosynthesis. By complementing double, triple or quadruple mutants of four Arabidopsis shoot-abundant auxin biosynthesis YUCCA genes with specific YUCCA promoters driven bacterial auxin biosynthesis iaaM gene, Cheng et al. provided unambiguous evidence that auxin biosynthesis is indispensable for embryo, flower and vascular tissue development.8,13 Importantly, it is clear that auxin synthesized by YUCCAs is not functionally interchangeable among different organs, supporting the notion that auxin synthesized by YUCCAs mainly functions locally or in a short range.6,8,13The central role of auxin in root meristem patterning and maintenance is well documented,1,2,16 but the source of such IAA is still unclear. When 14C-labeled IAA was applied to the five-day-old pea apical bud, the radioactivity could be detected in lateral root primordia but not the apical region of primary roots.17 Moreover, removal of the shoot only slightly affected elongation of the primary root, and localized application of auxin polar transport inhibitor naphthylphthalamic acid (NPA) at the primary root tip exerted more profound inhibitory effect on root elongation than at any other site.18 These results suggest that auxin generated near the root tip may play a more important role in primary root growth than that transported from the shoot. In line with this notion, Arabidopsis roots have been shown to harbor multiple auxin biosynthesis sites including root tips and the region upward from the tip.4Many steps of tryptophan synthesis and its conversion to auxin involve transamination reactions, which require the vitamin B6 pyridoxal 5-phosphate (PLP) as a cofactor. We previously reported that the Arabidopsis mutant pdx1 that is defective in vitamin B6 biosynthesis displays dramatically reduced primary root growth with smaller meristematic zone and shorter mature cortical cells.19 In the current investigation, we found that the root tips of pdx1 have reduced cell division capability and reduced DR5::GUS activity, although the induction of this reporter gene by exogenous auxin was not changed. Reciprocal grafting indicates that the short-root phenotype of pdx1 is caused by a root local rather than shoot generated factor(s). Importantly, pdx1 suppresses yucca mutant, an auxin overproducer, in root hair proliferation although it fails to suppress the hypocotyl elongation phenotype.20 Our work thus demonstrated that pdx1 has impaired root local auxin biosynthesis from tryptophan. To test whether the synthesis of tryptophan is also affected in pdx1 mutant, we planted pdx1 together with wild-type seeds on Murashige and Skoog (MS) medium supplemented with 5-mehtyl-anthranilate (5-MA), an analogue of the Trp biosynthetic intermediate anthranilate.21 Although pdx1 seedlings grew poorly under the control conditions, the growth of wild-type seedlings was more inhibited than that of the pdx1 seedlings on 10 µM 5-MA media (Fig. 1A–D). Compared with the elongated primary root on MS, wild-type seedlings showed very limited root growth on 5-MA (Fig. 1E). The relatively increased tolerance to 5-MA of pdx1 thus indicates that the pdx1 mutant may be defective in Trp biosynthesis, although amino acid analysis of the bulked seedlings did not find clear changes in Trp levels in the mutants (our unpublished data).Open in a separate windowFigure 1The pdx1 mutant seedlings are relatively less sensitive to toxic 5-methyl anthranilate (5-MA). (A and C) Five-day-old seedlings of the wild type (Col-0) (A) or pdx1 (C) on MS medium. (B and D) Five-day-old seedlings of the wild type (B) or pdx1 (D) on MS medium supplemented with 10 µM 5-MA. (E) Eight-day-old seedlings of the wild type or pdx1 on MS medium without or with 10 µM 5-MA supplement. Sterilized seeds were planted directly on the indicated medium and after two days of cold treatment, the plates were incubated under continuous light at 22–24°C before taking pictures.We reported that PDX1 is required for tolerance to oxidative stresses in Arabidopsis.19 Interestingly, redox homeostasis appears to play a critical role in Arabidopsis root development. The glutathione-deficient mutant root meristemless1 (rml1) and the vitamin C-deficient mutant vitamin C1 (vtc1) both have similar stunted roots.22,23 Nonetheless, pdx1 is not rescued by either glutathione or vitamin C19 suggesting that the pdx1 short-root phenotype may not be resulted from a general reduction of antioxidative capacity. Interestingly, ascorbate oxidase is found to be highly expressed in the maize root quiescent center.24 This enzyme can oxidatively decarboxylate auxin in vitro, suggesting that the quiescent center may be a site for metabolizing auxin to control its homeostasis.25 It is therefore likely that the reduced auxin level in pdx1 root tips could be partially caused by increased auxin catabolism resulted from reduced vitamin B6 level. We thus conducted experiments to test this possibility. A quiescent center-specific promoter WOX5 driven bacterial auxin biosynthetic gene iaaH26 was introduced into pdx1 mutant. The transgenic seeds were planted on media supplemented with different concentrations of indoleacetamide (IAM), the substrate of iaaH protein. Although promotion of lateral root growth was observed at higher IAM concentrations, which indicates increased tryptophan-independent auxin production from the transgene, no change in root elongation was observed between pdx1 with or without the WOX5::iaaH transgene at any concentration of IAM tested (data not shown), suggesting that the pdx1 short-root phenotype may not be due to increased auxin catabolism.Taken together, in addition to auxin transport; temporally, spatially or developmentally coordinated local auxin biosynthesis defines the plant growth and its response to environmental changes.8,14,15  相似文献   

9.
There is increasing evidence that immunophilins function as key regulators of plant development. One of the best investigated members, the multi-domain FKBP TWISTED DWARF1 (TWD1)/FKBP42, has been shown to reside on both the vacuolar and plasma membranes where it interacts in mirror image with two pairs of ABC transporters, MRP1/ MRP2 and PGP1/PGP19(MDR1), respectively. Twisted dwarf1 and pgp1/pgp19 mutants display strongly overlapping phenotypes, including reduction and disorientation of growth, suggesting functional interaction.In a recent work using plant and heterologous expression systems, TWD1 has been demonstrated to modulate PGP-mediated export of the plant hormone auxin, which controls virtually all plant developmental processes. Here we summarize recent molecular models on TWD1 function in plant development and PGP-mediated auxin tranport and discuss open questions.Key Words: Twisted Dwarf1, plant development, auxin, immunophilin, P-glycoprotein, ABC transporterFK506-binding Proteins (FKBPs), together with unrelated cyclophilins, belong to the immunophilins, an ancient and ubiquitous protein family.1,4,5 They were first described as receptors for immunosuppressive drugs in animal and human cells, FK506 and cyclosporin A, respectively.1 All FKBP-type immunophilins share a characteristic peptidyl-prolyl cis-trans isomerase domain (PPIase domain or FKBD, Fig. 2A) making protein folding a key feature among immunophilins.2 The best investigated example, the human cytosolic single-domain FKBP12, modulates Ca2+ release channels6,7 and associates with the cell cycle regulator TGF-β.8 Furthermore, the human FKBP12/FK506 complex is known to bind and inhibit calcineurin activity,9 leading to immune response inhibition. However, not all single- and multiple-domain FKBPs own folding activity and, interestingly, many form distinct protein complexes with diverse functions.35Open in a separate windowFigure 2Model of TWISTED DWARF 1 interacting proteins. (A) Domain structure of TWD1 and putative interacting proteins. FKBD, FK506-binding domain: TPR, tetratricopeptide repeat; CaM(-BD, calmodulin-binding domain; MA, membrane anchor. For details, see text. (B) Functional TWD1-ABC transporter complexes on both the vacuolar and plasma membrane. While for TWD1/PGP pairs, the positive regulatory role on auxin transport was demonstrated,18 the modulation of MRP-mediated vacuolar import of glutathion conjugates (GS-X) was established using mammalian test substrates17 because the in vivo substrates are unknown. Note that C-terminal nucleotide binding folds of MRP- and PGP-like ABC transporters interact with distinct functional domains of TWD1, the TPR and FKBD, respectively. The native auxin, IAAH, gets trapped by deprotonization upon uptake into the cell. Export is catalyzed by secondary active export via PIN-like efflux carriers15 and/or by primary active, ATP-driven P-glycoproteins (PGPs, right panel); loss-of TWD1 function abolishes PGP-mediated auxin export (left panel).  相似文献   

10.
11.
12.
The conserved eukaryotic protein SGT1 (suppressor of G2 allele of skp1) participates in diverse physiological processes such as cell cycle progression in yeast, plant immunity against pathogens and plant hormone signalling. Recent genetic and biochemical studies suggest that SGT1 functions as a novel co-chaperone for cytosolic/nuclear HSP90 and HSP70 molecular chaperones in the folding and maturation of substrate proteins. Since proteins containing the leucine-rich repeat (LRR) protein-protein interaction motif are overrepresented in SGT1-dependent phenomena, we consider whether LRR-containing proteins are preferential substrates of an SGT1/HSP70/HSP90 complex. Such a chaperone organisation is reminiscent of the HOP/HSP70/HSP90 machinery which controls maturation and activation of glucocorticoid receptors in animals. Drawing on this parallel, we discuss the possible contribution of an SGT1-chaperone complex in the folding and maturation of LRR-containing proteins and its evolutionary consequences for the emergence of novel LRR interaction surfaces.Key words: heat shock protein, SGT1, co-chaperone, HSP90, HSP70, leucine-rich repeat, LRR, resistance, SCF, ubiquitinThe proper folding and maturation of proteins is essential for cell viability during de novo protein synthesis, translocation, complex assembly or under denaturing stress conditions. A complex machinery composed of molecular chaperones (heat-shock proteins, HSPs) and their modulators known as co-chaperones, catalyzes these protein folding events.1,2 In animals, defects in the chaperone machinery is implicated in an increasing number of diseases such as cancers, susceptibility to viruses, neurodegenerative disease and cystic fibrosis, and thus it has become a major pharmacological target.3,4 In plants, molecular genetic studies have identified chaperones and co-chaperones as components of various physiological responses and are now starting to yield important information on how chaperones work. Notably, processes in plant innate immunity rely on the HSP70 and HSP9057 chaperones as well as two recently characterised co-chaperones, RAR1 (required for Mla12 resistance) and SGT1 (suppressor of G2 allele of skp1).811SGT1 is a highly conserved and essential co-chaperone in eukaryotes and is organized into three structural domains: a tetratricopeptide repeat (TPR), a CHORD/SGT1 (CS) and an SGT1-specific (SGS) domain (Fig. 1A). SGT1 is involved in a number of apparently unrelated physiological responses ranging from cell cycle progression and adenylyl cyclase activity in yeast to plant immunity against pathogens, heat shock tolerance and plant hormone (auxin and jasmonic acid) signalling.79,12,13 Because the SGT1 TPR domain is able to interact with Skp1, SGT1 was initially believed to be a component of SCF (Skp1/Cullin/F-box) E3 ubiquitin ligases that are important for auxin/JA signalling in plants and cell cycle progression in yeast.13,14 However, mutagenesis of SGT1 revealed that the TPR domain is dispensable for plant immunity and auxin signalling.15 Also, SGT1-Skp1 interaction was not observed in Arabidopsis.13 More relevant to SGT1 functions appear to be the CS and SGS domains.16 The former is necessary and sufficient for RAR1 and HSP90 binding. The latter is the most conserved of all SGT1 domains and the site of numerous disabling mutations.14,16,17Open in a separate windowFigure 1Model for SGT1/chaperone complex functions in the folding of LRR-containing proteins. (A) The structural domains of SGT1, their sites of action (above) and respective binding partners (below) are shown. N- and C-termini are indicated. TPR, tetratricopeptide repeat; CS, CHORD/SGT1; SGS, SGT1-specific. (B) Conceptual analogy between steroid receptor folding by the HOP/chaperone machinery and LRR protein folding by the SGT1/chaperone machinery. LRR motifs are overrepresented in processes requiring SGT1 such as plant immune receptor signalling, yeast adenylyl cyclase activity and plant or yeast SCF (Skp1/Cullin/F-box) E3 ubiquitin ligase activities. (C) Opposite forces drive LRR evolution. Structure of LRRs 16 to 18 of the F-box auxin receptor TIR1 is displayed as an illustration of the LRR folds.30 Leucine/isoleucine residues (side chain displayed in yellow) are under strong purifying selection and build the hydrophobic LRR backbone (Left). By contrast, solvent-exposed residues of the β-strands define a polymorphic and hydrophilic binding surface conferring substrate specificity to the LRR (Right) and are often under diversifying selection.We recently demonstrated that Arabidopsis SGT1 interacts stably through its SGS domain with cytosolic/nuclear HSP70 chaperones.7 The SGS domain was both necessary and sufficient for HSP70 binding and mutations affecting SGT1-HSP70 interaction compromised JA/auxin signalling and immune responses. An independent in vitro study also found interaction between human SGT1 and HSP70.18 The finding that SGT1 protein interacts directly with two chaperones (HSP90/70) and one co-chaperone (RAR1) reinforces the notion that SGT1 behaves as a co-chaperone, nucleating a larger chaperone complex that is essential for eukaryotic physiology. A future challenge will be to dissect the chaperone network at the molecular and subcellular levels. In plant cells, SGT1 localization appears to be highly dynamic with conditional nuclear localization7 and its association with HSP90 was recently shown to be modulated in vitro by RAR1.16A co-chaperone function suits SGT1 diverse physiological roles better than a specific contribution to SCF ubiquitin E3 ligases. Because SGT1 does not affect HSP90 ATPase activity, SGT1 was proposed rather as a scaffold protein.16,19 In the light of our findings and earlier studies,20 SGT1 is reminiscent of HOP (Hsp70/Hsp90 organizing protein) which links HSP90 and HSP70 activities and mediates optimal substrate channelling between the two chaperones (Fig. 1B).21 While the contribution of the HSP70/HOP/HSP90 to the maturation of glucocorticoid receptors is well established,21 direct substrates of an HSP70/SGT1/HSP90 complex remain elusive.It is interesting that SGT1 appears to share a functional link with leucine-rich repeat- (LRR) containing proteins although LRR domains are not so widespread in eukaryotes. For example, plant SGT1 affects the activities of the SCFTIR1 and SCFCOI1 E3 ligase complexes whose F-box proteins contain LRRs.13 Moreover, plant intracellular immune receptors comprise a large group of LRR proteins that recruit SGT1.8,9 LRRs are also found in yeast adenylyl cyclase Cyr1p and the F-box protein Grr1p which is required for SGT1-dependent cyclin destruction during G1/S transition.12,14 Yeast 2-hybrid interaction assays also revealed that yeast and plant SGT1 tend to associate directly or indirectly with LRR proteins.12,22,23 We speculate that SGT1 bridges the HSP90-HSC70 chaperone machinery with LRR proteins during complex maturation and/or activation. The only other structural motif linked to SGT1 are WD40 domains found in yeast Cdc4p F-box protein and SGT1 interactors identified in yeast two-hybrid screens.12What mechanisms underlie a preferential SGT1-LRR interaction? HSP70/SGT1/HSP90 may have co-evolved to assist specifically in folding and maturation of LRR proteins. Alternatively, LRR structures may have an intrinsically greater need for chaperoning activity to fold compared to other motifs. These two scenarios are not mutually exclusive. The LRR domain contains multiple 20 to 29 amino acid repeats, forming an α/β horseshoe fold.24 Each repeat is rich in hydrophobic leucine/isoleucine residues which are buried inside the structure and form the structural backbone of the motif (Fig. 1C, left). Such residues are under strong purifying selection to preserve structure. These hydrophobic residues would render the LRR a possible HSP70 substrate.25 By contrast, hydrophilic solvent- exposed residues of the β strands build a surface which confers ligand recognition specificity of the LRRs (Fig. 1C). In many plant immune receptors for instance, these residues are under diversifying selection that is likely to favour the emergence of novel pathogen recognition specificities in response to pathogen evolution.26 The LRR domain of such a protein has to survive such antagonist selection forces and yet remain functional. Under strong selection pressure, LRR proteins might need to accommodate less stable LRRs because their recognition specificities are advantageous. This could be the point at which LRRs benefit most from a chaperoning machinery such as the HSP90/SGT1/HSP70 complex. This picture is reminiscent of the genetic buffering that HSP90 exerts on many traits to mask mutations that would normally be deleterious to protein folding and/or function, as revealed in Drosophila and Arabidopsis.27 It will be interesting to test whether the HSP90/SGT1/HSP70 complex acts as a buffer for genetic variation, favouring the emergence of novel LRR recognition surfaces in, for example, highly co-evolved plant-pathogen interactions.28,29  相似文献   

13.
Legumes can establish a symbiosis with rhizobia and form root nodules that function as an apparatus for nitrogen fixation. Nodule development is regulated by several phytohormones including auxin. Although accumulation of auxin is necessary to initiate the nodulation of indeterminate nodules, the functions of auxin on the nodulation of determinate nodules have been less characterized. In this study, the functions of auxin in nodule development in Lotus japonicus have been demonstrated using an auxin responsive promoter and auxin inhibitors. We found that the lenticel formation on the nodule surface was sensitive to the auxin defect. Further analysis indicated that failure in the development of the vascular bundle of the determinate nodule, which was regulated by auxin, was the cause of the disappearance of lenticels.Key words: auxin, lenticel, Lotus japonicus, nodulation, symbiotic nitrogen fixationLegumes (Fabaceae) constitute the third largest plant family with around 700 genera and 20,000 species.1 Legume plants form root nodules through symbiosis with a soil microbe called rhizobia. This plant-microbe symbiosis in nodules mediates an harmonized exchange of chemical signals between host plants and rhizobia.2 Nodules are biologically divided into two different groups, i.e., indeterminate nodules and determinate nodules. Indeterminate nodules, represented by Trifolium repens (white clover) and Medicago truncatula, are initiated from the inner cortex to form a persistent nodule meristem, which allows continuous growth, and leads to the formation of elongated nodules, whereas in determinate legumes, nodules are mostly developed from outer cortical cells and form spherical nodules.3Auxin is one of the most important regulators for nodule development. Since the possible involvement of auxin in nodule formation was first reported by Thimann,4 auxin distribution during nodulation has been studied in particular with indeterminate nodules.5 However, little is known about auxin involvement in determinate nodule formation. To evaluate auxin functions in the determinate nodulation of legume plants, we performed an auxin-responsive promoter analysis in detail. Using GH3:GUS transformed Lotus japonicus (a kind gift from Dr. Herman P. Spaink, Leiden State University, Netherlands),6 we detected auxin signals throughout the nodulation process, e.g., at the basal and front part of the nodule primordia, circumjacent to the infection zone of the young developing nodules (Fig. 1), and at the nodule vascular bundle in mature nodules. We also investigated the effect of several auxin inhibitors, including newly synthesized auxin antagonist PEO-IAA (kindly provided by Dr. Hayashi, Okayama University of Science, Japan),7 on the nodulation of L. japonicus, and revealed that auxin was required for forming a nodule vascular bundle and lenticels (Fig. 2).8Open in a separate windowFigure 1GH3:GUS expression in determinate nodule at 6 dpi. (A) GUS staining was observed in the central cylinder of the root vascular bundle and in the nodule. (B) Cross section of (A). GUS expression was observed around the infection zone of the nodule. Bars = 100 µm.Open in a separate windowFigure 2The effect of auxin inhibitor on nodule surface. (A) Typical mature nodule of L. japonicus at 21 dpi. Lenticels are pointed out by yellow arrowheads. (B) The treatment of auxin inhibitor (NPA 100 µM) inhibited lenticel formation on the nodule surface. Bars = 500 µm.In indeterminate legumes, auxin is accumulated at the site of rhizobia inoculation.9 This is caused by the inhibition of polar auxin transport by accumulation of flavonoids around the infection site, which are known as regulators of auxin transport. When flavonoid biosynthesis is reduced by the gene silencing of chalcone synthase, which catalyzes the first step of flavonoid synthesis, M. truncatula was unable to inhibit polar auxin transport and resulted in reduced nodule number.10,11 A similar phenotype was observed when the auxin transporter gene was silenced.12 In addition, treatment of polar auxin transport inhibitors such as NPA and TIBA induce pseudonodule formation,9 suggesting that auxin accumulation is required for nodulation of indeterminate legumes. In contrast, the treatment of polar auxin transport inhibitors in determinate nodules did not induce a nodule-like structure, suggesting a different function of auxin between indeterminate and determinate nodules. It is, however, of interest to investigate the involvement of flavonoids in determinate nodule formation, because several genes in the flavonoid biosynthesis pathway are upregulated at 2 dpi (days post inoculation) in L. japonicus.13Lenticels regulate gas permeability of nodules.14 Under low oxygen or water-logged conditions, they develop more extensively, whereas they collapse, or develop very little during insufficient water conditions, or under high oxygen pressure.14,15 Because lenticel development on the nodule surface is accompanied with the nodule vascular bundle, growth regulators supplied from the vascular system likely facilitate lenticel development.15 Our data suggests that auxin is necessary to form the nodule vascular bundle, and in fact, auxin itself is one of the candidates of growth substances that control lenticel formation. It is necessary to analyze mutants, which lack in lenticel formation, but can form a nodule vascular bundle, for clarification of further mechanisms of lenticel development.  相似文献   

14.
15.
16.
17.
Plant-pathogen interaction induces a complex host response that coordinates various signaling pathways through multiple signal molecules. Besides the well-documented signal molecules salicylic acid (SA), ethylene and jasmonic acid, auxin is emerging as an important player in this response. We recently characterized an Arabidopsis activation-tagged mutant, bud1, in which the expression of the MAP kinase kinase 7 (AtMKK7) gene is increased. The bud1 mutant plants accumulate elevated levels of SA and display constitutive pathogenesis-related (PR) gene expression and enhanced resistance to pathogens. Additionally, increased expression of AtMKK7 in the bud1 mutant causes deficiency in polar auxin transport, indicating that AtMKK7 negatively regulates auxin signaling. Based on these results, we hypothesized that AtMKK7 may serve as a crosstalk point between auxin signaling and defense responses. Here we show that increased expression of AtMKK7 in bud1 results in a significant reduction in free auxin (indole-3-acetic acid) levels in the mutant plants. We propose three possible mechanisms to explain how AtMKK7 coordinates the growth hormone auxin and the defense signal molecule SA in the bud1 mutant plants. We suggest that AtMKK7 may play a role in cell death and propose that AtMPK3 and AtMPK6 may function downstream of AtMKK7.Key words: Arabidopsis, MAP kinase kinase 7, auxin signaling, defense responses, crosstalkPathogen invasion of a plant induces multiple physiological changes at the site of infection, including the accumulation of reactive oxygen species, nitric oxide and salicylic acid (SA).16 Jasmonic acid (JA) and ethylene (ET) are also produced in response to pathogen infection.711 Numerous reports have documented that SA, JA and ET work synergistically or antagonistically to fine-tune plant defense responses, based on a multitude of environmental, host and pathogen genetic factors that vary depending on the pathogen-host combinations.4,12The growth hormone auxin may also play an important role in plant defense responses. Many plant-pathogenic microorganisms have the ability to produce indole-3-acetic acid (IAA),13 which is important for the pathogenicity for some pathogens.1416 In the Arabidopsis-Xanthomonas campestris pv. campestris (Xcc) compatible interaction, Xcc triggers IAA synthesis in the host plants.17 Exogenous treatment of plants with the auxin analogs, NAA and 2,4-D, leads to disease susceptibility.18 A flagellin-derived-peptid e-induced microRNA (miRNA) was found to negatively regulate messenger RNAs for the F-box auxin receptors TIR1, AFB2 and AFB3, to repress auxin signaling, resulting in significantly enhanced host resistance.18 These results suggest that auxin likely functions as a virulence factor to suppress host defense.We previously identified an Arabidopsis activation-tagged mutant bud1 from a transgenic population generated by a sense/antisense RNA expression system.19 Further characterization indicated that bud1 is a semidominant mutant, in which the expression of the Arabidopsis MAP kinase kinase 7 (AtMKK7) gene is increased.20 The increased expression of AtMKK7 in bud1 causes deficiency in auxin transport, whereas reducing mRNA levels of AtMKK7 by antisense RNA expression leads to enhancement of auxin transport, indicating that AtMKK7 negatively regulates polar auxin transport (PAT).20 Recently, we have shown that the bud1 mutant plants accumulate elevated levels of SA and exhibit constitutive pathogenesis-related (PR) gene expression and enhanced resistance to both the bacterial pathogen Pseudomonas syringae pv. maculicola (Psm) ES4326 and the oomycete pathogen Hyaloperonospora parasitica Noco2.21 Reducing mRNA levels of AtMKK7 by antisense RNA expression not only compromises basal resistance but also blocks the induction of systemic acquired resistance (SAR), demonstrating that AtMKK7 is a positive regulator required for both basal resistance and SAR.21 Furthermore, we found that the free IAA levels in the bud1 mutant plants were significantly reduced, compared to those in wild-type plants (Fig. 1A). All these results taken together suggest that AtMKK7 may positively regulate SA signaling and negatively regulate auxin signaling.Open in a separate windowFigure 1(A) Free IAA levels in wild type (WT) and bud1 mutant plants. Thirty-day-old soil grown plants were used for free IAA measurement. (B) A schematic representation of three possible mechanisms through which MKK7 regulates host responses after pathogen invasion.Given that SA is a positive regulator of defense responses, whereas auxin is likely a negative regulator of defense responses, we propose three possible mechanisms through which AtMKK7 coordinates the growth hormone auxin and the defense signal molecule SA in the bud1 mutant plants (Fig. 1B): (1) AtMKK7 induces SA accumulation, which suppresses auxin signaling, leading to increased defense responses; (2) AtMKK7 independently induces SA accumulation and suppresses auxin signaling; (3) AtMKK7 suppresses auxin signaling, which relieves the repression of SA signaling by auxin, resulting in SA accumulation.We could test the hypotheses using different approaches. We can examine whether the expression of YUC1, YUC2, YUC4 and YUC6, genes that have been suggested to play essential roles in auxin biosynthesis,22 is altered in the bud1 mutant. We can also analyze the expression of YUC1, YUC2, YUC4 and YUC6, as well as the levels of free IAA in the double mutant bud1sid2 (sid2 is a SA deficient mutant) to test whether IAA biosynthesis is derepressed in the double mutant. Furthermore, polar auxin transport in the bud1sid2 plants should be determined. Finally, we can test whether exogenous application of auxin is able to suppress AtMKK7-induced constitutive defense responses in the bud1 mutant, including elevated levels of SA, constitutive PR gene expression and enhanced resistance to Psm ES4326 and H. parasitica Noco2.AtMKK7 belongs to the Group D of plant MAPKKs.23 Functions of two other members of this group, LeMKK4 and NbMKK1, have been described.24,25 LeMKK4 and NbMKK1 are orthologs of AtMKK7 in tomato and Nicotiana benthamiana, respectively. When overexpressed in leaves, wild-type LeMKK4 elicits cell death in both tomato and N. benthamiana.24 Overexpression of wild-type NbMKK1 also causes cell death on N. benthamiana leaves.25 We expected that overexpression of AtMKK7 would also result in cell death. However, neither increased expression of AtMKK7 in the bud1 mutant plants, nor overexpression of wild-type AtMKK7 from the dexamethasone-inducible promoter causes cell death.21 This is probably because the expression levels of AtMKK7 in these plants were below the threshold to induce cell death. Consistently, ectopic and constitutive expression of AtMKK7 driven by the cauliflower mosaic virus (CaMV) 35S promoter in wild-type plants leads to lethality of the transgenic plants.20 Therefore, to characterize the function of AtMKK7 in cell death, transgenic plants expressing a constitutively active form of AtMKK7 (AtMKK7S193A/S199D) from the dexamethasone-inducible promoter will be useful.What MAPK(s) acts downstream of AtMKK7? LeMKK4 directly phosphorylates LeMPK1, LeMPK2 and LeMPK3 in vitro, and activates LeMPK2 and LeMPK3 when expressed in tomato leaves,24 whereas NbMKK1 activates NbSIPK when expressed in N. benthamiana leaves.25 LeMPK2 and LeMPK3 are tomato orthologs of the well-studied tobacco proteins SIPK (salicylic acid-induced protein kinase) and WIPK (wound-induced protein kinase),26,27 respectively. The Arabidopsis orthologs of SIPK and WIPK are AtMPK6 and AtMPK3, respectively. Based on previous in-gel kinase assay results,21 we predict that both AtMPK3 and AtMPK6 may function downstream of AtMKK7. Characterization of double mutants bud1atmpk3 and bud1atmpk6, as well as atmpk3 and atmpk6 mutant plants expressing the constitutively active form of AtMKK7 from the dexamethasone-inducible promoter will shed light on this question.  相似文献   

18.
19.
Using time-lapse photography, we studied the response kinetics of low light intensity-induced upward leaf-movement, called hyponastic growth, in Arabidopsis thaliana. This response is one of the traits of shade avoidance and directs plant organs to more favorable light conditions. Based on mutant- and pharmacological data we demonstrated that among other factors, functional auxin perception and polar auxin transport (PAT) are required for the amplitude of hyponastic growth and for maintenance of the high leaf angle, upon low light treatment. Here, we present additional data suggesting that auxin and PAT antagonize the hyponastic growth response induced by ethylene treatment. We conclude that ethylene- and low light-induced hyponastic growth occurs at least partly via separate signaling routes, despite their strong similarities in response kinetics.Key words: hyponastic growth, petiole, Arabidopsis, ethylene, low light, auxin, polar auxin transport, differential growthUpward leaf movement (hyponastic growth) is a trait of several plant species to escape from growth-limiting conditions.1,2 Interestingly, Arabidopsis thaliana induces a marked hyponastic growth response triggered by various environmental stimuli, including complete submergence, high temperature, canopy shade and spectral neutral low light intensities (Fig. 1).36 The paper of Millenaar et al. in the New Phytologist 2009,7 provides a detailed analysis of low light intensity-induced hyponastic growth and components of the signal transduction are characterized using time-lapse photography. Low light intensity-induced hyponastic growth is a component of the so-called shade avoidance syndrome. Light-spectrum manipulations and mutant analyses indicated that predominantly the blue light wavelength region affects petiole movement and fast induction of hyponastic growth to low light conditions involves the photoreceptor proteins Cryptochrome 1 (Cry1), Cry2, Phytochrome-A (PhyA) and PhyB. Moreover, we show that also photosynthesis-derived signals can induce differential growth.Open in a separate windowFigure 1Typical hyponastic growth phenotype of Arabidopsis thaliana. Side view of Columbia-0 plants treated 10 h with ethylene (5 µl l−1) or low light (20 µmol m−2 s−1). Plants in control light conditions were in 200 µmol m−2 s−1. Both stimuli induce a clear leaf inclination (hyponasty) relative to the horizontal by differential growth of the petioles. Plants kept in control conditions only show modest diurnal leaf movement and leaf angles gradually decline over time due to maturation of the leaves. Note that the paint droplets were applied to facilitate quantitative measurement of leaf angle kinetics in a time-lapse camera setup.7The hyponastic growth response to low light intensity was not affected in several ethylene-insensitive mutant lines. Moreover, low light did not affect expression of ethylene inducible marker genes nor differences in ethylene release were noted. Therefore, we concluded that low light-induced hyponastic growth is independent of ethylene signaling. This is perhaps surprising, because ethylene is the main trigger of hyponastic growth induced by complete submergence in several species. Interestingly, both ethylene and low light can induce hyponastic growth in Arabidopsis with similar kinetics.3We showed that plants mutant in auxin perception components (transport inhibitor response1 (tir1) and tir1 afb1 afb2 afb3 quadruple, containing additional mutant alleles of TIR1 homologous F-box proteins) and plants mutant in (polar) auxin transport (tir3-1, pin-formed3 (pin3) and pin7) components had a lower hyponastic growth amplitude in low light conditions.7 Moreover, these mutants were less able to maintain the high leaf angles after the response maximum. Both characteristics were also noted in plants pre-treated with the polar auxin transport (PAT) inhibitor 2,3,5-triiodobenzoic acid (TIBA). We therefore concluded that auxin perception and PAT are involved in the regulation of low light-induced hyponastic growth.7 Interestingly, we observed that TIBA pretreatment did not inhibit ethylene-induced hyponastic growth. In fact, the response upon ethylene treatment was even modestly enhanced. In agreement with this observation, we show here that the above mentioned auxin perception and PAT mutants also showed a slightly enhanced hyponastic growth response upon ethylene treatment (Fig. 2).Open in a separate windowFigure 2Auxin involvement in ethylene induced hyponasty. Effect of exposure to ethylene (5 µl l−1) on the kinetics of hyponastic petiole growth (A) in Arabidopsis thaliana Columbia-0 plants treated with 50 µm TIBa (open circles) or a mock treatment (line) adapted from Supporting Information Figure S3 of Millenaar et al. (2009)7 and (B–F) in Arabidopsis auxin signaling and polar auxin transport mutants (closed circles), compared to the wild type response to low light (lines). Petiole angles are pair wise subtracted, which corrects for diurnal petiole movement in control conditions. For details on this procedure, growth conditions, treatments, data acquirement and analysis see.7,13 Error bars represent standard errors; n ≥ 12. mutants were obtained from the Nottingham Arabidopsis Stock Center (accession numbers are shown between brackets) or from the authors describing the lines. tir1-1 (n3798,14), tir1-1 afb1-1 afb2-1 (in a mixed Columbia/Wassilewskija background),15 tir3-1,14 pin3-4 (n9363,16) and pin7-1 (n9365,10).Despite that auxin and PAT are required for many differential growth responses such as phototropism and gravitropism,8,11 these data indicate that auxin perception and PAT are not obligatory for ethylene-induced hyponasty in Arabidopsis per se. In fact, one might even conclude that auxin and PAT antagonizes ethylene-induced hyponasty. These results are partly in agreement with observations on the wetland species Rumex palustris, were pretreatment with the auxin-efflux carrier 1-naphthylphthalamic acid (NPA) resulted in doubling of the lag-phase for hyponastic growth under water, but hardly affected the amplitude of the response.12Together, this indicates that auxin is not always a prerequisite for differential growth responses. Based on the apparent contrasting effects of auxin perception and PAT in low light- and ethylene-induced hyponastic growth, we conclude that ethylene and low light induce hyponastic growth, at least partly, via separate signaling routes.  相似文献   

20.
The prion hypothesis13 states that the prion and non-prion form of a protein differ only in their 3D conformation and that different strains of a prion differ by their 3D structure.4,5 Recent technical developments have enabled solid-state NMR to address the atomic-resolution structures of full-length prions, and a first comparative study of two of them, HET-s and Ure2p, in fibrillar form, has recently appeared as a pair of companion papers.6,7 Interestingly, the two structures are rather different: HET-s features an exceedingly well-ordered prion domain and a partially disordered globular domain. Ure2p in contrast features a very well ordered globular domain with a conserved fold, and—most probably—a partially ordered prion domain.6 For HET-s, the structure of the prion domain is characterized at atomic-resolution. For Ure2p, structure determination is under way, but the highly resolved spectra clearly show that information at atomic resolution should be achievable.Key words: prion, NMR, solid-state NMR, MAS, structure, Ure2p, HET-sDespite the large interest in the basic mechanisms of fibril formation and prion propagation, little is known about the molecular structure of prions at atomic resolution and the mechanism of propagation. Prions with related properties to the ones responsible for mammalian diseases were also discovered in yeast and funghi8,9 which provide convenient model system for their studies. Prion proteins described include the mammalian prion protein PrP, Ure2p,10 Rnq1p,11 Sup35,12 Swi1,13 and Cyc8,14 from bakers yeast (S. cervisiae) and HET-s from the filamentous fungus P. anserina. The soluble non-prion form of the proteins characterized in vitro is a globular protein with an unfolded, dynamically disordered N- or C-terminal tail.1518 In the prion form, the proteins form fibrillar aggregates, in which the tail adopts a different conformation and is thought to be the dominant structural element for fibril formation.Fibrills are difficult to structurally characterize at atomic resolution, as X-ray diffraction and liquid-state NMR cannot be applied because of the non-crystallinity and the mass of the fibrils. Solid-state NMR, in contrast, is nowadays well suited for this purpose. The size of the monomer, between 230 and 685 amino-acid residues for the prions of Figure 1, and therefore the number of resonances in the spectrum—that used to be large for structure determination—is now becoming tractable by this method.Open in a separate windowFigure 1Prions identified today and characterized as consisting of a prion domain (blue) and a globular domain (red).Prion proteins characterized so far were found to be usually constituted of two domains, namely the prion domain and the globular domain (see Fig. 1). This architecture suggests a divide-and-conquer approach to structure determination, in which the globular and prion domain are investigated separately. In isolation, the latter, or fragments thereof, were found to form β-sheet rich structures (e.g., Ure2p(1-89),6,19 Rnq1p(153-405)20 and HET-s(218-289)21). The same conclusion was reached by investigating Sup35(1-254).22 All these fragements have been characterized as amyloids, which we define in the sense that a significant part of the protein is involved in a cross-beta motif.23 An atomic resolution structure however is available presently only for the HET-s prion domain, and was obtained from solid-state NMR24 (vide infra). It contains mainly β-sheets, which form a triangular hydrophobic core. While this cross-beta structure can be classified as an amyloid, its triangular shape does deviate significantly from amyloid-like structures of smaller peptides.23Regarding the globular domains, structures have been determined by x-ray crystallography (Ure2p25,26 and HET-s27), as well as NMR (mammal prions15,2830). All reveal a protein fold rich in α-helices, and dimeric structures for the Ure2 and HET-s proteins. The Ure2p fold resembles that of the β-class glutathione S-transferases (GST), but lacks GST activity.25It is a central question for the structural biology of prions if the divide-and-conquer approach imposed by limitations in current structural approaches is valid. Or in other words: can the assembly of full-length prions simply be derived from the sum of the two folds observed for the isolated domains?  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