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.
Organelle movement in plants is dependent on actin filaments with most of the organelles being transported along the actin cables by class XI myosins. Although chloroplast movement is also actin filament-dependent, a potential role of myosin motors in this process is poorly understood. Interestingly, chloroplasts can move in any direction and change the direction within short time periods, suggesting that chloroplasts use the newly formed actin filaments rather than preexisting actin cables. Furthermore, the data on myosin gene knockouts and knockdowns in Arabidopsis and tobacco do not support myosins'' XI role in chloroplast movement. Our recent studies revealed that chloroplast movement and positioning are mediated by the short actin filaments localized at chloroplast periphery (cp-actin filaments) rather than cytoplasmic actin cables. The accumulation of cp-actin filaments depends on kinesin-like proteins, KAC1 and KAC2, as well as on a chloroplast outer membrane protein CHUP1. We propose that plants evolved a myosin XI-independent mechanism of the actin-based chloroplast movement that is distinct from the mechanism used by other organelles.Key words: actin, Arabidopsis, blue light, kinesin, myosin, organelle movement, phototropinOrganelle movement and positioning are pivotal aspects of the intracellular dynamics in most eukaryotes. Although plants are sessile organisms, their organelles are quickly repositioned in response to fluctuating environmental conditions and certain endogenous signals. By and large, plant organelle movements and positioning are dependent on actin filaments, although microtubules play certain accessory roles in organelle dynamics.1,2 Actin inhibitors effectively retard the movements of mitochondria,36 peroxisomes,5,711 Golgi stacks,12,13 endoplasmic reticulum (ER),14,15 and nuclei.1618 These organelles are co-aligned and associated with actin filaments.5,7,8,1012,15,18 Recent progress in this field started to reveal the molecular motility system responsible for the organelle transport in plants.19Chloroplast movement is among the most fascinating models of organelle movement in plants because it is precisely controlled by ambient light conditions.20,21 Weak light induces chloroplast accumulation response so that chloroplasts can capture photosynthetic light efficiently (Fig. 1A). Strong light induces chloroplast avoidance response to escape from photodamage (Fig. 1B).22 The blue light-induced chloroplast movement is mediated by the blue light receptor phototropin (phot). In some cryptogam plants, the red light-induced chloroplast movement is regulated by a chimeric phytochrome/phototropin photoreceptor neochrome.2325 In a model plant Arabidopsis, phot1 and phot2 function redundantly to regulate the accumulation response,26 whereas phot2 alone is essential for the avoidance response.27,28 Several additional factors regulating chloroplast movement were identified by analyses of Arabidopsis mutants deficient in chloroplast photorelocation.2932 In particular, identification of CHUP1 (chloroplast unusual positioning 1) revealed the connection between chloroplasts and actin filaments at the molecular level.29 CHUP1 is a chloroplast outer membrane protein capable of interacting with F-actin, G-actin and profilin in vitro.29,33,34 The chup1 mutant plants are defective in both the chloroplast movement and chloroplast anchorage to the plasma membrane,22,29,33 suggesting that CHUP1 plays an important role in linking chloroplasts to the plasma membrane through the actin filaments. However, how chloroplasts move using the actin filaments and whether chloroplast movement utilizes the actin-based motility system similar to other organelle movements remained to be determined.Open in a separate windowFigure 1Schematic distribution patterns of chloroplasts in a palisade cell under different light conditions, weak (A) and strong (B) lights. Shown as a side view of mid-part of the cell and a top view with three different levels (i.e., top, middle and bottom of the cell). The cell was irradiated from the leaf surface shown as arrows. Weak light induces chloroplast accumulation response (A) and strong light induces the avoidance response (B).Here, we review the recent findings pointing to existence of a novel actin-based mechanisms for chloroplast movement and discuss the differences between the mechanism responsible for movement of chloroplasts and other organelles.  相似文献   

2.
The newly defined phytohormones strigolactones (SLs) were recently shown to act as regulators of root development. Their positive effect on root-hair (RH) elongation enabled examination of their cross talk with auxin and ethylene. Analysis of wild-type plants and hormone-signaling mutants combined with hormonal treatments suggested that SLs and ethylene regulate RH elongation via a common regulatory pathway, in which ethylene is epistatic to SLs. The SL and auxin hormonal pathways were suggested to converge for regulation of RH elongation; this convergence was suggested to be mediated via the ethylene pathway, and to include regulation of auxin transport.Key words: strigolactone, auxin, ethylene, root, root hair, lateral rootStrigolactones (SLs) are newly identified phytohormones that act as long-distance shoot-branching inhibitors (reviewed in ref. 1). In Arabidopsis, SLs have been shown to be regulators of root development and architecture, by modulating primary root elongation and lateral root formation.2,3 In addition, they were shown to have a positive effect on root-hair (RH) elongation.2 All of these effects are mediated via the MAX2 F-box.2,3In addition to SLs, two other plant hormones, auxin and ethylene, have been shown to affect root development, including lateral root formation and RH elongation.46 Since all three phytohormones (SLs, auxin and ethylene) were shown to have a positive effect on RH elongation, we examined the epistatic relations between them by examining RH length.7 Our results led to the conclusion that SLs and ethylene are in the same pathway regulating RH elongation, where ethylene may be epistatic to SLs.7 Moreover, auxin signaling was shown to be needed to some extent for the RH response to SLs: the auxin-insensitive mutant tir1-1,8 was less sensitive to SLs than the wild type under low SL concentrations.7On the one hand, ethylene has been shown to induce the auxin response,912 auxin synthesis in the root apex,11,12 and acropetal and basipetal auxin transport in the root.4,13 On the other, ethylene has been shown to be epistatic to SLs in the SL-induced RH-elongation response.7 Therefore, it might be that at least for RH elongation, SLs are in direct cross talk with ethylene, whereas the cross talk between SL and auxin pathways may converge through that of ethylene.7 The reduced response to SLs in tir1-1 may be derived from its reduced ethylene sensitivity;7,14 this is in line with the notion of the ethylene pathway being a mediator in the cross talk between the SL and auxin pathways.The suggested ethylene-mediated convergence of auxin and SLs may be extended also to lateral root formation, and may involve regulation of auxin transport. In the root, SLs have been suggested to affect auxin efflux,3,15 whereas ethylene has been shown to have a positive effect on auxin transport.4,13 Hence, it might be that in the root, the SLs'' effect on auxin flux is mediated, at least in part, via the ethylene pathway. Ethylene''s ability to increase auxin transport in roots was associated with its negative effect on lateral root formation: ethylene was suggested to enhance polar IAA transport, leading to alterations in the quantity of auxin that unloads into the tissues to drive lateral root formation.4 Under conditions of sufficient phosphate, SL''s effect was similar to that of ethylene: SLs reduced the appearance of lateral roots; this was explained by their ability to change auxin flux.3 Taken together, one possibility is that the SLs'' ability to affect auxin flux and thereby lateral root formation in the roots is mediated by induction of ethylene synthesis.To conclude, root development may be regulated by a network of auxin, SL and ethylene cross talk.7 The possibility that similar networks exist elsewhere in the SLs'' regulation of plant development, including shoot architecture, cannot be excluded.  相似文献   

3.
4.
5.
6.
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   

7.
8.
The chemical cross talk between rice and barnyardgrass which is one of the most noxious weeds in rice cultivation was investigated. Allelopathic activity of rice was increased by the presence of barnyardgrass seedlings or barnyardgrass root exudates. Rice allelochemical, momilactone B, concentration in rice seedlings and momilactone B secretion level from rice were also increased by the presence of barnyardgrass seedlings or barnyardgrass root exudates. As momilactone B possesses strong growth inhibitory activity and acts as an allelochemical, barnyardgrass-induced rice allelopathy may be due to the increased momilactone B secretion. These results suggest that rice may respond to the presence of neighboring barnyardgrass by sensing the chemical components in barnyardgrass root exudates and increase allelopathic activity by elevated production and secretion levels of momilactone B. Thus, rice allelopathy may be one of the inducible defense mechanisms by chemical-mediated plant interaction between rice and barnyardgrass and the induced-allelopathy may provide a competitive advantage for rice through suppression of the growth of barnyardgrass.Key words: allelopathy, Echinochloa, chemical interaction, induced-allelopathy, momilactone, Oryza sativaThe chemical cross talk between host and symbiotic or parasitic plants is an essential process for the development of physical connections in symbiosis and parasitism.13 Barnyardgrass is one of the most common and noxious weeds in rice paddy fields.4 Although barnyardgrass is adapted rice production system due to its similarity in growth habit, the reason why barnyardgrass so often invades into the rice paddy fields is unknown. There might be some special interactions between both plant species.Plants are able to accumulate phytoalexins around infection sites of pathogens soon after sensing elicitors of pathogen origin. This accumulation of phytoalexins can protect the plants from further pathogen infection.5,6 Plants are also able to activate defense mechanisms against attacking herbivores by sensing volatile compounds, such as methacrolein and methyl jasmonate, released by herbivore-attacked plant cells. The volatile-sensed plants increase the production of phenolics, alkaloids, terpenes and defense proteins, which reduce herbivory attacks.7,8 Therefore, plants are able to elevate the defense mechanisms against several biotic stress conditions by detection of various compounds.Allelopathy is the direct influence of organic chemicals released from plants on the growth and development of other plants.911 Allelochemicals are such organic chemicals involved in the allelopathy.12,13 Allelochemicals can provide a competitive advantage for host-plants through suppression of soil microorganism and inhibition of the growth of competing plant species because of their antibacterial, antifungal and growth inhibitory activities.3,14,15Rice has been extensively studied with respect to its allelopathy as part of a strategy for sustainable weed management, such as breeding allelopathic rice strains. A large number of rice varieties were found to inhibit the growth of several plant species when these rice varieties were grown together with these plants under the field or/and laboratory conditions.1620 These findings suggest that rice may produce and release allelochemicals into the neighboring environments and may inhibit the growth of the neighboring plants by the allelochemicals.Potent allelochemical, momilactone B, was isolated from rice root exudates.21 Momilactone B inhibits the growth of typical rice weeds like barnyardgrass and Echinochloa colonum at concentrations greater than 1 µM and the toxicity of momilactone B to rice itself was very low.22 In addition, rice plants secrete momilactone B from the roots into the rhizosphere over their entire life cycle.22 The observations suggest rice allelopathy may be primarily dependant on the secretion levels of momilactone B from the rice seedlings.22,23Allelopathic activity of rice exhibited 5.3- to 6.3-fold increases when rice and barnyardgrass seedlings were grown together. Root exudates of barnyardgrass seedlings also increased allelopathic activity and momilactone B concentration in rice seedlings. The increasing the exudate concentration increased the allelopathic activity and momilactone B concentration in rice.24 Thus, the chemical components in barnyardgrass root exudates may affect gene expressions involved in momilactone B biosynthesis. However, effects of the barnyardgrass root exudates on the secretion level of mimilactone B from rice has not yet reported.Rice seedlings were incubated in the medium containing barnyardgrass root exudates for 10 d, and secretion level of momilactone B by rice was determined (Fig. 1). The root exudates increased the secretion level significantly at concentrations greater than 30 mg/L of barnyardgrass root exudates, and increasing the concentration increased the secretion level. At concentrations of 300 mg/L of the root exudates, the secretion level was 10-fold greater than that in control (0 mg of root exudate). There was no significant difference in the osmotic potential between the medium contained barnyardgrass root exudates and control medium (all about 10 mmol/kg), and pH value of the medium was maintained at 6.0 throughout the experiments.25 These results suggest that unknown chemical components in the barnyardgrass root exudates may induce the secretion of momilactone B from rice. As momilactone B possesses strong phytotoxic and allelopathic activities,2123,25 the elevated production and secretion of momilactone B in rice may provide a competitive advantage for root establishment through local suppression of pathogens and inhibition of the growth of competing plant species including barnyardgrass. Thus, barnyardgrass-induced rice allelopathy may be caused by the chemical components in the barnyardgrass root exudates.Open in a separate windowFigure 1Effects of barnyardgrass root exudates on momilactone B secretion level in rice. Rice seedlings were incubated in the medium containing barnyardgrass root exudates for 10 d, and secretion level of momilactone B was determined as described by Kato-Noguchi.24 The experiment was repeated six times with three assays for each determination. Different letters show significant difference (p < 0.01) according to Tukey''s HSD test.Although mechanisms of the exudation are not well understood, it is suggested that plants are able to secrete a wide variety of compounds from root cells by plasmalemma-derived exudation, endoplasmic-derived exudation and proton-pumping mechanisms.3,15 Through the root exudation of compounds, plants are able to regulate the soil microbial community in their immediate vicinity, change the chemical and physical properties of the soil, and inhibit the growth of competing plant species.3,14,15 The present research suggests that rice may be aware of the presence of neighboring barnyardgrass by detection of certain key in barnyardgrass root exudates, and this sensorial function may trigger a signal cascade resulting in increasing rice allelopathy through increasing production of momilactone B and secretion of momilactone B into the rhizosphere. Therefore, rice allelopathy may potentially be an inducible defense mechanism by chemical-mediated plant interactions between rice and barnyardgrass.  相似文献   

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

11.
12.
13.
14.
Strigolactones (SLs) have been recently identified as a new group of plant hormones or their derivatives thereof, shown to play a role in plant development. Evolutionary forces have driven the development of mechanisms in plants that allow adaptive adjustments to a variety of different habitats by employing plasticity in shoot and root growth and development. The ability of SLs to regulate both shoot and root development suggests a role in the plant''s response to its growth environment. To play this role, SL pathways need to be responsive to plant growth conditions, and affect plant growth toward increased adaptive adjustment. Here, the effects of SLs on shoot and root development are presented, and possible feedback loops between SLs and two environmental cues, light and nutrient status, are discussed; these might suggest a role for SLs in plants'' adaptive adjustment to growth conditions.Key words: strigolactones, light, nutrient status, root, shoot, branching, lateral roots, root hairsStrigolactones (SLs) are carotenoid-derived terpenoid lactones suggested to stem from the carotenoid pathway1 via the activity of various oxygenases.2,3 SLs production has been demonstrated in both monocotyledons and eudicotyledons (reviewed in ref. 4), suggesting their presence in many plant species.5 SLs are synthesized mainly in the roots and in some parts of the stem and then move towards the shoot apex (reviewed ref. 7).6,8,9SLs were first characterized more than 40 years ago as germination stimulants of the parasitic plants Striga and Orobanche and later, as stimulants of arbuscular mycorrhiza hyphal branching as well (reviewed in ref. 4, 1013). Recently, SLs or derivatives thereof, have been identified as a new group of plant hormones, shown to play a role in inhibition of shoot branching,2,3,8,9 thereby affecting shoot architecture; more recently they have also been shown to affect root growth by affecting auxin efflux.14Plants have developed mechanisms that allow adaptive adjustments to a variety of different habitats by employing plasticity in their growth and development.15 Shoot architecture is affected by environmental cues, such as light quality and quantity and nutrient status.1619 Root-system architecture and development are affected by environmental conditions such as nutrient availability (reviewed in ref. 20, 21). At the same time, plant hormones are known to be involved in the regulation of plant growth, development and architecture (reviewed in ref. 2224) and to be mediators of the effects of environmental cues on plant development; one classic example is auxin''s role in the plant''s shade-avoidance response (reviewed in ref. 25).The ability of SLs to regulate shoot and root development suggests that these phytohormones also have a role in the plant''s growth response to its environment. To play this putative role, SL pathways need to be responsive to plant growth conditions, and affect plant growth toward enhancing its adaptive adjustment. The present review examines the SLs'' possible role in adaptive adjustment of the plant''s response to growth conditions, by discussing their effect on plant development and the possible associations and feedback loops between SLs and two environmental cues: light and nutrient status.  相似文献   

15.
16.
17.
Environmental and developmental signals can elicit differential activation of membrane proton (H+) fluxes as one of the primary responses of plant and fungal cells. In recent work,1 we could determine that during the presymbiotic growth of arbuscular mycorrhizal (AM) fungi specific domains of H+ flux are activated by clover root factors, namely host root exudates or whole root system. Consequently, activation on hyphal growth and branching were observed and the role of plasma membrane H+-ATPase was investigated. The specific inhibitors differentially abolished most of hyphal H+ effluxes and fungal growth. As this enzyme can act in signal transduction pathways, we believe that spatial and temporal oscillations of the hyphal H+ fluxes could represent a pH signature for both early events of the AM symbiosis and fungal ontogeny.Key words: H+-specific vibrating probe, pH signatures, arbuscular mycorrhiza, pH signalling, Gigaspora margaritaThe 450-million-year-old symbiosis between the majority of land plants and arbuscular mycorrhizal (AM) fungi is one of the most ancient, abundant and ecologically important symbiosis on Earth.2,3The development of AM interaction starts before the physical contact between the host plant roots and the AM fungus. The hyphal growth and branching are induced by the root factors exudated by host plants, followed by the formation of appressorium leading to the hyphal penetration in the root system. These root factors seems to be specifically synthesized by host plants, since exudates from non-host plants are not able to promote neither hyphal differentiation nor appressorium formation.4,5 Most root exudates contain several host signals or better, active compounds including flavonoids6,19 and strigolactones,7,8 however many of them are not yet known.Protons (H+) may have an important role on the fungal growth and host signal perception.1 In plant and fungal cells, H+ can be pumped out through two different mechanisms: (1) the activity of the P-type plasma membrane (PM) H+-ATPase9 and (2) PM redox reactions.10 The proportional contribution from both mechanisms is not known, but in most plant cells the PM H+-ATPase seems to be the major responsible by the H+ efflux across plasma membrane. AM Fungal cells also energize their PM using P-type H+-pumps quite similar to the plant ones. Indeed, some genes codifying isoforms of P-type H+-ATPase have been isolated of AM fungi,1113 and AM fungal ATP hydrolysis activity was shown by cytochemistry, localized mainly in the first 70 µm from the germ tube tip.14 This structural evidence correlates with data obtained by H+-specific vibrating probe (Fig. 1A and B), which indicates that the H+ efflux in Gigaspora margarita is more intense in the subapical region of the lateral hyphae1 (Fig. 1A). Furthermore, the correlation between the cytosolic pH profile previously obtained by Jolicoeur et al.,15 with the H+ efflux pattern (erythrosine-dependent), seems to clearly indicate that an active PM H+-ATPase takes place at the subapical hyphal region. Using orthovanadate, we could show that those H+ effluxes are susceptible mainly in the subapical region, but no effect in the apical was found.1 Recently, a method to use fluorescent marker expression in an AM fungus driven by arbuscular mycorrhizal promoters was published.31 It could be adjusted as an alternative to measure “in vivo” PM H+-ATPase expression in AM fungal hyphae and their responses to root factors.31Open in a separate windowFigure 1(A) H+ flux profile along growing secondary hyphae of G. margarita in the presence (open squares) or absence (closed squares) of erythrosin B and its correlation with cytosolic pH (pHc) data described by Jolicoeur et al.,15 (dotted line). Dotted area depicts the region with higher susceptibility to erythrosin B. (B) ion-selective electrode near to AM fungal hyphae. (C) Stimulation on hyphal H+ efflux after incubation with root factors or whole root system. R, roots; RE, root exudates; CO2, carbon dioxide; CWP, cell wall proteins; GR24, synthetic strigolactone. The medium pH in all treatment was monitored and remained about 5.7, including with prior CO2 incubation. Means followed by the same letter are statistically equal by Duncan''s test at p < 5%.The H+ electrochemical gradient generated by PM H+-ATPases provides not only driving force for nutrient uptake,9,16 but also can act as an intermediate in signal transduction pathways.18 The participation of these H+ pumps in cell polarity and tip growth of plant cells was recently reported,27 addressing their crucial role on apical growth.28 Naturally, in the absence of root factors the AM fungi have basal metabolic8,2123 and respiratory activity.24 However when root signals are recognized and processed by AM fungal cells they might become activated.22 We thus searched for pH signatures that could reflect the alterations on fungal metabolism in response to external stimuli. In fact, preliminary analyses from our group demonstrate that AM fungal hyphae increase their H+ efflux in response not only to root exudates recognition, but also to other root factors (Fig. 1C). The incubation for 30 min of AM fungal hyphae with several root factors induces hyphal H+ efflux similar to the response to intact root system (5 days of incubation). The major increases were found with 1% CO2 (750%) followed by root cell wall proteins (221%), root exudates (130%) and synthetic strigolactone (5%) (Fig. 1C). Those stimulations could define the transition from the state without root signals to the presymbiotic developmental stage (Fig. 1C). In the case of CO2, the incorporation of additional carbon could represent a new source of energy, since CO2 dark fixation takes place in Glomus intraradices germ tubes.22,25Interestingly, after the treatment with synthetic strigolactone (10−5 M GR24), no significant stimulation was found compared to the remaining factors (Fig. 1C). It opens the question if the real effect of strigolactone is restrict to hyphal branching and does not intervene in very fast response pathways. Likewise, strigolactones need additional time to exhibit an effect, as recently discussed by Steinkellner et al.,26 However, at the moment, no comprehensive electrophysiological analyses are presently available separating the effects of strigolactone and some flavonoids in AM fungal hyphae.The next target of our work is the study of ionic responses of single germ tubes or primary hyphae to root factors (Fig. 2). As reported by Ramos et al.,1 we have been observing that the pattern of ion fluxes at the apical zone of primary hyphae is differentiated from secondary or lateral hyphae. In the primary, two interesting responses were detected in the absence of root factors: (1) a “dormant Ca2+ flux” and (2) Cl or anion fluxes at the same direction of H+ ions, suggesting a possible presence of H+/Cl symporters at the apex, similarly to what occurs in root hairs (Fig. 2).30 In the presence of root factors such as root exudates the stimulated influxes of Cl (anion), H+, Na+ and effluxes of K+ and Ca2+ are activated. It can explain why the AM fungi hyphal tips are depolarized20,29 during the period without root signals—“asymbiosis”—as long as K+ efflux and H+ influx occur simultaneously. Indeed, H+ as well as Ca2+ ions may act as second messengers, where extra and intracellular transient pH changes are preconditions for a number of processes, including gravity responses and possibly in plant-microbe interactions.17,30Open in a separate windowFigure 2Ion dynamics in the apex of primary hyphae of arbuscular mycorrhizal fungi. It represents the Stage 1 described in Ramos et al.1 After treatment with root factors, an activation of Ca2+ efflux is observed at the hyphal apex.Clearly, further data on the mechanism of action of signaling molecules such as strigolactones over the signal transduction and ion dynamics in AM fungi will be very important to improve our understanding of the molecular bases of the mycorrhization process. Future studies are necessary in order to provide basic knowledge of the ion signaling mechanisms and their role on the response of very important molecules playing at the early events of AM symbiosis.  相似文献   

18.
19.
Determination of the subcellular localization of an unknown protein is a major step towards the elucidation of its function. Lately, the expression of proteins fused to fluorescent markers has been very popular and many approaches have been proposed to express these proteins. Stable transformation using Agrobacterium tumefaciens generates stable lines for downstream experiments, but is time-consuming. If only colocalization is required, transient techniques save time and effort. Several methods for transient assays have been described including protoplast transfection, biolistic bombardment, Agrobacterium tumefaciens cocultivation and infiltration. In general colocalizations are preferentially performed in intact tissues of the same species, resembling the native situation. High transformation rates were described for cotyledons of Arabidopsis, but never for roots. Here we report that it is possible to transform Arabidopsis root epidermal cells with an efficiency that is sufficient for colocalization purposes.Key words: Arabidopsis, GFP-fusions, protein localization, root, transient transformationSince the release of the Arabidopsis thaliana genome sequence plant biologists set the goal to elucidate the functions of all coded genes. Apart from the spatio-temporal expression patterns of genes, the subcellular localization of gene products can play an essential role in deciphering their function. Classical immunological approaches to localize proteins can be hindered by cross-reactivity, time-consuming generation of antibodies and the low temporal resolution. Expression of tagged proteins forms a suitable alternative. Lately, fusions with fluorescent proteins in combination with confocal (CLSM)1 or spinning disc microscopy2 allow real time protein localization and even subcellular trafficking at high resolution. An overview of fluorescent tagging approaches can be found elsewhere.3Currently several techniques to introduce the coding region for a tagged protein in a plant are available. The generation of stable lines transformed by Agrobacterium tumefaciens offers a continuous source of plant material, but it is time-consuming especially when only colocalization experiments are required. Transient assays, on the other hand, offer the advantage of being fast and amenable to high throughput strategies. Each of these techniques, however, has some limitations and drawbacks. Particle bombardment (biolistics) 46 for example circumvents the host specificity of Agrobacterium strains, but requires expensive equipment. Moreover, it is rather disruptive and imposes a significant stress upon the plants, possibly influencing the results. Protoplasts lack a cell wall and protoplast transformation7,8 is therefore not suitable for certain experiments related to cell wall proteins or when interactions between cells on tissue level might be important.9 Moreover, protoplasts have lost their identity which might be critical for the correct functioning of certain transgenic constructs. Agrobacterium infiltration of tobacco leaves10 is regularly used and represents an efficient, fast and relatively easy transformation technique. However, tobacco leaves easily show autofluoresence due to tissue damage as a result of experimental manipulations. As it has been reported that some protein fusions expressed in an heterologous system localize to different subcellular localizations11 it is advisable not to use tobacco when localizing Arabidopsis proteins. Leaf infiltrations have been performed in Arabidopsis,12 but apparently their leaves are much more prone to mechanical damage and the leaf developmental stage is critical, complicating this technique. Cocultivation of Agrobacterium with seedlings offers a rapid and efficient approach applicable to many mono and dicot species. It was reported to work efficiently in Arabidopsis cotyledons, but not in roots.9 As an alternative method, Agrobacterium infiltration of Arabidopsis seedlings11 seems an efficient technique for transient expression. However, expression in root cells could not be obtained. Colocalizations are required in the native cells or tissue for the correct localization of an unknown protein or proteins that need interaction partners. As a consequence this technique can not be reliably used when root expressed gene products are studied. Here we show evidence that it is possible to use the described technique11 to induce transient expression in Arabidopsis roots.We used the Agrobacterium infiltration of Arabidopsis seedlings technique11 to colocalize several C-terminal (S65T)-sGFP fusions generated in the plant binary vector pGWB6.13 Each construct was transformed into Agrobacterium tumefaciens (C59C1RifR) containing the helper plasmid pMP90. Subsequently different stable marker lines, wild type Arabidopsis (Col-0) bearing mCherry fusion constructs,14 were transiently transformed.11 After 2 or 3 days seedlings were studied using CLSM. Besides being expressed in cotyledons fusion proteins were clearly observed in root epidermis and root cap cells (Fig. 1A and B). As reported11 the transformation efficiency in cotyledons was considerably higher than in root cells. However, in each experiment we obtained a considerable amount of transformed root epidermal cells which was more than sufficient for colocalization studies (Fig. 2). It was remarkable that transformation was repeatedly successful in groups of cells, adjacent or close to each other.Open in a separate windowFigure 1Transient transformation of Arabidopsis root cells. Expression of the protein-GFP fusion product can be seen in the epidermal (A) and root cap cells (B) on fluorescence/transmission merged images. As seen in (A) high efficiencies of root transformation can be reached.Open in a separate windowFigure 2Colocalization of mCherry and GFP constructs. Confocal image of the mCherry fluorescence (A), the GFP signal (B) and the merged image (C).In contrast to what was reported earlier we show here that the Agrobacterium infiltration technique11 is perfectly capable of transiently transforming Arabidopsis root epidermal cells. It allows the transient production and study of proteins in their native environment, considerably increasing the reliability of such experiments. Additionaly the use of RFP marker constructs in colocalisation studies in the root is free of interference by the red background autofluorescence of chlorophyll.  相似文献   

20.
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   

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