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 32 毫秒
1.
Plant hormone conjugation: A signal decision   总被引:1,自引:0,他引:1  
Tight regulation of the auxin hormone indole-3-acetic acid (IAA) is crucial for plant development. Newly discovered IAA antagonists are the amide-linked tryptophan conjugates of IAA and jasmonic acid (JA). JA-Trp and IAA-Trp interfered with root gravitropism in Arabidopsis, and inhibited several responses to exogenously supplied IAA. Relatively low concentrations of the inhibitors occurred in Arabidopsis, but Pisum sativum flowers contained over 300 pmole g−1 FW of JA-Trp. DihydroJA was an even more effective inhibitor than JA-Trp, suggesting that Trp conjugates with other JA derivatives may also be functional. JA-Trp and IAA-Trp add to the list of documented bioactive amide hormone conjugates. The only other example is JA-Ile, the recently discovered jasmonate signal. These examples establish that conjugation not only inactivates hormones, but in some cases creates novel compounds that function in hormone signaling.Key words: jasmonic acid, indole-3-acetic acid, auxin, tryptophan, conjugate, plant hormone, signaling, amino acid, antagonistPlants hold an amazing capacity to auto-regulate their growth and respond to a host of environmental challenges. Since the early discovery of the first plant hormone, indole-3-acetic acid (IAA),1 science has progressively unveiled ever more complex, and sometimes surprising, ways that plants manipulate hormones to optimize their growth and thwart their opponents. Until recently, the covalent coupling of hormones to sugars, amino acids and peptides was thought to be merely a way to dispose of excess hormone.2 The amide linkage of IAA to Asp and Glu does indeed result in IAA catabolism, while IAA-Ala and IAA-Leu are inactive stored forms of IAA.3 But the perception that all hormone conjugates are inactive changed abruptly with the discovery that the isoleucine conjugate of jasmonic acid (JA-Ile) is an active hormonal signal.  相似文献   

2.
The regulation of cellular auxin levels is a critical factor in determining plant growth and architecture, as indole-3-acetic acid (IAA) gradients along the plant axis and local IAA maxima are known to initiate numerous plant growth responses. The regulation of auxin homeostasis is mediated in part by transport, conjugation and deconjugation, as well as by de novo biosynthesis. However, the pathways of IAA biosynthesis are yet not entirely characterized at the molecular and biochemical level. It is suggested that several biosynthetic routes for the formation of IAA have evolved. One such pathway proceeds via the intermediate indole-3-acetamide (IAM), which is converted into IAA by the activity of specific IAM hydrolases, such as Arabidopsis AMIDASE1 (AMI1). In this article we present evidence to support the argument that AMI1-dependent IAA synthesis is likely not to be used during the first two days of seedling development.Key words: Arabidopsis thaliana, auxin biosynthesis, AMIDASE1, indole-3-acetic acid, indole-3-acetamide, LEAFY COTYLEDON1, seed developmentAuxins are versatile plant hormones that play diverse roles in regulating many aspects of plant growth and development.1 To enable auxins to develop their activity, a tight spatiotemporal control of cellular indole-3-acetic acid (IAA) contents is absolutely necessary since it is well-documented that auxin action is dose dependent, and that high IAA levels can have inhibitory effects on plant growth.2 To achieve this goal, plants have evolved a set of different mechanisms to control cellular hormone levels. On the one hand, plants possess several pathways that contribute to the de novo synthesis of IAA. This multiplicity of biosynthetic routes presumably facilitates fine-tuning of the IAA production. On the other hand, plants are equipped with a variety of enzymes that are used to conjugate free auxin to either sugars, amino acids or peptides and small proteins, respectively, or on the contrary, that act as IAA-conjugate hydrolases, releasing free IAA from corresponding conjugates. IAA-conjugates serve as a physiologically inactive storage form of IAA from which the active hormone can be quickly released on demand. Alternatively, conjugation of IAA can mark the first step of IAA catabolism. In general, conjugation and deconjugation of free IAA are ways to positively or negatively affect active hormone levels, which adds another level of complexity to the system. Additionally, IAA can be transported from cell to cell in a polar manner, which is dependent on the action of several transport proteins. All together, these means are used to form auxin gradients and local maxima that are essential to initiate plant growth processes, such as root or leaf primordia formation.3  相似文献   

3.
Auxin is an indispensable hormone throughout the lifetime of nearly all plant species. Several aspects of plant growth and development are rigidly governed by auxin, from micro to macro hierarchies; auxin also has a close relationship with plant-pathogen interactions. Undoubtedly, precise auxin levels are vitally important to plants, which have many effective mechanisms to maintain auxin homeostasis. One mechanism is conjugating amino acid to excessive indole-3-acetic acid (IAA; main form of auxin) through some GH3 family proteins to inactivate it. Our previous study demonstrated that GH3-2 mediated broad-spectrum resistance in rice (Oryza sativa L.) by suppressing pathogen-induced IAA accumulation and downregulating auxin signaling. Here, we further investigated the expression pattern of GH3-2 and other GH3 family paralogs in the life cycle of rice and presented the possible function of GH3-2 on rice root development by histochemical analysis of GH3-2 promoter:GUS reporter transgenic plants.Key words: auxin, GH3 gene, indole-3-acetic acid, Oryza sativa, rootThe phytohormone auxin regulates tropism and organ development and influences phyllotaxis, vascular canalization and root patterning by exerting its effect on cell division, elongation and differentiation in plants.1,2 Indole-3-acetic acid (IAA) is the most widespread form of auxin in most plants. Supraoptimal or insufficient concentration of auxin will cause plants to exhibit abnormal phenotypes. 3-9 Auxin homeostasis is partly sustained by the GH3 gene family, a supervisor of the fluctuation of auxin. Most GH3 genes contain auxin-responsive cis-acting elements (AuxRE) in their promoter regions and react rapidly and transiently to auxin signaling.1 Nineteen GH3 paralogs have been discovered in Arabidopsis.10 According to the phylogenetic relationship and acyl acid substrate preference, these genes are classified into three groups (I, II and III), which catalyze the formation of jasmonates, salicylic acid, 4-substituted benzoates or IAA acyl acid amido conjugates.11,12 The rice GH3 gene family includes 13 paralogs, 4 belonging to group I (GH3-3, -5, -6 and -12) and 9 to group II (GH3-1, -2, -4, -7, -8, -9, -10, -11 and -13); group III GH3 is absent in rice.10 Rice GH3-1, -2, -8 and -13 paralogs have been biochemically confirmed to have IAA-amido synthetase activity by in vivo or in vitro assays.69 It is believed that other GH3 group II paralogs in rice may also possess this enzymatic activity. But why does rice have such a functionally redundant group of GH3 proteins, which disobeys the economic principle? The explanation could be based on the different temporal and spatial expression of the genes encoding these proteins.  相似文献   

4.
One pathway leading to the bioactive auxin, indole-3-acetic acid (IAA), is known as the tryptamine pathway, which is suggested to proceed in the sequence: tryptophan (Trp), tryptamine, N-hydroxytryptamine, indole-3-acetaldoxime, indole-3-acetaldehyde (IAAld), IAA. Recently, this pathway has been characterized by the YUCCA genes in Arabidopsis (Arabidopsis thaliana) and their homologs in other species. YUCCA is thought to be responsible for the conversion of tryptamine to N-hydroxytryptamine. Here we complement the genetic findings with a compound-based approach in pea (Pisum sativum), detecting potential precursors by gas chromatography/tandem-mass spectrometry. In addition, we have synthesized deuterated forms of many of the intermediates involved, and have used them to quantify the endogenous compounds, and to investigate their metabolic fates. Trp, tryptamine, IAAld, indole-3-ethanol, and IAA were detected as endogenous constituents, whereas indole-3-acetaldoxime and one of its products, indole-3-acetonitrile, were not detected. Metabolism experiments indicated that the tryptamine pathway to IAA in pea roots proceeds in the sequence: Trp, tryptamine, IAAld, IAA, with indole-3-ethanol as a side-branch product of IAAld. N-hydroxytryptamine was not detected, but we cannot exclude that it is an intermediate between tryptamine and IAAld, nor can we rule out the possibility of a Trp-independent pathway operating in pea roots.Auxin is a key plant growth hormone, involved in processes as diverse as branching, gravitropism, phototropism, and seed development (Davies, 2004). However, the biosynthetic pathways leading to the main auxin in plants, indole-3-acetic acid (IAA), are not well understood. Although there is good evidence that the amino acid Trp is an early precursor (Gibson et al., 1972; Wright et al., 1991; Tsurusaki et al., 1997), several routes from Trp to IAA have been proposed, and for any given species it is not clear which route or routes occur. The possible Trp-dependent pathways in higher plants are the indole-3-pyruvic acid (IPyA) pathway (Stepanova et al., 2008; Tao et al., 2008), the tryptamine (YUCCA) pathway (Zhao et al., 2001), the indole-3-acetaldoxime (IAOx) pathway (Bartel et al., 2001), and the indoleacetamide pathway (Pollmann et al., 2002), on the basis of the first metabolite of Trp (Fig. 1). In addition, a possible Trp-independent pathway has been proposed (Normanly et al., 1993), bypassing Trp completely, further complicating the process of IAA biosynthesis.Open in a separate windowFigure 1.Left: The putative tryptamine (red), Trp-independent (light blue), IPyA (green), indoleacetamide (yellow), and IAOx (dark blue) biosynthetic pathways to IAA in Arabidopsis. The steps shown in gray appear not to occur in peas. Right: The simplified pathway scheme suggested to occur in pea based on the present results and Sugawara et al. (2009). N-hydroxytryptamine was not detected as a metabolite in this study, suggesting that tryptamine might be converted directly to IAAld in pea roots. The Trp-independent, indoleacetamide, and IPyA pathways were not studied.Since 2001, there has been renewed interest in the tryptamine route to IAA, after the discovery and functional analysis of the Arabidopsis (Arabidopsis thaliana) YUCCA gene, reported to encode the enzyme for converting tryptamine to N-hydroxytryptamine (Zhao et al., 2001, 2002). On the basis of the Zhao et al. (2001, 2002) reports, tryptamine pathways have generally been proposed in the sequence: Trp, tryptamine, N-hydroxytryptamine, IAOx, indole-3-acetaldehyde (IAAld), IAA (Fig. 1; Woodward and Bartel, 2005). Recently, however, Sugawara et al. (2009) suggested that IAOx be removed from the tryptamine pathway, and that IAOx-dependent IAA biosynthesis operates only in the Brassicaceae. In Arabidopsis, this pathway can be important, at least in some circumstances, because when a side branch is impaired, as in the sur1 mutant, IAA levels increase dramatically (Sugawara et al., 2009). On the other hand, the IAOx pathway is not the only pathway operating in Arabidopsis, because genetically blocking the step Trp to IAOx does not always reduce IAA content, compared with wild-type plants (Sugawara et al., 2009). This means that in Arabidopsis, the tryptamine and/or IPyA and/or indoleacetamide pathways compensate for the loss of the IAOx pathway. Interestingly, Sugawara et al. (2009) do not include IAAld in their tryptamine pathway, and their model implies instead that N-hydroxytryptamine is directly converted to IAA.Turning to other species, it has been reported that tryptamine is not present in pea (Pisum sativum; Schneider et al., 1972), despite being present in tomato (Solanum lycopersicum; Cooney and Nonhebel, 1991), rice (Oryza sativa; Ishihara et al., 2008), Arabidopsis (Sugawara et al., 2009), and barley (Hordeum vulgare; Schneider et al., 1972). In tomato, although tryptamine is relatively abundant, and early metabolism studies indicated the conversion of tryptamine to IAA via IAAld (Schneider et al., 1972), Cooney and Nonhebel (1991) cast doubt on the role of tryptamine after studying patterns of labeling after incubation of plants with deuterated water. Again, in tobacco (Nicotiana tabacum), Songstad et al. (1990) showed that while tobacco plants overexpressing a Trp decarboxylase accumulated very high levels of tryptamine, IAA levels were unaffected. Although this result has been interpreted as evidence against the involvement of tryptamine (Bartel et al., 2001), another explanation is that excess tryptamine is converted to compounds via a side branch or side branches, although these are not well studied. Finally, the compound N-hydroxytryptamine is relatively unknown, with no reports of its presence in plants to date.It is clear, therefore, that the tryptamine pathway to IAA remains poorly understood. In this article, we further characterize the pathway, using the garden pea as a model species. We report on the presence/absence and levels of the putative endogenous intermediates, as determined by gas chromatography/tandem mass spectrometry (GC/MS/MS), and investigate their metabolic fates using [14C] and deuterated versions of the compounds. Our evidence indicates that key elements of the tryptamine pathway are operative in pea roots.  相似文献   

5.
6.
Shoot elongation is a vital process for plant development and productivity, in both ecological and economic contexts. Auxin and bioactive gibberellins (GAs), such as GA1, play critical roles in the control of elongation,13 along with environmental and endogenous factors, including other hormones such as the brassinosteroids.4,5 The effect of auxins, such as indole-3-acetic acid (IAA), is at least in part mediated by its effect on GA metabolism,6 since auxin upregulates biosynthesis genes such as GA 3-oxidase and GA 20-oxidase and downregulates GA catabolism genes such as GA 2-oxidases, leading to elevated levels of bioactive GA1.7 In our recent paper,1 we have provided evidence that this action of IAA is largely independent of DELLA proteins, the negative regulators of GA action,8,9 since the auxin effects are still present in the DELLA-deficient la cry-s genotype of pea. This was a crucial issue to resolve, since like auxin, the DELLAs also promote GA1 synthesis and inhibit its deactivation. DELLAs are deactivated by GA, and thereby mediate a feedback system by which bioactive GA regulates its own level.10 However, our recent results,1 in themselves, do not show the generality of the auxin-GA relationship across species and phylogenetic groups or across different tissue types and responses. Further, they do not touch on the ecological benefits of the auxin-GA interaction. These issues are discussed below as well as the need for the development of suitable experimental systems to allow this process to be examined.Key words: auxin, gibberellins, DELLA proteins, interactions, elongation  相似文献   

7.
The plant hormone auxin plays an essential role in plant development. However, only a few auxin biosynthetic genes have been isolated and characterized. Here, we show that the TRANSPORT INHIBITOR RESPONSE2 (TIR2) gene is required for many growth processes. Our studies indicate that the tir2 mutant is hypersensitive to 5-methyl-tryptophan, an inhibitor of tryptophan synthesis. Further, treatment with the proposed auxin biosynthetic intermediate indole-3-pyruvic acid (IPA) and indole-3-acetic acid rescues the tir2 short hypocotyl phenotype, suggesting that tir2 may be affected in the IPA auxin biosynthetic pathway. Molecular characterization revealed that TIR2 is identical to the TAA1 gene encoding a tryptophan aminotransferase. We show that TIR2 is regulated by temperature and is required for temperature-dependent hypocotyl elongation. Further, we find that expression of TIR2 is induced on the lower side of a gravitropically responding root. We propose that TIR2 contributes to a positive regulatory loop required for root gravitropism.Auxin is known to play an important role in plant development (Davies, 1995). However, many aspects of auxin biology remain poorly understood. Auxin is synthesized primarily in young tissues, such as cotyledons, leaves, and roots (Ljung et al., 2001, 2005), and transported to other tissues where it is perceived by members of the TRANSPORT INHIBITOR RESPONSE1 (TIR1) auxin receptor family. Recent studies have dramatically increased our knowledge of auxin transport and signaling (Quint and Gray, 2006; Vieten et al., 2007). However, the pathways of auxin synthesis and their regulation are still relatively unclear.Several indole-3-acetic acid (IAA) biosynthetic pathways have been proposed in plants based on research in plant-associated bacteria (Patten and Glick, 1996; Woodward and Bartel, 2005; Spaepen et al., 2007). There are two major types of pathways: the Trp-dependent and Trp-independent pathways. It has been hypothesized that plants have four Trp-dependent pathways that are generally named after an intermediate. In bacteria, the indole-3-pyruvic acid (IPA) pathway, one of the Trp-dependent pathways, has been described in detail (Koga, 1995; Spaepen et al., 2007). The current model for the IPA pathway involves a Trp aminotransferase oxidatively transaminating Trp to IPA. Subsequently, an IPA decarboxylase converts IPA to indole-3-acetaldehyde, and indole-3-acetaldehyde is oxidized to IAA. The IPA pathway is considered a major IAA biosynthetic pathway in plants, since potential intermediates have been isolated from different species (Sheldrake, 1973; Cooney and Nonhebel, 1991; Koga, 1995; Tam and Normanly, 1998). In addition, Trp transamination activity has been found in many plants (Gamborg, 1965; Forest and Wightman, 1972; Truelsen, 1973). Recently, two groups reported the identification of a gene called TAA1. This gene encodes an aminotransferase that converts Trp to IPA and functions in IAA biosynthesis (Stepanova et al., 2008; Tao et al., 2008).To identify genes that are required for auxin synthesis, transport, and signaling, we previously screened for Arabidopsis (Arabidopsis thaliana) mutants that are resistant to auxin transport inhibitors, such as N-1-napthylpthalamic (NPA; Ruegger et al., 1997). The treatment of seedlings with NPA results in auxin accumulation in the root tip (Ljung et al., 2005). Thus, mutants that are resistant to NPA may have defects in synthesis, transport, or response because roots of these mutants are expected to have lower levels of IAA or reduced sensitivity to IAA. This screen succeeded in isolating mutations in seven genes with weak NPA-resistant phenotypes, including genes related to auxin signaling (TIR1), auxin transport (TIR3), and auxin synthesis (TIR7; Ruegger et al., 1997, 1998; Ljung et al., 2005).Here, we describe the characterization of TIR2, a gene whose function is required for auxin synthesis. Genetic and physiological analyses of the tir2 mutant suggest that TIR2 is required for the Trp-dependent auxin synthesis pathway and functions as a Trp aminotransferase. Molecular cloning of TIR2 reveals that the gene is identical to TAA1 (Stepanova et al., 2008; Tao et al., 2008). We show that auxin regulates expression of TIR2 in a tissue-specific manner. Furthermore, we show that TIR2 is required for temperature-dependent hypocotyl elongation and that high temperature positively regulates expression of the TIR2 gene, suggesting that temperature regulates hypocotyl elongation directly by stimulating auxin synthesis. Finally, we provide evidence that TIR2 functions in a positive regulatory loop required for root gravitropism.  相似文献   

8.
9.
10.
For proper development of plants auxin levels need to be tightly controlled. For this, several routes have evolved and it is plausible that different organisms use these differently. To determine whether members of the family of GH3 proteins, which partially act as auxin conjugate synthetases in Arabidopsis thaliana, have similar roles in the moss Physcomitrella patens, we have investigated the in vitro activity of the two GH3 members in moss. We showed that both proteins can form amino acid conjugates with indole-3-acetic acid (IAA) but also with jasmonic acid (JA). Confirming these findings, single and double knockout-mutants showed lower levels of IAA conjugates than wild type. We discuss the results in light of the possible functions of IAA conjugate formation in lower land plants.Key words: Arabidopsis thaliana, auxin metabolism, jasmonic acid, GH3 genes, moss, Physcomitrella patensAuxins play diverse roles in many aspects of plant growth and development. Their activity is relying on the correct concentration in a given tissue and developmental stage.1 If higher levels of indole-3-acetic acid (IAA) are present, the hormone can also have an inhibitory effect on growth processes.2 Therefore, the tight control of IAA concentrations is absolutely necessary. To this end plants have evolved different mechanisms.3 First, biosynthesis is contributing to increasing IAA concentrations, mostly in young tissues such as meristems. Second, IAA can be transported in a polar way, depending on transport molecules, from cell to cell, away from the site of synthesis, thereby forming an auxin gradient along the plant axis. Third, IAA can be degraded, and fourth, IAA can be reversibly inactivated by conjugation to small molecules such as amino acids or sugars, but also be linked to larger molecules such as peptides or proteins.4 The inactive IAA conjugates can be hydrolyzed to yield free (i.e., active) IAA if needed. In higher plants the levels of free IAA constitutes between 5 and 20% depending on the tissue or age of the plant, whereas the conjugated form constitutes the major part.4 However, it is not yet clear in which way auxin homeostasis has evolved. The hypothesis that auxin has to be present during the evolution of a body plan has been tested by using different lower land plants which were compared in their mechanism to control auxin homeostasis. In algae, e.g., charophytes, the major metabolic way of controlling IAA is via biosynthesis. In bryophytes, the formation of IAA conjugates has been shown, although the amount was lower than for example in seed plants such as Arabidopsis thaliana.5,6 Since the molecular biology of auxin homeostasis in Arabidopsis is most advanced, we will use this model plant to compare the knowledge on seed plants with that in the moss Physcomitrella patens. The recent publication of the Physcomitrella genome7 gives the possibility to investigate components of the machinery controlling IAA levels in a lower land plant.In general, there seem to be high levels of redundancy involved in the pathways leading to decrease or increase of IAA, respectively. In Figure 1 we compare the current knowledge about genes related to IAA concentrations in Physcomitrella with Arabidopsis. While in Arabidopsis many different biosynthetic routes leading to IAA were identified,8 in the Physcomitrella genome homologs of the YUCCA genes have been detected.7 The presence of auxin conjugate synthetases has been experimentally verified in the moss (see below) and additional evidence for ester conjugate synthesis comes from sequence homology to UDP-glucosyl transferases.7 There is also the possibility of degradation of either IAA or an amino acid conjugate with IAA9,10 as discussed below.Open in a separate windowFigure 1Comparison of possibilities to regulate auxin homeostasis in Physcomitrella (solid lines) and Arabidopsis (dotted lines). Biosynthesis—AO, aldehyde oxidase; AMI1, amidase; CYP, cytochrome P450; NIT, nitrilase; TAA1, tryptophan aminotransferase; YUCCA, flavin monooxygenase; transport—AUX/LAX, auxin influx facilitator family; PIN, auxin efflux carrier family; PGP, ABC transporter type auxin efflux carrier family; conjugation/hydrolysis—UGT, UDP-glucosyl transferase; GH3, auxin conjugate synthetase family; ILR/IAR, auxin conjugate hydrolase family; Ox-IAA, oxindole-3-acetic acid; Ox-IAAsp, oxindole-3-aspartic acid.So far our work has focussed on the characterization of two members of the so called GH3 family, of which several from Arabidopsis can form conjugates of IAA with a variety of amino acids.11 While 19 members of this family have been described in Arabidopsis, only two are present in Physcomitrella.12 The Arabidopsis family clusters in three groups: group I containing the jasmonic acid conjugate synthetase JAR1 and a few others with as yet unkown function, group II the auxin conjugate synthetases, and group III with mostly as yet uncharacterized members.11,13 Sequence similarity of the GH3 genes from Physcomitrella showed that both cluster within the JAR1 group.12 Therefore, we analyzed the enzymatic activity of the two Physcomitrella GH3 proteins (PpGH3-1 and PpGH3-2) in vitro14 and found that both were active on jasmonic acid and a variety of different amino acids, whereas PpGH3-2 was active mostly with IAA. PpGH3-1 showed only weak activity with IAA and only two amino acids. For this reason, it could be assumed that the two Physcomitrella genes evolved by gene duplication, from which the initial activities would be for IAA and jasmonic acid. One of these genes might have evolved into a jasmonate conjugate synthetase (maybe AtJAR1),13 thereby loosing its activity on IAA. The second may have given rise to the auxin conjugate synthetase family in Arabidopsis,11 but the conjugate synthetases of Physcomitrella have still activity with both hormones. Interestingly, there is no evidence as yet that jasmonic acid itself has a role during Physcomitrella development, although a possible function of JA-conjugates has not been closely investigated. Since in Arabidopsis the JA conjugate with isoleucine is the active compound to be recognized by the COI1 receptor protein,15 it could be the case that JA itself has no effect in Physcomitrella. However, in our growth experiments a small growth promoting effect of JA, independently on the presence of GH3 genes was found. Similar observations were made with gibberellins in Physcomitrella.16Further characterization of single and double KO mutants in each of the PpGH3 genes has led to the hypothesis that GH3 proteins are indeed involved in regulating the auxin homeostasis in Physcomitrella.14 Both single KO mutants were more sensitive to increasing IAA concentrations in the medium than the wild type. Furthermore, the levels of free IAA were higher and the levels of conjugated IAA concomitantly dropped. A double KO mutant had almost no IAA conjugates when compared to the wild type. However, this mutant was still able to synthesize ester conjugates with IAA. Interestingly, the role of GH3 proteins in auxin conjugation seemed to be only important in the gametophore stage, whereas protonema cultures of GH3 KO mutants did not show any changes in auxin homeostasis. Therefore, we hypothesize that the role of GH3 proteins is dependent on a certain developmental stage of the moss. Additionally, we propose other detoxification mechanisms for example, export or degradation, in protonema.In higher plants the ester conjugate formation of IAA has been shown to be dependent on UDP-glucosyl transferases (AtUGT84B1 for Arabidopsis17 and ZmIAGLU for maize18). In the genome of Physcomitrella we could detect candidate sequence(s) for these genes, indicating that Physcomitrella has indeed the potential to synthesise the ester conjugates found in the gametophores in addition to amide conjugates. However, in the Physcomitrella genome, no homolog for an auxin conjugate hydrolase was found. In higher plants, auxin conjugate hydrolysis is thought to contribute to free IAA and depending on the plant species, large gene families with overlapping but distinct substrate preferences for individual amino acid conjugates with IAA are present.19,20 Since this is not the case for Physcomitrella, one has to ask the question whether the conjugation of auxin is a one-way road for inactivation of excess auxin and whether auxin conjugate hydrolysis has evolved later during plant evolution.In the Selaginella moellendorffii genome (http://genome.jgi-psf.org/Selmo1/Selmo1.home.html), an auxin conjugate hydrolase sequence related to higher plant ones, has been found based on homology searches, but the completion of the genome has to be awaited to draw final conclusions. Likewise, it is not clear, if this effect is specific for Physcomitrella, or found in bryophytes in general. Therefore, additional sequenced bryophyte genomes are needed.21Since in Arabidopsis the degradation of the IAA-Aspartate conjugate to Ox-IAA-Asp (see Fig. 1) has been described,9,10 a similar scenario could be suggested to occur in Physcomitrella with the amino acid conjugates formed. Alternatively, the hydrolysis of IAA conjugates by members of the M20 dipeptidase family can be envisioned. However, this would need the activity of enzymes with very low sequence conservation to auxin conjugate hydrolases. These questions will be addressed in future research by studying the metabolism of IAA and IAA conjugates of Physcomitrella in more detail.  相似文献   

11.
12.
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   

13.
It is remarkable that although auxin was the first growth-promoting plant hormone to be discovered, and although more researchers work on this hormone than on any other, we cannot be definitive about the pathways of auxin synthesis in plants. In 2001, there appeared to be a dramatic development in this field, with the announcement of a new gene,1 and a new intermediate, purportedly from the tryptamine pathway for converting tryptophan to the main endogenous auxin, indole-3-acetic acid (IAA). Recently, however, we presented evidence challenging the original and subsequent identifications of the intermediate concerned.2Key words: auxin synthesis, YUCCA, tryptamine, N-hydroxytryptamineThe new gene was termed YUC, and the putative intermediate is N-hydroxytryptamine. It was claimed that the YUC protein, a flavin-containing monooxygenase, converts tryptamine (formed from tryptophan by decarboxylation) to N-hydroxytryptamine, which is converted via other intermediates to IAA. When the YUC gene was expressed in E. coli and the resulting protein incubated with tryptamine, a weak TLC spot resulted, which produced a mass spectrum said to match that expected from N-hydroxytryptamine.1 However, the authors did not report mass spectral data from authentic N-hydroxytryptamine, and their suggested fragmentation pattern breaks a fundamental rule of mass spectrometry (the even-electron rule).2 Nevertheless, N-hydroxytryptamine has been added to virtually all IAA synthesis schemes published since 2001.3In 2010, LeClere et al. expressed a maize YUC gene in E. coli,4 and again claimed that the resulting protein converted tryptamine to N-hydroxytryptamine. This time, the mass spectrum was of better quality, but we have shown that it does not match that of authentic N-hydroxytryptamine, synthesised in our laboratory.2,5 We have demonstrated by electrospray tandem mass spectrometry that the protonated molecule of N-hydroxytryptamine (m/z 177) fragments to give an abundant ion at m/z 144. This was the crucial piece of evidence that the product obtained by LeClere et al. was not N-hydroxytryptamine, since their compound gave an abundant ion at m/z 160, and no ion at m/z 144.4The m/z 144 ion is formed by loss of NH2OH (hydroxylamine), as shown by accurate mass determinations (Fig. 1). In other words, it is the alkyl-amine bond that is broken; this is also the case for tryptamine and serotonin. In the latter case, an m/z 160 ion results through loss of ammonia, because the hydroxyl group on the indole ring (at position 5) is retained in the fragment. The compound obtained by LeClere et al. when protonated, also had a mass of 177, consistent with a hydroxylated tryptamine, and the abundant m/z 160 ion indicates that this fragment, as in serotonin, retains the hydroxyl group.4 However, we believe that the LeClere et al. product is not serotonin, because of dissimilar behavior on thin layer chromatography. Apart from the probability that it is a hydroxylated tryptamine, the identity of the LeClere et al. product is not known.Open in a separate windowFigure 1Fragmentation of (A) N-hydroxytryptamine, (B) tryptamine and (C) 5-hydroxytryptamine (serotonin), as determined by MS/MS analysis.2 The m/z ratios of the fragments produced are indicated. The loss of neutral hydroxylamine (A) or ammonia (B and C) involves heterolytic cleavage and/or hydrogen atom rearrangement, and consequent retention of the positive charge on the remaining indole-containing fragment.14It is interesting to contrast the previous “identifications” of N-hydroxytryptamine1,4 with the identification of gibberellins, during the period when most of the gibberellins were identified (1970–1990). There were rigorous criteria for the identification of these compounds, imposed by a triumvirate of “Gibberellin Godfathers”: Jake MacMillan, Nobatuka Takahashi and the late Bernard Phinney, and more latterly by Caporegimes such as Peter Hedden and Yuji Kamiya. Three of the present authors (James B. Reid, Noel W. Davies and John J. Ross) experienced at first hand the rigour with which these criteria were applied.Essentially, any identification of an endogenous gibberellin was viewed with suspicion unless a synthesized form of that compound (a standard, confirmed by NMR) was available for comparison. For a firm identification, the retention time on GC should be identical between the standard and the putative compound, on the same GC instrument. Next, the electron ionization fragmentation patterns of the compound of interest and the standard should match, again on the one GC-MS system. It was not considered adequate to compare a spectrum of the compound of interest with published spectra from another laboratory. Often a spectrum from a plant extract might contain extra ions, contributed by “interfering” compounds and this was sometimes acceptable. However, the absence of ions that should be present was usually sufficient to render the identification unconvincing. Electrospray mass spectra are intrinsically much poorer in information than electron ionisation spectra since most or all of the signal is concentrated in the protonated molecule, and tandem mass spectrometry (MS/MS) is required to create diagnostic fragment ions. The MS/MS spectrum of another hydroxylated tryptamine that we have examined is dominated by a strong m/z 160 ion, and discrimination between hydroxylated tryptamines on the basis of MS alone could be problematic. N-Hydroxytryptamine is the exception in this regard, and it can be easily distinguished.Another technique that has been used extensively in gibberellin research, and in early auxin research as well, is “feeding” labelled compounds and determining the fate of the label concerned (often deuterium or 13C). This technique contributed strongly to the identification of most of the candidate auxin pathways.6 Its power should not be underestimated, and yet in the auxin field, it was under-utilised during much of the later 1990s and the 2000s. We have used this technique to demonstrate that tryptamine is not converted to N-hydroxytryptamine in pea roots or seeds.2,5 In fact, to our knowledge, N-hydroxytryptamine has not yet been identified in any plant species.N-Hydroxytryptamine has been the main link between the YUC genes and the tryptamine pathway, and this link is now called into question. In fact, there are some indications that YUCCAs might not be concerned with IAA synthesis at all. The floozy mutant of petunia has a strong phenotype but normal levels of IAA.7 The yuc1 yuc2 yuc4 yuc6 quadruple mutant of Arabidopsis also exhibits a whole-plant phenotype but again its IAA content is not reduced compared with WT plants.8 Indeed, as yet there is not a single instance where knocking out YUC function has been shown to significantly reduce IAA content. We should note also that while overexpression of YUC genes does lead to elevated IAA content, the increase is small (up to about 2-fold1,9) compared with the increases recorded for other IAA over-producing mutants; for example, sur1 (also known as rty) and sur2, which can contain 5 to 20 times more IAA than the WT.10-12Therefore, it is possible that YUC catalyses a reaction or reactions in another pathway leading to another compound that is required for normal plant development; hence the phenotypic consequences of loss-of-function yuc mutants. This compound might be another auxin or auxinlike compound, which might explain why elevating auxin content genetically sometimes rescues yuc phenotypes.13 The suggestion that YUC controls the synthesis of another compound was made as early as 2002,7 but has attracted little attention from auxin biologists. There seems little doubt that YUCCAs play essential roles in plant development, as evidenced by the phenotypes of knockout mutants, even though it is sometimes necessary to construct multiple mutants to observe strong phenotypes.13 Furthermore, YUC genes are found in a wide range species, and we recently extended the list to include pea.2 However, the almost universal placement of N-hydroxytryptamine in auxin synthesis schemes since 2001 is now called into question by the recently published evidence.2  相似文献   

14.
15.
Thioredoxin (NTR/TRX) and glutathione (GSH/GRX) are the two major systems that play a key role in the maintenance of cellular redox homeostasis. They are essential for plant development, cell division or the response to environmental stresses. In a recent article,1 we studied the interplay between the NADP-linked thioredoxin and glutathione systems in auxin signaling genetically, by associating TRX reductase (ntra ntrb) and glutathione biosynthesis (cad2) mutations. We show that these two thiol reduction pathways interfere with developmental processes. This occurs through modulation of auxin activity as shown by genetic analyses of loss of function mutations in a triple ntra ntrb cad2 mutant. The triple mutant develops almost normally at the rosette stage but fails to generate lateral organs from the inflorescence meristem, producing almost naked stems that are reminiscent of mutants affected in PAT (polar auxin transport) or biosynthesis. The triple mutant exhibits other defects in processes regulated by auxin, including a loss of apical dominance, vasculature defects and reduced secondary root production. Furthermore, it has lower auxin (IAA) levels and decreased capacity for PAT, suggesting that the NTR and glutathione pathways influence inflorescence meristem development through regulation of auxin transport and metabolism.Key words: arabidopsis, NTS pathway, NGS pathway, thioredoxin (TRX), glutaredoxine (GRX), polar auxin transport (PAT), auxin biosynthesis, pin-like phenotype, apical dominance, meristematic activityExposure of living organisms to environmental stresses triggers various defense and developmental responses. Redox signaling is involved in many aspects of these responses.26 The key players in these responses are the NADPH-dependent glutathione/glutaredoxin system (NGS) and the NADPH-dependent thioredoxin system (NTS). TRX and GRX play key roles in the maintenance of cellular redox homeostasis.710 Genetic approaches aiming to identify functions of TRX and GRX in knock-out plants have largely been limited by the absence of phenotypes of single mutants, presumably due to functional redundancies among members of the multigene families of TRX and GRX.11 Interplay between NTS and NGS pathways have been studied in different organisms1217 and association of mutants involved in these two pathways have recently revealed new functions in several aspects of plant development.46  相似文献   

16.
17.
18.
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   

19.
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   

20.
The molecularly well-characterized auxin signal transduction pathway involves two evolutionarily conserved families interacting through their C-terminal domains III and IV: the Auxin Response Factors (ARFs) and their repressors the Aux/IAAs, to control auxin-responsive genes, among them genes involved in auxin transport.1,2 We have developed a new genetic tool to study ARF function. Using MONOPTEROS (MP)/ARF5, we have generated a truncated version of MP (MPΔ),3 which has lost the target domains for repression by Aux/IAA proteins. Besides exploring genetic interactions between MP and Aux/IAAs, we used this construct to trace MP’s role in vascular patterning, a previously characterized auxin dependent process.4,5 Here we summarize examples of naturally occurring truncated ARFs and summarize potential applications of truncated ARFs as analytical tools.  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