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 15 毫秒
1.
Cells of henbane (Hyoscyamus muticus L.) grow indefinitely in culture without exogenous auxin. Cells of its temperature-sensitive variant XIIB2 grow like the wild type at 26[deg]C but die rapidly at 33[deg]C unless auxin is added to the medium. Despite this temperature-sensitive auxin auxotrophy, XIIB2 produces wild-type amounts of indole-3-acetic acid (IAA). IAA is the predominant auxin and is important for plant growth and development. Since the IAA production of the variant is functional, we investigated whether the synthesis or degradation of IAA metabolites, possibly active auxins themselves, is altered. The IAA metabolites were IAA-aspartate (IAAsp) and IAA-glucose. The wild type converted IAA mainly to IAAsp, whereas the variant produced mainly IAA-glucose. Exogenous auxin corrected the shunted IAA metabolism of the variant. The half-life of labeled IAAsp in the variant was reduced 21-fold, but in the presence of exogenous auxin it was not different from the wild type. The temperature sensitivity of XIIB2 was also corrected by supplying IAAsp. Pulse-chase experiments revealed that henbane rapidly metabolizes IAAsp to compounds not identical to IAA. The data show that the variant XIIB2 is a useful tool to study the function of IAA conjugates to challenge the popular hypothesis that IAA conjugates are merely slow-release storage forms of IAA.  相似文献   

2.
We have isolated from plant surfaces several bacteria with the ability to catabolize indole-3-acetic acid (IAA). One of them, isolate 1290, was able to utilize IAA as a sole source of carbon, nitrogen, and energy. The strain was identified by its 16S rRNA sequence as Pseudomonas putida. Activity of the enzyme catechol 1,2-dioxygenase was induced during growth on IAA, suggesting that catechol is an intermediate of the IAA catabolic pathway. This was in agreement with the observation that the oxygen uptake by IAA-grown P. putida 1290 cells was elevated in response to the addition of catechol. The inability of a catR mutant of P. putida 1290 to grow at the expense of IAA also suggests a central role for catechol as an intermediate in IAA metabolism. Besides being able to destroy IAA, strain 1290 was also capable of producing IAA in media supplemented with tryptophan. In root elongation assays, P. putida strain 1290 completely abolished the inhibitory effect of exogenous IAA on the elongation of radish roots. In fact, coinoculation of roots with P. putida 1290 and 1 mM concentration of IAA had a positive effect on root development. In coinoculation experiments on radish roots, strain 1290 was only partially able to alleviate the inhibitory effect of bacteria that in culture overproduce IAA. Our findings imply a biological role for strain 1290 as a sink or recycler of IAA in its association with plants and plant-associated bacteria.  相似文献   

3.
4.
The Indole-3-Acetic Acid Oxidase of Lupinus albus L   总被引:10,自引:10,他引:0       下载免费PDF全文
  相似文献   

5.
The metabolism of indole-3-acetic acid (IAA) was investigated in 14-d-old Arabidopsis plants grown in liquid culture. After ruling out metabolites formed as an effect of nonsterile conditions, high-level feeding, and spontaneous interconversions, a simple metabolic pattern emerged. Oxindole-3-acetic acid (OxIAA), OxIAA conjugated to a hexose moiety via the carboxyl group, and the conjugates indole-3-acetyl aspartic acid (IAAsp) and indole-3-acetyl glutamate (IAGlu) were identified by mass spectrometry as primary products of IAA fed to the plants. Refeeding experiments demonstrated that none of these conjugates could be hydrolyzed back to IAA to any measurable extent at this developmental stage. IAAsp was further oxidized, especially when high levels of IAA were fed into the system, yielding OxIAAsp and OH-IAAsp. This contrasted with the metabolic fate of IAGlu, since that conjugate was not further metabolized. At IAA concentrations below 0.5 μm, most of the supplied IAA was metabolized via the OxIAA pathway, whereas only a minor portion was conjugated. However, increasing the IAA concentrations to 5 μm drastically altered the metabolic pattern, with marked induction of conjugation to IAAsp and IAGlu. This investigation used concentrations for feeding experiments that were near endogenous levels, showing that the metabolic pathways controlling the IAA pool size in Arabidopsis are limited and, therefore, make good targets for mutant screens provided that precautions are taken to avoid inducing artificial metabolism.The plant hormone IAA is an important signal molecule in the regulation of plant development. Its central role as a growth regulator makes it necessary for the plant to have mechanisms that strictly control its concentration. The hormone is believed to be active primarily as the free acid, and endogenous levels are controlled in vivo by processes such as synthesis, oxidation, and conjugation. IAA has been shown to form conjugates with sugars, amino acids, and small peptides. Conjugates are believed to be involved in IAA transport, in the storage of IAA for subsequent use, in the homeostatic control of the pool of the free hormone, and as a first step in the catabolic pathways (Cohen and Bandurski, 1978; Nowacki and Bandurski, 1980; Tuominen et al., 1994; Östin et al., 1995; Normanly, 1997). It is generally accepted that in some species conjugated IAA is the major source of free IAA during the initial stages of seed germination (Ueda and Bandurski, 1969; Sandberg et al., 1987; Bialek and Cohen, 1989), and there is also evidence that in some plants (but not all; see Bialek et al., 1992), the young seedling is entirely dependent on the release of free IAA from conjugated pools until the plant itself is capable of de novo synthesis (Epstein et al., 1980; Sandberg et al., 1987).The function of conjugated IAA during vegetative growth is somewhat less clear. It has been shown that conjugated IAA constitutes as much as 90% of the total IAA in the plant during vegetative growth (Normanly, 1997). However, the role of the IAA conjugates at this stage of the plant''s life cycle remains unknown. Analysis of endogenous IAA conjugates in vegetative tissues has revealed the presence of a variety of different compounds, including indole-3-acetyl-inositol, indole-3-acetyl-Ala, IAAsp, and IAGlu (Anderson and Sandberg, 1982; Cohen and Baldi, 1983; Chisnell, 1984; Cohen and Ernstsen, 1991; Östin et al., 1992). Studies of vegetative tissues have indicated that IAAsp, one of the major conjugates in many plants, is the first intermediate in an irreversible deactivation pathway (Tsurumi and Wada, 1986; Tuominen et al., 1994; Östin, 1995). Another mechanism that is believed to be involved in the homeostatic control of the IAA pool is catabolism by direct oxidation of IAA to OxIAA, which has been shown to occur in several plant species (Reinecke and Bandurski, 1983; Ernstsen et al., 1987).One area in the study of IAA metabolism in which our knowledge is increasing is the analysis of the homeostatic controls of IAA levels in plants. It has been possible, for instance, to increase the levels of IAA in transgenic plants expressing iaaM and iaaH genes from Agrobacterium tumefaciens. Analysis of these transgenic plants has indicated that plants have several pathways that can compensate for the increased production of IAA (Klee et al., 1987; Sitbon, 1992). It is expected that future studies using now-available genes will provide further insight into IAA metabolism. For example, a gene in maize encoding IAA-Glc synthetase has been identified, and several genes (including ILR1, which may be involved in hydrolysis of the indole-3-acetyl-Leu conjugate) have been cloned from Arabidopsis (Szerszen et al., 1994; Bartel and Fink, 1995). Furthermore, Chou et al. (1996) identified a gene that hydrolyzes the conjugate IAAsp to free IAA in the bacterium Enterobacter aggloremans.Because of its small genome size, rapid life cycle, and the ease of obtaining mutants, Arabidopsis is increasingly used as a genetic model system to investigate various aspects of plant growth and development. IAA signal transduction is also being investigated intensively in Arabidopsis in many laboratories (Leyser, 1997). Mutants with altered responses to externally added auxins or IAA conjugates have been identified in Arabidopsis. The identified mutants are either signal transduction mutants such as axr1-4 (Lincoln et al., 1990), or have mutations in genes involved in auxin uptake or transport, such as aux1 and pin1 (Okada et al., 1991; Bennett et al., 1996). A few mutants that are unable to regulate IAA levels or are unable to hydrolyze IAA conjugates, sur1-2 and ilr1, respectively, have also been identified (Bartel and Fink, 1995; Boerjan et al., 1995). To our knowledge, no mutant that is auxotrophic for IAA has been identified to date, which may reflect the redundancy in IAA biosynthetic pathways or the lethality of such mutants.In spite of the work reported thus far, many aspects of the metabolism of IAA in Arabidopsis require further investigation, because few details of the processes involved in IAA regulation are known. This lack of knowledge puts severe constraints on genetic analysis of IAA metabolism in Arabidopsis. For example, it is essential to have prior knowledge of IAA metabolism to devise novel and relevant screens with which to identify mutants of IAA metabolism. We have sought to address this issue by identifying the metabolic pathways involved in catabolism and conjugation under conditions that minimally perturb physiological processes. In this investigation we studied the conjugation and catabolic pattern of IAA by supplying relatively low levels of labeled IAA and identifying the catabolites and conjugates by MS. Different feeding systems were tested to optimize the application of IAA and to avoid irregularities in metabolism attributable to culturing, feeding conditions, or microbial activity. It is well documented that IAA metabolism is altered according to the amount of exogenous auxin applied; therefore, we placed special emphasis on distinguishing between catabolic routes that occur at near-physiological concentrations and those that occur at the high auxin concentrations commonly used in mutant screens.  相似文献   

6.
Both indole-3-acetamide (IAM) and indole-3-acetic acid (IAA)were identified in extracts of the hypocotyls of Japanese cherryby GC/MS. Exogenous IAA and IAM promoted the elongation of segmentsof these hypocotyls and the effect of IAA applied together withIAM was the same as that of IAA alone. (Received July 29, 1992; Accepted October 19, 1992)  相似文献   

7.
Indole-3-acetic acid is oxidized to oxindole-3-acetic acid by Zea mays tissue extracts. Shoot, root, and endosperm tissues have enzyme activities of 1 to 10 picomoles per hour per milligram protein. The enzyme is heat labile, is soluble, and requires oxygen for activity. Cofactors of mixed function oxygenase, peroxidase, and intermolecular dioxygenase are not stimulatory to enzymic activity. A heat-stable, detergent-extractable component from corn enhances enzyme activity 6- to 10-fold. This is the first demonstration of the in vitro enzymic oxidation of indole-3-acetic acid to oxindole-3-acetic acid in higher plants.  相似文献   

8.
9.
A general gas chromatography/mass spectrometry (MS)-based screen was performed to identify catabolites and conjugates of indole-3-acetic acid (IAA) during vegetative growth of Arabidopsis. This experiment revealed the existence of two new conjugates: N-(indole-3-acetyl)-alfa-alanine (IA-Ala) and N-(indole-3-acetyl)-alfa-leucine (IA-Leu). A method for quantitative analysis of IAA metabolites in plant extracts by liquid chromatography-electrospray tandem MS has been developed. The accuracy and precision of the new method are better than 10% for standards close to the detection limit, and are between 6% and 16% for the entire protocol applied to plant extracts. The low detection limits, 0.02 to 0.1 pmol for the different metabolites, made it possible to use as little as 50 to 100 mg of tissue for quantitative analysis. The analysis was performed on different tissues of an Arabidopsis plant at two stages of development, using heavy labeled internal standards of the catabolite 2-oxoindole-3-acetic acid as well as IAA conjugated to amino acids: aspartate, glutamate, Ala, and Leu. Expanding leaves and roots that generally contain high amounts of the free hormone also contained the highest levels of IA-aspartate, IA-glutamate, and 2-oxoindole-3-acetic acid, supporting their role as irreversible catabolic products. The levels of IA-Leu and IA-Ala did not follow the general distribution of IAA. Interestingly, the level of IA-Leu was highest in roots and IA-Ala in the aerial tissues.  相似文献   

10.
Immuno-Gold Localization of Indole-3-Acetic Acid in Peach Seedlings   总被引:3,自引:0,他引:3  
The localization of indole-3-acetic acid (IAA) in peach seedlings(Prunus persica [L.] Batsch ‘Momo Daigi Tsukuba 4’)was investigated using immunocytochemical technique. In meristematiccells of root tip, the gold particles were accumulated in nucleolus,while in leaf cells, they were mainly associated to chloroplastsand mitochondria. Physiological meaning of these localizationswas discussed. (Received December 13, 1989; Accepted April 12, 1990)  相似文献   

11.
Indole-3-acetic acid (IAA) was detected in the ether extracts of culture filtrates of indigotin-producing strains of the basidiomycete Schizophyllum commune. Several solvents, known to give distinctly different RF values for IAA, and 3 location reagents gave identical results with synthetic IAA and IAA found in the extract. Confirmation was obtained by the Avena straight growth test, split pea test, and ultraviolet absorption spectrum.  相似文献   

12.
A soil culture experiment was conducted to determine whether a plant hormone, indole-3-acetic acid (IAA), could influence fluoranthene (Flu) removal from soil. Four treatments were utilized: (i) unplanted soil (CK), (ii) soil planted with ryegrass (P), (iii) soil planted with ryegrass and treated with 0.24 mg kg?1 IAA (P+0.24), (iv) soil planted with ryegrass and treated with 2.4 mg kg?1 IAA (P+2.4). The Flu initial concentration was 200 mg kg?1. After 3 months, the percentage of Flu removal and plant root biomass were significantly increased under the P+2.4 and the removal rate was 35.68%. The total Flu content in plants was higher than that in the other treatments. The Flu concentration was significantly increased in the shoots, but not significantly altered in the roots. The highest translocation factor was observed in the P+2.4. Increase in number of bacteria, actinomycetes and fungi were observed in the planted treatments, and the amount of fungi was significantly increased in P+2.4. Flu removal was related to the Flu in ryegrass, and was insignificantly correlated with the stimulation of soil microflora, which suggesting that IAA may work mainly on improving plant growth, the Flu uptake, and eventually leading to enhanced remediation of Flu polluted soil.  相似文献   

13.
Genetic evidence in Arabidopsis (Arabidopsis thaliana) suggests that the auxin precursor indole-3-butyric acid (IBA) is converted into active indole-3-acetic acid (IAA) by peroxisomal β-oxidation; however, direct evidence that Arabidopsis converts IBA to IAA is lacking, and the role of IBA-derived IAA is not well understood. In this work, we directly demonstrated that Arabidopsis seedlings convert IBA to IAA. Moreover, we found that several IBA-resistant, IAA-sensitive mutants were deficient in IBA-to-IAA conversion, including the indole-3-butyric acid response1 (ibr1) ibr3 ibr10 triple mutant, which is defective in three enzymes likely to be directly involved in peroxisomal IBA β-oxidation. In addition to IBA-to-IAA conversion defects, the ibr1 ibr3 ibr10 triple mutant displayed shorter root hairs and smaller cotyledons than wild type; these cell expansion defects are suggestive of low IAA levels in certain tissues. Consistent with this possibility, we could rescue the ibr1 ibr3 ibr10 short-root-hair phenotype with exogenous auxin. A triple mutant defective in hydrolysis of IAA-amino acid conjugates, a second class of IAA precursor, displayed reduced hypocotyl elongation but normal cotyledon size and only slightly reduced root hair lengths. Our data suggest that IBA β-oxidation and IAA-amino acid conjugate hydrolysis provide auxin for partially distinct developmental processes and that IBA-derived IAA plays a major role in driving root hair and cotyledon cell expansion during seedling development.The auxin indole-3-acetic acid (IAA) controls both cell division and cell expansion and thereby orchestrates many developmental events and environmental responses. For example, auxin regulates lateral root initiation, root and stem elongation, and leaf expansion (for review, see Davies, 2004). Normal plant morphogenesis and environmental responses require modulation of auxin levels by controlling biosynthesis, regulating transport, and managing storage forms (for review, see Woodward and Bartel, 2005a). In some storage forms, the carboxyl group of IAA is conjugated to amino acids or peptides or to sugars, and free IAA can be released by hydrolases when needed (Bartel et al., 2001; Woodward and Bartel, 2005a). A second potential auxin storage form is the side chain-lengthened compound indole-3-butyric acid (IBA), which can be synthesized from IAA (Epstein and Ludwig-Müller, 1993) and is suggested to be shortened into IAA by peroxisomal β-oxidation (Bartel et al., 2001; Woodward and Bartel, 2005a).Genetic evidence suggests that the auxin activity of both IAA-amino acid conjugates and IBA requires free IAA to be released from these precursors (Bartel and Fink, 1995; Zolman et al., 2000). Mutation of Arabidopsis (Arabidopsis thaliana) genes encoding IAA-amino acid hydrolases, including ILR1, IAR3, and ILL2, reduces plant sensitivity to the applied IAA-amino acid conjugates that are substrates of these enzymes, including IAA-Leu, IAA-Phe, and IAA-Ala (Bartel and Fink, 1995; Davies et al., 1999; LeClere et al., 2002; Rampey et al., 2004), which are present in Arabidopsis (Tam et al., 2000; Kowalczyk and Sandberg, 2001; Kai et al., 2007).Unlike the simple one-step release of free IAA from amino acid conjugates, release of IAA from IBA is suggested to require a multistep process (Zolman et al., 2007, 2008). Conversion of IBA to IAA has been demonstrated in a variety of plants (Fawcett et al., 1960; for review, see Epstein and Ludwig-Müller, 1993) and may involve β-oxidation of the four-carbon carboxyl side chain of IBA to the two-carbon side chain of IAA (Fawcett et al., 1960; Zolman et al., 2000, 2007). Mutation of genes encoding the apparent β-oxidation enzymes INDOLE-3-BUTYRIC ACID RESPONSE1 (IBR1), IBR3, or IBR10 results in IBA resistance, but does not alter IAA response or confer a dependence on exogenous carbon sources for growth following germination (Zolman et al., 2000, 2007, 2008), consistent with the possibility that these enzymes function in IBA β-oxidation but not fatty acid β-oxidation.Both conjugate hydrolysis and IBA β-oxidation appear to be compartmentalized. The IAA-amino acid hydrolases are predicted to be endoplasmic reticulum localized (Bartel and Fink, 1995; Davies et al., 1999) and enzymes required for IBA responses, including IBR1, IBR3, and IBR10, are peroxisomal (Zolman et al., 2007, 2008). Moreover, many peroxisome biogenesis mutants, such as peroxin5 (pex5) and pex7, are resistant to exogenous IBA, but remain IAA sensitive (Zolman et al., 2000; Woodward and Bartel, 2005b).Although the contributions of auxin transport to environmental and developmental auxin responses are well documented (for review, see Petrášek and Friml, 2009), the roles of various IAA precursors in these processes are less well understood. Expansion of root epidermal cells to control root architecture is an auxin-regulated process in which these roles can be dissected. Root epidermal cells provide soil contact and differentiate into files of either nonhair cells (atrichoblasts) or hair cells (trichoblasts). Root hairs emerge from trichoblasts as tube-shaped outgrowths that increase the root surface area, thus aiding in water and nutrient uptake (for review, see Grierson and Schiefelbein, 2002). Root hair length is determined by the duration of root hair tip growth, which is highly sensitive to auxin levels (for review, see Grierson and Schiefelbein, 2002). Mutants defective in the ABCG36/PDR8/PEN3 ABC transporter display lengthened root hairs and hyperaccumulate [3H]IBA, but not [3H]IAA, in root tip auxin transport assays (Strader and Bartel, 2009), suggesting that ABCG36 functions as an IBA effluxer and that IBA promotes root hair elongation. The related ABCG37/PDR9 transporter also can efflux IBA (Strader et al., 2008b; Růžička et al., 2010) and may have some functional overlap with ABCG36 (Růžička et al., 2010). In addition to lengthened root hairs, abcg36/pdr8/pen3 mutants display enlarged cotyledons, a second high-auxin phenotype. Both of these developmental phenotypes are suppressed by the mildly peroxisome-defective mutant pex5-1 (Strader and Bartel, 2009), suggesting that IBA contributes to cell expansion by serving as a precursor to IAA, which directly drives the increased cell expansion that underlies these phenotypes. However, whether IBA-derived IAA contributes to cell expansion events during development of wild-type plants is not known.Here, we directly demonstrate that peroxisome-defective mutants are defective in the conversion of IBA to IAA, consistent with previous reports that these genes are necessary for full response to applied IBA. We found that a mutant defective in three suggested IBA-to-IAA conversion enzymes displays low-auxin phenotypes, including decreased root hair expansion and decreased cotyledon size. We further found that these mutants suppress the long-root-hair and enlarged cotyledon phenotypes of an abcg36/pdr8 mutant, suggesting that endogenous IBA-derived IAA drives root hair and cotyledon expansion in wild-type seedlings.  相似文献   

14.
Nonhcbcl, H. M. 1986. Measurement of the rates of oxindole-3-aceticacid turnover and indole-3-acetic acid oxidation in Zea maysseedlings.—J. exp. Bat. 37: 1691–1697. Oxindole-3-acetic acid is the pnncipal catabolite of indole-3-aceticacid in Zea mays seedlings. In this paper measurements of theturnover of oxindole-3-acetic acid are presented and used tocalculate the rate of indole-3-acetic acid oxidation. [3H]Oxindolc-3-acetic acid was applied to the endosperm of Zeamays seedlings and allowed to equilibrate for 24 h before thestart of the experiment. The subsequent decrease in its specificactivity was used to calculate the turnover rate. The averagehalf-life of oxindole-3-acetic acid in the shoots was foundto be 30 h while that in the kernels had an average half-lifeof 35 h. Using previously published values of the pool sizesof oxindole-3-acetic acid in shoots and kernels from seedlingsof the same age and variety, and grown under the same conditions,the rate of indole-3-acetic acid oxidation was calculated tobe I-I pmol plant–1 h–1 in the shoots and 7·1pmol plant–1 h–1 in the kernels. Key words: Oxindole-3-acetic acid, indole-3-acetic acid, turnover, Zea mays  相似文献   

15.
The major ether-soluble, growth-stimulating substance detected by the Avena coleoptile straight-growth test in extract from sprouting buds of Scots pine (Pinus silvestris L.) was identified as indole-3-acetic acid by Rf values in 5 solvent systems and by its elution volume in ethanol on a Sephadex LH-20 column. When the substance was applied to the growth solution of wheat roots in a special test the growth in length of the roots was at first inhibited, but growth was recovered after about 6 hours in the same manner as when small quantities of IAA were applied. The extracts also contained large amounts of growth inhibitors which interfered with the auxin response if they were not removed.  相似文献   

16.
《植物生理学报》2013,(6):1992-1995
Dear Editor, The plant hormone indole-3-acetic acid (IAA) has long been used in plant culture media for practical applications and sci- entific inquiries. The use of IAA is complicated by the fact that IAA is a photo-labile compound. In Murashige and Skoog (MS) plant media (Murashige and Skoog, 1962), the concen- trations of salts and mineral nutrients are known to hasten the photodegradation of IAA under white light (Dunlap and Robacker, 1988). This degradation can be virtually eliminated by the use of a yellow-colored light filter that removes UV, violet, and some of the blue wavelengths from the incident light (Stasinopoulos and Hangarter, 1990). However, the use of yellow light clearly affects the quality of light that the plants under study receive. In addition to applications in plants, IAA has been used in human health applications.  相似文献   

17.
18.
Evidence for Trp-independent IAA synthesis is critically reevaluated in the light of tryptophan synthase proteome data, local IAA synthesis and Trp, indole-3-pyruvate, and IAA turnover.Trp-independent synthesis of indole-3-acetic acid (IAA) was proposed back in the early 1990s based on observations from Trp auxotrophs in maize (Zea mays; Wright et al., 1991) and Arabidopsis (Arabidopsis thaliana; Normanly et al., 1993). Recently, Wang et al. (2015) published new data suggesting that a cytosolic indole synthase (INS) may catalyze the first step separating the Trp-dependent and Trp-independent pathways in Arabidopsis. If this is the case, it would be a major breakthrough; however, in this article, I critically evaluate both recent and older evidence for the Trp-independent route and suggest that the INS is more likely to participate in Trp-dependent IAA production.The original work supporting Trp-independent IAA production was carried out prior to the availability of genome/proteome data and before the discovery that the final step of Trp-dependent IAA synthesis is carried out by a large number of YUCCA homologs operating in a highly localized manner (Zhao, 2008). I argue that experimental data supporting the Trp-independent route needs to be reconsidered in light of complete proteome data. Further, the evidence from feeding labeled compounds should be critically evaluated in light of recent data on the highly localized nature of IAA synthesis as well as older quantitative data on Trp, indole-3-pyruvic acid (IPA), and IAA turnover from my own laboratory (Cooney and Nonhebel, 1991). I conclude that evidence for the Trp-independent route is at best equivocal, and that it is not a conserved source of IAA in angiosperms.Figure 1 shows the major Trp-dependent route for IAA production whereby Trp, produced by the concerted action of Trp synthase α- and β-subunits, is converted to IAA in a further two steps catalyzed by Trp aminotransferase and the flavin monooxygenases commonly known as YUCCA (Mashiguchi et al., 2011; Won et al., 2011). This is compared with the Trp-independent route in which IAA may be produced from free indole by an unknown route (Ouyang et al., 2000; Wang et al., 2015).Open in a separate windowFigure 1.Outline of the major pathway for Trp-dependent IAA synthesis and the proposed Trp-independent route. The role proposed for Trp synthase beta (TSB) homologs discussed in the present paper is also shown. For clarity, reactions are simplified to show only the major compounds relevant to IAA synthesis.The Trp-independent route was originally based on data from Trp auxotrophs that have mutations in genes encoding either the α- or β-subunits of Trp synthase. The α-subunit catalyzes the removal of the side chain from indole-3-glycerol phosphate, passing the indole product directly to the β-subunit where the Trp side chain is created from a Ser substrate (Pan et al., 1997). In plants, this is a chloroplast-localized enzyme. Elevated levels of IAA have been reported in Trp auxotrophs of both maize and Arabidopsis. However, the trp3-1 and trp2-1 mutants of Arabidopsis, deficient in the α- and β-subunits, respectively, only showed an increase in total IAA measured following conjugate hydrolysis. No difference in free IAA levels was found (Normanly et al., 1993). The orange pericarp (orp) maize mutant was reported to have 50 times more IAA than the wild type (Wright et al., 1991). However, this was also total IAA; no data on free IAA were published. Work by Müller and Weiler (2000) indicated that IAA measured following conjugate hydrolysis could have originated via the degradation of indole-3-glycerol phosphate that accumulates in trp3-1 mutants. Further doubt regarding the accuracy of IAA measurements following conjugate hydrolysis has recently been published. Yu et al. (2015) have shown that conjugate hydrolysis treatment substantially overestimates the actual conjugated IAA due to degradation of glucobrassicin and proteins. In addition, neither report (Wright et al., 1991; Normanly et al., 1993) described a high auxin phenotype for the Trp auxotrophs. This contrasts with the superroot1 (sur1) and sur2 mutants, where the accumulation of indole intermediates resulted in a high level of free IAA as well as a high auxin phenotype (Boerjan et al., 1995; Delarue et al., 1998). It is therefore doubtful that Trp auxotrophs actually accumulate more IAA than the wild-type plants.In addition, proteome data have revealed new homologs of TSB in both Arabidopsis and maize that may contribute to Trp production in TSB mutants; these have not been considered in arguments supporting Trp-independent IAA synthesis. Maize orp has mutations in two TSB genes, resulting in a seedling lethal phenotype with high levels of accumulated indole. However, proteome sequence information now indicates that maize has three TSB genes. Plants and bacteria have divergent forms of TSB, type 1 and type 2 (Xie et al., 2001); the major TSB genes responsible for Trp synthase activity in maize and Arabidopsis are type 1. The third maize TSB gene, maize locus ID GRMZM2G054465, is a member of the TSB type 2 group. Its product is reported not to interact directly with a Trp synthase alpha (TSA) subunit but has experimentally demonstrated catalytic activity converting indole and Ser to Trp (Yin et al., 2010). This type 2 TSB may allow orp plants to make sufficient Trp for IAA production from the accumulated free indole.When the original work on trp2 mutants of Arabidopsis was carried out, two TSB genes were known (Last et al., 1991). As the trp2 plants were deficient only in TSB1, they were able to make sufficient Trp to survive under low-light conditions. Full proteome data now indicate that Arabidopsis has four TSB-like genes; in addition to TSB1 and TSB2, there is a third type 1 TSB gene, Arabidopsis locus ID AT5G28237. The product of this gene has not been experimentally characterized. The fourth gene, AT5G38530, encodes a type 2 TSB with demonstrated catalytic activity similar to ZmTSB type 2 mentioned above (Yin et al., 2010). Thus, the trp2 plants may also make enough Trp for IAA production. It is even possible that one of the minor forms of TSB has a specific role in IAA production. Type 2 TSBs are conserved throughout the plant kingdom, and the biological role for this protein is not known (Xie et al., 2001).A phylogenetic analysis of type 1 TSBs is shown in Figure 2. This indicates that the product of AT5G28237 belongs to a eudicot-conserved TSB type 1-like clade, divergent from that containing major experimentally characterized TSBs. A multiple sequence alignment (not shown) reveals that members of this divergent clade have a shortened N terminus with respect to the major chloroplast-localized TSB proteins. A localization prediction carried out in CELLO (Yu et al., 2006) suggests a cytosolic location for these proteins. Examination of EST databases indicates that the genes encoding these proteins are expressed. It is possible that the product of AT5G28237 could interact with the cytosolic INS studied by Wang et al. (2015), or separately with its indole product, to produce Trp that is further converted to IAA.Open in a separate windowFigure 2.Phylogeny of TSB type 1 homologs from Oryza sativa (LOC_Os), Sorghum bicolor (Sobic), Z. mays (GRMZM), Arabidopsis (AT), Brassica rapa (Brara), Solanum lycopersicum (Solyc), Populus trichocarpa (Potri), and Physcomitrella patens (Phpat). Protein sequences were downloaded from Phytozome v10.2 (Goodstein et al., 2012). The phylogenetic analysis was conducted in MEGA6 (http://megasoftware.net; Tamura et al., 2013) with multiple sequence alignment by MUSCLE (Edgar, 2004) and evolutionary history inferred using the neighbor-joining method (Saitou and Nei, 1987). The optimal tree is shown; the percentage of replicate trees in which the associated sequences clustered together in the bootstrap test (500 replicates) is shown next to the branches (Felsenstein, 1985). The tree is drawn to scale; the scale bar indicates the number of amino acid substitutions per site. It is rooted with type 1 TSBs from the moss P. patens.The second major line of evidence for Trp-independent IAA synthesis comes from isotopic labeling experiments. Wright et al. (1991) observed greater incorporation of 2H into IAA than Trp in orp seedlings grown on 2H2O. Normanly et al. (1993) reported higher enrichment of 15N in IAA than Trp in trp2-1 mutants of Arabidopsis grown on 15N anthranilate; very poor incorporation of deuterium from 2H-Trp into IAA was reported in the trp2-1 plants. A number of similar reports relating to other plants have been published showing differences in the incorporation of label from Trp into IAA depending on experimental tissue and environmental conditions (e.g. Michalczuk et al., 1992; Rapparini et al., 2002; Sztein et al., 2002). This evidence has been persuasive; however, it assumes a single pool of Trp to which 15N anthranilate and 2H-Trp contribute and from which IAA is made. If Trp is made at different rates in different parts of the plant, and/or exogenous 2H-Trp does not equilibrate with newly synthesized Trp, then the ratio of 15N to 2H in Trp will vary in different plant organs/tissues/cells. Trp turnover and thus incorporation of label from 15N anthranilate are likely to differ substantially throughout the plant, with the highest rates of labeling occurring in cells with high rates of protein synthesis. This would not be a problem for the experiment if IAA is made at equal rates in different parts of the plant, but we know it is not. The Trp aminotransferase/YUCCA pathway of IAA synthesis elegantly shown to be responsible for the bulk of IAA synthesis (Mashiguchi et al., 2011; Won et al., 2011) appears to be locally controlled in Arabidopsis via 11 different YUCCA-encoding genes that have highly localized expression (Zhao, 2008). Adding to the complication is the need for 15N anthranilate and 2H-Trp to move into and through the plant to regions of Trp and IAA synthesis, respectively. This is likely to occur at different rates due to differing transporter requirements.Data from my own laboratory (Cooney and Nonhebel, 1991) is particularly relevant to this discussion. We monitored incorporation of 2H from deuterated water into IAA and Trp in tomato (S. lycopersicum) shoots. Unlike the other studies, we also measured the incorporation of label into IPA. Our data showed that IPA became labeled at a rate consistent with this compound acting as the major/sole precursor of IAA. Crucially, the proportion of labeled Trp was lower than 2H-IPA. Our interpretation of these data was that IPA and IAA were produced from newly synthesized Trp, and that Trp was not uniformly labeled throughout the shoot. At the time, we suggested different subcellular pools of Trp; this may be the case, but in light of new knowledge of localized IAA synthesis, it is most likely that substantial differences in Trp and IAA turnover in different cells/tissues may be the reason for these observations.The arguments above cast doubt on the existence of the Trp-independent route; however, a recent publication by Wang et al. (2015) claims to provide new evidence for its importance. They present the interesting finding that Arabidopsis plants with a null mutation in INS, a cytosolic TSA homolog previously shown to have indole-3-glycerol phosphate lyase (IGL) activity (Zhang et al., 2008), had reduced levels of IAA. The mutation particularly affected early embryo development. I suggest that the INS may make a contribution to IAA synthesis, but the only specific evidence that it does so via a Trp-independent route is the observation that the ins-1 mutation has an additive effect with the weakly ethylene insensitive8-1 Trp aminotransferase mutation. This evidence is indicative rather than conclusive. The possibility that INS may act in concert with a minor TSB homolog, as suggested in Figure 1, needs to be considered.In addition, Wang et al. (2015) focus on Arabidopsis alone. If INS has a key role in IAA synthesis, then evolutionary theory predicts a conserved protein with wide taxonomic distribution. On the contrary, an exhaustive BLAST search (Altschul et al., 1997) of diverse taxa in Phytozome v10.2 (Goodstein et al., 2012) and GenBank (Benson et al., 2013) revealed that INS orthologs with cytosolic prediction and shortened N terminus occur only in members of the Brassicaceae (Eutrema salsugineum, Arabis alpina, Camelina sativa, Capsella rubella, Brassica napus, Boechera stricta, Arabidopsis lyrata, B. rapa) and in Tarenaya hassleriana from the Brassicaceae sister family, the Cleomaceae. The phylogenetic tree in Figure 3 shows relationships between INS and TSA homologs from several plant species and indicates the separate clade of cytosolic INS homologs in the Brassicaceae. In this diagram, the sequence most closely related to INS from another group is that from tomato. This protein is the only TSA found in tomato and has an unambiguous chloroplast signal peptide. P. trichocarpa and M. truncatula as well as other eudicots outside the Brassicaeae and Cleomaceae also lack cytosolic TSA homologs. Furthermore, INS and its orthologs are phylogenetically distinct from the other experimentally characterized indole-3-glycerol phosphate lyases benzoxazin1 and IGL (Frey et al., 2000) and their orthologs. The latter are restricted to the grasses where they are involved in the production of cyclic hydroxamic acid defense compounds (Frey et al., 2000). The grasses also have an additional separate clade of cytosolic TSA homologs, although work by Kriechbaumer et al. (2008) did not detect any catalytic activity for the product of GRMZM2G046191_T01. The phylogeny of INS and its orthologs would suggest the major role of these proteins may be the production of lineage-specific metabolites such as the indole-derived defense compounds produced in grasses; any role in IAA synthesis may be incidental and restricted to the Brassicaeae and Cleomaceae.Open in a separate windowFigure 3.Phylogeny of TSA homologs from O. sativa (LOC_Os), S. bicolor (Sobic), Z. mays (GRMZM), Arabidopsis (AT), A. lyrata (Alyrata), B. rapa (Brara), S. lycopersicum (Solyc), Medicago truncatula (Medtr), P. trichocarpa (Potri), and P. patens (Phpat). Protein sequences were downloaded from Phytozome v10.2 (Goodstein et al., 2012). The phylogenetic analysis was conducted in MEGA6 (Tamura et al., 2013) with multiple sequence alignment by MUSCLE (Edgar, 2004) and evolutionary history inferred using the neighbor-joining method (Saitou and Nei, 1987). The rooted optimal tree is shown; the percentage of replicate trees in which the associated sequences clustered together in the bootstrap test (500 replicates) is shown next to the branches (Felsenstein, 1985). The tree is drawn to scale; the scale bar indicates the number of amino acid substitutions per site. It is rooted with the TSA ortholog from the moss P. patens.In conclusion, I contend that experimental data relating to IAA synthesis in Arabidopsis, including that suggesting the involvement of a cytosolic INS, can be explained by the Trp-dependent IAA synthesis pathway. I show that INS and its orthologs are not found outside the Brassicaceae and a closely related sister clade; any alternative IAA synthesis pathway in which they may be involved is likely to have similar limited taxonomic occurrence. Furthermore, Arabidopsis and its relatives contain two additional TSB homologs that could convert free indole into Trp. Curiously, both of these proteins have a wider taxonomic distribution. A priority for further experimental work should be testing the involvement of minor TSB homologs in IAA synthesis, including the highly conserved type 2 TSBs as well as a eudicot-specific clade of possibly cytosolic type 1 TSBs. Work would also have to establish whether free indole exists in plants other than the Brassicaceae and the grasses. Finally, I argue that isotope-labeling experiments do not provide strong support for the Trp-independent route, as IAA production is highly localized. Previously published data from my laboratory clearly show that the main Trp-dependent IAA precursor IPA becomes more highly labeled from 2H2O than Trp, even though the latter is produced from Trp in a single reaction. Thus, it cannot be argued that differences in isotope enrichment between Trp and IAA demonstrate the existence of a Trp-independent route.  相似文献   

19.
When indole-3-acetic acid (IAA) is applied to the cotyledonsof broad bean seedlings (Vicia faba L. cv Chukyo), the majormetabolites found in the roots are 3-(O-ß-glucosyl)-2-indoIone-3-acetylaspartic acid (Glc-DIA-Asp) and 3-hydroxy-2-indolone-3-acetylasparticacid (DIA-Asp). In this report, the metabolic pathway from IAAto the two dioxindole-3-acetic acid (DIA) conjugates was investigatedby using [14C]IAA, [14C]DIA, [14C]indole-3-acetylaspartic acid(IAA-Asp), and [14C]IAA-[3H]Asp. The precursor of DIA-Asp wasfound to be IAA-Asp but not DIA. Incorporation of the doublelabeled IAA-Asp into the DIA conjugates demonstrated that hydrolysisof IAA-Asp was not involved in the formation of the DIA conjugates.DIA-Asp was further metabolized to Glc-DIA-Asp in the cotyledons,while formation of Glc-DIA-Asp in the roots was very low. Glc-DIA-Aspformed in the cotyledons was transported to the roots. (Received April 21, 1986; Accepted September 10, 1986)  相似文献   

20.
We used coleoptile sections of Avena sativa, Sorghum bicolor,and Zea mays seedlings to examine interactions between epidermalgrowth factor (EGF) and indole-3-acetic acid (IAA) that mayaffect plant growth and development. Our 24-h bioassays employedthree controls ranging in dilution from 10–4 to 10–8g ml–1: (1) 50 mM potassium-phosphate buffer solution(pH=6.0), (2) bovine serum albumin, a nonspecific protein; and(3) IAA; plus two treatments: (1) mouse epidermal growth factor(EGF) ranging from 10–6 to 10–10gml–1, and(2) EGF + IAA. In all three species growth in IAA, EGF, andEGF + IAA treatments showed significant increases over controls;EGF+IAA showed significant increases in growth over IAA alone.As the concentrations of IAA decreased, the EGF and IAA interactionbecame more pronounced. At the highest IAA concentrations, EGF+ IAA increased growth rates ca. 2% to 39%, whereas at lowerIAA concentrations EGF + IAA promoted growth as much as 121%,thereby lowering the normal IAA physiological set point up tothree or four orders of magnitude. Our data suggest that aninteraction between EGF and IAA may allow plants to recognizeand respond to animal biochemical messengers, resulting in changesin plant cell elongation that ultimately may alter plant growthpatterns. (Received April 27, 1994; Accepted September 5, 1994)  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