A Novel Auxin Conjugate Hydrolase from Wheat with Substrate ...

0 downloads 0 Views 191KB Size Report
... Molecular Biology, Montclair State University, Montclair, New Jersey 07043 ..... Seeds for sources of root, upper and lower leaf, and stem tissue were sown.
A Novel Auxin Conjugate Hydrolase from Wheat with Substrate Specificity for Longer Side-Chain Auxin Amide Conjugates1 James J. Campanella*, Adebanke F. Olajide, Volker Magnus, and Jutta Ludwig-Mu¨ller Department of Biology and Molecular Biology, Montclair State University, Montclair, New Jersey 07043 (J.J.C., A.F.O.); Rudjer Boskovic Institute, 10002 Zagreb, Croatia (V.M.); and Institut fu¨r Botanik, Technische Universita¨t Dresden, 01062 Dresden, Germany (J.L.-M.)

This study investigates how the ILR1-like indole acetic acid (IAA) amidohydrolase family of genes has functionally evolved in the monocotyledonous species wheat (Triticum aestivum). An ortholog for the Arabidopsis IAR3 auxin amidohydrolase gene has been isolated from wheat (TaIAR3). The TaIAR3 protein hydrolyzes negligible levels of IAA-Ala and no other IAA amino acid conjugates tested, unlike its ortholog IAR3. Instead, TaIAR3 has low specificity for the ester conjugates IAA-Glc and IAAmyoinositol and high specificity for the conjugates of indole-3-butyric acid (IBA-Ala and IBA-Gly) and indole-3-propionic-acid (IPA-Ala) so far tested. TaIAR3 did not convert the methyl esters of the IBA conjugates with Ala and Gly. IBA and IBA conjugates were detected in wheat seedlings by gas chromatography-mass spectrometry, where the conjugate of IBA with Ala may serve as a natural substrate for this enzyme. Endogenous IPA and IPA conjugates were not detected in the seedlings. Additionally, crude protein extracts of wheat seedlings possess auxin amidohydrolase activity. Temporal expression studies of TaIAR3 indicate that the transcript is initially expressed at day 1 after germination. Expression decreases through days 2, 5, 10, 15, and 20. Spatial expression studies found similar levels of expression throughout all wheat tissues examined.

In vascular plants, auxins, primarily indole-3-acetic acid (IAA), regulate gene expression, cell division, and cell elongation and differentiation in plant tissue. Auxins also affect vascularization, phototropism, geotropism, fruit development, flower development, and apical dominance (Davies, 1995). While IAA in low concentrations stimulates growth and development, higher concentrations can be toxic to the plant (Bandurski et al., 1995). Therefore, tight control of IAA concentration is necessary for proper plant development. IAA is stored in conjugated forms that are mostly considered to be inactive. Two main types of conjugated molecules have been studied: the amide-linked IAA forms bound to one or more amino acids and the ester-linked forms primarily bound to a sugar(s). These two types of conjugates appear to be found at varying concentrations in the diverse tissues of angiosperms (Domagalski et al., 1987). On average, 95% of all IAA in a plant is conjugated into these storage forms (Cohen and Bandurski, 1982; Bandurski et al., 1995; Campanella et al., 1996; Walz et al., 2002). There have been a variety of amide conjugates found in the plants studied to date. IAA-Asp has been identified as a natural conjugate in Scots pine 1 This work was supported in part by a Sokol grant for undergraduate research and in part by the Croatian Ministry of Science and Technology (grant no. 0098080). * Corresponding author; e-mail james.campanella@montclair. edu; fax 419–791–9834. Article, publication date, and citation information can be found at www.plantphysiol.org/cgi/doi/10.1104/pp.104.043398.

2230

(Pinus sylvestris; Andersson and Sandberg, 1982) and, together with IAA-Glu, in cucumber (Sonner and Purves, 1985) and soybean (Cohen, 1982). IAA-Ala ¨ stin et al., has been detected in Picea abies Karst (O 1992). Additionally, IAA-Ala, IAA-Asp, IAA-Leu, and IAA-Glu have been detected in Arabidopsis L. Heynh (Tam et al., 2000; Kowalczyk and Sandberg, 2001), although recent data suggest that IAA peptides may account for the majority of amide conjugates in this and other plant species (Walz et al., 2002; Park et al., 2003). Common IAA-ester conjugates include IAAmyoinositol glycosides, IAA-myoinositol, and IAAGlc (Bandurski et al., 1969; Kopcewicz et al., 1974; Bandurski and Schulze, 1977; Domagalski et al., 1987). These ester conjugates have been identified in many species over the last 60 years (Bandurski et al., 1969; Slovin et al., 1999). The ILR1-like IAA amidohydrolase gene family is thought to be involved in the regulation of free IAA concentrations. This gene family was originally characterized in the model plant Arabidopsis (Bartel and Fink, 1995; Davies et al., 1999; LeClere et al., 2002). The ILL1, ILL2, and IAR3 enzymes cleave IAA-Ala preferentially, while ILR1 prefers IAA-Leu and IAA-Phe as substrates. Another ILR1-like family member, sILR1, whose gene product has substrate specificity for IAAGly and IAA-Ala, has been isolated from the related species Arabidopsis suecica Norrl. ex O.E. Schulz (Campanella et al., 2003a, 2003d). Amidohydrolase homologs have also been isolated from Brassica rapa (Schuller and Ludwig-Mu¨ller, 2002). Additionally,

Plant Physiology, August 2004, Vol. 135, pp. 2230–2240, www.plantphysiol.org Ó 2004 American Society of Plant Biologists

Isolation of an Auxin Amidohydrolase from Wheat

several IAR3-like orthologs (MtIAR32, MtIAR33, MtIAR34, and MtIAR35) have been cloned from the non-Brassicaceae species Medicago truncatula Gaertn. (J.J. Campanella, J. Ludwig-Mu¨ller, and S. Wexler, unpublished data). Since many of the IAA conjugates in monocots are described as ester conjugates (Cohen and Bandurski, 1982), we hypothesized that IAA conjugate hydrolases isolated from a grass might primarily be ester conjugate hydrolases. The first ester conjugate hydrolase activity was found in maize (Hall and Bandurski, 1986). This work was followed by whole-seedling studies examining movement and hydrolysis of IAAinositol (Chisnell and Bandurski, 1988). More recently, Kowalczyk et al. (2003) identified an ester conjugate hydrolase in maize, but the corresponding gene has yet to be cloned. Therefore, our goal became the identification and isolation of a putative IAA conjugate hydrolase from a major model monocot. Our phylogenomic studies identified in wheat (Triticum aestivum) a full-length ortholog of the Arabidopsis IAA amidohydrolase IAR3 (Campanella et al., 2003c). We employed this sequence data to clone the full-length putative wheat hydrolase (TaIAR3) and characterize it. Our goals were to better understand how the functional molecular evolution of this important gene family has diverged between dicots and monocots, and which auxin conjugates play a specific role in monocots. Upon expression and enzymatic characterization of the TaIAR3 enzyme, we found that the enzymatic regulation of auxin in wheat may be more intricate than we anticipated. TaIAR3 is an amidohydrolase that preferably hydrolyzes amino acid conjugates of the long-side-chain auxins indole butyric acid (IBA) and indole propionic acid (IPA). This article reports the first functional characterization of an amidohydrolase able to recognize these long-side-chain conjugate species.

RESULTS Substrate Specificity of TaIAR3

The DNA and protein sequences of TaIAR3 show higher homology to the Arabidopsis ortholog IAR3 with 49% and 79% identity, respectively, than to any other member of the ILR1-like family. A phylogenetic analysis comparing the TaIAR3 protein against all the members of the ILR1-like family confirms that IAR3 and TaIAR3 are most similar (Fig. 1). Both the IAR3 and TaIAR3 products contain putative endoplasmic reticulum localization signals (data not shown), a hallmark of many of the auxin amidohydrolases (Davies et al., 1999; LeClere et al., 2002; Campanella et al., 2003a, 2003c). After cloning the complete cDNA of TaIAR3 (sequence obtained from the The Institute of Genomic Research [TIGR] Expressed Sequence Tag Index) into Plant Physiol. Vol. 135, 2004

Figure 1. Neighbor-joining phylogram of Arabidopsis ILR1-like protein family members orthologous to TaIAR3. The wheat hydrolase (bold) demonstrates the closest homology to the IAR3 protein sequence (bold); the high bootstrap value at the node of the tree (bold) indicates a probability score of approximately 98% for the validity of this prediction. The scale at the bottom of the figure indicates relative genetic distance.

pETBlue-2, the enzyme activity was determined in vitro (Table I; Fig. 2). Enzyme activity from cells containing an empty vector was compared with that from cells containing the peTaIAR3 vector. These cells were either incubated with Glc because the promoter from pETBlue-2 is leaky, or induced with isopropyl b–D-thiogalactopyranoside (IPTG). Relative expression levels of the TaIAR3 protein in Escherichia coli were determined by direct comparison of induced and uninduced bacterial proteins on SDS-PAGE (data not shown). Densitometry analysis indicated a 2.7-fold increase in expression of TaIAR3 in the induced cells over background expression. The level of enzyme activity was calculated after comparing auxin release in uninduced and IPTGinduced cells. We observed a novel substrate specificity for longer side-chain auxins, especially IBA-Ala. In addition, the amino acid attached to the auxin moiety was also important, as shown by comparing IBA-Ala and IBA-Gly as substrates. Methylation of their carboxyl groups abolished recognition by the enzyme almost completely. Interestingly, the Ala conjugate of an auxin with a side chain shorter by one methylene group (IPA) was also accepted as a good substrate. Further shortening of the side chain of the indole derivative (IAA) almost completely abolished enzyme activity. 2231

Campanella et al.

Table I. Enzyme activity of TaIAR3 expressed in E. coli For each assay, approximately 2.5 mg total E. coli protein was used. Mean values of three to five independent experiments 6SE are given. The values for which no SE is given have been measured only once. The activity is expressed as IAA formed in cells induced by IPTG, subtracted by values from noninduced cells. Substrate

Enzyme Activity nmol auxin released min21 mg21 protein

IAA-Asp IAA-Ala IAA-Gly IAA-Leu IAA-Ile IAA-Phe IAA-Val IAA-Glc IAA-inositol IPA-Ala IBA-Ala IBA-Gly Me-IBA-Ala Me-IBA-Gly

0.04 6 0.03 0.02 6 0.01 0.03 0.12 0.06 0 0 0.5 6 0.1 0.2 6 0.04 8.8 6 2.2 14.8 6 3.6 4.9 6 1.5 0 0.8

IBA-Ala were used. IAA-Ala was negligibly hydrolyzed in 3-d-old seedlings at approximately 20% of the enzyme activity toward the IBA-Ala substrate (Table II). As seedlings aged to 6, 15, and 17 d, no IAA-Ala hydrolase activity was detectable. IBA-Ala hydrolyzing enzyme activity was detected in all developmental stages investigated, being highest in 15- and 17-d-old seedlings. However, the analysis of whole seedlings could mask localized concentration changes in these compounds and in the hydrolysis activity.

Activity was specifically observed in cells after induction with IPTG as shown for IBA-Ala as substrate (Fig. 2). Cells containing an empty vector as well as uninduced cells showed only a very small peak at the retention time (Rt) of IBA. This demonstrates the specificity of the reaction. There were no other metabolites detectable (data not shown). The reaction product obtained after enzymatic hydrolysis of IBA-Ala with the Rt of IBA was collected from HPLC, subsequently methylated and analyzed by gas chromatography-mass spectrometry (GC-MS). The methylated putative IBA showed the same Rt as a methylated IBA standard in the ion chromatogram, and the mass spectrum showed the characteristic ions of the molecular ion at m/z 217 and the quinolinium ion at m/z 130 (data not shown). Other potential substrates for TaIAR3, such as IBA and IPA conjugates with Asp, will be investigated in the future. Determination of Free and Conjugated IAA and IBA and Amidohydrolase Activity in Wheat Seedlings

To test whether the endogenous auxin content would reflect the expression pattern of TaIAR3, the endogenous content of free and total IAA and IBA was determined during seedling growth (Fig. 3). While total IAA and IBA decreased as expected during the first few days after germination (IAA more than IBA), the concentration of free IAA did not change significantly; however, free IBA was significantly enhanced in 17-dold seedlings compared to 6-d-old seedlings, as determined by ANOVA (95% confidence). This is reflected in the amidohydrolase activity measured in vitro in extracts of seedlings when the substrates IAA-Ala and 2232

Figure 2. HPLC separation of the substrate, IBA-Ala, and the reaction product, IBA, after incubation of the substrate with E. coli cell lysate for 1 h. Chromatograms were recorded at 280 nm. A, Standards of IBA and IBA-Ala. B, Lysate of cells containing an empty vector. C, Lysate of cells containing the peTaIAR3 vector uninduced. D, Lysate of cells containing the peTaIAR3 vector induced with IPTG. Plant Physiol. Vol. 135, 2004

Isolation of an Auxin Amidohydrolase from Wheat

both young and mature plants (Fig. 4B). There is a slightly higher expression level of the transcript seen in coleoptiles during early development, and the same is seen in adult stem tissues. Presumably, the TaIAR3 gene product is required to help stimulate early growth of the coleoptile and continues to be of regulatory importance throughout the life of the seedling and into adulthood. Expression of the 18S control showed no difference between samples examined (data not shown).

DISCUSSION

Figure 3. Free and total IAA and IBA concentrations in wheat during seedling development. The values presented are means 6 SE of three independent experiments.

Identification of IBA-Ala in Wheat Seedlings

Extracts of wheat seedlings, as well as TaIAR3 in vitro, hydrolyzed IBA-Ala to a large extent. We therefore analyzed extracts from wheat seedlings for IBA-Ala as an endogenous compound. Following treatment with diazomethane, a peak was found in the ion chromatogram with the Rt of IBA-Ala methyl ester (data not shown). This peak showed a mass spectrum that is in accordance with the presence of IBA-Ala in wheat seedlings. The molecular ion at m/z 289, as well as the quinolinium ion at m/z 130, was present, although the spectrum contained several impurities. The latter are due to the problem that only very small amounts of IBA-Ala were present. Under the same conditions, no peak corresponding to IPA-Ala was detectable in wheat extracts, indicating that either very low concentrations are present or IPAAla is not a native substance in wheat seedlings.

Auxin conjugates play a major role in the regulation of auxin homeostasis. While, for two Arabidopsis species, conjugate hydrolases have been identified that can hydrolyze amino acid conjugates of IAA (LeClere et al., 2002; Campanella et al., 2003a), we describe here the activity of a novel auxin conjugate hydrolase activity from a monocot species (T. aestivum) toward amino acid conjugates with IBA. The isolation of an amidohydrolase from wheat became possible with the availability of the sequenced T. aestivum cDNAs from TIGR database, providing the opportunity to identify novel plant genes on the basis of homology searching. In this study, the sequence of the Arabidopsis IAR3 gene was employed to identify an orthologous gene (TaIAR3) in the wheat genome. It was found that the TaIAR3 protein, although it has high amino acid sequence homology to its Arabidopsis ortholog, is able to cleave conjugates of auxin molecules with longer side chains, IBA and IPA, while all other auxin conjugate hydrolases so far identified cleaved amino acid conjugates with IAA (LeClere et al., 2002; Campanella et al., 2003a). In addition, the TaIAR3 mRNA shows a different spatial and temporal regulation compared to IAR3 from Arabidopsis (Davies et al., 1999; Rampey et al., 2003), suggesting that the IAR3 and TaIAR3 enzymes may be employed in different developmental pathways in Arabidopsis and wheat. IBA is a naturally occurring auxin commonly used to induce rooting (Hartmann et al., 1990). Several studies have demonstrated within the last decade that IBA is present in vascular plants (Epstein et al., 1989; Ludwig-Mu¨ller and Epstein, 1991, 1993; Epstein and

Temporal and Spatial Expression of TaIAR3

The TaIAR3 transcript is first detectable in seedlings at day 1 after germination and appears to fall steadily in concentration until day 20 (Fig. 4A). It may be that the high expression of the enzyme is required for growth very early in plant development, but the expression is down-regulated after that significant phase of growth has passed. Expression of the 18S control showed no difference between samples examined (data not shown). Spatial expression of the TaIAR3 transcript appears fairly homogeneous throughout the wheat tissues in Plant Physiol. Vol. 135, 2004

Table II. Auxin conjugate hydrolase activity from wheat seedlings during development using IAA-Ala and IBA-Ala as substrates Mean values of three different enzyme preparations 6SE are given. Enzyme Activity

Age of Seedlings IBA-Ala days after germination

0 3 6 15 17

IAA-Ala

pM auxin released min21 mg21 protein

1.0 1.4 3.7 1.8

0 6 6 6 6

0.1 0.1 0.3 0.2

0 0.4 6 0.06 0 0 0 2233

Campanella et al.

Figure 4. Temporal (A) and spatial (B) expression of TaIAR3 transcripts. Graphs were generated using the relative standard curve method (Applied Biosystems, 2001). A, Calibrator was the 1-d-old plant expression data. B, Calibrator was the coleoptile expression data.

Ludwig-Mu¨ller, 1993; Ludwig-Mu¨ller et al., 1993). IBA appears to be stored like IAA in a conjugated form with amino acids or sugars (Epstein and LudwigMu¨ller, 1993; Ludwig-Mu¨ller et al., 1993), allowing its slow hydrolysis and release. Moreover, IBA may be an endogenous precursor to IAA (Epstein and Lavee, 1984; Nordstro¨m et al., 1991; van der Krieken et al., 1992; Baraldi et al., 1993). Conversely, IAA can be converted to IBA in several species, including maize and Arabidopsis (Epstein and Ludwig-Mu¨ller, 1993; Ludwig-Mu¨ller and Hilgenberg, 1995; Ludwig-Mu¨ller et al., 1995a; Ludwig-Mu¨ller, 2000). IBA has been mostly disregarded in plant physiology studies until recently because there were questions of 2234

whether or not it was a natural plant hormone. Several studies suggest that IBA is important during the rooting process (Nordstro¨m et al., 1991; van der Krieken et al., 1992). Ludwig-Mu¨ller et al. (1993) found that up to 30% of auxin in Arabidopsis may be IBA. Some experimental evidence suggests that IBA may act as an endogenous growth hormone on both roots and stems without conversion to IAA (Ludwig-Mu¨ller et al., 1995b, 1995c; Chhun et al., 2003), although the recent analysis of IBA-resistant Arabidopsis mutants revealed that at least some of the auxin activity of IBA results from its conversion to IAA by peroxisomal b-oxidation (Zolman et al., 2000, 2001; Zolman and Bartel, 2004). IPA was first found in decomposing organic matter, likely as a Trp metabolite of microorganisms participating in the decay process. Bacteria can indeed produce IPA (Mohammed et al., 2003), but this capability is limited to certain species (Elsden et al., 1976), while the compound is inhibitory to others (Matsuda et al., 1998; Walker et al., 2003). Some microbes have acquired the ability to detoxify IPA by conjugation. Bacillus megatherium, for instance, couples the acid to Glc, Ala, Ser, and Thr (Tabone and Tabone, 1953; Tabone, 1958). Seed plants appear to take advantage of the sensitivity of some bacteria to IPA to suppress pathogens; the compound was secreted by the roots of axenically grown Arabidopsis treated with salicylic acid, a compound that is usually formed in response to microbial attack (Walker et al., 2003). Further research will show if the occurrence of IPA in (nonsterile) Cucurbita pepo hypocotyls (Segal and Wightman, 1982) and in the roots of Pisum sativum seedlings (Schneider et al., 1985) and the preliminary evidence for its presence in Brassica oleracea (Linser et al., 1954), Nicotiana tabacum leaves (Bayer, 1969), and Lycopersicon esculentum Mill cotyledons (Aung, 1972) should be viewed as a response to bacterial attack, as a result of bacterial contamination, or as evidence for a role in plant development. Exogenous IPA is somewhat less active than IAA in stem-elongation assays (Fawcett et al., 1960; Katekar, 1979), but more active in the promotion of lateral root growth on the primary root of pea seedlings (Wightman et al., 1980). Conjugated forms of IPA have, to our knowledge, not been identified in seed plants, but it is likely that their roots are consistently exposed to these compounds as metabolites of soil bacteria. Three factors suggest that, in wheat, IBA-amide conjugates are more likely to be the endogenous target of the enzyme. First, we have identified the presence of IBA and IBA conjugates in wheat seedlings, where the enzymatic synthesis of IBA has also been demonstrated (Ludwig-Mu¨ller et al., 1995c), but we have been unable to detect IPA conjugates. The very presence of IBA conjugates strongly supports the hypothesis that they are in vivo substrates. Second, although TaIAR3 is able to cleave IPA-Ala conjugates efficiently, it has a higher activity for IBA-Ala. Third, IBA, but not IPA, synthesis was shown to occur in maize and IBA synthesis was also demonstrated in wheat (LudwigMu¨ller and Epstein, 1991; Ludwig-Mu¨ller et al., 1995c). Plant Physiol. Vol. 135, 2004

Isolation of an Auxin Amidohydrolase from Wheat

Although the enzymatic hydrolysis of IBA conjugates has not been previously described in the literature, it is unclear whether this activity is unique to monocots. Hydrolases from other plant species, such as A. suecica (sILR1; Campanella et al., 2003d) and M. truncatula (MtIAR32; J. J. Campanella, J. LudwigMu¨ller, and S. Wexler, unpublished data) tested so far have little or no activity against IBA-Ala and others have not been reported. Monocots diverged from dicots approximately 200 to 250 million years ago (Wolfe et al., 1989; Troitsky et al., 1991; Kellogg, 2001). Therefore, IAR3 must predate this point of divergence, since it is conserved not only in the angiosperms but in the gymnosperms as well (Campanella et al., 2003c). The gymnospermangiosperm divergence took place 300 to 360 million years ago, suggesting that the ILR1-like family is quite ancient and that its role is intimately tied to growth and development in vascular plants (Troitsky et al., 1991; Oliviusson et al., 2001; Albert et al., 2002; Cooke et al., 2002, 2004). From the experimental evidence provided so far, it seems that different strategies to regulate auxin content have evolved. While in charophytes and liverworts the auxin conjugation rates are very slow and therefore biosynthesis is a major contributor to free auxin, the conjugation rate increases from hornworts and mosses to vascular plants and thus also its importance in auxin homeostasis (Cooke et al., 2002, 2004). With the characterization of the TaIAR3 gene product, we are faced with additional questions of how monocots and dicots differ in their regulation of auxin homeostasis. Mapelli et al. (1995) reported that the primary conjugates in wheat appear to be the IAAester type; however, the TaIAR3 enzyme appears to have little to do with the release of free IAA from the sugar esters tested, although a small activity was observed using IAA-ester conjugates. This result may indicate an important parallel or redundant role for IBA conjugate hydrolysis in growth induction. One strategy to regulate IAA-ester conjugates occurs in maize, where it was demonstrated that the enzyme catalyzing the conjugation of IAA to myoinositol was also able to catalyze the hydrolysis of the IAA-ester conjugate (Kowalczyk et al., 2003). Amide-type auxin conjugates, including acyl peptides and proteins (Walz et al. 2002), were not found in wheat seeds (Bandurski and Schulze, 1977), but their occurrence in vegetative tissues cannot be ruled out. There appears to be a contradiction between the temporal expression data of TaIAR3, suggesting the transcript decreases in expression over 20 d, and the enzymatic analysis data, suggesting an increase in enzymatic activity as the seedlings age. This may simply indicate a difference in how plants grown under varying environmental conditions express the enzyme, or this phenomenon could result from the disparities in the growth conditions causing different physiological states between various types of experiments. In addition, the TaIAR3 protein may be Plant Physiol. Vol. 135, 2004

stable and accumulate in seedlings while the transcript level is being down-regulated, or there may be posttranscriptional regulatory mechanisms involved. Spatial expression of TaIAR3 indicates that, in seedlings and adult plants, the transcript is expressed at approximately the same level in most tissues examined, supporting the belief that the enzyme is required in some maintenance capacity in all tissues over the life of the plant. However, loss-of-function mutants for IAR3 in Arabidopsis are phenotypically wild type, suggesting that IAR3 is not required for early development or mature growth (Davies et al., 1999). Experiments are under way to isolate and characterize additional wheat hydrolase orthologs and also orthologs from maize so that we may determine the roles of each and their particular substrate specificity. Functional characterization in future studies will provide essential information on how concentrations of the different auxins may be regulated in our model wheat and how this reflects on other plant species. MATERIALS AND METHODS Plant Materials and Plant Growth Winter wheat (Triticum aestivum cv Caledonia) seeds were obtained commercially from Johnny’s Selected Seeds (Albion, Maine). Plants were raised under conditions optimized to fit the needs of the individual experiments performed, avoiding, in the interpretations, quantitative comparisons between batches cultivated in different ways. Seeds for sources of root, upper and lower leaf, and stem tissue were sown in soil (perlite:sphagnum peat moss:vermiculite, 1:1:1, v/v/v) saturated with liquid minimal medium (5 mM KNO3, 2.5 mM K2HPO4, pH 5.5, 2 mM MgSO4, 2 mM Ca(NO3)2.H2O, 50 mM Fe-EDTA, 70 mM H3BO3, 14 mM MnCl2, 0.5 mM CuSO4, 1 mM ZnSO4, 0.2 mM Na2MoO4.2H2O, 10 mM NaCl, 0.01 mM CoCl2). Plants were slowly hardened off over 1 week and fertilized with liquid minimal medium every 2 to 3 weeks as needed. The plants were grown at 23°C in constant light (cool-white, fluorescent, approximately 100 mmol s21 m22) in a plant growth chamber (Model E-30B; Percival Scientific, Perry, IA). Wheat seeds for harvest of coleoptile and radicle tissues were germinated in 250-mL flasks with 50 mL of liquid Murashige and Skoog medium (SigmaAldrich, St. Louis). The flasks were agitated at approximately 100 rpm at 23°C in constant light (cool-white, fluorescent, approximately 100 mmol s21 m22) in a plant growth chamber (Model E-30B; Percival Scientific). Leaves, roots, and stems for spatial expression analysis were harvested 70 d after germination. Coleoptiles and radicles were harvested at 4 d after germination. All tissues were stored frozen at 280°C until RNA extraction. For developmental expression studies in seedlings, seeds were germinated in 250-mL flasks with 50 mL of liquid Murashige and Skoog medium. The flasks were agitated at approximately 100 rpm at 23°C in constant light in a plant growth chamber. Seedlings were collected 1, 2, 3, 4, 5, 10, 15, and 20 d after germination and stored frozen at 280°C until RNA extraction. Wheat seeds used in enzyme activity measurement and endogenous auxin determination were placed on moist filter paper, covered with plastic, and incubated for 48 h at 4°C. The seeds were then placed at 23°C (60% humidity) with a 16-h day/8-h night cycle. Day 1 after germination was specified as the point when the radicles were first visible. The seedlings were then placed in moist lecaton (expanded clay substrate) and harvested after the appropriate time period after germination. Seedlings were watered every 2 d.

RNA Extraction Total RNA was extracted from approximately 0.2 g of plant tissue, using the RNeasy RNA extraction kit (Qiagen, Valencia, CA). Before extraction, micropestles and all microfuge tubes were treated with an 8% solution of RNA Secure (Ambion, Austin TX) for 10 min at 65°C. RNA concentration was determined by UV absorbance (Spectronic Genesys 5 spectrophotometer;

2235

Campanella et al.

Thermo Electron, Waltham, MA) and samples stored at 280°C as separate aliquots until real-time reverse transcription (RT)-PCR analysis could be performed.

Cloning of Full-Length TaIAR3 cDNA The cDNA sequence for TaIAR3 (T. aestivum TIGR accession no. TC150449, GenBank accession no. AY701776) was obtained from TIGR sequence database and first identified as an IAR3 ortholog (Sambrook et al., 1989). Total wheat mRNA was extracted from whole seedlings and RT-PCR was then used to amplify the specific cDNA of TaIAR3. The transcript primers used to amplify TaIAR3 for cloning were 5#-AGAACGTCACGGCGGAATT-3# (TaIAR3F) and 5#-CAAACAGTTGCATAATCACCTG-3# (TaIAR3R). A single 50-mL reaction was carried out in an RNase-free 0.5-mL microfuge tube using a Mastercycler gradient thermocycler (Eppendorf Scientific,Westbury, NY). The RT reaction was incubated at 50°C for 1 h, followed by 95°C for 15 min. The PCR step was performed for 36 cycles at the following times and temperatures: 45 s at 95°C, 45 s at 57°C, and 1 min at 72°C. The resulting amplified cDNA was blunt-end ligated, in-frame, into the EcoRV cloning site of the pETBlue-2 expression vector (Novagen, Madison, WI) using T4 DNA ligase (Novagen) in such a way that the cDNA was terminated by its own stop codon and the His-tag of the pETBlue-2 vector was not expressed. The resulting construct, peTaIAR3, was then transformed into Escherichia coli (NovaBlue) using heat shock (Sambrook et al., 1989) and putative transformants selected on the basis of ampicillin resistance and bluewhite selection (Sambrook et al., 1989). Plasmids were obtained from transformants by alkaline lysis (Sambrook et al., 1989). Insert orientation and size were determined by endonuclease double digestion with BglI and BglII, electrophoretic analyses through 1% agarose gels, and DNA sequencing of the insert region. The DNA/protein alignment and similarity calculations were performed using the MatGAT v2.0 computer program (Campanella et al., 2003b). All analyses were performed using default values.

Real-Time RT-PCR Total RNA from wheat plants was used for real-time RT-PCR to examine the expression of the TaIAR3 hydrolase gene. Analyses were performed on two biological replicates for each treatment. The transcript-specific primers used to amplify TaIAR3 were 5#-GTGGTGGAGCCTTCAATGTT-3# (TaIAR3 Exp F) and 5#-CAAACAGTTGCATAATCACCTG-3# (TaIAR3R). As an expression control for use in quantitation, universal 18S primers (Ambion) were included in the same reaction mixes. This mixture of two sets of primers constituted a duplexed RT-PCR reaction in which the primers were able to amplify two different transcripts without interfering with each other (data not shown). All quantitative real-time analyses were performed as single-tube reactions using approximately 1,500 ng of wheat mRNA with components of a One-Step RT-PCR kit (Qiagen).This single 50-mL reaction was carried out in an RNasefree 0.5-mL microfuge tube using a Mastercycler gradient thermocycler (Eppendorf). The RT reaction was incubated at 50°C for 1 h, followed by 95°C for 15 min. At the end of the RT reaction, 18S competimer primers (Ambion) were added to the reaction tube to ensure that the abundant 18S transcript did not overwhelm the hydrolase transcripts in amplification. The ratio of 18S primers to 18S competimers in the reaction was 3:7. The PCR step was performed for 32 to 44 cycles at the following times and temperatures: 45 s at 95°C, 45 s at 57°C, and 1 min at 72°C. Ten 5-mL aliquots were removed from the reaction tubes at the even-numbered cycles, starting at cycle 14, 18, 20, 22, or 24, depending on the profile of initially detected expression in each sample studied. After sampling, the reaction tube was placed back on the thermocycler, during temperature ramping between cycles, to proceed with PCR. Each 5-mL aliquot was frozen at 220°C and stored for analysis. The aliquots were analyzed by agarose gel electrophoresis and stained with ethidium bromide. The RT-PCR products were imaged using an Ultralum gel documentation system (Ultralum, Claremont, CA) and Scion computer software (Scion Pharmaceuticals, Medford, MA). Densitometry was performed on each cDNA band by application of the ImageTool Analysis program (University of Texas Health Science Center, San Antonio). Each experiment was repeated two to three times and densitometry values averaged.

2236

The bar graphs indicating TaIAR3 expression were generated from the real-time RT-PCR data. The standard curve method described by Applied Biosystems (2001) was employed for this analysis. In this method, product amounts were normalized to the 18S standard and a relative standard curve was generated employing a basis sample, called a calibrator. For all experimental samples, target quantity was determined from the standard curve and divided by the target quantity of the calibrator. The calibrator acted as the 13 sample, and all other samples were expressed in differences relative to the calibrator.

Enzyme Preparation from Bacteria The peTaIAR3 strain (containing the wheat homolog of IAR3 in the EcoRV site of the pETBlue-2 vector) was grown overnight in 5 mL Luria-Bertani medium containing 100 mg mL21 ampicillin. From this culture, 2 mL were transferred to a flask containing 50 mL Luria-Bertani medium including 100 mg mL21 ampicillin and 1 mM IPTG for gene induction. Induction was performed for 4 h under continuous shaking of the cultures. Uninduced controls were grown under the same conditions, but without IPTG. Instead, 0.5% Glc was included in the medium since the promoter of pETBlue-2 is leaky. Enzyme preparation and enzyme assays with TaIAR3 were expressed in E. coli (Campanella et al., 2003d). After collecting the bacterial cells by centrifugation for 10 min at 8,000g, the pellet was resuspended in 100 mL lysozyme buffer per initial 1 mL bacterial culture (30 mM Tris-HCl, pH 8.0, containing 1 mM EDTA, 20% Suc, and 1 mg mL21 lysozyme [Sigma]) for cell lysis and incubated for 10 min at 4°C. To break the cells, three freeze-thaw cycles were performed where the cells were frozen in liquid N2 and thawed at 30°C. After the last thawing, 2 mL of a 10 mg mL21 DNase solution (in 150 mM NaCl, 50% glycerol) were added, and the mixture was incubated for 15 min at RT. The extract (100-mL volume per assay) was then directly used for the enzyme assay.

Enzyme Assay with TaIAR3 The enzyme assay for the hydrolysis of auxin conjugates was performed in a 500-mL reaction mixture containing 395 mL assay buffer, 100 mL bacterial enzyme extract (corresponding to approximately 2.5 mg total protein), and 5 mL of a 10 mM stock solution (dissolved in a small volume of ethanol, then diluted with H2O) of each substrate (final concentration 100 mM; ethanol concentration was always less then 0.1%). The substrates used in this study were the IAA-amide conjugates IAA-Asp, IAA-Ala, IAA-Gly, IAA-Leu, IAAIle, IAA-Phe, and IAA-Val (all from Sigma), the IAA-ester conjugates IAA-Glc and IAA-myoinositol (both gifts from Dr. Jerry D. Cohen), the amide conjugate of IPA with Ala (synthesis described below), the amide conjugates of IBA with Ala (Sigma) and Gly (obtained after demethylation of methyl-IBA-Gly, see below), and the methyl esters of IBA-Ala and IBA-Gly (obtained from Dr. Joseph Riov, The Volcani Center, Bet Dagan, Israel). The assay buffer consisted of 100 mM Tris, pH 8.0, 10 mM MgCl2, 100 mM MnCl2, 50 mM KCl, 100 mM PMSF, 1 mM DTT, and 10% Suc (Ludwig-Mu¨ller et al., 1996). The reaction was routinely incubated for 1 h at 40°C. The reaction was stopped by adding 100 mL 1 N HCl, and the aqueous phase was then extracted with 600 mL ethyl acetate. The organic phase was removed, evaporated to dryness, and resuspended in 100 mL CH3OH for HPLC analysis. For GC-MS analysis, the samples were evaporated to dryness, redissolved in 100 mL ethyl acetate, methylated with diazomethane (Cohen, 1984), and resuspended in ethyl acetate. The experiments were repeated three to five times using different enzyme preparations. All results represent means of independent experiments 6SE. As controls, the uninduced cultures were evaluated. The enzymatic activity was calculated as nanomolar IAA released from cultures induced with IPTG, subtracted by the values obtained in noninduced cultures.

HPLC Analysis The total CH3OH extract (100 mL, see above) was subjected to HPLC (BT 8100 pumps; Jasco, Easton, MD) coupled to an autosampler (AS-1550; Jasco), equipped with a 125-mm 3 4.6-mm i.d. Lichrosorb C18 (particle size 5 mm) reverse-phase column and a multiwavelength diode array detector (MC-919; Jasco) set at 280 nm. As solvent, 1% aqueous CH3COOH (solvent A) and 100% CH3OH (solvent B) were used. The program used was: 0 min, 25% B; 20 min, 25% B; 10 min, 100% B; 5 min, 25% B; 5 min equilibration, 25% B. Flow rate was

Plant Physiol. Vol. 135, 2004

Isolation of an Auxin Amidohydrolase from Wheat

1 mL min21. The BORWIN chromatography software (JMBS Developments Software for Scientists; Varian, Palo Alto, CA) was used. Identification of IAA, IBA, and IPA was achieved by comparison with authentic standards. IAA was separated in this system from all conjugates used as substrates with a Rt of 23.4 min (Ludwig-Mu¨ller et al., 1996). The amount of the free IAA, IPA, or IBA released was determined using a standard curve for the respective standard substance.

GC-MS Analysis GC-MS analysis was carried out on a Varian Saturn 2100 ion-trap mass spectrometer using electron impact ionization at 70 eV, connected to a Varian CP-3900 gas chromatograph equipped with a CP-8400 autosampler (Varian). For the analysis, 2.5 mL of the methylated sample dissolved in 20 mL ethyl acetate was injected in the splitless mode (splitter opening 1:100 after 1 min) onto a Phenomenex ZB-5 column, 30 m 3 0.25 mm i.d. 3 0.25 mm using He carrier gas at 1 mL min21. The injector temperature was 250°C and the temperature program was 70° for 1 min, followed by an increase of 20° min21 to 280°C, then 5 min isothermically at 280°C. Transfer line temperature was 280°C. Either full-scan mass spectra were recorded or, for higher sensitivity, the mSIS mode (Varian Manual) was used monitoring ions 130 and 189 (for IAA methyl ester) or 217 (for IBA methyl ester). The scan rate was 0.6 s scan21, the multiplier offset voltage was 200 V, the emission current was 30 mA, and the trap temperature was 200°C. To identify the IBA enzymatically released from its conjugates, the respective peak from HPLC was collected, evaporated, methylated, and analyzed by GC-MS. IBA from the sample was identified according to the Rt on GC compared with an authentic methylated standard and by recording the respective mass spectrum.

Methylation and Demethylation of IBA Conjugates To test whether the methyl esters of IBA amino acid conjugates were also hydrolyzed, IBA-Ala was methylated using diazomethane. Briefly, the appropriate concentration of the conjugate was dissolved in 50 mL ethyl acetate, then 950 mL of a freshly prepared solution of diazomethane in diethyl ether were added, and the mixture was incubated for 15 min at room temperature (Cohen, 1984) The residual ether was evaporated and the methylated compound dissolved in a small volume of CH3OH, or directly in ethyl acetate, for GC-MS. The methanolic extract of Me-IBA-Ala was checked on HPLC in an isocratic system (1% aqueous CH3COOH:CH3OH, 40:60, v/v), at a flow rate of 0.7 mL min21 (Rt IBA-Ala, 8.8 min; Rt IBA-Ala methyl ester, 22.8 min). It was shown that no unreacted IBA-Ala was present as an impurity (data not shown). The methyl ester of IBA-Gly was demethylated to yield IBA-Gly by dissolving 100 mM methyl ester of IBA-Gly in 70% CH3OH. To this solution, 20 mL of a 2 N NaOH solution was added (final pH approximately 10). The pH was monitored over the reaction time and was readjusted, if necessary. The progress of the reaction was checked by thin-layer chromatography (TLC), using silica gel plates (Merck, Rahway, NJ) and ethyl acetate:isopropanol: NH4OHconc (45:35:20, v/v/v) as solvent. The methyl ester of IBA-Gly had an RF value of approximately 1, IBA-Gly approximately 0.5. After 6-h reaction time, the reaction did not proceed further and was therefore stopped by evaporating the CH3OH and adding 2 N HCl to the aqueous phase until the pH was adjusted to 1.5. The H2O phase was extracted four times with equal volumes of ethyl acetate. The organic phases were combined, dried over anhydrous Na2SO4, evaporated to dryness, and resuspended in CH3OH. The IBA-Gly was purified by HPLC. The HPLC system used for the purification of IBA-Gly was an isocratic system (1% aqueous CH3COOH:CH3OH, 40:60, v/v), at a flow rate of 0.7 mL min21 (Rt methyl ester of IBA-Gly, 21.4 min; Rt IBA-Gly, 7.3 min). The peak corresponding to IBA-Gly was collected, evaporated to dryness, and resuspended in a small volume of CH3OH. For enzymatic assays, this methanolic extract was diluted with H2O as for the other auxin conjugates.

Synthesis of N-[3-(Indol-3-yl)propionyl]-L-Ala In the syntheses, commercial, analytical grade chemicals and solvents were used without further purification. Only dioxane was redistilled over sodium, immediately before use, to ensure the absence of peroxides. Melting points were determined in open capillaries and were not corrected. High-resolution mass spectra were obtained on a Micromass Q-TOF2 instrument using electrospray ionization (positive ion mode; capillary voltage, 3 kV; cone

Plant Physiol. Vol. 135, 2004

voltage, 30 V; source temperature, 80°C; desolvation temperature, 150°C; nitrogen flow, 480 L h21; sample applied in acetonitrile:water, 2:8, containing 0.1% formic acid). Nuclear magnetic resonance (NMR) spectra were taken on a Bruker AV600 instrument operating at 600 MHz for 1H and at 150 MHz for 13 C. Chemical shifts are reported in parts per million relative to internal tetramethylsilane. TLC was performed on glass plates (5 3 10 cm), coated with silica gel GF254 (Merck), to a thickness of 0.3 mm. Chromatograms were developed with solvents A (2-propanol:ethyl acetate:NH4OHconc, 35:45:20, v/v/v) and B (CH2Cl2:CH3OH:CH3COOH, 90:10:5, v/v/v). Detection was by UV absorbance and by spraying with Ehrlich’s reagent (1% [w/v] 4-dimethylaminobenzaldehyde in a 1:1 [v/v] mixture of ethanol and concentrated HCl), which affords blue to purple spots with the indole derivatives studied here. To a stirred solution of 3-(indol-3-yl) propionic acid (1 g, 5.29 mmol) and N-hydroxysuccinimide (639 mg, 5.55 mmol) in a mixture of dioxane (6 mL) and anhydrous ethyl acetate (4 mL), 1,3-dicyclohexylcarbodiimide (1.20 g, 5.81 mmol) was added in small aliquots, for 20 min, at 28°C. After 2 h of further stirring in an ice-water bath, TLC (solvents A and B) indicated that the reaction was complete. The precipitated 1,3-dicyclohexylurea was removed by filtration and rinsed with ethyl acetate. The filtrate was evaporated to yield the crude 3-(indol-3-yl) propionic acid N-hydroxysuccinimide ester, RF 5 0.6 (solvent A) and 0.4 (solvent B). The latter was redissolved in dioxane (20 mL), a solution of L-Ala (471 mg, 5.29 mmol) in 10% aqueous NaHCO3 (20 mL) was added, and the resulting suspension was stirred at room temperature, for 1 h, when TLC (solvent A) indicated that the reaction was complete. The mixture was diluted with water (20 mL, pH after dilution, 9.6) and nonacidic byproducts were extracted with diethyl ether (2 3 30 mL). The aqueous phase was acidified to pH 2.5 and partitioned against ethyl acetate (5 3 50 mL). The organic phase was dried over anhydrous sodium sulfate and evaporated to yield the crude title compound. Repeated crystallization from ethyl acetate: cyclohexane (approximately 1:1, v/v) afforded off-white crystals (1.02 g, 74%), RF 5 0.4 (solvent A), melting point 179°C to 181°C. High-resolution mass spectrum (time-of-flight instrument, electron spray ionization with positive ion monitoring): 261.1218 (MH1; C14H17N2O3 requires 261.1239), 172.0701 ([M - Ala]1; C11H10NO requires 172.0762), 130.0655 (quinolinium ion; C9H8N requires 130.0657) amu. 1H-NMR spectrum [(CD3)2SO]: 3-(indol-3-yl) propionyl moiety: d 10.71 (1 H, s, H-1), 7.06 (1 H, narrow m, H-2), 7.47 (1 H, d, J4,5 5 7.8 Hz, H-4), 6.91 (1 H, t, J5,6 5 7.1 Hz, H-5), 7.00 (1 H, t, H-6), 7.27 (1 H, d, J6,7 5 8.1 Hz, H-7), 2.42 (1 H, dt, Jgem 5 14.7 Hz, Jvic 5 7.6 Hz, ArCHI), 2.46 (1 H, dt, ArCHII), 2.87 (2 H, t, CH2CO); L-Ala moiety: d 8.16 (1 H, d, JCH,NH 5 7.2 Hz, NH), 4.19 (1 H, quintet, JCH2,CH3 5 7.3 Hz, CH), 1.20 (1 H, t, CH3), 12.46 (1 H, broad s, COOH). 13C-NMR spectrum [(CD3)2SO]: 3-(indol-3-yl) propionyl moiety: d 122.2 (C-2), 113.9 (C-3), 127.1 (C-3a), 118.3 (C-4), 118.2 (C-5), 120.9 (C-6), 111.3 (C-7), 136.3 (C-7a), 20.9 (ArCH2), 36.0 (CH2CO), 171.8 (CONH); L-Ala moiety d: 17.3 (CH3), 47.5 (CHNH), 174.3 (COOH).

Determination of Endogenous Auxins Free and Conjugated IAA and IBA Wheat seedlings were harvested after different times of culture (3, 6, 10, 15, and 17 d after germination), the roots were washed thoroughly and the tissue was blotted dry between sheets of filter paper. The plant material (approximately 0.1 g fresh weight per analysis) was extracted as previously described (Chen et al., 1988; Ludwig-Mu¨ller and Cohen, 2002). Each extract was divided into two equal parts for determination of free auxins and for alkaline hydrolysis. To each sample, 200 ng 13C6-IAA (Cambridge Isotope Laboratories, Andover, MS), and 100 ng 13C1-IBA (Sutter and Cohen, 1992) were added. For each sample, three independent extractions were performed. Free auxins were purified on NH2 columns (Chen et al., 1988) and, after elution from the column, the samples were evaporated to dryness, directly methylated with diazomethane (Cohen, 1984), and resuspended in ethyl acetate for GC-MS analysis. Hydrolysis of conjugated auxins was performed with 7 N NaOH at 100°C under N2 for 3 h. The hydrolysate was filtered, the pH adjusted to 2.5, and the auxins were extracted twice with equal volumes of ethyl acetate. The organic phase was evaporated and the extract methylated as described above. GC-MS analysis was performed with the program described above for the analysis of enzymatically released IBA. Endogenous auxin concentrations were calculated using the isotope dilution equation (Ilic´ et al., 1996). The experiments were performed three times using different plant extracts. All results present the arithmetic means of independent experiments 6SE. Statistical analysis of the data was done using analysis of variance (ANOVA)

2237

Campanella et al.

employing a Web-based program (http://www.physics.csbsju.edu/stats/ anova.html).

Identification of Endogenous IBA-Ala The nonhydrolyzed methanolic extracts from wheat seedlings were subjected to GC-MS analysis after methylation with diazomethane (see above). As a standard, the methyl ester of IBA-Ala was subjected to the same GC-MS procedure. The settings used for separation were as described above, with the following modifications: Injector temperature was 250°C and the temperature program was 70°C for 2 min, followed by an increase of 10°C min21 to 300°C, then 10 min isothermically at 300°C. Transfer line temperature was 300°C. Endogenous IBA-Ala was identified according to the Rt of its methyl ester in the ion chromatogram (26.1 min) and the respective mass spectrum was recorded and compared to that of an authentic IBA-Ala standard.

In Vitro Assay for Auxin Conjugate Hydrolysis Using Wheat Seedlings The determination of enzymatic activity in plant extracts was carried out as in Ludwig-Mu¨ller et al. (1996). The same buffer as for the enzymatic assays with bacteria was used. Briefly, the seedlings were harvested for enzyme preparation, the roots were thoroughly washed and then blotted dry between sheets of filter paper. The plants were extracted in a 100 mM Tris-HCl buffer, pH 8, containing 10 mM MgCl2, 100 mM MnCl2, 50 mM KCl, 1 mM DTT, 100 mM PMSF, and 10% Suc. The homogenate was filtered and then centrifuged for 30 min at 50,000g. The supernatant was used for the hydrolase assays. The enzyme assay for the conversion of IAA conjugates to free IAA was performed in a 500-mL reaction mixture containing 300 mL of enzyme extract and 50 mM of individual substrate (IAA-Ala and IBA-Ala). The enzyme reaction was carried out at pH 8.0 for 60 min at 40°C, as described above. Inactivated extracts (boiling for 10 min) were also incubated and used as controls to determine nonenzymatic hydrolysis of auxin conjugates. These values were subtracted from the values determined with active enzyme extract. The reaction was stopped by adding 100 mL 1 N HCl, and the aqueous phase was then extracted with 500 mL ethyl acetate. The organic phase was removed, evaporated to dryness, and resuspended in 20 mL CH3OH. The sample was kept in liquid nitrogen prior to HPLC analysis. Protein was determined with the Bradford protein assay reagent (Sigma) using bovine serum albumin as standard (Bradford, 1976). The experiments were performed three times using different enzyme preparations. All results represent the arithmetic means of independent experiments 6SE.

Phylogenetic Tree Construction Amino acid alignments of TaIAR3, ILR1, IAR3, ILL1, ILL2, ILL3, and ILL6 orthologs were performed using ClustalX v1.8 software (Thompson et al., 1997). Alignment settings were employed at default values. Phylogenetic trees were generated from the distances provided by the ClustalX analysis using the neighbor-joining method (Saitou and Nei, 1987). Bootstrap analyses (Felsenstein, 1985) consisted of 1,000 replicates. The neighbor-joining trees were visualized with the TREEVIEW program (Page, 1996). The protein sequence of a bacterial M20 peptidase from Campylobacter jejuni (GenBank accession no. Z36940) was used as an outgroup. Sequence data from this article have been deposited with the EMBL/ GenBank data libraries under accession number AY701776.

ACKNOWLEDGMENTS We thank Dr. Joseph Riov, The Volcani Center, Bet Dagan, Israel, for the gift of the methyl ester of IBA-Gly, Dr. Ellen G. Sutter, University of California, Davis, for 13C1-IBA, Dr. Jerry D. Cohen, University of Minnesota, St. Paul, for IAA-ester conjugates with Glc and myoinositol, and Igor Bratos, PLIVA Pharmaceuticals, Zagreb, Croatia, for high-resolution mass spectra. The helpful discussions with Dr. Ephraim Epstein are gratefully acknowledged. We thank Silvia Heinze and Dr. Campanella’s graduate molecular biology laboratory class for technical assistance. Finally, we wish to thank Lisa Campanella for her generous editorial help.

2238

Received March 23, 2004; returned for revision June 4, 2004; accepted June 7, 2004.

LITERATURE CITED Albert VA, Oppenheimer DG, Lindqvist C (2002) Pleiotropy, redundancy and the evolution of flowers. Trends Plant Sci 7: 297–301 Andersson B, Sandberg G (1982) Identification of endogenous N-(3indoleacetyl)aspartic acid in Scots pine (Pinus sylvestris L.) by combined gas chromatography-mass spectrometry, using high-performance liquid chromatography for quantification. J Chromatogr 238: 151–156 Applied Biosystems Incorporated (2001) User Bulletin #2: ABI PRISM 7700 Sequence Detection System. http://www.appliedbiosystems.com Aung LH (1972) The nature of root-promoting substances in Lycopersicon esculentum seedlings. Physiol Plant 26: 306–309 Bandurski RS, Cohen JD, Slovin JP, Reinecke DM (1995) Auxin biosynthesis and metabolism. In PJ Davies, ed, Plant Hormones: Physiology, Biochemistry and Molecular Biology, Ed 2. Kluwer Academic Publishers, Dordrecht, The Netherlands, pp 39–65 Bandurski RS, Schulze A (1977) Concentrations of IAA and its derivatives in plants. Plant Physiol 60: 211–213 Bandurski RS, Ueda M, Nicholls PB (1969) Esters of indole-3-acetic acid and myo-inositol. Ann N Y Acad Sci 165: 655–667 Baraldi R, Bertazza G, Predieri S, Bregoli AM, Cohen JD (1993) Uptake and metabolism of indole-3-butyric acid during the in vitro rooting phase in pear cultivars (Pyrus communis). Acta Hortic 329: 289–291 Bartel B, Fink G (1995) ILR1, an amidohydrolase that releases active indole-3-acetic acid from conjugates. Science 268: 1745–1748 Bayer M (1969) Gas chromatographic analysis of acidic indole auxins in Nicotiana. Plant Physiol 44: 267–271 Bradford MM (1976) A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal Biochem 72: 248–254 Campanella JJ, Bakllamaja V, Restieri T, Vomacka M, Herron J, Patterson M, Shahtaheri S (2003a) Isolation of an ILR1 auxin conjugate hydrolase homolog from Arabidopsis suecica. Plant Growth Regul 39: 175–181 Campanella JJ, Bitincka L, Smalley J (2003b) MatGAT: an application that generates similarity/identity matrices using protein or DNA sequences. BMC Bioinformatics 4: 29 Campanella JJ, Larko D, Smalley J (2003c) A molecular phylogenomic analysis of the ILR1-like family of IAA amidohydrolase genes. Comp Funct Genomics 4: 584–600 Campanella JJ, Ludwig-Mu¨ller J, Bakllamaja V, Sharma V, Cartier A (2003d) ILR1 and sILR1 IAA amidohydrolase homologs differ in expression pattern and substrate specificity. Plant Growth Regul 41: 215–223 Campanella JJ, Ludwig-Mu¨ller J, Town CD (1996) Isolation and characterization of mutants of Arabidopsis thaliana with increased resistance to growth inhibition by indoleacetic acid-amino acid conjugates. Plant Physiol 112: 735–745 Chen KH, Miller AN, Patterson GW, Cohen JD (1988) A rapid and simple procedure for purification of indole-3-acetic acid prior to GC-SIM-MS analysis. Plant Physiol 86: 822–825 Chhun T, Taketa S, Tsurumi S, Ichii M (2003) The effects of auxin on lateral root initiation and root gravitropism in a lateral rootless mutant Lrt1 of rice (Oryza sativa L.). Plant Growth Regul 39: 161–170 Chisnell JR, Bandurski RS (1988) Translocation of radiolabeled indole-3acetic acid and indole-3-acetyl-myo-inositol from kernel to shoot in Zea mays L. Plant Physiol 86: 79–84 Cohen JD (1982) Identification and quantitative analysis of indole-3-acetylL-aspartate from seeds of Glycine max L. Plant Physiol 70: 749–753 Cohen JD (1984) Convenient apparatus for the generation of small amounts of diazomethane. J Chromatogr 303: 193–196 Cohen JD, Bandurski RS (1982) The chemistry and physiology of the bound auxins. Annu Rev Plant Physiol 33: 403–430 Cooke TJ, Poli DB, Cohen JD (2004) Did auxin play a crucial role in the evolution of novel body plans during the Late-Silurian-Early Devonian radiation of land plants? In AR Hemsley, I Poole, eds, The Evolution of Plant Physiology: From Whole Plants to Ecosystems. Linnean Society of London and Elsevier Academic Press, Amsterdam, pp 85–107 Cooke TJ, Poli DB, Sztein AE, Cohen JD (2002) Evolutionary patterns in auxin action. Plant Mol Biol 49: 319–338

Plant Physiol. Vol. 135, 2004

Isolation of an Auxin Amidohydrolase from Wheat

Davies PJ (1995) The plant hormones: their nature, occurrence, and functions, In PJ Davies, ed, Plant Hormones: Physiology, Biochemistry, and Molecular Biology. Kluwer Academic Publishers, Dordrecht, The Netherlands, pp 13–38 Davies RT, Goetz DH, Lasswell J, Anderson MN, Bartel B (1999) IAR3 encodes an auxin conjugate hydrolase from Arabidopsis. Plant Cell 11: 365–376 Domagalski W, Schulze A, Bandurski RS (1987) Isolation and characterization of esters of indole-3-acetic acid from liquid endosperm of the horse chestnut (Aesculus species). Plant Physiol 84: 1107–1113 Elsden SR, Hilton MG, Waller JM (1976) The end products of the metabolism of aromatic amino acids by Clostridia. Arch Microbiol 107: 283–288 Epstein E, Chen K-H, Cohen JD (1989) Identification of indole-3-butyric acid as an endogenous constituent of maize kernels and leaves. Plant Growth Regul 8: 215–223 Epstein E, Lavee S (1984) Conversion of indole-3-butyric acid to indole-3acetic acid by cuttings of grapevine (Vitis vinifera) and olive (Olea europea). Plant Cell Physiol 25: 692–703 Epstein E, Ludwig-Mu¨ller J (1993) Indole-3-butyric acid in plants: occurrence, synthesis, metabolism and transport. Physiol Plant 88: 382–389 Fawcett CH, Wain RI, Wightman F (1960) The metabolism of 3-indolylalkanecarboxylic acids, and their amides, nitriles and methyl esters in plant tissues. Proc R Soc Lond B Biol Sci 152: 231–254 Felsenstein J (1985) Confidence limits on phylogenies: an approach using the bootstrap. Evolution 39: 783–791 Hall PJ, Bandurski RS (1986) [3H]Indole-3-acetyl-myo-inositol hydrolysis by extracts of Zea mays L. vegetative tissue. Plant Physiol 80: 374–377 Hartmann HT, Kester DE, Davies FT (1990) Plant Propagation: Principles and Practices. Prentice-Hall, Englewood Cliffs, NJ, pp 246–247 Ilic´ N, Normanly J, Cohen JD (1996) Quantification of free plus conjugated indole-3-acetic acid in Arabidopsis requires correction for the nonenzymatic conversion of indolic nitriles. Plant Physiol 111: 781–788 Katekar G (1979) Auxins: on the nature of the receptor site and molecular requirements for auxin activity. Phytochemistry 18: 223–233 Kellogg EA (2001) Evolutionary history of the grasses. Plant Physiol 125: 1198–1205 Kopcewicz J, Ehmann A, Bandurski RS (1974) Enzymatic esterification of indole-3-acetic acid to myo-inositol and glucose. Plant Physiol 54: 846–851 Kowalczyk M, Sandberg G (2001) Quantitative analysis of indole-3-acetic acid metabolites of Arabidopsis. Plant Physiol 127: 1845–1853 Kowalczyk S, Jakubowska A, Zielinska E, Bandurski RS (2003) Bifunctional indole-3-acetyl transferase catalyses synthesis and hydrolysis of indole-3-acetyl-myo-inositol in immature endosperm of Zea mays. Physiol Plant 119: 165–174 LeClere S, Tellez R, Rampey RA, Matsuda SPT, Bartel B (2002) Characterization of a family of IAA-amino acid conjugate hydrolases from Arabidopsis. J Biol Chem 277: 20446–20452 Linser H, Mayr H, Maschek F (1954) Papierchromatographie von zellstreckend wirksamen Indolko¨rpern aus Brassica-.Arten. Planta 44: 103–120 Ludwig-Mu¨ller J (2000) Indole-3-butyric acid in plant growth and development. Plant Growth Regul 32: 219–230 Ludwig-Mu¨ller J, Cohen JD (2002) Identification and quantification of three active auxins in different tissues of Tropaeolum majus. Physiol Plant 115: 320–329 Ludwig-Mu¨ller J, Epstein E (1991) Occurrence and in vivo biosynthesis of indole-3-butyric acid in corn (Zea mays L.). Plant Physiol 97: 765–770 Ludwig-Mu¨ller J, Epstein E (1993) Indole-3-butyric acid in Arabidopsis thaliana. II. In vivo metabolism. Plant Growth Regul 13: 189–195 Ludwig-Mu¨ller J, Epstein E, Hilgenberg W (1996) Auxin-conjugate hydrolysis in Chinese cabbage: characterization of an amidohydrolase and its role during infection with clubroot disease. Physiol Plant 97: 627–634 Ludwig-Mu¨ller J, Hilgenberg W (1995) Characterization and partial purification of indole-3-butyric acid synthetase from maize (Zea mays). Physiol Plant 94: 651–660 Ludwig-Mu¨ller J, Hilgenberg W, Epstein E (1995a) The in vitro biosynthesis of indole-3-butyric acid in maize. Phytochemistry 40: 61–68 Ludwig-Mu¨ller J, Raisig A, Hilgenberg W (1995b) Uptake and transport of indole-3-butyric acid in Arabidopsis thaliana: comparison with other natural and synthetic auxins. J Plant Physiol 147: 351–354

Plant Physiol. Vol. 135, 2004

Ludwig-Mu¨ller J, Sass S, Sutter EG, Wodner M, Epstein E (1993) Indole-3butyric acid in Arabidopsis thaliana. I. Identification and quantification. Plant Growth Regul 13: 179–187 Ludwig-Mu¨ller J, Schubert B, Pieper K (1995c) Regulation of IBA synthetase from maize (Zea mays L.) by drought stress and ABA. J Exp Bot 46: 423–432 Mapelli S, Locatelli F, Bertani A (1995) Effect of anaerobic environment on germination and growth of rice and wheat: endogenous levels of ABA and IAA. Bulg J Plant Physiol 21: 33–41 Matsuda K, Toyoda H, Nishio H, Nishida T, Dohgo M, Bingo M, Matsuda Y, Yoshida S, Harada S, Tanaka H, et al (1998) Control of the bacterial wilt of tomato plants by a derivative of 3-indolepropionic acid based on selective actions on Ralstonia solanacearum. J Agric Food Chem 46: 4416– 4419 Mohammed N, Onodera R, Or-Rashid MM (2003) Degradation of tryptophan and related indolic compounds by ruminal bacteria, protozoa and their mixture in vitro. Amino Acids 24: 73–80 Nordstro¨m AC, Jacobs FA, Eliasson L (1991) Effect of exogenous indole-3acetic acid and indole-3-butyric acid on internal levels of the respective auxins and their conjugation with aspartic acid during adventitious root formation in pea cuttings. Plant Physiol 96: 856–861 Oliviusson P, Salaj J, Hakman I (2001) Expression pattern of transcripts encoding water channel-like proteins in Norway spruce. Plant Mol Biol 46: 289–299 ¨ stin A, Moritz T, Sandberg G (1992) Liquid chromatography/mass O spectrometry of conjugates and oxidative metabolites of indole-3-acetic acid. Biol Mass Spectrom 21: 292–298 Page RD (1996) TREEVIEW: an application to display phylogenetic trees on personal computers. Comput Appl Biosci 12: 357–358 Park S, Walz A, Momonoki YS, Ludwig-Mu¨ller J, Slovin JP, Cohen JD (2003) Partial characterization of major IAA conjugates in Arabidopsis. Proceedings of the American Society of Plant Biologists Conference, July 25–30, 2003, Honolulu, HI. http://abstracts.aspb.org/pb2003/public/ P46/0961.html Rampey RA, LeClere S, Bartel B (2003) The roles of Arabidopsis IAA conjugate hydrolases in auxin homeostasis. Proceedings of the American Society of Plant Biologists Conference, July 25–30, 2003, Honolulu, HI. http://abstracts.aspb.org/pb2003/public/P46/0446.html Saitou N, Nei M (1987) The neighbor-joining method: a new method for reconstructing phylogenetic trees. Mol Biol Evol 4: 406–425 Sambrook J, Maniatis T, Fritsch EF (1989) Molecular Cloning: A Laboratory Manual, Ed 2. Cold Spring Harbor Laboratory, Cold Spring Harbor, NY Schneider EA, Kazakoff CW, Wightman F (1985) Gas chromatographymass spectrometry evidence for several endogenous auxins in pea seedling organs. Planta 165: 232–241 Schuller A, Ludwig-Mu¨ller J (2002) Isolation of differentially expressed genes involved in clubroot disease. Plant Prot Sci 38: 483–486 Segal LM, Wightman F (1982) Gas chromatographic and GC-MS evidence for the occurrence of 3-indolylpropionic acid and 3-indolylacetic acid in seedlings of Cucurbita pepo. Physiol Plant 56: 367–370 Slovin JP, Bandurski RS, Cohen JD (1999) Auxins. In PJJ Hooykaas, MA Hall, KR Libbenga, eds, Biochemistry and Molecular Biology of Plant Hormones. Elsevier, Amsterdam, The Netherlands, pp 115–140 Sonner JM, Purves WK (1985) Natural occurrence of indole-3-acetyl aspartate and indole-3-acetyl glutamate in cucumber shoot tissue. Plant Physiol 77: 784–785 Sutter EG, Cohen JD (1992) Measurement of indolebutyric acid in plant tissues by isotope dilution gas chromatography-mass spectrometry analysis. Plant Physiol 99: 1719–1722 Tabone D (1958) Biosynthe`se par B. megatherium de combinaisons de l’acide indol propionique avec certains acides amine´s. Bull Soc Chim Biol (Paris) 40: 965–969 Tabone J, Tabone D (1953) Bio-este´rification du glucose. V. Bio-synthe`se par Bacillus megatherium de l’ester b glucosidique de l’acide indolpropionique. C R Hebd Seances Acad Sci 237: 943–944 Tam YY, Epstein E, Normanly J (2000) Characterization of auxin conjugates in Arabidopsis. Low steady-state levels of indole-3-acetyl aspartate, indole-3-acetyl glutamate, and indole-3-acetyl glucose. Plant Physiol 123: 589–595 Thompson JD, Gibson TJ, Plewniak F, Jeanmougin F, Higgins DG (1997) The ClustalX Windows interface – flexible strategies for multiple sequence alignment aided by quality analysis tools. Nucleic Acids Res 25: 4876–4882

2239

Campanella et al.

Troitsky AV, Melekhovets YF, Rakhimova GM, Bobrova VK, ValiejoRoman KM, Antonov AS (1991) Angiosperm origin and early stages of seed plant evolution deduced from rRNA sequence comparisons. J Mol Evol 32: 253–261 Van der Krieken W, Bretler H, Visser M (1992) The effect of the conversion of indolebutyric acid into indoleacetic acid on root formation on microcuttings of Malus. Plant Cell Physiol 33: 709–713 Walker TS, Bais HP, Halligan KM, Stermitz FR, Vivanco JM (2003) Metabolic profiling of root exudates of Arabidopsis thaliana. J Agric Food Chem 51: 2548–2554 Walz A, Park S, Slovin JP, Ludwig-Mu¨ller J, Momonoki YS, Cohen JD (2002) A gene encoding a protein modified by the phytohormone indoleacetic acid. Proc Natl Acad Sci USA 99: 1718–1723 Wightman F, Schneider EA, Thimann KV (1980) Hormonal factors controlling the initiation and development of lateral roots. II. Effects

2240

of exogenous growth factors on lateral root formation in pea roots. Physiol Plant 49: 304–314 Wolfe KH, Gouy M, Yang YW, Sharpe PM, Li WH (1989) Date of the monocot-dicot divergence estimated from chloroplast DNA sequence data. Proc Natl Acad Sci USA 86: 6201–6205 Zolman BK, Bartel B (2004) An Arabidopsis indole-3-butyric-acid-response mutant defective in PEROXIN6, an apparent ATPase implicated in peroxisomal function. Proc Natl Acad Sci USA 101: 1786–1791 Zolman BK, Silva ID, Bartel B (2001) The Arabidopsis pxa1 mutant is defective in an ATP-binding cassette transporter-like protein required for peroxisomal fatty acid b-oxidation. Plant Physiol 127: 1266–1278 Zolman BK, Yoder A, Bartel B (2000) Genetic analysis of indole-3-butyric acid responses in Arabidopsis thaliana reveals four mutant classes. Genetics 156: 1323–1337

Plant Physiol. Vol. 135, 2004