Approaching Cellular and Molecular Resolution of ...

2 downloads 0 Views 900KB Size Report
From mu- tant analysis, it appears that indolic secondary metabolites such as IGs and camalexin are a major metabolic sink for Trp (Bender and. Celenza 2009).
Approaching Cellular and Molecular Resolution of Auxin Biosynthesis and Metabolism Jennifer Normanly Department of Biochemistry and Molecular Biology, University of Massachusetts, Amherst, Massachusetts 01003 Correspondence: [email protected]

There is abundant evidence of multiple biosynthesis pathways for the major naturally occurring auxin in plants, indole-3-acetic acid (IAA), and examples of differential use of two general routes of IAA synthesis, namely Trp-dependent and Trp-independent. Although none of these pathways has been completely defined, we now have examples of specific IAA biosynthetic pathways playing a role in developmental processes by way of localized IAA synthesis, causing us to rethink the interactions between IAA synthesis, transport, and signaling. Recent work also points to some IAA biosynthesis pathways being specific to families within the plant kingdom, whereas others appear to be more ubiquitous. An important advance within the past 5 years is our ability to monitor IAA biosynthesis and metabolism at increasingly higher resolution.

he topic of auxin biosynthesis and metabolism in plants was comprehensively reviewed in 2005 (Woodward and Bartel 2005). Since then, more genes involved in IAA biosynthesis and metabolism have been identified. A combination of numerous valuable mutants, the manipulation of IAA synthesis in specific cell types, and direct measurement of IAA levels at tissue and cellular resolution now point to localized IAA biosynthesis and metabolism as playing key roles in specific developmental events (reviewed in Cheng and Zhao 2007; Lau et al. 2008; Zhao 2008; Chandler 2009). With apologies to any authors who were not included because of space constraints, this review summarizes those recent findings that require us to rethink yet again, the role

T

of IAA biosynthesis and metabolism in auxin biology. IAA SYNTHESIS OCCURS THROUGHOUT PLANTS

Within the past decade, the long-held hypothesis that IAA is synthesized primarily in the apical region and transported throughout the plant to form morphogenic gradients has undergone revision and refinement, concomitant with the ability to monitor de novo IAA synthesis in specific tissues at high resolution. Deuterium oxide (2H2O) is readily taken up by seedlings, and deuterium will exchange with hydrogen in the noncyclized intermediates early in the shikimate pathway (Fig. 1). Once

Editors: Mark Estelle, Dolf Weijers, Karin Ljung, and Ottoline Leyser Additional Perspectives on Auxin Signaling available at www.cshperspectives.org Copyright # 2010 Cold Spring Harbor Laboratory Press; all rights reserved; doi: 10.1101/cshperspect.a001594 Cite this article as Cold Spring Harb Perspect Biol 2010;2:a001594

1

J. Normanly COO– PEP

H2O

2 2

H

2

C

COO–

HO

O

C

H

H

2

H [2H]chorismate

+ 2

E4P

HO H2O

H

H

OH

H

OH

OH

O

O

OH

CH2OPO3H2

OH

NH2

2

DHQ

H

DAHP

2

H

2

H

2

H

2

[ H]ANA

OPO3H2 OH

2

H

2

H

2

OH N H

H 2

2 2

H

H

2

[ Hn]IGP

H O

H

[2Hn]indole

C H 2

H

2

H

N H 2

H

OH 2

2

H

H

H

O

2

H OH

[2Hn]IAA

H 2

H 2

N H

2

NH2

H

H [2Hn]Trp

Figure 1. Monitoring de novo IAA synthesis by stable isotope labeling. Live plants readily take up deuterium

oxide (2H2O) from the medium, and deuterium readily exchanges with hydrogens in the course of normal metabolism. In the case of IAA, it has been demonstrated that indole-ring deuteriums will not exchange with hydrogens, even in the presence of strong base (Magnus et al. 1980), so measuring the rate of incorporation of deuterium into IAA over a short time period (typically 24 h) indicates the rate of de novo IAA synthesis, as opposed to the hydrolysis of IAA conjugates. n refers to a variable number of deuteriums, from one to five. (E4P) Erythrose-4-phospate, (PEP) phosphoenol pyruvate, (DAHP) 3-deoxy-D-arabino-heptulosonate 7-phosphate, (DHQ) dehydroquinate, (ANA) anthranilate, (IGP) indole-3-glycerol phosphate, (Trp) tryptophan, (IAA) indole-3-acetic acid.

2

Cite this article as Cold Spring Harb Perspect Biol 2010;2:a001594

Auxin Biosynthesis and Metabolism

incorporated, deuterium in the 4, 5, 6, and 7 positions, and to a lesser extent the 2 position of the indole ring in IAA are stable and will not exchange with hydrogen, even under alkaline conditions (Magnus et al. 1980). Thus, measuring the incorporation of deuterium into the indole ring of IAA is a measure of de novo IAA synthesis, as opposed to hydrolysis of IAA from pre-existing IAA conjugates. This approach has revealed that all parts of young seedlings are capable of synthesizing IAA (Ljung et al. 2001a; Bhalerao et al. 2002) with newly fertilized embryos, young leaves, and roots exhibiting high IAA synthesis activity (Ljung et al. 2001a; Ljung et al. 2001b; Ribnicky et al. 2002). Removal of aerial tissues early in seedling development and inclusion of the auxin transport inhibitor naphthalene acetic acid (NPA) in 2H2O labeling experiments reveals that shoot-derived IAA is important for lateral root emergence early in development. Later in development, auxin synthesis in the primary root is sufficient to sustain lateral root initiation and can drive lateral root emergence independently of shoot derived IAA (Bhalerao et al. 2002; Ljung et al. 2005). A very recent study using multiple Arabidopsis lines expressing celltype specific GFP constructs allowed fluorescence activated cell sorting (FACS) of roots into cell types. 2H2O labeling of these sorted cells reveals that virtually all cells in the root apex are capable of synthesizing IAA (Petersson et al. 2009). These and other experiments described later support the hypothesis that it is a combination of IAA transport and localized IAA synthesis that form and maintain IAA gradients throughout the plant. PROGRESS IN FILLING IN THE GAPS FOR MULTIPLE BIOSYNTHETIC ROUTES TO IAA

The biosynthetic pathways for IAA (Fig. 2) can be classified as Trp-dependent if IAA is derived via metabolism of Trp, or as Trp-independent if IAA is derived from an early indolic precursor of Trp (reviewed in Normanly et al. 2004; Woodward and Bartel 2005), most likely indole3-glycerol phosphate (IGP) (Ouyang et al. 2000). Stable isotope labeling studies conducted in

a variety of plant species have consistently demonstrated the differential utilization of Trpdependent and Trp-independent IAA synthesis pathways at critical times in plant development (Ljung et al. 2002), including embryogenesis (Michalczuk et al. 1992; Ribnicky et al. 2002), fruit ripening (Epstein et al. 2002), in response to changes in temperature (Rapparini et al. 2002), during germination and early seedling growth (Ljung et al. 2001b), and following wounding (Sztein et al. 2002). Decades of investigation have established that there are multiple Trp-dependent biosynthetic routes to IAA in plants (Normanly et al. 2004; Woodward and Bartel 2005) and microbes (Glick et al. 1999). Microbial IAA biosynthesis pathways are well defined, but none of the proposed IAA biosynthetic pathways in plants have been completely defined (dashed arrows in Fig. 2 represent pathway steps that have not been confirmed). Substantial progress has been made in the identification of IAA biosynthesis genes. An update follows. More Evidence for the Indole-Pyruvic Acid (IPA) Pathway in Plants

The IPA pathway (Fig. 2) has been well characterized in microbes (Koga et al. 1991; Koga et al. 1992; Koga 1995). IPA has been identified as a native compound in Arabidopsis (Tam and Normanly 1998). Recently, TAA1 (Trp aminotransferase of Arabidopsis) has been shown to catalyze the formation of IPA from L-Trp in vitro (Stepanova et al. 2008; Tao et al. 2008), and mutants of TAA1 accumulate less IAA in response to simulated shade (Tao et al. 2008), ethylene (Stepanova et al. 2008), and probably high temperature (Yamada et al. 2009). As discussed later, the phenotypes of TAA1 and related TAR mutants provide specific examples of the role of localized IAA biosynthesis in the response to environmental and developmental cues. Whether the subsequent steps of this IAA synthesis pathway are similar to those in the microbial IPA pathway remains to be determined. Sequence analysis, including motif identification, predicts five pyruvate decarboxylases that are candidates for an indole pyruvate

Cite this article as Cold Spring Harb Perspect Biol 2010;2:a001594

3

J. Normanly ASA1,2 ASB1,2,3 PAT1 PAI1,2,3 IGS ANA Chorismate PANA CADP trp1 wei2 wei7 jdl1

OH O

TSA-like?

O

IGP N H

trp 3-1

OH

P

OH

indole

OH

N H

TSA1 indole TSB1,2

OH

CYP83B1 (CYP83A1)

O

trp 2-1

N

IAOx N H

OH

CYP79B2,3

TRP

cyp79B2cyp79B3

sur2 N-OH-TRM

IAOx N-oxide

YUCCA

NH2

N H

TRM wei8 sav3 tir2

TAA1 TAR1,2 S-IAH-L-cys C-S lyase

sur1

OH

IAM

indole-3-T-OH

OH

IPA

pad4 CYP71A13

UGT74B1 ST5a

OH

N H

AMI1

OSO–3 N

IG

S-glucose O

N H

TGG1,2 ESP ESM1

IAAld N H N

N

N

IAN

IAN

IAN

N H

N H

N H

CYP71B15

CysIAN pad3

Aldehyde oxidase NIT 1,2,3

nit 1

OH

DHCA

IAA

CYP71B15 pad3

O

N H

S N

camalexin N H

Figure 2. Proposed IAA biosynthesis pathways for Arabidopsis. Dashed arrows indicate that neither a gene nor

enzyme activity has been identified in Arabidopsis. (ANA) anthranilate, (PANA) 5-phosphoribosylanthranilate, (CADP) 1-(o-carboxyphenylamino)-1-deoxyribulose-5-phosphate, (IGP) indole-3-glycerol phosphate, (TRP) tryptophan, (IAM) indole-3-acetamide, (IPA) indole-3-pyruvic acid, (IAAld) indole-3-acetaldehyde, (IAOx) indole-3-acetaldoxime, (S-IAH-L-cys) S-(indolylacetohydroximoyl)-L-cysteine, (indole-3-T-OH) indole-3thiohydroximate, (IG) indole-3-methylglucosinolate, (TRM) tryptamine, (IAN) indole-3-acetonitrile, (DHCA) dihydrocamalexic acid. Gene abbreviations are given in upper case italics. Mutant alleles are indicated in lower case italics. In the case of aldehyde oxidase, this enzyme activity has been proposed for the corresponding conversion, but a definitive gene assignment has not been made (Sekimoto et al. 1998; Seo et al. 1998). Arabidopsis encodes at least four myrosinases, TGG1 and TGG2 (Barth and Jander 2006), and two others (reviewed in Halkier and Gershenzon 2006), and at least one epithiospecifier (ESP) locus (Lambrix et al. 2001) and an epithiospecifier modifier (ESM1) gene has been identified (Burow et al. 2008). Mutant phenotypes and references therein for trp2, trp3, cyp79b2cyp79b3, sur1, sur2, and nit 1 are described in Woodward and Bartel 2005 and in Last and Fink 1988 for trp1. Mutants identified since the Woodward and Bartel review include wei2, wei7 (Stepanova et al. 2005), jdl1 (Sun et al. 2009), wei8 (Stepanova et al. 2008), sav3 (Tao et al. 2008), tir2 (Yamada et al. 2009), pad3 (Zhao et al. 1999; Bottcher et al. 2009), and pad4 (Nafisi et al. 2007).

4

Cite this article as Cold Spring Harb Perspect Biol 2010;2:a001594

Auxin Biosynthesis and Metabolism

decarboxylase (IPDC) in Arabidopsis. All five genes were expressed in Escherichia coli, and a sensitive in vitro assay failed to demonstrate IPDC activity for any of the genes. Additionally, IPDC activity was not detected in protein extracts from wild-type Arabidopsis (Ye and Cohen 2009), so the enzymatic conversion of IPA to IAA may be very different from what is seen in microbes.

Is the Indole-3-Acetaldoxime (IAOx) Pathway Limited to Indole Glucosinolate (IG)-producing Species?

Although the molecular genetic tools associated with Arabidopsis have played a large role in the identification of many of the proposed IAA biosynthesis genes, dissection of IAA biosynthesis pathways in Arabidopsis is complicated by the way in which Trp metabolism and secondary metabolism intersect in this species. Specifically, some plants in the order Capparales, to which Arabidopsis belongs, uniquely make indole glucosinolates (IGs) and other indole-derived defense compounds (e.g., camalexin) (Kjaer 1974) from IAOx, a metabolite of Trp (Fig. 2). From mutant analysis, it appears that indolic secondary metabolites such as IGs and camalexin are a major metabolic sink for Trp (Bender and Celenza 2009). IAOx can be diverted to IAA synthesis from pathways that (a) primarily function in defense against biotic stress (Hull et al. 2000; Bak and Feyereisen 2001; Bak et al. 2001; Zhao et al. 2002; Bednarek et al. 2009; Clay et al. 2009; Mikkelsen et al. 2009), and (b) are not ubiquitous in the plant kingdom (Sugawara et al. 2009). IAOx has been measured in tobacco, rice, maize (Sugawara et al. 2009), peas (Quittenden et al. 2009), Brachypodium distachyon, and the aliphatic glucosinolate-producing watercress (J. Normanly, unpubl.), and found to be below the limits of detection (5–10pg). Based upon these observations, the IAOx pathway, which appears to have both indole-3-acetonitrile (IAN) and indole-3-acetamide (IAM) as parallel intermediates (Fig. 2), has been proposed to be specific to the IG-producing plant species (Sugawara et al. 2009).

Does the “YUCCA Pathway” Employ IAOx as an Intermediate?

Members of the YUCCA family of flavin monooxygenases have been implicated in IAA biosynthesis in maize (Gallavotti et al. 2008), rice (Yamamoto et al. 2007), petunia (Toben˜aSantamaria et al. 2002), and tomato (Exposito-Rodrıguez et al. 2007), in addition to Arabidopsis (Zhao et al. 2001; Kim et al. 2007), where these genes were first discovered. The proposed activity for YUCCA, conversion of tryptamine (TRM) to N-hydroxyl-TRM (Fig. 2), is based upon in vitro evidence (Zhao et al. 2001; Kim et al. 2007), and genetic validation of this pathway has not included profiling TRM levels in YUCCA overexpression lines (Zhao et al. 2001) or loss-of-function mutants (Cheng et al. 2006; Chen et al. 2007). N-hydroxyl TRM has been proposed to be a precursor for IAOx (Zhao et al. 2001), but whereas label from [2H5]TRM was incorporated into indole-3-acetaldehyde (IAAld), indole-3-ethanol, and IAA in pea roots, neither N-OH TRM nor IAOX were identified as labeled intermediates (Quittenden et al. 2009). IAOx levels were measured in the Arabidopsis yuc1 yuc2 yuc4 yuc6 quadruple mutant and found to be indistinguishable from wild type (Sugawara et al. 2009); however, given the restricted expression patterns of these four genes (Cheng et al. 2006), it is likely that any localized differences in IAOx levels would have been masked by measuring bulk tissue, even though the sampled tissue exhibited a visible phenotype. The Arabidopsis YUCCA overexpression mutant does not accumulate IAOx, but does accumulate IAN (J. Celenza, J. Cohen, and J. Normanly, unpubl.), which could be the result of ectopically expressing YUCCA. That is, the product of YUCCA may normally be restricted to cells that do not have enzymes capable of converting it to IAOx and subsequently IAN. Additionally, there may be separate IAOx pools in IGproducing plants. IAOx destined for IG synthesis is derived from the CYP79B2 and CYP79B3 pathway, and if IAOx is also derived from YUCCA, it does not contribute to IG synthesis (reviewed in Bender and Celenza 2009). The

Cite this article as Cold Spring Harb Perspect Biol 2010;2:a001594

5

J. Normanly

observation that [15N]TRM did not label IAOx in vivo in wild-type Arabidopsis, but did label IAA (Sugawara et al. 2009) leaves us with more questions about the role of IAOx in the YUCCA pathway. Nitrilase and IAN

IAN has been a challenging intermediate to work with in Arabidopsis because of the complexity of IG metabolism (reviewed in Bones and Rossiter 2006; Bender and Celenza 2009). Conversion of IAOx to IAA in Arabidopsis was recently proposed to occur with IAN and IAM as parallel intermediates, based on incorporation of label from IAOx (Sugawara et al. 2009). Label from IAOx was incorporated into both IAN and IAM, but to different degrees, seeming to rule out a precursor product relationship between IAN and IAM. There may be multiple pools of IAN (Bender and Celenza 2009), not all of which are available for synthesis of IAA, and this may complicate the interpretation of the labeling data. For example, IAN is implicated in the synthesis of the defense compound camalexin (Nafisi et al. 2007). The P450 CYP71A13 converts IAOX to IAN in vitro, and exogenous IAN is required to restore camalexin levels in mutants with defects in CYP71A13. IAN levels were not measured in the CYP71A13 mutant, but the fact that exogenous IAN was required implies that if there were indeed other endogenous sources of IAN, they were not available. Tobacco does not accumulate detectable levels of IAOx or IAN (Sugawara et al. 2009), yet ectopic expression of AtCYP79B2 or AtCYP79B3 leads to an accumulation of both and to increases in IAA as well (H. Nonhebel, J. Celenza, J. Cohen, J. Normanly, unpubl.). This implies that enzyme activities exist to convert IAOx to IAN and IAN to IAA in nonglucosinolate plants. Maize, which also does not appear to make IAOx, has at least two nitrilases, Zmnit1 and Zmnit2, that are proposed to have dual functions in kernels and seedlings as a heterodimer in the hydrolysis of b-cyanoalanine (Kriechbaumer et al. 2007), and Zmnit2 is proposed to convert IAN to IAA ( purified Zmnit2 does so at a rate

6

of 100 nmol/mg protein/min) (Kriechbaumer et al. 2006; Kriechbaumer et al. 2007). In contrast, IAN is not detectable in maize endosperm, which is a source of highly efficient Trpdependent IAA synthesis activity; 32 pmol/ mg/min (Kriechbaumer et al. 2006) or 140 pmol/mg protein/min (Hendrickson Culler and Cohen 2007). IAN is not a substrate for Trp-dependent IAA synthesis in this system (Hendrickson Culler and Cohen 2007). Interestingly, none of the other proposed intermediates in Trp-dependent IAA synthesis (Fig. 2) were good substrates in this system. Adding radiolabeled Trp to the endosperm assay yielded a small membrane protein to which Trp was covalently attached through a thioester bond (Hendrickson Culler and Cohen 2007). This protein-Trp conjugate was a substrate for IAA synthesis, releasing radiolabeled IAA, and upon hydrolysis of the thioester, racemic Trp.

Trp-independent IAA Synthesis

The Trp-independent IAA synthesis pathway has been established in numerous plant species by way of stable isotope labeling studies (reveiewed in Woodward and Bartel 2005), yet remains undefined genetically. New alleles of iar4, a putative E1a subunit of mitochondrial pyruvate dehydrogenase, provide a possible link to the regulation of this pathway. The phenotypes of iar4-3 and iar 4-4 suggest a dual role for this protein in auxin homeostasis and the normal metabolic function of the acetyl-CoA-generating pyruvate dehydrogenase complex (Quint et al. 2009). The auxin response phenotypes in these mutants can be suppressed by increasing endogenous IAA, either by raising the temperature as in Gray et al. 1998, or ectopic expression of YUCCA, suggesting a deficiency in IAA levels. IAA biosynthesis was not impaired, but dual labeling that was able to distinguish between Trp-dependent and Trp-independent IAA synthesis showed that the Trp-independent pathway had been up-regulated. Levels of IAAglutamate, a presumed storage form of IAA, were elevated in the mutants, implying a possible defect in conversion to free IAA.

Cite this article as Cold Spring Harb Perspect Biol 2010;2:a001594

Auxin Biosynthesis and Metabolism

LOCALIZED IAA BIOSYNTHESIS CONTRIBUTES TO MAINTENANCE OR CHANGES IN AUXIN MAXIMA ESTABLISHED BY AUXIN TRANSPORT

IAA exerts its effects through the formation of local gradients and maxima/minima of IAA; thus, the regulation of IAA concentrations is central to the many roles that it plays in plant growth and development. The formation of IAA gradients in plant organs has been demonstrated directly in several systems by measuring IAA concentrations in segments of tissues, using stable isotope labeled internal standards and gas chromatography-mass spectrometry (GC-MS) (reviewed in Bhalerao and Bennett 2003). Direct measurement of IAA gradients include a 30-fold concentration differential across developing xylem in the wood-forming section of trees (Uggla et al. 1996), along a 15 mm tobacco leaf from petiole (high) to tip (low) (Ljung et al. 2001a), a broad range of concentrations in 3-week old soil grown plants, with siliques having the highest and mature leaves the lowest (Muller et al. 2002) and Arabidopsis roots from the root-shoot junction (high) to the root tip (low) (Bhalerao et al. 2002). Here, the source of the IAA in the gradient is proposed to be shoot-derived. Interestingly, further dissection of the root tip (1 – 2 mm sections) reveals a much shallower gradient of high to low from the root tip upwards, implying local synthesis, which is supported by measurement of auxin synthesis rates in stable isotope labeling experiments (Ljung et al. 2001a; Ljung et al. 2005; Ikeda et al. 2009). Concomitant with these sorts of measurements, which are painstaking, IAA gradients have been inferred from the observation of a polar auxin transport (PAT) system in plants and subsequently from correlations between the distribution of auxin transport components (reviewed in Vieten et al. 2007) and the gene expression output from either colorimetric or fluorescent reporter constructs driven by the synthetic auxin-responsive promoter DR5 (Ulmasov et al. 1997), which responds to the TIR1-mediated signal transduction pathway (reviewed in Lau et al. 2008; see also an animation in Laskowski 2006). The readout from the

DR5-reporter system has frequently been interpreted to be an indicator of IAA location or amount; however, as pointed out in at least two recent reviews (Ljung et al. 2004; Chandler 2009), this is a risky extrapolation, considering the complexity of auxin signal transduction. In only a very few cases (Nacry et al. 2005; Ruzicka et al. 2007; Swarup et al. 2007; Ikeda et al. 2009), notably all in Arabidopsis roots, the DR5-reporter output has been directly correlated with measurement of IAA in the same tissues. The output from reporter constructs generally correlated with the measured levels, but with much less precision. In the Arabidopsis IAA-overproducing mutant sur2, the only correlation between DR5-GUS staining and IAA levels was in the base of the hypocotyl, even though more dramatic increases in IAA were measured in the apical portions of the plant compared to wt (Ljung et al. 2004). The most definitive example of the inadequacy of DR5reporters as measures of IAA levels or location comes from IAA measurements on cells that expressed DR5-GFP and were separated from nonexpressing cells by FACS (Petersson et al. 2009). These measurements were compared with those of other specific cell types, isolated in the same manner. DR5-GFP was expressed only in a very few cells in the root tip, whereas IAA levels measured in the other root cell types revealed a much more nuanced and broad range of IAA levels in the root. Another important consideration for the DR5-reporter system is that it depends upon IAA/AUX and ARF interactions, and expression of several ARFs involved in auxin signaling has been shown recently to depend upon a microRNA (Mallory et al. 2005; Wang et al. 2005). In maize, the ARF3-regulating siRNA forms a gradient from the upper to lower sides of leaves, which patterns ARF3 (Chitwood et al. 2009). Interpretation of DR5-reporter readout as an indicator of auxin response, combined with direct measurement of IAA and/or localized manipulation of auxin levels in a variety of mutants, has consistently supported the hypothesis that auxin transport, biosynthesis, and metabolism work in concert to establish, maintain, or alter auxin maxima to affect changes

Cite this article as Cold Spring Harb Perspect Biol 2010;2:a001594

7

J. Normanly

in plant developmental programs. Some recent examples of this follow. Interactions between IAA Biosynthesis and Ethylene in the Formation of Auxin Gradients that Affect Lateral Root Elongation and Root Hair Positioning

IAA transport from the shoot is proposed to be sufficient to generate the gradients that control lateral root initiation and emergence (Grieneisen et al. 2007). This model is being re-examined in light of the evidence for significant IAA biosynthesis activity in the root (Ljung et al. 2001a; Ljung et al. 2005; Petersson et al. 2009). Ethylene response mutants have been particularly instructive here. The stimulation of ethylene synthesis by high concentrations of auxin is well established (reviewed in Stepanova and Alonso 2005), and recently, a reciprocal interaction has been demonstratedwith thediscovery that two root-specific Arabidopsis ethylene-insensitive mutants, wei 2 and wei 7, are alleles of genes encoding the anthranilate synthase a (ASA) subunit and the AS b (ASB) subunit, respectively (Stepanova et al. 2005), both of which are required for IAA synthesis (Fig. 2). Direct measurement of IAA levels in the root, combined with either genetic orchemical disruption of auxin response and transport (Ruzicka et al. 2007; Swarup et al. 2007), establish that ethylene stimulates IAA synthesis and transport in root tips, which creates IAA gradients in the root elongation zone, inhibiting cell elongation. The identification of the Arabidopsis TAA1 gene, which encodes the Trp aminotransferase proposed to catalyze the first step in the IPA pathway, adds detail to this model (Stepanova et al. 2008). TAA1 is expressed in the quiescent center of roots. The double mutant combination of wei8 (defective in TAA1) and tar2 (defective in a TAA-related gene) exhibit ethylene insensitivity in the root and in the ethylene-induced apical hook response in etiolated seedlings. The double mutants appear to have wild-type auxin transport and response, but lower levels of IAA, measured directly in roots and hypcotyls. Exogenous IAA restores ethylene sensitivity in

8

roots only. These findings are consistent with a role for the IPA pathway in ethylene-responsive IAA synthesis that is involved in regulating lateral root growth. Interestingly, in triple mutant combinations with wei8tar2, the sur1/rty, and sur2 alleles (Fig. 2), which themselves accumulate IAA, alleviate the root-specific phenotypes of the TAA1 and TAR2 defects, implying convergence of the IAOx and IPA pathways in the root. The TAA1 gene was also identified in two other mutant screens; one for mutants unable to elongate in simulated shade conditions (Tao et al. 2008) and another for hypocotyl length (Yamada et al. 2009). TAA1 is proposed to provide a rapid local increase in IAA in response to changes in R/FR ratios (Tao et al. 2008) and in response to temperature during hypocotyl elongation (Yamada et al. 2009). Arabidopsis mutants with defects in CTR1, a negative regulator of ethylene signaling, examined in combination with the ASA and ASB mutants, wei2 and wei7, respectively, as well as other ethylene response and auxin transport mutants, demonstrate that IAA synthesized in the root tip contributes to an auxin transportdependent auxin maxima distal to the root tip that establishes planar polarity of epidermal cells, specifically, the position of root hairs (Ikeda et al. 2009). Another twist to this story is the finding that ASA1 is responsive to methyl jasmonate (MeJA) in roots. MeJA inhibits primary root growth and induces LR formation. The jdl1/asa1 mutant is not responsive to MeJA induction of LR, but was responsive to MeJA with regard to primary root growth (Sun et al. 2009). MeJA-induced LR formation requires functional COI1, a regulator of MeJA response, whereas ethylene responses in COI1 with regard to LR formation were normal, implying that the response of ASA1 to ethylene and MeJA may be independent. Sugars act as signaling molecules (reviewed in Gibson 2005) and intersect with ethylene regulation of lateral root production. An Arabidopsis T-DNA mutant insensitive to the nonmetabolizable sucrose analogue turanose, is defective in the WUSCHEL-related homeobox gene WOX5 and has root phenotypes indicative of disrupted auxin homeostasis (Gonzali et al.

Cite this article as Cold Spring Harb Perspect Biol 2010;2:a001594

Auxin Biosynthesis and Metabolism

2005). WOX5 is expressed in the quiescent center and is auxin inducible. The WOX5 mutant up-regulates ethylene production, SUR2 expression, and IAA conjugation activity, suggesting that WOX5 normally functions as a negative regulator of IAA inactivation and promotes IAA biosynthesis through the CYP79B2 CYP79B3 IAOx pathway, thereby contributing to the maintenance of the auxin maximum required for lateral root formation.

Localized IAA Synthesis is Required for Embryogenesis and Vegetative and Flower Development

Arabidopsis mutants with defects in TAA1/TAR2 or a combination of several of the YUCCA genes result in poorly developed gynoecium, defective embryos, and infertility (reviewed in Cheng and Zhao 2007; Lau et al. 2008). The overlap in both the expression of the TAA1 and YUCCA gene family members and in embryo defects in corresponding mutants implies that these two IAA biosynthetic pathways, both incompletely defined at this point, may converge. A map of auxin response and YUCCA1 and YUCCA2 expression throughout seven defined developmental stages of the Arabidopsis female gametophyte (embryo sac) predicts dynamic and asymmetric distribution of auxin gradients for the duration of embryo sac development (Pagnussat et al. 2009). Down-regulation of auxin response during these seven stages within the embryo sac only, was achieved with a synthetic microRNA that targeted a subset of ARFs. The results included changes in cell identity at the cellularization stages and mis-expression of cell-specific marker genes. Nuclear positioning in the early stages of embryo sac development was not affected. Similar results were obtained by overexpressing YUCCA1 throughout the embryo sac over the seven developmental stages. PIN localization was observed only during early stages, implying the formation of gradients early in the process. These results are consistent with the hypothesis that localized IAA synthesis contributes to the gradients that determine cell patterns in the embryo sac.

Arabidopsis STY1 has a zinc-binding RINGfinger-like domain belonging to the SHI gene family (Sohlberg et al. 2006). STY1 activates transcription of YUCCA4. The gynoecia of sty1-1 and sty1-1 sty2-1 mutants exhibit abnormal style morphology and vascular patterning, are hypersensitive to chemical inhibition of PAT, and exacerbate the gynoecia phenotypes of PAT/signaling mutants pin1-5, pid-8, and ett-1. These phenotypes are consistent with a role for STY1 and STY2 in maintaining auxin gradients through regulating localized IAA biosynthesis by way of the YUCCA pathway. Lastly, a maize mutant, defective in a putative YUCCA ortholog is unable to properly initiate axillary meristems and lateral organs during vegetative and inflorescence development (Gallavotti et al. 2008). Contrary to these observations, localized manipulation of IAA biosynthesis and conjugation in Arabidopsis epidermal cells with the A. tumefaciens IAA biosynthesis gene, iaaM, and the P. syringae IAA-Lys synthase gene, iaaL, respectively, had no effect on embryo pattern formation (Weijers et al. 2005). The conclusion was that localized IAA biosynthesis and metabolism did not contribute to gradients in the embryo. Instead, a functional PAT system was proposed to buffer against changes in IAA levels and to maintain IAA gradients in the embryo. Direct measurement of IAA levels in these lines revealed significant changes in whole leaves and hypocotyls, and the substrate for the ectopic IAA synthesis was not rate limiting in the embryo. However, the IAA gradient in the embryowas extrapolated from DR5-reporter activity, which is not a direct measurement of IAA. This leaves open the possibility that IAA gradients were unchanged in the embryos because of rapid conjugation by endogenous enzymes and thus inactivation of IAA. IAA-METABOLISM: ARE IAA-PROTEINS A MEANS OF IAA INACTIVATION OR DOES THE IAA TAG THE PROTEIN FOR A SPECIFIC FATE?

Most of what we know about IAA metabolism in plants is from work done in Arabidopsis and

Cite this article as Cold Spring Harb Perspect Biol 2010;2:a001594

9

J. Normanly

maize (Fig. 3) (reviewed in Normanly 1997; Slovin et al. 1999; Cohen and Gray 2006). The most abundant IAA metabolites are IAA conjugated to a variety of small molecules, peptides, or proteins (Fig. 3) (reviewed in Seidel et al. 2006). IAA-conjugates are presumed to serve as biologically inactive, storage forms of IAA, or in some cases (IAA-Asp in Arabidopsis), as the first step in a degradation pathway (Ostin et al. 1998). IAA conjugates are generally classified by the type of conjugate linkage; esterlinked for IAA-glucose, IAA-myoinositol, and large molecular weight IAA-glycans, and amidelinked for IAA-amino acids, IAA-peptides, and IAA-proteins. The conjugate moieties vary somewhat by species (see Seidel et al. 2006). The predominant IAA conjugates in Arabidopsis are predicted to be amide-linked (Ljung et al. 2002) based upon the amount of free IAA that is released when amide linkages are broken by treatment of plant extracts with 7N NaOH at 100 8C versus treatment with 2N NaOH at room temperature, which releases ester-linkages (see, for example, Chen et al. 1988). However, in IG-producing species such as Arabidopsis, the presence of high levels of IAN can complicate this analysis. IAN is nonenzymatically converted to IAA in the presence of 7N NaOH, thereby overestimating the amount of amide-linked IAA (Ilic et al. 1996). When measured directly, levels of IAA-amino acids in Arabidopsis are quite low, within the same order of magnitude as free IAA (Tam et al. 2000; Kowalczyk and Sandberg 2001; Rampey et al. 2004), and it is presumed that most of the amide conjugates are peptides and small proteins (Ljung et al. 2002), which are not straight forward to quantify (reviewed in Seidel et al. 2006). Other metabolites of IAA include indole-3-butyric acid (IBA), which is typically present in levels lower than free IAA (Ludwig-Muller and Epstein 1994; Jones et al. 2005) and possibly methyl-IAA, which is below the limits of detection in Arabidopsis (Qin et al. 2005), so possibly turned over very rapidly. Thus, inclusion of quantitative data about IAA metabolite levels in studies of IAA homeostasis is fairly infrequent.

10

The role of IAA-peptides and IAA-proteins is still under investigation. IAA-protein IAP1 was originally isolated from bean (reviewed in Seidel et al. 2006), and antibodies to bean IAP1 have been used to isolate ortholog genes from Arabidopsis. One of the Arabidopis IAP proteins is a seed storage protein and another is present in the embryo. Rather than a storage form of IAA, it is tempting to speculate that IAA may be a “tag” for these proteins, directing them to specific locations in the cell or interactions with other proteins, similarly to other small molecule protein modifiers. Further characterization of the Arabidopsis IAA-proteins should provide new hypotheses regarding the mechanisms of IAA action. The genes for amidohydrolases specific to IAA-amino acids have been identified (Bartel and Fink 1995; Davies et al. 1999; LeClere et al. 2002) and collectively their expression patterns encompass areas of auxin biosynthesis activity (Rampey et al. 2004), although direct correlations between the expression of these genes and that of CYP79B2, YUCCA, or TAA1, for example, have not been described and should be informative about the interplay between IAA biosynthesis and conjugation in the maintenance of IAA gradients. For IAA-conjugate formation, the GH3 gene family has been identified, some members of which are amidosynthases and can form amide conjugates of IAA with amino acids (Staswick et al. 2002; Staswick et al. 2005). GH3.2 through GH3.8 and GH3.17 are active with several amino acids in vitro. So far, only GH3.6 and GH3.8 have been overexpressed in vivo (Staswick et al. 2005; Ding et al. 2008). IAA-Asp is the predominant conjugate in both cases, but profiling of all possible IAA-amino acid conjugates was not done. The Arabidopsis IAMT1 gene encodes a methyltransferase that methylates IAA in vitro (Qin et al. 2005). A mutant allele that overexpresses this gene exhibited hyponastic leaf phenotypes, suggesting that methylation of IAA plays a role in regulating plant development and auxin homeostasis. Members of an Arabidopsis family of methyl ester transferases are able to hydrolyze MeIAA in vitro (Yang et al. 2008).

Cite this article as Cold Spring Harb Perspect Biol 2010;2:a001594

Auxin Biosynthesis and Metabolism COOH

O N H O

R

R

Amide-linked IBA R = Asp, Ala, Phe, Trp

COOH

GH3.2–GH3.6 O

O

IBA

amide-linked IAA R = peptide, protein

O

ester-linked IAA R = myoinositol, glycans

N H

O

R

N H

OH

N H

R COOH

O N H

O

Zm iaglu At UGT84B1

N H

ester-linked IAA R = Glc IAA-Glc hydrolase O

O

N H

GIc

GH3.5,GH3.6 (Asp) GH3.17 (Glu) GH3.2–GH3.4

O N H

Ox-IAA-R R = Asp, Glu, Val, Phe

OH

IAA

O

R

O

IAR3 (Ala) ILR1 (Leu, Phe)

COOH

Ox-IAA-Glc IAMT1

O

AtMET

R

N H

amide-linked IAA R = Glu, Ala, Leu, Phe, Val, Asp,Trp

OH

O N H

O O

R

OxIAA N H

O

ester-linked IAA R = CH3

COOH

O OH

R

N H

6-OH-IAA-R R = Val, Phe

Figure 3. IAA metabolism in Arabidopsis. Updated from Woodward and Bartel 2005. Genes identified since the Woodward and Bartel review include GS3.2-GS3.6 and GS3.17 (Staswick et al. 2002; Staswick et al. 2005). Recently identified IAA-metabolites include IAA-Leu, IAA-Ala (Kowalczyk and Sandberg 2001; Rampey et al. 2004), IAA-Phe, IAA-Val (Kai et al. 2007a), IAA-Trp (Staswick 2009), oxIAA-Asp, oxIAA-hexose (Ostin et al. 1998), oxIAA-Glu, -Val, -Phe and –glucose (Kai et al. 2007a), 6-OH-IAA-Val and -Phe (Kai et al. 2007a), and methyl-IAA (Qin et al. 2005). IBA is a substrate for some of the GH3 genes, as indicated by in vitro assay (Staswick et al. 2005). See Cohen and Bandurski 1982, Normanly 1997, and Seidel et al. 2006 for IAA metabolites identified in other species. Dashed arrows indicate that the enzyme activity has been observed in vitro only, or neither gene nor enzyme activity have been identified.

Cite this article as Cold Spring Harb Perspect Biol 2010;2:a001594

11

J. Normanly

IAA CONJUGATION LINKS IAA HOMEOSTASIS TO STRESS AND PATHOGEN RESPONSES

Members of the GH3 family conjugate amino acids to IAA, SA, and JA (Staswick et al. 2002), which in the case of IAA and presumably SA, inactivates the signaling molecule (Staswick et al. 2005) and in the case of JA, activates it (Staswick and Tiryaki 2004). Some GH3 family members play dual roles in conjugating signaling molecules, providing a link between signaling pathways. For example, GH3.5 is able to conjugate both SA and IAA, and this impacts resistance or susceptibility to the bacterial pathogen P. syringae (Zhang et al. 2007). Overexpression of GH3.5 resulted in increases in IAA-Asp and enhanced resistance to biotic and abiotic stress, whereas the converse was observed in a GH3.5 T-DNA knockout (Park et al. 2007). Overexpression of GH3.8 in rice gives rise to enhanced disease resistance and growth defects related to decreases in free IAA levels (Ding et al. 2008).

IAA-TRP IS AN IAA ANTAGONIST

Several of the GH3 IAA amidosynthases show activity in vitro with Trp as a substrate, and IAA-Trp is a native compound in Arabidopsis (Staswick 2009). Unlike other IAA conjugates, IAA-Trp is not an inactive form of IAA; rather, it is an IAA antagonist, rendering resistance to the root growth inhibitory effects of IAA, IBA, and 2,4-dichlorophenoxyacetic acid (2,4-D). The mechanism by which IAA-Trp counteracts IAA is not clear, but IAA-Trp is probably not a source of free IAA based on the substrate specificities of known Arabidopsis amidohydrolases (LeClere et al. 2002), and its antagonistic effects require functional TIR1, but not auxin transport. With the identification of a Trpprotein intermediate in Trp-dependent IAA synthesis (Hendrickson Culler and Cohen 2007), it would be interesting to determine the effects, if any of IAA-Trp on this reaction. IAA is proposed to be enzymatically converted to IBA and vice versa (reviewed in Woodward and Bartel 2005). IBA is classified

12

as an auxin and it remains to be determined if IBA itself is an auxin, or whether IBA effects are through its metabolism to IAA. IBA quantification in Arabidopsis reveals that it is generally present in lower amounts than free IAA (Ludwig Mu¨ller et al. 1993), in one case as low as 2% of free IAA levels (Jones et al. 2005). IBA synthetase activity, measured in vitro in a variety of Arabidopsis ecotypes, revealed varying activities (Ludwig Mu¨ller 2007). In the original report on the presence of IBA in maize, some cultivars had detectable levels of IBA and others did not (Epstein et al. 1989). A similar situation may exist for Arabidopsis, as inconsistent levels have been observed by groups who have critically examined IBA levels analytically (Ludwig Mu¨ller et al. 1993; Jones et al. 2005; K. Ljung, unpubl.; J. Normanly, unpubl.; J. Cohen, unpubl.). We do not as yet understand what developmental, tissue, environmental, or genetic differences might account for this variation in amounts found in the various investigations. One IBA resistant mutant, ibr5, identified an auxin response pathway that acts independently of the TIR1-mediated proteosome degradation of repressor proteins (Strader et al. 2008a), and a suppressor screen of ibr5 has yielded an ATP-binding cassette transporter implicated in cellular efflux of the synthetic auxin 2,4-dichlorophenoxyacetic acid (Strader et al. 2008b). This screen identified a number of loci that are likely to be new components of auxin signaling, transport, and metabolism. THE IDENTIFICATION OF NEW IAA METABOLITES IN ARABIDOPSIS

The catabolism of IAA is proposed to occur either by enzymatic oxidation of the indole nucleus of IAA (to form ox-IAA or ox-IAA-conjugates) (Fig. 3), or through oxidative decarboxylation of the IAA side chain. Most of this work has been done in vitro (see, e.g., Nonhebel et al. 1985; Beffa et al. 1990), and none of the genes for IAA catabolism have been definitively identified. Recently, three new oxidative metabolites of IAA have been identified in Arabidopsis using a sensitive MS screen of HPLC fractions from extracts

Cite this article as Cold Spring Harb Perspect Biol 2010;2:a001594

Auxin Biosynthesis and Metabolism

of 2-week-old plants (Fig. 3) (Kai et al. 2007a). Two new IAA conjugates, IAA-Phe and IAA-Val, were identified indirectly, which are good substrates for the GH3 family of amidosynthases. OxIAA-Glc is proposed to be a primary oxidative metabolite of IAA in Arabidopsis based on quantification with deuterium-labeled internal standard (Kai et al. 2007a; Kai et al. 2007b). This recent observation is consistent with previous work in which OxIAA-hexose was identified as a major metabolite of IAA in Arabidopsis (Ostin et al. 1998). In the case of OxIAA-Glc measurement, the deuterium, label was incorporated into the side chain (20 ) (Fig. 1) of OxIAA-Glc, which is subject to exchange with hydrogen, not just in alkaline conditions, but likely in plant extracts as well. It has been known for some time (Caruso and Zeisler 1983) that using side chain deuteriumlabeled IAA as an internal standard overestimates endogenous IAA concentrations by as much as 10-fold. Similar results have been observed with side chain deuterium-labeled IAN (threefold overestimation) (J. Normanly, unpubl.). CONCLUDING REMARKS

The identification of new genes involved in IAA biosynthesis and metabolism has continued at a brisk pace, and the use of mutants to characterize their function has been invaluable in revealing the complexity of interactions between IAA and other signaling molecules. The ability to quantify IAA at high resolution has always been a bottleneck for auxin biology, but significant progress has been made here as well, and the next challenge will be to more broadly profile IAA precursors and metabolites along with those of other signaling molecules. The combination of these sort of data with the molecular genetic approaches that have brought us to this point will only aid us in finally answering the long-standing question of how IAA is made in plants. ACKNOWLEDGMENTS

The author is supported by National Science Foundation (NSF) MCB 0517420 and 0725192.

REFERENCES Bak S, Feyereisen R. 2001. The involvement of two P450 enzymes, CYP83B1 and CYP83A1, in auxin homeostasis and glucosinolate biosynthesis. Plant Physiol 127: 108 –118. Bak S, Tax FE, Feldmann KA, Galbraith DW, Feyereisen R. 2001. CYP83B1, a cytochrome P450 at the metabolic branch point in auxin and indole glucosinolate biosynthesis in Arabidopsis. Plant Cell 13: 101 –111. Bartel B, Fink G. 1995. ILR1, an amidohydrolase that releases active indole-3-acetic acid from conjugates. Science 268: 1745– 1748. Barth C, Jander G. 2006. Arabidopsis myrosinases TGG1 and TGG2 have redundant function in glucosinolate breakdown and insect defense. The Plant Journal 46: 549 –562. Bednarek P, Pis´lewska-Bednarek M, Svatosˇ A, Schneider B, Doubsky´ J, Mansurova M, Humphry M, Consonni C, Panstruga R, Sanchez-Vallet A, et al. 2009. A glucosinolate metabolism pathway in living plant cells mediates broad-spectrum antifungal defense. Science 323: 101 –106. Beffa R, Martin HV, Pilet P-E. 1990. In vitro oxidation of indoleacetic acid by soluble auxin-oxidases and peroxidases from maize roots. Plant Physiol 94: 485– 491. Bender J, Celenza JL. 2009. Indole glucosinolates at the crossroads of tryptophan metabolism. Phytochemistry 8: 25–37. Bhalerao R, Bennett MJ. 2003. The case for morphogens in plants. Nature Cell Biology 5: 939 –943. Bhalerao RP, Eklof J, Ljung K, Marchant A, Bennett M, Sandberg G. 2002. Shoot-derived auxin is essential for early lateral root emergence in Arabidopsis seedlings. Plant J 29: 325–32. Bones A, Rossiter J. 2006. The enzymic and chemically induced decomposition of glucosinolates. Phytochemistry 67: 1053– 1067. Bottcher C, Westphal L, Schmotz C, Prade E, Scheel D, Glawischnig E. 2009. The multifunctional enzyme CYP71B15 (PHYTOALEXIN DEFICIENT3) converts aysteine-indole-3-acetonitrile to camalexin in the indole-3-acetonitrile metabolic network of Arabidopsis thaliana. Plant Cell: doi: 10.1105/tpc.109.066670. Burow M, Zhang Z-Y, Ober JA, Lambrix V, Wittstock U, Gershenzon J, Kliebenstein DJ. 2008. ESP and ESM1 mediate indol-3-acetonitrile production from indol3-ylmethyl glucosinolate in Arabidopsis. Phytochemistry 69: 663– 671. Caruso JL, Zeisler CS. 1983. Indole-3-acetic acid in douglas fir seedlings: a reappraisal. Phytochemistry 22: 589– 590. Chandler JW. 2009. Local auxin production: a small contribution to a big field. Bioessays 31: 60–70. Chen Y, Dai X, Zhao Y. 2007. Auxin synthesized by the YUCCA flavin monooxygenases is essential for embryogenesis and leaf formation in Arabidopsis. Plant Cell 19: 2430–2439. Chen K-H, 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.

Cite this article as Cold Spring Harb Perspect Biol 2010;2:a001594

13

J. Normanly Cheng Y, Zhao Y. 2007. A role for auxin in flower development. J Integr Plant Biol 49: 99–104. Cheng Y, Dai X, Zhao Y. 2006. Auxin biosynthesis by the YUCCA flavin monooxygenases controls the formation of floral organs and vascular tissues in Arabidopsis. Genes Dev 20: 1790–1799. Chitwood DH, Nogueira FTS, Howell MD, Montgomery TA, Carrington JC, Timmermans MCP. 2009. Pattern formation via small RNA mobility. Genes Dev 23: 549 –554. Clay NK, Adio AM, Denoux C, Jander G, Ausubel FM. 2009. Glucosinolate metabolites required for an Arabidopsis innate immune response. Science 323: 95–101. Cohen JD, Bandurski RS. 1982. Chemistry and physiology of the bound auxins. Annu Rev Plant Physiol 33: 403 –430. Cohen JD, Gray WM. 2006. Auxin metabolism and signaling. In Plant Hormone Signaling. (eds. P. Hedden, S.G. Thomas), 37–66. Blackwell, Oxford. Davies RT, Goetz DH, Lasswell J, Anderson MN, Bartel B. 1999. IAR3 encodes an auxin conjugate hydrolase from Arabidopsis. Plant Cell 11: 365– 76. Ding X, Cao Y, Huang L, Zhao J, Xu C, Li X, Wang S. 2008. Activation of the indole-3-acetic acid-amido synthetase GH3–8 suppresses expansin expression and promotes salicylate- and jasmonate-independent basal immunity in rice. Plant Cell 20: 228– 240. 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 Regulation 8: 215–223. Epstein E, Cohen J, Slovin J. 2002. The biosynthetic pathway for indole-3-acetic acid changes during tomato fruit development. Plant Growth Regulation 38: 15–20. Exposito-Rodrıguez M, Borges AA, Borges-Perez A, Hernandez M, Perez JA. 2007. Cloning and biochemical characterization of ToFZY, a tomato gene encoding a flavin monooxygenase involved in a tryptophan-dependent auxin biosynthesis pathway. J Plant Growth Regul 26: 329 –340. Gallavotti A, Barazesh S, Malcomber S, Hall D, Jackson D, Schmidt RJ, McSteen P. 2008. sparse inflorescence1 encodes a monocot-specific YUCCA-like gene required for vegetative and reproductive development in maize. Proc Natl Acad Sci 105: 15196– 15201. Gibson SI. 2005. Control of plant development and gene expression by sugar signaling. Current Opin Plant Biol 8: 93– 102. Glick BR, Patten CL, Holguin G, Penrose DM. 1999. Auxin production. In Biochemical and Genetic Mechanisms Used by Plant Growth Promoting Bacteria. (ed. B.R. Glick et al.), 86– 133. Imperial College Press, London. Gonzali S, Novi G, Loreti E, Paolicchi F, Poggi A, Alpi A, Pierdomenico P. 2005. A turanose-insensitive mutant suggests a role for WOX5 in auxin homeostasis in Arabidopsis thaliana. The Plant Journal 44: 633–645. ¨ stin A, Sandberg G, Romano CP, Estelle M. Gray WM, O 1998. High temperature promotes auxin-mediated hypocotyl elongation in Arabidopsis. Proc Natl Acad Sci 95: 7197– 7202. Grieneisen VA, Xu J, Maree AF, Hogeweg P, Scheres B. 2007. Auxin transport is sufficient to generate a maximum and gradient guiding root growth. Nature 449: 1008– 1013.

14

Halkier BA, Gershenzon J. 2006. Biology and biochemistry of glucosinolates. Annu Rev Plant Biol 57: 303– 333. Hendrickson Culler A, CohenJD. 2007. Novel Pathway for IAA Biosynthesis in Maize Endosperm. ASPB 2007. Canada. http://2007.botanyconference.org/engine/search/index. php?func¼detail&aid¼692. Hull AK, Vij R, Celenza JL. 2000. Arabidopsis cytochrome P450s that catalyze the first step of tryptophandependent indole-3-acetic acid biosynthesis. Proc Natl Acad Sci 97: 2379–84. Ilic N, Normanly J, Cohen JD. 1996. Quantification of free plus conjugated indoleacetic acid in Arabidopsis requires correction for the nonenzymatic conversion of indolic nitriles. Plant Physiol 111: 781– 788. Ikeda Y, Men S, Fischer U, Stepanova AN, Alonso JM, Ljung K, Grebe M. 2009. Local auxin biosynthesis modulates gradient-directed planar polarity in Arabidopsis. Nature Cell Biology 11: 731 –738. Jones SE, DeMeo JS, Davies NW, Noonan SE, Ross JJ. 2005. Stems of the Arabidopsis pin1– 1 mutant are not deficient in free indole-3-acetic acid. Planta 222: 530 –534. Kai K, Horita J, Wakasa K, Miyagawa H. 2007a. Three oxidative metabolite of indole-3-acetic acid from Arabidopsis thaliana. Phytochemistry 68: 1651– 1663. Kai K, Nakamura S, Wakasa K, Miyagawa H. 2007b. Facile preparation of deuterium-labeled standards of indole3-acetic acid (IAA) and its metabolites to quantitatively analyze the disposition of exogenous IAA in Arabidopsis thaliana. Biosci Biotechnol Biochem 71: 1946–1954. Kim JI, Sharkhuu A, Jin JB, Li P, Jeong JC, Baek D, Lee SY, Blakeslee JJ, Murphy AS, Bohnert HJ, et al. 2007. yucca6, a dominant mutation in Arabidopsis, affects auxin accumulation and auxin-related phenotypes. Plant Phsysiol 145: 722 –735. Kjaer A. 1974. The natural distribution of glucosinolates: a uniform group of sulfur-containing glucosides. In Chemistry in Botanical Classification. (eds. G. Bendz, J. Santesson), 229– 234. Academic Press, New York. Koga J. 1995. Structure and function of indolepyruvate decarboxylase, a key enzyme in indole-3-acetic acid biosynthesis. Biochimica et Biophysica Acta 1249: 1 –13. Koga J, Adachi T, Hidaka H. 1991. Molecular cloning of the gene for indolepyruvate decarboxylase from Enterobacter cloacae. Molecular and General Genetics 226: 10– 16. Koga J, Adachi T, Hidaka H. 1992. Purification and characterization of indolepyruvate decarboxylase. Journal of Biological Chemistry 267: 15823–15828. Kowalczyk M, Sandberg G. 2001. Quantitative analysis of indole-3-acetic acid metabolites in Arabidopsis. Plant Physiol 127: 1845– 53. Kriechbaumer V, Park WJ, Gierl A, Glawischnig E. 2006. Auxin biosynthesis in maize. Plant Biology 8: 334– 339. Kriechbaumer V, Park WJ, Piotrowski M, Meeley RB, Gierl A, Glawischnig E. 2007. Maize nitrilases have a dual role in auxin homeostasis and -cyanoalanine hydrolysis. J Exp Bot 58: 4225– 4233. Lambrix V, Reichelt M, Mitchell-Olds T, Kliebenstein DJ, Gershenzon J. 2001. The Arabidopsis epithiospecifier protein promotes the hydrolysis of glucosinolates to

Cite this article as Cold Spring Harb Perspect Biol 2010;2:a001594

Auxin Biosynthesis and Metabolism nitriles and influences Trichoplusia ni herbivory. Plant Cell 13: 2793–2807. Laskowski M. 2006. Model of the TIR1 pathway for auxinmediated gene expression. Sci STKE 2006: tr1. doi: 10.1126/stke.3222006tr1, http://stke.sciencemag.org/cgi/ content/abstract/sigtrans;2006/322/tr1. Last RL, Fink GR. 1988. Tryptophan-requiring mutants of the plant Arabidopsis thaliana. Science 240: 305 –310. Lau S, Jurgens G, De Smet I. 2008. The evolving complexity of the auxin pathway. Plant Cell 20: 1738–1746. LeClere S, Tellez R, Rampey RA, Matsuda SP, Bartel B. 2002. Characterization of a family of IAA-amino acid conjugate hydrolases from Arabidopsis. J Biol Chem 277: 20446– 20452. Ljung K, Bhalerao RP, Sandberg G. 2001a. Sites and homeostatic control of auxin biosynthesis in Arabidopsis during vegetative growth. Plant J 28: 465– 74. Ljung K, Sandberg G, Moritz T. 2004. Methods of plant hormone analysis. In Plant Hormones: Biosynthesis, Signal Transduction, Action! (ed. P.J. Davies), 671 –694. Kluwer Academic Publishers, Dordrecht. Ljung K, Hull A, Celenza J, Yamada M, Estelle M, Normanly J, Sandberg G. 2005. Sites and regulation of auxin biosynthesis in Arabidopsis roots. Plant Cell 17: 1090– 1140. Ljung K, Hull AK, Kowalczyk M, Marchant A, Celenza J, Cohen JD, Sandberg G. 2002. Biosynthesis, conjugation, catabolism and homeostasis of indole-3-acetic acid in Arabidopsis thaliana. Plant Mol Biol 49: 249– 72. Ljung K, Ostin A, Lioussanne L, Sandberg G. 2001b. Developmental regulation of indole-3-acetic acid turnover in Scots pine seedlings. Plant Physiol 125: 464– 75. Ludwig Mu¨ller J. 2007. Indole-3-butyric acid synthesis in ecotypes and mutants of Arabidopsis thaliana under different growth conditions. Journal of Plant Physiology 164: 47–59. Ludwig-Muller J, Epstein E. 1994. Indole-3-butyric acid in Arabidopsis thaliana III: in vivo biosynthesis. Plant Growth Regulation 14: 7 –14. Ludwig Mu¨ller J, Sass S, Sutter EG, Wodner M, Epstein E. 1993. Indole-3-buytyric acid in Arabidopsis thaliana 1. Identification and quantification. Plant Growth Regulation 179: 179 –187. Magnus V, Bandurski RS, Schulze A. 1980. Synthesis of 4,5,6,7 and 2,4,5,6,7 deuterium-labeled indole-3-acetic acid for use in mass spectrometric assays. Plant Physiology 66: 775– 781. Mallory AC, Bartel DP, Bartel B. 2005. MicroRNA-directed regulation of Arabidopsis AUXIN RESPONSE FACTOR17 is essential for proper development and modulates expression of early auxin response genes. Plant Cell 17: 1360– 1375. Michalczuk L, Ribnicky DM, Cooke TJ, Cohen JD. 1992. Regulation of indole-3-acetic acid biosynthetic pathways in carrot cell cultures. Plant Physiol 100: 1346– 1353. Mikkelsen MD, Fuller VL, Hansen BG, Nafisi M, Olsen CE, Nielsen HL, Halkier BA. 2009. Controlled indole-3acetaldoxime production through ethanol-induced expression of CYP79B2. Planta, DOI 10.1007/s00425– 009-0907 –5. Muller A, Duchting P, Weiler E. 2002. A mulitplex GCMS/MS technique for the sensitive and quantitative

single-run analysis of acidic phytohormones and related compounds and its application to Arabidopsis thaliana. Planta 216: 44–56. Nacry P, Canivenc G, Muller B, Azmi A, Van Onckelen H, Rossignol M, Doumas P. 2005. A role for auxin redistribution in the responses of the root system architecture to phosphate starvation in Arabidopsis. Plant Physiol 138: 2061– 2074. Nafisi M, Goregaoker S, Botanga CJ, Glawischnig E, Olsen CE, Halkier BA, Glazebrook J. 2007. Arabidopsis cytochrome P450 monooxygenase 71A13 catalyzes the conversion of indole-3-Acetaldoxime in camalexin synthesis. Plant Cell 19: 2039– 2052. Nonhebel H, Kruse L, Bandurski R. 1985. Indole-3-acetic acid catabolism in Zea mays seedlings. Journal of Biological Chemistry 260: 12685–12689. Normanly J. 1997. Auxin metabolism. Physiol Plant 100: 431 –442. Normanly J, Slovin JP, Cohen JD. 2004. Auxin Metabolism. In Plant Hormones: Biosynthesis, Signal Transduction, Action! (ed. P.J. Davies), 36–62. Kluwer Academic Publishers, Dordrecht. Ostin A, Kowalyczk M, Bhalerao RP, Sandberg G. 1998. Metabolism of indole-3-acetic acid in Arabidopsis. Plant Physiol 118: 285 –296. Ouyang J, Shao X, Li J. 2000. Indole-3-glycerol phosphate, a branchpoint of indole-3-acetic acid biosynthesis from the tryptophan biosynthetic pathway in Arabidopsis thaliana. Plant J 24: 327 –333. Pagnussat GC, Alandete-Saez M, Bowman JL, Sundaresan V. 2009. Auxin-dependent patterning and gamete specification in the Arabidopsis female gametophyte. Science: doi:10.1126/science.1167324. Park J-E, Park J-Y, Kim Y-S, Staswick PE, Jeon J, Yun J, Kim S-Y, Kim J, Lee Y-H, Park C-M. 2007. GH3-mediated auxin homeostasis links growth regulation with stress adaptation response in Arabidopsis. J Biol Chem 282: 10036– 10046. Petersson SV, Johansson AI, Kowalczyk M, Makoveychuk A, Wang JY, Moritz T, Grebe M, Benfey PN, Sandberg G, Ljung K. 2009. An auxin gradient and maximum in the Arabidopsis root apex shown by high-resolution cellspecific analysis of IAA distribution and synthesis. Plant Cell 21: 1659–1668. Qin G, Gu H, Zhao Y, Ma Z, Shi G, Yang Y, Pichersky E, Chen H, Liu M, Chen Z, Qu L-J. 2005. An indole-3-acetic acid carboxyl methyltransferase regulates Arabidopsis leaf development. Plant Cell 17: 2693–2704. Quint M, Barkawi LS, Fan K-T, Cohen JD, Gray WM. 2009. Arabidopsis IAR4 modulates auxin response by regulating auxin homeostasis. Plant Physiol 150: 748 –758. Quittenden LJ, Davies NW, Smith JA, Molesworth PP, Tivendale ND, Ross JJ. 2009. Auxin biosynthesis in Pisum sativum: Characterisation of the tryptamine pathway. Plant Physiol: doi:10.1104/pp.109.141507. Rampey RA, LeClere S, Kowalcyk S, Ljung K, Sandberg G, Bartel B. 2004. A family of auxin-conjugate hydrolases that contributes to free indole-3-acetic acid levels during Arabidopsis germination. Plant Physiol 135: 978 –988.

Cite this article as Cold Spring Harb Perspect Biol 2010;2:a001594

15

J. Normanly Rapparini F, Tam YY, Cohen JD, Slovin JP. 2002. Indole3-acetic acid metabolism in Lemna gibba undergoes dynamic changes in response to growth temperature. Plant Physiol 128: 1410–1416. Ribnicky DM, Cohen JD, Hu WS, Cooke TJ. 2002. An auxin surge following fertilization in carrots: a mechanism for regulating plant totipotency. Planta 214: 505– 509. Ruzicka K, Ljung K, Vanneste S, Podhorska R, Beeckman T, Friml J, Benkova E. 2007. Ethylene regulates root growth through effects on auxin biosynthesis and transportdependent auxin distribution. Plant Cell 19: 2197– 2212. Seidel C, Walz A, Park S, Cohen JD, Ludwig Mu¨ller J. 2006. Indole-3-acetic acid protein conjugates: novel players in auxin homeostasis. Plant Biology 8: 340– 345. Sekimoto H, Seo M, Kawakami N, Komano T, Desloire S, Liotenberg S, Marion-Poll A, Caboche M, Kamiya Y, Koshiba T. 1998. Molecular cloning and characterization of aldehyde oxidases in Arabidopsis thaliana. Plant Cell Physiology 39: 433– 442. Seo M, Akaba S, Oritani T, Delarue M, Bellini C, Caboche M, Koshiba T. 1998. Higher activity of an aldehyde oxidase in the auxin-overproducing superroot1 mutant of Arabidopsis thaliana. Plant Physiol 116: 687– 693. Slovin J, Bandurski R, Cohen J. 1999. Auxin. In Biochemistry and Molecular Biology of Plant Hormones. (ed. P. Hooykaas et al.), 115–140. Elsevier Science, Oxford. Sohlberg JJ, Myrenas M, Kuusk S, Lagercrantz U, Kowalczyk M, Sandberg G, Sundberg E. 2006. STY1 regulates auxin homeostasis and affects apical-basal patterning of the Arabidopsis gynoecium. The Plant Journal 47: 112 –123. Staswick PE. 2009. Trp conjugates inhibit auxin. Plant Physiol: DOI:10.1104/pp.109.138529. Staswick PE, Tiryaki I. 2004. The oxylipin signal jasmonic acid is activated by an enzyme that conjugates it to isoleucine in Arabidopsis. Plant Cell 16: 2117–2127. Staswick PE, Tiryaki I, Rowe ML. 2002. Jasmonate response locus JAR1 and several related Arabidopsis genes encode enzymes of the firefly luciferase superfamily that show activity on jasmonic, salicylic, and indole-3-acetic acids in an assay for adenylation. Plant Cell 14: 1405– 1415. Staswick PE, Serban B, Rowe M, Tiryaki I, Maldonado MT, Maldonado MC, Suza W. 2005. Characterization of an Arabidopsis enzyme family that conjugates amino acids to indole-3-acetic acid. Plant Cell 17: 616–627. Stepanova AN, Alonso J. 2005. Ethylene signaling and response pathway: A unique signaling cascade with a multitude of inputs and outputs. Physiol Plant 123: 195–206. Stepanova AN, Hoyt JM, Hamilton AA, Alonso J. 2005. A link between ethylene and auxin uncovered by the characterization of two root-specific ethylene-insensitive mutants in Arabidopsis. Plant Cell 17: 2230–2242. Stepanova AN, Robertson-Hoyt J, Yun J, Benavente LM, Xie DY, Dolezal K, Schlereth A, Jurgens G, Alonso JM. 2008. TAA1-mediated auxin biosynthesis is essential for hormone crosstalk and plant development. Cell 133: 177 –191. Strader L, Monroe-Augustus M, Bartel B. 2008a. The IBR5 phosphatase promotes Arabidopsis auxin responses through a novel mechanism distinct from TIR1-mediated

16

repressor degradation. BMC Plant Biology 8: doi:10.1186/ 1471–2229-8–41. Strader LC, Monroe-Augustus M, Rogers KC, Lin GL, Bartel B. 2008b. Arabidopsis iba response5 suppressors separate responses to various hormones. Genetics 180: 2019– 2031. Sugawara S, Hishiyama S, Jikumaru Y, Hanada A, Nishimura T, Koshiba T, Zhao Y, Kamiya Y, Kasahara H. 2009. Biochemical analyses of indole-3-acetaldoximedependent auxin biosynthesis in Arabidopsis. Proceedings of the National Academy of Sciences 106: 5430–5435. Sun J, Xu Y, Ye S, Jiang H, Chen Q, Liu F, Zhou W, Chen R, Li X, Tietz O, et al. 2009. Arabidopsis ASA1 is important for jasmonate-mediated regulation of auxin biosynthesis and transport during lateral root formation. Plant Cell: doi: 10.1105/tpc.108.064303. Swarup R, Perry P, Hagenbeek D, Van Der Straeten D, Beemster GTS, Sandberg G, Bhalerao R, Ljung K, Bennett MJ. 2007. Ethylene upregulates auxin biosynthesis in Arabidopsis seedlings to enhance inhibition of root cell elongation. Plant Cell 19: 2186–2196. Sztein A, Cohen J, Cooke T. 2002. Indole-3-acetic acid biosynthesis in isolated axes from germinating bean seeds: The effect of wounding on the biosynthetic pathway. Plant Growth Regulation 36: 201 –207. Tam YY, Normanly J. 1998. Determination of indole-3pyruvic acid levels in Arabidopsis thaliana by gas chromatography-selected ion monitoring-mass spectrometry. J Chromatogr A 800: 101–108. 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– 596. Tao Y, Ferrer JL, Ljung K, Pojer F, Hong F, Long JA, Li L, Moreno JE, Bowman ME, Ivans LJ, et al. 2008. Rapid synthesis of auxin via a new tryptophan-dependent pathway is required for shade avoidance in plants. Cell 133: 164 –176. Toben˜a-Santamaria R, Bliek M, Ljung K, Sandberg F, Mol JNM, Souer E, Koes R. 2002. FLOOZY of petunia is a flavin mono-oxygenase-like protein required for the specification of leaf and flower architecture. Genes Dev 16: 753 –763. Uggla C, Moritz T, Sanberg G, Sundberg B. 1996. Auxin as a positional signal in pattern formation in plants. Proc Natl Acad Sci 93: 9282–9286. Ulmasov T, Murfett J, Hagen G, Guilfoyle T. 1997. AUX/IAA proteins repress expression of reporter genes containing natural and highly active synthetic auxin response elements. Plant Cell 9: 1963– 1971. Vieten A, Sauer M, Brewer PB, Friml J. 2007. Molecular and cellular aspects of auxin-transport-mediated development. Trends in Plant Science 12: 160 –168. Wang J-W, Wang L-J, Mao Y-B, Cai W-J, Xue H-W, Chen X-Y. 2005. Control of root cap formation by microRNA-targeted auxin response factors in Arabidopsis. Plant Cell 17: 2204–2216. Weijers D, Sauer M, Meurette O, Friml J, Ljung K, Sandberg G, Hooykaas P, Offringa R. 2005. Maintenance of embryonic auxin distribution for apical-basal patterning by PIN-FORMED-dependent auxin transport in Arabidopsis. Plant Cell 17: 2517– 2526.

Cite this article as Cold Spring Harb Perspect Biol 2010;2:a001594

Auxin Biosynthesis and Metabolism Woodward AW, Bartel B. 2005. Auxin: regulation, action, and interaction. Ann Bot 95: 707–735. Yamada M, Greenham K, Prigge MJ, Jensen PJ, Estelle M. 2009. The TRANSPORT INHIBITOR RESPONSE2 (TIR2) gene is required for auxin synthesis and diverse aspects of plant development. Plant Physiol: doi: 10.1104/pp.109.138859. Yamamoto Y, Kamiya N, Morinaka Y, Matsuoka M, Sazuka T. 2007. Auxin biosynthesis by the YUCCA genes in rice. Plant Phsysiol 143: 1362– 1371. Yang Y, Xu R, Ma CJ, Vlot AC, Klessig DF, Pichersky E. 2008. Inactive methyl indole-3-acetic acid ester can by hydrolyzed and activated by several esterases belonging to the AtMES esterase family of Arabidopsis thaliana. Plant Phsysiol 147: 1034–1045. Ye S, Cohen JD. 2009. Reevaluation of the role of the indole pyruvate decarboxylase (IPyA) pathway in auxin biosynthesis in plants. ASPB 2009. Honolulu, Hawaii, http://abstracts.aspb.org/pb2009/public/P34/P34020. html.

Zhang Z, Li Q, Li Z, Staswick PE, Wang M, Zhu Y, He Z. 2007. Dual regulation role of GH3.5 in salicylic acid and auxin signaling during Arabidopsis-Pseudomonas syringae interaction. Plant Physiol 145: 450 – 464. Zhao Y. 2008. The role of local biosynthesis of auxin and cytokinin in plant development. Current Opinion in Plant Biology 11: 16–22. Zhao N, Tootle TL, Glazebrook J. 1999. Arabidopsis PAD3, a gene required for camalexin biosynthesis encodes a putative cytochrome P450 monooxygenase. Plant Cell 11: 2419–2428. Zhao Y, Christensen SK, Fankhauser C, Cashman JR, Cohen JD, Weigel D, Chory J. 2001. A role for flavin monooxygenase-like enzymes in auxin biosynthesis. Science 291: 306 –309. Zhao Y, Hull AK, Gupta NR, Goss KA, Alonso J, Ecker JR, Normanly J, Chory J, Celenza JL. 2002. Trp-dependent auxin biosynthesis in Arabidopsis: involvement of cytochrome P450s CYP79B2 and CYP79B3. Genes Dev 16: 3100–3112.

Cite this article as Cold Spring Harb Perspect Biol 2010;2:a001594

17