* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
Download Brassinosteroid biosynthesis and signalling in Arabidopsis thaliana
Venus flytrap wikipedia , lookup
Plant physiology wikipedia , lookup
Plant morphology wikipedia , lookup
Cryptochrome wikipedia , lookup
Glossary of plant morphology wikipedia , lookup
Plant disease resistance wikipedia , lookup
Sustainable landscaping wikipedia , lookup
Brassinosteroid biosynthesis and signalling in Arabidopsis thaliana and Petunia hybrida The research presented in this thesis was performed at the Department of Genetics, Institute of Molecular and Cell Biology, Vrije Universiteit, De Boelelaan 1085, 1081 HV Amsterdam. The research was funded by the Netherlands Organisation for Scientific Research (NWO). Cover design: Nathalie Verhoef & Sylvia Klop Printing was done by Wöhrman print service, Zutphen VRIJE UNIVERSITEIT Brassinosteroid biosynthesis and signalling in Arabidopsis thaliana and Petunia hybrida ACADEMISCH PROEFSCHRIFT ter verkrijging van de graad Doctor aan de Vrije Universiteit Amsterdam, op gezag van de rector magnificus prof.dr. L.M. Bouter, in het openbaar te verdedigen ten overstaan van de promotiecommissie van de faculteit der Aard- en Levenswetenschappen op woensdag 6 april 2011 om 15.45 uur in de aula van de universiteit, De Boelelaan 1105 door Nathalie Verhoef geboren te Sliedrecht promotor: prof.dr. R.E. Koes copromotor: dr. E. Souer Voor mijn ouders Contents Chapter 1 General introduction: Brassinosteroid biosynthesis and signalling in plants 9 Chapter 2 Group II GSK3/SHAGGY-like kinases act redundantly in the brassinosteroid signalling pathway 35 Chapter 3 Towards identification of group II ASKs containing complexes in Arabidopsis 57 Chapter 4 Brassinosteroid biosynthesis and signalling in petunia 81 Summary 113 Samenvatting 117 Dankwoord 121 Chapter 1 Brassinosteroid biosynthesis and signalling in plants General introduction Nathalie Verhoef, Ronald Koes and Erik Souer General introduction: Brassinosteroid biosynthesis and signalling in plants Introduction Early in the 1970’s a new plant hormone was described as a substance present in pollen of Brassica napus that, when applied in minute concentrations, could induce growth of bean internodes (Mitchell et al., 1970). It took till 1979 before the chemical nature of this growth-promoting factor was elucidated as a steroid compound. This discovery was the birth of a new class of plant hormones, since then known as brassinosteroids (BRs). Pioneering work using feeding experiments in Catharanthus roseus and Arabidopsis thaliana lead to insight in the biosynthesis pathway (Suzuki et al., 1993; Suzuki et al., 1994). Subsequent analysis of mutants from Arabidopsis combined with detailed biochemical analysis of identified proteins led to a near complete picture of BR biosynthesis and signalling (Fujioka and Yokota, 2003; Gendron and Wang, 2007; Kim et al., 2009). In this chapter, we will give an overview of the current knowledge on BR biosynthesis and signalling in plants. Biosynthesis of brassinosteroids The first BR compound, brassinolide, was discovered in bee-collected pollen of Brassica napus L. (Grove et al., 1979). To date more than 65 free BRs and 5 BR conjugates have been isolated from plants with brassinolide as the most bioactive compound (Bajguz, 2007). BRs are widely distributed among the plant kingdom and so far they have been found in gymnosperms, monocots, dicots, a pteridophyte, a bryophyte and a chlorophyte (Bajguz and Tretyn, 2003). They are found in almost every part of the plant, but are most abundant in the reproductive organs (Bajguz and Tretyn, 2003). The sterol biosynthesis pathway can be divided into three biosynthetic domains (Fig. 1). In the first domain the mevalonate derivative squalene is converted into 24methylenelophenol. Next, the pathway diverges into two branches, one branch producing the more abundant sterols that are used to produce membrane components, while the second part produces brassinosteroids from the precursor campesterol. Detailed studies in Arabidopsis thaliana and Catharanthus roseus revealed that synthesis of catasterone from the precursor campesterol can be accomplished via two pathways, the early and late C-6 oxidation pathways (Choi et al., 1996; Nomura et al., 2001; Fujioka and Yokota, 2003). Although both pathways are found throughout the plant kingdom, it seems that for species such as tobacco and tomato the late C-6 oxidation pathway is more prevalent (Fujioka and Yokota, 2003). In Arabidopsis and tomato, these two pathways are linked to each other via the cytochrome P450 enzyme CYP85A1/A2, implying that they are not totally autonomous 11 Chapter 1 (Shimada et al., 2001). Fujioka et al. (2002) revealed upstream of the C-6 oxidation pathways an early C-22 oxidation pathway. Epoxidase SMT1 Lupeol synthase Squalene-2,3epooxide Squalene Mevalonate 24-Methylenecycloartenol Cycloartenol CYP51G1 CYP710A Stigmasterol DIM Sitosterol DWF5 DWF7 5-Dehydroavenasterol Isofucosterol Cycloeucalenol δ 7-Avenasterol Cyclopropyl isomerase Obutusifoliol 24-Ethylidenelophenol membranes CYP51G1 δ8,14-Sterol SMT2 FK 24-Methylenecholestrol DWF5 5-Dehydroepisterol DWF7 24-Methylene lophenol Episterol HYD1 4α-Methylfecosterol DIM Campesterol Campest-4-en3β-one DWF4 22α-Hydroxy-5− campesterol Campest-4en-3-one DWF4 SAX1 22α-Hydroxycampest-4-en-3-one DET2 5α-Campestan- 6-Oxocampestenol DWF4 DWF4 DET2 22α-Hydroxy-5α- DWF4 6-Deoxocathasterone campestan-3-one CYP90 C1/D1 CYP90C1/D1 6α-Hydroxy campestenol Campestenol 3-one Cathasterone CYP90C1/D1 6-Deoxoteasterone CYP90C1/D1 CYP90C1/D1 CYP85A1/A2 Teasterone 3-Epi-6-deoxocathasterone 22,23-Dihydroxy campesterol 22,23-Hydroxycampest-4-en-3-one 3-Dehydro-6-deoxoteasterone CYP85A1/A2 3-Dehydroteasterone CYP85A2 7-Oxateasterone CYP90C1/D1 6-Deoxotyphasterol 6-Deoxycatasterone CYP85A1/A2 6α-Hydroxycatasterone CYP85A1/A2 Typhasterol CYP85A2 7-Oxatyphasterol Catasterone CYP85A2 Brassinolide Figure 1. The proposed sterol/brassinosteroid biosynthesis pathway in Arabidopsis thaliana. The metabolic products are shown in boxes and the enzymes involved in the pathway are depicted in red. Mevalonate is converted via multiple steps into squalene (black dashed arrow). The early C-22 oxidation pathway is represented by light green arrows. The late C-6 oxidation pathway is visualized by red arrows and the early C-6 oxidation pathway by dark green arrows. Sterol/BR biosynthesis genes have been cloned from a number of species such as Arabidopsis, rice and tomato (Choe et al., 1998; Choe et al., 1999a; Choe et al., 1999b; Koka et al., 2000; Shimada et al., 2001; Tanabe et al., 2005). Most of the identified BR biosynthesis genes encode for cytochrome P450 monooxygenases. BR biosynthesis mutants and sterol mutants with defects in late steps of the sterol pathway such as dwarf7 (dwf7), dim (diminuto) and dwarf5 (dwf5) have a dwarfish appearance and can be rescued by exogenous application of BRs. Adult sterol mutants with defects in genes functioning in the early steps of the sterol biosynthesis pathway, such as fackel and hydra1 (hyd1), have 12 General introduction: Brassinosteroid biosynthesis and signalling in plants phenotypes similar to BR-biosynthesis mutants. However, mostly these mutants are arrested early in development as during embryogenesis they show distinctive defects (except for sterol methyl transferase 2; smt2) that cannot be rescued by exogenous application of BR. This suggests that these early sterols or their metabolites are involved in developmental processes during embryogenesis that are not dependent on BRs. There are multiple indications for a role of these sterols in correct auxin signalling. For instance, hyd1 and fackel mutants were shown to be defective in auxin and ethylene signalling (Lindsey et al., 2003). Schrick et al. (2004) showed that sterols are also important for cellulose synthesis and play a crucial role in cell elongation and cell wall expansion during the building of the primary cell wall. Although the BR biosynthesis pathway has been intensively studied still some questions remain. For some of the enzymes it is not completely clear in which BR conversion they are involved. For example, the BR biosynthesis enzyme CPD (CONSTITUTIVE PHOTOMORPHOGENIC DWARF, CYP90A1,) was suggested to act as a C-23 hydroxylase in Arabidopsis (Szekeres et al., 1996), but recently Ohnishi et al. (2006) proposed that CPD rather acts in the early C-22 oxidation pathway and performs a function different from C-23 hydroxylation. BRs are structurally similar to mammalian steroids. Two of the BR biosynthetic enzymes, the cytochrome P450 monooxygenase CPD and the 5α-reductase DE-ETIOLATED2 (DET2), are very similar to enzymes involved in animal steroid biosynthesis. Arabidopsis DET2 protein was able to catalyse the 5α reduction of several animal steroid substrates in human embryonic kidney cells (Li et al., 1997). Moreover, expression of the human steroid 5α-reductase types 1 or 2 both complemented the Arabidopsis det2 mutant. In health sciences there are multiple claims on the beneficial action of plant sterols (often refered to as phytosterols) as anti-cancer and cholesterol-lowering substances. However, many conflicting reports have been published. Nevertheless, plant sterols as well as BRs have been shown to inhibit the growth of several cancer lines (Malikova et al., 2008; Woyengo et al., 2009) and were also shown to reduce the levels of cholesterol and possess anti-inflammatory properties (Bouic, 2002; Woyengo et al., 2009). Brassinosteroid signalling In Arabidopsis, BRs are perceived by the membrane-localized leucine-rich repeat (LLR) receptor-like kinase BRASSINOSTEROID INSENSITIVE 1 (BRI1, Li and Chory, 1997). BRI1 has an extracellular domain containing 25 LRRs, a transmembrane domain and a cytoplasmic serine/threonine kinase domain (Li and Chory, 1997). The non-repetitive 13 Chapter 1 island found between the 21st en 22nd LRR in conjunction with LRR22 is of crucial importance for binding to BR (Kinoshita et al., 2005). BRI1 orthologs have been identified in a number of plant species such as tomato (CU-3), pea (LKA), rice (OsBRI1), barley (HvBRI1), cotton (GhBRI1) and petunia (PhBRI1) (Yamamuro et al., 2000; Montoya et al., 2002; Chono et al., 2003; Nomura et al., 2003; Sun et al., 2004) suggesting that plants perceive BRs on a similar manner. - BR + BR BRI1 BRI1 BAK1 BAK1 BSK1 BSK1 BSK1 p p BKI1 p p p BIN2 BSU1 BSU1 p 14-3-3 p BKI1 BZR1/BES1 BZR1/BES1 BIN2 proteasome BSU1 BZR1/BES1 p p p BZR1/BES1 BIN2 BIN2 BSU1 BZR1 BES1 BR taget gene BR taget gene Figure 2. Predicted BR signalling pathway. In the absence of BR (left), BKI1 suppresses BRI1 and BSK1; BAK1 and BSU1 are inactive. BIN2 is active and phosphorylates the transcription factors BZR1 and BES1, resulting in inhibition of their activity through proteasome mediated degradation, cytoplasmic retention by 14-3-3 proteins and inhibition of their DNA-binding and transcriptional activities. In the presence of BR (right), BR binds to the extracellular domain of BRI1 to activate its kinase, which in turn leads to disassociation of BKI1 from BRI1 and dimerization with and activation of BAK1. Subsequently, BRI1 phosphorylates BSK1, which in turn activates the plant-specific serine/threonine protein phosphatase BSU1 and BSU1 inactivates BIN2 by dephosphorylation at pTyr200. Unphosphorylated BES1 and BZR1 accumulate in the nucleus where they bind to the promoters of BRtarget genes, leading to diverse cellular en developmental responses. 14 General introduction: Brassinosteroid biosynthesis and signalling in plants BR binding induces autophosphorylation of BRI1 in vivo and a number of the phosphorylation sites have been identified and studied (Wang et al., 2005a). Characterization of these sites revealed that phosphorylation of the kinase domain is very important for kinase function and downstream BR signalling (Wang et al., 2005a). Furthermore, phosphorylation of the carboxy-terminal domain plays a critical role in the activation of events downstream of BRI1 (Wang et al., 2005b). Binding of BR also leads to dimerization of BRI1 with the co-receptor BRI1-ASSOCIATED RECEPTOR KINASE 1 (BAK1) that is subsequentely phosphorylated (Fig. 2). These phosporylation events lead to the dissociation of the repressor BRI1 KINASE INHIBITOR 1 (BKI1) from BRI1 (Li et al., 2002; Nam and Li, 2002; Wang and Chory, 2006). BRI1 and BAK1 are not only localized to the plasma membrane but also in endosomes. Co-expression of BRI1 and BAK1 in protoplasts of cowpea and Arabidopsis led to endocytosis of both receptors (Russinova et al., 2004). In addition, Geldner et al. (2007) observed localization of BRI1GFP in the endosomes of root cells of Arabidopsis that was independent of BR. Plants treated with Brefeldin A (BFA), which inhibits trafficking from the early to late endosomal compartments and vacuoles, enhanced the accumulation of BRI1 in endosomal compartments. Interestingly, similar to a BR treatment, BFA on its own was able to induce strong dephosphorylation of BES1 and suppressed DWARF4 (DWF4) gene expression. This suggests that BRI1 can signal from these endosomal compartments and is not just transported. BRI1 in the endosome might exist in an activated form because the negative regulator BKI1 is only present at the plasma membrane (Geldner et al., 2007). Recently, Song et al. (2009) demonstrated that the MEMBRANE STEROID-BINDING PROTEIN 1 (MSBP1) interacts with the extracellular domain of BAK1 in a BL-independent manner. Like BAK1, MSBP1 is localized in the plasma membrane and endosomal compartments. MSBP1 accelerates BAK1 endocytosis and MSBP1 expression reduced the interaction between BRI1 and BAK1 in vivo. This suggests that MSBP1 acts as a negative factor in the BR signalling pathway. Another class of BRI1 substrates that have been identified are members of the BRSIGNALING KINASES (BSKs) and represent the next step in the signalling cascade towards the nucleus (Tang et al., 2008). Recently Kim et al. (2009) showed that phosphorylation of BSK1 a member of the BSK family, facilitates interaction with and presumably activation of the plant-specific serine/threonine protein phosphatase BRI1 SUPPRESSOR 1 (BSU1). BSU1 is regarded as a positive regulator of the BR signalling pathway and encodes a serine-threonine protein phosphatase with a N-terminal Kelchrepeat domain and a C-terminal phosphatase domain (Mora-Garcia et al., 2004). 15 Chapter 1 Dephosphorylation of the GSK3/SHAGGY-like kinase BR-INSENSITIVE 2 (BIN2) by BSU1 inactivates BIN2 that acts as a negative regulator of the BR signalling pathway. BIN2’s role as negative regulator is performed through the phosphorylation of the closely related transcription factors BRI1-EMS-SUPPRESSOR 1 (BES1) and BRASSINAZOLE RESISTANT 1 (BZR1, He et al., 2002; Yin et al., 2002). Peng et al. (2010) recently showed that BIN2 interacts directly with BZR1 via a 12 amino acid BIN2-docking motif (DM) adjacent to the C-terminus of BZR1. A treatment with the proteasome inhibitor MG132 increased the accumulation of phosphorylated BZR1 protein and this suggests that phosphorylation of BZR1 and BES1 by BIN2 promotes the degradation of BZR1 and BES1 (He et al., 2002). Peng et al. (2010) reinforced this finding by demonstrating that Arabidopsis lines expressing BZR1-GFP without the BIN2-DM could not be phosphorylated and led to accumulation of BZR1. Vert & Chory (2006) demonstrated that phosphorylation inhibited the DNA-binding activity and transcription activation potential of BES1. A study in Arabidopsis demonstrated that 14-3-3 proteins interact with phosphorylated BZR1 (Gampala et al., 2007). Binding by 14-3-3 proteins was shown to increase the cytoplasmic retention of phosphorylated BES1 and mutation of a BIN2 phosphorylation site in BZR1 was shown to abolish 14-3-3 binding, leading to increased nuclear localization of the BZR1 protein. Similar results have been obtained with the rice homolog OsBZR1 (Bai et al., 2007). Taken together this suggests that BIN2-mediated phosphorylation inhibits BZR1 and BES1 by at least three ways: degradation by the proteasome, loss of DNA binding activity and cytoplasmic retention. Until recently, it was believed that BSU1 activated BES1 and BZR1 directly by dephosphorylation. However Kim et al. (2009) demonstrated that BSU1 does not dephosphorylate the BES1 and BZR1 transcription factors. In contrary, it dephosphorylates BIN2 and as a result, unphosphorylated BZR1/BES1 accumulate in the nucleus where they regulate the expression of BR-responsive genes. Thus, the effect of BSU1 on the phosphorylation status of BES1 and BZR1 is indirect and runs via inhibition of BIN2. BES1 and BZR1 belong to a group of transcription factors which are only found in plants (Li and Deng, 2005). BZR1 represses the transcription of several BR biosynthetic genes such as CPD and DWF4 by directly binding to their promoters via a motif called the BR response element (BRRE), thereby creating a feedback mechanism (He et al., 2005). BES1 has been found to interact with the basic helix-loop-helix transcription factor BES1Interacting Myc-like protein 1 (BIM1) to synergistically bind to E box sequences present in many BR-induced promoters such as the SAUR-AC1 promoter (Yin et al., 2005). Another target of BES1 is MYB30 (Li et al., 2008). Interestingly, MYB30 was shown to bind to a conserved MYB-binding site in BR target gene promoters creating a new mechanism in which BES1 and MYB30 cooperatively regulate BR gene expression during the early 16 General introduction: Brassinosteroid biosynthesis and signalling in plants stages of plant development. Yu et al. (2008) identified two JmjN/C domain-containing histone demethylases as putative BES1 partners, ELF6 (EARLY FLOWERING 6) and REF6 (RELATIVE OF EARLY FLOWERING 6). It therefore appears that the BES1 and BZR1 proteins act as hubs, interacting with proteins from diverse developmental pathways to activate and/or repress various target genes. This central role of BES1 and BZR1 is also reflected by the pleiotropic phenotype of BR mutants. Several proteins of the BR signalling pathway such as the GSK3-like kinases and 14-3-3 proteins are also conserved in mammals. Striking parallels with WINGLESS (Wnt), TRANSFORMING GROWTH FACTOR BETA (TGF beta), and RECEPTOR TYROSINE KINASE (RTK) signalling in animals have been proposed (Haubrick and Assmann, 2006). In similarity to the above described interactions between BIN2, BES1/BZR1 and 14-3-3 proteins in BR signaling, Wang et al. (2004) showed that phosphorylation of the HEATSHOCK TRANSCRIPTION FACTOR 1 (HSF1) by EXTRACELLULAR SIGNALREGULATED KINASE 1 (ERK1) and GSK3, and subsequent association of 14-3-3 proteins led to cytoplasmic sequestration of HSF1 and inhibition of HSF1 to bind DNA under certain growth conditions. The BRI1 receptor is a serine/threonine kinase that lacks tyrosine kinase activity. Most animal receptor-like kinases (RLK) have tyrosine kinase activity, with the exception of the TRANSFORMING GROWTH FACTOR β (TGF-β) receptors. Thus a potential split of RLKs is suggested prior to the evolutionary separation of plants and animals (Bishop and Koncz, 2002). Ehsan et al. (2005) reported that Arabidopsis TGF-β RECEPTOR INTERACTING PROTEIN (TRIP-1), a protein with high sequence similarity to a mammalian substrate of the TGF-β RII receptor kinase, as another phosphorylation target of BRI1. Plant GSK3-like kinases The initial goal of our project was to clarify the role of the GSK3-like kinase GSK1 from Arabidopsis in tolerance against sodium stress. Therefore, we will take a closer look on what is known about GSK3-like kinases, with emphasis on their roles in plants. The Arabidopsis genome encodes ten GSK3-like kinases and together with the other plant GSK3-like kinases they can be classified into four subgroups based on phylogenetic analyses of the amino acid- and cDNA sequences (Dornelas et al., 1998; Dornelas et al., 1999; Jonak and Hirt, 2002; Yoo et al., 2006). In plants, members of the GSK3-like kinase family seem to have both similar and distinct functions. The protein sequences of Arabidopsis GSK3-like kinases (ASKs) are highly conserved throughout the kinase 17 Chapter 1 domain, in contrast to the N- and C-terminal regions that are highly variable. Possibly this variability in the terminal regions might lead to specific functions of ASKs. For many plant GSK3-like kinases it is still elusive in what kind of pathway(s) they act (Table 1). A role for the three group II members in BR signalling is well established nowadays (Chapter 2; Vert and Chory, 2006; Koh et al., 2007; Kim et al., 2009; Yan et al., 2009). However, even in triple knock-outs for group II GSKs a substantial amount of BES1 remains phosphorylated, suggesting that other ASKs play a role in BR signalling as well (Vert and Chory, 2006; Yan et al., 2009). De Rybel et al. (2009) discovered a small molecule, called bikinin, which inhibits the activity of all members of group I and II and the group III member ASKθ. Inhibition of these seven GSKs activates BR responses, indicating the involvement of additional Arabidopsis GSKs in BR signalling. Recently Rozhon et al. (2010) provided stong evidence that ASKθ acts as a negative regulator in the BR signalling pathway as well. Interestingly, the group I members ASK11 and ASK12 also seem to be involved in floral patterning at several developmental stages (Dornelas et al., 2000). Group II GSK3-like kinases might also be involved in NaCl stress, although results of Piao et al. (1999; 2001) conflicts with results obtained by others (Chapter 2; Koh et al., 2007; Zeng et al., 2009). Piao et al. (2001) demonstrated that ectopic expression of Glycogen Synthase Kinase 1 (GSK1) in Arabidopsis leads to expression of multiple stress-induced genes and increased salt tolerance . However Koh et al. (2007) showed that an insertion mutant of rice OsGSK1, is more tolerant to salt stress and Zeng et al. (2009) showed that the semi-dominant mutant bin2-1 is very sensitive to salt. In Medicago sativa WIG seems to be involved in wounding and MSK4 in carbohydrate metabolism (Jonak et al., 2000; Kempa et al., 2007). It can be concluded that the majority of GSK proteins in Arabidopsis are involved in BR signalling. It is tempting to speculate that other roles inferred for GSKs in Arabidopsis and other species are direct and indirect effects of changes in BR signalling as well. Regulation of brassinosteroid levels BRs can be classified into C27, C28 and C29 steroids depending on the length of the sidechain. Not only the steroidal skeleton, but also its side-chain is metabolized by modifications such as hydroxylation, dehydrogenation and glycosylation (Bajguz, 2007). Both BR metabolism and catabolism seem to play an important role in the control of bioactive BR levels. For instance the cytochrome P450 monooxygenases CYP72B1 (BAS1) and CYP72C1 both catalyze the hydroxylation of the bioactive BRs brassinolide and catasterone generating inactive products (Turk et al., 2003; Nakamura et al., 2005). 18 ASKκ/ASK41 ASKδ/ASK42 MSK4 ASKβ/ASK31 ASKθ/ASK32 BSKθ NSK6 NSK59 NSK91 NSK111 PSK6 PSK7 WIG IV III II Arabidopsis thaliana Arabidopsis thaliana Brassica napus Nicotiana tabacum Nicotiana tabacum Nicotiana tabacum Nicotiana tabacum Petunia hybrida Petunia hybrida Medicago sativa Arabidopsis thaliana Arabidopsis thaliana Medicago sativa AJ002280 Y07822 Y12674 Y08607 AJ002315 AJ224163 AJ002314 AJ224164 AJ224165 AJ295939 X79279 AC079732 AF432225 Y13437 DQ060684 Arabidopsis thaliana Arabidopsis thaliana Arabidopsis thaliana Gossypium hirsutum Gossypium hirsutum Gossypium hirsutum Gossypium hirsutum Oryza sativa Oryza sativa Petunia hybrida Petunia hybrida X94938 X94939 X99696 ASKζ/ASK23 BIN2/ASKη/ASK21 GSK1/ASKι/ASK22 GhBIN2-B GhBIN2-C GhBIN2-D GhBIN2-E OSKη OSKGSK1 PSK8 PSK9 Dornelas et al., 1999; Vert & Chory 2006; Chapter 2 Dornelas et al., 1999; Li et al. Dornelas et al., 1999; Vert & Chory 2006; Chapter 2 Sun et al., 2004 Sun et al., 2004 Sun et al., 2004 Sun et al., 2004 Dornelas et al., 1999 Koh et al., 2007 Chapter 4 Chapter 4 Tichtisnky et al., 1998; Charrier et al., 2002 Tichtisnky et al., 1998; De Rybel et al., 2009 Tichtisnky et al., 1998 Tichtisnky et al., 1998 Tichtisnky et al., 1998 Tichtisnky et al., 1998 Tichtisnky et al., 1998 Tichtisnky et al., 1998 Tichtisnky et al., 1998 Jonak et al., 2000 Charrier et al., 2002 Charrier et al., 2002 Kempa et al., 2007 high expression in aerial organs, embryogenesis, expression tip cotyledon and weak in roots weak expression in aerial organs, embryogenesis and roots weak expression in aerial organs, tip cotyledon and roots consistently expressed throughout the plant highest expression in hypocotyls, leaf buds, immature bolls and ovules consistently expressed throughout the plant weak expression in aerial organs, tip cotyledon and roots not determined young shoots and roots, calli, developing seeds and highest in panicles not determined not determined flower buds, pollen; upregulated in the dark flower buds, pollen, roots not determined mainly in pollen and anther not determined not determined not determined stamen (highest), ovary, pollen stamen (highest),leaves,sepals,petals,styles,ovary,stigma, pollen roots,stems flowers, leaves (low) after wounding strongly upregulated very low expression in roots, leaves,inflorescence stems,flowers very low expression in roots, leaves,inflorescence stems,flowers; upregulated by NaCl or PEG in roots: amyloplasts, plastids, in epidermis and cortex cells; plastid of leaf mesophyll cells not determined possibly in BR signalling not determined not determined not determined not determined not determined probably in development, reproduction probably in development wounding probably in reproduction probably in abiotic stress NaCl stress, carbohydrate metabolism Einzenberger et al., 1995 Decroocq-Ferrant et al., 1995 Pay et al., 1993 Pay et al., 1993 Pay et al., 1993 BR signalling BR signalling, NaCl stress BR signalling, NaCl stress BR signalling BR signalling BR signalling BR signalling not determined BR signalling, abiotic stress probably in BR signalling not determined probably development development Nicotiana tabacum Petunia hybrida probably development probably development probably development Medicago sativa Medicago sativa Medicago sativa Dornelas et al., 1999; Dornelas et al., 2000; Kim et al., 2009 Dornelas et al., 1999; Dornelas et al., 2000; Kim et al., 2009 Charrier et al., 2002 BR signalling, flower patterning BR signalling, flower patterning BR signalling Arabidopsis thaliana Arabidopsis thaliana Arabidopsis thaliana X77763 X83619 X68410 X68409 Reference Expressed weak expression in aerial organs and hypocotyl, high floral meristem weak expression in aerial organs, tip cotyledon and hypocotyl, high floral meristem consititutively expressed in roots, leaves infloresescence stems, siliques, seedlings, flowers, upregulated by NaCL constituitve expression in leaves, petioles,roots, stems, nodes, flower buds, developing flowers, (un)opened mature flowers,seed pods, various stage of somatic embryos constituitve expression in roots, stems and nodes, flower buds, globular embryos weak expression in roots and stems, moderately in nodes, high in early flower development gradually decreased in later stages and absent after fertilization, moderate in globular embryos and decreased in later stages high expression corolla, ovary, during pollen development; low expression in leaves stems,roots, ovules, embry sac and highest expression in leaves, flowers, ovaries and flower buds Function Species NTK-1 PSK4 I X68411 MSK2 MSK3 MSK1 Accesion no. X75432 X75431 AL163792 Group ASKα/ASK11 ASKγ/ASK12 ASKε/ASK13 GSK3 like kinases Table 1. Plant GSK3-like kinases General introduction: Brassinosteroid biosynthesis and signalling in plants 19 Chapter 1 Nomura et al. (2001) suggested that C-6 oxidation as well as steroid side chain hydroxylation at C-22 and C-23 are potential rate-limiting reactions in BR biosynthesis in Arabidopsis, tomato and pea. Several experiments have shown that the BR signalling pathway tightly controls the biosynthesis of BRs. In a bri1-5 mutant background, the BR biosynthesis genes CPD, DET2 and DWF4 are upregulated, while CYP72B1 is downregulated (Choe et al., 2001). Severe bri1 alleles as well as the semi-dominant allele of BIN2 accumulate high levels of catasterone and brassinolide (Noguchi et al., 1999; Choe et al., 2002). Wang et al. (2002) demonstrated that a dominant mutation in BZR1 reduced the level of BRs and CPD. In addition, He et al. (2005) demonstrated that BZR1 repressed the transcription of several BR biosynthetic genes such as CPD and DWF4. This suggests that BZR1 acts as a repressor on the BR synthesis pathway. Taken together these data indicate that BR levels are regulated by a feedback mechanism mediated by the BR signalling pathway. Transport of brassinosteroids The distribution of BRs in plants is widespread; with higher levels found in young growing tissue (such as shoot tip and young internodes) in comparison to older vegetative tissue. The highest levels of bioactive BRs are found in pollen and immature seeds (Bajguz and Tretyn, 2003; Symons and Reid, 2004). Several studies have shown that a number of BR biosynthesis enzymes are membranebound and since steroid hormones are relatively hydrophobic, BR synthesis is thought to take place on or near internal cell membranes such as the endoplasmatic reticulum (Klahre et al., 1998; Shimada et al., 2001). Assuming that BRs are synthesized within the cell, transport out of the cell is required if BRs are perceived at the exterior cell surface by the BRI1 receptor. Yamamoto et al. (2001) observed that cultured Zinnia elegans callus tissue secreted BRs in the medium, supporting the idea that BRs are transported out of the cell. At the moment it is unclear how BR transport takes place, but perhaps through conjugation with fatty acids or glucose or via proteinaceous transporters BRs are transported out of the cell (reviewed by Symons et al., 2008). Since signalling can also occur via endosomal BRI1, there seems no absolute requirement for BR transport (Geldner et al., 2007). Although short-distance transport of BRs appears likely, is it questionable whether BRs are transported over a longer distance. Numerous studies have tried to answer this question, but the obtained results are often contrary. Application of radio-labelled BRs to the roots of rice, wheat and cucumber plantlets showed that exogenous applied BRs were transported into the shoot (Nishikawa et al., 1994). Furthermore, Arabidopsis BR biosynthesis mutants were rescued when they were grown on media containing bioactive BRs (Choe et al., 20 General introduction: Brassinosteroid biosynthesis and signalling in plants 1998). On the other hand exogenous BRs applied to leaves of pea, rice, cucumber and wheat were relatively immobile (Yokota et al., 1992; Nishikawa et al., 1994). With respect to the feeding experiments it is doubtful whether exogenous applied BRs do reflect endogenous BRs in plants as grafting experiments in pea showed that BRs are not transported from the root to shoot or vice versa. In addition, removing the apical bud did not result in any significant changes in BR levels in either the stem or leaf tissues of pea and removal of the three youngest expanded leafs did not substantially alter endogenous BR levels in the stem or apical tissues (Symons and Reid, 2004). More evidence that plants do not undergo long distance transport of BRs is derived from the unstable BR-deficient dx mutant of tomato. Revertant sectors in dx did not restore the mutant phenotype. Again, the immobility of BRs was also revealed by reciprocal grafting experiments using dx and wild type plants (Bishop et al., 1996; Montoya et al., 2005). Interestingly, depletion of active BRs through overexpression of the catabolic enzyme BAS1 in the epidermis of wild type Arabidopsis reduced growth, suggesting that growth depends on local BR biosynthesis (Savaldi-Goldstein et al., 2007). Considering all studies, it seems most likely that BRs are produced at the location where they are needed and that the BR levels in plant tissues are rather regulated through tissue-specific control of BR biosynthesis, catabolism and inactivation than via long-distance travelling. Physiological responses to brassinosteroids BRs are involved in various growth- and developmental processes such as cell division and elongation, vascular differentiation, abiotic- and biotic stresses, photomorphogenesis, senescence and fertility (Clouse et al., 1996; Li et al., 1996; Szekeres et al., 1996; Bajguz and Hayat, 2009). The major cause of dwarfism is reduced cell size and several studies proved the involvement of BRs in cell elongation (Kauschmann et al., 1996; Azpiroz et al., 1998). For example, BR application promotes the elongation of stems, petioles and flower peduncles in dicotyledons as well as coleoptiles and mesocotyls in monocots (reviewed by Clouse and Sasse, 1998). Savaldi-Goldstein et al. (2007) showed that expression of the BRI1 receptor or BR biosynthesis gene CPD in the epidermis of Arabidopsis BR mutants rescues their dwarf phenotype, but not when expressed in underlying tissues. Their results suggest that BR signalling in the epidermis is an important factor in leaf growth. Reinhardt et al. (2007) showed that DWF4 expression in the leaf margin cells of a dwf4 mutant restored leaf shape, but not leaf size. This insinuates that BRs play a role in leaf shape. 21 Chapter 1 In several plant species BRs were shown to upregulate the expression of XYLOGLUCAN ENDOTRANSGLYCOLASES/HYDROLASES (XETs or XHTs), which are a group of enzymes with cell-wall modifying activities (Zurek and Clouse, 1994; Xu et al., 1996; Koka et al., 2000). As (BL-induced) cell expansion requires cell wall loosening it is imaginable that these enzymes are induced. Microarray analysis in Arabidopsis showed that also other genes involved in cell expansion and cell wall organization such as expansins, extensins and pectin modifying enzymes are upregulated after BL treatment (Goda et al., 2002). These data suggest that BRs play an important role in the regulation of cell wall expansion, possibly through controlling the expression or the activity of wall modifying factors. Also a role of BRs in microtubule organization has been implicated, as brassinolide was able to restore the microtubule organisation and rescue cell elongation in the Arabidopsis bul1 mutant (Catterou et al., 2001). BRs also seem to play a role in cell division. Nakajima et al. (1996) showed that addition of epi-brassinolide to Chinese cabbage mesophyll protoplasts increased the cell division rate. This is further supported by Oh & Clouse (1998) who observed acceleration of the cell division rate in leaf mesophyll protoplasts of Petunia hybrida under optimal auxin and cytokinin conditions. They showed that a high concentration of BR rather had an inhibitory effect on cell division and that the effect of BR is dependent on the auxin/cytokinin ratio. However, microscopic examination of leaves of BR biosynthesis and BR signalling mutants of Arabidopsis led to some suspicion as the dwarfish phenotype was due to a reduced cell size and not a consequence of a decreased number of cells (Kauschmann et al., 1996). Further studies need to clarify the role of BRs in cell division. Several experiments suggest that BRs are involved in vascular differentiation. Cano Delgado et al. (2004) identified three genes, designated BRI1-LIKE RECEPTOR KINASE 1 (BRL1), BRL2 and BRL3, which have a high sequence identity with BRI1. BRL1 and BRL3 were shown to complement a bri1 mutant when expressed under control of the BRI1 promoter and both can bind to brassinolide with high affinity. A brl1 mutant showed an increase in phloem and a decrease in xylem differentiated cells. Knocking out of brl1 in a weak bri1 mutant background enhanced the vascular defects. A bri1brl1brl3 mutant was smaller than the strong bri1 allele (bri1-101) and the vasculature defects were enhanced. Altogether this implies that BR signalling in provascular cells through BRL1, BRL3 and BRI1 receptors contributes to the maintenance of proper xylem/phloem ratios. In addition, Nagata et al. (2001) reported that brassinazole, a BR biosynthesis inhibitor, inhibited the development of secondary xylem in cress. Brassinolide was shown to accelerate senescence in leaves of mung bean seedlings and in the detached cotyledons of cucumber seedlings (reviewed by Clouse and Sasse, 1998). BR biosynthesis and signalling mutants have an expanded lifespan when compared to normal 22 General introduction: Brassinosteroid biosynthesis and signalling in plants plants, supporting the idea that BRs accelerate senescence in plants. To clarify the role of BRs in senescence it would be valuable to examine the effect of BR application on senescence-associated mutants of Arabidopsis and to study the expression of senescenceassociated genes in BR mutants. Another common characteristic of most BR biosynthesis and signalling mutants is reduced fertility or male sterility. In Prunus avium brassinolide was shown to improve elongation of pollen tubes (Hewitt et al., 1985). The cpd mutant was shown to be male sterile as its pollen were unable to elongate during germination (Szekeres et al., 1996). The dwf4 mutant on the other hand has viable pollen, but they cannot reach the stigma because the stamen filaments failed to elongate and are shorter than the carpels which results in the deposition of pollen on the ovary wall rather than onto the stigmatic surface (Choe et al., 1998). The only mutant with stamens longer than the gynoecium with wild-type fertility is dwf5-1 (Choe et al., 2000). All other identified dwf5 alleles and other BR biosynthesis mutants have stamens which are shorter than the gynoecium. Although dwf5-1 plants had siliques with more seeds than the other BR mutants, seeds did not develop normally and exogenous application of BR was shown to improve germination. Another prominent feature of several BR biosynthesis mutants is their de-etiolated phenotype; dark grown seedlings having the appearance of light-grown seedlings with short hypocotyls and open cotyledons. Arabidopsis BR biosynthesis mutants such as cpd and det2 possess a de-etiolated phenotype when grown in the dark and they stimulate the expression of light-regulated genes (Chory et al., 1991; Szekeres et al., 1996). BR biosynthesis mutants of tomato and rice also develop a de-etiolated phenotype in the dark, but this is not the case for BR biosynthesis mutants of pea (reviewed by Fujioka and Yokota, 2003). This suggests that the role of BRs in photomorphogenesis might differ among plant species. Bancos et al. (2006) demonstrated that the expression of the Arabidopsis BR biosynthesis genes CPD and CYP85A2 are under diurnal regulation. The expression profile of CPD is circadian regulated and dependent on light. In addition, the BR-catabolyzing enzymes CYP72B1 and CYP72C1 are repressed by light (Turk et al., 2003; Nakamura et al., 2005). As brassinolide was shown to accumulate in the light, it might be that the level of active BRs is influenced by the diurnal expression of both ratelimiting biosynthetic enzymes and enzymes that inactivate BRs. There are several reports published describing the relationship between BRs and abiotic stresses such as salt, heavy metals, high and low temperature stress and drought (reviewed by Bajguz and Hayat, 2009). The role of BRs in pathogenic defence has also been investigated. For instance the addition of BR containing extract of Lychnis viscaria L. seeds yielded an increased resistance of tobacco, cucumber and tomato to viral and fungal 23 Chapter 1 pathogens (Roth et al., 2000). Furthermore, Nakashita et al. (2003) demonstrated that brassinolide independently of salicylic acid (SA) enhanced the resistance to the viral pathogen tobacco mosaic virus (TMV), the fungal pathogen Odium sp. and the bacterial pathogen Pseudomonas syringae pv. tabaci. In rice, brassinolide enhanced the resistance against the fungal pathogen Magnaporthe grisea and the bacterial pathogen Xanthomonas oryzae pv. Oryzae. This suggests that brassinolide acts as inducer of a broad range of disease resistances in rice and tobacco and possibly other plant species. Cross talk with other plant hormones Interactions of BRs with other plant hormones are well known. BRs were shown to induce the expression of OPDA-REDUCTASE 1 (OPR3), an enzyme involved in jasmonate biosynthesis and induction of OPR3 gene expression by membrane damage was impaired in the BR- biosynthesis mutant cabbage 1 (cbb1, Schaller et al., 2000). This suggests intertwinement of the jasmonate and BR biosynthesis pathway. BRs and gibberellins (GA) were shown to antagonistically regulate the expression of the GA responsive gene GASA1 and the GA-repressible gene GA5 (Bouquin et al., 2001). Shimada et al. (2006) reported that mutation of the rice SPINDLY gene (OsSPY), a negative regulator of the GA signalling pathway, reduced the expression of several BR biosynthesis genes and OsBRI1. Wang et al. (2009) reported that GA-induced OsGSR1 was repressed upon BR treatment. Transgenic plants in which OsGSR1 expression was down regulated by RNA interference had phenotypes similar to BR mutants, had a reduced level of endogenous BRs and were rescued upon a brassinolide treatment. Interestingly, OsGSR1 interacted with DIM, a BR biosynthetic enzyme, suggesting that GA directly activate BR biosynthesis via OsGSR1. Further analysis should reveal the cross talk mechanisms between BRs and GA. Microarray analysis has shown that several BR responsive genes are also regulated by ABA (Nemhauser et al., 2006). Germination of det2 and bri1-1 mutants is more strongly inhibited by abscisic acid (ABA) than wild type seeds (Steber and McCourt, 2001). This suggest that ABA and BRs work antagonistic in seed germination and that a BR signal is needed to overcome the inhibition of germination by ABA. Subsequently, Zhang et al. (2009) showed that in Arabidopsis seedlings ABA rapidly induced the expression of the BR-suppressed genes CPD and DWF4 and also phosphorylation of BES1 was enhanced. In the absence of BIN2, ABA failed to induce BES1 phosphorylation. Taken together, their results suggest that the ABA signalling pathway cross talks with the BR signalling pathway downstream of the BRI1 receptor but upstream of BES1. 24 General introduction: Brassinosteroid biosynthesis and signalling in plants The synergistic action of BRs and auxin in cell growth and development is a well-known phenomenon (reviewed by Mandava, 1988; Halliday, 2004; Hardtke et al., 2007). Recently Ibañes et al. (2009) showed that auxin transport coupled to BR signalling is required to set the number and arrangement of plant-shoot vasculature. Several studies showed that BRs and auxins regulate a similar set of target genes such as genes belonging to the AUX/IAA, SAUR and GH3 families (Nakamura et al., 2003; Goda et al., 2004; Nemhauser et al., 2004). Most of the BR/auxin target genes are coordinately regulated, so if a gene is repressed by auxin it is usually also repressed by BR. AUX/IAA genes were expressed at a lower level in the BR biosynthesis mutant det2 and in the BR signalling mutant bri1 and AUX/IAA expression could be restored after BL application in the det2 mutant. In addition some auxin-responsive genes were up regulated in the BES1-D mutant (Yin et al., 2002). Nemhauser et al. (2004) observed that the TGTCTC auxin-responsive element is enriched in brassinolide-regulated gene promoter sequences. The core ARF-binding element TGTC was enriched in promoters of brassinolide, auxin and dual regulated genes. DR5, a wellcharacterized synthetic element containing ARF binding sites, was not only induced by auxin, but also by BRs and both hormones are required for full induction (Nemhauser et al., 2004). Like auxin, BRs promote primary root growth at low concentrations (Mussig et al., 2003). In the Arabidopsis mutant brevis radix (brx) impaired auxin-responsive gene expression was observed which could be restored upon a brassinolide treatment. BRX expression was strongly induced by auxin, but mildly repressed by brassinolide. This suggests that BR levels are rate-limiting for auxin-induced transcriptional responses. BRX was shown to regulate the expression of the BR biosynthesis gene CPD, suggesting a feedback loop that involves the BRX transcriptional regulator. Chandler et al. (2009) showed that BIM1, formerly known to interact with BES1, controls embryonic patterning via its interaction with DORNROESCHEN (DRN) and DORNROESCHEN-LIKE (DRNL). Given that DRN is also involved in auxin signalling it is conceivable that the auxin and BR signalling pathways control embryonic patterning via DRN. Work by Vert et al. (2008) showed that the negative regulator of the BR pathway, BIN2, binds and phosphorylates the AUXIN RESPONSE FACTOR 2 (ARF2). They suggest that repressor ARFs, such as ARF2 may compete with activator ARFs for binding auxin responsive elements (AuxRes). Phosphorylation of repressor ARFs by BIN2 would inactivate them leaving the AuxRes more accessible for activator ARFs. De Grauwe et al. (2005) proposed that ethylene, auxin and BRs all play a role in the formation of the apical hook. Gendron et al. (2008) discovered that a compound named brassinopride (BRP) inhibits BR action and promotes ethylene action. Their results suggest that active downstream BR signalling is required for ethylene-induced apical hook formation (Li et al., 2004). ARF2 protein levels 25 Chapter 1 were also decreased upon ethylene treatment but accumulate in the light and transcription of HOOKLESS1 (HLS1), a N-actetyltransferase, was up regulated by ethylene but down regulated by light. All together this insinuates that ARF2 and possibly other ARF repressors are at the convergence of BRs, ethylene, light, and auxin signalling networks. Aims and outline of this thesis The goal of this PhD research project was to investigate the role of the GSK3-like kinase GSK1 and other group II GSK3-like kinases (BIN2 and ASKζ) in Na+ stress. In Chapter 2 we analysed the effect of Na+ stress in several insertion mutants. In contrast to results of others we could not detect a role in tolerance against Na+ stress for GSK1 or for the other group II members. In a yeast two hybrid assay, GSK1 interacted with a member of the BES1/BZR1 family. As BIN2 negatively regulates the BR signalling pathway by phosphorylating the transcription factors BES1 and BZR1, it prompted us to investigate whether the group II GSK3-like kinases act redundantly in the BR signalling pathway. In order to learn more about the mode of actions of GSK1 and the other group II GSK3-like kinases in BR signalling and possibly in other signal transduction pathways we tried to isolate the protein complexes with help of various Tandem Affinity Purification (TAP) methods described in Chapter 3. Within our petunia collection we found several mutant families that resemble sterol/BR biosynthesis and signalling mutants. In Chapter 4 we characterized and analysed petunia sterol/BR biosynthesis mutants. For four families, the mutant phenotype was caused by the insertion of transposable elements in the BR biosynthesis gene CPD. Furthermore we report the identification of a bri1 mutant in petunia and isolated several homologs of key regulators of the Arabidopsis BR signalling pathway from petunia. References Azpiroz, R., Wu, Y., LoCascio, J.C. and Feldmann, K.A. (1998) An Arabidopsis brassinosteroid-dependent mutant is blocked in cell elongation. Plant Cell, 10, 219-230. Bai, M.Y., Zhang, L.Y., Gampala, S.S., Zhu, S.W., Song, W.Y., Chong, K. and Wang, Z.Y. (2007) Functions of OsBZR1 and 14-3-3 proteins in brassinosteroid signaling in rice. Proc Natl Acad Sci U S A, 104, 13839-13844. Bajguz, A. and Tretyn, A. (2003) The chemical characteristic and distribution of brassinosteroids in plants. Phytochemistry, 62, 1027-1046. Bajguz, A. (2007) Metabolism of brassinosteroids in plants. Plant Physiol Biochem, 45, 95-107. Bajguz, A. and Hayat, S. (2009) Effects of brassinosteroids on the plant responses to environmental stresses. Plant Physiol Biochem, 47, 1-8. 26 General introduction: Brassinosteroid biosynthesis and signalling in plants Bancos, S., Szatmari, A.M., Castle, J., Kozma-Bognar, L., Shibata, K., Yokota, T., Bishop, G.J., Nagy, F. and Szekeres, M. (2006) Diurnal regulation of the brassinosteroid-biosynthetic CPD gene in Arabidopsis. Plant Physiol, 141, 299-309. Bishop, G.J., Harrison, K. and Jones, J.D. (1996) The tomato Dwarf gene isolated by heterologous transposon tagging encodes the first member of a new cytochrome P450 family. Plant Cell, 8, 959-969. Bishop, G.J. and Koncz, C. (2002) Brassinosteroids and plant steroid hormone signaling. Plant Cell, 14 Suppl, S97-110. Bouic, P.J. (2002) Sterols and sterolins: new drugs for the immune system? Drug Discov Today, 7, 775-778. Bouquin, T., Meier, C., Foster, R., Nielsen, M.E. and Mundy, J. (2001) Control of specific gene expression by gibberellin and brassinosteroid. Plant Physiol, 127, 450-458. Cano-Delgado, A., Yin, Y., Yu, C., Vafeados, D., Mora-Garcia, S., Cheng, J.C., Nam, K.H., Li, J. and Chory, J. (2004) BRL1 and BRL3 are novel brassinosteroid receptors that function in vascular differentiation in Arabidopsis. Development, 131, 5341-5351. Catterou, M., Dubois, F., Schaller, H., Aubanelle, L., Vilcot, B., Sangwan-Norreel, B.S. and Sangwan, R.S. (2001) Brassinosteroids, microtubules and cell elongation in Arabidopsis thaliana. II. Effects of brassinosteroids on microtubules and cell elongation in the bul1 mutant. Planta, 212, 673-683. Chandler, J.W., Cole, M., Flier, A. and Werr, W. (2009) BIM1, a bHLH protein involved in brassinosteroid signalling, controls Arabidopsis embryonic patterning via interaction with DORNROSCHEN and DORNROSCHEN-LIKE. Plant Mol Biol, 69, 57-68. Charrier, B., Champion, A., Henry, Y. and Kreis, M. (2002) Expression profiling of the whole Arabidopsis shaggy-like kinase multigene family by real-time reverse transcriptase-polymerase chain reaction. Plant Physiol, 130, 577-590. Choe, S., Dilkes, B.P., Fujioka, S., Takatsuto, S., Sakurai, A. and Feldmann, K.A. (1998) The DWF4 gene of Arabidopsis encodes a cytochrome P450 that mediates multiple 22alpha-hydroxylation steps in brassinosteroid biosynthesis. Plant Cell, 10, 231-243. Choe, S., Dilkes, B.P., Gregory, B.D., Ross, A.S., Yuan, H., Noguchi, T., Fujioka, S., Takatsuto, S., Tanaka, A., Yoshida, S., Tax, F.E. and Feldmann, K.A. (1999) The Arabidopsis dwarf1 mutant is defective in the conversion of 24-methylenecholesterol to campesterol in brassinosteroid biosynthesis. Plant Physiol, 119, 897907. Choe, S., Noguchi, T., Fujioka, S., Takatsuto, S., Tissier, C.P., Gregory, B.D., Ross, A.S., Tanaka, A., Yoshida, S., Tax, F.E. and Feldmann, K.A. (1999) The Arabidopsis dwf7/ste1 mutant is defective in the delta7 sterol C-5 desaturation step leading to brassinosteroid biosynthesis. Plant Cell, 11, 207-221. Choe, S., Tanaka, A., Noguchi, T., Fujioka, S., Takatsuto, S., Ross, A.S., Tax, F.E., Yoshida, S. and Feldmann, K.A. (2000) Lesions in the sterol delta reductase gene of Arabidopsis cause dwarfism due to a block in brassinosteroid biosynthesis. Plant J, 21, 431-443. Choe, S., Fujioka, S., Noguchi, T., Takatsuto, S., Yoshida, S. and Feldmann, K.A. (2001) Overexpression of DWARF4 in the brassinosteroid biosynthetic pathway results in increased vegetative growth and seed yield in Arabidopsis. Plant J, 26, 573-582. Choe, S., Schmitz, R.J., Fujioka, S., Takatsuto, S., Lee, M.O., Yoshida, S., Feldmann, K.A. and Tax, F.E. (2002) Arabidopsis brassinosteroid-insensitive dwarf12 mutants are semidominant and defective in a glycogen synthase kinase 3beta-like kinase. Plant Physiol, 130, 1506-1515. Choi, Y., Fujioka, S., Nomura, T., Harada, A., Yokota, T., Takatsuto, S. and Sakurai, A. (1996) An alternative brassinolide biosynthetic pathway via late C-6 oxidation Phytochemistry, 44, 609-613. 27 Chapter 1 Chono, M., Honda, I., Zeniya, H., Yoneyama, K., Saisho, D., Takeda, K., Takatsuto, S., Hoshino, T. and Watanabe, Y. (2003) A semidwarf phenotype of barley uzu results from a nucleotide substitution in the gene encoding a putative brassinosteroid receptor. Plant Physiol, 133, 1209-1219. Chory, J., Nagpal, P. and Peto, C.A. (1991) Phenotypic and Genetic Analysis of det2, a New Mutant That Affects Light-Regulated Seedling Development in Arabidopsis. Plant Cell, 3, 445-459. Clouse, S.D., Langford, M. and McMorris, T.C. (1996) A brassinosteroid-insensitive mutant in Arabidopsis thaliana exhibits multiple defects in growth and development. Plant Physiol, 111, 671-678. Clouse, S.D. and Sasse, J.M. (1998) BRASSINOSTEROIDS: Essential Regulators of Plant Growth and Development. Annu Rev Plant Physiol Plant Mol Biol, 49, 427-451. De Grauwe, L., Vandenbussche, F., Tietz, O., Palme, K. and Van Der Straeten, D. (2005) Auxin, ethylene and brassinosteroids: tripartite control of growth in the Arabidopsis hypocotyl. Plant Cell Physiol, 46, 827-836. De Rybel, B., Audenaert, D., Vert, G., Rozhon, W., Mayerhofer, J., Peelman, F., Coutuer, S., Denayer, T., Jansen, L., Nguyen, L., Vanhoutte, I., Beemster, G.T., Vleminckx, K., Jonak, C., Chory, J., Inze, D., Russinova, E. and Beeckman, T. (2009) Chemical inhibition of a subset of Arabidopsis thaliana GSK3-like kinases activates brassinosteroid signaling. Chem Biol, 16, 594-604. Decroocq-Ferrant, V., Van Went, J., Bianchi, M.W., de Vries, S.C. and Kreis, M. (1995) Petunia hybrida homologues of shaggy/zeste-white 3 expressed in female and male reproductive organs. Plant J, 7, 897-911. Dornelas, M.C., Lejeune, B., Dron, M. and Kreis, M. (1998) The Arabidopsis SHAGGY-related protein kinase (ASK) gene family: structure, organization and evolution. Gene, 212, 249-257. Dornelas, M.C., Wittich, P., von Recklinghausen, I., van Lammeren, A. and Kreis, M. (1999) Characterization of three novel members of the Arabidopsis SHAGGY-related protein kinase (ASK) multigene family. Plant Mol Biol, 39, 137-147. Dornelas, M.C., Van Lammeren, A.A. and Kreis, M. (2000) Arabidopsis thaliana SHAGGY-related protein kinases (AtSK11 and 12) function in perianth and gynoecium development. Plant J, 21, 419-429. Ehsan, H., Ray, W.K., Phinney, B., Wang, X., Huber, S.C. and Clouse, S.D. (2005) Interaction of Arabidopsis BRASSINOSTEROID-INSENSITIVE 1 receptor kinase with a homolog of mammalian TGF-beta receptor interacting protein. Plant J, 43, 251-261. Einzenberger, E., Eller, N., Heberle-Bors, E. and Vicente, O. (1995) Isolation and expression during pollen development of a tobacco cDNA clone encoding a protein kinase homologous to shaggy/glycogen synthase kinase-3. Biochim Biophys Acta, 1260, 315-319. Fujioka, S., Takatsuto, S. and Yoshida, S. (2002) An early C-22 oxidation branch in the brassinosteroid biosynthetic pathway. Plant Physiol, 130, 930-939. Fujioka, S. and Yokota, T. (2003) Biosynthesis and metabolism of brassinosteroids. Annu Rev Plant Biol, 54, 137-164. Gampala, S.S., Kim, T.W., He, J.X., Tang, W., Deng, Z., Bai, M.Y., Guan, S., Lalonde, S., Sun, Y., Gendron, J.M., Chen, H., Shibagaki, N., Ferl, R.J., Ehrhardt, D., Chong, K., Burlingame, A.L. and Wang, Z.Y. (2007) An essential role for 14-3-3 proteins in brassinosteroid signal transduction in Arabidopsis. Dev Cell, 13, 177-189. Geldner, N., Hyman, D.L., Wang, X., Schumacher, K. and Chory, J. (2007) Endosomal signaling of plant steroid receptor kinase BRI1. Genes Dev, 21, 1598-1602. Gendron, J.M. and Wang, Z.Y. (2007) Multiple mechanisms modulate brassinosteroid signaling. Curr Opin Plant Biol, 10, 436-441. Gendron, J.M., Haque, A., Gendron, N., Chang, T., Asami, T. and Wang, Z.Y. (2008) Chemical genetic dissection of brassinosteroid-ethylene interaction. Mol Plant, 1, 368-379. 28 General introduction: Brassinosteroid biosynthesis and signalling in plants Goda, H., Shimada, Y., Asami, T., Fujioka, S. and Yoshida, S. (2002) Microarray analysis of brassinosteroidregulated genes in Arabidopsis. Plant Physiol, 130, 1319-1334. Goda, H., Sawa, S., Asami, T., Fujioka, S., Shimada, Y. and Yoshida, S. (2004) Comprehensive comparison of auxin-regulated and brassinosteroid-regulated genes in Arabidopsis. Plant Physiol, 134, 1555-1573. Grove, M.D., Gayland, F.S., Rohwedder, W.K., Nagabhushanam, M., Worley, J.F., Warthen Jr, J.D., Steffens, G.L., Flippen-Anderson, J.L. and Cook Jr, J.C. (1979) Brassinolide, a plant growth-promoting steroid isolated from Brassica napus pollen. Nature, 281, 2. Halliday, K.J. (2004) Plant hormones: the interplay of brassinosteroids and auxin. Curr Biol, 14, R1008-1010. Hardtke, C.S., Dorcey, E., Osmont, K.S. and Sibout, R. (2007) Phytohormone collaboration: zooming in on auxin-brassinosteroid interactions. Trends Cell Biol, 17, 485-492. Haubrick, L.L. and Assmann, S.M. (2006) Brassinosteroids and plant function: some clues, more puzzles. Plant Cell Environ, 29, 446-457. He, J.X., Gendron, J.M., Yang, Y., Li, J. and Wang, Z.Y. (2002) The GSK3-like kinase BIN2 phosphorylates and destabilizes BZR1, a positive regulator of the brassinosteroid signaling pathway in Arabidopsis. Proc Natl Acad Sci U S A, 99, 10185-10190. He, J.X., Gendron, J.M., Sun, Y., Gampala, S.S., Gendron, N., Sun, C.Q. and Wang, Z.Y. (2005) BZR1 is a transcriptional repressor with dual roles in brassinosteroid homeostasis and growth responses. Science, 307, 16341638. Hewitt, F.R., Hough, T., O'Neill, P., Sasse, J.M., Williams, E.G. and Rowan, K.S. (1985) Effect of Brassinolide and other Growth Regulators on the Germination and Growth of Pollen Tubes of Prunus avium using a Multiple Hanging-drop Assay. Aust. J. Plant Physio, 12, 201-211. Ibanes, M., Fabregas, N., Chory, J. and Cano-Delgado, A.I. (2009) Brassinosteroid signaling and auxin transport are required to establish the periodic pattern of Arabidopsis shoot vascular bundles. Proc Natl Acad Sci U S A, 106, 13630-13635. Jonak, C., Beisteiner, D., Beyerly, J. and Hirt, H. (2000) Wound-induced expression and activation of WIG, a novel glycogen synthase kinase 3. Plant Cell, 12, 1467-1475. Jonak, C. and Hirt, H. (2002) Glycogen synthase kinase 3/SHAGGY-like kinases in plants: an emerging family with novel functions. Trends Plant Sci, 7, 457-461. Kauschmann, A., Jessop, A., Koncz, C., Szekeres, M., Willmitzer, L. and Altmann, T. (1996) Genetic evidence for an essential role of brassinosteroids in plant development. The Plant Journal, 9, 701-713. Kempa, S., Rozhon, W., Samaj, J., Erban, A., Baluska, F., Becker, T., Haselmayer, J., Schleiff, E., Kopka, J., Hirt, H. and Jonak, C. (2007) A plastid-localized glycogen synthase kinase 3 modulates stress tolerance and carbohydrate metabolism. Plant J, 49, 1076-1090. Kim, T.W., Guan, S., Sun, Y., Deng, Z., Tang, W., Shang, J.X., Burlingame, A.L. and Wang, Z.Y. (2009) Brassinosteroid signal transduction from cell-surface receptor kinases to nuclear transcription factors. Nat Cell Biol, 11, 1254-1260. Kinoshita, T., Cano-Delgado, A., Seto, H., Hiranuma, S., Fujioka, S., Yoshida, S. and Chory, J. (2005) Binding of brassinosteroids to the extracellular domain of plant receptor kinase BRI1. Nature, 433, 167-171. Klahre, U., Noguchi, T., Fujioka, S., Takatsuto, S., Yokota, T., Nomura, T., Yoshida, S. and Chua, N.H. (1998) The Arabidopsis DIMINUTO/DWARF1 gene encodes a protein involved in steroid synthesis. Plant Cell, 10, 1677-1690. 29 Chapter 1 Koh, S., Lee, S.C., Kim, M.K., Koh, J.H., Lee, S., An, G., Choe, S. and Kim, S.R. (2007) T-DNA tagged knockout mutation of rice OsGSK1, an orthologue of Arabidopsis BIN2, with enhanced tolerance to various abiotic stresses. Plant Mol Biol, 65, 453-466. Koka, C.V., Cerny, R.E., Gardner, R.G., Noguchi, T., Fujioka, S., Takatsuto, S., Yoshida, S. and Clouse, S.D. (2000) A putative role for the tomato genes DUMPY and CURL-3 in brassinosteroid biosynthesis and response. Plant Physiol, 122, 85-98. Li, J., Nagpal, P., Vitart, V., McMorris, T.C. and Chory, J. (1996) A role for brassinosteroids in lightdependent development of Arabidopsis. Science, 272, 398-401. Li, J., Biswas, M.G., Chao, A., Russell, D.W. and Chory, J. (1997) Conservation of function between mammalian and plant steroid 5alpha-reductases. Proc Natl Acad Sci U S A, 94, 3554-3559. Li, J. and Chory, J. (1997) A putative leucine-rich repeat receptor kinase involved in brassinosteroid signal transduction. Cell, 90, 929-938. Li, J., Nam, K.H., Vafeados, D. and Chory, J. (2001) BIN2, a new brassinosteroid-insensitive locus in Arabidopsis. Plant Physiol, 127, 14-22. Li, J., Wen, J., Lease, K.A., Doke, J.T., Tax, F.E. and Walker, J.C. (2002) BAK1, an Arabidopsis LRR receptor-like protein kinase, interacts with BRI1 and modulates brassinosteroid signaling. Cell, 110, 213-222. Li, H., Johnson, P., Stepanova, A., Alonso, J.M. and Ecker, J.R. (2004) Convergence of signaling pathways in the control of differential cell growth in Arabidopsis. Dev Cell, 7, 193-204. Li, L. and Deng, X.W. (2005) It runs in the family: regulation of brassinosteroid signaling by the BZR1-BES1 class of transcription factors. Trends Plant Sci, 10, 266-268. Li, L., Yu, X., Thompson, A., Guo, M., Yoshida, S., Asami, T., Chory, J. and Yin, Y. (2008) Arabidopsis MYB30 is a Direct Target of BES1 and Cooperates with BES1 to Regulate Brassinosteroid-Induced Gene Expression. Plant J. Lindsey, K., Pullen, M.L. and Topping, J.F. (2003) Importance of plant sterols in pattern formation and hormone signalling. Trends Plant Sci, 8, 521-525. Malikova, J., Swaczynova, J., Kolar, Z. and Strnad, M. (2008) Anticancer and antiproliferative activity of natural brassinosteroids. Phytochemistry, 69, 418-426. Mandava, N.B. (1988) Plant growth-promoting brassinosteriods. Annu. Rev. Plant Physiol. Plant Mol. Biol 39, 23-52. Mitchell, J.W., Mandava, N., Worley, J.F., Plimmer, J.R. and Smith, M.V. (1970) Brassins--a new family of plant hormones from rape pollen. Nature, 225, 1065-1066. Montoya, T., Nomura, T., Farrar, K., Kaneta, T., Yokota, T. and Bishop, G.J. (2002) Cloning the tomato curl3 gene highlights the putative dual role of the leucine-rich repeat receptor kinase tBRI1/SR160 in plant steroid hormone and peptide hormone signaling. Plant Cell, 14, 3163-3176. Montoya, T., Nomura, T., Yokota, T., Farrar, K., Harrison, K., Jones, J.D., Kaneta, T., Kamiya, Y., Szekeres, M. and Bishop, G.J. (2005) Patterns of Dwarf expression and brassinosteroid accumulation in tomato reveal the importance of brassinosteroid synthesis during fruit development. Plant J, 42, 262-269. Mora-Garcia, S., Vert, G., Yin, Y., Cano-Delgado, A., Cheong, H. and Chory, J. (2004) Nuclear protein phosphatases with Kelch-repeat domains modulate the response to brassinosteroids in Arabidopsis. Genes Dev, 18, 448-460. Mussig, C., Shin, G.H. and Altmann, T. (2003) Brassinosteroids promote root growth in Arabidopsis. Plant Physiol, 133, 1261-1271. 30 General introduction: Brassinosteroid biosynthesis and signalling in plants Nagata, N., Asami, T. and Yoshida, S. (2001) Brassinazole, an inhibitor of brassinosteroid biosynthesis, inhibits development of secondary xylem in cress plants (Lepidium sativum). Plant Cell Physiol, 42, 1006-1011. Nakajima, N., Shida, A. and Toyoma, S. (1996) Effect of brassinosteroid on cell division and colony formation of Chinese cabbage mesophyll protoplasts. Jpn J Crop Sci, 65, 114-118. Nakamura, A., Higuchi, K., Goda, H., Fujiwara, M.T., Sawa, S., Koshiba, T., Shimada, Y. and Yoshida, S. (2003) Brassinolide induces IAA5, IAA19, and DR5, a synthetic auxin response element in Arabidopsis, implying a cross talk point of brassinosteroid and auxin signaling. Plant Physiol, 133, 1843-1853. Nakamura, M., Satoh, T., Tanaka, S., Mochizuki, N., Yokota, T. and Nagatani, A. (2005) Activation of the cytochrome P450 gene, CYP72C1, reduces the levels of active brassinosteroids in vivo. J Exp Bot, 56, 833-840. Nakashita, H., Yasuda, M., Nitta, T., Asami, T., Fujioka, S., Arai, Y., Sekimata, K., Takatsuto, S., Yamaguchi, I. and Yoshida, S. (2003) Brassinosteroid functions in a broad range of disease resistance in tobacco and rice. Plant J, 33, 887-898. Nam, K.H. and Li, J. (2002) BRI1/BAK1, a receptor kinase pair mediating brassinosteroid signaling. Cell, 110, 203-212. Nemhauser, J.L., Mockler, T.C. and Chory, J. (2004) Interdependency of brassinosteroid and auxin signaling in Arabidopsis. PLoS Biol, 2, E258. Nemhauser, J.L., Hong, F. and Chory, J. (2006) Different plant hormones regulate similar processes through largely nonoverlapping transcriptional responses. Cell, 126, 467-475. Nishikawa, N., Toyama, S., Shida, A. and Fatatsuya, F. (1994) The uptake and transport of 14C-labeled epibrassinolide in intact seedlings of cucumber and wheat. Journal of Plant Research, 107. Noguchi, T., Fujioka, S., Choe, S., Takatsuto, S., Yoshida, S., Yuan, H., Feldmann, K.A. and Tax, F.E. (1999) Brassinosteroid-insensitive dwarf mutants of Arabidopsis accumulate brassinosteroids. Plant Physiol, 121, 743-752. Nomura, T., Sato, T., Bishop, G.J., Kamiya, Y., Takatsuto, S. and Yokota, T. (2001) Accumulation of 6deoxocathasterone and 6-deoxocastasterone in Arabidopsis, pea and tomato is suggestive of common rate-limiting steps in brassinosteroid biosynthesis. Phytochemistry, 57, 171-178. Nomura, T., Bishop, G.J., Kaneta, T., Reid, J.B., Chory, J. and Yokota, T. (2003) The LKA gene is a BRASSINOSTEROID INSENSITIVE 1 homolog of pea. Plant J, 36, 291-300. Oh, M.H. and Clouse, S.D. (1998) Brassinolide affects the rate of cell division in isolated leaf protoplasts of Petunia hybrida. Plant Cell Reports, 17, 921-924. Ohnishi, T., Szatmari, A.M., Watanabe, B., Fujita, S., Bancos, S., Koncz, C., Lafos, M., Shibata, K., Yokota, T., Sakata, K., Szekeres, M. and Mizutani, M. (2006) C-23 hydroxylation by Arabidopsis CYP90C1 and CYP90D1 reveals a novel shortcut in brassinosteroid biosynthesis. Plant Cell, 18, 3275-3288. Pay, A., Jonak, C., Bogre, L., Meskiene, I., Mairinger, T., Szalay, A., Heberle-Bors, E. and Hirt, H. (1993) The MsK family of alfalfa protein kinase genes encodes homologues of shaggy/glycogen synthase kinase-3 and shows differential expression patterns in plant organs and development. Plant J, 3, 847-856. Peng, P., Zhao, J., Zhu, Y., Asami, T. and Li, J. (2010) A direct docking mechanism for a plant GSK3-like kinase to phosphorylate its substrates. J Biol Chem, 285, 24646-24653. Piao, H.L., Pih, K.T., Lim, J.H., Kang, S.G., Jin, J.B., Kim, S.H. and Hwang, I. (1999) An Arabidopsis GSK3/shaggy-like gene that complements yeast salt stress-sensitive mutants is induced by NaCl and abscisic acid. Plant Physiol, 119, 1527-1534. 31 Chapter 1 Piao, H.L., Lim, J.H., Kim, S.J., Cheong, G.W. and Hwang, I. (2001) Constitutive over-expression of AtGSK1 induces NaCl stress responses in the absence of NaCl stress and results in enhanced NaCl tolerance in Arabidopsis. Plant J, 27, 305-314. Reinhardt, B., Hanggi, E., Muller, S., Bauch, M., Wyrzykowska, J., Kerstetter, R., Poethig, S. and Fleming, A.J. (2007) Restoration of DWF4 expression to the leaf margin of a dwf4 mutant is sufficient to restore leaf shape but not size: the role of the margin in leaf development. Plant J, 52, 1094-1104. Roth, U., Friebe, A. and Schnabl, H. (2000) Resistance induction in plants by a brassinosteroid-containing extract of Lychnis viscaria L. Zeitschrift Naturforschung 55c 552-559. Rozhon, W., Mayerhofer, J., Petutschnig, E., Fujioka, S. and Jonak, C. (2010) ASKtheta, a group-III Arabidopsis GSK3, functions in the brassinosteroid signalling pathway. Plant J, 62, 215-223. Russinova, E., Borst, J.W., Kwaaitaal, M., Cano-Delgado, A., Yin, Y., Chory, J. and de Vries, S.C. (2004) Heterodimerization and endocytosis of Arabidopsis brassinosteroid receptors BRI1 and AtSERK3 (BAK1). Plant Cell, 16, 3216-3229. Savaldi-Goldstein, S., Peto, C. and Chory, J. (2007) The epidermis both drives and restricts plant shoot growth. Nature, 446, 199-202. Schaller, F., Biesgen, C., Mussig, C., Altmann, T. and Weiler, E.W. (2000) 12-Oxophytodienoate reductase 3 (OPR3) is the isoenzyme involved in jasmonate biosynthesis. Planta, 210, 979-984. Schrick, K., Fujioka, S., Takatsuto, S., Stierhof, Y.D., Stransky, H., Yoshida, S. and Jurgens, G. (2004) A link between sterol biosynthesis, the cell wall, and cellulose in Arabidopsis. Plant J, 38, 227-243. Shimada, Y., Fujioka, S., Miyauchi, N., Kushiro, M., Takatsuto, S., Nomura, T., Yokota, T., Kamiya, Y., Bishop, G.J. and Yoshida, S. (2001) Brassinosteroid-6-oxidases from Arabidopsis and tomato catalyze multiple C-6 oxidations in brassinosteroid biosynthesis. Plant Physiol, 126, 770-779. Shimada, A., Ueguchi-Tanaka, M., Sakamoto, T., Fujioka, S., Takatsuto, S., Yoshida, S., Sazuka, T., Ashikari, M. and Matsuoka, M. (2006) The rice SPINDLY gene functions as a negative regulator of gibberellin signaling by controlling the suppressive function of the DELLA protein, SLR1, and modulating brassinosteroid synthesis. Plant J, 48, 390-402. Song, L., Shi, Q.M., Yang, X.H., Xu, Z.H. and Xue, H.W. (2009) Membrane steroid-binding protein 1 (MSBP1) negatively regulates brassinosteroid signaling by enhancing the endocytosis of BAK1. Cell Res, 19, 864-876. Steber, C.M. and McCourt, P. (2001) A role for brassinosteroids in germination in Arabidopsis. Plant Physiol, 125, 763-769. Sun, Y., Fokar, M., Asami, T., Yoshida, S. and Allen, R.D. (2004) Characterization of the brassinosteroid insensitive 1 genes of cotton. Plant Mol Biol, 54, 221-232. Suzuki, H., Fujioka, S., Takatsuto, S., Yokota, T., Murofushi, N. and Sakurai, A. (1993) Biosynthesis of brassinolide from castasterone in cultured cells of Catharanthus roseus. Journal of Plant Growth Regulation, 12, 101-106. Suzuki, H., Fujioka, S., Takatsuto, S., Yokota, T., Murofushi, N. and Sakurai, A. (1994) Biosynthesis of brassinolide from teasterone via typhasterol and castasterone in cultured cells of Catharanthus roseus. Journal of Plant Growth Regulation, 13, 21-26. Symons, G.M. and Reid, J.B. (2004) Brassinosteroids do not undergo long-distance transport in pea. Implications for the regulation of endogenous brassinosteroid levels. Plant Physiol, 135, 2196-2206. Symons, G.M., Ross, J.J., Jager, C.E. and Reid, J.B. (2008) Brassinosteroid transport. J Exp Bot, 59, 17-24. 32 General introduction: Brassinosteroid biosynthesis and signalling in plants Szekeres, M., Nemeth, K., Koncz-Kalman, Z., Mathur, J., Kauschmann, A., Altmann, T., Redei, G.P., Nagy, F., Schell, J. and Koncz, C. (1996) Brassinosteroids rescue the deficiency of CYP90, a cytochrome P450, controlling cell elongation and de-etiolation in Arabidopsis. Cell, 85, 171-182. Tanabe, S., Ashikari, M., Fujioka, S., Takatsuto, S., Yoshida, S., Yano, M., Yoshimura, A., Kitano, H., Matsuoka, M., Fujisawa, Y., Kato, H. and Iwasaki, Y. (2005) A novel cytochrome P450 is implicated in brassinosteroid biosynthesis via the characterization of a rice dwarf mutant, dwarf11, with reduced seed length. Plant Cell, 17, 776-790. Tang, W., Kim, T.W., Oses-Prieto, J.A., Sun, Y., Deng, Z., Zhu, S., Wang, R., Burlingame, A.L. and Wang, Z.Y. (2008) BSKs mediate signal transduction from the receptor kinase BRI1 in Arabidopsis. Science, 321, 557560. Tichtinsky, G., Tavares, R., Takvorian, A., Schwebel-Dugue, N., Twell, D. and Kreis, M. (1998) An evolutionary conserved group of plant GSK-3/shaggy-like protein kinase genes preferentially expressed in developing pollen. Biochim Biophys Acta, 1442, 261-273. Turk, E.M., Fujioka, S., Seto, H., Shimada, Y., Takatsuto, S., Yoshida, S., Denzel, M.A., Torres, Q.I. and Neff, M.M. (2003) CYP72B1 inactivates brassinosteroid hormones: an intersection between photomorphogenesis and plant steroid signal transduction. Plant Physiol, 133, 1643-1653. Vert, G. and Chory, J. (2006) Downstream nuclear events in brassinosteroid signalling. Nature, 441, 96-100. Vert, G., Walcher, C.L., Chory, J. and Nemhauser, J.L. (2008) Integration of auxin and brassinosteroid pathways by Auxin Response Factor 2. Proc Natl Acad Sci U S A, 105, 9829-9834. Wang, Z.Y., Nakano, T., Gendron, J., He, J., Chen, M., Vafeados, D., Yang, Y., Fujioka, S., Yoshida, S., Asami, T. and Chory, J. (2002) Nuclear-localized BZR1 mediates brassinosteroid-induced growth and feedback suppression of brassinosteroid biosynthesis. Dev Cell, 2, 505-513. Wang, X., Grammatikakis, N., Siganou, A., Stevenson, M.A. and Calderwood, S.K. (2004) Interactions between extracellular signal-regulated protein kinase 1, 14-3-3epsilon, and heat shock factor 1 during stress. J Biol Chem, 279, 49460-49469. Wang, X., Goshe, M.B., Soderblom, E.J., Phinney, B.S., Kuchar, J.A., Li, J., Asami, T., Yoshida, S., Huber, S.C. and Clouse, S.D. (2005) Identification and functional analysis of in vivo phosphorylation sites of the Arabidopsis BRASSINOSTEROID-INSENSITIVE1 receptor kinase. Plant Cell, 17, 1685-1703. Wang, X., Li, X., Meisenhelder, J., Hunter, T., Yoshida, S., Asami, T. and Chory, J. (2005) Autoregulation and homodimerization are involved in the activation of the plant steroid receptor BRI1. Dev Cell, 8, 855-865. Wang, X. and Chory, J. (2006) Brassinosteroids regulate dissociation of BKI1, a negative regulator of BRI1 signaling, from the plasma membrane. Science, 313, 1118-1122. Wang, L., Wang, Z., Xu, Y., Joo, S.H., Kim, S.K., Xue, Z., Xu, Z., Wang, Z. and Chong, K. (2009) OsGSR1 is involved in crosstalk between gibberellins and brassinosteroids in rice. Plant J, 57, 498-510. Woyengo, T.A., Ramprasath, V.R. and Jones, P.J. (2009) Anticancer effects of phytosterols. Eur J Clin Nutr, 63, 813-820. Xu, W., Campbell, P., Vargheese, A.K. and Braam, J. (1996) The Arabidopsis XET-related gene family: environmental and hormonal regulation of expression. Plant J, 9, 879-889. Yamamoto, R., Fujioka, S., Demura, T., Takatsuto, S., Yoshida, S. and Fukuda, H. (2001) Brassinosteroid levels increase drastically prior to morphogenesis of tracheary elements. Plant Physiol, 125, 556-563. Yamamuro, C., Ihara, Y., Wu, X., Noguchi, T., Fujioka, S., Takatsuto, S., Ashikari, M., Kitano, H. and Matsuoka, M. (2000) Loss of function of a rice brassinosteroid insensitive1 homolog prevents internode elongation and bending of the lamina joint. Plant Cell, 12, 1591-1606. 33 Chapter 1 Yan, Z., Zhao, J., Peng, P., Chihara, R.K. and Li, J. (2009) BIN2 Functions Redundantly with Other Arabidopsis GSK3-Like Kinases to Regulate Brassinosteroid Signaling. Plant Physiol, 150, 710-721. Yin, Y., Wang, Z.Y., Mora-Garcia, S., Li, J., Yoshida, S., Asami, T. and Chory, J. (2002) BES1 accumulates in the nucleus in response to brassinosteroids to regulate gene expression and promote stem elongation. Cell, 109, 181-191. Yin, Y., Vafeados, D., Tao, Y., Yoshida, S., Asami, T. and Chory, J. (2005) A new class of transcription factors mediates brassinosteroid-regulated gene expression in Arabidopsis. Cell, 120, 249-259. Yokota, T., Higuchi, K., Kosaka, Y. and Takahashi, N. (1992) Transport and metabolism of brassinosteroids in rice. In: Karssen CM, van Loon LC, Vreugdenhil D, eds. Progress in plant growth regulation, 298-305. Yoo, M.J., Albert, V.A., Soltis, P.S. and Soltis, D.E. (2006) Phylogenetic diversification of glycogen synthase kinase 3/SHAGGY-like kinase genes in plants. BMC Plant Biol, 6, 3. Yu, X., Li, L., Guo, M., Chory, J. and Yin, Y. (2008) Modulation of brassinosteroid-regulated gene expression by Jumonji domain-containing proteins ELF6 and REF6 in Arabidopsis. Proc Natl Acad Sci U S A, 105, 76187623. Zeng, H., Tang, Q. and Hua, X. (2009) Arabidopsis Brassinosteroid Mutants det2-1 and bin2-1 Display Altered Salt Tolerance. J Plant Growth Regul. Zhang, S., Cai, Z. and Wang, X. (2009) The primary signaling outputs of brassinosteroids are regulated by abscisic acid signaling. Proc Natl Acad Sci U S A, 106, 4543-4548. Zurek, D.M. and Clouse, S.D. (1994) Molecular cloning and characterization of a brassinosteroid-regulated gene from elongating soybean (Glycine max L.) epicotyls. Plant Physiol, 104, 161-170. 34 Chapter 2 Group II GSK3/SHAGGY-like kinases act redundantly in the Arabidopsis brassinosteroid signalling pathway Nathalie Verhoef, Annegret Roß, Ronald Koes and Erik Souer Group II GSK3/SHAGGY-like kinases act redundantly in the brassinosteroid signalling pathway Abstract Brassinosteroids (BR) are steroid hormones that are essential for plant growth and development. In Arabidopsis, BRs bind to the leucine-rich repeat (LRR) receptor serine/threonine kinase BRASSINOSTEROID INSENSITIVE 1 (BRI1), which initiates a phosphorylation cascade leading to a final BR response. The group II GSK3-like kinase BR-INSENSITIVE 2 (BIN2) negatively regulates the BR signalling pathway by phosphorylating the transcription factors BRI1-EMS-SUPPRESSOR 1 (BES1) and BRASSINAZOLE RESISTANT 1 (BZR1). In this study we characterized two other group II GSK3-like kinases from Arabidopsis, named Glycogen Synthase Kinase 1 (GSK1) and Arabidopsis GSK3-like kinase ζ (ASKζ). Mutant analysis and their interaction with members of the BES1/BZR1 family allude to a redundant role of group II GSK3-like kinases in the BR signalling pathway. In contrast to prior research we could not detect a role for GSK1 in tolerance against Na+ stress, neither for the other group II members. Introduction GLYCOGEN SYNTHASE KINASE-3 (GSK3) is a highly conserved serine/threonine protein kinase. It was originally identified as a key regulator in glycogen synthesis in mammals (Embi et al., 1980), but subsequent work showed its involvement in an array of processes such as cell proliferation, cell fate specification, differentiation and apoptosis (Frame and Cohen, 2001). Purification and molecular cloning revealed two isomers, GSK3α and GSK3β, that have 97% homology within their catalytic kinase domains. It was generally assumed that GSK3α and GSK3β have the same biochemical and physiological functions, but with the year’s closer examination invalidate this statement (Ruel et al., 1993; Hoeflich et al., 2000). Although the role of GSK3 in mammals and its homolog SHAGGY in Drosophila melanogaster (Siegfried et al., 1992) has been extensively studied, little is known about the specific functions of GSK3-like kinases in plants. In higher plant species such as Arabidopsis thaliana (Dornelas et al., 1999; Li et al., 2001), Medicago sativa (Pay et al., 1993), Oryza sativa (Koh et al., 2007), Nicotiana tabacum (Einzenberger et al., 1995) and Petunia hybrida (Decroocq-Ferrant et al., 1995) several GSK3-like kinases have been identified. In Arabidopsis ten GSK3-like kinases, also known as Arabidopsis GSK3-like kinases (ASKs), have been identified. These ten ASK genes can be classified into four subgroups based on phylogenetic analyses of amino acid- and cDNA sequences (Dornelas et al., 1998; Dornelas et al., 1999; Jonak and Hirt, 2002). The protein sequences of ASKs 37 Chapter 2 are highly conserved throughout the kinase domain, in contrast to the N- and C- terminal regions that are highly variable. This variability might point to specific functions for each of the GSK3-like kinases in a variety of plant processes. For some of the GSK3-like kinases specific roles were assigned. ASK11 and ASK12 have been reported to play a role in flower development (Dornelas et al., 2000), while GSK1 seems to play a role in Na+ signalling (Piao et al., 1999; 2001). The best-studied GSK3-like kinase in plants is BRINSENSITIVE2 (BIN2 also known as ASKη/ UCU1/DWF12/AtSK2-1), a member of subgroup II, which is involved in BR signalling in Arabidopsis (Li et al., 2001). Brassinosteroids (BRs) are steroidal hormones that are important for growth and development in plants. Deficiency in BR biosynthesis or signalling causes dramatic growth defects including dwarfism, reduced apical dominance, male sterility, delayed flowering and senescence, and photomorphogenesis in the dark (Li et al., 1996; Szekeres et al., 1996; Azpiroz et al., 1998). In Arabidopsis, BRs are perceived at the plasma membrane by BRASSINOSTEROID INSENSITIVE 1 (BRI1), which encodes a leucine-rich repeat (LRR) receptor serine/threonine kinase. As a result, BRI1 can dimerize and transphosphorylate the co-receptor BRI1-ASSOCIATED RECEPTOR KINASE 1 (BAK1) and dissociate from BRI1 KINASE INHIBITOR 1 (BKI1), a repressor of BRI1 (Li et al., 2002; Nam and Li, 2002; Russinova et al., 2004; Gampala et al., 2007; Wang et al., 2008a). Recently, Kim et al. (2009) showed that BRI1 mediated phosphorylation of the BRSIGNALING KINASE 1 (BSK1) results in interaction and presumably activation of the plant-specific serine/threonine protein phosphatase BRI1 SUPPRESSOR 1 (BSU1). BSU1 in turn inactivates BIN2 by dephosphorylation at pTyr200. BIN2 negatively regulates the BR signalling pathway by phosphorylating closely related members of a plant-specific transcription factor family BES1/BZR1. Phosphorylated BES1 and BZR1 is inactivated by protein degradation, cytoplasmic retention and/or inhibition of their DNA-binding activities (He et al., 2002; Yin et al., 2002; Vert and Chory, 2006; Gampala et al., 2007; Ryu et al., 2007; Peng et al., 2010). The last decade a number of papers was published on the mode of action of BIN2, but the other ASKs fell into oblivion. In this study we have analysed the group II ASKs from Arabidopsis. Mutant analysis and interaction with the BES1 and BZR1 transcription factors point to a redundant role for these kinases in BR signalling, results that were confirmed by other scientists as well (Vert and Chory, 2006; Kim et al., 2009; Yan et al., 2009). Furthermore, in contrast to previous reports (Piao et al., 1999; 2001) we could not establish a role for these kinases in tolerance against Na+ stress. 38 Group II GSK3/SHAGGY-like kinases act redundantly in the brassinosteroid signalling pathway Results Redundancy between GSK1, ASKζ and BIN2 Within this project we intended to study the role of the GSK3-like kinase GSK1 in salt tolerance. Previous work by Piao et al. (1999; 2001) implied a role for GSK1 in salt tolerance as over-expression led to an increased tolerance in Arabidopsis. Phylogenetic analysis placed GSK1 in subgroup II together with ASKζ and BIN2 (Jonak and Hirt, 2002). To unravel the role(s) of GSK1 and its closest homolog ASKζ, we identified lossof-function alleles for both genes (Fig. 1a and b). Single loss-of-function mutants in GSK1 and ASKζ did not show any visible phenotypes neither did the double mutant. However, a triple insertion mutant for all group II ASKs, GSK1, ASKζ and BIN2, showed aberrant growth of the leaves and long bending petioles (Fig. 1c). Since a phenotype is only apparent in a insertion mutant of all group II ASKs, we assume that BIN2, GSK1 and ASKζ act redundantly. BIN2 plays an important role in the brassinosteroid (BR) signalling pathway and thus GSK1 and ASKζ might function in this pathway too. Asami et al. (2000) reported the finding of a specific BR biosynthetic inhibitor brassinazole (brz) and we would expect that single or double group II ask loss-of-function mutants would be sensitive to brz whereas the triple mutant should show resistance to brz. Indeed, Arabidopsis wild type, single and double mutant seedlings treated with brz showed a strong reduction in hypocotyl length (Fig. 1d) and the thickness of the cell wall was strongly increased (Fig. 1e). Inhibition of BR biosynthesis by brz in the triple mutant background did hardly affect hypocotyl elongation. However, brz still had some reducing effect on the longitudinal growth of the triple insertion mutant, suggesting that other GSK3s still partly inhibited the BR signalling pathway. Temporarily switching on genes with help of the estrogen receptor-based inducible XVE system was shown to be an effective molecular tool (Zuo et al., 2000; Papdi et al., 2008; Pre et al., 2008). As the triple insertion mutant displayed strong resistance against brz (Fig. 1d), we tested whether induction of either one of the transgenes in a triple mutant background would re-establish sensitivity to brz. Group II ASK transcripts were clearly detected after induction by estradiol, although gene expression varied between independent lines (Fig. 1f). In the non-induced transgenic lines no transcript was observed, indicating that the XVE system is reliable to switch on ASKs. In contrast to our expectations, not all transgenic lines of pER8:BIN2dom, pER8:BIN2 and pER8:GSK1 were sensitive to brz after induction (Fig. 1g). In fact none of the transgenic pER8:ASKζ lines analyzed in figure 1f were sensitive to brz and for pER8:BIN2 only one line was found (Fig. 1g). 39 Chapter 2 A D BIN2 Wt Col gsk1 askζ gsk1/askζ bin2/gsk1/askζ 200 bp ATG STOP DMSO ASKζ ATG STOP GSK1 1 μM brz STOP W t gs Col as k1 k gs ζ k bi 1/a n2 sk W /gs ζ t k bi Ws 1/a n2 sk ζ ATG B E + - 1 μM brz GSK1 Wt Col ASKζ BIN2 bin2/gsk1/askζ ACTIN C F pER8:BIN2L8137 pER8:BIN2domL8138 Wt Col gsk1 askζ gsk1/askζ 14 17 18 9 19 14 21 19 - +- +- +- + - +- +- +- + pER8:ASKζL8140 pER8:GSK1L8139 24 20 34 37 39 - + - +- +- + - + 5 6 9 11 12 - + - + - + - + - + estradiol ASK ACTIN Wt Ws bin2/gsk1/askζ bin2 bin2/gsk1/askζ background G Wt Col + - + + bin2/gsk1/askζ pER8:BIN2doma + - + - + + + + pER8:BIN2b + - + + pER8:GSK1c + - + + estradiol 1 µM brz Figure 1. Analysis of insertion mutants in group II ASKs A) Schematic representation showing the T-DNA integration site (triangle) in BIN2 (At4G18710), GSK1 (At1G06390) and ASKζ (At2G30980). Solid bars indicate exons, thin lines introns. B) RT-PCR analysis on total RNA isolated from leaves showing expression of BIN2, GSK1 and ASKζ in various mutant backgrounds. ACTIN was used as control. C) Phenotype of bin2, gsk1 and askζ single loss-of-function mutants, the gsk1/askζ double mutant and the group II ask triple insertion mutant grown under short-day conditions. D) The effect of brassinazole (brz) on cell elongation in hypocotyls of ask mutant and wild type seedlings. E) Light microscopy of stems of brassinazole treated and non-treated wild type and triple ask mutant seedlings. Bar = 100 µM. F) Independent transgenic lines were screened for their ability to express group II ASK transgenes in the presence or absence of the inducer estradiol. RT-PCR analysis on total RNA isolated from seedlings was performed to determine ASK expression (Upper row of panels). ACTIN was used as control (Lower row of panels). G) Restored brassinazole sensitivity in a subset of lines after induced expression of various group II ASK genes in mutant seedlings. a, L8138-9; b, L8137-14; c, L8139-13. 40 Group II GSK3/SHAGGY-like kinases act redundantly in the brassinosteroid signalling pathway A gain of function mutation in GSK1 and ASKζ results into a dwarfish phenotype In a screen to search for BR insensitive (semi)dwarf mutants, several semi-dominant mutants in BIN2 were identified (Li et al., 2001; Choe et al., 2002; Li and Nam, 2002; Perez-Perez et al., 2002). bin2-1 carries a missense mutation in the part encoding the kinase domain that results in the conversion of a glutamic acid (E) at amino acid 263 into a lysin (K), leading to a more active BIN2 kinase (Li and Nam, 2002). In order to test if this amino acid change also increases the activity of GSK1 or ASKζ, a similar mutation was introduced in GSK1 and ASKζ (Fig. 2a). These dominant constructs were expressed under control of a CaMV35S promoter in Arabidopsis. Indeed, some transformants expressing GSK1dom or ASKζdom have a dwarfish phenotype that is morphologically comparable to the previously characterized bin2-1 mutant (Fig. 2b). Furthermore, the level of transgene expression correlates with the severity of the phenotype (Fig. 2c). These results suggest that all three group II kinases can negatively regulate the BR signalling pathway. A B KpnI L G T P T R E E cttggtactccaactcgtgaagaa cttggtaccccaactcgcaaagaa L G T P T R K E a b c d e f Wt Col 35S GSK1dom 35S GSK1dom KpnI L G T P T R E E cttggtactccaactcgcgaagaa cttggtaccccaactcgcaaagaa L G T P T R K E C a b c g d e f Wt Col 35S ASKζdom 35S ASKζdom g g GSK1/ASKζ Wt Col ACTIN Figure 2. Gain-of-function mutants of GSK1 and ASKζ have a dwarfish phenotype. A) Sequence comparison of wild type and dominant constructs of GSK1 and ASKζ. The introduced KpnI site used for cloning (see Material and Methods) as well as the nucleotide(s) resulting in the desired amino acid change, glutamic acid (E) into lysin (K), are indicated in green. B) Over-expression of GSK1dom and ASKζdom under control of a 35S promoter in Arabidopsis leads to dwarfism. C) RT-PCR analysis on total RNA isolated from leaves of the plants shown in B (a-g). ACTIN was used as control. 41 Chapter 2 GSK1 and ASKζ bind to members of the BZR1/BES1 family Our data suggest that GSK1 and ASKζ function redundantly with BIN2 as negative regulators of the BR signalling pathway. To further explore the function(s) of GSK1, a yeast two-hybrid approach was used to find interacting partners. A full-length GSK1 protein was used as bait to screen a two-hybrid cDNA library made from mature Arabidopsis leaves and roots. Only one interacting clone was found and interestingly this clone contained the C-terminal part of the BEH2 protein (Fig. 3a). BEH2 is a homolog of BES1 and BZR1, two plant specific transcription factors that have been shown to interact with BIN2 (Zhao et al., 2002). As the BES1/BZR1 family only consists of six genes, twohybrid constructs of the complete family were generated and screened for interaction with BIN2, GSK1 and ASKζ. Like BIN2, also GSK1 and ASKζ strongly interact with BZR1 and BES1 (Fig. 3b). Noteworthy, in contrast to GSK1 and BIN2, ASKζ seem to (weakly) B A MAAGGGGGGG RRRRAITAKI VLKALCLEAG ISGTPTNFST HGSPVSSSFP IASSIPANLP GSKRKLTSEQ SSPTRRAGHQ RWINFQSTAP DWGMSGMNGR EVGVEDLELT GSSSGRTPTW YSGLRAQGNY WIVEDDGTTY NSSIQPSPQS SPSRYDGNPS PLRISNSAPV LPNGGSLHVL TPPTIPECDE TSPTFNLVQQ GAEFEFENGT LGGTKARCX KERENNKKRE KLPKHCDNNE RKGFKPPASD SAFPSPAPSY SYLLLPFLHN TPPLSSPTSR RHPLFAISAP SEEDSIEDSG TSMAIDMKRS VKPWEGEMIH G a G l4B al D G 4BD al G 4BD AS al - K 4B G ζ S G D-B K1 al G 4B IN2 al D G 4BD al G 4BD AS al - K 4B G ζ D SK -B 1 G IN al 2 G 4B al D 4 G BD al G 4BD AS al - K 4B G ζ D SK -B 1 IN G al 2 G 4B al D 4 G BD al G 4BD AS al - K 4B G ζ D SK -B 1 IN 2 interact with all members of the BZR1/BES1 family, but most strongly to BEH2, BES1 and BZR1. Gal4 AD-BZR1 Gal4 AD-BES1 Gal4 AD-BEH2 Gal4 AD-BEH1 Gal4 AD-BEH3 Gal4 AD-BEH4 Gal4 AD-δBEH2 Gal4 AD -LT -LTH + 3AT -LTHA -LT X-gal Figure 3. GSK1 and ASKζ bind to members of the BES1/BZR1 transcription factor family. A) The complete amino acid sequence of BEH2 is shown. The underlined part of the sequence indicates the part of BEH2 fished from the yeast two-hybrid library using full-length GSK1 as bait. B) Yeast two-hybrid interaction of members of the BES1/BZR1 family with GSK1, ASKζ and BIN2. For BEH2, two clones were used, the full-length sequence (BEH2) and the part of the sequence that was picked up from the initial library screening with GSK1 (δBEH2, see panel A). Interactions were measured by growth on medium lacking histidine with addition of 3-amino-1,2,4triazole (-LTH + 3-AT), lacking histidine and adenine (-LTHA) and by blue coloring using 5-bromo-4-chloro-3indolyl-β-D-galactoside (-LT X-gal). The group II ASKs seem not involved in NaCl stress signalling Previously it has been suggested that GSK1 plays an important role in NaCl signalling (Piao et al., 1999; 2001). Over-expression of GSK1 enhanced the resistance to NaCl stress and induced the expression of salt responsive genes like RD29 and AtCP1. We therefore 42 Group II GSK3/SHAGGY-like kinases act redundantly in the brassinosteroid signalling pathway expected that the gsk1 mutant would be more sensitive to salt stress and considered this might be the same for mutants in the other group II members, ASKζ and BIN2. In order to check whether group II ASKs play a role in NaCl signalling, three-week-old wild type and bin2/gsk1/askζ plants were exposed to a high NaCl concentration (Fig. 4a). Whereas the leaves of untreated wild type and bin2/gsk1/askζ plants remained green, NaCl stress caused yellowing of the rosette leaves and growth was retarded. However no clear difference in response to NaCl between wild type and the triple bin2/gsk1/askζ plants was observed. Since this experiment was performed on soil, the actual exposure to NaCl is uncertain. A more accurate experiment to check the influence of NaCl was performed by growing wild type, gsk1 and bin2/gsk1/askζ plants on hydroponics for three weeks, followed by exposure to various NaCl concentrations (Fig. 4b). As a control we used a mutant in the HKT1 gene (Rus et al., 2001). HKT1 encodes a Na+ transporter, which is involved in the recirculation of Na+ from shoots to roots (Berthomieu et al., 2003). As predicted, hkt1 is hypersensitive to NaCl treatment. In contrast, wild type Arabidopsis, gsk1 and bin2/gsk1/askζ plants did not show any morphological change upon a mild NaCl treatment. A 0 mM C 200 mM 4,0 Wt Col gsk1 mutant bin2/gsk1/askζ 3,5 bin2/gsk1/askζ B Wt Col gsk1 bin2/gsk1/askζ rootgrowth in cm Wt Col 3,0 2,5 2,0 1,5 1,0 hkt1 0,5 0 mM 0,0 0mM 25mM NaCL 25 mM 50 mM Figure 4. GSK1 and the other two members of the group II ASK family seem not involved in NaCl stress signalling. A) Phenotypic comparison of three-week-old soil grown wild type and bin2/gsk1/askζ plants treated with 0 or 200 mM NaCl for a period of 10 days. B) NaCl treatment of wild type, gsk1, bin2/gsk1/askζ and hkt1 plantlets grown on hydroponics under short-day conditions. C) Root growth measurements of NaCl treated wild type, gsk1 and bin2/gsk1/askζ plantlets on hydroponics under short-day conditions. Error-bars indicate the standard error. 43 Chapter 2 Raising the NaCl concentrations affected plant growth, but no difference in phenotype was encountered between the tested mutants and wild type (results not shown). As these experiments are not very accurate, depending on visual inspection only, minor differences in sensitivity to NaCl would have been missed. Therefore, classical root bending assays (Wu et al., 1996) were performed to assess the NaCl tolerance of wild type and bin2, gsk1, askζ, gsk1/askζ, bin2/gsk1/askζ mutant seedlings, but no obvious difference in root length inhibition was observed (results not shown). Next a sensitive quantitative root assay experiment was set up in which three week old wild type, gsk1 and bin2/gsk1/askζ plants were grown on hydroponics, treated with 25 mM NaCl for 7 days, which incompletely inhibited root growth, and root lengths were measured (Fig. 4c). The root growth of wild type, gsk1 and bin2/gsk1/askζ was greatly reduced upon salt treatment, but no significant difference in response to NaCl was encountered between wild type and the tested mutants. Neither the single and double mutant seedlings of bin2, askζ and gsk1/askζ responded differently when compared to wild type (data not shown). The fact that we consistently found no difference in response to NaCl between wild type and the ask mutants questioned the involvement of GSK1 and other Group II ASKs in salt stress signalling. Discussion The Arabidopsis genome contains ten GSK3-like kinases. In phylogenetic trees GSK1 clusters together with ASKζ and BIN2 to form the group II GSK3-like kinases (Jonak and Hirt, 2002). BIN2 is well known as a negative regulator of BR signalling (Li et al., 2001). We analysed loss-of-function alleles of all three genes, but in single mutants no visible phenotype was detected. However, an insertion mutation of all group II ASKs displayed a peculiar phenotype with abnormal stretched bending leaves and long petioles. Insensitivity of the triple mutant to the BR biosynthesis inhibitor brz indicates that this phenotype is due to constitutive BR sensing. We also transformed estradiol inducible versions (Zuo et al., 2000) of group II GSKs into the mutant background. Although most transformants nicely respond to the estradiol treatment with expression of the transgene, only a subset of the transformants gained sensitivity to brz treatment. It is unclear at the moment where this variability comes from. Our initial goal was to use these transformants to study downstream effects of specific group II GSK3-like kinase genes. A comparison by micro-array between GSK1, ASKζ and BIN2 induced in the triple mutant background might reveal gene-specific down-stream events. 44 Group II GSK3/SHAGGY-like kinases act redundantly in the brassinosteroid signalling pathway Gain of function mutations in GSK1 and ASKζ resulted in a dwarf phenotype similar to the previously described semi-dominant bin2-1 mutant (Li et al., 2001). Simultaneously, similar results were obtained by Vert and Chory (2006) and Yan et al. (2009). In addition, like ΒΙΝ2, GSK1 and ASKζ interacted with members of the BES1/BZR1 transcription factor family. These data suggest that group II ASKs act redundantly as negative regulators of BR signalling. The triple insertion mutant turned out to be not completely insensitive to a treatment with brz (Fig. 1d). This suggests that other GSK3s may be involved in the BR signalling pathway. This idea is further supported by the observation that BES1 proteins are still phosphorylated in a triple insertion mutant for all group II GSK3-like kinases (Vert and Chory, 2006). Just recently Kim and co-workers (2009) demonstrated that not only group II ASKs bind and phosphorylate BZR1, but also members of group I (ASK11-13). Furthermore transgenic plants over-expressing ASK12 as well as a semi-dominant mutant displayed similar dwarf phenotypes as seen before with BIN2. Altogether this suggests that not only group II ASKs are involved in BR signalling, but at least also members of group I. This is further supported by De Rybel et al. (2009) who identified a small molecule called bikinin, that was shown to inhibit the activity of all group I and II ASK members and ASKθ from group III. Inhibition of all seven ASKs activated BR responses and furthermore a large overlap of downstream transcriptional regulation was observed between a bikinin and BL treatment. Recently Rozhon et al. (2010) showed that bikinin could rescue the growth phenotype of Arabidopsis seedlings with an enhanced ASKθ activity and inhibited ASKθ mediated phosphorylation of BEH2. A negative role for ASKθ in the BR signalling pathway was further confirmed by several phenotypic and molecular analyses that showed that ASKθ is negatively regulated by BRI1 and that it phosphorylates several members of the BES1/BZR1 family. The initial aim of this project was to study the role of the GSK3-like kinase GSK1 in tolerance towards salt. In contrast to previous research by Piao et al. (1999; 2001) we could not establish a role for GSK1 in tolerance against Na+ stress. Neither was this the case for the other group II GSK3-like kinases BIN2 and ASKζ. Piao et al. demonstrated that ectopic expression of GSK1 in Arabidopsis leads to expression of multiple stress-induced genes and salt tolerance was increased. Diametrically opposed, Koh et al. (2007) showed that a insert mutation of rice OsGSK1, an orthologue of Arabidopsis BIN2, is more tolerant to salt stress. Furthermore, NaCl significantly inhibited germination of BR deficient (det2-1) and signalling mutants (bri1-301) in Arabidopsis (Zhang et al., 2009). This was also observed when the repressor of the BRI1 receptor BKI1 was over-expressed. Plants over-expressing BRI1 on the other hand were more resistant to NaCl. Moreover, Zeng et al. (2009) showed 45 Chapter 2 that the semi-dominant mutant bin2-1 was very sensitive to salt. Therefore, all data point to a positive role for BRs in the tolerance to salt stress. It therefore is puzzling why Piao et al. observed increased Na+ tolerance upon inhibition of BR signalling. One reason could be that over-expression of GSK1 might have caused unforeseen sight effects. Given the general picture that increased BR signalling improved Na+ tolerance we expected that knocking out ASKs would result in a more salt resistant plant. However, even a triple insertion mutation of the group II GSK3-like kinases did not result in more salt resistant plants. It might be that also in salt stress signalling, additional redundancy among GSK3like kinases exists. In this respect it is noteworthy that in Medicago sativa a group IV GSK3-like kinase was identified that is involved in stress tolerance (Kempa et al., 2007). Analysis of higher order mutant combinations might shed light on the function of all GSK3-like kinases in Arabidopsis. Experimental procedures Plant material and growth Unless stated otherwise, Arabidopsis Columbia was used as wild type control for phenotype comparisons in RNA analyses, brassinazole experiments and in NaCl stress assays. Seeds of Arabidopsis Colombia, mutant and transgenic lines were sterilized in 10% bleach, washed three times with sterile water, dissolved in 0.1% agarose and sewed on 0.8% (w/v) agar plates containing 0.5X Murashige and Skoog (MS) salts, 1.5% sucrose at pH 5.7-5.9. After stratification for 2 days at 4 °C, seeds were germinated in a climate chamber at 22 °C under a 10-hr-light/14-hr-dark photoperiod. Later, seedlings were transferred to soil or grown in hydroculture in a half strength Hoagland solution (3 mM KNO3, 2 mM Ca(NO3)2 x 4 H2O, 1 mM NH4H2PO4, 0.5 mM MgSO4 x 7 H2O, 1 µM KCl, 25 µM H3BO3, 2 µM MnSO4 x 4 H2O, 2 µM ZnSO4 x 7 H2O, 0.1 µM CuSO4 x 5 H2O, 0.1 µM (NH4)6Mo7O24 x 24 H2O, 20 µM Fe(Na)EDTA, 2 µM MnSO4 x H2O and 0.39 g/l MES adjusted to pH 5.5 with KOH). Plants were kept in either a climate chamber at 22 °C under a 10-hr-light/14hr-dark photoperiod or grown under standard greenhouse conditions as indicated. Mutant isolation The knock-out mutant for BIN2 (At4G18710, FLAG_593C09, Wassilawskija background) was isolated from the FLAGdb Collection at INRA Versailles (http://urgv. evry.inra.fr/projects/FLAGdb++/HTML/index.shtml). The knock-out mutants for GSK1 (At1G06390, SALK_062147, Colombia background) and ASKζ (At2G30980, SALK_ 46 Group II GSK3/SHAGGY-like kinases act redundantly in the brassinosteroid signalling pathway 060220, Colombia ecotype) were obtained from the SALK Institute Collection (http:// signal.salk.edu/cgi-bin/tdnaexpress). The double gsk1/askζ and triple bin2/gsk1/askζ mutants were obtained via cross-pollination and confirmed by PCR. In case of the triple insertion mutant, offspring of several triple mutants that originated from independent crossings were analyzed to circumvent any unwanted side effects of potential modifiers originating from mixing different ecotypes. Generation of gain-of-function mutants The full length coding sequences of GSK1 (101-FW and 102-RV) and ASKζ (121-FW and 122-RV) were amplified via RT-PCR on Arabidopsis Colombia leaf mRNA. Amplification products were cloned into the pMON999 (PET24; GSK1 and PET23; ASKζ, BgIII/Xba1) and completely sequenced. The generation of the dominant constructs of GSK1 and ASKζ into the Gateway pENTR4 vector (Invitrogen) was performed into two steps. First, the 5´part of GSK1 and ASKζ was amplified by PCR (M544-FW and M549-RV on PET24; 121-FW and M549-RV on PET23). These fragments were cloned as Nco1/Kpn1 fragments into pENTR4 (PET69; 5´ GSK1dom and PET70; 5´ASKζdom). Then the 3’ part, containing the glutamic acid (E) into lysin (K) substitution was generated (M550-FW and M551-RV on PET24; M550-FW and M546-RV on PET23). These PCR fragments were cloned into PET69 and PET70 to create PET72; GSK1dom and PET73; ASKζdom. Next, a Gateway LR recombination reaction transferred the dominant cDNA constructs into destination vector pK2GW7, which contains the 35S promoter and terminator (Karimi et al., 2002) and the inserts were fully sequenced (PET75; 35SCaMV:GSK1dom and PET76; 35SCaMV:ASKζdom). Nucleotide sequences of primers are listed in table 1. The constructs were introduced in Arabidopsis using Agrobacterium tumefaciens strain C58C1 (pMP90) and the floral dip method (Clough and Bent, 1998). Generation of CaMV35S constructs The full length coding sequences of GSK1 (101-FW and 102-RV) and ASKζ (121-FW and 122-RV) were amplified via RT-PCR on Arabidopsis Colombia leaf mRNA. BIN2 was amplified (144-FW and 145-RV) from a clone that was kindly gifted by Dr. Jianming Li. Amplification products were cloned into the Gateway pENTR4 vector (Nco1/BamH1, PET43; GSK1, PET44; ASKζ and PET78; BIN2) and fully sequenced. Next, a Gateway LR reaction was performed to transfer the constructs into the destination vector pH2WG7 (Karimi et al., 2002), containing the 35S promoter and terminator (PET142; 47 Chapter 2 35SCaMV:GSK1, PET143; 35SCaMV:ASKζ and PET145; 35SCaMV:BIN2). The inserts were sequenced and transformed via electroporation into the Escherichia coli strain XL1blue (Stragene) and afterwards in the Agrobacterium tumefaciens strain C58C1 (pMP90). The plasmids were introduced in the Arabidopsis bin2/gsk1/askζ mutant using the floral dip method (Clough and Bent, 1998). Nucleotide sequences of primers are listed in table 1. Generation of the transient-expressing constructs The full length coding sequences of GSK1 (101-FW and 102-RV) and ASKζ (121-FW and 122-RV) were amplified via RT-PCR on Arabidopsis leaf mRNA. Amplification products were cloned into the pMON999 (PET24; GSK1 and PET23; ASKζ, BgIII/Xba1). BIN2 was amplified from a clone (144-FW and 145-RV, PET42; BIN2) that was kindly gifted by Dr. Jianming Li. Generation of BIN2dom into the Gateway pENTR4 vector (Invitrogen) was performed into two steps. First, the 5´part of BIN2 was amplified by PCR using primers 144-FW and M549-RV on PET42 . This fragment was cloned as a Nco1/Kpn1 fragment into pENTR4 (PET71; 5´BIN2dom). The 3’ part, containing the glutamic acid (E) into lysin (K) substitution, was amplified with M550-FW and 145-RV. The PCR fragment was cloned into PET71 to create PET74. The full length coding sequences of GSK1, ASKζ, BIN2 and BIN2dom were amplified from PET24 (GSK1; M623-FW and 80-RV, cut with XhoI/XbaI), PET23 (ASKζ; M624-FW and 85-RV, cut with XhoI/XbaI), PET42 (BIN2; M621-FW and M622-RV, cut with SalI/SpeI) and PET74 (BIN2dom; M621-FW and M622RV, cutted with SalI/SpeI) and cloned in pER8 (XhoI/Spe1). The inserts were sequenced and transformed via electroporation into the Escherichia coli strain XL1-blue (Stragene) and afterwards in the Agrobacterium tumefaciens strain C58C1 (pMP90). The plasmids were introduced in the Arabidopsis bin2/gsk1/askζ mutant using the floral dip method (Clough and Bent, 1998). Nucleotide sequences of primers are listed in table 1. Estradiol treatments with pER8 transgenic lines T2 generation transgenic plants carrying the XVE cassette were selected on 0.8% (w/v) agar plates containing 0.5X Murashige and Skoog (MS) salts, 1.5% sucrose and 30 mg/l hygromycin for selection. After stratification for 2 days at 4 °C, seeds were germinated in a climate chamber at 22 °C under a 16-hr-light/8-hr-dark photoperiod for 10 days, after which 15-20 seedlings were transferred to 50 ml polypropylene tubes (Sarstedt) containing 10 ml 0.5X MS medium with 1.5% sucrose. The tubes were incubated on a shaker at 120 48 Group II GSK3/SHAGGY-like kinases act redundantly in the brassinosteroid signalling pathway rpm for 4 additional days before treatment. Plants were either treated with 4 µM estradiol (Sigma, dissolved in DMSO) or DMSO for 24 hours unless stated otherwise. Brassinazole experiment Brassinazole (Tokyo Chemical Industry) was dissolved in DMSO (1 mM) and stored at -20 °C. Seeds of Arabidopsis Colombia, mutant and transgenic lines were sterilized in 10% bleach, washed three times with sterile water, dissolved in 0.1% agarose and sown in a straight line on 0.8% (w/v) agar plates containing 0.5X Murashige and Skoog (MS) salts, 1.5% sucrose and 1 µM brassinazole or DMSO as control. The plates were wrapped in aluminium foil and incubated for 48 hrs at 4 °C to stimulate equal germination. Plates were placed in a vertical orientation for 7 days at 22 °C. The seedlings that were used for microscopy were made transparent in 100% EtOH. A similar experiment was performed with the pER8 transgenic lines, only seeds were sown on various plates with or without 4 µM estradiol, 1 µM brassinazole or DMSO as a control. RT PCR analysis on ASK insertion mutants Plant tissue was homogenized in liquid nitrogen and 1 ml Trizol (38% phenol, 0.8 M guanidine thiocyanate, 0.4 M ammonium thiocyanate, 0.1 M NaAc (pH5) and 5% glycerol) was added after which the samples were extracted with 0.3 ml chloroform. The nucleic acids were precipitated with an equal volume 2-propanol and washed with 75% EtOH. The pellet was dissolved in 500 µl RNase free H2O. A DNase treatment was performed by adding 50 µl DNase buffer (0.5M Tris/HCl pH 7.6; 0.1 M MgCl2) and 3 µl RNase free DNase (Roche), followed by an incubation step for 30 min at 37°C. Next, a phenolchloroform extraction was performed and the RNA was precipitated again using 2propanol. The pellet was washed with 75% EtOH and dissolved in RNase free H2O. Total RNA was quantified using the Nanodrop ND1000. cDNA was synthesized from total RNA (2 µg, boiled for 1 min) using 100 Units M-MLV Reverse Transcriptase (Invitrogen), 2 mM dNTPs, 100 mM DTT, 10X RT buffer and 10 µM oligo-dT primer (P0007) at 42°C for 1 hr. cDNA products were amplified using primers specific for GSK1 (M558–FW and M561RV) , ASKζ (M559-FW and M562-RV), BIN2 (M560-FW and M563-RV) and ACTIN (M554-FW and M555-RV). The nucleotide sequences of the primers used in figure 1b are shown in table 1. It was ensured that the amplification rate was in a linear range by using a reduced number of amplification cycles (27 cycles for GSKs, 21 cycles for ACTIN). All PCR products were separated by electrophoresis in a 1% (w/v) agarose. 49 Chapter 2 Table 1. Description of primers used for PCR Primer Gene Sequence 5´-3´ M554-FW M555-RV 85-FW 121-FW 122-RV M546-RV M559-FW M562-RV M624-FW M726-FW M727-RV 140-FW 141-RV M730-FW M731-RV 144-FW 145-RV M560-FW M563-RV M621-FW M622-RV M728-FW M729-RV M732-FW M733-RV M549-RV M550-FW 80-FW 101-FW 102-RV M544-FW M551-RV M558–FW M561-RV M623–FW M724-FW M725-RV P0007-RV ACTIN ACTIN ASKζ ASKζ ASKζ ASKζ ASKζ ASKζ ASKζ ASKζ ASKζ BEH3 BEH3 BES1 BES1 BIN2 BIN2 BIN2 BIN2 BIN2 BIN2 BIN2 BIN2 BZR1 BZR1 GSK GSK GSK1 GSK1 GSK1 GSK1 GSK1 GSK1 GSK1 GSK1 GSK1 GSK1 Oligo-dT GGCCAACAGAGAGAAGATGACTC CGAAAGCAGATTCGGGAAATGAAC GCTCTAGACTAGGGTCCAGCTTGAAATG GCCCATGGCTTCGATACCATG GGGATCCCTAGGG TCCAGCTTGAAATG GGGGTCGACCTAGGGTCCAGCTTGAAATGG CGTCGTCCCGATATAGACAAC CTAGGGTCCAGCTTGAAATGGA GGCTCGAGATGACTTCGATACCATTGGGG GTAATAGAGAGAGAACAAAATAGTCT GTTGTCTATATCGGGACGACGTTTC GGAATTCATGACGTCGGGGACTAGAACG CCTCGAGTCATCTGGTTCTTGAGTTTCC AAACGTACTTCTTCTTCAAAACGCG GACGTCATCTTCTTCGAGGAATACG GCCCATGGCTGATGATAAGGAGATGCCTG GGGATCCTTAAGTTCCAGATTGATTCAAGAAG CTGTGTCTCTATCGCCATGGCTG GATTGATTCAAGAAGCTTAGAC GGGTCGACATGGCTGATGATAAGGAGATG GGACTAGTTTAAGTTCCAGATTGATTCAAG AGAGAAGAGCTGGATTCACATGGTCT TCAGCCATGGCGATAGAGACACAG GTTGGAGAAGAAAGAGAGATTCTTC TCATCGGGAAAACCAACAACAAACC TTCGCGAGTTGGGGTACCAAGAAC GTTCTTGGTACCCCAACTCGCAAAGAAAT GCTCTAGATTAACTGTTTTGTAATCCTGTG CCCCATGGCCTCATTACCATTG GGGATCCTTAACTGTTTTGTAATCCTGTG GGCCATGGCTCATTACCATTGGGG GGCTCGAGTTAACTGTTTTGTAATCCTG CGCCGTCCTGAATTGGATTCT TTAACTGTTTTGTAATCCTGTG GGCTCGAGATGGCCTCATTACCATTGGGG GCAAAGAGAGAGGAAAAATTAGACT GAATCCAATTCAGGACGGCGTTTC GACTCGAGTCGACATCGATTTTTTTTTTTTTTT Yeast two-hybrid The full length coding sequences of GSK1 (101-FW and 102-RV, listed in table 1) and ASKζ (121-FW and 122-RV, listed in table 1) were amplified via RT-PCR on Arabidopsis leaf mRNA and cloned in pAS1 (Nco1/BamH1, Clontech). Dr. Jianming Li kindly gifted BIN2 in pAS1 and BZR1 and BES1 in pACT. BEH1 (ABRC clone U17510), BEH2 (ABRC 50 Group II GSK3/SHAGGY-like kinases act redundantly in the brassinosteroid signalling pathway clone U13164) and BEH4 (ABRC clone U13344) were cloned EcoRI/SalI in pAD-GAL4 (Stratagene), cut with EcoRI and Klenow was used to get it in frame. The full length coding sequence of BEH3 (140-FW and 141-RV, listed in table 1) were amplified via RT-PCR on Arabidopsis leaf mRNA and cloned in pAD-GAL4 (EcoRI/XhoI). The clones were sequenced to ensure that no mutations were introduced after PCR amplification. The plasmids were introduced into yeast strain PJ69 (James et al., 1996), which carries HIS3, ADE2 and LACZ reporter genes driven by distinct GAL4-responsive promoters. For the yeast two-hybrid screen, GSK1 was used as a bait against a cDNA library from the ABRC stock center (CD4-10) that was prepared from mature Arabidopsis leaves and roots. Library screen was performed as described previously (Quatrocchio et al., 2006) yielding approximately 1 x 105 transformants. The two-hybrid screen was optimized by adding 3amino-triazole, an enzymatic inhibitor of His3. Plasmid DNA was isolated from positive (HIS, ADE) colonies, transformed in Escherichia coli, sequenced and subsequently reintroduced in PJ69, either with the empty pBD-GAL4 vector or with the bait plasmid expressing GSK1. Only plasmids that specified a HIS, ADE phenotype when cotransformed with the bait vector, but not when co-transformed with the empty pBD-GAL4, were considered as positives and analyzed further. The yeast two hybrid assay between the BES1/BZR1 family and members of the GSK3/SHAGGY like family was performed on the same manner. NaCl stress assays In order to analyse the effect of NaCl stress on Arabidopsis Colombia and bin2/gsk1/askζ plants grown on soil, three-week-old plantlets were soaked every three days in 0 or 200 mM NaCl for a period of 10 days under greenhouse conditions. To test the effect of a NaCl treatment on Arabidopsis Colombia, gsk1, bin2/gsk1/askζ and hkt1 in hydroculture, seeds were sown directly in agarplugs (Araponics) in a hydroponic growing system (Araponics). After a cold treatment of 48 hrs at 4°C to synchronise germination, seeds were germinated in a climate chamber at 22 °C under a 10-hr-light/14hr-dark photoperiod in half strength Hoagland solution. After 3 weeks, plants of approximate equal size were transferred to one litre pots and subjected to half strength Hoagland solution supplemented with 0, 25 or 50 mM NaCl. Hoagland solutions were refreshed every week. Pictures were taken after an additional growth of 4 weeks. For the root measurement experiment, Arabidopsis Colombia, gsk1, bin2/gsk1/askζ plants of approximate equal size were transferred to one litre pots after 3 weeks. After a total of 30 days, plants were subjected to half strength Hoagland solution supplemented with 25 mM 51 Chapter 2 NaCl or further cultivated on standard Hoagland solution. After 3 days of treatment, root tips were dipped in active coal powder. After another 3 days, root growth was measured as the length of the unstained segment. Per treatment 12 plants were analysed, which were divided over a total of 4 pots. Acknowledgements We thank Maartje Kuijpers for plant care. BIN2, BES1 and BZR1 clones were a kind gift of Jianming Li. This work was supported by a grant of the Netherlands Organisation for Scientific Research (NWO). References Asami, T., Min, Y.K., Nagata, N., Yamagishi, K., Takatsuto, S., Fujioka, S., Murofushi, N., Yamaguchi, I. and Yoshida, S. (2000) Characterization of brassinazole, a triazole-type brassinosteroid biosynthesis inhibitor. Plant Physiol, 123, 93-100. Azpiroz, R., Wu, Y., LoCascio, J.C. and Feldmann, K.A. (1998) An Arabidopsis brassinosteroid-dependent mutant is blocked in cell elongation. Plant Cell, 10, 219-230. Berthomieu, P., Conejero, G., Nublat, A., Brackenbury, W.J., Lambert, C., Savio, C., Uozumi, N., Oiki, S., Yamada, K., Cellier, F., Gosti, F., Simonneau, T., Essah, P.A., Tester, M., Very, A.A., Sentenac, H. and Casse, F. (2003) Functional analysis of AtHKT1 in Arabidopsis shows that Na(+) recirculation by the phloem is crucial for salt tolerance. EMBO J, 22, 2004-2014. Charrier, B., Champion, A., Henry, Y. and Kreis, M. (2002) Expression profiling of the whole Arabidopsis shaggy-like kinase multigene family by real-time reverse transcriptase-polymerase chain reaction. Plant Physiol, 130, 577-590. Choe, S., Schmitz, R.J., Fujioka, S., Takatsuto, S., Lee, M.O., Yoshida, S., Feldmann, K.A. and Tax, F.E. (2002) Arabidopsis brassinosteroid-insensitive dwarf12 mutants are semidominant and defective in a glycogen synthase kinase 3beta-like kinase. Plant Physiol, 130, 1506-1515. Clough, S.J. and Bent, A.F. (1998) Floral dip: a simplified method for Agrobacterium-mediated transformation of Arabidopsis thaliana. Plant J, 16, 735-743. De Rybel, B., Audenaert, D., Vert, G., Rozhon, W., Mayerhofer, J., Peelman, F., Coutuer, S., Denayer, T., Jansen, L., Nguyen, L., Vanhoutte, I., Beemster, G.T., Vleminckx, K., Jonak, C., Chory, J., Inze, D., Russinova, E. and Beeckman, T. (2009) Chemical inhibition of a subset of Arabidopsis thaliana GSK3-like kinases activates brassinosteroid signaling. Chem Biol, 16, 594-604. Decroocq-Ferrant, V., Van Went, J., Bianchi, M.W., de Vries, S.C. and Kreis, M. (1995) Petunia hybrida homologues of shaggy/zeste-white 3 expressed in female and male reproductive organs. Plant J, 7, 897-911. Dornelas, M.C., Lejeune, B., Dron, M. and Kreis, M. (1998) The Arabidopsis SHAGGY-related protein kinase (ASK) gene family: structure, organization and evolution. Gene, 212, 249-257. Dornelas, M.C., Wittich, P., von Recklinghausen, I., van Lammeren, A. and Kreis, M. (1999) Characterization of three novel members of the Arabidopsis SHAGGY-related protein kinase (ASK) multigene family. Plant Mol Biol, 39, 137-147. 52 Group II GSK3/SHAGGY-like kinases act redundantly in the brassinosteroid signalling pathway Dornelas, M.C., Van Lammeren, A.A. and Kreis, M. (2000) Arabidopsis thaliana SHAGGY-related protein kinases (AtSK11 and 12) function in perianth and gynoecium development. Plant J, 21, 419-429. Einzenberger, E., Eller, N., Heberle-Bors, E. and Vicente, O. (1995) Isolation and expression during pollen development of a tobacco cDNA clone encoding a protein kinase homologous to shaggy/glycogen synthase kinase-3. Biochim Biophys Acta, 1260, 315-319. Embi, N., Rylatt, D.B. and Cohen, P. (1980) Glycogen synthase kinase-3 from rabbit skeletal muscle. Separation from cyclic-AMP-dependent protein kinase and phosphorylase kinase. Eur J Biochem, 107, 519-527. Frame, S. and Cohen, P. (2001) GSK3 takes centre stage more than 20 years after its discovery. Biochem J, 359, 1-16. Gampala, S.S., Kim, T.W., He, J.X., Tang, W., Deng, Z., Bai, M.Y., Guan, S., Lalonde, S., Sun, Y., Gendron, J.M., Chen, H., Shibagaki, N., Ferl, R.J., Ehrhardt, D., Chong, K., Burlingame, A.L. and Wang, Z.Y. (2007) An essential role for 14-3-3 proteins in brassinosteroid signal transduction in Arabidopsis. Dev Cell, 13, 177-189. He, J.X., Gendron, J.M., Yang, Y., Li, J. and Wang, Z.Y. (2002) The GSK3-like kinase BIN2 phosphorylates and destabilizes BZR1, a positive regulator of the brassinosteroid signaling pathway in Arabidopsis. Proc Natl Acad Sci U S A, 99, 10185-10190. Hoeflich, K.P., Luo, J., Rubie, E.A., Tsao, M.S., Jin, O. and Woodgett, J.R. (2000) Requirement for glycogen synthase kinase-3beta in cell survival and NF-kappaB activation. Nature, 406, 86-90. James, P., Halladay, J. and Craig, E.A. (1996) Genomic libraries and a host strain designed for highly efficient two-hybrid selection in yeast. Genetics, 144, 1425-1436. Jonak, C. and Hirt, H. (2002) Glycogen synthase kinase 3/SHAGGY-like kinases in plants: an emerging family with novel functions. Trends Plant Sci, 7, 457-461. Karimi, M., Inze, D. and Depicker, A. (2002) GATEWAY vectors for Agrobacterium-mediated plant transformation. Trends Plant Sci, 7, 193-195. Kempa, S., Rozhon, W., Samaj, J., Erban, A., Baluska, F., Becker, T., Haselmayer, J., Schleiff, E., Kopka, J., Hirt, H. and Jonak, C. (2007) A plastid-localized glycogen synthase kinase 3 modulates stress tolerance and carbohydrate metabolism. Plant J, 49, 1076-1090. Kim, T.W., Guan, S., Sun, Y., Deng, Z., Tang, W., Shang, J.X., Burlingame, A.L. and Wang, Z.Y. (2009) Brassinosteroid signal transduction from cell-surface receptor kinases to nuclear transcription factors. Nat Cell Biol, 11, 1254-1260. Koh, S., Lee, S.C., Kim, M.K., Koh, J.H., Lee, S., An, G., Choe, S. and Kim, S.R. (2007) T-DNA tagged knockout mutation of rice OsGSK1, an orthologue of Arabidopsis BIN2, with enhanced tolerance to various abiotic stresses. Plant Mol Biol, 65, 453-466. Li, J., Nagpal, P., Vitart, V., McMorris, T.C. and Chory, J. (1996) A role for brassinosteroids in lightdependent development of Arabidopsis. Science, 272, 398-401. Li, J., Nam, K.H., Vafeados, D. and Chory, J. (2001) BIN2, a new brassinosteroid-insensitive locus in Arabidopsis. Plant Physiol, 127, 14-22. Li, J. and Nam, K.H. (2002) Regulation of brassinosteroid signaling by a GSK3/SHAGGY-like kinase. Science, 295, 1299-1301. Li, J., Wen, J., Lease, K.A., Doke, J.T., Tax, F.E. and Walker, J.C. (2002) BAK1, an Arabidopsis LRR receptor-like protein kinase, interacts with BRI1 and modulates brassinosteroid signaling. Cell, 110, 213-222. Nam, K.H. and Li, J. (2002) BRI1/BAK1, a receptor kinase pair mediating brassinosteroid signaling. Cell, 110, 203-212. 53 Chapter 2 Papdi, C., Abraham, E., Joseph, M.P., Popescu, C., Koncz, C. and Szabados, L. (2008) Functional identification of Arabidopsis stress regulatory genes using the controlled cDNA overexpression system. Plant Physiol, 147, 528-542. Pay, A., Jonak, C., Bogre, L., Meskiene, I., Mairinger, T., Szalay, A., Heberle-Bors, E. and Hirt, H. (1993) The MsK family of alfalfa protein kinase genes encodes homologues of shaggy/glycogen synthase kinase-3 and shows differential expression patterns in plant organs and development. Plant J, 3, 847-856. Peng, P., Zhao, J., Zhu, Y., Asami, T. and Li, J. (2010) A direct docking mechanism for a plant GSK3-like kinase to phosphorylate its substrates. J Biol Chem, 285, 24646-24653. Perez-Perez, J.M., Ponce, M.R. and Micol, J.L. (2002) The UCU1 Arabidopsis gene encodes a SHAGGY/GSK3-like kinase required for cell expansion along the proximodistal axis. Dev Biol, 242, 161-173. Piao, H.L., Pih, K.T., Lim, J.H., Kang, S.G., Jin, J.B., Kim, S.H. and Hwang, I. (1999) An Arabidopsis GSK3/shaggy-like gene that complements yeast salt stress-sensitive mutants is induced by NaCl and abscisic acid. Plant Physiol, 119, 1527-1534. Piao, H.L., Lim, J.H., Kim, S.J., Cheong, G.W. and Hwang, I. (2001) Constitutive over-expression of AtGSK1 induces NaCl stress responses in the absence of NaCl stress and results in enhanced NaCl tolerance in Arabidopsis. Plant J, 27, 305-314. Pre, M., Atallah, M., Champion, A., De Vos, M., Pieterse, C.M. and Memelink, J. (2008) The AP2/ERF domain transcription factor ORA59 integrates jasmonic acid and ethylene signals in plant defense. Plant Physiol, 147, 1347-1357. Rozhon, W., Mayerhofer, J., Petutschnig, E., Fujioka, S. and Jonak, C. (2010) ASKtheta, a group-III Arabidopsis GSK3, functions in the brassinosteroid signalling pathway. Plant J, 62, 215-223. Ruel, L., Bourouis, M., Heitzler, P., Pantesco, V. and Simpson, P. (1993) Drosophila shaggy kinase and rat glycogen synthase kinase-3 have conserved activities and act downstream of Notch. Nature, 362, 557-560. Rus, A., Yokoi, S., Sharkhuu, A., Reddy, M., Lee, B.H., Matsumoto, T.K., Koiwa, H., Zhu, J.K., Bressan, R.A. and Hasegawa, P.M. (2001) AtHKT1 is a salt tolerance determinant that controls Na(+) entry into plant roots. Proc Natl Acad Sci U S A, 98, 14150-14155. Russinova, E., Borst, J.W., Kwaaitaal, M., Cano-Delgado, A., Yin, Y., Chory, J. and de Vries, S.C. (2004) Heterodimerization and endocytosis of Arabidopsis brassinosteroid receptors BRI1 and AtSERK3 (BAK1). Plant Cell, 16, 3216-3229. Ryu, H., Kim, K., Cho, H., Park, J., Choe, S. and Hwang, I. (2007) Nucleocytoplasmic shuttling of BZR1 mediated by phosphorylation is essential in Arabidopsis brassinosteroid signaling. Plant Cell, 19, 2749-2762. Siegfried, E., Chou, T.B. and Perrimon, N. (1992) wingless signaling acts through zeste-white 3, the Drosophila homolog of glycogen synthase kinase-3, to regulate engrailed and establish cell fate. Cell, 71, 1167-1179. Szekeres, M., Nemeth, K., Koncz-Kalman, Z., Mathur, J., Kauschmann, A., Altmann, T., Redei, G.P., Nagy, F., Schell, J. and Koncz, C. (1996) Brassinosteroids rescue the deficiency of CYP90, a cytochrome P450, controlling cell elongation and de-etiolation in Arabidopsis. Cell, 85, 171-182. Tichtinsky, G., Tavares, R., Takvorian, A., Schwebel-Dugue, N., Twell, D. and Kreis, M. (1998) An evolutionary conserved group of plant GSK-3/shaggy-like protein kinase genes preferentially expressed in developing pollen. Biochim Biophys Acta, 1442, 261-273. Vert, G. and Chory, J. (2006) Downstream nuclear events in brassinosteroid signalling. Nature, 441, 96-100. Wang, X., Kota, U., He, K., Blackburn, K., Li, J., Goshe, M.B., Huber, S.C. and Clouse, S.D. (2008) Sequential transphosphorylation of the BRI1/BAK1 receptor kinase complex impacts early events in brassinosteroid signaling. Dev Cell, 15, 220-235. 54 Group II GSK3/SHAGGY-like kinases act redundantly in the brassinosteroid signalling pathway Wu, S.J., Ding, L. and Zhu, J.K. (1996) SOS1, a Genetic Locus Essential for Salt Tolerance and Potassium Acquisition. Plant Cell, 8, 617-627. Yan, Z., Zhao, J., Peng, P., Chihara, R.K. and Li, J. (2009) BIN2 Functions Redundantly with Other Arabidopsis GSK3-Like Kinases to Regulate Brassinosteroid Signaling. Plant Physiol, 150, 710-721. Yin, Y., Wang, Z.Y., Mora-Garcia, S., Li, J., Yoshida, S., Asami, T. and Chory, J. (2002) BES1 accumulates in the nucleus in response to brassinosteroids to regulate gene expression and promote stem elongation. Cell, 109, 181-191. Zeng, H., Tang, Q. and Hua, X. (2009) Arabidopsis Brassinosteroid Mutants det2-1 and bin2-1 Display Altered Salt Tolerance. J Plant Growth Regul. Zhang, S., Cai, Z. and Wang, X. (2009) The primary signaling outputs of brassinosteroids are regulated by abscisic acid signaling. Proc Natl Acad Sci U S A, 106, 4543-4548. Zhao, J., Peng, P., Schmitz, R.J., Decker, A.D., Tax, F.E. and Li, J. (2002) Two putative BIN2 substrates are nuclear components of brassinosteroid signaling. Plant Physiol, 130, 1221-1229. Zuo, J., Niu, Q.W. and Chua, N.H. (2000) Technical advance: An estrogen receptor-based transactivator XVE mediates highly inducible gene expression in transgenic plants. Plant J, 24, 265-273. 55 Chapter 3 Towards identification of group II ASKs containing complexes in Arabidopsis Nathalie Verhoef and Erik Souer Towards identification of group II ASKs containing complexes in Arabidopsis Abstract The two mammalian glycogen synthase kinase 3 (GSK3) genes play a role in diverse processes like metabolism, development and disease. In Arabidopsis there are ten GSK3/SHAGGY-like kinases (ASKs) that have been implicated in stress and brassinosteroid (BR) signalling. The group II ASKs, including its most famous member BIN2, seem to be involved in the repression of BR signalling (Chapter 2). At the start of this project it was unknown by which mechanism group II ASKs are regulated in the BR signalling pathway. Furthermore group II ASKs might have additional roles in other signalling pathways. The setting in which group II ASK proteins reside within the cell might give answers to these questions. Therefore, we sought to reveal group II ASK complexes within the plant cell. For this purpose we fused the group II ASK members to various Tandem Affinity Purification (TAP) tags that were available. Although in most cases we found fusion protein expression, the final isolation of the protein complex of plant extracts appeared problematic. Introduction Several methods have been developed to discover interactions between a known protein and unknown partner proteins. Commonly used methodology to identify protein-protein interactions are complex immuno-precipitation (Co-IP) and yeast two-hybrid analysis. Yeast two-hybrid interaction requires protein processing and binding in a setting that is incomparable with the natural habitat where the studied protein normally resides in. A disadvantage of Co-IP is that for every protein a new antibody needs to be generated. Specificity and capacity of these antibodies to detect low levels of protein is sometimes insufficient. The development of several synthetic affinity tags for which high quality antibodies are available circumvents these disadvantages. Purification of protein complexes using antibodies against affinity tags can be performed directly from plant extracts. Furthermore, indirect interactions can be revealed, whereas yeast two-hybrid only identifies direct interactions between bait and prey protein. However, fusion of an affinity tag to a protein yields an artificial molecule, for which biological activity has to be proven. Full complementation of the corresponding mutant by the fusion protein is the preferred method to do so. The ideal affinity tag is small, allows rapid purification to maintain protein (complex) integrity and conserve post-translational modifications. Examples of epitope tags that are used nowadays are Green Fluorescent Protein (GFP), Glutathione-S transferase (GST), 59 Chapter 3 FLAG, myc and HIS (Reviewed by Terpe, 2003). Although many of these tags are very well suited to recognise the studied protein in extracts or even in living cells, purification of a complex where the protein of interest belongs to, is often hampered by low yield and copurification of proteins that are not genuine interactors. A D contaminant TAPi Target protein Protein A partners contaminant TAPa TEV cleavage site Target protein Protein A partners 6x HIS 9x myc CBP domain B TAPi purification C First affinity step Reversed TAPi purification First affinity step IgG beads E TAPa purification First affinity step IgG beads Cal beads TEV cleavage step Native elution (EGTA) Second affinity step Second affinity step 3C cleavage site 3C cleavage step Second affinity step Cal beads α-Myc Ni beads Native elution (EGTA) Elution via low pH Figure 1. Schematic representation of TAPi and TAPa purification. A) The TAPi tag consists of two IgGbinding units of protein A from Staphylococcus aureus (ProtA) and the Calmodulin Binding Peptide (CBP) that are fused in tandem and separated by a TEV protease cleavage site. B) TAPi tagged proteins can be recovered by affinity selection on an IgG-matrix. After washing, TEV protease is added to release the bound material. The eluate is then incubated with calmodulin-coated beads in the presence of calcium. After washing, the bounded material is released with EGTA. C) An alternative Reversed TAPi method has been used by us (see Results). Here, protein extracts are first incubated with calmodulin-coated beads in the presence of calcium. After washing, the bounded material is released with EGTA. Next the eluate is incubated with IgG beads, washed and elution is performed with a low pH buffer. D) The TAPa tag consists of two IgG-binding domains of protein A, a 3C cleavage site (3C), nine myc epitopes and six histidine repeats. E) Protein extracts are first incubated with IgG beads. After washing, TAPa-tagged proteins are cleaved by the low-temperature active rhinovirus 3C protease (3C). The eluate is either incubated with α-myc beads or in our case with nickel coated beads. In the latter case the complex can be eluted with a imidazole containing buffer. 60 Towards identification of group II ASKs containing complexes in Arabidopsis The generation of Tandem Affinity Purification (TAP) tags containing two or more epitopes separated by a cleavable protease site was believed to improve complex purification significantly. The first TAP purification was demonstrated in yeast and later this approach was used to purify protein complexes from many other organisms including plants (Puig et al., 2001; Rohila et al., 2004; Witte et al., 2004; Rubio et al., 2005; Rohila et al., 2006). The original TAP tag consists of the IgG binding domain of protein A from Staphylococcus aureus and the Calmodulin-Binding Peptide (CBP) which are fused in tandem and separated by a tobacco etch virus (TEV) cleavage site (Puig et al., 2001). Later on this tag (TAPi) was renovated to improve protein purification from plants (Rohila et al., 2004). This tag still contained the same affinity tags and cleavage site, but the duplicated regions of the tandem protein A domains were made non-identical and a nuclear localization signal (NLS) was removed from the CBP domain. In addition, catalase intron 1 from castor bean was introduced for higher gene expression in plants. As the TEV cleavage should be performed at a rather undesirable high temperature (16 ˚C), Rubio et al. (2005) developed an alternative TAP tag (TAPa) in which the TEV cleavage site was replaced by the human rhinovirus 3C protease cleavage site which is optimally cleaved at 4°C, allowing purification under conditions that usually stabilize protein complexes (Fig. 1d,e). The 3C cleavage site is also more specific as it consists of an eight amino acid sequence instead of the six residues present in the TEV cleavage site. Furthermore they replaced the CBP domain by nine myc epitopes and six histidine repeats, enabling protein purification via two different strategies. Protein-protein interaction studies proved to be useful in the study of several signalling pathways in plants (Xinga and Chen, 2006; Batelli et al., 2007; Van Leene et al., 2008). The signalling cascade downstream of the plant hormone brassinosteroid (BR) now constitutes a pathway that runs from the plasma membrane to activation of target promoters in the nucleus (Kim et al., 2009). BRs are perceived at the plasma membrane by the leucine-rich repeat receptor-like kinase BRASSINOSTEROID INSENSITIVE 1 (BRI1, Li and Chory, 1997). Upon BR activation, BRI1 dimerize and transphosphorylate with the coreceptor BRI1-ASSOCIATED RECEPTOR KINASE 1 (BAK1) and dissociate from the repressor BRI1 KINASE INHIBITOR 1 (BKI1), leading to BRI1 activation and phosphorylation of BR SIGNALING KINASES (BSKs; Li et al., 2002; Nam and Li, 2002; Wang and Chory, 2006; Tang et al., 2008). Downstream of the BSKs the GSK3/SHAGGYlike kinase BR-INSENSITIVE 2 (BIN2) acts a negative regulator by phosphorylating two closely related transcription factors BRI1-EMS-SUPPRESSOR 1 (BES1) and BRASSINAZOLE RESISTANT 1 (BZR1) (He et al., 2002; Yin et al., 2002). Kim et al. (2009) showed that phosphorylation of BSK1 a member of the (BSKs), interacts with and 61 Chapter 3 presumably activates the plant-specific serine/threonine protein phosphatase BRI1 SUPPRESSOR 1 (BSU1), which in turn inactivates BIN2 by dephosphorylation at pTyr200. As a result unphosphorylated BZR1/BES1 accumulate in the nucleus where they regulate the expression of BR-responsive genes. Next to BIN2, 5 other Arabidopsis SHAGGY-like protein kinases (ASKs) act redundantly with BIN2 in the Arabidopsis BR signalling pathway (Chapter 2; Vert and Chory, 2006; Kim et al., 2009; Yan et al., 2009; Rozhon et al., 2010). In order to learn more about the mode of action of GSK3/SHAGGY-like kinases, we transformed Arabidopsis with different TAP tagged versions of group II members for subsequent analysis of interacting proteins in plants. Here, we present an analysis of the efficiency of complex purification of various TAP fusions. Results Molecular and phenotypical characterization of TAPi-GSK1/ASKζ transgenic lines In order to isolate proteins interacting with ASKζ and GSK1 we used the improved Tandem Affinity Purification tag (TAPi tag) made by Rohila et al. (2004) as this tag had already successfully been used for the isolation of protein complexes from plants. Transgenic Arabidopsis lines were generated containing a N-terminal fusion of the TAPi tag to both ASKζ and GSK1 (N-TAPiASKζ and N-TAPiGSK1, Fig. 2a). GSK1 was also fused with the C-TAPi tag at its carboxyl end (C-TAPiGSK1, Fig. 2a). As a negative control to use during complex isolation along with the TAPi tagged fusions, the C-TAPi tag alone was transformed to Arabidopsis. All TAPi derived constructs were over-expressed in Arabidopsis Colombia under control of the strong constitutive CaMV35S promoter. Independent lines of the T2 generation were screened for their ability to express the fusion protein and for several transgenic lines containing N-TAPiGSK1, N-TAPiASKζ and CTAPi constructs high levels of fusion-protein were detected (Fig. 2b-d). The PAP antibody cross-reacted with proteins of the expected size; the TAPi tag alone had a size of ~20 kDa and N-TAPiGSK1 and N-TAPiASKζ had a size of ~78 kDa. In the C-TAPiGSK1 transgenic lines no accumulation of fusion-protein could be detected (data not shown). Since we could detect the fusion mRNA in the C-TAPiGSK1 transformed T2 lines, it seems that the tag is somehow detrimental for synthesis or stability of the C-terminal tagged fusion protein (data not shown). Visually, all characterized transgenic lines were indistinguishable from untransformed Arabidopsis Colombia (Fig. 2e). 62 T35S TEV CBP domain attR1::Cmr::ccdB::attR2 08 2x IgG BD 35S:C-TAPi 35S T35S attR1::Cmr::ccdB::attR2 CBP domain TEV C N -T AP iA N SK -T ζ J8 AP 10 iA 0N SK 3 -T ζ J8 AP 10 iA 0N S 6 -T Kζ J AP 81 00 W iAS -7 tC Kζ J 81 ol 0 35S D C -T A C Pi J -T J A 81 C Pi J 01-T 81 1 0 A C Pi J 1-4 -T 81 0 A C Pi J 1-5 -T 8 AP 101 C -T i J81 -6 0 A W Pi J 1-8 t C 81 ol 01-9 N -T AP B 35S:N-TAPi N A iG SK -T 1 J8 AP 09 iG 7N 6 SK -T 1 J8 AP 09 iG 7W 10 SK tC 1 J8 ol 09 7 -1 3 Towards identification of group II ASKs containing complexes in Arabidopsis 2x IgG BD Figure 2. Molecular and phenotypic analysis of TAPi-GSK1/ASKζ transgenic lines. A) Schematic drawing of the N-TAPi and C-TAPi expression cassettes under control of the CaMV35S promoter. ASKζ and GSK1 will replace the region between the two attR recombination sites after Gateway recombination. B-D) From several independent N-TAPiGSK1 (B), N-TAPiASKζ (C) and C-TAPi (D) lines protein was extracted from leaves and loaded on a SDS-PAGE gel followed by Western blotting. The Western blot was immuno-probed with the peroxidase anti peroxidase antibody (PAP) that recognizes the protein A part of the TAPi tag. TAPi purification with N-TAPiGSK1 Initially, a small-scale TAP tag purification (~3 gram leaf tissue) was performed with plants expressing N-TAPiGSK1 or C-TAPi. Protein extracts were incubated with IgG beads and after washing incubated with TEV protease. In the second purification step, the TEV eluate was incubated with calmodulin-coated beads and subsequently eluted by using an EGTAcontaining buffer (Fig. 1b). In order to follow the purification process, samples were taken after each step and subjected to Western blot analysis using the peroxidase anti peroxidase antibody (PAP) or the anti calmodulin binding protein antibody (CBP) to follow the presence or absence of the epitopes. N-TAPiGSK1 as well as C-TAPi proteins bound quite efficient to the IgG beads, but after the TEV protease cleavage both proteins could not be detected (Fig. 3a,b). In the case of C-TAPi, cleavage did work as we could detect the protein A part that remained attached to IgG beads after cleaving and, after concentration of the final C-TAPi eluate, the CBP part that was clipped of the original fusion protein (Fig. 3a). After N-TAPiGSK1 purification no N-TAPiGSK1 protein was observed, neither could we detect the cleaved protein A part bound to the IgG beads (Fig. 3b). Although we succeeded to perform the TAPi purification with C-TAPi, analysis of flow through fractions and eluates revealed that a lot of the fusion protein was lost during the purification, especially during the TEV protease cleavage step. During purification of N-TAPiGSK1 all fusion protein seems to be lost after the TEV protease cleavage and second purification step. 63 Chapter 3 B First affinity step Ex tra c Bo t (1 un ) d Fi ed to na Ig le G lu at (2 e ) (5 ) Ex tra c Bo t (1 un ) de d El to ua Ig te G af (2 te Ig ) r G be TEV ad cl s ea (4 va ) g e (3 ) e Ex tra c Bo t (1 un ) d Fi ed to na Ig le G lu at (2 e ) (5 ) tra c Bo t (1 un ) de d El to ua Ig te G af (2 te Ig ) r G be TEV ad cl s ea (4 va ) g Ex Input 1 C (3 ) A IgG beads 2 4 TEV cleavage step TAPiGSK1 TAPiGSK1 3 Second affinity step TAPi Protein A part Cal beads TAPi * ~ 35S:TAPiJ8321-5 CBP part * ~ Native elution (EGTA) 35S:N-TAPiGSK1J8300-4 5 Figure 3. Inmuno-blot of different fractions obtained during TAPi purification of N-TAPiGSK1 and CTAPi. A-B) A TAPi purification was performed using protein extracts from C-TAPi (A) and N-TAPiGSK1 (B) transformants. Several fractions, indicated above each lane, were separated on a 12,5% SDS-PAGE gel. The Western blot was probed with either CBP (*) or PAP (~) antibodies. The numbering corresponds to fractions that were collected during purification as depicted in C (C) TAPi purification scheme. The numbering refers to samples on the immuno-blots shown in A and B : 1. Input protein extract (40 µg); 2. IgG bound fraction prior to TEV cleavage (10x); 3. Eluate after protease TEV cleavage (20x); 4. IgG beads after the TEV protease cleavage step (20x); 5. Eluate from calmodulin beads using EGTA containing buffer followed by a concentration step (250x). x indicate the amount of protein loaded relative to input sample. In order to find out why a lot of fusion protein was lost during the TEV protease step, several experiments were performed. Rohila et al. (2004) reported that binding of a certain protease to the IgG beads might have a negative effect on the purification of certain TAP tagged proteins. Addition of the protease inhibitor E64 increased their protein recovery, but in our case the amount of fusion protein was not visibly changed with or without the addition of E64 (data not shown) at the stage when the C-TAPi or N-TAPiGSK1 was bound to the IgG beads. Next, we determined how efficiently TEV protease cleaved C-TAPi and N-TAPiGSK1. Protein extracts were incubated with IgG coated beads and after washing TEV protease was added. At several time points a sample was taken and immuno-probed with the PAP or CBP antibody (Fig. 4). Looking at the C-TAPi cleavage, most of the fusion proteins were cleaved within 1 hour. Cleavage of N-TAPiGSK1 on the other hand was less efficient. After 1 hour of treatment still a lot of fusion protein was not cleaved. Due to the high background of the immuno-blot incubated with CBP antibody we failed to detect the cleaved N-TAPiGSK1 proteins. So, although a large fraction of the N-TAPiGSK1 fusion protein is cleaved it could be that the released CBP-GSK1 fusion is degraded by a protease in the extract. 64 Towards identification of group II ASKs containing complexes in Arabidopsis N-TAPiGSK1J8293-6 60 30 45 15 pu 0 In 60 30 45 15 0 In pu t t C-TAPiJ8321-5 min TAPGSK1 CBP TAP Protein A part Protein A part 0 CBP part TAPGSK1 PAP TAP Protein A part Protein A part Figure 4. TEV does not cleave N-TAPiGSK1 efficiently. Protein extracts from transgenic plants containing NTAPiGSK1 and C-TAPi (negative control) were incubated with IgG coated beads. During the TEV cleavage several samples were taken and separated on a 12,5% SDS-PAGE gel. The Western blot was immuno-probed with either the CBP or PAP antibody. The input extract contained 40 µg protein. In the TEV with IgG beads fractions, 50x the amount of protein was loaded relative to the input. Reverse TAPi purification with N-TAPiGSK1 As optimilization of the TAPi purification was unsuccesfull we reasoned it might be possible to perform TAPi purification in reverse ommiting the TEV cleavage step (Fig. 5). The protein extract was first incubated with the calmodulin coated beads and then the second affinity purification step was performed with IgG beads. To release the proteins from the beads without using TEV cleavage, other ways of eluting were tested such as boiling and by using the gentle Ag/Ab buffer (Pierce). The best results were obtained with a low pH buffer (Fig. 5). The reverse TAPi purification was more efficient in comparison to the TAPi purification. C-TAPi binding and elution from the calmodulin beads went very well, but in the case of N-TAPiGSK1 not all fusion protein did bind to beads or was released from the beads. Elution from the IgG beads went successful, but the amount of fusion protein at the end was too low to detect it on SDS PAGE after Coomassie staining. 65 Chapter 3 (6 ) B In pu t( C 1) al be Fi ad na s e le lu at lu Fi na ate e ( le 1 2) Pr lua (3) ot te A 2 (4 C ) al be ad C al s be (5) Ig ad s G be flow ad th r s (7 oug ) h In pu t( C 1) al be Fi ads na l e elua lu te Fi (2 na ate ) le 1 Pr lua (3) te ot 2 A (4 C ) al be ad C s al ( 5) be Ig ads G be flow ad th r s (7 oug ) h (6 ) A Input 1 C First affinity step Cal beads 6 5 Native elution (EGTA) TAPiGSK1 Cal beads 2 second affinity step IgG beads TAP Elution via low pH 7 C:TAPiJ8321-5 N-TAPiGSK1J8293-6 IgG beads 3,4 Figure 5. Inmuno-blot of different fractions obtained during the reverse TAPi purification of N-TAPiGSK1 and C-TAPi. A-B) A reverse TAPi purification was performed with protein extracts from N-TAPiGSK1 (A) and C-TAPi (B) transformants. Several fractions, indicated above each lane, were separated on a 12,5% SDS-PAGE gel. The Western blot was probed with the CBP antibody. The numbering corresponds to fractions that were collected during purification as depicted in C (C) Reverse TAPi purification scheme. The numbering refers to samples on the immuno-blots shown in A and B : 1. Input protein extract (40 µg); 2. Eluate after from calmodulin beads using EGTA containing buffer (6,25x); 3,4. 2x Eluate from IgG beads using a low pH buffer followed by a concentration step (250x); 5. Calmodulin beads after elution (6,25x); 6. Unbound protein after calmodulin bead incubation (18,75x); 7. IgG beads after elution (100x). x indicate the amount of protein loaded relative to extract sample. Prot A (1 ng) was loaded to quantify the amount of detected protein. Molecular and phenotypic characterization of TAPa-GSK1/ASKζ transgenic lines Rubio et al. (2005) constructed another version of the TAP tag, the TAPa tag. They improved the tag by substituting the TEV cleavage site by the human rhinovirus 3C protease cleavage site and replacing the CBP domain by nine myc epitopes and six histidine repeats, enabling protein purification via two different purification strategies (Fig. 6a). We generated transgenic plants expressing N-TAPaGSK1, C-TAPaGSK1 and C-TAPaASKζ. As a negative control we kindly received N-TAPaGFP seeds from Alois Schweighofer (Max-Planck-Institute of Molecular Plant Physiology). All TAPa derived constructs were expressed in Arabidopsis Colombia under control of the CaMV35S promoter. Transformants were selected, self-pollinated and all transgenic lines had a similar phenotype as wild type Colombia (Fig. 6b). Independent lines of the T2 generation were screened for their ability to express the fusion protein. Several transgenic lines containing N-TAPaGSK1, C-TAPaGSK1 or C-TAPaASKζ constructs produced fusion-proteins, however protein levels were lower compared to the amounts produced in TAPi expressing lines (Fig. 6c). The ASK fusions had a size of ~100 kDa while N-TAPaGFP had a size of ~75 kDa (Fig. 6c,d) which corresponds well to the expected size. 66 Towards identification of group II ASKs containing complexes in Arabidopsis A B N-TAPa 2x 35S TMVU1Ω 2x IgG BD 6x HIS 3C N-TAPaGSK1L8079 Nos ter 9x myc attR1::Cmr::ccdB::attR2 C-TAPa 2x 35S TMVU1Ω 3C C-TAPaASKζL8082 Nos ter 2x IgG BD Wt Col 0 D C C N -T AP aG SK -T AP 1 L8 07 N aAS 9 -T AP Kζ M 80 aG 34 W t C FP M ol 80 7 08 3 9x myc W tC o N l -T A C PiG -T A SK N Pi L8 1 L80 -T 77 07 A 8 N PaG -T AP SK a 1L C -T GS 807 9 A K C PaA 1 L80 -T 80 AP SK C a ζ L8 08 -T AS 1 AP K aG ζ L8 SK 082 1 L8 attR1::Cmr::ccdB::attR2 6x HIS Figure 6. Molecular and phenotypic analysis of TAPa-GSK1/ASKζ transgenic lines. A) Schematic drawing of the N-TAPa and C-TAPa expression cassettes under control of the CaMV35S promoter. ASKζ and GSK1 will replace the region between the two AttR recombination sites after Gateway recombination. B) Phenotype of transgenic lines expressing N-TAPaGSK1, C-TAPaASKζ and wild type Colombia grown under short-day conditions. C-D) Immunoblots containing leaf protein extracts from several selected lines containing NTAPaGSK1, C-TAPaASKζ or N-TAPaGFP constructs compared to transformants expressing TAPi fusions. The immunoblot was probed with the PAP antibody. Are TAP tagged versions of ASKs still functional? In order to check whether ASK fusions were still functional, we expressed N-TAPaGSK1, C-TAPaGSK1 or C-TAPaASKζ in triple bin2/gsk1/askζ mutant background (Chapter 2) and checked whether they could rescue the mutant phenotype. Several independent lines of the T2 generation containing N-TAPaGSK1, C-TAPaGSK1 or C-TAPaASKζ constructs produced fusion-proteins (Fig. 7a). Unfortunately, all transgenic lines still appeared to have a similar phenotype as the triple insertion mutant (Fig. 7b). As the bin2/gsk1/askζ mutant phenotype is not always easy to score visually, we used sensitivity towards the BR inhibitor brassinazole (brz) as an assay for complementation of the mutant phenotype. Again, in none of the transgenic plants sensitivity to brz was restored (Fig. 7c). This suggests that either the TAP tag disturbs the function of the ASKs or the used CaMV35S promoter did not express the transgene at the proper time and/or place. The latter could be ruled out as expression of the ASKζ and BIN2 cDNAs under control of the CaMV35S promoter in the triple mutant background restored their sensitivity towards brz (Fig. 7d). Unfortunately, due to poor plant transformation we only possessed one transformant that expressed GSK1. 67 Chapter 3 A N-TAPaGSK1M8111 3 5 6a 7b B Wt a b e f d c C-TAPaASKζΜ8112 2 4 10c 13d Wt bin2/gsk1/askζ Wild type C-TAPaGSK1M8113 2 D 3 7e 10f Wild typeL8211-2 - + Wt 35S:ASKζL8251-8a - Wild typeL8211-2 - 1 µM brz 1 µM brz - + - 35S:BIN2L8253-2a + Wild typeL8211-2 + - + 35S:GSK1L8250-1a - + 35S:ASKζL8251-12a + bin2/gsk1/askζL8215-2 - 35S:BIN2L8253-3a - + + bin2/gsk1/askζL8215-2 - + bin2/gsk1/askζL8215-2 - + Figure 7. The TAPa tag disturbs the function of ASKs. A) Immunoblots containing leaf protein extracts from several independent T2 generation lines expressing N-TAPaGSK1M8111, C-TAPaASKζ M8112 or C-TAPaGSK1M8113 in triple mutant background. Wt Col was used as control. The immunoblot was probed with the myc antibody. B) Phenotype of transgenic lines expressing N-TAPaGSK1M8111, C-TAPaASKζ M8112 or C-TAPaGSK1M8113 in triple mutant background and wild type Colombia grown under short-day conditions. C) No restored brassinazole sensitivity after expression of N-TAPaGSK1M8111, C-TAPaASKζM8112 or C-TAPaGSK1M8113 genes in triple mutant seedlings. The characters (a-f) depicted in B and C refers to samples on the immuno-blots shown in A. D) Restored brassinazole sensitivity after expression of individual group II ASK genes in triple mutant seedlings. Constructs were over-expressed in triple mutant background. 68 a Towards identification of group II ASKs containing complexes in Arabidopsis C Wild type bin2/gsk1/askζ a b Wild type bin2/gsk1/askζ c d Wild type bin2/gsk1/askζ e f DMSO 1 µM brz DMSO 1 µM brz DMSO 1 µM brz 69 Chapter 3 Although GSK1 expression was high (data not shown), the T2 generation seedlings were less sensitive to a brz treatment than wild type seedlings and thus CaMV35S GSK1 did not fully complemented the mutant phenotype. (Fig. 7d). Taken together, at least in the case of ASKζ and BIN2 it seems that the TAP tag itself has a detrimental effect on the function of tra ct (1 ) be El ads ua te bef or a Fi na fter e e le 3C l u t io lu cl n at e (2 e ) (4 ava ) ge (3 ) IgG beads First affinity step 2 G TAPaASKζ TAPaGSK1 Input 1 D Ig C Ex B Ex tra c Ig t (1 G ) be El ads ua te bef or a Fi na fter e e le 3C l u t io lu cl n at e (2 e ) (4 ava ) ge (3 ) A Ex tra c Ig t (1 G ) be El ads ua te bef or a Fi na fter e e le 3C lut io lu cl n at e (2 e ) (4 ava ) ge (3 ) these ASKs. 3C cleavage step TAPaGFP IgG beads 3 N-TAPaGSK1L8079-1 C-TAPaASKζL8082-5 N-TAPaGFP M8071 Second affinity step Ni beads Ni beads 4 Figure 8. Inmuno-blot of different fractions obtained during the TAPa purification of N-TAPaGSK1, CTAPaASKζ and N-TAPaGFP. A-C) A TAPa purification was performed with protein extracts from NTAPaGSK1 (A) and C-TAPaASKζ (Β) and N-TAPaGFP (C) transformants. Several fractions, indicated above each lane, were separated on a 12,5% SDS-PAGE gel. The Western blot was probed with α-myc. The numbering corresponds to fractions that were collected during purification as depicted in D (D) TAPa purification scheme. The numbering refers to samples on the immuno-blots shown in A , B and C: 1. Input protein extract (40 µg); 2. IgG bound fraction prior to 3C protease cleavage (2,5x); 3. Eluate after 3C protease cleavage (6,25x); 4. Eluate from Ni beads using imidazole containing buffer followed by a concentration step (100x). x indicate the amount of protein loaded relative to extract sample. TAPa purification with N-TAPaGSK1 and C-TAPaASKζ Although the TAPa tagged proteins do not seem to be functional we still pursued to test if we could purify the fusion protein and anything that might be associated with it. A largescale TAPa tag purification (~15 gram leaf tissue) was performed with plants expressing NTAPaGSK1, C-TAPaASKζ and N-TAPaGFP. Total protein was isolated and incubated with IgG beads. After washing the beads were incubated with 3C protease and the collected eluate was incubated with nickel-coated resin. Final elution was performed with an imidazole-containing buffer and the eluate was concentrated. After each step samples were taken to evaluate the efficiency of the purification (Fig. 8). Fusion protein was lost in the IgG binding step and during the 3C cleavage step, but most protein seem to be lost during the second affinity purification step (Fig. 8). Although we succeeded to purify both ASKs 70 Towards identification of group II ASKs containing complexes in Arabidopsis and silver staining revealed bands that most probably correspond to both ASKs, the protein level was still too low to detect it by mass spectrometry. Discussion Almost a decade ago Puig et al. (2001) developed the Tandem Affinity Purification (TAP) method to rapidly purify protein complexes under native conditions from yeast. Although this tag has been proven to be useful, only a limited number of articles have been published about TAP purifications in plants suggesting that this method is not totally trouble-free. In order to optimize TAP purification from plants, scientists developed over the years several plant-adapted versions of the TAP tag, such as TAPi, TAPa and GS (Rohila et al., 2004; Rubio et al., 2005; Van Leene et al., 2008). In chapter 2 we showed that group II GSK3-like kinases are involved in BR signalling and discussed about a possible function of GSK3-like kinases in salt stress signalling. In order to learn more about the functions of group II GSK3-like kinases we wanted to perform TAP purifications with the GSK3-like kinases GSK1 and ASKζ. In transgenic plants harbouring the C-TAPiGSK1 construct we could not detect any fusion-protein. Probably the TAPi tag at the C-terminus of GSK1 disturbs the synthesis or stability of the fusion-protein emphasising the necessity for making both N and C-terminal fusions when using epitope tags. In chapter 2 we observed that gain of function mutants of GSK1 and ASKζ develop a dwarfish phenotype. All transgenic plants carrying TAP tagged versions of GSK1 and ASKζ turned out to have a normal phenotype. Li et al. (2002) observed for a fraction of transformants a dwarfish phenotype when they over-expressed BIN2. In our hands, expression of BIN2 under control of the CaMV35S promoter only sporadically leads to dwarfism, depending on high transcript levels. Taken this into account it could be that the level of expression in the transgenic lines was too low to generate a dwarfish phenotype or that we should screen more transformants. Nevertheless, the lack of phenotypic effects in lines over-expressing group II GSK3-like kinases seems not to be a good indication for functionality of the protein. TAPa tagged GSK1 or ASKζ had a normal phenotype too. Here we checked the functionality of the tagged ASKs by expressing all three constructs in the triple bin2/gsk1/askζ mutant background (described in chapter 2). None of the transgenic lines could rescue the mutant phenotype and insensitivity towards the BR inhibitor brz was comparable to the triple mutant control. The CaMV35S promoter worked properly as sole expression of ASKζ and BIN2 under control of the CaMV35S promoter in the triple mutant 71 Chapter 3 background resulted in sensitivity towards a brz treatment. It thus seems that the TAPa tag has a detrimental effect on the function and/or folding of the ASKs and might disrupt certain ASK-protein interactions. We did not test fusions with TAPi for activity because similar antibiotic resistance markers in the T-DNA tagged mutants and the TAPi vectors made generation of these transformants difficult. At least our results show that testing complementation of the corresponding mutant by the fusion proteins is a required experiment. Although there are doubts about the activity of the fusion-proteins it is not to be said that these proteins do not end up in their natural complexes or at least contact a subset of their natural partners. Also, as the fusion protein itself is translated well as indicated by Western blots they can serve at least as pilots for testing the feasibility of different TAP tag purifications. In a pilot TAPi purification with N-TAPiGSK1 and C-TAPi as a control we succeeded to purify the CBP part of the TAPi tag. The N-TAPiGSK1 purification however failed; we could not detect any purified CBP-GSK1 in the final eluate nor in the eluate after the TEV cleavage. Somehow the fusion-protein completely disappeared during the TEV cleavage, as we could not detect any fusion-protein in the flow through after binding to the IgG beads, neither attached to IgG beads after the TEV protease cleavage. This experiment was repeated several times with the same result. Specificity of the TEV protease has been questioned as some proteins might contain an endogenous TEV protease cleavage site or a degenerate site that also can be recognized by the TEV protease. In the peptide sequence of GSK1 we did not spot a TEV protease cleavage site, so it seems not likely that GSK1 is cleaved. Rohila et al. (2004) encountered that some TAP purifications were less successful due to binding of a certain protease to the IgG beads. To overcome this problem they added the protease inhibitor E64 during the TEV cleavage step. The addition of E64 in our TAPi purifications did not increase protein recovery. As we could detect cleavage of CBP-GSK1 from the IgG coated beads, it is most likely that GSK1 is somehow degraded after it is cleaved. Several attempts to optimize the TAPi purification did not help and our last shot to use the TAPi tag was to try to reverse the purification steps thereby omitting cleavage with TEV protease. Although we were able to purify our protein of interest, the amount was too low to detect it on a SDS PAGE gel stained with Coomassie blue. In addition, silver staining revealed a high background of contaminating proteins. As the normal and reverse TAPi purifications did not look promising, we switched to the TAPa system developed by Rubio (2005). In this tag the rather undesirable TEV cleavage site is replaced by the human rhinovirus 3C protease cleavage site which is efficiently cleaved at 4 °C. Furthermore, the cleavage site is more specific as it consists of an eight amino acid sequence instead of six, thereby lowering the chance that your protein of interest and its bindings partners are cleaved. In addition, they replaced the CBP domain by 72 Towards identification of group II ASKs containing complexes in Arabidopsis nine myc epitopes and six histidine repeats, allowing one to chose between metal affinity chromatography or myc epitope immunoprecipitation. The tag-protein level was considerably lower in plants expressing N-TAPaGSK1, C-TAPaGSK1 and C-TAPaASKζ, compared to those containing the TAPi fusions. Although we succeeded to purify our ASKs, the purification was not optimal as a lot of our fusion protein was lost during the purification steps. Silver staining of the final TAP eluates revealed bands that corresponded to the size of the ASKs. However mass spectrometry used to identify the ASKs and possible interactors failed because the protein levels were too low. Unfortunately there was no time to further optimize the TAPa purification. Although the TAPa tag seems to have a detrimental effect on the ASKs it still may be partially functional and thus might be used for identification of binding partners. In general, there are several ideas that might help to successfully purify TAP tagged complexes in the future. First, purification of binding partners might work better when performed in a mutant background as there no competition with the endogenous protein. Second, the amount of protein extract should be increased and the best way to do this is to perform multiple TAPa purifications in parallel rather then to perform one large purification. Third, although it is preferable to perform the TAPa purification from whole plantlets it might be easier to perform a purification from cell suspensions. Cell suspensions are easy to handle, grow fast and one has an unlimited supply of cells that grow equally. However a disadvantage could be that your protein of interest is involved in certain biological processes that do not take place in cell suspensions. In that case certain bindings partners will not be revealed. Fourth, it might help to perform protein cross-linking to stabilize the protein in its complex before the start of the purification (Rohila et al., 2004). This is especially worthwhile when the interaction is weak or transient. Although there are still several possibilities that might help to improve the TAPa purification it is questionable whether the TAP purification technique is the best way to reveal protein complexes. The fact that in these ten years only a few articles are published about TAP purifications, mainly by the scientist that developed the tags themselves, it is questionable whether this technology is a foregone conclusion. At least the TAP tag seems not superior to other, single tag methods like GFP, MYC or FLAG that have been used successfully by a large number of researchers nowadays. 73 Chapter 3 Experimental procedures Plant material and growth Seeds of Arabidopsis Colombia and transgenic lines were sterilized in 10% bleach, washed three times with sterile water, dissolved in 0.1% agarose and sewed on 0.8% (w/v) agar plates containing 0.5X Murashige and Skoog (MS) salts, 1.5% sucrose at pH 5.7-5.9. The seeds were germinated in the climate chamber at 22 °C under a 10-hr-light/14-hr-dark photoperiod. Later, seedlings were transferred to soil and growth was continued in either a climate chamber at 22 °C under a 10-hr-light/14-hr-dark photoperiod or under standard greenhouse conditions. Brassinazole experiment Brassinazole (Tokyo Chemical Industry) was dissolved in DMSO (1 mM) and stored at -20 °C. Seeds of Arabidopsis Colombia, mutant and transgenic lines were sterilized in 10% bleach, washed three times with sterile water, dissolved in 0.1% agarose and sown in a straight line on 0.8% (w/v) agar plates containing 0.5X Murashige and Skoog (MS) salts, 1.5% sucrose and 1 µM brassinazole or DMSO as control. The plates were wrapped in aluminium foil and incubated for 48 hrs at 4 °C to stimulate equal germination. Plates were placed in a vertical orientation for 7 days at 22 °C. Generation of TAPi/TAPa fusion constructs To obtain N-tagged GSK1/ASKζ fusion proteins, full length coding sequences of GSK1 (NTAPi, 101-FW and 102-RV; N-TAPa M544-FW and M551-RV) and ASKζ (121-FW and 122-RV) were amplified by RT-PCR on Arabidopsis leaf total RNA. In order to generate the C-tagged GSK1/ASKζ fusions, the coding sequences of GSK1 (101-FW and 123-RV) and ASKζ (121-FW and M547-RV) were amplified without STOP codon. The amplification products were cloned into the Gateway pENTR4 vector (Nco1/BamH1; PET43 GSK1+ ASKζ −STOP , PET44 ASKζ + STOP , PET63 GSK1 + STOP STOP , PET46 GSK1- STOP and Nco1/Xho; PET65 ). Next, a Gateway LR reaction was performed to transfer PET43 (PET50) and PET44 (PET51) in the pN-TAPi destination vector and PET46 (PET49) in the pC-TAPi destination vector. The pN-TAPi and pC-TAPi constructs were kindly provided by Dr. Michael Fromm (University of Nebraska, Lincoln, USA). The pNTAPa and pC-TAPa plasmids were obtained from the Arabidopsis Biological Resource 74 Towards identification of group II ASKs containing complexes in Arabidopsis Center (ABRC). PET63 (NcoI/BamHI, PET66) was cloned in pN-TAPa and PET46 (NcoI/XhoI, PET68) and PET65 (NcoI/XhoI, PET67) in pC-TAPa. The complete inserts of all clones and the fusion boundaries with the epitopes in the final clones were sequenced. Clones were transformed via electroporation into the Escherichia coli strain XL1-blue (Stragene) and afterwards in the Agrobacterium tumefaciens strain C58C1 (pMP90). The plasmids were introduced in the Arabidopsis Colombia using the floral dip method (Clough and Bent, 1998). The nucleotide sequences of the primers used in this experiment are listed in table 1. As a control for the TAPa system, we kindly received N-TAPaGFP seeds from Alois Schweighofer (Max-Planck-Institute of Molecular Plant Physiology). Fast plant protein extraction method The fast plant protein extraction method was used to screen individual transformants. Tissue was ground in liquid nitrogen and transferred into a frozen 1.5 ml eppendorf tube to a volume of ~ 50 μl. To the tissue, 100 μl of 1 x SDS sample buffer (0.0625 M Tris-HCl, 2% SDS, 10% glycerol, 0.15 M β-mercapto-ethanol, 0.025% bromophenol blue, pH adjusted to 6.8 with HCl) was added. After vigorous mixing, the extract was boiled for 10 min. After a short centrifugation, 10 μl of extract was loaded on a 12.5 % (w/v) SDS PAGE gel. Protein extracts were stored at -20 ºC. Western analysis, immuno-blotting and detection Proteins were separated by SDS-PAGE and subsequently transferred to Immobilon PVDF membrane (Millipore) using the BIO-RAD semi-dry blotting system. In case of using the peroxidase anti peroxidase antibody (PAP, Sigma) and α-myc antibodies, blots were blocked first for 1 hour in 5% milk powder (Marvel) in 1x TBST (20 mM Tris-HCl at pH 7.6, 140 mM NaCl and 0.5 ml Tween20). Blots incubated with the anti calmodulin binding protein antibody (CBP, Upstate) were blocked first o/n. After blocking, blots were briefly washed in 1x TBST and incubated with the PAP antibody (1/7500, 5% milk) or the a-myc antibody (1/2000, 5% milk in 1x TBST) overnight at 4°C or with the CBP antibody (1/5000 in 5% milk) for 1 hr at room temperature. Blots immunoprobed with CBP were incubated with the secondary anti-rabbit IgG-peroxidise antibody (1/5000, Promega) and those immunoprobed with α-myc with the secondary anti-mouse IgG- peroxidase antibody (1/7500, Promega) or anti-mouse IgG-AP (1/7500, Promega) for 1 hour in 1x TBST and 5 % milk. After washing, proteins were visualized according the manufacturers protocol by either using the Amersham ECL kit or the BCIP/NBT color development substrate (Promega). 75 Chapter 3 TAPi purification method The TAPi purification was based on the protocol described by Rohila et al. (2004). Leaf tissue (15 gram, fresh weight) was ground in liquid nitrogen and thawed in ~25 ml TAPi buffer (20 mM Tris HCl pH 8.0, 150 mM NaCl, 2.5 mM EDTA, 20 mM NaF, 10 mM βmercaptoethanol, 0.1% IGEPAL and 1x complete protease inhibitor cocktail EDTA free from Roche), filtered through miracloth (Calbiochem), and centrifuged for 10 min at 4 °C. The supernatant was incubated with 500 µl washed IgG beads (IgG Sepharose 6 Fast Flow beads, GE Healthcare) for 2 hrs at 4 °C with gentle rotation. The beads were loaded onto a disposable column (Altech Extract-Clean Reservoirs) and washed three times with 10 ml TAPi buffer lacking protease inhibitors and once with 10 ml of cleavage buffer (TAPi buffer with 1 mM DTT). Next a cleavage step was performed for 2 hrs at 16 °C with gentle rotation, in which the IgG beads were incubated with 0.5 ml cleavage buffer and 100 U AcTEV protease (Intvitrogen) and 1 µM E-64 protease inhibitor (Calbiochem). The supernatant was recovered by centrifugation and mixed with 100 µl washed calmodulin beads (Stratagene), 3 volumes of calmodulin binding buffer (10 mM Tris HCl pH 8.0, 150 mM NaCl, 1 mM MgAc, 1 mM imidazole, 2 mM CaCl2, 10 mM β-mercaptoethanol, 0.1% IGEPAL) and 2 mM CaCl2. The beads were washed three times with calmodulin binding buffer for 5 min and eluted two times with 0.5 ml calmodulin elution buffer (10 mM Tris HCl pH 8.0, 150 mM NaCl, 1 mM MgAc, 1 mM imidazole, 10 mM β-mercaptoethanol, 0.1% IGEPAL and 10 mM EGTA). The eluates were concentrated to 50 µl with help of an Amicon centrifugal filter unit (Millipore) separated on a 12.5% SDS-PAGE gel, blotted and immunoprobed with PAP or CBP antibody. To test cleavage of the fusion proteins (Fig. 4), total protein was isolated from 7.5 gram leaf tissue. Supernatant was collected filtered through miracloth (Calbiochem), and centrifuged for 10 min at 4 °C. The extract was incubated with 150 µl washed IgG beads for 2 hrs at 4 °C with gentle rotation. The beads were washed and 0.5 ml TEV buffer was added. Prior to the addition of 100 U AcTEV protease and 1 µM E-64 a sample was taken (0 min) The cleavage step was performed at 16 °C with gentle rotation and every 15 minutes a sample was taken. The samples were loaded on a 12.5% SDS-PAGE gel, blotted and incubated with PAP or CBP antibodies. 76 Towards identification of group II ASKs containing complexes in Arabidopsis Reverse TAPi purification method Leaf tissue (15g, fresh weight) was grind in liquid nitrogen and thawed in ~25 ml calmodulin binding buffer (10 mM Tris HCl pH 8.0, 150 mM NaCl, 1 mM MgAc, 1 mM imidazole, 2 mM CaCl2, 10 mM β-mercaptoethanol, 0.1% IGEPAL and 1x complete protease inhibitor cocktail EDTA free from Roche), filtered through miracloth (Calbiochem), and centrifuged for 10 min at 4 °C. The supernatant was incubated with 500 µl washed calmodulin beads (Stratagene) for 2 hrs at 4 °C with gentle rotation. The supernatant including the beads were loaded onto a disposable column (Altech ExtractClean Reservoirs) and washed five times with 10 ml calmodulin binding buffer. Next, the beads were incubated for 5 min with 2 ml calmodulin elution buffer (10 mM Tris HCl pH 8.0, 150 mM NaCl, 1 mM MgAc, 1 mM imidazole, 10 mM β-mercaptoethanol, 0.1% IGEPAL and 10 mM EGTA). The eluate was collected and the elution step was repeated. To the eluate (4 ml) 20 mM NaF, 2 mM EDTA and 1x complete protease inhibitor cocktail (EDTA free) was added. After addition of 400 µl washed IgG beads (IgG Sepharose 6 Fast Flow beads, GE Healthcare) the mixture was rotated for 2 hrs at 4 °C. The beads were loaded onto a column (Altech Extract-Clean Reservoirs) and washed three times with 10 ml TAPi buffer (20 mM Tris HCl pH 8.0, 150 mM NaCl, 2.5 mM EDTA, 20 mM NaF, 10 mM β-mercaptoethanol, 0.1% IGEPAL and 1x complete protease inhibitor cocktail EDTA free). The beads were incubated for a few minutes with 2 ml low pH elution buffer (0.1M glycine-HCl pH 3.0) and this step was repeated. The eluate was concentrated to 50 µl on an Amicon centrifugal filter unit (Millipore), separated on a 12.5% SDS-PAGE gel, blotted and immuno-detected using an PAP or CBP antibody. TAPa purification method The alternative TAP was based on the protocol described by Rubio et al (2005). Leaf tissue (15 gram fresh weight) was ground in liquid nitrogen and thawed in ~25 ml TAPa buffer (50 mM Tris HCl pH 7.5, 150 mM NaCl, 10% glycerol, 0.1% IGEPAL and 1x complete protease inhibitor cocktail EDTA free from Roche), filtered through miracloth (Calbiochem), and centrifuged for 10 min at 4 °C. The extracts were incubated with 500 µl washed IgG beads (IgG Sepharose 6 Fast Flow beads, GE Healthcare) for 2 hrs at 4 °C with gentle rotation. After centrifugation at 300 rpm for 2 min at 4 °C, the IgG beads were recovered and washed three times with 10 ml TAPa buffer without protease inhibitors and once with 10 ml of cleavage buffer (TAPa buffer with 1 mM DTT). Next, a cleavage step was performed for 2 hrs at 4 °C with gentle rotation, in which the IgG beads were incubated 77 Chapter 3 with 5 ml cleavage buffer and 25 µl (50 U) 3C protease (PreScission Protease, GE Healthcare). The supernatant was recovered by centrifugation and mixed with 1 ml washed Ni-Nt agarose beads (Qiagen). The beads were washed three times with 10 ml TAPa buffer without protease inhibitors and the eluate was recovered by adding 5 ml of imidazole containing buffer (TAPa buffer with 0.05 M imidazole), concentrated to 50 µl by using an Amicon centrifugal filter unit (Millipore), separated on a 12.5% SDS-PAGE gel and immuno-detected using a PAP or α-myc antibody. Table 1. Description of primers used for PCR Primer Gene Sequence 5´-3´ 121-FW 122-RV M547-RV 101-FW 102-RV 123-RV M544-FW M551-RV ASKζ ASKζ ASKζ GSK1 GSK1 GSK1 GSK1 GSK1 GCCCATGGCTTCGATACCATG GGGATCCCTAGGG TCCAGCTTGAAATG GGGGTCGACGGTCCAGCTTGAAATGGAAAG CCCCATGGCCTCATTACCATTG GGGATCCTTAACTGTTTTGTAATCCTGTG GGGATCCAGGGTCCAGCTTGAAATGG GGCCATGGCTCATTACCATTGGGG GGCTCGAGTTAACTGTTTTGTAATCCTG Acknowledgements We thank Maartje Kuijpers for plant care. This work was supported by a grant of the Netherlands Organisation for Scientific Research (NWO). References Batelli, G., Verslues, P.E., Agius, F., Qiu, Q., Fujii, H., Pan, S., Schumaker, K.S., Grillo, S. and Zhu, J.K. (2007) SOS2 promotes salt tolerance in part by interacting with the vacuolar H+-ATPase and upregulating its transport activity. Mol Cell Biol, 27, 7781-7790. Clough, S.J. and Bent, A.F. (1998) Floral dip: a simplified method for Agrobacterium-mediated transformation of Arabidopsis thaliana. Plant J, 16, 735-743. He, J.X., Gendron, J.M., Yang, Y., Li, J. and Wang, Z.Y. (2002) The GSK3-like kinase BIN2 phosphorylates and destabilizes BZR1, a positive regulator of the brassinosteroid signaling pathway in Arabidopsis. Proc Natl Acad Sci U S A, 99, 10185-10190. Kim, T.W., Guan, S., Sun, Y., Deng, Z., Tang, W., Shang, J.X., Burlingame, A.L. and Wang, Z.Y. (2009) Brassinosteroid signal transduction from cell-surface receptor kinases to nuclear transcription factors. Nat Cell Biol, 11, 1254-1260. Li, J. and Chory, J. (1997) A putative leucine-rich repeat receptor kinase involved in brassinosteroid signal transduction. Cell, 90, 929-938. Li, J. and Nam, K.H. (2002) Regulation of brassinosteroid signaling by a GSK3/SHAGGY-like kinase. Science, 295, 1299-1301. 78 Towards identification of group II ASKs containing complexes in Arabidopsis Li, J., Wen, J., Lease, K.A., Doke, J.T., Tax, F.E. and Walker, J.C. (2002) BAK1, an Arabidopsis LRR receptor-like protein kinase, interacts with BRI1 and modulates brassinosteroid signaling. Cell, 110, 213-222. Nam, K.H. and Li, J. (2002) BRI1/BAK1, a receptor kinase pair mediating brassinosteroid signaling. Cell, 110, 203-212. Puig, O., Caspary, F., Rigaut, G., Rutz, B., Bouveret, E., Bragado-Nilsson, E., Wilm, M. and Seraphin, B. (2001) The tandem affinity purification (TAP) method: a general procedure of protein complex purification. Methods, 24, 218-229. Rohila, J.S., Chen, M., Cerny, R. and Fromm, M.E. (2004) Improved tandem affinity purification tag and methods for isolation of protein heterocomplexes from plants. Plant J, 38, 172-181. Rohila, J.S., Chen, M., Chen, S., Chen, J., Cerny, R., Dardick, C., Canlas, P., Xu, X., Gribskov, M., Kanrar, S., Zhu, J.K., Ronald, P. and Fromm, M.E. (2006) Protein-protein interactions of tandem affinity purificationtagged protein kinases in rice. Plant J, 46, 1-13. Rozhon, W., Mayerhofer, J., Petutschnig, E., Fujioka, S. and Jonak, C. (2010) ASKtheta, a group-III Arabidopsis GSK3, functions in the brassinosteroid signalling pathway. Plant J, 62, 215-223. Rubio, V., Shen, Y., Saijo, Y., Liu, Y., Gusmaroli, G., Dinesh-Kumar, S.P. and Deng, X.W. (2005) An alternative tandem affinity purification strategy applied to Arabidopsis protein complex isolation. Plant J, 41, 767778. Tang, W., Kim, T.W., Oses-Prieto, J.A., Sun, Y., Deng, Z., Zhu, S., Wang, R., Burlingame, A.L. and Wang, Z.Y. (2008) BSKs mediate signal transduction from the receptor kinase BRI1 in Arabidopsis. Science, 321, 557560. Terpe, K. (2003) Overview of tag protein fusions: from molecular and biochemical fundamentals to commercial systems. Appl Microbiol Biotechnol, 60, 523-533. Van Leene, J., Witters, E., Inze, D. and De Jaeger, G. (2008) Boosting tandem affinity purification of plant protein complexes. Trends Plant Sci, 13, 517-520. Vert, G. and Chory, J. (2006) Downstream nuclear events in brassinosteroid signalling. Nature, 441, 96-100. Wang, X. and Chory, J. (2006) Brassinosteroids regulate dissociation of BKI1, a negative regulator of BRI1 signaling, from the plasma membrane. Science, 313, 1118-1122. Witte, C.P., Noel, L.D., Gielbert, J., Parker, J.E. and Romeis, T. (2004) Rapid one-step protein purification from plant material using the eight-amino acid StrepII epitope. Plant Mol Biol, 55, 135-147. Xinga, D. and Chen, Z. (2006) Effects of mutations and constitutive overexpression of EDS1 and PAD4 on plant resistance to different types of microbial pathogens Plant Science, 171, 251-262. Yan, Z., Zhao, J., Peng, P., Chihara, R.K. and Li, J. (2009) BIN2 Functions Redundantly with Other Arabidopsis GSK3-Like Kinases to Regulate Brassinosteroid Signaling. Plant Physiol, 150, 710-721. Yin, Y., Wang, Z.Y., Mora-Garcia, S., Li, J., Yoshida, S., Asami, T. and Chory, J. (2002) BES1 accumulates in the nucleus in response to brassinosteroids to regulate gene expression and promote stem elongation. Cell, 109, 181-191. 79 Chapter 4 Brassinosteroid biosynthesis and signalling in petunia Nathalie Verhoef, Michiel Vandenbussche, Gert-Jan de Boer, Tom Gerats, Ronald Koes and Erik Souer Brassinosteroid biosynthesis and signalling in petunia Abstract Brassinosteroids (BRs) are steroidal plant hormones that play an important role in the growth and development of plants. The biosynthesis of sterols and BRs and the signalling cascade they induce in plants has been elucidated largely through metabolic studies and the analysis of mutants in Arabidopsis and rice. Only fragmentary details about BR signalling in other plant species are known. Here we used a forward genetics strategy in petunia, by which we identified 19 petunia families with phenotypic alterations typical for BR deficiency mutants. In five families we disclosed the tagged genes as the petunia BR biosynthesis gene CONSTITUTIVE PHOTOMORPHOGENESIS AND DWARF (CPD) and the BRASSINOSTEROID INSENSITIVE 1 (BRI1) receptor. In addition, we cloned several homologs of key regulators of the BR signalling pathway from petunia based on homology with their Arabidopsis counterparts, including the BRI1 receptor, a member of the BRI1EMS-SUPPRESSOR 1 (BES1) / BRASSINOZOLE RESISTANT 1 (BZR1) transcription factor family (PhBEH2) and two GSK3-like kinases (PSK8 and PSK9). PhBEH2 was shown to interact with PSK8 and 14-3-3 proteins in yeast revealing similar interactions as during BR signalling in Arabidopsis. Interestingly, in yeast PhBEH2 also interacted with proteins implicated in other signalling pathways. This suggests that PhBEH2 might function as an important hub in the cross talk between diverse signalling pathways. Introduction The importance of steroidal hormones in mammalian development is already known for almost eight decades. The idea that plant steroids have hormonal functions arose when scientists reported the isolation of brassinolide, a steroid with strong plant growthpromoting activity, from bee-collected pollen of Brassica napus L (Grove et al., 1979). The identification of several brassinosteroid (BR) biosynthesis- and signalling mutants helped to some extent to unravel the BR biosynthesis and signalling pathway and nowadays BRs are accepted as the sixth class of plant phytohormones. BRs are widely distributed among the plant kingdom and so far they have been found in gymnosperms, monocots, dicots, a pteridophyte, a bryophyte and a chlorophyte (Bajguz and Tretyn, 2003). BRs are involved in various growth- and developmental processes such as cell division and elongation, vascular differentiation, abiotic- and biotic stresses, photomorphogenesis, germination, senescence and fertility (Clouse et al., 1996; Li et al., 1996; Szekeres et al., 1996; Bajguz and Hayat, 2009). 83 Chapter 4 Epoxidase SMT1 Lupeol synthase Squalene-2,3epooxide Squalene Mevalonate 24-Methylenecycloartenol Cycloartenol CYP51G1 CYP710A Stigmasterol DIM Sitosterol DWF5 DWF7 5-Dehydroavenasterol Isofucosterol Cycloeucalenol δ 7-Avenasterol Cyclopropyl isomerase Obutusifoliol 24-Ethylidenelophenol membranes CYP51G1 δ8,14-Sterol SMT2 FK 24-Methylenecholestrol DWF5 5-Dehydroepisterol DWF7 24-Methylene lophenol Episterol HYD1 4α-Methylfecosterol DIM Campesterol Campest-4-en3β-one DWF4 22α-Hydroxy-5− campesterol Campest-4en-3-one DWF4 SAX1 DET2 5α-Campestan- 6-Oxocampestenol DWF4 DWF4 DET2 22α-Hydroxy-5α22α-Hydroxycampest-4-en-3-one campestan-3-one DWF4 6-Deoxocathasterone CYP90 C1/D1 CYP90C1/D1 6α-Hydroxy campestenol Campestenol 3-one Cathasterone CYP90C1/D1 6-Deoxoteasterone CYP90C1/D1 CYP90C1/D1 CYP85A1/A2 Teasterone 3-Epi-6-deoxocathasterone 22,23-Dihydroxy campesterol 22,23-Hydroxycampest-4-en-3-one 3-Dehydro-6-deoxoteasterone CYP85A1/A2 3-Dehydroteasterone CYP85A2 7-Oxateasterone CYP90C1/D1 6-Deoxotyphasterol 6-Deoxycatasterone CYP85A1/A2 6α-Hydroxycatasterone CYP85A1/A2 Typhasterol CYP85A2 7-Oxatyphasterol Catasterone CYP85A2 Brassinolide Figure 1. The proposed sterol/brassinosteroid biosynthesis pathway in Arabidopsis thaliana. The metabolic products are shown in boxes and the enzymes involved are depicted in red. Mevalonate is converted via multiple steps into squalene (black dashed arrow). The light green arrows represent the early C-22 oxidation pathway. The late C-6 oxidation pathway is visualized by red arrows and the early C-6 oxidation pathway by dark green arrows. The sterol biosynthesis pathway can be divided into three domains (Fig. 1). In the first part, the mevalonate derivative squalene is converted via multiple steps into 24-methylene lophenol. Subsequently, the pathway diverges into two downstream branches, one branch produces cellular membrane components, while in the second branch brassinosteroids are synthesized from the precursor campesterol (Lindsey et al., 2003). Detailed studies on the BR biosynthesis route in Catharanthus roseus and Arabidopsis thaliana led to the identification of two parallel branched routes, better known as the early and late C-6 oxidation pathways. Recent studies demonstrated that these two pathways are linked to each other, implying that they are not totally autonomous. Although the early and late C-6 oxidation pathways are found throughout the plant kingdom, it seems that for species such as tomato and tobacco the late C-6 oxidation pathway is more prevalent (Nomura et al., 2001). Upstream of the C-6 oxidation pathways Fujioka et al. (2002) revealed an early C- 84 Brassinosteroid biosynthesis and signalling in petunia 22 oxidation pathway. It thus seems that, due to the existence and intertwinement of several biosynthesis pathways, plants are able to synthesize BRs in several ways. A number of sterol and BR biosynthesis mutants have been described, predominantly in Arabidopsis, and many of the genes have been cloned and studied (Choe et al., 1998; Klahre et al., 1998; Mathur et al., 1998; Choe et al., 1999a; Choe et al., 2001; Clouse, 2002b). BR biosynthesis mutants and some sterol mutants such as dwarf7 (dwf7), diminuto (dim) and dwarf5 (dwf5) can be rescued by exogenous application of BR. However, mutants that are affected in genes functioning in early steps of the sterol biosynthesis pathway such as FACKEL (FK), STEROL METHYL TRANFERASES 1 & 2 (SMT1, SMT2) and HYDRA1 (HYD1) are embryo lethal and can not be rescued by application of exogenous BR. Presumably, the absence of these sterols interferes with membrane integrity. BR signalling has largely been elaborated by studies in Arabidopsis. Yet, it is still unclear to what extent these findings can be extrapolated to other plant species. The currently known BR signalling pathway in Arabidopsis runs from membrane to nucleus (Kim et al., 2009). Despite the identification of these key components, it seems unlikely that all components of the Arabidopsis BR signalling pathway are identified. Moreover, how the well-known crosstalk between BR signalling and other hormone signalling pathways is achieved is not completely clear. In Arabidopsis, BRs are perceived by the plasma membrane located leucine-rich repeat receptor-like kinase BRASSINOSTEROID INSENSITIVE 1 (BRI1) that dimerizes with the co-receptor BRI1-ASSOCIATED RECEPTOR KINASE 1 (BAK1) thereby inducing transphosphorylation. BRI1 then dissociates from its repressor BRI1 KINASE INHIBITOR 1 (BKI1, Li and Chory, 1997; Li et al., 2002; Nam and Li, 2002; Russinova et al., 2004; Wang et al., 2008a). Subsequently, BRI1 phosphorylates BR SIGNALING KINASE 1 (BSK1), which in turn phosporylates and activates the plant-specific serine/threonine protein phosphatase BRI1 SUPPRESSOR1 (BSU1, Tang et al., 2008). BSU1 inactivates a negative regulator of BR signalling, the Arabidopsis GSK3/SHAGGY-like kinase BRASINOSTEROID INSENSITIVE 2 (BIN2) by dephosphorylation at pTyr200 (Kim et al., 2009). BIN2 performs its negative effect on BR signalling by phosphorylating two closely related plant-specific transcription factors BRI1-EMS-SUPPRESSOR 1 (BES1) and BRASSINOZOLE RESISTANT 1 (BZR1) resulting in inhibition of their activity through proteasome mediated degradation, cytoplasmic retention by 14-3-3 proteins and inhibition of their DNA-binding and transcriptional activities (Li et al., 2001; He et al., 2002; Yin et al., 2002; Vert and Chory, 2006; Gampala et al., 2007; Ryu et al., 2007; Peng et al., 2010). Interestingly, other Arabidopsis GSK3-like protein kinases (ASKs) play a redundant role with BIN2 in the BR signalling pathway (Chapter 2; Vert and Chory, 2006; Kim et al., 85 Chapter 4 2009; Yan et al., 2009; Rozhon et al., 2010). In other plant species such as Brassica napus, Medicago sativa, Nicotiana tabacum, Oryza sativa, Gossypium hirsutum and Petunia hybrida several GSK3-like kinases have been found. In Oryza sativa and Gossypium hirsutum group II GSK3-like kinases, to which BIN2 belongs, have been described although their role in the BR signalling pathway still need to be demonstrated (Sun et al., 2004; Sun and Allen, 2005; Koh et al., 2007). To reveal to which extent the biosynthesis of steroids and BRs and the signalling cascades downstream in plants are conserved we browsed our petunia collection for mutants that exhibit a typical BR deficiency phenotype. Here, we describe the identification and partial characterisation of a number of these petunia mutants. To reveal if our current knowledge on BR biosynthesis and signalling can be extrapolated to petunia we initiated an analysis of BR signalling in petunia. We isolated several petunia homologs of key components of the BR signalling pathway in Arabidopsis and determined if a similar network of interaction between these factors can occur. In addition, we identified several components of other hormone pathways that interacted with the petunia homolog of a member of the BES1/BZR1 family of transcription factors. These interactions reveal new mechanisms, not recognised from Arabidopsis research that suggest extensive cross talk between BR signalling and other hormone signalling pathways. Results Identification of petunia sterol/BR biosynthesis and signalling mutants In order to study BR biosynthesis and signalling in petunia we visually screened populations of the line W138. This petunia line contains more than 200 copies of the active transposable element dTph1, leading to a high incidence of mutations (Gerats et al., 1990; van Houwelingen et al., 1998). We identified four mutants (Table 1) that phenocopy sterol/BR biosynthesis and signalling mutants from other species. Apart from these four, we received 15 mutant families with similar phenotypes that were identified in a running large scale transposon tagging experiment using petunia W138 at Enza Zaden (Enkhuizen) (Table 1). Due to their appearance with reduced stature and dark round shiny leaves we named these mutants compact disc (cd). In order to identify whether these mutants are BR biosynthesis or signalling mutants, individual mutant plants were sprayed 2-3 times a week with 1 µM 24-epibrassinolide or with a control solution under standard greenhouse conditions. Out of 19 mutant families, 18 clearly responded to spraying with 24-epibrassinolide by elongation of internodes, petioles and leaves (Fig. 2). Complementation of the mutant phenotype in these plants indicates that 86 Brassinosteroid biosynthesis and signalling in petunia these are sterol/BR biosynthesis mutants. Only one mutant family, cd10P2020, did not respond to exogenous application of 24-epibrassinolide and thus likely harbours a dTph1 insertion in a BR signalling gene. Table 1 Overview of cd alleles and results of complementation experiment by brassinolide treatment Locus Allele Origin Complemented by BRa cd1 cd1-W503 VU yes cd1-K2078 Nijmegen yes cd2-W2254 VU yes cd2-P2013 ENZA yes cd2-P2015 ENZA yes cd2-P2018 ENZA yes cd3-D2213 VU yes cd3-P2014 ENZA yes cd9-P2016 ENZA yes cd9-P2019 ENZA yes cd9-P2021 ENZA yes cd9-P2022 ENZA yes cd9-P2023 ENZA yes cd9-P2024 ENZA yes cd9-P2025 ENZA yes cd9-P2026 ENZA yes cd9-P2027 ENZA yes cd9-P2028 ENZA yes cd9-P2016 ENZA yes cd10-P2020 ENZA no cd2b cd3 cd9c cd10 a b Complementation was tested by spraying with 1 µM 24-epibrassinolide. Due to reduced fertility, cd2 allelism was confirmed molecularly as independent insertions in PhCPD (see text). c cd9P2016, cd9P2019, cd9P2022, cd9P2024, cd9P2025, cd9P2026, cd9P2027 and cd9P2028 are derived from a common ancestor and thus might represent identical alleles. Cross-pollination between the mutant families revealed that the 19 mutants could be assigned to five different complementation groups (see Table 1). Ten out of 16 families received from Enza Zaden harbour a dTph1 insertion in the same BR biosynthesis gene, CD9. As some of these families are derived from a common ancestor (see Table 1) it is likely that they represent only three different alleles. As the mutated gene in cd2W2254 is known (see below), we were able to detect independent transposon insertions in the same gene in cd2P2013, cd2P2015 and cd2P2018 and consider these allelic. As can be derived from Table 1, the total number of sterol/BR biosynthesis genes hit is four, while the remaining mutant cd10 represents a signalling mutant. All mutants have a short robust stature with thick, small, round and dark-green leaves. The severity of the dwarfish phenotype slightly differed among the complementation groups. The cd10 mutants remained extremely small and did not flower after staying for months in the greenhouse. 87 Chapter 4 cd2 mutants were intermediate in size and occasionally flowered. The growth of the cd3 and cd9 mutants was clearly slower compared to wild type, but they ultimately flowered and fertility appeared to be normal. In comparison to cd3 and cd9, cd1 mutants had a more severe phenotype and were less fertile. - + - P2013 + P2023 - + Brassinolide P2020 Figure 2. Effect of 24-epibrassinolide on BR biosynthesis/signalling mutants. Plants were sprayed with 1 µM 24-epibrassinolide (+) or with a control solvent (-). Two BR biosynthesis mutants cd2 shown. The signalling mutant cd10 P2020 P2013 and cd9 P2023 are did not respond to a brassinolide treatment. Isolation of BR biosynthesis genes We presumed that at least the majority of our cd mutants are insertions in genes known from research on BR mutants in other species. Therefore, a candidate gene approach might lead to identification of the mutated genes. As our knowledge about the molecular players involved in BR biosynthesis in petunia is blank we started out to clone gene fragments of known BR biosynthesis genes. Using a combination of database mining and primer design based on sequences of Solanaceae species (eg. tomato, tobacco, potato etc. Supplemental data, Table 1) we were able to clone gene and/or cDNA fragments of petunia homologs of the Arabidopsis BR biosynthesis genes CPD, DWF5, CYP85A, DIM, DWF7, SMT2, DETIOLATED2 (DET2), DWARF4 (DWF4) and CYP90C1 (ROT3) and a number of signalling genes (see below). Four BR biosynthesis mutant families are cpd mutants Most mutant alleles in W138 are due to insertions of dTph1 transposons. We used the obtained sequence data of petunia BR biosynthesis genes to determine in each cd mutant if 88 Brassinosteroid biosynthesis and signalling in petunia 13 20 2 P 15 2 0 W ild 18 -ty pe PhCPD 2P D P2013 cd 2P cd -4 * 94 94 21 21 2M 2M cd cd cd 20 C A -4 ~ any of these known genes is disrupted by a dTph1 insertion. A PCR was performed on DNA isolated from these mutants using gene-specific and/or transposon specific primers When DNA isolated from cd2W2254 (Fig. 3a) and three mutants from Enza Zaden (cd2P2013, P2015, P2018 ) was used as a template, these PCRs identified dTph1 insertions at different sites in the petunia homolog of the BR biosynthesis gene CPD (Fig. 3c,d,e). The dTph1 transposon insertion in allele cd2W2254 was found in the first exon, about 80 nucleotides after the ATG. The other three mutants (cd2P2013, P2015, P2018) carry a dTph1 insertion at different sites in the second exon (Fig. 3d). Plants in progeny of cd2W2254 occasionally gave rise to revertant shoots (Fig. 3b). PCR and sequence analysis on genomic DNA extracted from a revertant shoot (cd2M2194-4) revealed excision of dTph1 from CPD, leaving a 6 bp footprint (Fig. 3c,d,e). Furthermore, a stable cpd mutant (cd2M2194-1) was identified in which excision of dTph1 created an 8bp footprint (Fig. 3d,e). dTph1 insertion W2254 P2015 E P2018 200 bp cd2W2254 Wild-type cd2K2056 cd2M2194-1 cd2M2194-4~ TCTCCTCCTCCG TGCA TCTCCTCCTCCG------dTph1-------CTCCTCCGTGCA TCTCCTCCTCCC GTCCTCCGTGCA TCTCCTCCTAG TCCTCCGTGCA + 8bp + 6bp cd2P2013 B revertant shoot Wild-type cd2P2013 CCATGCGAGTGG ACTG CCATGCGAGTGG------dTph1-------GCGAGTGGACTG cd2P2015 Wild-type cd2P2015 ATATGTGCAGAT AACG ATATGTGCAGAT------dTph1-------GTGCAGATAACG cd2P2018 dwarf part Wild-type cd2P2018 CCTACCGCAAGG CCAT CCTACCGCAAGG------dTph1-------CCGCAAGGCCAT Figure 3. Four BR biosynthesis mutant families are defective in the CPD gene. A) Mutant phenotype of a W2254 W2254 M2194-4 plant B) A cd2 plant that develops a revertant shoot (cd2 ) C) PCR with CPD specific primers cd2 on genomic DNA extracted from plants carrying various cpd alleles. D) Schematic drawing of CPD and mutant alleles. Boxes represent exons and the thin lines represent introns. The triangles indicate the dTph1 transposon insertions in the indicated alleles. E) Sequence analysis of the dTph1 insertions in various cpd alleles. The red sequence indicates the target site duplication (TSD). * indicates dwarf part and ~ the revertant shoot. For the other three loci cd1, cd3 and cd9 we were unable to detect transposon insertions in the putative petunia homologs of DWARF5, CYP85, DET3, DIM, DWARF7, SMT2, DET2, DWARF4 or ROT3. It must be said that we currently lack the complete sequences of these 89 Chapter 4 genes. As we miss 3’ sequences (DIM), 5’ sequences (DWARF5, CYP85, DWARF7, SMT2) or both (DET3, DET2, DWARF4, ROT3) it cannot be excluded that there are dTph1 insertions present in these parts of the gene. Identification of mutant alleles of BR biosynthesis genes by reverse genetics To identify other mutants in petunia homologs of BR biosynthesis genes from Arabidopsis we screened the 3D indexed petunia insertion database as described by Vandenbussche et al. (2008). A mutant plant was identified that harbours a dTph1 insertion in the first exon of a gene homologous to the Arabidopsis sterol biosynthesis gene DIM. Remarkably, in progeny, plants homozygous for the mutant dimL2006 allele could not be identified. All seeds seemed to germinate when sown on agar plates and no aberrant plantlets could be identified in progeny. Furthermore, reciprocal crosses revealed that the allele is passed on via the mother and father. At the moment it is unclear whether DIM interferes with seed development or that the DIM gene is linked to a lethal allele of another gene. Another possible explanation could be that DIM is recently duplicated in petunia and thus our PCR primers amplify two genes. Apart from this we also identified insertions in genes homologous to the Arabidopsis BR biosynthesis genes DET2 and CYP85A1 (BR6OX1). The insertion in phcyp85A1L2228 is located in intron 2. As no phenotypic alterations are seen in plants homozygous for the insertion allele, we presume that the small dTph1 element is spliced from the pre-mRNA along with intron 2. The insertion in the PhDET2 gene (phdet2L2224) is located in exon 1. However, also in this case plants homozygous for the insertion allele do not show any morphological aberrations, suggesting that DET2 is redundant in petunia. Isolation of a BRI1 homolog from petunia In Arabidopsis, BRs are perceived by the leucine-rich repeat receptor-like kinase BRI1 (Li and Chory, 1997). BRI1 orthologs have been identified in species such as tomato (CU-3), pea (LKA), rice (OsBRI1), barley (HvBRI1) and cotton (GhBRI1) (Yamamuro et al., 2000; Montoya et al., 2002; Chono et al., 2003; Nomura et al., 2003; Sun et al., 2004). From the Petunia hybrida EST collection we retrieved a sequence that showed high similarity to the 3´ end of BRI1 (GenBank accession number CV294320). The full-length mRNA sequence of W138 PhBRI1 was obtained via RACE-PCR and somatic insertion-mediated-PCR (SOTI-PCR, see Materials and Methods). Comparison of the derived amino acid sequence of PhBRI1 with AtBRI1, OsBRI1, LeBRI1 and NbBRI1 revealed 63,3%, 53,4%, 84,1% and 91,3% identity, respectively (Fig. 4a). Similar to the other identified BRIs, PhBRI1 has 90 Brassinosteroid biosynthesis and signalling in petunia an Leucine-Rich Repeat (LRR) domain, a putative leucine-zipper motif, a transmembrane domain to anchor the protein in the plasma membrane, and a cytoplasmic kinase domain (Fig. 4b and Supplementary data Fig. 1). The BRI1 extracellular domain of petunia contains 25 tandem amino-terminal LRRs, which is interrupted by a 68 amino acid non-repetitive island between the 21st and 22nd LRR. Identification of a petunia mutant defected in BRI1 As described above, only one of the identified mutant families (cd10 P2020) did not respond to exogenous application of 24-epibrassinolide and thus most likely harboured a dTph1 insertion in a BR signalling gene (Fig. 2a). One of the prime candidates for being hit in cd10P2020 is the homolog of the Arabidopsis BR receptor, BRI1. Indeed, PCR analysis on genomic DNA with numerous PhBRI1 specific primers resulted in a larger PCR fragment in cd10 plants in comparison to wild type when using primers that amplified the 5’ part of the gene (Fig. 4c). Sequence analysis of the 5’ part of PhBRI1 in cd10 revealed that the dTph1 transposable element was inserted in the 3rd leucine-rich repeat sequence motif (Fig. 4b, c). A B PhBRI1 A tC LV 1 Putative Leucine zipper motif 68 amino acid (aa) Transmembrane domain non-repetitive island kinase domain Leucine-rich repeats (LRR) 1-21 LRR 22-25 H vB R I1 1000 O sB R I1 A tB R I1 100 aa 968 Cd10 P hB R I1 1000 N bB R I1 C W 1000 13 cd 8 1 cd 0 P2 10 020 P2 02 4 05 P sB R I1 dTph1 insertion 906 N tB R I1 D Wild-type cd10P2020-4 TAATGATGTTTC CAGC TAATGATGTTTC--dTph1--GATGTTTCCAGC BRI1 1000 LeB R I1 0,1 Figure 4. Characterization of the BRI1 receptor from petunia. A) Phylogenetic tree constructed using derived amino acid sequences of the BRI1 homolog isolated from petunia and various other species (At: Arabiopsis thaliana; Hv: Hordeum vulgare; Os: Oryzia sativa; Le: Lycoperscicon esculentum; Nb: Nicotiana benthamiana; Nt: Nicotiana tabacum; Ps: Pisum sativa and Ph: Petunia hybrida) The derived amino acid sequence of the leucine-rich repeat receptor-like kinase CLAVATA1 from Arabidopsis was used as an outgroup. Genbank accession numbers are recited in Supplemental Table 2. B) Schematic drawing of PhBRI1 with its predicted functional domains. The triangle indicates the dTph1 transposon insertion in cd10. C) PCR with BRI1 specific primers on genomic DNA extracted from cd10P2020and wild type plants. D) Sequence analysis of the dTph1 insertion in cd10. The red sequence indicates the target site duplication (TSD). 91 Chapter 4 Identification of a BEH2 homolog in petunia To further explore the BR signalling cascade in petunia, we were aiming to identify signalling components downstream of PhBRI1. In Arabidopsis, the BES1/BZR1 transcription factor family plays an important role in the regulation of BR-responsive genes (Zhao et al., 2002). A PCR-based approach was performed to identify potential petunia homologs. Primers were developed based on the sequences of all six BES1/BZR1 family members of Arabidopsis and a nested RACE-PCR on petunia leaf cDNA yielded a PCR fragment of the desired size. Sequence analysis revealed high sequence identity with members of the BES1/BZR1 family. The 5´ end was isolated by SOTI-PCR (see material and methods, data not shown). The derived amino acid sequence has highest sequence identity to the BES1/BZR1 family member BES1 HOMOLOG 2 (AtBEH2) (64,2%) and therefore we named the clone PhBEH2 (Fig. 5a,b). Like the previously characterized BES1/BZR1 family members from Arabidopsis, PhBEH2 has a putative bipartite nuclear localization signal (NLS) at its N-terminus (Fig. 5a). The central part of PhBEH2 is rich in serine/threonine residues that match with the phosphorylation sites for GSK3 kinases (Wang et al., 2002). At the C-terminus of PhBEH2 a WEGE core sequence was found which is similar to the BIN2-docking motif (DM) of BZR1 (Peng et al., 2010). Furthermore a PEST motif was identified that seems to play an important role in protein degradation as BES1 and BZR1 proteins were stabilized in bes1-D and bzr1-D mutants carrying a mutation in this motif (Wang et al., 2002; Yin et al., 2002). PhBEH2 bind to a GSK3/SHAGGY like kinase and 14-3-3 proteins To identify proteins that interact with PhBEH2, a yeast two-hybrid screen was performed. As many transcription factors carry activation domains, auto-activation of the full-length cDNA of PhBEH2 on yeast reporters was tested first. Indeed, introduction of the PhBES1GAL4 binding domain fusion resulted in some activation of the histidine reporter, but selection on histidine and adenine resulted in no growth (Supplemental data Fig. 3) Subsequently, the full-length cDNA of PhBEH2 was used as bait to screen a two-hybrid cDNA library made from young petunia inflorescences and twelve interacting proteins were found (Fig. 5c,d). Sequence analysis showed that five of these proteins were 14-3-3 proteins. Another interesting interacting partner had high similarity to the GSK3-like kinase BIN2 from Arabidopsis (Fig. 5c,d). Other interaction partners include proteins homologous to SECRET AGENT (SEC) that encodes an O-linked N-acetylglucosamine transferase (OGT), ARIA (ARMADILLO Repeat protein Interacting with ABF2), TPL (TOPLESS) 92 Brassinosteroid biosynthesis and signalling in petunia and two RRM (RNA Recognition Motif) proteins of which one is a homolog of FPA (Flowering Protein A, Fig. 5c,d). A B AtBES1 AtBZR1 PhBEH2 AtBEH2 AtBEH1 AtBEH3 AtBEH4 1 1 1 1 1 1 1 MKRFFYNSSEEERKKKAYSSKKMTSDGATSTS-AAAAAAAMATRRKPSWRERENNRRRER ----------------------MTSDGATSTSAAAAAAAAAAARRKPSWRERENNRRRER ----------------------MT---------AAGGGGGSATGRLPTWKERENNKRRER ----------------------MAAG-------GGGGGGGSSSGRTPTWKERENNKKRER ----------------------MT---------ASGGGSTAATGRMPTWKERENNKKRER ---------------------------------------MTSGTRTPTWKERENNKRRER ---------------------------------------MTSGTRMPTWRERENNKRRER AtBES1 AtBZR1 PhBEH2 AtBEH2 AtBEH1 AtBEH3 AtBEH4 60 39 30 32 30 22 22 RRRAVAAKIYTGLRAQGNYNLPKHCDNNEVLKALCSEAGWVVEEDGTTYRKGHKPLPGDM RRRAVAAKIYTGLRAQGDYNLPKHCDNNEVLKALCVEAGWVVEEDGTTYRKGCKPLPGEI RRRAIAAKIFTGLRAQGNFKLPKHCDNNEVLKALCIEAGWIVEDDGTTYRKGHRRPPIEY RRRAITAKIYSGLRAQGNYKLPKHCDNNEVLKALCLEAGWIVEDDGTTYRKGFKPPASDI RRRAIAAKIFTGLRSQGNYKLPKHCDNNEVLKALCLEAGWIVHEDGTTYRKGSRP----RRRAIAAKIFAGLRIHGNFKLPKHCDNNEVLKALCNEAGWTVEDDGTTYRKGC-KPMDRRRRAIAAKIFTGLRMYGNYELPKHCDNNEVLKALCNEAGWIVEPDGTTYRKGCSRPVER- AtBES1 AtBZR1 PhBEH2 AtBEH2 AtBEH1 AtBEH3 AtBEH4 120 99 90 92 85 80 81 AGSSSRATPYSSHNQSPLSSTFDSPILSYQVSPSSSSFPSPSRV--GDPHN--ISTIFPF AGTSSRVTPYSSQNQSPLSSAFQSPIPSYQVSPSSSSFPSPSR---GEPNNNMSSTFFPF AVSSMDISACSSIQPSPMSSSFPSPVPSYHASPTSSSFPSPSR---CDGNP--SSYILPF SGTPTNFSTNSSIQPSPQSSAFPSPAPSYHGSPVSSSFPSPSR---YDGNPS-SYLLLPF ---TETTVPCSSIQLSPQSSAFQSPIPSYQASPSSSSYPSPTR---FDPNQS-STYLIPY -----MDLMNGSTSASPCSSYQHSPRASYNPSPSSSSFPSPTN------PFGDANSLIPW -----MEIGGGSATASPCSSYQPSPCASYNPSPGSSNFMSPASSSFANLTSGDGQSLIPW AtBES1 AtBZR1 PhBEH2 AtBEH2 AtBEH1 AtBEH3 AtBEH4 176 156 145 148 138 129 136 LRN--GGIPS------SLPP----LRISNSAPVTPPVSSPTSRNPKP------------LRN--GGIPS------SLPS----LRISNSCPVTPPVSSPTSKNPKP------------LHNL-ASIPF------SLPP----LRISNSAPVTPPPSSPT-RGSQP------------LHNIASSIPA------NLPP----LRISNSAPVTPPLSSPTSRGSKR------------LQN--LASSG------NLAP----LRISNSAPVTPPISSPRRSNPR-------------LKNLSSNSPS------KLP---FFHGNSISAPVTPPLAR--------------------LKHLSTTSSSSASSSSRLPNYLYIPGGSISAPVTPPLSSPTARTPRMNTDWQQLNNSFFV AtBES1 AtBZR1 PhBEH2 AtBEH2 AtBEH1 AtBEH3 AtBEH4 211 191 180 185 172 159 196 ---LPTWESFTKQSMSMAAKQSMTSLNYPFYAVSAPASPTHHRQFHAPATIPECDESDSS ---LPNWESIAKQSMAIA-KQSMASFNYPFYAVSAPASPTHRHQFHTPATIPECDESDSS ---KPVWESFSRG--------PLHSFHHPIFAASAPSSPT-RRQYSKPATIPECDESDAA ---KLTSEQLPNGG-------SLHVLRHPLFAISAPSSPTRRAGHQTPPTIPECDESEED ---LPRWQS-------------------SNFPVSAPSSPTRRLHHYT--SIPECDESDVS ---SPTRDQVTIPD--------SGWLSGMQTPQSGPSSPTFSLVSRNPF-FDKEAFKMGSSTPPSPTRQIIPD--------SEWFSGIQLAQSVPASPTFSLVSQNPFGFKEEAASAAG AtBES1 AtBZR1 PhBEH2 AtBEH2 AtBEH1 AtBEH3 AtBEH4 268 247 228 235 208 206 248 A tB E H 1 A tB E S 1 1000 A tB Z R 1 AtBES1 AtBZR1 PhBEH2 AtBEH2 AtBEH1 AtBEH3 AtBEH4 316 295 281 277 244 248 292 * * * * * * 686 A tB E H 3 1000 A tB E H 4 * * * * P hB EH2 840 0,1 A tB E H 2 ** * * * * C * * 1 8 16 17 1 8 16 17 1 8 16 17 25 36 41 60 25 36 41 60 25 36 41 60 67 73 78 81 67 73 78 81 67 73 78 81 AD AD -LT X-gal AD -LTH -LTHA D *TVD-SGHWISFQKFAQQQPFSASMVPTSPTFNLVKPAPQQLSPNTAAIQ----------* * * Clone Type Arabidopsis homolog TVD-SGHWISFQKFAQQQPFSASMVPTSPTFNLVKPAPQQMSPNTAAFQ----------SVE-SARWVSFQTLAP------SAAPTSPTFNLVKPVIPQQNILLDALSGRGAFGWGEAA SIEDSGRWINFQ----------STAPTSPTFNLVQQTSMAIDMKRSDWGMSG-------TVD-SCRWGNFQSVNVS-----QTCPPSPTFNLVGKSVSSVG------------------DCNSPMWTPGQSGNCS-----PAIPAGVDQNSDVPMADGMT-AEFAFG----------GGGGSRMWTPGQSGTCS-----PAIPPGADQTADVPMSEAVAPPEFAFG----------- 1 14-3-3 GRF1/GF14χ (AT4G09000) 8 14-3-3 16 RNA binding GRF1/GF14κ (AT5G65430) FPA (AT2G43410) 17 OGT SEC (AT3g04240) 25 ARMADILLO ARIA (AT5G16050) 36 WD40 TOPLESS/related1 (AT1G80490) 41 14-3-3 GRF5/GF14ε (AT5G16050) 60 unknown 67 14-3-3 GRF2/GF14ω (AT1G78300) 73 14-3-3 78 GSK3/SHAGGY GRF5/GF14ε (AT5G16050) BIN2 (AT1G06390) 81 RNA binding AT4G00830 EIGQSSEFKFENSQVKPWEGERIH-DVAMEDLELTLGN-GKAHS EIGQSSEFKFENSQVKPWEGERIH-DVGMEDLELTLGN-GKARG QRGHASEFDLESCKVKAWEGERIH-EVAVDDLELTLGA-GKAHA MNGRGAEFEFENGTVKPWEGEMIH-EVGVEDLELTLGG-TKARC ----------VDVSVKPWEGEKIH-DVGIDDLELTLGHNTKGRG -----CNAMAANGMVKPWEGERIHGECVSDDLELTLGN-SRTR-----SN---TNGLVKAWEGERIHEESGSDDLELTLGN-SSTR- Figure 5. Characterization of PhBEH2 and the identification of binding partners by a yeast two-hybrid screen. A) Deduced amino acid alignment of PhBEH2 and the six members of the BES1/BZR1 family from Arabidopsis. Identical residues in four out of seven sequences are indicated in black, similar amino acids in gray. Dashed lines represent gaps to facilitate alignment. A putative nuclear localization signal (straight black line), putative phosphorylation sites by GSK-3 kinases (*), 14-3-3 binding site (red dashed line), PEST motif (black dashed line) and BIN2-DM (straight red line) are shown. Ph, petunia; At, Arabidopsis. B) Phylogenetic tree constructed using derived amino acid sequences of PhBEH2 and the BES1/BZR1 family from Arabidopsis. Genbank accession numbers are recited in Supplemental Table 2. C) The full-length sequence of PhBEH2 was fused to the yeast GAL4-binding domain and screened against a cDNA library made from young petunia inflorescences. Interactions were measured by growth on medium lacking histidine (-LTH), histidine and adenine (-LTHA) and blue colouring using 5-bromo-4-chloro-3-indolyl-β-D-galactosidase (-LT X-gal). D) Classification of the positive clones from (C). As indicated, for some proteins a homolog with known function is found in Arabidopsis. 93 Chapter 4 Isolation and characterization of two group II members of the GSK3/SHAGGY-like family from petunia One of the interacting partners of PhBEH2 was a protein that has strong homology to the GSK3-like kinase BIN2 from Arabidopsis and we named it PSK8 (Fig. 6a and Supplementary data Fig. 2). From a previous yeast two-hybrid screen performed in our laboratory, using one of the transcription factors involved in flower pigmentation as bait, another GSK3-like kinase was identified as an artefact (W. Verweij and F. Quatrocchio, personal communication). As this clone missed the 5’ half of the cDNA, we performed SOTI-PCRs (see materials and methods) to isolate the remaining part. We renamed this clone PSK9. Both PSK8 and PSK9 are related to BIN2 as they group together with the other group II ASKs (Fig. 6a). Kim et al. (2009) showed that the phosphatase BSU1 inactivates BIN2 by dephosphorylation at pTyr 200. This tyrosine residue is also conserved in PSK8 and PSK9 (Supplemental Fig. 2). To get more insight in the function of PSK8, a yeast two-hybrid screen was performed, where full-length PSK8 was used as bait to screen a cDNA library made from young petunia inflorescences. We identified eight different cDNAs that interacted with PSK8. The interacting partners include two different LRR-extensins, REMORIN and several unknown proteins. (Fig. 4b,c). III A B A S K 32 1 3 4 1 3 4 1 3 4 1 3 4 6 16 21 6 16 21 6 16 21 6 16 21 28 93 AD 28 93 AD 28 93 AD 28 93 AD A S K 31 PSK6 MPK1 414 1000 PSK7 487 A S K 41 PSK4 I 1000 A S K 13 A S K 12 1000 A S K 42 A S K 11 A tA S K dzeta 0.1 1000 891 861 1000 910 PSK9 734 1000 -LTH -LT 298 A tB IN 2 A tG S K 1 PSK8 C IV -LTHA -LT X-gal Clone type Arabidopsis homolog 1 unknown At3g11590 3 GCN5-related N-acetyl transferase At2g39020 4 REMORIN AtREM6.3 (At1g53860) 6 LRR-extensin AtLRX3 (At4g13340) 16 LRR-extensin At1g26250 21 Dynein light chain weak homology, no clear Arabidopsis homolog 28 unknown no clear plant homolog 93 unknown At3g63095 II Figure 6. Phylogenetic analysis of PSK8/9 and the identification of PSK8 binding partners by a yeast twohybrid screen. A) Phylogenetic tree constructed of derived amino acid sequences of PSK8/9 and GSK3-like kinases from Arabidopsis and petunia. MPK1 from Arabidopsis was used as an outgroup to construct the tree. Genbank accession numbers are shown in Supplemental data Table 2. B) Interaction of PSK8 with eight different clones in a yeast two-hybrid assay. Interactions were measured by growth on medium lacking histidine (-LTH), histidine and adenine (-LTHA) and blue colouring using 5-bromo-4-chloro-3-indolyl-β-D-galactosidase (-LT Xgal). C) Classification of the positive clones from (B). When a clear homolog in Arabidopsis is found their AGI identifier is listed. 94 Brassinosteroid biosynthesis and signalling in petunia Key regulators of the petunia/Arabidopsis BR pathway are functionally homologous H BZ 2 At R1 BE AD S1 BE At Ph H BZ 2 At R1 BE AD S1 BE At Ph At Ph BE H BZ 2 At R1 BE AD S1 PSK8 and PSK9 are highly homologous to the group II GSK3/SHAGGY-like kinases from Arabidopsis and this prompted us to examine whether they also have functional homology. In order to investigate this, a yeast two-hybrid between members of the BES1/BZR1 family and the group II GSK3/SHAGGY like kinases from Arabidopsis and petunia was performed. AtBIN2 interacted with AtBES1, AtBZR1 and also with BEH2 from petunia (Fig. 7). A similar behaviour was found for the petunia BIN2 homolog PSK8. However, PSK9 did not bind to PhBEH2 nor to the Arabidopsis members. As PhBEH2 is able bind to PSK8, it seems that PSK9 does not interact with the BES1/BZR1 family and thus may not be involved in BR signalling. However it seems more likely that the fusion between PSK9 and the GAL4 binding domain is somehow not functional. Our yeast-two hybrid results suggest that group II GSK3/SHAGGY like kinases and members of the BES/BZR1 family from Arabidopsis and petunia are structurally highly homologous. PSK9 PSK8 AtBIN2 BD -LT -LTH -LTHA Figure 7. The BES1/BZR1 family and group II GSK3/SHAGGY like kinases from petunia and Arabidopsis are functionally homologous. Full length sequences of AtBIN2, PSK8 and PSK9 were fused to the GAL4binding domain and AtBZR1, AtBES1 and PhBEH2 to the GAL4-activation domain and tested for interaction. Interactions were measured by growth on medium lacking histidine (-LTH), histidine and adenine (-LTHA). Discussion The synthesis of BRs and the signalling cascade downstream the bioactive compounds has been largely elucidated over the past ten years (Clouse, 2002a; Fujioka and Yokota, 2003; Kim et al., 2009). However, as in many plant research areas these findings basically originate from Arabidopsis. Partial conformation of evolutionary conservation as well as additions came from several reports on BR biosynthesis and signalling in other species (eg. Nomura et al., 2001; Montoya et al., 2002; Bai et al., 2007; Koh et al., 2007). We have 95 Chapter 4 used the resources that are available in petunia to study BR biosynthesis and signalling. The initial results point to conserved genes and interactions, but also revealed potential new relations between hormone pathways. From our mutant collection we identified five different loci of which four are BR biosynthesis genes. One locus, CD2, appeared to correspond to the Arabidopsis homolog of CPD (Szekeres et al., 1996). At the moment it is unclear what the nature is of the remaining three loci. As transposon insertions in the petunia homologs of DIM and DET2 did not cause the typical BR deficiency phenotype, it seems unlikely that cd1, cd3 or cd9 mutants carry insertions in these genes. By PCR analysis on DNA isolated from the three mutants we were unable to detect dTph1 insertions in DWARF5, CYP85, DET3, DIM, DWARF7, SMT2, DET2, DWARF4 or ROT3. However, due to the lack of complete gene sequences we do not dare to claim that there are no insertions in those genes. If true however, the chance that we have hit an unknown enzyme in BR biosynthesis seems high. Cloning of these genes can then be accomplished using AFLP based transposon display (Van den Broeck et al., 1998). In addition, we (in collaboration with Dr. Takao Yokota, Teikyo University, Japan) are in the process of identifying the BR biosynthesis reaction intermediates that accumulate in these mutants. These data might also give us clues about the nature of the affected genes. The single signalling locus that we identified, CD10, was shown to harbour an insertion in the petunia homolog of the BR receptor, BRI1 (Clouse et al., 1996). The cd10 mutant exhibits an extremely dwarfed stature and did not flower at all. BRI1 orthologs have been identified in a number of species such as tomato, pea, rice, cotton and barley (Yamamuro et al., 2000; Montoya et al., 2002; Chono et al., 2003; Nomura et al., 2003; Sun et al., 2004). All bri1 mutants do flower except the osbri1 mutant from rice (Nakamura et al., 2006). The strong mutants from rice and petunia might reflect complete loss of BRI1 function whereas in other species either only partial loss-of-function alleles were identified and/or partial redundancy in BRI1 function might exist. PhBRI1 is most homologous to BRI1 from Lycoperscicon esculentum, Nicotiana benthamiana and Nicotiana tabacum. In an amino acid alignment of BRI1 from petunia, tomato, Arabidopsis and Nicotiana benthamiana the highest similarity was observed in the crucial kinase domain (Supplementary data Fig. 1; Tang et al., 2008; Wang et al., 2008a). The extracellular domain of PhBRI1 contains 25-tandem amino-terminal LRRs and a 68 amino acid non-repetitive island between the 21st and 22nd LRR. This island together with LRR22 has proved to be important for BR binding by BRI1 in Arabidopsis (Kinoshita et al., 2005). Thus it seems that the perception of the BR hormone is evolutionary well conserved. 96 Brassinosteroid biosynthesis and signalling in petunia Apart from BRI1, other key regulators of the Arabidopsis BR signalling pathway have been identified from a number of species. In rice as well as in cotton gain-/loss-of-function of BIN2 homologs suggest that they are involved in BR signalling (Sun et al., 2004; Sun and Allen, 2005; Koh et al., 2007). In addition four genes (OsBZR1-4) were isolated from rice that are homologous to the Arabidopsis BES1 and BZR1 transcription factors (Bai et al., 2007). In order to reveal conservation in the BR signalling pathway among plant species, we cloned several petunia homologs of key components of the BR signalling pathway in Arabidopsis including the BRI1 receptor, a homolog of the BES1/BZR1 family and GSK3like kinases. The BES1/BZR1 family homolog that we isolated from petunia showed highest sequence identity to the BES1/BZR1 family member BEH2 (64,2%) and therefore we named this clone PhBEH2. In Arabidopsis a BR treatment increased the accumulation of dephosphorylated BZR1, BES1 and also BEH2 (Yin et al., 2005). Furthermore, the bzr1/bes1 double mutant only exhibits a semi-dwarf phenotype. Therefore, it seems likely that BEH2 and maybe other members of the BES1/BZR1 family are redundant to BES1 and BZR1. In chapter 2 we showed that group II GSK3-like kinases interact with several members of the BES1/BZR1 family, including AtBEH2, again insinuating that AtBEH2 also plays a role in the Arabidopsis BR signalling pathway. The interaction spectrum of PhBEH2 in yeast also implies a role in BR signalling. First, PhBEH2 is capable of interacting with Arabidopsis BIN2, a proven negative regulator of BR signalling (Li et al., 2001). Second, the PhBEH2 protein interacted with several 14-3-3 proteins. Previous research in Arabidopsis showed that 14-3-3 proteins increase the cytoplasmic retention of phosphorylated BES1. Mutation of a BIN2 phosphorylation site in BZR1 was shown to abolish 14-3-3 binding, leading to increased nuclear localization of the BZR1 protein (Gampala et al., 2007). Similar results have been obtained with the rice homolog OsBZR1 (Bai et al., 2007). Like AtBZR1 and AtBES1, PhBEH2 also possesses the conserved putative 14-3-3 binding site, suggesting that the regulation of the BZR1/BES1 family by 14-3-3 proteins might be conserved among plants. Third, PhBEH2 interacted with a GSK3-like kinase (PSK8) that clustered together with the group II GSK3like kinases. Arabidopsis group II GSK3-like kinases have been shown to negatively regulate BR signalling by phosphorylating BES1 and BZR1 (Chapter 2; Vert and Chory, 2006; Yan et al., 2009). Although PSK8 interacted with a member of the BES1/BZR1 family, more proof is needed to conclude whether this kinase plays a role in BR signalling. For instance, phosphorylation of PhBEH2 by PSK8/9 should be demonstrated and mutants might ultimately reveal its function. Next to PSK8 we also identified another group II member (PSK9) in petunia. Solely based on homology it seems likely that PSK9 is involved in BR signalling as well. The lack of interaction with PhBEH2 and with 97 Chapter 4 Arabidopsis BES1 and BZR1 is likely due to aberrant processing of the PSK9-GAL4 fusion RNA and/or protein in yeast. Our yeast two-hybrid screen with PhBEH2 revealed several interacting proteins that function in diverse signal transduction pathways. This suggests that PhBEH2 could be an important hub between BR and other signalling pathways. For instance, PhBEH2 interacted with a protein that has high sequence homology to ARIA (ARMADILLO REPEAT PROTEIN INTERACTING WITH ABF2), a positive regulator of the ABA signalling cascade (Kim et al., 2004). ARIA modulates the transcriptional activity of ABF2, which interacts with ABA-responsive elements in target promoters. ARIA seems only involved in a subset of ABA-dependent processes, namely germination and growth of young seedlings, but not for example in stomatal closure. Zhang et al. (2009) demonstrated that ABA rapidly induced the expression of the BR-suppressed genes CPD and DWF4 and also enhanced the phosphorylation of BES1. However in the absence of BIN2, ABA failed to induce BES1 phosphorylation. The interaction between PhBEH2 and PhARIA could be the link between the ABA and BR signalling pathways. Numerous reports point to a synergistic interaction between auxin and BR signalling. Specificity of the identified interaction of PhBEH2 with a TOPLESS-related protein, a transcriptional co-repressor of AUXIN RESPONSE FACTORS (ARFs) (Szemenyei et al., 2008) seems doubtful, as the same protein was already identified as a possible partner in other unrelated yeast 2-hybrid screens using various baits. However, recent findings point to a more general role for TOPLESS(-related) proteins as transcriptional co-repressors in diverse unrelated processes (Pauwels et al., 2010) and thus the discovered interactions might be meaningful. PhBEH2 interacts with two proteins containing RNA Recognition Motifs (RRM). One of the RRM proteins is a homolog of FPA, a gene that regulates flowering time in Arabidopsis via a pathway that is independent of day length. FPA negatively regulates the expression of the floral repressor FLOWERING LOCUS C (Schomburg et al., 2001). Interestingly, Domagalska et al. (2007) demonstrated that the BR signalling pathway plays an assisting role in the repression of FLC and promotes flowering. This is consistent with the finding that BR biosynthesis mutants as well as signalling mutants are late-flowering or, like our cd10 (petunia bri1) mutant, non-flowering. In Arabidopsis two JmjN/C domain-containing histone demethylases, ELF6 (early flowering 6) and REF6 (relative of early flowering 6), were identified as BES1 partners (Yu et al., 2008). ELF6 was found to be a repressor in the photoperiodic flowering pathway and REF6 negatively regulates the expression of FLC. In addition many induced BR-induced genes such as cell wall-modifying enzymes involved in cell elongation (like xyloglucan endotransglucosylases/hydrolases (XETs/XTHs), expansins, and pectate lyases,) were down-regulated in ref6 and elf6 insertion mutants, 98 Brassinosteroid biosynthesis and signalling in petunia suggesting that the interaction of BES1 with REF6/ELF6 positively regulates BR target gene expression and promotes cell elongation. It could be that BRs control developmental processes such as flowering and cell elongation via the interaction between BES1, REF6/ELF6 and FPA. Additionally it has been shown that rice SVP-group MADS-box proteins that repress floral meristem identity, work as negative regulators in BR responses, however their roles are diversified and they function at different developmental stages (Lee et al., 2008a). Further research by Lee et al. (2008b) suggest that SVP-group MADS-box proteins repress BR responses by binding to the promoters of BR target genes. Maybe one of the most interesting partners of PhBEH2 is a protein homologous to the Arabidopsis gene SECRET AGENT (SEC) encoding an O-linked N-acetylglucosamine transferase (OGT) (Hartweck et al., 2002; Hartweck et al., 2006). SEC catalyses the transfer of a single O-linked-N-acetylglucosamine to specific serine/threonine residues of proteins. SEC is partly redundant with another OGT in Arabidopsis called SPINDLY (SPY). In Arabidopsis SPY negatively regulates the gibberellin (GA) pathway, but also promotes cytokinin responses and interacts with the circadian clock protein GIGANTEA in yeast (Tseng et al., 2004; Maymon et al., 2009). It seems that SEC is not involved in GA signalling or only has a limited role, (Hartweck et al., 2006). Shimada et al. (2006) reported that mutation of the rice SPINDLY gene (OsSPY) reduced the expression of several BR biosynthesis genes and OsBRI1. As expected, the levels of sterol and BR compounds were slightly up regulated in the OsSPY knock-out mutant. O-GlcNAcylation of target proteins might affect their activity, localization, stability, and interactions with other proteins. The fact that OGTs add acetylglucosamine to serine/threonine residues of proteins envisages a situation where OGTs and GSKs compete for modification of these residues in the BES/BZR family of proteins. Such a scenario, where an OGT competes with a kinase is a widely recognised mechanism in animal research (Kamemura et al., 2002; Wang et al., 2008b; Butkinaree et al., 2009). Most striking, there are specific examples where GSK3 phosphorylation and O-GlcNAcylation occurs on the same site of proteins (Wang et al., 2007). The addition of acetylglucosamine by OGTs to BES1/BZR1 family members might have role in localization, stability or activity of these proteins. Although we identified some interesting links between BR signalling and other pathways via PhBEH2 (and possibly via the other members of the BES1/BZR1 family) further research should confirm these findings. For example, it would be interesting to see whether PhSEC adds acetylglucosamine to PhBEH2 and if so what the consequences are for the function of PhBEH2. Furthermore, it would be interesting to reveal whether ABA is capable of inducing phosphorylation of BES1 in an aria mutant background and if the expression of BR-suppressed genes is induced. 99 Chapter 4 Remarkably, we did not find any known homologs of the BR signalling pathway in the yeast two-hybrid screen performed with PSK8. Using PhBEH2 as bait we isolated PSK8, but using the same cDNA library the reverse interaction was not identified. It might be that none of the BES/BZR1 family members present in the library is in frame with the GAL4 binding domain. Nonetheless, our yeast two-hybrid screen with PSK8 might have revealed other roles of group II GSK kinases that have not been recognised in other model species yet. PSK8 interacted with a number of proteins that are not very well characterised yet, but at least some seem to be cell membrane or wall localized. Two proteins showed high homology to the LLR-extensin family (LRX), which are localized in the cell wall. It has been suggested that LRXs might control the development of the cell by regulating cell wall expansion, cell polarization and the modification of localized cell wall domains during differentiation (Baumberger et al., 2003). As BRs are involved in cell elongation, such an interaction is not unlikely. PSK8 mediated phosphorylation of LRR-extensins might affect their stability, activity or disrupt interaction with other proteins. PSK8 also interacted with a homolog of AtREM6.3, a member of the REMORIN family. REMORINS are plant specific and are located in the plasma membrane and in lipid rafts (Raffaele et al., 2007). The precise biological roles of REMORINS are not clear, but according to the available expression data they may be involved in various signal transduction pathways triggered by abiotic and biotic stressors. The expression level of AtREM6.3 is the highest in the embryo and interestingly, a BR treatment decreased the expression level of AtREM6.3 (Raffaele et al., 2007). Experimental procedures Brassinolide experiment 24- epibrassinolide was purchased from Wako Biochemistry, dissolved in DMSO (2mM) and stored at -20°C. Young plantlets were sprayed each tow or three days until runoff with 1 µM brassinolide or with a solvent control (0.01% ethanol and 0.1% Tween 20) for a couple of weeks. Genomic DNA isolation and PCR Petunia DNA was isolated by grinding one young leaf in liquid nitrogen. To this 400 µl DNA extraction buffer was added (0.1 M Tris pH 8.0; 0.5 M NaCl; 50 mM EDTA and 10 mM β mercapto-ethanol) and samples were incubated for 15-30 min. at 65°C. Next 250 µl 100 Brassinosteroid biosynthesis and signalling in petunia phenol/chloroform was added, samples were shaken and subsequently centrifuged for 5 minutes at maximum speed in a microcentrifuge. The supernatant was transferred to a clean tube and DNA was precipitated using 2-propanol and washed with 70% EtOH. The pellet was dissolved in 100 µl H2O containing DNase free RNase (10 mg/ml) followed by an incubation step for 15-30 min. at 65°C. PCR reactions were performed by using 2 µl genomic DNA in a standard PCR reaction. The nucleotide sequences of the primers used are listed in Supplemental Table 1. All PCR products were separated by electrophoresis in a 1% (w/v) agarose gel. Isolation of 3’ gene fragments by RACE PCR In order to isolate cDNA fragments, total RNA was isolated from leaves followed by first strand cDNA synthesis using the RACE1 primer as described previously (Quattrocchio et al., 2006). PCR amplification was carried out using gene-specific primers or, to isolate 3’ ends of cDNAs, by using one gene-specific primer in combination with the RACE2 primer under standard PCR conditions (annealing temperature 55˚C). All used primer sequences are listed in Supplemental Table 1. PCR fragments were directly sequenced or cloned in pGEM-Teasy (Promega) and sequenced. Isolation of 5’ gene fragments by SOTI PCR To complete sequences at the 5´ end, for some genes a somatic insertion-mediated PCR was performed (Rebocho et al., 2008). Here, we make use of somatic dTph1 insertions in the target gene that may occur in some cells of petunia W138 plants. Multiple W138 genomic DNAs from leaves of individual plants or leaves of pooled plants were subjected to nested PCR using gene-specific primers and dTph1 transposon primers (see Supplemental Table 1 for primer sequences). In the first PCR the dTph1 transposon primers out11 or out12 were used in a standard PCR reaction with annealing temperature 55˚C. In the second, nested PCR, 1 µl of the out11 reaction was re-amplified with out15 and a nested gene-specific primer while the out12 reaction was re-amplified with out6 and a nested gene-specific primer. The second reaction was a touch-down PCR where annealing started at 70˚C and decreased with one ˚C every cycle until reaching 55 ˚C, followed by 25 additional cycles at 55 ˚C. PCR fragments were directly sequenced or cloned in pGEM-Teasy (Promega) and sequenced. 101 Chapter 4 Phylogenetic analysis Phylogenetic trees were generated using the neighbour-joining method of ClustalX (ftp://ftp.ebi.ac.uk/pub/software/clustalw2/). Bootstrap mode (1000 replications) was used for estimating the level of confidence assigned to the particular nodes in the tree. The tree was visualized with help of Treeviewer 1.6.6 (http://taxonomy.zoology.gla.ac.uk/rod/rod. html) and rooted using ClustalW (http://www.bioinformatics.nl/tools/clustalw.html). Alignments were coloured using Boxshade (http://www.ch.embnet.org/software/BOX_ form.html). Yeast two-hybrid The full length coding sequences of PhBEH2 was amplified by RT-PCR on total RNA isolated from leaves using Phusion polymerase (Finnzymes) and primers M736 and M737. The amplification product was cloned EcoRI/SalI in pBD-GAL4 and pAD-GAL4 (Stratagene). The PSK8-GAL4BD fusion was made by cutting PhBEH2 yeast two-hybrid interactor 78 with EcoRI/SalI and ligation of the fragment in pBD-GAL4. The plasmids were introduced into the yeast strain PJ69 (James et al., 1996), which carries HIS3, ADE2 and LACZ reporter genes driven by distinct GAL4-responsive promoters. For the yeast two-hybrid screen, PhBEH2/PSK8 were used as bait against a cDNA library from young petunia inflorescences (Souer et al., 2008). Library screen was performed as described previously (Quattrocchio et al., 2006). A total of 100.000 and 300.000 transformed yeast cells were screened for PhBEH2 and PSK8, respectively. Plasmid DNA was isolated from positive (HIS, ADE) colonies, transformed to Escherichia coli, sequenced and subsequently reintroduced in PJ69, either with the empty pBD-GAL4 vector or with the bait plasmid expressing PhBEH2/PSK8. Only plasmids that grew on media lacking histidine and adenine when co-transformed with the bait vector, but not when co-transformed with the empty pBD-GAL4, were considered as positives and analyzed further. For the yeast two-hybrid assay between members of the BES1/BZR1 family and GSK3/SHAGGY-like family members from Arabidopsis and petunia were amplified by RT-PCR using Phusion polymerase (Finnzymes) on Petunia leaf total RNA or from existing plasmid clones. The amplification products were cloned EcoRI/SalI in pBD-GAL4 and/or pAD-GAL4 (Stratagene). 102 Brassinosteroid biosynthesis and signalling in petunia Acknowledgements We thank Allan Awanijan and Longinus Igbojionu for their assistance during some of the cloning experiments and Maartje Kuijpers for plant care. This work was supported by a grant of the Netherlands Organisation for Scientific Research (NWO). References Bai, M.Y., Zhang, L.Y., Gampala, S.S., Zhu, S.W., Song, W.Y., Chong, K. and Wang, Z.Y. (2007) Functions of OsBZR1 and 14-3-3 proteins in brassinosteroid signaling in rice. Proc Natl Acad Sci U S A, 104, 13839-13844. Bajguz, A. and Tretyn, A. (2003) The chemical characteristic and distribution of brassinosteroids in plants. Phytochemistry, 62, 1027-1046. Bajguz, A. and Hayat, S. (2009) Effects of brassinosteroids on the plant responses to environmental stresses. Plant Physiol Biochem, 47, 1-8. Baumberger, N., Doesseger, B., Guyot, R., Diet, A., Parsons, R.L., Clark, M.A., Simmons, M.P., Bedinger, P., Goff, S.A., Ringli, C. and Keller, B. (2003) Whole-genome comparison of leucine-rich repeat extensins in Arabidopsis and rice. A conserved family of cell wall proteins form a vegetative and a reproductive clade. Plant Physiol, 131, 1313-1326. Butkinaree, C., Park, K. and Hart, G.W. (2009) O-linked beta-N-acetylglucosamine (O-GlcNAc): Extensive crosstalk with phosphorylation to regulate signaling and transcription in response to nutrients and stress. Biochim Biophys Acta, 1800, 96-106. Choe, S., Dilkes, B.P., Fujioka, S., Takatsuto, S., Sakurai, A. and Feldmann, K.A. (1998) The DWF4 gene of Arabidopsis encodes a cytochrome P450 that mediates multiple 22alpha-hydroxylation steps in brassinosteroid biosynthesis. Plant Cell, 10, 231-243. Choe, S., Dilkes, B.P., Gregory, B.D., Ross, A.S., Yuan, H., Noguchi, T., Fujioka, S., Takatsuto, S., Tanaka, A., Yoshida, S., Tax, F.E. and Feldmann, K.A. (1999) The Arabidopsis dwarf1 mutant is defective in the conversion of 24-methylenecholesterol to campesterol in brassinosteroid biosynthesis. Plant Physiol, 119, 897907. Choe, S., Fujioka, S., Noguchi, T., Takatsuto, S., Yoshida, S. and Feldmann, K.A. (2001) Overexpression of DWARF4 in the brassinosteroid biosynthetic pathway results in increased vegetative growth and seed yield in Arabidopsis. Plant J, 26, 573-582. Chono, M., Honda, I., Zeniya, H., Yoneyama, K., Saisho, D., Takeda, K., Takatsuto, S., Hoshino, T. and Watanabe, Y. (2003) A semidwarf phenotype of barley uzu results from a nucleotide substitution in the gene encoding a putative brassinosteroid receptor. Plant Physiol, 133, 1209-1219. Clouse, S.D., Langford, M. and McMorris, T.C. (1996) A brassinosteroid-insensitive mutant in Arabidopsis thaliana exhibits multiple defects in growth and development. Plant Physiol, 111, 671-678. Clouse, S.D. (2002) Arabidopsis mutants reveal multiple roles for sterols in plant development. Plant Cell, 14, 1995-2000. Clouse, S.D. (2002) Brassinosteroid signaling: novel downstream components emerge. Curr Biol, 12, R485-487. Domagalska, M.A., Schomburg, F.M., Amasino, R.M., Vierstra, R.D., Nagy, F. and Davis, S.J. (2007) Attenuation of brassinosteroid signaling enhances FLC expression and delays flowering. Development, 134, 28412850. 103 Chapter 4 Fujioka, S., Takatsuto, S. and Yoshida, S. (2002) An early C-22 oxidation branch in the brassinosteroid biosynthetic pathway. Plant Physiol, 130, 930-939. Fujioka, S. and Yokota, T. (2003) Biosynthesis and metabolism of brassinosteroids. Annu Rev Plant Biol, 54, 137-164. Gampala, S.S., Kim, T.W., He, J.X., Tang, W., Deng, Z., Bai, M.Y., Guan, S., Lalonde, S., Sun, Y., Gendron, J.M., Chen, H., Shibagaki, N., Ferl, R.J., Ehrhardt, D., Chong, K., Burlingame, A.L. and Wang, Z.Y. (2007) An essential role for 14-3-3 proteins in brassinosteroid signal transduction in Arabidopsis. Dev Cell, 13, 177-189. Gerats, A.G., Huits, H., Vrijlandt, E., Marana, C., Souer, E. and Beld, M. (1990) Molecular characterization of a nonautonomous transposable element (dTph1) of petunia. Plant Cell, 2, 1121-1128. Grove, M.D., Gayland, F.S., Rohwedder, W.K., Nagabhushanam, M., Worley, J.F., Warthen Jr, J.D., Steffens, G.L., Flippen-Anderson, J.L. and Cook Jr, J.C. (1979) Brassinolide, a plant growth-promoting steroid isolated from Brassica napus pollen. Nature, 281, 2. Hartweck, L.M., Scott, C.L. and Olszewski, N.E. (2002) Two O-linked N-acetylglucosamine transferase genes of Arabidopsis thaliana L. Heynh. have overlapping functions necessary for gamete and seed development. Genetics, 161, 1279-1291. Hartweck, L.M., Genger, R.K., Grey, W.M. and Olszewski, N.E. (2006) SECRET AGENT and SPINDLY have overlapping roles in the development of Arabidopsis thaliana L. Heyn. J Exp Bot, 57, 865-875. He, J.X., Gendron, J.M., Yang, Y., Li, J. and Wang, Z.Y. (2002) The GSK3-like kinase BIN2 phosphorylates and destabilizes BZR1, a positive regulator of the brassinosteroid signaling pathway in Arabidopsis. Proc Natl Acad Sci U S A, 99, 10185-10190. James, P., Halladay, J. and Craig, E.A. (1996) Genomic libraries and a host strain designed for highly efficient two-hybrid selection in yeast. Genetics, 144, 1425-1436. Kamemura, K., Hayes, B.K., Comer, F.I. and Hart, G.W. (2002) Dynamic interplay between O-glycosylation and O-phosphorylation of nucleocytoplasmic proteins: alternative glycosylation/phosphorylation of THR-58, a known mutational hot spot of c-Myc in lymphomas, is regulated by mitogens. J Biol Chem, 277, 19229-19235. Kim, S., Choi, H.I., Ryu, H.J., Park, J.H., Kim, M.D. and Kim, S.Y. (2004) ARIA, an Arabidopsis arm repeat protein interacting with a transcriptional regulator of abscisic acid-responsive gene expression, is a novel abscisic acid signaling component. Plant Physiol, 136, 3639-3648. Kim, T.W., Guan, S., Sun, Y., Deng, Z., Tang, W., Shang, J.X., Burlingame, A.L. and Wang, Z.Y. (2009) Brassinosteroid signal transduction from cell-surface receptor kinases to nuclear transcription factors. Nat Cell Biol, 11, 1254-1260. Kinoshita, T., Cano-Delgado, A., Seto, H., Hiranuma, S., Fujioka, S., Yoshida, S. and Chory, J. (2005) Binding of brassinosteroids to the extracellular domain of plant receptor kinase BRI1. Nature, 433, 167-171. Klahre, U., Noguchi, T., Fujioka, S., Takatsuto, S., Yokota, T., Nomura, T., Yoshida, S. and Chua, N.H. (1998) The Arabidopsis DIMINUTO/DWARF1 gene encodes a protein involved in steroid synthesis. Plant Cell, 10, 1677-1690. Koh, S., Lee, S.C., Kim, M.K., Koh, J.H., Lee, S., An, G., Choe, S. and Kim, S.R. (2007) T-DNA tagged knockout mutation of rice OsGSK1, an orthologue of Arabidopsis BIN2, with enhanced tolerance to various abiotic stresses. Plant Mol Biol, 65, 453-466. Lee, S., Choi, S.C. and An, G. (2008) Rice SVP-group MADS-box proteins, OsMADS22 and OsMADS55, are negative regulators of brassinosteroid responses. Plant J, 54, 93-105. Lee, S., Jeong, D.H. and An, G. (2008) A possible working mechanism for rice SVP-group MADS-box proteins as negative regulators of brassinosteroid responses. Plant Signal Behav, 3, 471-474. 104 Brassinosteroid biosynthesis and signalling in petunia Li, J., Nagpal, P., Vitart, V., McMorris, T.C. and Chory, J. (1996) A role for brassinosteroids in lightdependent development of Arabidopsis. Science, 272, 398-401. Li, J. and Chory, J. (1997) A putative leucine-rich repeat receptor kinase involved in brassinosteroid signal transduction. Cell, 90, 929-938. Li, J., Nam, K.H., Vafeados, D. and Chory, J. (2001) BIN2, a new brassinosteroid-insensitive locus in Arabidopsis. Plant Physiol, 127, 14-22. Li, J., Wen, J., Lease, K.A., Doke, J.T., Tax, F.E. and Walker, J.C. (2002) BAK1, an Arabidopsis LRR receptor-like protein kinase, interacts with BRI1 and modulates brassinosteroid signaling. Cell, 110, 213-222. Lindsey, K., Pullen, M.L. and Topping, J.F. (2003) Importance of plant sterols in pattern formation and hormone signalling. Trends Plant Sci, 8, 521-525. Mathur, J., Molnar, G., Fujioka, S., Takatsuto, S., Sakurai, A., Yokota, T., Adam, G., Voigt, B., Nagy, F., Maas, C., Schell, J., Koncz, C. and Szekeres, M. (1998) Transcription of the Arabidopsis CPD gene, encoding a steroidogenic cytochrome P450, is negatively controlled by brassinosteroids. Plant J, 14, 593-602. Maymon, I., Greenboim-Wainberg, Y., Sagiv, S., Kieber, J.J., Moshelion, M., Olszewski, N. and Weiss, D. (2009) Cytosolic activity of SPINDLY implies the existence of a DELLA-independent gibberellin-response pathway. Plant J, 58, 979-988. Montoya, T., Nomura, T., Farrar, K., Kaneta, T., Yokota, T. and Bishop, G.J. (2002) Cloning the tomato curl3 gene highlights the putative dual role of the leucine-rich repeat receptor kinase tBRI1/SR160 in plant steroid hormone and peptide hormone signaling. Plant Cell, 14, 3163-3176. Nakamura, A., Fujioka, S., Sunohara, H., Kamiya, N., Hong, Z., Inukai, Y., Miura, K., Takatsuto, S., Yoshida, S., Ueguchi-Tanaka, M., Hasegawa, Y., Kitano, H. and Matsuoka, M. (2006) The role of OsBRI1 and its homologous genes, OsBRL1 and OsBRL3, in rice. Plant Physiol, 140, 580-590. Nam, K.H. and Li, J. (2002) BRI1/BAK1, a receptor kinase pair mediating brassinosteroid signaling. Cell, 110, 203-212. Nomura, T., Sato, T., Bishop, G.J., Kamiya, Y., Takatsuto, S. and Yokota, T. (2001) Accumulation of 6deoxocathasterone and 6-deoxocastasterone in Arabidopsis, pea and tomato is suggestive of common rate-limiting steps in brassinosteroid biosynthesis. Phytochemistry, 57, 171-178. Nomura, T., Bishop, G.J., Kaneta, T., Reid, J.B., Chory, J. and Yokota, T. (2003) The LKA gene is a BRASSINOSTEROID INSENSITIVE 1 homolog of pea. Plant J, 36, 291-300. Pauwels, L., Barbero, G.F., Geerinck, J., Tilleman, S., Grunewald, W., Perez, A.C., Chico, J.M., Bossche, R.V., Sewell, J., Gil, E., Garcia-Casado, G., Witters, E., Inze, D., Long, J.A., De Jaeger, G., Solano, R. and Goossens, A. (2010) NINJA connects the co-repressor TOPLESS to jasmonate signalling. Nature, 464, 788-791. Peng, P., Zhao, J., Zhu, Y., Asami, T. and Li, J. (2010) A direct docking mechanism for a plant GSK3-like kinase to phosphorylate its substrates. J Biol Chem, 285, 24646-24653. Quattrocchio, F., Verweij, W., Kroon, A., Spelt, C., Mol, J. and Koes, R. (2006) PH4 of Petunia is an R2R3 MYB protein that activates vacuolar acidification through interactions with basic-helix-loop-helix transcription factors of the anthocyanin pathway. Plant Cell, 18, 1274-1291. Raffaele, S., Mongrand, S., Gamas, P., Niebel, A. and Ott, T. (2007) Genome-wide annotation of remorins, a plant-specific protein family: evolutionary and functional perspectives. Plant Physiol, 145, 593-600. Rebocho, A.B., Bliek, M., Kusters, E., Castel, R., Procissi, A., Roobeek, I., Souer, E. and Koes, R. (2008) Role of EVERGREEN in the development of the cymose petunia inflorescence. Dev Cell, 15, 437-447. Rozhon, W., Mayerhofer, J., Petutschnig, E., Fujioka, S. and Jonak, C. (2010) ASKtheta, a group-III Arabidopsis GSK3, functions in the brassinosteroid signalling pathway. Plant J, 62, 215-223. 105 Chapter 4 Russinova, E., Borst, J.W., Kwaaitaal, M., Cano-Delgado, A., Yin, Y., Chory, J. and de Vries, S.C. (2004) Heterodimerization and endocytosis of Arabidopsis brassinosteroid receptors BRI1 and AtSERK3 (BAK1). Plant Cell, 16, 3216-3229. Ryu, H., Kim, K., Cho, H., Park, J., Choe, S. and Hwang, I. (2007) Nucleocytoplasmic shuttling of BZR1 mediated by phosphorylation is essential in Arabidopsis brassinosteroid signaling. Plant Cell, 19, 2749-2762. Schomburg, F.M., Patton, D.A., Meinke, D.W. and Amasino, R.M. (2001) FPA, a gene involved in floral induction in Arabidopsis, encodes a protein containing RNA-recognition motifs. Plant Cell, 13, 1427-1436. Shimada, A., Ueguchi-Tanaka, M., Sakamoto, T., Fujioka, S., Takatsuto, S., Yoshida, S., Sazuka, T., Ashikari, M. and Matsuoka, M. (2006) The rice SPINDLY gene functions as a negative regulator of gibberellin signaling by controlling the suppressive function of the DELLA protein, SLR1, and modulating brassinosteroid synthesis. Plant J, 48, 390-402. Souer, E., Rebocho, A.B., Bliek, M., Kusters, E., de Bruin, R.A. and Koes, R. (2008) Patterning of inflorescences and flowers by the F-Box protein DOUBLE TOP and the LEAFY homolog ABERRANT LEAF AND FLOWER of petunia. Plant Cell, 20, 2033-2048. Sun, Y., Fokar, M., Asami, T., Yoshida, S. and Allen, R.D. (2004) Characterization of the brassinosteroid insensitive 1 genes of cotton. Plant Mol Biol, 54, 221-232. Sun, Y. and Allen, R.D. (2005) Functional analysis of the BIN 2 genes of cotton. Mol Genet Genomics, 274, 5159. Szekeres, M., Nemeth, K., Koncz-Kalman, Z., Mathur, J., Kauschmann, A., Altmann, T., Redei, G.P., Nagy, F., Schell, J. and Koncz, C. (1996) Brassinosteroids rescue the deficiency of CYP90, a cytochrome P450, controlling cell elongation and de-etiolation in Arabidopsis. Cell, 85, 171-182. Szemenyei, H., Hannon, M. and Long, J.A. (2008) TOPLESS mediates auxin-dependent transcriptional repression during Arabidopsis embryogenesis. Science, 319, 1384-1386. Tang, W., Kim, T.W., Oses-Prieto, J.A., Sun, Y., Deng, Z., Zhu, S., Wang, R., Burlingame, A.L. and Wang, Z.Y. (2008) BSKs mediate signal transduction from the receptor kinase BRI1 in Arabidopsis. Science, 321, 557560. Tseng, T.S., Salome, P.A., McClung, C.R. and Olszewski, N.E. (2004) SPINDLY and GIGANTEA interact and act in Arabidopsis thaliana pathways involved in light responses, flowering, and rhythms in cotyledon movements. Plant Cell, 16, 1550-1563. Van den Broeck, D., Maes, T., Sauer, M., Zethof, J., De Keukeleire, P., D'Hauw, M., Van Montagu, M. and Gerats, T. (1998) Transposon Display identifies individual transposable elements in high copy number lines. Plant J, 13, 121-129. van Houwelingen, A., Souer, E., Spelt, K., Kloos, D., Mol, J. and Koes, R. (1998) Analysis of flower pigmentation mutants generated by random transposon mutagenesis in Petunia hybrida. Plant J, 13, 39-50. Vandenbussche, M., Janssen, A., Zethof, J., van Orsouw, N., Peters, J., van Eijk, M.J., Rijpkema, A.S., Schneiders, H., Santhanam, P., de Been, M., van Tunen, A. and Gerats, T. (2008) Generation of a 3D indexed Petunia insertion database for reverse genetics. Plant J, 54, 1105-1114. Vert, G. and Chory, J. (2006) Downstream nuclear events in brassinosteroid signalling. Nature, 441, 96-100. Wang, Z.Y., Nakano, T., Gendron, J., He, J., Chen, M., Vafeados, D., Yang, Y., Fujioka, S., Yoshida, S., Asami, T. and Chory, J. (2002) Nuclear-localized BZR1 mediates brassinosteroid-induced growth and feedback suppression of brassinosteroid biosynthesis. Dev Cell, 2, 505-513. Wang, Z., Pandey, A. and Hart, G.W. (2007) Dynamic interplay between O-linked N-acetylglucosaminylation and glycogen synthase kinase-3-dependent phosphorylation. Mol Cell Proteomics, 6, 1365-1379. 106 Brassinosteroid biosynthesis and signalling in petunia Wang, X., Kota, U., He, K., Blackburn, K., Li, J., Goshe, M.B., Huber, S.C. and Clouse, S.D. (2008) Sequential transphosphorylation of the BRI1/BAK1 receptor kinase complex impacts early events in brassinosteroid signaling. Dev Cell, 15, 220-235. Wang, Z., Gucek, M. and Hart, G.W. (2008) Cross-talk between GlcNAcylation and phosphorylation: sitespecific phosphorylation dynamics in response to globally elevated O-GlcNAc. Proc Natl Acad Sci U S A, 105, 13793-13798. Yamamuro, C., Ihara, Y., Wu, X., Noguchi, T., Fujioka, S., Takatsuto, S., Ashikari, M., Kitano, H. and Matsuoka, M. (2000) Loss of function of a rice brassinosteroid insensitive1 homolog prevents internode elongation and bending of the lamina joint. Plant Cell, 12, 1591-1606. Yan, Z., Zhao, J., Peng, P., Chihara, R.K. and Li, J. (2009) BIN2 Functions Redundantly with Other Arabidopsis GSK3-Like Kinases to Regulate Brassinosteroid Signaling. Plant Physiol, 150, 710-721. Yin, Y., Wang, Z.Y., Mora-Garcia, S., Li, J., Yoshida, S., Asami, T. and Chory, J. (2002) BES1 accumulates in the nucleus in response to brassinosteroids to regulate gene expression and promote stem elongation. Cell, 109, 181-191. Yin, Y., Vafeados, D., Tao, Y., Yoshida, S., Asami, T. and Chory, J. (2005) A new class of transcription factors mediates brassinosteroid-regulated gene expression in Arabidopsis. Cell, 120, 249-259. Yu, X., Li, L., Guo, M., Chory, J. and Yin, Y. (2008) Modulation of brassinosteroid-regulated gene expression by Jumonji domain-containing proteins ELF6 and REF6 in Arabidopsis. Proc Natl Acad Sci U S A, 105, 76187623. Zhang, S., Cai, Z. and Wang, X. (2009) The primary signaling outputs of brassinosteroids are regulated by abscisic acid signaling. Proc Natl Acad Sci U S A, 106, 4543-4548. Zhao, J., Peng, P., Schmitz, R.J., Decker, A.D., Tax, F.E. and Li, J. (2002) Two putative BIN2 substrates are nuclear components of brassinosteroid signaling. Plant Physiol, 130, 1221-1229. 107 Chapter 4 Supplemental data Supplemental Table 1. Primers used for PCR Gene Primer Sequence 5´-3´ Gene Primer Sequence 5´-3´ RACE1 P007 GACTCGAGTCGACATCGATTTTTTTTTTTTTTT PhCYP85A1 M673 GGACTGGTCCCAGGATATCCACAGTC RACE2 P008 GACTCGAGTCGACATCGA M674 CATAAGGTAACCAAGCATATCATGTTG dTph1 OUT6 P016 CCTTTGCACCAAGGAGCTCCGCC M698 TGCTGCTGTCCATGGTTCTGCTCACA dTph1 OUT11 P423 CGAAGGGGTGTCAATGCTG M699 TCAATTAGTGTCCTTAGCAAGCTGAC dTph1 OUT12 P630 MCAGCATTGACACCCCTTC M810 GGCATGAATTCTTGAGCTAAGGAG dTph1 OUT15 P1114 TATTTGAACRGTTGTCCTCTTGAA M811 CTTCCCAATACTAACTATCGTTG PhBEH2 M711 GCTCTTTGTACTGAAGCTGGTTGG M815 CAATCCCAGCAATTTGTTTCAATAGTG M712 GATGGTACCACTTATCGCAAGG PhBRI1 PSK8 PSK9 PhCPD 108 M656 CATTTCCCACCTTCATCTTATGTC M717 GATGTATGACGACGGATTTCCGTCACAG M657 GCTCGAGGAACCAAATTAGCACA M718 CTGGCATGGTAAGAAGGTACAGGA M679 CTTACGGCAAACACTACACTTCAGCTG M736 GGGAATTCACAGCTGCCGGCGGTGGAGGTG M703 AGCCAATTCGAGACCTCTGCCTATG M737 GGGGTCGACGAAGCTCCTTCAAGCATGTGCC M704 CAGCAAAAGACCATGTCACTAACCATTC M631 GTTGAAGGATGGCAGTGTCGTAG M662 GAATAATGGCTTCCAGGTATTGGGT M632 GAGTGTCCATAGCACTCATGAG M663 CGAGAATGGAACGTACTTTCTTCTC M694 TCAGGTCAGCGGACAGGGTGATCG M680 CATTGAAGGGGTCTGCTCCAAGACAC M695 TGTAGGAAAGCCAACCCTCTAGCA M706 TGTTACAGTATAGAGAGCATACTC M709 GGCTCGAGGGATGGACATTCTCATTCTGCAAC M658 GACCTTCTTGTCCAGTTCAGATG M714 CAACCCAAAGATGCAGAAGAGGGAG M659 GATCAATCTTCAGGCTCATCAAC M719 CCTGCTCCATGGCACTCCTTACTCC M678 GACGCCAACATCAGTATACATTTGAG M723 GGCAATTGGATGGTCATTCTCATTC M689 GCTCAAATGTATACTGATGTTGGCGTC M741 TCCTTCAGCTTTGACAATGACC M705 GGGAGCTTGATGATCTGACCGTTG M743 CACCAAGTTGGCTTGCCAACACACC M747 CTCTTTGAGGCGCTTCACGACCTC M746 GAGGACACCAATGTAAAGTCAACAC M753 GAACGAGCACGCTTATAGTCAC M746B TATTTGAACRGTTGTCCTCTTGAA M775 TTGAACGGTTGTCCTCTTGAAC M826 CTTTTCTACGACTAGTAGAGATGAGC M814 GACGCAGTTGGTGGACTTTCTTGTC M834 CAAGGTCCTTGGTACTCCTACTCG M654 CTCCAACTGGACATAGGCCTGTCT M738 GGGAATTCCTGTTAAGGGATGTGAATGATG M655 CTCTTGCCTCTGCCTGTCACAATC M744 CTGAACGCGTAGTGGGAACTGGATC M696 GGGATGGAGTTATATCCTCGTATTG M745 GGGTCGACCTTCACGTCATGGCATCGTGTG M697 GTGCAATATCCATTGTGTACCAGTAC PhDET2 PhDET3 PhDIM PhDWARF5 PhDWARF7 M660 CTTGAAGCTATTCGTGGAGGAGAC M781 GCTTCGGCTGGCCATTCTTCCCTC M782 GGTTGTGGAAATTATATGACCAGTGAC M661 GAACAACAAAAGAGATGAACGTTAC M809 GGGAATTCTCTGCTCCTGTTATGGATGTGAATG M677 CTAAGTACATGATTGCAAATGGATGGAC M664 GAAAATCCTGAACCCTTCATCGATGA M701 GGCAATGAGTTATCACAGCCTTAGC M665 CTCAAACAAGCTTCGATGTTGGAC M675 M671 CTGATAATGTTCCTGCTTTTGTTGACAC CACCACCCATATATTTGGTGAACCGAC M672 GAAAGATTGACCCCAACCCCATTC M676 GGTTTCATAGCCAGCCACAAGTAATGC M702 CTTGTTACTGATATTTATTGAATGG M700 GGACCATTTGTTGGTCGATATAGAC M720 CTCTTCAGGTTTTCTGCAGAGAATC M748 AGCGAACCAGGATAACTAGATTC M797 CAACAGATGCAGAGGTGAACAAAGTGGT M752 CACACACTAGTGGAGTATATAAGAGG M798 CCTTCACCTCAGTTCACATGGATGAAG M774 CAAGTCCATGCCCTTCACTCAATG M799 CAAGCCTTGAAAGTTCTAAACCAGGAC M812 CTTCCCACTGGTAAACACTCAACTC M866 CTTCTGATACTTTTTATGTGTCTTC M867 GTTATACCATTCCAAAAGGGTGGAAG M868 CTCAACGCATCTGCCACCTTCTTTCTC M869 CTCATAACCAGGACATCGCCGTGGTC PhSMT2 PhROT3 Brassinosteroid biosynthesis and signalling in petunia Supplemental Table 2. Genbank accession numbers Gene Genbank accession number BES1/BZR1 family AtBES1 AtBZR1 AtBEH1 AtBEH2 AtBEH3 AtBEH4 NP_973863 NP_565099 NP_190644 Q94A43 NP_193624 NP_565187 BRI1 AtBRI1 LeBRI1 PsBRI1 NbBRI1 NtBRI1 OsBRI1 HvBRI1 Gene Genbank accession number GSK3/SHAGGY kinases NP_195650 AAN85409 BAC99050 ABO27628 ABR18799 NP_001044077 BAD01654 AtSK11 ASK12 ASK13 ASK31 ASK32 ASK41 ASK42 AtBIN2 AtGSK1 AtASKζ PSK4 PSK6 PSK7 NP_568486 NP_187235 NP_196968 NP_974471 NP_191981 NP_172455 NP_176096 NP_193606 NP_172127 NP_180655 CAA58594 CAA11861 CAA11862 Other AtCLV1 AtMPK1 D83257 Q39021 109 Chapter 4 putative leucine PhBRI1 NbBRI1 LeBRI1 AtBRI1 1 1 1 1 PhBRI1 NbBRI1 LeBRI1 AtBRI1 46 60 51 43 MNSQNNVIYN-------LFLLSFCLH-------LIFFLPPASPAS-INGLFKDTQQLLSF MKPHKSALYQHHCSLKKIFLLSFYLQPLFILLLIIFFLPPASPAS-VNGLLKDSQQLLSF MKAHKTVFNQHPLS---LNKLFFVLL-------LIFFLPPASPAASVNGLYKDSQQLLSF MKTFSSFFLS--------VTTLFFFS----------FFSLSFQASPSQSLYREIHQLISF 24 698 715 707 698 LGGLKNVAILDLSYNRLNGSIPNSLTSLTLLGEIDLSNNNLSGLIPESAPFDTFPDYRFA LGGLKNVAILDLSYNRLNGSIPNSLTSLTLLGELDLSNNNLTGPIPESAPFDTFPDYRFA LGGLKNVAILDLSYNRFNGTIPNSLTSLTLLGEIDLSNNNLSGMIPESAPFDTFPDYRFA VGDLRGLNILDLSSNKLDGRIPQAMSALTMLTEIDLSNNNLSGPIPEMGQFETFPPAKFL PhBRI1 NbBRI1 LeBRI1 AtBRI1 758 775 767 758 * * NN-SLCGYPL-TPCN-SGASNANLHQKSH-RKQASWQG-VAMGLLFSLFCIFGLIIVAVE NT-SLCGYPL-QPCGSVGNSNSSQHQKSH-RKQASLAGSVAMGLLFSLFCIFGLIIVAIE NN-SLCGYPLPIPCSSGPKSDANQHQKSH-RRQASLAGSVAMGLLFSLFCIFGLIIVAIE NNPGLCGYPL-PRCDPSNADGYAHHQRSHGRRPASLAGSVAMGLLFSFVCIFGLILVGRE PhBRI1 NbBRI1 LeBRI1 AtBRI1 813 832 825 817 MKKRRKKKEAALEAYMDGHSHS---ATANSAWKFTSAREALSINLAAFEXPLRKLTFADL TKKRRKKKEAALEAYMDGHSNS---ATANSAWKFTSAREALSINLAAFEKPLRKLTFADL TKKRRRKKEAALEAYMDGHSHS---ATANSAWKFTSAREALSINLAAFEKPLRKLTFADL MRKRRRKKEAELEMYAEGHGNSGDRTANNTNWKLTGVKEALSINLAAFEKPLRKLTFADL PhBRI1 NbBRI1 LeBRI1 AtBRI1 870 889 882 877 LEATNGFHNDSLIGSGGFGDVYRAQLKDGSVVAIKKLIQVSGQGDREFTAEMETIGKIKH LEATNGFHNDSLIGSGGFGDVYKAQLKDGSVVAIKKLIHVSGQGDREFTAEMETIGKIKH LEATNGFHNDSLVGSGGFGDVYKAQLKDGSVVAIKKLIHVSGQGDREFTAEMETIGKIKH LQATNGFHNDSLIGSGGFGDVYKAILKDGSAVAIKKLIHVSGQGDREFMAEMETIGKIKH PhBRI1 NbBRI1 LeBRI1 AtBRI1 930 949 942 937 RNLVPLLXYCKVGEERLLVYEYMKYGSLEDVLHDRKKNGIKLNWAARRKIAIGAARGLAF RNLVPLLGYCKVGEERLLVYEYMKYGSLEDVLHDRKKNGIKLNWHARRKIAIGAARGLAF RNLVPLLGYCKVGEERLLVYEYMKYGSLEDVLHDRKKIGIKLNWPARRKIAIGAARGLAF RNLVPLLGYCKVGDERLLVYEFMKYGSLEDVLHDPKKAGVKLNWSTRRKIAIGSARGLAF PhBRI1 NbBRI1 LeBRI1 AtBRI1 990 1009 1002 997 LHHNCIPHIIHRDMKSSNVLLDENLEARVSDFGMARLMSAMDTHLSVSTLAGTPGYVPPE LHHNCIPHIIHRDMKSSNVLLDENLEARVSDFGMARLMSAMDTHLSVSTLAGTPGYVPPE LHHNCIPHIIHRDMKSSNVLLDENLEARVSDFGMARLMSAMDTHLSVSTLAGTPGYVPPE LHHNCSPHIIHRDMKSSNVLLDENLEARVSDFGMARLMSAMDTHLSVSTLAGTPGYVPPE PhBRI1 NbBRI1 LeBRI1 AtBRI1 1050 1069 1062 1057 YYQSFRCSTKGDVYSYGVVLLELLTGRQPTDSADFGDNNLVGWVKQQ-KMKISDVFDREL YYQSFRCSTKGDVYSYGVVLLELLTGRTPTDSADFGDNNIVGWVRQHAKLKISDVFDREL YYQSFRCSTKGDVYSYGVVLLELLTGKQPTDSADFGDNNLVGWVKLHAKGKITDVFDREL YYQSFRCSTKGDVYSYGVVLLELLTGKRPTDSPDFGDNNLVGWVKQHAKLRISDVFDPEL PhBRI1 NbBRI1 LeBRI1 AtBRI1 1109 1129 1122 1117 LKEDPTIEIELLQHLKVARACLDDRHWKRPTMIQVMAMFKEIQAGSGIDSSSTIA-TDDC LKEDPSIEIELLQHLKVACACLDDRHWKRPTMIQVMAMFKEIQAGSGIDSSSTIA-ADDV LKEDASIEIELLQHLKVACACLDDRHWKRPTMIQVMAMFKEIQAGSGMDSTSTIG-ADDV MKEDPALEIELLQHLKVAVACLDDRAWRRPTMVQVMAMFKEIQAGSGIDSQSTIRSIEDG PhBRI1 NbBRI1 LeBRI1 AtBRI1 1168 1188 1181 1177 NFNAVEGGIEMGINESIKEGNELSKHL NFSAVEGGIEMGISESIKEGNELSKHL NFSGVEGGIEMGINGSIKEGNELSKHL GFSTIEM-----VDMSIKEVPEGKL-- transmembrane domain zipper motif 1 2 * * KSSLP--STTLQGLAASTDPCSYTGVSCKNSRVVSIDLSNTLLSVDFTLVSSYLLTLSNL KSSLPNTQAQLQNWLSSTDPCSFTGVSCKNSRVSSIDLTNTFLSVDFTLVSSYLLGLSNL KAALPPTPTLLQNWLSSTGPCSFTGVSCKNSRVSSIDLSNTFLSVDFSLVTSYLLPLSNL KDVLPD-KNLLPDWSSNKNPCTFDGVTCRDDKVTSIDLSSKPLNVGFSAVSSSLLSLTGL 3 25 PhBRI1 NbBRI1 LeBRI1 AtBRI1 4 PhBRI1 104 ETLVLKNANLSGSLTSASKSQCGVSLNSLDLSENTISGPVNDVSSLGSCSNLKSLNLSRN NbBRI1 120 ESLVLKNANLSGSLTSAAKSQCGVSLNSIDLAENTISGSVSDISSFGPCSNLKSLNLSKN LeBRI1 111 ESLVLKNANLSGSLTSAAKSQCGVTLDSIDLAENTISGPISDISSFGVCSNLKSLNLSKN AtBRI1 102 ESLFLSNSHING---SVSGFKCSASLTSLDLSRNSLSGPVTTLTSLGSCSGLKFLNVSSN kinase domain 5 6 PhBRI1 164 LMDSPLKEAKFQSFSLSLQVLDLSYNNISGQNLFPWLFFLRFYELEYFSVKGNKLAGTIP * NbBRI1 180 LMDPPSKEIKAS--TLSLQVLDLSFNNISGQNLFPWLSSMRFVELEYFSLKGNKLAGNIP LeBRI1 171 FLDPPGKEMLKAA-TFSLQVLDLSYNNISGFNLFPWVSSMGFVELEFFSLKGNKLAGSIP AtBRI1 159 TLDFPGKVSGGLK-LNSLEVLDLSANSISGANVVGWVLSDGCGELKHLAISGNKISGDVD 7 8 9 PhBRI1 224 ELDFKNLSYLDLSANNFSTGFPLFKDCGNLQHLDLSSNKFVGDIGGSLAACVKLSFVNLT NbBRI1 238 ELDYKNLSYLDLSANNFSTGFPSFKDCSNLEHLDLSSNKFYGDIGASLSSCGRLSFLNLT LeBRI1 230 ELDFKNLSYLDLSANNFSTVFPSFKDCSNLQHLDLSSNKFYGDIGSSLSSCGKLSFLNLT AtBRI1 218 VSRCVNLEFLDVSSNNFSTGIPFLGDCSALQHLDISGNKLSGDFSRAISTCTELKLLNIS 10 11 12 PhBRI1 284 NNMFVGFVPKLQSESLEFLYLRGNDFQGVLASQLGDLCKSLVELDLSFNNFSGFVPETLG NbBRI1 298 SNQFVGLVPKLPSESLQFMYLRGNNFQGVFPSQLADLCKTLVELDLSFNNFSGLVPENLG LeBRI1 290 NNQFVGLVPKLPSESLQYLYLRGNDFQGVYPNQLADLCKTVVELDLSYNNFSGMVPESLG AtBRI1 278 SNQFVGPIPPLPLKSLQYLSLAENKFTGEIPDFLSGACDTLTGLDLSGNHFYGAVPPFFG 13 14 PhBRI1 344 ACSKLELLDVSNNNFSGKLPVDTLLKLSNLKTLVLSFNNFIGGLPESLSSLVK-LETLDV NbBRI1 358 ACSSLELLDISNNNFSGKLPVDTLLKLSNLKTMVLSFNNFIGGLPESFSNLLK-LETLDV LeBRI1 350 ECSSLELVDISYNNFSGKLPVDTLSKLSNIKTMVLSFNKFVGGLPDSFSNLLK-LETLDM AtBRI1 338 SCSLLESLALSSNNFSGELPMDTLLKMRGLKVLDLSFNEFSGELPESLTNLSASLLTLDL 15 16 PhBRI1 403 SSNNLTGLIPSGICKDPLNSLKVLYLQNNLFTGPIPDSLGNCSRLVSLDLSFNYLTERIP NbBRI1 417 SSNNITGVIPSGICKDPMSSLKVLYLQNNWLTGPIPDSLSNCSQLVSLDLSFNYLTGKIP LeBRI1 409 SSNNLTGVIPSGICKDPMNNLKVLYLQNNLFKGPIPDSLSNCSQLVSLDLSFNYLTGSIP AtBRI1 398 SSNNFSGPILPNLCQNPKNTLQELYLQNNGFTGKIPPTLSNCSELVSLHLSFNYLSGTIP 17 18 19 PhBRI1 463 SSLGSLSKLKDLVLWLNQLSGEIPQELMYLKSLENLILDFNDLSGSIPASLSNCTNLNWI NbBRI1 477 SSLGSLSKLKDLILWLNQLSGEIPQELMYLKSLENLILDFNDLTGSIPASLSNCTNLNWI LeBRI1 469 SSLGSLSKLKDLILWLNQLSGEIPQELMYLQALENLILDFNDLTGPIPASLSNCTKLNWI AtBRI1 458 SSLGSLSKLRDLKLWLNMLEGEIPQELMYVKTLETLILDFNDLTGEIPSGLSNCTNLNWI 21 20 PhBRI1 523 SLSNNMLSGEIPASLGRLVNLAILKLK-ITQSQEYPAEWG-CQSLIWLDLNNNFLNGSIR NbBRI1 537 SMSNNLLSGEIPASLGGLPNLAILKLGNNSISGNIPAELGNCQSLIWLDLNTNLLNGSIP LeBRI1 529 SLSNNQLSGEIPASLGRLSNLAILKLGNNSISGNIPAELGNCQSLIWLDLNTNFLNGSIP AtBRI1 518 SLSNNRLTGEIPKWIGRLENLAILKLSNNSFSGNIPAELGDCRSLIWLDLNTNLFNGTIP 68 amino acid non-repetitive island PhBRI1 581 -RHVKQSGKIAVAFLTGKRYVYIKNDGSK-ECHGAGNLLEFGGIRQEQLDRISTRHPCNF NbBRI1 597 GPLFKQSGNIAVALLTGKRYVYIKNDGSK-ECHGAGNLLEFGGIRQEQLDRISTRHPCNF LeBRI1 589 PPLFKQSGNIAVALLTGKRYVYIKNDGSK-ECHGAGNLLEFGGIRQEQLDRISTRHPCNF AtBRI1 578 AAMFKQSGKIAANFIAGKRYVYIKNDGMKKECHGAGNLLEFQGIRSEQLNRLSTRNPCNI 22 23 PhBRI1 639 T-RVYRGITQPTFNHNGSMIFLDLSYNKLEGSIPKELGSMFYLSILNLGHNDLSSAIPQE NbBRI1 656 T-RVYRGITQPTFNHNGSMIFLDLSYNKLEGSIPKELGSMYYLSILNLGHNDLSGVIPQE LeBRI1 648 T-RVYRGITQPTFNHNGSMIFLDLSYNKLEGSIPKELGAMYYLSILNLGHNDLSGMIPQQ AtBRI1 638 TSRVYGGHTSPTFDNNGSMMFLDMSYNMLSGYIPKEIGSMPYLFILNLGHNDISGSIPDE Supplemental Figure 1. Alignment of the deduced amino acid sequences of BRI1 and homologs. Ph, petunia; Nb, Nicotiana benthamiana; Le, tomato and At, Arabidopsis. Identical residues in three out of four sequences are indicated in black and similar amino acids in gray. Dashed lines represent gaps to facilitate alignment. A putative leucine zipper motif, 68 amino acid non-repetitive island, N- and C- terminal LRR-flanking cysteine pairs (*), transmembrane and kinase domain are shown. Numbers 1-25 depict the first amino acid of each Leucine Rich Repeat (LRR). 110 Brassinosteroid biosynthesis and signalling in petunia PSK8 AtBIN2 PSK9 AtGSK1 AtASKdzeta 1 1 1 1 1 MASLPLAPQHHHVPPDNHHPLQHDQPPAAVKHAVAGARPEMESDKEMSAAVVEGNDAVTG ----------------------------------------MADDKEMPAAVVDGHDQVTG ----------------------------------------------MSAPVMDVNDAVTG MASLPLGP---------QPHALAP-PLQLHDGDALKRRPELDSDKEMSAAVIEGNDAVTG MTSIPLGPP--------QPPSLAPQPPHLHGGDSLKRRPDIDNDKEMSAAVIEGNDAVTG PSK8 AtBIN2 PSK9 AtGSK1 AtASKdzeta 241 201 195 231 233 ICSRYYRAPELIFGATEYTTSIDIWSAGCVLAELLLGQPLFPGENAVDQLVEIIKVLGTP ICSRFYRAPELIFGATEYTTSIDIWSAGCVLAELLLGQPLFPGENAVDQLVEIIKVLGTP ICSRFYRAPELIFGATEYTDSIDIWSAGCVLAELLLGQPLFPGENAVDQLVEIIKVLGTP ICSRYYRAPELIFGATEYTASIDIWSAGCVLAELLLGQPLFPGENSVDQLVEIIKVLGTP ICSRYYRAPELIFGATEYTSSIDIWSAGCVLAELLLGQPLFPGENSVDQLVEIIKVLGTP PSK8 AtBIN2 PSK9 AtGSK1 AtASKdzeta 61 21 15 51 53 HIISTTIGGKNGEPKRTISYMAERVVGTGSFGIVFQAKCLETGETVAIKKVLQDKRYKNR HIISTTIGGKNGEPKQTISYMAERVVGTGSFGIVFQAKCLETGETVAIKKVLQDRRYKNR HIISTTIGGKNGQPKQTVSYMAERVVGTGSFGVVFQAKCLENGETVAIKKVLQDRRYKNR HIISTTIGGKNGEPKQTISYMAERVVGTGSFGIVFQAKCLETGESVAIKKVLQDRRYKNR HIISTTIGGKNGEPKQTISYMAERVVGTGSFGIVFQAKCLETGESVAIKKVLQDRRYKNR PSK8 AtBIN2 PSK9 AtGSK1 AtASKdzeta 301 261 255 291 293 TREEIRCMNPNYTDFRFPQIKAHPWHKVFHKRMPPEAIDLASRLLQYSPSLRCTALEACA TREEIRCMNPHYTDFRFPQIKAHPWHKIFHKRMPPEAIDFASRLLQYSPSLRCTALEACA TREEIRCMNPNYTDFRFPQIKAHPWHKVFHKRMPPEAIDLASRLLQYSPSLRCTALDACA TREEIRCMNPNYTDFRFPQIKAHPWHKVFHKRMPPEAIDLASRLLQYSPSLRCTALEACA TREEIRCMNPNYTDFRFPQIKAHPWHKVFHKRMPPEAIDLASRLLQYSPSLRCTALEACA PSK8 AtBIN2 PSK9 AtGSK1 AtASKdzeta 121 81 75 111 113 ELQLMRLMDHPNVICLKHCFFSTTSRDELFLNLVMDYVPESLYKVLKHYSNSNQRMPLIY ELQLMRVMDHPNVVCLKHCFFSTTSKDELFLNLVMEYVPESLYRVLKHYSSANQRMPLVY ELQLMRTMDHPNVVSLKHCFYSTTSKNELFLNLVMEYVPETMYRMLKHYSNMNQRMPLIY ELQLMRPMDHPNVISLKHCFFSTTSRDELFLNLVMEYVPETLYRVLRHYTSSNQRMPIFY ELQLMRLMDHPNVVSLKHCFFSTTTRDELFLNLVMEYVPETLYRVLKHYTSSNQRMPIFY PSK8 AtBIN2 PSK9 AtGSK1 AtASKdzeta 361 321 315 351 353 HSFFDELREPNARLPNGRPFPPLFNFKQELTGASPELVNKLIPEHVWRQIGLNFPYPGAT HPFFDELREPNARLPNGRPFPPLFNFKQEVAGSSPELVNKLIPDHIKRQLGLSFLNQSGT HPFFDELREPNARLPNGRPLPPLFNFKQELSGSSPDLINRLIPEHIKRQMGLQFFTSHDA HPFFNELREPNARLPNGRPLPPLFNFKQELGGASMELINRLIPEHVRRQMSTGLQNS--HPFFNELREPNARLPNGRPLPPLFNFKQELSGASPELINRLIPEHVRRQMNGGFPFQAGP PSK8 AtBIN2 PSK9 AtGSK1 AtASKdzeta 181 141 135 171 173 VKLYMYQIFRGLAYIHNVPRVCHRDVKPQNLLVDPLTHQVKLCDFGSAKVLVNGEANISY VKLYMYQIFRGLAYIHNVAGVCHRDLKPQNLLVDPLTHQVKICDFGSAKQLVKGEANISY VKLYIYQVFRGLAYMHTVAGVCHRDLKPQNILVDPVTHQVKICDFGSAKLLVKGEANISY VKLYTYQIFRGLAYIHTVPGVCHRDVKPQNLLVDPLTHQVKLCDFGSAKVLVKGEPNISY VKLYTYQIFRGLAYIHTAPGVCHRDVKPQNLLVDPLTHQCKLCDFGSAKVLVKGEANISY PSK8 AtBIN2 PSK9 AtGSK1 AtASKdzeta -381 -375 MT -413 -- * PS H2 K9 Ph 14 BD -33ε BE Ph Ph BR I1 cy t Supplemental Figure 2. Alignment of the deduced amino acid sequences of class II GSK3/SHAGGY-like kinases from Arabidopsis and petunia. Ph, petunia; At, Arabidopsis. Identical residues in three out of five sequences are indicated in black and similar amino acids in gray. Dashed lines represent gaps to facilitate alignment. The red asterisk marks the phospho-tyrosine residue (pTyr200, Kim et al., 2009). - TX-gal - TH - THA Supplemental Figure 3. Yeast one-hybrid to test for auto-activation of baits. The cytoplasmic domain of PhBRI1 and the full-length sequences of PhBEH2, PSK9 and Ph14-3-3ε were fused to the GAL4-binding domain. Auto-activation was determined by growth on medium lacking histidine (-TH), histidine and adenine (-THA) and by blue colouring using 5-bromo-4-chloro-3-indolyl-β-D-galactosidase (-TX-gal). 111 Summary As the world population is growing continuously, food production needs to keep pace. Apart from improved fertilization and cultivation techniques, plant breeding greatly improved productivity. One of the major abiotic stresses that affects productivity in the agricultural sector is soil salinity. Desertification, floods and unsuitable irrigation methods increase the amount of salt all over the world. Removal of salt from soil is costly and therefore it is economically important to find solutions that allow crops to grow in salty habitats. Currently two genetic approaches are used to improve salt stress tolerance in crops: 1) direct selection of existing plant varieties that are salt tolerant and cross them into cultivars and 2) the generation of transgenic plants to introduce novel genes or alter the level of expression of existing genes that play a role in salt stress tolerance. The initial goal of my PhD research project was to investigate the role of the GSK3-like kinase GSK1 in salt stress signalling by using the model plant Arabidopsis thaliana. The Arabidopsis genome contains ten GSK3-like kinases that based on phylogenetic analysis can be classified into four subgroups. BIN2, a group II member, is the best-studied GSK3like kinase and plays a negative role in the brassinosteroid (BR) signalling pathway. Based on previous research by another group we expected that mutation of the GSK1 gene would result in a plant more susceptible to salt stress. We considered this might be the same for two other GSK3-like kinases (ASKζ and BIN2) as they are very similar to GSK1, all belonging to the group II GSK3-like kinases. In Chapter 2, we checked whether mutation of these genes altered the response of Arabidopsis towards salt. In contrast to prior research, we could not detect a role for GSK1 in tolerance against Na+ stress, neither for the other group II members ASKζ and BIN2. However, based on research by others, the involvement of not only group II GSK3-like kinases but also GSK3-like kinases of other subgroups in salt tolerance seems likely. Instead of playing a positive role in salt stress resistance, GSK3-like kinases seem to play a negative role. The potential role of GSK3-like kinases in salt stress signalling might be an indirect effect that originates from the role that we and others discovered for GSK1 as well as ASKζ in BR signalling. Brassinosteroids (BR) are steroid hormones that are essential for plant growth and development. Chapter 1 reviews the current knowledge about BR biosynthesis, BR signalling and its involvement in various growth- and developmental processes such as cell 113 Summary division and elongation, vascular differentiation, abiotic- and biotic stresses, photomorphogenesis, germination, senescence and fertility. BIN2 negatively regulates the BR signalling pathway by phosphorylating the transcription factors BES1 and BZR1. In Chapter 2 we show that GSK1 as well as ASKζ interact with the BES1/BZR1 transcription factor family. The fact that group II GSK3-like kinases have redundant roles in the BR signalling pathway was confirmed by studying several kinds of mutants. In order to learn more about the mode of actions of GSK1 and the group II GSK3-like kinases in BR signalling and possibly in other signal transduction pathways we tried to purify protein complexes containing group II GSK3-like kinases via various Tandem Affinity Purification (TAP) methods which are described in Chapter 3. Till date, TAP purification was unsuccessful due to technical complications and maybe also because the TAP tagged versions of GSK3-like kinases seem to be biologically inactive. BR biosynthesis and signalling has been extensively studied in Arabidopsis and to some extent in other plant species such as tomato, rice and Catharanthus roseus. Within our petunia collection we found several mutant families that resemble sterol/BR biosynthesis and signalling mutants and this tempted us to research whether the BR biosynthesis and signalling pathways are also conserved in petunia (Chapter 4). Cross-pollination between the mutant families revealed five different complementation groups. Four mutant groups clearly responded to spraying with the most biologically active BR, 24-epibrassinolide, by elongation of their internodes, petioles and leaves. Complementation of the mutant phenotype in these plants indicated that these are sterol/BR biosynthesis mutants. For one mutant group we could reveal that this mutant phenotype was caused by the insertion of a transposable dTph1 element in the BR biosynthesis gene CPD. One mutant group did not respond to exogenous application of 24-epibrassinolide and therefore it was considered to be a BR signalling mutant. Interestingly, this family turned out to have a dTph1 transposon insertion in a gene homologous to the BR receptor BRI1 from Arabidopsis. With help of a PCR-based approach we succeeded to isolate a member of the BES1/BZR1 transcription factor family (PhBEH2) and a group II GSK3-like kinase (PSK9) from petunia. In a yeast two-hybrid experiment PhBEH2 interacted with another group II GSK3like kinase (PSK8) and 14-3-3 proteins, supporting the idea that the BR signalling pathway is conserved among the plant kingdom. This was further supported by the fact that PSK8 could interact with members of the BES1/BZR1 family from Arabidopsis and BIN2 from Arabidopsis with PhBEH2 from petunia. Interestingly PhBEH2 also interacted with proteins involved in gibberellin (GA), auxin (IAA) and abscisic acid (ABA) signalling and a protein that plays a role in the autonomous flowering pathway. These interactions reveal new mechanisms, not recognised from 114 Summary Arabidopsis research that suggest extensive cross-talk between BR signalling and other (hormone) signalling pathways. This thesis provides some knowledge about BR biosynthesis and signalling in Arabidopsis and petunia. For a long time BRs are applied in the agriculture and horticulture to increase crop yield and quality. With help of the acquired knowledge about the BR pathway, superior crops could be developed in the future with a desirable stature, high yield and good quality. 115 Samenvatting Brassinosteroïden biosynthese en signalering in Arabidopsis thaliana en Petunia hybrida Aangezien de wereldbevolking alsmaar toeneemt, is het van belang om de voedselproductie op peil te houden. Naast verbeterde bemestingstechnieken en teeltmethoden, speelt ook plantenbiotechnologie een grote rol in de vergroting van de gewassenopbrengst. Door slechte irrigatiemethoden, overstromingen en woestijnvorming wordt de landbouwgrond steeds zouter en dit is erg nadelig voor het verbouwen van gewassen. Het verwijderen van zout uit de bodem is erg duur en daarom is het van economisch belang om oplossingen te vinden die er voor kunnen zorgen dat gewassen in verzilte gebieden kunnen groeien. Op dit moment worden twee technieken gebruikt om zouttolerantie in gewassen te vergroten: 1) directe selectie van bestaande variëteiten die al zouttolerant zijn en deze tolerantie door middel van kruisen over te brengen in landbouwgewassen en 2) de generatie van transgene planten door nieuwe genen te introduceren of het expressieniveau te wijzigen van genen die een rol spelen in zoutstress tolerantie. Het aanvankelijke doel van dit PhD onderzoeksproject was het onderzoeken van de rol van de GSK3 kinase GSK1 in zoutstress signalering in de modelplant Arabidopsis thaliana. In het genoom van Arabidopsis zijn 10 GSK3 kinasen geïdentificeerd die op basis van hun fylogenie in 4 subgroepen kunnen worden verdeeld. BIN2 is de best bestudeerde GSK3 kinase en speelt een negatieve rol in de brassinosteroïden (BR) signaal transductie. Op basis van voorafgaand onderzoek door een Koreaanse onderzoeksgroep kregen wij het idee dat het uitschakelen van het GSK1 gen zou resulteren in een zoutgevoelige plant. Aangezien BIN2 en een andere GSK3 kinase ASKζ heel erg veel op GSK1 lijken, zou dit ook voor hun kunnen gelden. In hoofdstuk 2 hebben we bekeken of mutaties van deze genen invloed hadden op de zouttolerantie in Arabidopsis. In tegenstelling tot het Koreaanse onderzoek vonden wij geen rol voor GSK1 in zoutstress tolerantie en ook niet voor de andere GSK3 subgroep II leden BIN2 en ASKζ. Recente onderzoeksresultaten van andere onderzoekers suggereren dat niet alleen subgroep II betrokken is in zouttolerantie, maar ook andere subgroepen. Daarnaast lijkt het er ook sterk op dat GSK3 kinasen in plaats van een positieve rol juist een negatieve rol spelen in zoutstress signalering. Een vraag die nog beantwoord dient te worden is of GSK3 kinasen direct de zout tolerantie reguleren of indirect via de BR signaal transductie cascade waar GSK3 kinases ook een belangrijke rol spelen. 117 Samenvatting BRs zijn steroïde hormonen die van essentieel belang zijn voor de groei en ontwikkeling van een plant. Hoofdstuk 1 geeft een samenvatting weer over de huidige kennis omtrent BR biosynthese, BR signalering en de rol van BRs in verschillende groei en ontwikkelingsprocessen zoals celdeling en celstrekking, vaatbundel ontwikkeling, abiotische en biotische stress, fotomorfogenese, kieming, veroudering en vruchtbaarheid. BIN2 reguleert op een negatieve wijze de BR signaal transductie cascade door de transcriptiefactoren BES1 en BZR1 te fosforyleren. In hoofdstuk 2 laten wij zien dat GSK1 zowel als ASKζ ook een interactie kunnen aangaan met leden van de BES1/BZR1 transcriptiefactor familie. Dat BIN2, GSK1 en ASKζ een mutuele rol spelen in BR signalering wordt verder bevestigd door een onderzoek aan verschillende soorten mutanten. Om meer te weten te komen over de functie van GSK1 en de andere groep II GSK3 kinasen in BR signalering en mogelijk in andere signaal transductie ketens hebben we groep II GSK3 kinasen gemerkt met verschillende Tandem Affinity Purification (TAP) tags in Arabidopsis en getracht bindingspartners te isoleren (hoofdstuk 3). Het is ons niet gelukt om bindingspartners te identificeren, dit vanwege technische complicaties en mogelijk ook doordat de GSK3 kinasen inactief lijken te zijn door de TAP tag die er aan vast zit. BR biosynthese en signalering is uitgebreid bestudeerd in Arabidopsis en in enigermate in plantensoorten zoals tomaat, rijst en roze maagdenpalm. In onze petunia collectie vonden wij verschillende mutanten families die defect lijken te zijn in de sterol/BR biosynthese of in de BR signaal transductie keten en dit stimuleerde ons om te onderzoeken of in petunia de BR biosynthese en BR signaal transductie keten geconserveerd zijn (hoofdstuk 4). Kruisbestuiving tussen de verschillende mutanten families leverde vijf verschillende complementatie groepen op. Vier van die groepen bleken sterol/BR biosynthese mutanten te zijn, omdat ze complementeerden na gesprayd te zijn met de meest actieve BR, 24epibrassinolide. Voor één groep mutanten konden we achterhalen dat het dwerg fenotype was ontstaan door een dTph1 transposon insertie in het BR biosynthese gen CPD. Eén groep mutanten herstelde niet na gesprayd te zijn met 24-epibrassinolide en zou dus een BR signaleringsmutant moeten zijn. Inderdaad, deze groep mutanten bleek een transposon insertie te hebben in een gen dat homoloog is aan de BR receptor BRI1 uit Arabidopsis. Met behulp van PCR gebaseerde methoden hebben wij een lid van de BES1/BZR1 transcriptiefactor familie (PhBEH2) en een groep II GSK3 kinase (PSK9) uit petunia weten te isoleren. In een gist ´two-hybrid´ systeem, een techniek om eiwit-eiwit interacties te bepalen, vertoonde PhBEH2 interactie met een andere GSK3 kinase (PSK8) en 14-3-3 eiwitten. De vondst van deze eiwitten insinueert dat de BR signaal transductie route geconserveerd is in het plantenrijk en dit wordt verder bekrachtigd door het feit dat PSK8 van petunia in gist een eiwit-eiwit interactie kan aangaan met leden van de BES1/BZR1 familie uit Arabidopsis en BIN2 van Arabidopsis met PhBEH2 van petunia. 118 Samenvatting PhBEH2 ging ook interacties aan met eiwitten die betrokken zijn in bijvoorbeeld gibberelline, auxine en abscisinezuur signalering. Deze interacties zijn nog niet eerder gevonden en mogelijk dat de BR signaleringsroute en andere (hormoon) signaal transductie ketens met elkaar kunnen communiceren via deze eiwit-eiwit interacties. Dit proefschrift levert informatie op over BR biosynthese en signalering in Arabidopsis en petunia. BR bevattende oplossingen worden al een tijd gebruikt in de horticultuur en agricultuur om de opbrengst en kwaliteit van gewassen te verhogen. Met behulp van alle kennis die is en wordt opgedaan omtrent de BR biosynthese- en signaleringsroute, kan in de toekomst de BR route zodanig genetisch gemanipuleerd worden dat dit kan leiden tot de komst van superieure gewassen met een juist formaat, een hoge productie en kwaliteit. 119 Dankbetuiging Vanzelfsprekend zou dit proefschrift niet tot stand zijn gekomen zonder de bijdrage en steun van velen en hiervoor wil ik iedereen van harte bedanken, maar een aantal mensen toch nog even in het speciaal. Mijn dank gaat allereerst uit naar mijn co-promotor Erik en promotor Ronald die mij in de gelegenheid hebben gesteld om het onderzoek te verrichten dat tot dit proefschrift leidde. Erik, omdat we samen aan het project werkten waren we echt op elkaar aangewezen. Gedurende die periode heb ik veel van je geleerd. Je hebt me in mijn werk veel vrijheid en vertrouwen gegeven, waardoor ik de gelegenheid heb gehad om me goed te ontplooien. Daarnaast heb ik veel waardering voor de opbouwende kritiek die je me gaf tijdens het schrijven van mijn proefschrift. Zonder jou was dit proefschrift nooit zo geworden zoals het nu is. Alhoewel mijn naaste collega´s niet direct betrokken waren bij het onderzoek, hebben zij altijd voor me klaar gestaan om me bijvoorbeeld nieuwe technieken aan te leren of me raad te geven. Dus Jan, Toon, Francesca, Tarcies , Xana, Walter, Elske, Rob, Tijs, Kees ,Ilja etc. heel erg bedankt voor alles! Voor de goede zorg van mijn plantjes in de kas wil ik Maartje, Daisy, Pieter en Martina bedanken. En Joke, ik kwam maar al te graag bij je op het kantoortje langs om effe lekker te beppen. Bedankt voor alle administratieve rompslomp waar je me bij hebt geholpen en bij het voorbereiden van mijn promotiefeest. Natuurlijk wil ik ook nog alle MolMic en H0 mensen bedanken voor de hulp en gezelligheid. Henk, wil ik in het bijzonder bedanken voor zijn hulp met betrekking tot de zoutstressexperimenten en voor het gebruik maken van je klimaatkamers. Ik ben zeer vereerd dat je deel uitmaakt van mijn promotiecommissie. Tijdens mijn VU tijd heb ik ook studenten mogen begeleiden. Over het algemeen pakte dit positief uit. Annegret, many thanks for your good work. I wish you all the luck with your PhD. Allan, bedankt voor het deelnemen aan het petuniaonderzoek. Jammer, dat je niet door bent gegaan voor je Master. Ik wens je heel veel succes toe met je opleiding tot piloot. Katja, destijds de stagiaire van Jan, wil ik bedanken voor de gezellige tijd op het lab. Heel veel succes met je AIO! Mies, ik had nooit gedacht dat ik zo’n goede vriend zou tegenkomen op mijn werk. Ik kan niet alleen verschrikkelijk met je lachen, maar ook zoveel met je delen. Ook al spreek en zie ik je nog regelmatig, ik mis ons snackie van de week, de foute plaatjes die we draaiden 121 Dankbetuiging en onze dagelijkse onderonsjes. Als ik het even niet meer zag zitten, wist je me altijd op te beuren. Ik ben ingenomen met het feit, dat je als paranimf aan mijn zijde zult staan. Een andere collega moet ik ook niet vergeten. Lieve Bets, ik heb me van begin af aan altijd op mijn gemak gevoeld bij jou. Ik vond het heerlijk om met jou op het lab te werken en lekker tussendoor te babbelen over onze poezenbeesten en natuurlijk niet te vergeten Expeditie Robinson. Ik vind het een hele grote eer dat jij mijn paranimf wilt zijn. Als ik terugdenk aan de vier jaren op de VU roept dit warme herinneringen op. Ik denk dan terug aan de sinterklaasavondjes, kerstdiners, het heerlijke eten van Francesca tijdens de vele meetings, het labtuinieren van Tarcies, de fietspuzzeltocht en niet te vergeten alle fijne kletspraatjes die ik op het lab, in de wandelgangen, koffiekamer, kas tot in de kelders van de VU heb gehad. Ook buiten de VU om heb ik veel steun gehad tijdens het schrijven van mijn proefschrift. Sylvia, ik wil jou in het bijzonder bedanken voor jouw hulp bij het werken met Photoshop en Illustrator. Als ik jou niet had gehad, was ik nu nog bezig geweest! Vrienden en collega´s van NSure, bedankt voor jullie belangstelling en luisterend oor, maar vooral ook voor jullie gezelligheid waar ik mijn energie uit kon putten. Roland, ik wil je bedanken voor je steun, geduld en begrip. Je interesse lag niet zo zeer bij de aard van mijn werk, maar wel hoe het met je grote zus ging. Dit was heel belangrijk voor mij. Pap en mam, jullie onvoorwaardelijke steun en liefde heeft er voor een heel groot deel aan bijgedragen dat dit proefschrift nu af is. Dankzij jullie bleef ik met beide benen op de grond staan en heb ik me kunnen ontwikkelen tot de persoon die ik nu ben. Ik draag dit proefschrift dan ook graag aan jullie op. Ten slotte, wil ik mijn opa en oma bedanken voor hun morele en financiële steun. Wat fijn zou het zijn geweest als ze mijn promotie nog hadden mogen meemaken. 122