CHARLES UNIVERSITY IN PRAGUE, FACULTY OF SCIENCE Department of Biochemistry REGULATION OF CYTOKININ METABOLISM IN TOBACCO PLANTS AND CHLOROPLASTS Ph.D. THESIS MARIE HAVLOVÁ Supervisor: RNDr. Radomíra Vaňková, CSc. Prague 2011 Declaration I hereby declare that I completed this PhD. thesis independently, except where explicitly indicated otherwise. It documents my own work, carried out under the supervision of RNDr. Radomíra Vaňková, CSc. Throughout, I have properly acknowledged and cited all sources used. Neither this thesis nor its substantial part have been submitted to obtain this or other academic degree. Prague................................ ........................................... Marie Havlová Declaration of co-authors I hereby declare that Marie Havlová/Nováková has substantially contributed to all 4 articles which represent an integral part of this Ph.D. thesis. She performed most of the experiments, especially in the papers where she is the first author and she actively participated in the set-up of the experiments, in the interpretation of the results and in the preparation of the manuscripts. Prague................................ ........................................... RNDr. Radomíra Vaňková, CSc. Aknowledgement: I am heartily thankful to my supervisor, RNDr. Radomíra Vaňková, CSc., without whose encouragement, guidance and support from the initial to the final level I would not have been able to successfully complete my studies. I would like to thank all my dear collegues who helped me with particular experiments and stimulated to my further work, namely RNDr. Alena Gaudinová and Eva Kobzová. Special thanks belongs to ing. Petre I. Dobrev, CSc. for HPLC and to ing. Jiří Malbeck, CSc. for mass spectroscopy analysis. I also would like to make a special reference to Prof. RNDr. Danuše Sofrová, CSc., who was always so kind and prepared to give me an advice. Finally, and most importantly, I would like to thank my husband Jan and children Jindřich and Alžběta, who had to be very, very patient with me during last few months. And last but not least, I thank my parents for their assistance during my whole study. The work was supported by Grant Agency of CR, grant no: 522/04/0549 and 206/06/1306. ABSTRACT Cytokinins (CKs) are one of the most important group of phytohormones influencing many processes throughout the whole plant. As many processes are regulated both by the light and phytohormones, the first part of this work has been focused on evaluation of diurnal rhytmicity in levels of cytokinins and other cooperating hormones like auxin (indol-3-acetic acid, IAA), abscisic acid (ABA) and polyamines (PA). The changes in activity of selected enzymes participating in metabolism of the above mentioned phytohormones were followed as well. Diurnal variation of phytohormones was tested in tobacco leaves (Nicotiana tabacum L. cv. Wisconsin 38) grown under a 16/8 h (light/dark) period. The main peak of the physiologically active forms of CKs, found after the middle of the light period, coincided well with the maximum of IAA and PA levels and with activity of the corresponding enzymes. The achieved data indicate that metabolism of CKs, IAA and PAs is tightly regulated by the circadian clock. The other part of the study has been focused on changes in the contents of CKs, IAA and ABA in transgenic tobacco plants with altered cytokinin metabolism, achieved via the over-expression of particular enzymes participating in CK metabolism (biosynthesis, degradation and reversible conjugation). As CKs are known to be involved in plastid differentiation and function as well as part of CK biosynthetic pathway is localized in plastids, the impact of modulation of CK metabolism in cytoplasm on CK levels in the whole leaves and isolated chloroplasts was compared. The achieved results suggest that plant hormone compartmentation plays an important role in hormone homeostasis and that the chloroplasts are relatively autonomous organelles with respect to regulation of CK metabolism. Finally, we studied the impact of water deficit progression on CK, IAA and ABA levels in upper, middle and lower leaves and roots of wild type tobacco plants and two transformants overexpressing trans-zeatin O-glucosyltransferase gene (ZOG1) either under constitutive 35S or senescence-inducible SAG12 promoters. During drought stress, ABA content strongly increased, especially in upper leaves. A delay in ABA elevation was observed in plants with constitutive overexpression of ZOG1, which indicates that high CK levels might postpone stress sensing. As drought progressed, content of bioactive CKs in leaves gradually decreased, being maintained longer in the upper leaves of all tested genotypes. Establishment of active CK gradient in the favour of upper leaves suggests preferential protection of these leaves. During drought, significant accumulation both of CKs and IAA occured in roots, indicating the role of these hormones in root response to severe drought, which involves the stimulation of primary root growth and inhibition of lateral root initiation and formation. ABSTRAKT Cytokininy (CK) jsou jednou z nejvýznamnějších skupin rostlinných hormonů, které ovlivňují mnoho fyziologických dějů probíhajících v rostlinách. Jelikož je známo, že mnoho procesů je regulováno jak fytohormony, tak světlem, byla první část této práce zaměřena na diurnální rytmy regulace hladin jednotlivých cytokininů a dalších fytohormonů - auxinu (kyseliny indol-3-octové, IAA), kyseliny abscisové (ABA) a polyaminů (PA). Současně byly sledovány i změny v aktivitách enzymů účastnících se metabolismu CK a PA. Diurnální změny byly měřeny v listech rostlin tabáku (Nicotiana tabacum L. cv. Wisconsin 38), které byly pěstovány v režimu osvětlení 16/8 h světlo/tma. Maximální hladina fyziologicky aktivních CK, naměřená po 9 h světla (tj., uprostřed světelné fáze), velmi dobře odpovídá maximům IAA a PA a je rovněž v souladu s aktivitami odpovídajících enzymů. Tyto údaje nasvědčují tomu, že metabolismus výše uvedených hormonů vykazuje diurnální rytmus. Další část práce byla zaměřena na změny hladin CK, IAA a ABA v transgenních rostlinách tabáku, které měly pozměněný metabolismus cytokininů způsobený zvýšenou expresí vybraných enzymů zapojených do jejich metabolismu (biosyntézy, degradace a reverzibilní konjugace).Vzhledem k tomu, že CK hrají důležitou roli ve funkci a diferenciaci plastidů a zároveň velká část biosyntézy CK probíhá právě v plastidech, porovnávali jsme dopad modulace metabolismu CK v cytoplazmě na hladiny CK na úrovni listů a izolovaných chloroplastů. Výsledky svědčí o tom, že kompartmentalizace CK hraje velmi významnou roli v jejich homeostázi a zároveň, že chloroplasty jsou relativně autonomními organelami vzhledem k regulaci metabolismu cytokininů. Poslední část této práce se zabývala dopadem stresu z nedostatku vody na hladiny CK, IAA a ABA na horní, střední a dolní listy a kořeny kontrolních rostlin tabáku a dvou transformantů s nadprodukcí genu pro trans-zeatin O-glukosyltransferasu, a to buď pod konstitutivním 35S nebo indukovatelným SAG12 promotorem, který spouští expresi genu až během postupující senescence. Během sucha hladiny ABA velmi výrazně vzrostly, zvláště v horních listech. U rostlin s konstitutivní expresí ZOG1, a tedy se zvýšeným „turn-over“ CK již před začátkem stresu, byl tento nástup pozorován se zpožděním. S prohlubujícím se stresem obsah bioaktivních CK v listech postupně klesal, nejpomaleji v horních listech, a to u všech studovaných genotypů. Ustavení gradientu aktivních CK ve prospěch nejmladších listů naznačuje jejich prioritní ochranu. Během sucha byla dále pozorována významná akumulace CK a IAA v kořenech, což poukazuje na jejich roli při změně morfologie kořenového systému v reakci na silný stres suchem, zejména stimulaci růstu primárních kořenů a inhibici zakládání a růstu laterálních kořenů. ABBREVIATIONS ABA ADC ADP AHK AHK4/CRE1/WOL AHP AMP APRR ARR AtCKX AtIPT ATP cAMP Cad Cdc2 CDK CK CKX CYCD3 cZ DAO DHZ DMAPP DNA dSAM GARP domain HPt IAA i6Ade i6Ado IPT iP iP7G iPMP iPR iPRMP iPRDP iPRTP LOG MEP MVA mRNA NADPH ODC PA Put PCA abscisic acid arginine decarboxylase adenosine 5´-diphosphate Arabidopsis thaliana histidine kinases cytokinin receptor (histidine kinase 4/cytokinin response/wooden leg) Arabidopsis thaliana HPt adenosine 5´-monophosphate Arabidopsis thaliana pseudo response regulator Arabidopsis thaliana response regulator Arabidopsis thaliana cytokinin oxidase/dehydrogenase Arabidopsis thaliana isopentenyl transferase adenosine 5´-triphosphate cyclic adenosine 5´-monophosphate cadaverine gene encoding the major CDK cyclin-dependent protein kinase cytokinin cytokinin oxidase/dehydrogenase D-type cyclin cis-zeatin diamine oxidase dihydrozeatin dimethylallyl diphosphate deoxyribonucleic acid decarboxylated S-adenosylmethionine a plant-specific, conserved DNA binding domain protein with a histidine-containing phosphotransmitter domain indol-3-acetic acid N6-(2-isopentenyl)-adenine N6-(2-isopentenyl)-adenosine isopentenyl transferase N6-(2-isopentenyl)-adenine N6-(2-isopentenyl)-adenine 7-glucoside isopentenyladenosine-5´-monophosphate iP riboside iP riboside 5´-monophosphate iP riboside 5´-diphosphate iP riboside 5´-triphosphate cytokinin nucleoside 5´-monophosphate phosphorylase (called “lonely guy”) methylerythritol phosphate mevalonate messenger ribonucleic acid nicotinamide adenine dinucleotide phosphate (reduced form) ornithine decarboxylase polyamine putrescine perchloric acid Pssu-ipt RT-PCR RuBisCO SAG Sho Spm Spd T-DNA Ti-plasmid tRNA tRNA-IPT tZ UDPG UDPX ZOG1 Zm-p60.1 non-rooting transgenic tobacco with ipt under the control of the promoter of small subunit of RuBisCO reverse transcription polymerase chain reaction ribulose-1,5-bisphosphate carboxylase/oxygenase senescence-associated gene („shooting“) IPT homologue gene from Petunia hybrida spermine spermidine transfer deoxyribonucleic acid from „tumor inducing“ plasmid transfer ribonucleic acid tRNA-isopentenyl transferase trans-zeatin uridine diphosphoglucose uridine diphosphoxylose zeatin O-glucosyltransferase gene for β-glucosidase CONTENT 1. INTRODUCTION.............................................................................................1 1. 1. Discovery of cytokinins..........................................................................................1 1. 2. Structure of cytokinins.........................................................................................2 1. 3. Cytokinin metabolism............................................................................................2 1. 4. Degradation of cytokinins.....................................................................................6 1. 5. Biosynthesis of cytokinins.....................................................................................7 1. 6. Cytokinins and biological functions.....................................................................11 1. 6. 1. Cytokinins and cell cycle regulation...............................................................11 1. 6. 2. Cytokinins and leaf senescence.......................................................................11 1. 6. 3. Cytokinins and chloroplast development........................................................12 1. 6. 4. Cytokinins and movement of nutrients...........................................................12 1. 6. 5. Cytokinins and signalling................................................................................13 1. 7. Polyamines – plant growth regulators.................................................................15 1. 8. Polyamine metabolism...........................................................................................16 2. AIMS AND SCOPES.........................................................................................17 3. PLANT MATERIALS AND METHODS........................................................18 3.1. Growing conditions and collecting samples..........................................................18 3. 1. 1. Isolation of CKs and study of diurnal rhythms...............................................18 3. 1. 2. Isolation of PAs and study of diurnal rhythms................................................18 3. 1. 3. Isolation of CKs and study of altered CK metabolism....................................18 3. 1. 4. Isolation of CKs and study of water deficit.....................................................19 3. 1. 5. Incubation of samples with tritiated trans-zeatin............................................19 3. 1. 6. Chloroplast isolation........................................................................................19 3. 2. Plant hormone extraction and purification.........................................................19 3. 2. 1. Extraction and purification of CKs.................................................................19 3. 2. 2. Extraction and purification of PAs..................................................................19 3. 3. Chromatographic methods...................................................................................19 3. 3. 1. HPLC of radiolabelled trans-zeatin and its metabolic products.....................19 3. 3. 2. HPLC of auxin and abscisic acid....................................................................19 3. 3. 3. HPLC/MS........................................................................................................19 3. 3. 4. PA analysis.......................................................................................................19 3. 3. 5. Chlorophyl determination................................................................................19 3. 3. 6. Chloroplast integrity........................................................................................19 3. 4. Determination of enzymes activities.....................................................................19 3. 4. 1. Determination of CKX activity.......................................................................19 3. 4. 2. Determination of β-glucosidase activity.........................................................19 3. 4. 3. Determination of ODC, ADC activity.............................................................19 3. 4. 4. Determination of DAO activity.......................................................................20 3. 5. Microscopic methods.............................................................................................20 3. 5. 1. Transmission electron microscopy..................................................................20 3. 6. Molecular methods.................................................................................................20 3. 6. 1. RNA extraction and reverse transcription.......................................................20 3. 6. 2. qRT-PCR..........................................................................................................20 3. 7. Determination of relative water content..............................................................20 4. RESULTS AND DISCUSSION........................................................................21 4. 1. Diurnal rhythms of cytokinins, auxin and abscisic acid.....................................21 4. 2. Diurnal rhythms of polyamines............................................................................23 4. 3. The study of chloroplast autonomy in regulation of CK levels..........................25 4. 4. The role of CK in responses to water deficit........................................................27 5. SUMMARY........................................................................................................31 LITERATURE.......................................................................................................32 SUPPLEMENT 1...................................................................................................48 SUPPLEMENT 2...................................................................................................55 SUPPLEMENT 3...................................................................................................64 SUPPLEMENT 4...................................................................................................77 1. Introduction At the end of the 19th century, German botanist Julius von Sachs proposed a theory concerning the existence of chemical signals, which would play an important role in communication among plant organs. Since that time, a lot of information has been collected on substances, which could influence and regulate growth and development in plants. These substances are generally called plant growth regulators and can be separated into two groups: plant hormones (phytohormones) and other substances with growth regulatory activity. Phytohormones are naturally occuring substances which belong into nine major groups: auxins, cytokinins, abscisic acid, ethylene, gibberellins, brassinosteroids, jasmonic acid, salicylic acid and strigolactones. Recently, a group of peptide hormones has emerged. Other molecules with regulatory activity, recognized as polyamines or phenolics, are not classified as phytohormones because they are effective in relatively high concentrations. Due to the sessile character of plants and their cell totipotency, mechanisms of plant hormone functions differ from those in animals and humans. Phytohormones are not produced in special glands, but in a wide range of tissues. In contrast to animal hormones, each phytohormone influences several physiological processes and on the other hand, one process is influenced (positively or negatively) by more than one phytohormone or growth regulator. Plant growth and development is thus controlled by a very subtile balance among levels of individual phytohormones. 1. 1. Discovery of cytokinins Cytokinins have been defined as a class of plant hormones crucial to cell division (“cytokinesis”). Discovery of the first cytokinin, kinetin, by Skoog, Miller and associates in 1955 in herring sperm DNA (Miller et al. 1955) triggered an avalanche of experiments with the aim to isolate and identify other compounds with CK activity. This effort led to discovery of diphenylurea as a synthetic compound with CK activity (Shantz et al. 1955), trans-zeatin, the most active naturally occuring CK in plants (Letham 1963), and CKs with aromatic side chain (Horgan et al. 1973). 1 1. 2. Structure of cytokinins Naturally occuring CKs are N6-substituated adenine derivatives with either isoprenoid or aromatic side chain (Strnad 1997, Mok and Mok 2001). The abundance of individual CK derivatives varies significantly in the dependence on plant species, tissue or developmental stage. Isoprenoid CKs include derivatives of N 6-(Δ2isopentenyl)-adenine (iP), trans-zeatin (tZ), cis-zeatin (cZ) and dihydrozeatin (DHZ) (Fig. 1). CK bases and, to a lesser extent, also ribosides are considered to be physiologically active CK forms, as they are recognised by CK receptors (Inoue et al. 2001, Yamada et al. 2001, Spichal et al. 2004). There are nevertheless big differences in biological activity of individual bases and ribosides due to the structure and conformation of the side chain, trans-zeatin being the most active naturally occurring CK. Fig. 1. The structure of bioactive cytokinins. 1. 3. Cytokinin metabolism Regulation of the level of bioactive CKs involves their metabolic conversions. CK modifications include changes of either purine ring or of the side chain. Those reactions may significantly affect physiological activity of the respective CK. Interconversion of CK 2 phosphates to ribosides and CK bases is probably a major feature of CK metabolism (Chen 1981). While CK ribosides (especially trans-zeatin riboside) have been considered as CK transport forms, CK phosphates were identified as biosynthetic intermediates (Kakimoto 2001a, Takei et al. 2001). However, they may be also the product of metabolic conversion of CK bases and ribosides by adenosinekinase (EC 2. 7. 1. 20), as was shown in Physcomitrella patens (von Swartzenberg et al. 1998) or tobacco cells (Kwade et al. 2005), or adenine phosphoribosyltransferase (EC 2. 4. 2. 7; Chen et al. 1982, Shukla and Sawhney 1997). Due to the structure similarity with other, highly abundant, purine derivatives, CKs may share some metabolic steps with the purine metabolic pathway, although amount and type of metabolites can be influenced by incubation conditions and the type of cells used in experiment (Chen et al. 1985, Letham 1994). These enzymes, participating in CK interconversions have higher affinity for adenine, adenosine and AMP than for corresponding CK. 5´-nucleotidase (EC 3. 1. 3. 5), isolated from wheat germ and tomato (Chen and Kristopeit 1981a, Burch and Stuchbury 1987) can convert cytokinin nucleotide to its corresponding nucleoside, which can be further metabolized to CK free base by for example adenosine nucleosidase (EC 3. 2. 2. 7), which is responsible for deribosylation of CK nucleoside (i6Ado) to CK free base (i6Ade) (Chen and Kristopeit 1981b). Similarly, there exist enzymes catalyzing conversion of free bases to corresponding nucleosides or nucleotides. For example, adenosine phosphorylase (EC 2. 4. 2. 1), an enzyme requiring ribose-1-phosphate for ribosylation of CK base strongly favoured CK nucleoside formation (Chen and Petschow 1978b). Nucleotides can be formed either directly from CK bases due to the enzyme adenine phosphoribosyltranferase (EC 2. 4. 2. 8) in presence of 5´-phosphoribose-1´-pyrophosphate (Chen et al. 1982) or from corresponding nucleoside by adenosine kinase (EC 2. 7. 1. 20). Modifications of the purine ring of CK include N-glycosylation at the 3, 7, or 9 position (Fox et al. 1971, 1973, McGaw et al. 1984) and conjugation of alanine at the 9 position of zeatin, forming lupinic acid. Although N-glucosylation in 3 position has been reported in some plant species (Letham 1978), specific N-glucosyltransferase has not been identified yet. N-glucosylation in 7 and 9 position leads to metabolically stable compounds with very low biological activity, which are resistant to glucosidases (Brzobohatý et al. 1993), although recent data show that N-glucoside in position 9 could be cleaved by some isoforms of CKX (Galuszka et al. 2007, Kowalska et al., 2010). For these reasons N-glucosides are considered 3 to be involved in deactivation and detoxification (Letham and Palni 1983), especially in case of sudden excess of bioactive CKs (e.g. by exogenous application). Two Nglucosyltransferases from A. thaliana were identified in 2004 by Hou et al.. Both of them recognized all classical CKs and form N-glucoside in 7 and 9 position in in vitro assay. Fig. 2. Enzymes (E) and genes (G) involved in the modifications of N6-isoprenoid side chain. According to Mok and Mok (2001). 4 Modifications of the isoprenoid side chain of CKs have been also very often observed and were detected in almost all plant tissues (Fig. 2). The most common modification is glycosylation of the hydroxyl group of the side chain of t-Z, DHZ and cis-Z, resulting in CK O-glucosides or O-xylosides. O-glycosylation is a reversible process, the deglycosylation, forming the active aglycone, is catalyzed by β-glucosidase (EC 3. 2. 1. 21; Brzobohatý et al. 1993). Thus O-glucosylation may function as an important mechanism in the regulation of levels of active CKs. The gene for β-glucosidase (Zm-p60.1) was isolated from maize coleoptiles (Campos et al. 1992; Esen 1992). Recently, it was shown that overexpression of Zm-p60.1 can disturb trans-zeatin metabolism (Kiran et al. 2006). Trans-zeatin O-glucosyltransferase (EC 2. 4. 1. 203) was isolated from Phaseolus lunatus (Dixon et al. 1989) and trans-zeatin O-xylosyltransferase (EC 2. 4. 1. 204) was found in P. vulgaris (Turner et al. 1987). The former enzyme utilizes as donor substrates UDPG or UDPX, but the affinity to UDPG and UDPX differs and that to UDPG is much higher. The latter enzyme accepts only UDPX. The only CK substrate which is recognized by both enzymes is DHZ and it forms dihydrozeatin O-xyloside in presence of UDPX (Dixon et al. 1989, Martin et al. 2000, Mok et al. 2000). As DHZ O-glucoside was found in P. vulgaris (Wang et al. 1977), but reported trans-zeatin O-glucosyltransferase from beans does not catalyze the formation of this compound in the presence of UDPG, it seems that other O-glucosyltransferases converting DHZ to DHZ O-glucoside exist. In 2001, an O-glucosyltransferase (EC 2. 4. 1. 215) with 41% identity to the Phaseolus transzeatin O-glucosyltransferase, but highly specific to cis-zeatin, was discovered in maize (Martin et al. 2001b). Substrates of this enzyme are cis-zeatin (but not trans-zeatin, DHZ or cis-zeatin riboside) and UDPG (not UDPX). The discovery of the enzyme with function parallel to those for trans-zeatin led to a new view on the importance and relevance of cis-CK derivatives in plants. Existence of cis-specific regulatory mechanisms in plants suggests that cis-zeatin and its derivatives may play a more important role in CK homeostasis than it was thought. Till now, cis-CK derivatives were isolated from a number of plant species, including rice (Takagi et al. 1989), potato (Suttle and Banowetz 2000), wheat and oats (Parker et al. 1989), lupins (Emery et al. 2000), tobacco (Dobrev et al. 2002). Among other side chain modifications of CKs pertain reduction of trans-zeatin to form DHZ and cis-trans isomeration. Reduction of the trans-zeatin side chain is catalyzed by a zeatin reductase (EC 1.3.1.69) isolated from immature seeds of P. vulgaris (Martin et al. 1989) and pea (Gaudinová et al. 2005) forming dihydrozeatin. The enzyme shows high specificity for 5 trans-zeatin and does not recognize other CK derivatives (cis-zeatin, iP) as substrates and requires NADPH as a cofactor. As cytokinin oxidase/dehydrogenase does not use DHZ as a substrate, reduction of trans-zeatin to DHZ may preserve a cytokinin activity (Gaudinová et al. 2005). A cis-trans zeatin isomerase was partially purified from Phaseolus (Bassil et al. 1993). As the enzyme favours conversion from cis- to trans- isomer, isomeration may be a way to increase amount of highly active trans-zeatin in plant tissues. Also another enzyme, hydroxylase, catalyzing a conversion of iP to zeatin and iPR to transzeatin riboside was found in a microsomal fraction of cauliflower (Chen and Leisner 1984). 1. 4. Degradation of cytokinins The last but definitely not least side chain modification is CK degradation mediated by cytokinin oxidase/dehydrogenase (CKX), which irreversibly inactivates free CK bases and nucleosides with an unsaturated N6-side chain in a single enzymatic step (Fig. 3). Changes in CKX activity alter the concentrations of bioactive cytokinins in tissues, therefore CKX enzymes play an important role in controlling CK levels and thus influence cytokinin dependent processes. The first report on degradation of CKs by an enzyme system came in 1968 by Chen et al. (1968). They observed degradation of iPR to adenosine by an enzyme preparation from tobacco callus culture. A few years later, conversion of labelled i6Ade to adenine in crude tobacco cell culture extracts was reported (Pačes et al. 1971). In 1974, Whitty and Hall described similar activity in Zea mays kernels and named the enzyme cytokinin oxidase. Since then, CKX activity was observed in many other plant species like beans, poplar, wheat (Jones and Schreiber 1997). The various CKX could differ in their biochemical properties like substrate affinity, etc. CKX from majority of plants are glycoproteins (Armstrong 1994), however, non-glycosylated CKX or CKX with low degree of glycosylation were observed in several plants including tobacco and Phaseolus (Motyka et al. 1996, 2003, Kamínek and Armstrong 1990). CKX glycosylation affects enzyme localization, the apoplastic forms being glucosylated. The occurrence of several CKX isoenzyme (e.g. 7 members in Arabidopsis), which differ in their localization as well as in their regulation, indicate the involvement of CKX in regulation of different CK controlled processes (Schmülling et al. 2003). For a long time, CKX was supposed to require copper in its active center and molecular oxygen as an electron acceptor for its activity. For this reason the enzyme was named 6 cytokinin oxidase and classified as copper amine oxidase (EC 1. 4. 3. 6) (Whitty and Hall 1974, Hare and Van Staden 1994a). After cloning of maize gene, the view was corrected, because a flavin-binding site was observed (Morris et al. 1999) and a few years later atomic absorption analysis did not prove the presence of copper (Bilyeu et al. 2001). Also, the opinion that molecular oxygen is the electron acceptor was changed after finding that CKX can use a variety of artificial electron acceptors under anaerobic conditions (Galuszka et al. 2001, Frébort et al. 2002). Regarding these evidences that CKX are flavine enzymes not using oxygen as electron acceptors, enzyme was reclassified as cytokinin dehydrogenase (EC 1. 5. 99. 12). Degradation of aromatic CKs still remains to be solved, as CKX exhibit only very low reaction rate towards aromatic CKs, e.g. benzyladenine (Frébortová et al. 2004). Fig. 3. CKX function scheme 1. 5. Biosynthesis of cytokinins CK homestasis is regulated not only by CK interconversion, transport and degradation, but also by the control of the rate of their biosynthesis. The occurence of modified bases, including the isopentenylated ones, in plant tRNAs led to the hypothesis that the breakdown of tRNA (which results in the release of iP or zeatin derivatives) could be a potential mechanism for biosynthesis of CKs as free hormones (Hall 1970, Murai 1994, Morris 1995). CK moieties in plant tRNA include iPR, cis-zeatin riboside, trans-zeatin riboside, 2methylthio-iPR, 2-methylthio-zeatin riboside (Horgan 1984). It is generally thought that purine moieties in tRNAs are first isopentenylated and then can be further modified by side chain hydroxylation and/or ring methylation to give CK derivatives (Cherayil and Lipsett 1977). An enzyme that catalyzes the reaction between Δ2-isopentenyl diphosphate and unmodified tRNA is Δ2-isopentenyl diphosphate:tRNA- Δ2-isopentenyltransferase (EC 2. 5. 1. 7 8) and has been partially purified from various sources (Hall 1971). Recently, the corresponding gene was identified in Arabidopsis thaliana (Golovko et al. 2002). Also the theory that tRNA is a source especially of cis-type CKs was supported by finding that isoprenoide moiety of cis-zeatin is a product of mevalonate pathway localized in cytoplasm as well as tRNA-IPT2 (Kasahara et al. 2004). The role of tRNA-IPT in synthesis of cis-zeatin was demonstrated recently (Miyawaki et al. 2006). However, the slow turnover rate of tRNAs does not seem sufficient to be a sole source of CKs, especially when the necessity of a precise control of the level of bioactive CKs is taken into consideration. The first demonstration of an enzyme catalyzing transfer of the isopentenyl moiety from dimethylallyl diphosphate (DMAPP) to AMP was reported in 1978 in Dictyostelium discoideum (Taya et al. 1978). The enzyme was named DMAPP:AMP isopentenyltransferase. ATP, ADP or cAMP did not function as isopentenyl acceptors. The reaction product, isopentenyladenosine-5´-monophosphate (iPMP) was converted to iP. The first characterization of ipt (tmr) gene came from Agrobacterium tumefaciens, a crown gallforming bacterium (Akiyoshi et al. 1984, Barry et al. 1984). The tmr is encoded on the TDNA region of the Ti-plasmid and during the infection is introduced into the host plant chromosome. The tmr (ipt) gene was cloned and expressed in E. coli and its extracts were shown to catalyze production of iPMP from DMAPP and AMP. It has been assumed that similar biosynthetic pathway functions also in higher plants. Many attempts have been performed to isolate and characterize plant ipt genes, but due to their low expression levels and enzyme instability, only little biochemical information was reported. In 1979, Chen and Melitz partially purified an enzyme from cytokinin-autotrophic cultured cells of tobacco and a few years later, Blackwell and Horgan (1994) did the same from kernels of Zea mays. Recently, ipt genes have been identified in Arabidopsis thaliana (Kakimoto 2001a, Takei et al. 2001a), petunia (Zubko et al. 2002), hop (Sakano et al. 2004), soybean (Ye et al. 2006), Chinese cabbage (Ando et al. 2005), rice (Sakamoto et al. 2006), pingyitiancha (Malus hupehensis, Peng et al. 2008), maize (Brugiere et al. 2008) and finally, a prokaryote-type tRNA-IPT gene from the moss Physcomitrella patens was characterised (Yevdakova and von Schwartzenberg 2007). The completion of Arabidopsis genomic sequence showed a total of nine ipt-homologues described as AtIPT1 to AtIPT9. A phylogenetic analysis revealed that AtIPT2 (Golovko et al. 2002) and AtIPT9 (Kakimoto 2001a) encode a putative tRNAisopentenyltransferases, while the other seven AtIPTs participate in de novo CK synthesis (Kakimoto 2001a, Takei et al. 2001a). It had been assumed, according to the studies with prokaryotes, that the first step in CK biosynthesis in plants could be isopentenylation of AMP. 8 But biochemical studies showed that plant IPTs use ADP or ATP rather than AMP as prenyl acceptors, resulting in production of iP riboside 5´-diphosphate (iPRDP) or iP riboside 5´triphosphate (iPRTP) (Kakimoto 2001a, Sakano et al. 2004) (Fig. 4). Plants have two possible biosynthetic pathways for the prenyl group of CKs – the methylerythritol phosphate (MEP) pathway in plastids (Lichtenhaler 1999) and the mevalonate (MVA) pathway in the cytosol (Rohmer 2003). Both pathways produce the isoprenoid precursors isopentenyl diphosphate and DMAPP. In spite of the fact that MEP and MVA pathways exist in different subcellular compartments, exchange of precursors between these two pathways was found (Rohmer 1999, Kasahara et al. 2002). It has been observed that isoprene units from both pathways were incorporated into a single downstream isoprenoid (Kasahara et al. 2002). Using a 13C labeled precursor specific to the MEP and MVA pathways, it was demonstrated that both pathways are able to supply DMAPP to CKs in Arabidopsis (Kasahara et al. 2004). They proved the dominant role of the MEP pathway in the biosynthesis of iP- and tZ- type CK. Kasahara et al. (2004) postulated that the majority of the prenyl group of cZ-type CK comes from MVA pathway. They also used fusion of AtIPTs with green fluorescent protein and showed that a large number of the AtIPTs (AtIPT1, 3, 5, 8) are located in plastids (suggesting plastids to be a major subcellular compartment for iP- and tZ-type CK biosynthesis in higher plants), AtIPT4 and 7 are localized in the cytosol and mitochondria. Recently, using in vivo isotope labeling, an iPRMP-independent pathway for the biosynthesis of tZ-type CK was proposed, which involves binding of hydroxylated derivate of DMAPP directly to AMP (Åstot et al. 2000). In order to obtain a physiological activity, CK nucleotides must be converted to bases. It was assumed that this process has two steps operated by nucleosidases and nucleotidases participating in general metabolism of purine (Chen and Kristopeid 1981a,b). Recently, a new metabolic pathway releasing a free base directly from nucleotide was identified. The reaction is catalyzed by cytokinin nucleoside 5´-monophosphate phosphorylase (called LOG - “lonely guy”), which utilizes only mononucleotides (Kurakawa et al. 2007, Kuroha et al. 2009). In spite of the fact that aromatic CK exhibit strong physiological activity, their biosynthesis and degradation pathways remain still unclear. Regulation of CK biosynthesis can be studied at different levels. Using constructs of AtIPT promoter:reporter genes, tissue- and organ-specific patterns of CK biosynthesis were revealed (Miyawaki et al. 2004, Takei et al. 2004). CK biosynthesis is significantly affected by other phytohormones, namely auxin. In Arabidopsis, accumulation of AtIPT5 and 7 is promoted by 9 auxin in roots, while high CK levels affect negatively expression of AtIPT1, 3, 5 and 7 (Miyawaki et al. 2004). In shoots, auxin down-regulate IPT expression (Nordström et al. 2004, Tanaka et al. 2006). Genes for CKX in maize are upregulated by CK and abscisic acid (ABA) (Brugiere et al. 2003). This kind of hormone cross-talk may play a role in regulation of plant development and in the plant responses to environmental conditions. CKs are also pivotal signalling substances delivering information about nitrogen availibility from root to shoot via the xylem vessels (Takei et al. 2001, 2002). Recently, AtIPT3 and 5 were revealed to be regulated differently, depending on the nitrogen sources available. Whereas AtIPT3 specifically responses to NO3- under nitrogen-limited conditions, AtIPT5 reacts to both NO3- and NH4+ under long-term treatment (Miyawaki et al. 2004, Takei et al. 2004). Fig. 4. Scheme of cytokinin biosynthesis (according to Kamada-Nobusada and Sakakibara, 2009) 10 1. 6. Cytokinins and biological functions 1. 6. 1. Cytokinins and cell cycle regulation Regulation of cell division and growth involves tight control of cell cycle progression. The eukaryotic cell cycle consists of four phases, which lead to the reproduction of genetic material and its transport to the next cell generation. CKs were discovered due to their ability to promote cell division in presence of optimal level of auxin and it seems that CKs affect and control various activities through the cell cycle. There is an evidence that CKs play an important role especially in G2/M and G1/S transition (Hare and Van Staden 1997, Riou-Khamlichi et al. 1999, Dobrev et al. 2002). It is suggested that both auxin and CK regulate the cell cycle by controlling the activity of cyclin-dependent protein kinases (CDKs), enzymes regulating eukaryotic cell cycle. Expression of the gene encoding the major CDK (Cdc2) is regulated by auxin, but this CDK is inactive until an inhibitory phosphate group from the Cdc2 is removed. And this is the role for CKs, which activate phosphatase catalyzing removal of the phosphate group from Cdc2 (Zhang et al. 1996) and enable the cell transition from the G2 to M-phase. CKs also increase expression of the CYCD3 gene encoding a D-type cyclin that leads to elevation of cell proliferation and enables transition from G1 to S-phase (Riou-Khamlichi et al. 1999, Soni et al. 1995). 1. 6. 2. Cytokinins and leaf senescence Leaf senescence is a form of programmed cell death and the final phase of leaf development during which the nutrients contained in the macromolecules (e.g. nucleic acids, proteins) are metabolized to the basic components (e.g. amino acids) and transported from old, senescing leaves to the young productive leaves and developing seeds. This strategy is very advantageous, especially under adverse conditions where nutrients are limiting factors (Noodén 1988). Senescence is a highly regulated process that requires de novo synthesis of many nuclear genes, referred as senescence-associated genes (SAGs) (Gan and Amasino 1997, BuchananWollaston 1997, Weaver et al. 1998). Senescence can be triggered by different environmental 11 and endogenous stimuli such as drought, nutrient deficiency, temperature, pathogen attack, age, concentration of plant hormones etc. CKs have been showed to delay and, in some cases, reverse the leaf senescence process (Gan and Amasino 1996), as revealed by re-greening of tissues after CK elevation. To study leaf senescence and overcome problems with CK overproduction in transgenic plants, Gan and Amasino (1995) developed a gene construct using highly senescence specific promoter from SAG12. This gene is thought to be the most specific for natural senescence, showing no detectable expression in young leaves and not accumulating until the leaf is 20 % chlorotic (Weaver et al. 1998). When SAG12 promoter is used to drive IPT expression, transcription of IPT is triggered and production od functional enzyme is activated only when the leaf senescence is stimulated. In our work, we used transgenic plants in which the SAG12 promoter drove expression of ZOG1, gene for another key enzyme participating in CK metabolism, as described earlier. We used also plants with 35S promoter driven expression of ZOG1 to compare drought stress response of plants with senescence-induced CK elevation and those with constitutive stimulation of CK turn-over. 1. 6. 3. Cytokinins and chloroplast development Already 10 years after discovery of kinetin as a plant hormone (Miller et al. 1955), it was shown that kinetin has been able to promote chloroplast differentiation from proplastids (Stetler and Laetsch 1965). It has been proved that exogenously applied CK can induce partial development of chloroplasts from proplastids, amyloplasts and etioplasts (Chory et al. 1994, Kusnetsov et al. 1994). The importance of CKs for plastid development can be deduced from partial localization of the CK biosynthetic pathway to this compartment (Kasahara et al. 2004). CKs are also known to function in the protection of photosynthetic machinery (Chernyadev 2009). 1. 6. 4. Cytokinins and movement of nutrients In general, CKs have been shown to play a major role in the regulation of many processes associated with the supply of nutrients to growing tissues and with regulation of the sink strength. 12 Several reports show the effect of CKs on the expression of genes for nitrate reductase (Lu et al. 1990, Hänsch et al. 2001), a hexose transporter (Ehneß and Roitsch 1997) and an extracellular invertase, catalyzing the hydrolytic cleavage of the transport sugar sucrose released into the apoplast, which seems to be an essential component of the molecular mechanism of the delay of senescence by CKs (Ehneß and Roitsch 1997, Lara et al. 2004). It is suggested that de novo synthesis and recycling of storage forms (O-glucosides) might contribute to CK accumulation in the roots, followed by translocation to the leaves via the xylem during recovery from nitrogen deficiency in maize (Takei et al. 2001, 2002). Nitrogendependent accumulation of CKs was also observed in roots of Arabidopsis thaliana (Takei et al. 2002) and barley (Samuelson and Larsson 1993). The main CKs accumulated in response to nitrate supply in roots are generally isopentenyladenine-type CKs, which are transformed into trans-zeatin riboside, when entering the xylem, and into free base trans-zeatin, which accumulates in the leaves (Takei et al. 2001). 1. 6. 5. Cytokinins and signalling As plants cannot escape from adverse environmental conditions, their lives depend on their ability to react quickly and efficiently. Therefore they have developed different signal perception and transduction systems. In higher plants, the major role in CK perception and signalling is played by the twocomponent system. Until recently, signal transduction via the two-component system was thought to be only a prokaryotic concern. But now it was proved that this mechanism participates in higher plants, too (Lohrmann and Harter 2002, Hass et al. 2004a). The twocomponent system serves as a sensing/responding mechanism. It is composed of hybrid histidine kinases, histidine-containing phosphotransfer domain proteins and response regulators that are biochemically linked by His-to-Asp phosphorelay (Fig. 5). As a reaction to endogenous or exogenous stimuli, autophosphorylation of the histidine kinase at a conserved histidine residue is induced and signalling is initiated, followed by a multistep system with additional phosphorylation steps. The hybrid histidine kinase carries an additional receiver domain, usually at the COOH-terminal end. Then, the phosphoryl group is transferred to a conserved Asp residue of its own receiver domain instead of transferring the phosphoryl group directly to the response regulator. 13 Additional proteins with a histidine-containing phosphotransmitter (HPt) domain then percieve the phosphoryl group from the receiver domain of the hybrid histidine kinase and transfer it to the receiver domain of the response regulator, what induces a conformational change in output domain that alters its biological activity (West and Stock 2001, Hass et al. 2004a). The plant species intensively studied in most detail is Arabidopsis thaliana. Sequence analysis of Arabidopsis genome has revelead 8 histidine kinases (AHKs), 5 HPt proteins (AHPs), 23 response regulators (ARRs) and 9 pseudo response regulators (APRRs) (Urao et al. 2000a, Hwang et al. 2002). Arabidopsis AHKs can be divided into several subfamilies, where one subfamily functions as CK receptors. It is composed of AHK2, AHK3 and AHK4/CRE1/WOL (histidine kinase 4/cytokinin response/wooden leg) (Kakimoto 2003). The Arabidopsis HPt proteins (AHPs) are expressed ubiquitously and their transcription in not affected by CK treatment (Tanaka et al. 2004). They are suggested to play a role as intermediates in multistep phosphorelay to transfer the signal from AHKs to corresponding response regulators. Fig. 5. Cytokinin signalling. According to Werner and Schmülling (2009). 14 The Arabidopsis response regulators are divided into three groups: 1) Type-A response regulators (type-A ARRs) are relatively small and contain a receiver domain and a short carboxyl terminus and their transcription is rapidly elevated in response to exogenous cytokinin. They are considered to be primary response genes (D´Agostino et al. 2000, Kiba et al. 2002, Rashotte et al. 2003). They are negative regulators of CK signalling. 2) Type-B response regulators (type-B ARRs) are comprised of a receiver domain, carboxy-terminal output domain that has a DNA-binding GARP domain and a transcriptional activation domain. This type of ARRs transfers the CK signal to CK early response genes (Imamura et al. 1999, Kiba et al. 1999). 3) Pseudo response regulators (APRRs) contain a receiver-like domain, but the invariant amino acids including the phosphorylated Asp are substituted by other amino acids (Hwang et al. 2002). Some observations suggest that APRRs may function as nuclear transcription factors (Strayer et al. 2000). Other evidences suggested the crucial role of APRRs as intrinsic elements of Arabidopsis circadian clock (Hayama and Coupland 2003, Eriksson et al. 2003). 1. 7. Polyamines – plant growth regulators Polyamines (PA) are small organic, ubiquitous polycations with several aminogroups in their molecules, found in all living organisms. They play an important role in regulation of growth and development in prokaryotes and eukaryotes (Tiburcio et al. 1997). In plants, they have been implicated in a wide range of biological processes, including stimulation of DNA replication, transcription and translation, response to senescence, environmental stress and infection by viruses and fungi. These biological activities are fastened on their cationic nature (Galston et al. 1997, Bouchereau et al. 1999, Thomas and Thomas 2001, Theiss et al. 2002, Capell et al. 2004). The major PAs found in plants are diamine putrescine (Put), triamine spermidine (Spd) and tetramine spermine (Spm). They occur in the free form or as conjugates bound to phenolic acids or to macromolecules such as proteins and nucleic acids. PAs were reported to participate in stabilization of membranes and protection against free radicals. They influence protein and nucleic acid synthesis and enzyme activities (RNAse, protease, etc.). They have intensive cross-talk with phytohormones such as CKs, auxin, ethylene. 15 1. 8. Polyamine metabolism The PAs are synthetized from arginine or ornithine by arginine decarboxylase (ADC) and ornithine decarboxylase (ODC), respectively. The intermediate agmatine is then converted to Put, which can be further transformed to Spd and Spm by transfer of aminopropyl groups from decarboxylated S-adenosylmethionine (dSAM) catalysed by specific Spd and Spm synthases (Slocum 1991a, Martin-Tanguy 2001). The catabolism of PAs is performed by oxidases. It has been found that each PA is catabolized by a specific oxidase (Smith and Marshall 1988). In order to understand their biological roles, many studies have been done to localize PAs and their biosynthetic enzymes in plants. They have been found in cell wall vacuoles, mitochondria and chloroplasts (Torrigiani et al. 1986, Slocum 1991b, Tiburcio et al. 1997). PAs are involved in many plant developmental processes such as cell division, reproductive organ development, root growth, floral initiation and development, fruit development and ripening, leaf senescence and abiotic stress response. Changes in free and conjugated PAs occur during these developmental processes. In general, cells undergoing division contain high levels of free PAs synthetized via ODC, whereas cells undergoing expansion and elongation contain low levels of free PAs synthetized via ADC (reviewed by Galston and Kaur-Sawhney 1995). 16 2. Aims and scope Cytokinins (CKs) are one of the most important plant hormones, which exhibit, in cooperation with other phytohormones, multiple physiological functions. This work has been focused on different aspects of cytokinin research, namely on: 1) monitoring of diurnal changes in CK levels (in correlation with other phytohormones - auxin and abscisic acid) and polyamine content in tobacco plants, in order to contribute to better understanding of the complex network of light and hormone signal transduction pathways. 2) elucidation of the extent of chloroplast autonomy in the maintenance of CK homeostasis determined by comparison of the effect of modulation of CK metabolism in cytoplasm on CK content at chloroplast and whole leaf. Manipulation of CK metabolism was achieved by over-expression of the genes for CK biosynthesis, degradation and CK glucosylation/deglucosylation, respectively. This study included monitoring of CK effect on changes in chloroplast ultrastructure in tobacco plants. 3) evaluation of the role of CKs (together with other hormones - auxin and ABA) in the drought response of tobacco plants, comparing the wild-type plants with those exhibiting increased CK turn-over before drought stress initiation or during the stress progression (achieved by the expression of trans-zeatin O-glucosyltransferase gene, ZOG1, under 35S or SAG12 promoters, respectively). 17 3. Plant material and methods Plant material 1) Plants of Nicotiana tabacum L. cv. W 38 (prof. Machteld and David Mok, Oregon State University) 2) Tobacco in vitro lines expressing an IPT homologue Sho (Shooting) gene from P. hybrida under the control of a constitutive cauliflower mosaic virus (CaMV) 35S promoter or construct with four 35S enhancer elements (Prof. Břetislav Brzobohatý, Mendel University in Brno) 3) Eleven-week-old tobacco plants (35S:AtCKX3) overexpressing a gene for CK oxidase/dehydrogenase from A. thaliana under a constitutive CaMV 35S promoter (Prof. Thomas Schmülling, Free University, Berlin) 4) Nine-week-old Nicotiana tabacum L. cv. Samsun NN plants (Prof. Thomas Schmülling, Free University, Berlin). 5) Eight-to nine-week-old tobacco plants (35S:P60) overexpressing a maize βglucosidase Zm-p60.1 naturally targeted to plastids under a CaMV 35S promoter (Prof. Břetislav Brzobohatý, Mendel University in Brno). 6) Nicotiana tabacum L. cv. Petit Havana SR1 plants (Prof. Břetislav Brzobohatý, Mendel University in Brno). 7) Eight-to nine-week-old tobacco plants (35S:ZOG1) harbouring a trans-zeatin O- glucosyltransferase gene from P. lunatus under a constitutive CaMV 35S promoter (prof. Machteld and David Mok, Oregon State University) Methods 3. 1. Growing conditions and collecting samples 3. 1. 1. For isolation of CKs and study of diurnal rhythms see supplement 1, p. 49 3. 1. 2. For isolation of PAs and study of diurnal rhythms see suplement 2, p. 56 3. 1. 3. For isolation of CKs and study of altered CK metabolism see supplement 3, p. 65 18 3. 1. 4. For isolation of CKs and study of water deficit see supplement 4, p. 78 3. 1. 5. Incubation of samples with tritiated trans-zeatin see supplement 1, p. 49 3. 1. 6. Chloroplast isolation see supplement 3, p. 66 3. 2. Plant hormone extraction and purification 3. 2. 1. Extraction and purification of CKs see supplement 1, p. 49 ; supplement 3, p. 66 and supplement 4, p. 79 3. 2. 2. Extraction and purification of PAs see supplement 2, p. 56 3. 3. Chromatographic methods 3. 3. 1. HPLC of radiolabelled trans-zeatin and its metabolic products see supplement 1, p. 49 3. 3. 2. HPLC of auxin and abscisic acid see supplement 1, p. 49 ; supplement 3, p. 66 and supplement 4, p. 80 3. 3. 3. HPLC/MS see supplement 1, p. 49 ; supplement 3, p. 66 and supplement 4, p. 80 3. 3. 4. PA analysis see supplement 2, p. 56 3. 3. 5. Chlorophyl determination see supplement 3, p. 66 3. 3. 6. Chloroplast integrity see supplement 3, p. 66 3. 4. Determination of enzymes activities 3. 4. 1. Determination of CKX activity see supplement 1, p. 49 and supplement 4, p. 80 3. 4. 2. Determination of β−glucosidase activity see supplement 1, p. 50 3. 4. 3. Determination of ODC, ADC activity see supplement 2, p. 57 19 3. 4. 4. Determination of DAO activity see supplement 2, p. 57 3. 5. Microscopic method 3. 5. 1. Transmission electron microscopy see supplement 3, p. 66 3. 6. Molecular methods 3. 6. 1. RNA extraction and reverse transcription see supplement 4, p. 79 3. 6. 2. qRT-PCR see supplement 4, p. 79 3. 7. Determination of relative water content see supplement 4, p. 79 20 4. Results and Discussion 4. 1. Diurnal rhythms of cytokinins, auxin and abscisic acid Plants growth and development, as well as their interactions with the environment are mediated by plant hormones. One of the most important environmental factors is light. In order to cope with regular rhythm of light/dark periods, plants have evolved circadian regulatory system, endogenous mechanism that allows organisms to correlate their physiological activities with predictable day/night cycles. Circadian rhythms are the subset of biological rhythms with period defined as the time to complete one cycle of ~ 24 h (Dunlap et al. 2004). These rhythms have been found in a wide range of organisms from plants to mammals, indicating its importance in life processes, including germination, developmental processes as hypocotyl elongation and flowering, stomatal opening and gas exchange, photosynthesis and control of the contents of various metabolites. Also many enzymes like glutamate dehydrogenase, nitrate reductase, sucrose-phosphate synthase are liable to circadian and/or diurnal rhythms. Several studies on circadian rhythms were done using the model plant Arabidopsis thaliana (Alabadi et al. 2001, Mizoguchi et al. 2005, Nakamichi et al. 2009), followed by soybean plants (Rufty et al. 1983), tomato (Tucker et al. 2004, Mizoguchi et al. 2007), maize (Bowsher et al. 1991), tobacco (Masclaux-Daubresse et al. 2002) and barley (Kurapov et al. 2000). In presented work, we focused on elucidation of diurnal rhythms of regulations of plant hormone levels, including mechanisms governing their metabolism. Analysis of CKs, auxin (IAA), ABA (publication 1) and polyamines (publication 2) was performed using leaves of tobacco plants (Nicotiana tabacum L. cv. Wisconsin 38) grown under 16/8h photoperiod. The main peak of physiologically active CKs, including trans-zeatin, isopentenyladenine, dihydrozeatin and their ribosides,was detected 1 h after the middle of the light period (Fig. 1, p. 50). This finding is in agreement with data of Kurapov et al. (2000), who measured diurnal variation of phytohormones in barley leaves. The midday peak of CKs was accompanied by the maximum level of IAA (Fig. 6, p. 51). Since IAA and CKs are necessary for the progression of cell division, these results indicate that cell division may be initiated at this time, as demonstrated Kurapov et al. (2000), where the maximal plant growth coincided with time right after the midday peak of CKs and IAA, whereas the minimal growth was measured right before dawn. The behaviour of cis-zeatin and its riboside was similar to the time course 21 of active CKs, the minor peak coincided with the maximum of active CKs, whereas the major peak detected 3 h after light/dark transition had 2 h delay in relation to the bioactive CKs. We also measured concentration of deactivation metabolites of CKs (N-glucosides, Fig. 2, p. 50) with two peaks coinciding with active CKs. This result seems to indicate precise control of the content of bioactive CKs. Another way of regulation of CK levels is their conversion to storage forms (O-glucosides, Fig. 3, p. 51). They reached the maximum 1 h after the dark/light transition, then the concentration started declining to exhibit another minor peak 1 h before the middle of the light period. These data correspond well with time course of bioactive CKs - while levels of O-glucosides were decreasing probably due to the activity of β-glucosidase, concentration of bioactive CKs started rising to reach the maximum after the vehement decline of Oglucosides, shortly after the middle of the light period. To evaluate the role of reversible conjugation in regulation of CK levels, β-glucosidase activity was measured (Fig. 3, p. 51). Right after the light was switched on, enzyme activity sharply increased, which might coincide enhanced transport of CKs (especially of CK O-glucosides) from the roots by transpiration stream. The occurrence of β-glucosidase peak right before and after the maximum of CKs might indicate that hydrolysis of CK O-glucosides could be involved in the fine-tuning of bioactive CK concentration at midday, even if this deglucosylation could not be the only source of the midday peak of bioactive CKs. Elevation of CK phosphates, the immediate CK biosynthesis precursors indicates that CK biosynthesis is likely to play the crucial role in sharp midday increase of bioactive CKs. As determination of total β-glucosidase activity was done, hydrolysis of other glucoconjugates (e.g., ABA-glucosylester) should be also taken into consideration. Measuring ABA levels (Fig. 6, p. 51), we found one minor peak following the middle day maximum of IAA and bioactive CKs, where ABA probably serves to counterbalance the levels of bioactive CKs to regulate the stomatal aperture by CK/ABA ratio (Fusseder et al. 1992). The sharp maximum detected at the beginning of the dark phase was consistent with the data of Tallman (2004), who reported a stimulation of ABA biosynthesis at dark, to keep the stomata closed during the night to ensure maximal tissue rehydration. The decrease in ABA concentration found after the beginning of the light period is probably associated with the exhaustion of endogenous ABA and the activation of its catabolism (Tallman 2004, Kraepiel et al. 1994). The decline in ABA concentration in the guard cells and surrounding apoplast in the morning allows stomatal opening with possitive effects on photosynthesis. 22 As the roots are the main site of CK biosynthesis, it was suggested that high concentration of xylem CKs in the morning originated from the steady rate of CK biosynthesis in the roots and of their exudation into the xylem sap during the night, when the transpiration rate is low (Beck and Wagner, 1994). In the morning, the accumulated CKs are delivered throughout the plant as the flow of the xylem sap increases. Due to the enhanced volume of the transpiration stream during the day and the steady rate of CK biosynthesis in roots, CK concentration in the xylem sap gradually decreases. The importance of this process in translocation of rootsynthetised CKs was recently shown by Aloni et al. (2005). When the activity of the main CK degrading enzyme was followed, the main peak of CKX activity (Fig. 4, p. 51) occurred early in the light period. The timing of the peak might coincide with enhanced CK delivery in the morning by increased xylem flow. Other two minor CKX peaks (detected after midday and after light/dark transition) coincided well with the decrease of endogenous CKs. Another approach in the study of CK diurnal variations was the use of radiolabelled transzeatin to follow its metabolic conversion (Fig. 5, p. 51). We found a sharp maximum of adenine and adenosine, which are products of CK breakdown, right after the dark/light transition and one minor peak around 9 h of light. Both of them preceded the maxima of CKX activity. Also the peak of non-metabolized trans-zeatin around 6 h of light corresponds well with the period od the diminished trans-zeatin breakdown as well as with the minimum of the ribosylated and phosphoribosylated forms. 4. 2. Diurnal rhythms of polyamines Our next paper was focused on diurnal patterns of polyamines and their metabolic enzymes (publication 2). Polyamines play an imporatnt role in the regulation of different processes in plants due to their cationic nature, which enables interaction with anionic biomolecules, such as nucleic acids, proteins or phospholipids. They are also known to affect many processes regulated by CKs and auxins, which is the reason for the study of diurnal rhythms of CKs, auxins and PAs simultaneously. The results presented in the publication 2 show detailed changes in the contents both free and conjugated (soluble and insoluble) forms of Put, Spd, Spm and Cad in tobacco leaves grown under the 16/8 h light/dark period, as well as in the activities of main enzymes participating in PA metabolism as ADC, ODC and diamine oxidase (DAO). The ratio among PA forms was changing during the 24 h cycle (Fig. 2, p. 58). Whereas the most abundant PA in free and PCA-soluble form was Put and the least abundant Cad, in PCAinsoluble conjugates predominated Cad before Put, Spm, Spd (Fig. 3, p. 58). The endogenous 23 levels of free PAs can be regulated also by conjugation with hydroxycinnamic acid (MartinTanguy 1985, Creus et al. 1991, Bagni and Tassoni 2001, Biondi et al. 2001), which seems to be the common phenomena in Nicotiana (Martin-Tanguy 1985, Pfosser et al. 1990, Gemperlová et al. 2005), because about 75 % of the total Put and Spd pool detected in the midday peak was presented in PCA-soluble conjugates. Nevertheless, the biological function of conjugated PAs remains unclear, speculating the reversible conjugates may supply the cells with an additional amine reserve (Bonneau et al. 1994) and/or protect cells from harmful effects of intracellular accumulation of free PAs on the maintenance of cellular pH, ion homeostasis and on many other processes influenced by PAs. The time-course of the level of PCA-soluble conjugates of Put and Spd was quite similar (Fig. 2A, B, p. 58) showing the first maximum 1 h after the middle of the light period. Further transient peaks were detected 2 h before the light was switched off and at the beginning of the dark period. A minor peak was observed at the end of the dark phase. It is very interesting but not surprising that the observed midday maximum of PAs coincided well with the main peak of CKs and IAA, as was shown in our previous work (Fig. 6, p. 51). These results only confirm our suggestion that cell division occurs at this time period. In tobacco leaves grown at the 16/8 h light/dark period, the level of free and PCA-soluble PAs seems to be regulated especially by the activities of biosynthetic enzyme ODC and catabolic enzyme DAO. Light dramatically stimulated activity of ODC (Fig. 5, p. 59 ), whereas no increase in activity of the second biosynthetic enzyme ADC was observed at the beginning of the light phase. On the contrary, the enzyme activity declined during the first hours of light and started increasing at midday, with main peak at the light/dark transition. This peak was accompanied by the accumulation of free PAs and PCA-soluble conjugates and also by the increase in ABA levels (Nováková et al. 2005). Those results seems to be in accordance with findings of Koetje et al. (1993) and Galston et al. (1997) who suggested that the pathway via ADC is generally linked to stress responses and morphogenetic processes, whereas ODC is involved in the regulation of cell division in actively growing plant tissue. Activity of DAO, the catabolic enzyme, was induced by light in the similar manner as ODC activity. We measured a sharp maximum 1 h after the dark/light transition, which may be the reason why only a slight increase in levels of free Put was observed after the photoinduction of ODC activity at the beginning of light phase. The second peak measured shortly after the middle of the light period is in agreement with results of Andreadakis and Kotzabasis (1996) who observed maximum of enzyme activity in isolated etioplasts from Hordeum vulgare after light exposure for 9 h. 24 We must also remark that according to information from literature, catabolic degradation of Put by DAO is not only a way of regulation of cellular content of this diamine (Kumar et al. 1997, Bagni and Tassoni 2001), as the other reaction products, like pyrroline and hydrogen peroxide, play important roles in subsequent processes. Pyrroline can be further catabolised to γ-aminobutyric acid (GABA) and finally incorporated into the Krebs cycle (Tiburcio et al. 1997) and/or function in signal transduction pathways during stress response (Shelp et al. 1999). Hydrogen peroxide might trigger a programmed cell death (Walters 2003) or might be utilised in lignification both during the normal growth and stress response (Angelini et al. 1993). 4. 3. The study of chloroplast autonomy in regulation of CK levels To study the impact of manipulation of CK metabolism on the endogenous CK pool in leaves and isolated intact chloroplast, and on chloroplast ultrastructure in tobacco (Nicotiana tabacum L.) we used plants overproducing CKs (plants were expressing the Sho gene coding for a Petunia hybrida IPT), plants with decreased CK levels (overexpressing cytokinin oxidase/dehydrogenase AtCKX3 from Arabidopsis thaliana), and plants with altered CK glucoconjugation (overexpressing maize β-glucosidase Zm-p60.1 or trans-zeatin Oglucosyltransferase ZOG1 from Phaseolus lunatus). The total CK content in leaves after Sho induction was eight times higher than in the corresponding WT (SR1) (Fig. 2, p. 67) and the most elevated CK metabolites were phosphates and N-glucosides. The differences between Sho transformants and the WT was much lower in isolated chloroplasts than in leaves. Following Sho induction in leaves, major increase was observed in iP7G and iPMP, which is in line with previous results reported for petunia and tobacco constitutively overexpressing Sho (Zubko et al. 2002). The CK content was also increased in chloroplasts but to a relatively lower extent. Recent analysis of Pssu-ipt tobacco (Synková et al. 2006) showed that, similarly to Sho plants, CK levels increased in chloroplasts to much lower levels compared with leaf samples, indicating that chloroplasts are relatively autonomous and keep their CK homeostasis quite well in both Z- and iP-type CKoverproducing plants. The content of IAA in leaves of Sho-expressing plants did not differ significantly from that in WT, whereas chloroplasts contained a lower level of IAA than WT chloroplasts (Table 1, p. 68). A higher ABA level was found in chloroplasts and leaves (when expressed per FW) of Sho plants compared with WT (Table 1, p. 68). It was shown that constitutive overexpression of AtCKX3 significantly delayed development of transgenic plants (Werner et al. 2001, 2003). Retarded ontogenesis was the reason why the 25 hormone content and chloroplast structure of transgenic plants were analysed in older plants than corresponding WT (SNN) plants. The total CK content was markedly reduced in leaves (Fig. 3, p. 68), particularly levels of N-glucosides and to a lesser extent bases and ribosides. It is obvious that after a tremendous increase of the activity of CKX, the activity of the other CK deactivation pathway (N-glucosylation) was suppressed. No significant differences were found between the IAA levels (expressed on FW basis), in leaves of AtCKX3 and WT plants (Table 2, p. 69). A lower ABA content was observed in AtCKX3 leaves compared with WT leaves, which might be associated with prolonged life span and retarded senescence of transgenic plants (Table 2, p. 69). Maize β-glucosidase Zm-p60.1 was the first plant enzyme shown to release active CKs from CK metabolites O- and N3-glucosides in vivo (Brzobohatý et al. 1993). In tobacco plants overexpressing maize β-glucosidase Zm-p60.1 targeted to plastids, we observed an increase in CK bases and ribosides and CK N-glucosides in leaves (Fig. 4A,B, p. 69 ). These results are in accordance with recent data by Kiran et al. (2006). Accumulation of CK bases and ribosides was found also in chloroplasts (Fig. 4C, p. 69). The very dramatic decrease in CK O-glucosides levels shown here for chloroplasts isolated from Zm-p60.1 plants corresponded well with plastid/chloroplast location of Zm-p60.1 (Kristoffersen et al. 2000). Measurements of IAA levels revealed a significantly lower concentration of IAA both in leaves and chloroplasts of transgenic tobacco plants compared with the WT IAA level (Table 3, p. 70), which is in agreement with Kiran et al. (2006), who measured a fall in IAA gradient from the youngest leaves downward and a decreased IAA content in the apex and first internodes in Zm-p60.1 tobacco plants. In ABA levels, we observed lower levels in leaves and chloroplasts of transgenic plants than in WT, which is in contrast with previous work of Kiran et al. (2006), who found higher ABA accumulation especially in leaves of Zm-p60.1 tobacco plants. Tobacco transformed with ZOG1 gene from Phaseolus lunatus under the control of the CaMV 35S promoter was constructed by Martin et al. (2001a). When a ZOG1 gene was constitutively expressed in tobacco, CK O-glucosides accumulated in leaves (Fig. 5A, p. 70). No significant difference in other CK metabolites was observed. By contrast, in chloroplasts a very low levels of CK O-glucosides were found (Fig. 5C, p. 70). This indicates that CK Oglucosides are accumulated mainly out of the plastids, which is in agreement with the previous data that transgenic plants overexpressing Zm-p60.1 can accumulate high levels of tZOG when grown on a medium supplemented with t-Z despite high Zm-p60.1 β-glucosidase activity in chloroplasts (Kiran et al. 2006). As the main storage compartment for CK O- 26 glucosides was proposed vacuole (Fusseder and Ziegler 1988). In leaves of transgenic plants, a 2-fold higher IAA level was measured compared with WT (Table 4, p. 71), whereas the ABA content was not significantly affected. The IAA and ABA levels in chloroplasts were higher in ZOG1 plants than in WT. A significant changes were observed in the chloroplast ultrastructure of tobacco plants constitutively expressing the Sho gene, including an increased number of plastoglobuli and the appearance of crystalloids in plants grown in vitro (Fig. 6, p. 71). Except for Sho tobacco transformants, marked changes in the ultrastructure of chloroplasts were not found in the other CK transformants analysed. The most obvious feature observed was the different starch accumulation associated with a change in chloroplast shape (Fig. 7, p. 72) which is in accordance with Wilhelmová and Kutík (1995) and Stoynova et al. (1996). All these results indicate that chloroplasts are relatively independent organelles with respect to the regulation of CK metabolism and that different compartmentation of individual CK metabolites should be taken into account. 4. 4. The role of CK in responses to water deficit Water stress represents one of the main limiting factors of plant growth and development. Water availability has become one of the most discussed problems. Detailed understanding of mechanisms of plant response to water deficit is a necessary prerequisite for the elucidation of an optimal strategy for enhancement of plant drought stress tolerance. We studied the impact of water deficit progression on CKs, auxin and ABA, using tobacco plants over-expressing trans-zeatin O-glucosyltransferase gene under 35S or SAG12 promoters and the corresponding wild-type. The SAG12 promoter has been, since the pioneering work of Gan and Amasino (1995), successfully used for targeted manipulation of CK levels. The transformants were characterized by ZOG1 expression profile. The amount of ZOG1 mRNA was quantified in the individual leaves and roots of control (well-watered) and drought-treated plants by real-time RT-PCR. ZOG1 expression in constitutive transformants was relatively high and even within the plant. The SAG12 promoter was significantly less active. As a promoter of senescence associated gene, it exhibited strict dependence on the leaf position. The highest expression levels of ZOG1 were found in SAG::ZOG1 transformants in older (lower) leaves (Table 1, p. 81). For better orientation in results of endogenous CK analysis, we divided individual CKs into seven groups according to their structure and function. Namely bioactive CKs, CK 27 deactivation forms (N-glucosides), CK storage forms (O-glucosides), CK phosphates (CK biosynthesis intermediates), cis-zeatin derivatives, cis-N-glucosides and cis-O-glucosides. In plants not suffering from water deficiency, the concentration of bioactive CKs is usually quite high in the young leaves and declines as the leaves start senescing (e.g. Singh et al. 1992). In accordance with these results, we observed in young plants at the beginning of the experiment a similar content of bioactive CKs in all leaves. With stress induced senescence progression, a gradient of bioactive CKs in favour of upper leaves was observed (Fig. 5, p. 82). In senescing well-watered SAG12::ZOG1 plants, we found an elevation of CK content, especially in lower leaves, which is in accordance with Cowan et al. (2005), who reported an increase in trans-zeatin-type CKs in senescing well-watered SAG12::ipt plants. After stimulation of SAG12::ZOG1 expression, we observed, apart of the expected increase of the level of CK O-glucosides, also an elevation of bioactive CKs. Both control and stressed 35S::ZOG1 transformants had considerably higher total CK levels than WT (Fig. 4, p. 82), predominantly of O-glucosides, both of the trans- and cis-zeatin type. As the ZOG1 expressing plants have elevated pool of CK O-glucosides, we compared the β-glucosidase activity in WT and ZOG1 transformants to check the ability of these plants to utilise the elevated levels of CK O-glucosides. Mild elevation of β-glucosidase activity in ZOG1 plants (results not shown) seemed to prove their capacity to hydrolase the CK storage pool. In order to elucidate CK function during the drought stress, we compared the water deficit response of WT plants, those with a uniform elevation of CK content before stress initiation and those exhibiting CK elevation targeted to senescing tissues. The mild water stress caused in WT plants formation of a gradient of bioactive CKs in favour of the upper leaves (Fig. 5, p. 82). This observation reflects necessity for the preferential protection of upper leaves under adverse conditions. The importance of CKs in sink-source polarisation during mild stress was recently discussed by Cowan et al. (2005). SAG::ZOG1 exhibited similar CK response as WT, however, the content of bioactive CKs was in lower leaves less suppressed by the drought. During the prolonged drought, decrease in bioactive CK levels was in 35S::ZOG1 plants more profound than in WT. The elevated CK content before stress initiation might be the reason of the delay in the response of 35S::ZOG1 plants to unfavourable environmental conditions, which might have negative consequence in case of severe, prolonged drought. The ABA content of plants correlated well with the strength of applied water stress (Fig. 8, p. 85). When mild stress was applied, only a slight increase in ABA concentration occurred in 28 the upper and middle leaves of WT plants and in the middle leaves of SAG::ZOG1 plants. In constitutive ZOG1 transformants we observed no increase of ABA levels. Under mild stress, an elevated level of free IAA was found in WT lower leaves (Fig. 9, p. 85), whereas in roots of both transgenics we observed slight increase of IAA levels. In all genotypes, the CKX activity was much higher in roots than in leaves (Fig. 7, p. 84). Moderate drought stress had two main effects on the concentrations of bioactive CKs. Firstly, the progressive drought led to a decrease of the bioactive CK levels of (especially of transzeatin/riboside) in all leaves of all genotypes tested, which is in agreement with Goicoechea et al. (1995). Secondly, the steepness of the apex-to-base gradient in bioactive CKs increased substantially, especially in 35S::ZOG1 plants, in which only minor response to mild stress was observed. The relatively small difference in bioactive CKs levels between the middle and lower leaves of SAG::ZOG1 plants might be the consequence of enhanced SAG promoter activity in stressed lower leaves (Fig. 3, p. 81). Under moderate drought stress conditions, ABA levels were increasing substantially in all stressed plants (Fig. 8, p. 85), which is in agreement with previous reports (e.g. Mansfield and McAinsh 1995, Riccardi et al. 2004). As ABA is an important component of the plant stress defence system, its higher levels, observed in the upper leaves, might reflect their preferential protection. As far as IAA is considered, moderate drought stress resulted in all genotypes in a substantial increase in IAA levels, both in the lower leaves and roots (Fig. 9, p. 85). Severely stressed plants had all the leaves completely wilty. The bioactive CK content decreased in all genotypes, compared with that in moderate stress and CK gradient was suppressed in the WT and SAG::ZOG1 plants, while it was still visible in 35S::ZOG1 plants. The flattening of the CK gradient might indicate that the strength of the stress nearly reached the limits of the plant's ability to cope with the drought. Nevertheless, the decrease of CKX activity in the upper leaves and, especially, the large increase in CKX activity in the lower leaves seems to indicate a tendency for the maintenance of the gradient of bioactive CKs (Fig. 7, p. 84). In all genotypes tested, severe stress led to the further accumulation of total CKs in the roots. Severe stress also led to a further increase in ABA content (by more than one order magnitude in comparison with control conditions). The highest concentration occurred in the upper leaves (Fig. 8, p. 85). Large elevation of IAA content, observed in the stressed lower leaves of all genotypes and also in the severely stressed middle leaves of 35S::ZOG1 plants (Fig. 9, p. 85), is in the agreement with the results of Quirino et al. (1999), who found an increase in the free IAA 29 levels in senescing Arabidopsis leaves. The function of the increased IAA levels in senescing leaves is not clear. As senescence is a strongly regulated process, high levels of IAA might be necessary for its control. A possible explanation could be that the increased protein breakdown in senescing leaves leads to an increase in the content of tryptophan, which may be converted to IAA. After re-watering, a decrease in the total CK content was observed in the roots in all tested genotypes. Simultaneously, the level of bioactive CKs in the leaves increased significantly in comparison with severely stressed plants (Fig. 5, p. 82 ), which might be, at least partially, result of the re-establishment of xylem sap flow from the roots. The importance of the transpiration stream in CK transport from the roots was recently demonstrated by Aloni et al. (2006). After rehydration, the stress indicators (SAG12 promoter activity and ABA levels) sharply decreased (Table 1, p. 81 and Fig. 8, p. 85). The slightly faster growth initiation in SAG::ZOG1 plants in comparison with WT is in agreement with the reported positive effect of CKs on plant recovery (e.g. Itai et al. 1978). The slower recovery in the case of 35S::ZOG1 plants might indicates that elevated CK levels need not be advantageous in case of prolonged, severe drought stress, which is in accordance with Synková et al. (1999). To conclude, data on SAG12 promoter activity demonstrated that drought strongly accelerated senescence of tobacco plants, which coincided with the gradient of bioactive CKs in favour of the upper leaves. The gradient was maintained by the stimulation of CKX activity in the lower leaves and by its suppression in the upper leaves. Targeted, senescence-induced elevation of ZOG1 expression exhibited a positive effect on both the delay of senescence of the stressed lower leaves and the speeding up of the recovery. 30 5. Summary ➔ Cytokinins, auxin, abscisic acid and polyamines levels undergo significant diurnal variations in leaves. These rhythms have physiological relevance and correlate with circadian changes of plant growth and development. ➔ The activity of cytokinin oxidase/dehydrogenase and cytokinin N- and O-glucosylation are also undergoing diurnal changes. ➔ Arginine decarboxylase and ornithine decarboxylase activities, as well as putrescine oxidizing activity, correlates well with the alternations in polyamine content during the day/night period. ➔ Highly non-uniform distribution of cytokinins found between chloroplasts and bulk leaf tissue in different transgenic tobacco lines with modulated cytokinin content indicates that chloroplasts are relatively independent organelles with respect to the regulation of cytokinin metabolism. ➔ Data on SAG12 promoter activity demonstrated that drought strongly accelerated the senescence of tobacco plants, which coincided with the establishment of the gradient of bioactive cytokinins in favour of the upper leaves. ➔ Targeted, senescence-induced elevation of expression of gene for trans-zeatin Oglucosyltransferase exhibited a positive effect both on the delay of senescence of the stressed lower leaves and during the recovery. ➔ Uniform elevation of cytokinin content in constitutive 35S::ZOG1 transformants before stress initiation might have a negative effect during the prolonged, severe drought. 31 Literature Akiyoshi DE, Klee H, Amasino RM, Nester EW, Gordon MP. 1984. T-DNA of Agrobacterium tumefaciens encodes an enzyme of cytokinin biosynthesis. Proc Natl Acad Sci USA 81: 5994-5998. Alabadi D, Oyama T, Yanovsky MJ, Harmon FG, Mas P, Kay SA. 2001. Reciprocal regulation between TOC1 and LHY/CCA1 within the Arabidopsis circadian clock. Science 293: 880–883. Aloni R, Langhans M, Aloni E, Dreieicher E, Ullrich CI. 2005. Root-synthetized cytokinin in Arabidopsis is distributed in the shoot by the transpiration stream. J Exp Bot 56: 15351544. Aloni R, Aloni E, Langhans M, Ullrich CI. 2006. Role of cytokinin and auxin in shaping root architecture: regulating vascular differentiation, lateral root initiation, root apical dominance and root gravitropism. Ann Bot 97: 883-893. Ando S, Asano T, Tsushima S, Kamachi S, Haigo T, Tabei Y. 2005. Changes in gene expression of putative isopentenyltransferase during clubroot development in Chinese cabbage (Brassica rapa L.). Physiol Mol Plant Pathol 67: 59-67. Andreadakis A, Kotzabasis K. 1996. Changes in the biosynthesis and catabolism of polyamines in isolated plastids during chloroplast photodevelopment. J Photochem Photobiol B: Biology 33: 163-170. Angelini R, Bragaloni M, Federico R, Infantino A, Portapuglia A. 1993. Involvment of polyamines, diamine oxidase and peroxidase in resistance of chickpea to Ascochyta rabiei. J Plant Physiol 142: 704-709. Armstrong DJ. 1994. Cytokinin oxidase and regulation of cytokinin degradation. In: Cytokinin: chemistry, activity, and function. Mok DWS, Mok MC (eds), CRC, Boca Raton, pp 139-154. Åstot C, Dolezal K, Nordström A, Wang Q, Kunkel T, Moritz T, Chua N-H, Sandberg G. 2000. An alternative cytokinin biosynthesis pathway. Proc Natl Acad Sci USA 97: 1477814783. Bagni N, Tassoni A. 2001. Biosynthesis, oxidation and conjugation of aliphatic polyamines in higher plants. Amino Acids 20: 301-317. Barry GF, Rogers SG, Fraley RT, Brand L. 1984. Identification of a cloned cytokinin biosynthetic gene. Proc Natl Acad Sci USA 81: 4776-4780. 32 Bassil NV, Mok DWS, Mok MC. 1993. Partial purification of a cis-trans-isomerase of zeatin from immature seed of Phaseolus vulgaris L. Plant Physiol 102: 867-872. Beck E, Wagner BM. 1994. Quantification of the daily cytokinin transport from the root to shoot of Urtica dioica L. Bot Acta 107: 342-348. Bilyeu KD, Cole JZ, Laskey JG, Riekhof WR, Esparza TJ, Kramer MD, Morris RO. 2001. Molecular and biochemical characteization of a cytokinin oxidase from maize. Plant Physiol 125: 378-386. Biondi S, Scaramagli S, Capitani F, Altamura MM, Torrigiani P. 2001. Methyl jasmonate upregulates biosynthetic gene expression, oxidation and conjugation of polyamines, and inhibits shoot formation in tobacco thin layers. J Exp Bot 52: 231-242. Blackwell JR, Horgan R. 1994. Cytokinin biosynthesis by extracts of Zea mays. Phytochemistry 35: 339-342. Bonneau L, Carre M, Martin-Tanguy J. 1994. Polyamines and related enzymes in rice seeds differing in germination potential. Plant Growth Regul 15: 75-82. Bouchereau A, Aziz A, Larher F, Martin-Tanguy J. 1999. Polyamines and environmental challenges: recent development. Plant Sci 140: 103-125. Bowsher CG, Long DM, Oaks A, Rothstein SJ. 1991. Effect of light/dark cycles on expression of nitrate assimilatory genes in maize shoots and roots. Plant Physiol 95: 281285. Brugiere N, Jiao S, Hantke S, Zinselmeier C, Roessler JA, Niu X, Jones RJ, Habben JE. 2003. Cytokinin oxidase gene expression in maize is located to the vasculature, and is induced by cytokinins, abscisic acid, and abiotic stress. Plant Physiol 132: 1228-1240. Brugiere N, Humbert S, Rizzo N, Bohn J, Habben JE. 2008. A member of the maize isopentenyl transferase gene family, Zea mays isopentenyl transferase 2 (ZmIPT2), encodes a cytokinin biosynthetic enzyme expressed during kernel development. Plant Mol Biol 67: 215-229. Brzobohatý B, Moore I, Kristoffersen P, Bako L, Campos N, Schell J, Palme K. 1993. Release of active cytokinin by a beta-glucosidase localized to the maize root meristem. Science 262: 1051-1054. Buchanan-Wollaston V. 1997. The molecualr biology of leaf senescence. J Exp Bot 48: 181199. Burch LR, Stuchbury T. 1987. Activity and distribution of enzymes that interconvert purine bases, ribosides for cytokinin metabolism. Plant Physiol 69: 283-288. 33 Campos N, Bako L, Feldwisch J, Schell J, Palme K. 1992. A protein from maize labeled with azido-IAA has novel β-glucosidase activity. Plant J 2: 675-684. Capell T, Bassie L, Christou P. 2004. Modulation of the polyamine biosynthetic pathway in transgenic rice confers tolerance to drought stress. Proc Natl Acad Sci USA 101: 99099914. Chen CM. 1981. Biosynthesis and enzymatic regulation of the interconversion of cytokinin.In Metabolism and Molecular Activities of Cytokinins. J. Guern and C. Peaud-Lenoel, (eds). Springer-Verlag, Berlin. pp. 34-43. Chen CM, Ertl JR, Leisner SM, Chang CC. 1985. Localization of cytokinin biosynthetic sites in pea plants and carrot roots. Plant Physiol 78: 510-513. Chen CM, Kristopeit SM. 1981a. Dephosphorylation of cytokinins ribonucleotide by 5´nucleotidases from wheat germ cytosol. Plant Physiol 67: 494-498. Chen CM, Kristopeit SM. 1981b. Deribosylation of cytokinin ribonucleoside by adenosine nucleotidase from wheat germ cells. Plant Physiol 68: 1020-1023. Chen CM, Leisner SM. 1984. Modification of cytokinins by cauliflower microsomal enzymes. Plant Physiol 75: 442-446. Chen CM, Logan DM, McLennan BD, Hall RH. 1968. Studies of the metabolism of a cytokinin, N6-(Δ2-isopentenyl)adenosine. Plant Physiol 43 (suppl.): 18. Chen CM, Melitz DK. 1979. Cytokinin biosynthesis in a cell-free systém from cytokininautotrophic tobacco tissue cultures. FEBS Lett 107: 15-20. Chen CM, Melitz DK, Clough FW. 1982. Metabolism of cytokinins: phosphorylation of cytokinin bases by adenine phosphoribosyl transferase from wheat germ. Arch Biochem Biophys 214: 634-641. Chen CM, Petchow B. 1978b. Metabolism of cytokinin. Ribosylation of cytokinin base by adenosine phosphorylase from wheat germ. Plant Physiol 62: 872-874. Cherayil JD, Lipsett MN. 1977. Zeatin ribonucleosides in the transfer ribonucleic acid of Rhizobium leguminosarum, Agrobacteruim tumefaciens, Corynebacterium fascians, and Erwinia amylovora. J Bacteriol 131: 741-744. Chernyadev II. 2009. The protective action of cytokinins on the photosynthetic machinery and productivity of plants under stress. App Biochem Microbiol 45: 351-362. Chory J, Reinecke D, Sim S, Washburn T, Brenner M. 1994. A role for cytokinins in deetiolation in Arabidopsis. Plant Physiol 104: 339-347. 34 Cowan AK, Freeman M, Björkman PO, Nicander B, Sitbon F, Tillberg E. 2005. Effects of senescence-induced alternation in cytokinin metabolism on source-sink relationships and ontogenic and stress-induced transitions in tobacco. Planta 221: 801-814. Creus JA, Eucuentra A, Gavalda EG, Barcelo J. 1991. Binding of polyamines to different macromolecules in plants. In: Polyamines as modulators of plant development. Galston AW, Tiburcio AF, (eds). Madrid: Ediciones Peninsular, pp. 30-34. D´Agostino I, Deruère J, Kieber JJ. 2000. Characterization of the response of the Arabidopsis ARR gene family to cytokinin. Plant Physiol 124: 1706-1717. Dixon SC, Martin RC, Mok MC, Shaw G, Mok DWS. 1989. Zeatin glycosylation enzymes in Phaseolus: isolation of O-glucosyltransferase from P. lunatus and comparison to Oxylosyltransferase from P. vulgaris. Plant Physiol 90: 1316-1321. Dobrev PI, Motyka V, Gaudinová A, Malbeck J, Trávníčková A, Kamínek M, Vaňková R. 2002. Transient accumulation of cis- and trans-zeatin type cytokinins and its relations to cytokinin oxidase activity during cell cycle of synchronized tobacco cells. Plant Physiol Biochem 40: 333-337. Dunlap JC, Loros JJ, Denault D, Lee K, Froehlich AF, Colot H, Shi M, Pregueiro A. 2004. Genetics and molecular biology of circadian rhythms. In: Biochemistry and molecular biology: The Mycota III, Brambl R, Marzluf GA (eds), Springer-Verlag, Berlin, pp. 209. Ehneß R, Roitsch T. 1997. Co-ordinated induction of mRNAs for extracellular invertase and a glucose transporter in C. rubrum by cytokinins. Plant J 11: 539-548. Emery RJN, Ma Q, Atkins CA. 2000. The forms and sources of cytokinins in developing white lupine seeds and fruits. Plant Physiol 123: 1593-1604. Eriksson ME, Hanano S, Southern MM, Hall A, Millar AJ. 2003. Response regulator homologues have complementary light-dependent functions in the Arabidopsis circadian clock. Planta 218: 59-62. Esen A. 1992. Purification and partial characterization of maize (Zea mays L) β-glucosidase. Plant Physiol 98: 174-182. Fox JE, Cornette J, Deleuze G, Dyson W, Giersak C, McChesney JD. 1973. The formation, isolation and biological activity of a cytokinin 7-glucoside. Plant Physiol 52: 627-632. Fox JE, Sood CK, Buckwalter B, McChesney JD. 1971. Metabolism and biological activity of 9-substituated cytokinins. Plant Physiol 47: 275-281. Frébort I, Šebela M, Galuszka P, Werner T, Schmülling T, Peč P. 2002. Cytokinin oxidase/cytokinin dehydrogense assay: optimized procedures and applications. Anal Biochem 306: 1-7. 35 Frébortová J, Fraaije MW, Galuszka P, Šebela M, Peč P, Hrbáč J, Novák O, Bilyeu KD, English JT, Frébort I. 2004. Catalytic reaction of cytokinin dehydrogenase: preference for quinones as electron acceptores. Biochem J 380: 121-130. Fusseder A, Ziegler P. 1988. Metabolism and compartmentation of dihydrozeatin exogenously supplied to photoautotrophic suspension cultures of Chenopodium rubrum. Planta 173: 104-109. Fusseder A, Wartinger A, Hartung W, Schulze ED, Heilmeier H. 1992. Cytokinins in the xylem sap of desert-grown almond (Prunus dulcis) trees: daily courses and their possible interactions with abscisic acid and leaf conductance. New Phytol 122: 45-52. Galston AW, KaurSawhney R, Altabella T, Tiburcio AF. 1997. Plant polyamines in reproductive activity and response to abiotic stress. Bot Acta 110: 197-207. Galston AW, KaurSawhney R. 1995. Polyamines as endogenous growth regulators. In: Plant hormones, physiology, biochemistry and molecular biology(2nd edn.). Davies PJ (ed), Kluwer Academic Publishers, Dordrecht, The Netherlands, pp. 158-178. Galuszka P, Frébort I, Šebela M, Sauer P, Jacobsen S, Peč P. 2001. Cytokinin oxidase or dehydrogenase? Mechanism of cytokinin degradation in cereals. Eur J Biochem 268: 450461. Galuszka P, Popelková H, Werner T, Frébortová J, Pospíšilová H, Mik V, Köllmer I, Schmülling T, Frébort I. 2007. Biochemical characterization and histochemical localization of cytokinin oxidases/dehydrogenases from Arabidopsis thaliana expressed in Nicotiana tabaccum L. J Plant Growth Regul 26: 255-267. Gan S, Amasino RM. 1995. Inhibition of leaf senescence by autoregulated production of cytokinin. Science 270: 1986-1988. Gan S, Amasino RM. 1996. Cytokinins in plant senescence: From spray and pray to clone and play. Bio Essays 18: 557-565. Gan S, Amasino RM. 1997. Making sense of senescence. Molecular genetic regulation of leaf senescence. Plant Physiol 113: 313-319. Gaudinová A, Dobrev PI, Šolcová B, Novák O, Strnad M, Fridecký D, Motyka V. 2005. The involvement of cytokinin oxidase/dehydrogenase and zeatin reductase in regulation of cytokinin levels in pea (Pisum sativum L.) leaves. J Plant Growth Regul 24: 188-200. Gemperlová L, Eder J, Cvikrová M. 2005. Polyamine metabolism during the growth cycle of tobacco BY-2 cells. Plant Physiol Biochem 43: 375-381. 36 Goicoechea N, Dolezal K, Antolin MC, Strnad M, Sanchez-Diaz M. 1995. Influence of mycorrhizae and Rhizobium on cytokinin content in drought-stressed alfalfa. J Exp Bot 46: 1543-1549. Golovko A, Sitbon F, Tillberg E, Nicauder B. 2002. Identification of a tRNA isopentenyltransferase gene from Arabidopsis thaliana. Plant Mol Biol 49: 161-169. Hall RH. 1970. N6-(Δ2-isopentenyl)adenosine: chemical reactions, biosynthesis, metabolism and significance to the structure and function of tRNA. Prog Nucleic Acid Res Mol Biol 10: 57-86. Hall RH. 1971. The modified nucleosides. In: Nucleic acid, Columbia University Press, New York, NY, pp 317-346. Hänsch R, Fessel DG, Witt C, Hesberg C, Hoffmann G, Walch-Liu P, Engels C, Kruse J, Rennenberg H, Kaiser WM, Mendel RR. 2001. Tobacco plants that lack expression of functional nitrate reductase in roots show changes in growth rates and metabolite accumulation. J Exp Bot 52: 1251-1258. Hare PD, Van Staden J. 1994a. Cytokinin oxidase: bichemical features and physiological significance. Physiol Plant 91: 128-136. Hare PD, Van Staden J. 1997. The molecular basis of cytokinin action. Plant Growth Regul 23: 41-78. Hass C, Weinl S, Harter K, Kudla J. 2004a. Functional genomics of protein phosphorylation in Arabidopsis thaliana. Plant Functional Genomics, Haworth, Binghampton, NY. Hayama R, Coupland G. 2003. Shedding light on the circadian clock and the photoperiodic control of flowering. Curr Opin Plant Biol 6: 3-9. Horgan R. 1984. Cytokinins. In: Advanced plant physiology, Wilkins MB (ed), Longmans, London, pp. 89-101. Horgan R, Hewett EW, Purse JG, Wareing PF. 1973. A new cytokinin from Populus robusta. Tetrahedron Lett. 14: 2827-22828. Hou B, Lim EK, Higgins GS, Bowles DJ. 2004. N-glucosylation of cytokinins by glycosyltransferase of Arabidopsis thaliana. FEBS Lett. 554: 373-380. Hwang I, Chen HC, Sheen J. 2002. Two-component signal transduction pathways in Arabidopsis. Plant Physiol 129: 500-515. Imamura A, Hanaki N, Nakamura A, Suzuki T, Taniguchi M, Kiba T, Ueguchi C, Sugiyama T, Mizuno T. 1999. Compilation and characterization of Arabidopsis thaliana response regulators implicated in His-Asp phosphorelay signal transduction. Plant Cell Physiol 40: 733-742. 37 Inoue T, Higuchi M, Hashimoto Y, Seki M, Kobayashi M, Kato T, Tabata S, Shinozaki K, Kakimoto T. 2001. Identification of CRE1 as a cytokinin receptor from Arabidopsis. Nature 409: 1060-1063. Itai C, Benzioni A, Munz S. 1978. Heat stress: effects of abscisic acid and kinetin on response and recovery of tobacco leaves. Plant Cell Physiol 19: 453-459. Jones RJ, Schreiber BMN. 1997. Role and function of cytokinin oxidase in plants. Plant Growth Regul 23: 122-134. Kakimoto T. 2001a. Identification of plant cytokinin biosynthetic enzymes as dimethylallyl diphosphate: ATP/ADP isopentenyltransferases. Plant Cell Physiol 42: 677-685. Kakimoto T. 2003. Perception and signal transduction of cytokinins. Annu Rev Plant Biol 54: 605-627. Kamada-Nobusada T, Sakakibara H. 2009. Molecular basis for cytokinin biosynthesis. Phytochemistry 70: 444-449. Kamínek M, Armstrong DJ. 1990. Genotypic variation in cytokinin oxidase from Phaseolus callus cultures. Plant Physiol 93: 1530-1538. Kasahara H, Hanada A, Kuzuyama T, Takagi M, Kamiya Y, Yamaguchi S. 2002. Contribution of the mevalonate and methylerythritol phosphate pathways to the biosynthesis of gibberellins in Arabidopsis. J Biol Chem 277: 45188-45194. Kasahara H, Takei K, Ueda N, Hishiyama S, Yamaya T, Kamiya Y, Yamaguchi S, Sakakibara H. 2004. Distinct isoprenoid origins of cis- and trans-zeatin biosynthesis in Arabidopsis. J Biol Chem 279: 14049-14054. Kiba T, Taniguchi M, Imamura A, Ueguchi C, Mizuno T, Sugiyama T. 1999. Differential expression of genes for response regulators in response to cytokinins and nitrate in Arabidopsis thaliana. Plant Cell Physiol 40: 767-771. Kiba T, Yamada H, Mizuno T. 2002. Characterization of the ARR15 and ARR16 response regulators with special reference to the cytokinin signalling patheway mediated by the AHK4 histidine kinase in roots of Arabidopsis thaliana. Plant Cell Physiol 43: 1059-1066. Kiran NS, Polanská L, Fohlerová R, Mazura P, Válková M, Šmeral M, Zouhar J, Malbeck J, Dobrev PI, Macháčková I, Brzobohatý B. 2006. Ectopic expression of the maize bglucosidase Zm-p60.1 perturbs cytokinin homeostasis in transgenic tobacco. J Exp Bot 57: 985-996. Koetje DS, Kononowicz H, Hodges TK. 1993. Polyamine metabolism associated with growth and embryogenic potential of rice. J Plant Physiol 141: 215-221. 38 Kowalska M, Galuszka P, Frebortova J, Sebela M, Beres T, Hluska T, Smehilova M, Bilyeu KD, Frebort I. 2010. Vacuolar and cytosolic cytokinin dehydrogenases of Arabidopsis thaliana. Heterologous expression, purification and properties. Phytochemistry 71: 19701978. Kraepiel Y, Rousselin P, Sotta B, Kerhoas L, Einhorn J, Caboche M, Miginiac E. 1994. Analysis of phytochrome- and ABA-deficient mutants suggests that ABA degradation is controlled by light in Nicotiana plumbaginifolia. Plant J 6: 665-672. Kristoffersen P, Brzobohatý B, Hohfeld I, Bako L, Melkonian M, Palme K. 2000. Developmental regulation of the maize Zm-p60.1 gene encoding a β-glucosidase localized to plastids. Planta 210: 407-415. Kumar A, Altabella T, Taylor MA, Tiburcio AF. 1997. Recent advances in polyamine research. Trends Plant Sci 2: 124-136. Kurakawa T, Ueda N, Maekawa M, Kobayashi K, Kojima M, Nagato Y, Sakakibara H, Kyozuka J. 2007. Direct control of shoot meristem activity by a cytokinin-activating enzyme. Nature 445: 652-655. Kurapov PB, Skorobogatova IV, Sal'nikova EI, Sorkina GL, Siusheva AG. 2000. Diurnal courses of endogenous phytohormones in barley. Izvestiya Akademii Nauk Seriya Biologicheskaya 1: 108-114. Kuroha T, Tokunaga H, Kojima M, Ueda N, Ishida T, Nagawa S, Fukuda H, Sugimoto K, Sakakibara H. 2009. Functional analysis of LONELY GUY cytokinin-activating enzymes reveal the importance of the direct activation pathway in Arabidopsis. Plant Cell 21: 31523169. Kusnetsov VV, Oelmüller R, Sarwatt MI, Porfirova SA, Cherepneva GN, Herrmann RG, Kulaeva ON. 1994. Cytokinins, abscisic acid and light affect accumulation of chloroplast proteins in Lupinus luteus cotyledons without notable effect on steady-state mRNA levels. Planta 194: 318-327. Kwade Z, Swiatek A, Azmi A, Goossens A, Inze D, van Onckelen H, Roef L. 2005. Identification of four adenosine kinase isoforms in tobacco BY-2 cells and their putative role in the cell cycle-regulated cytokinin metabolism. J. Biol. Chem. 280: 17512-17519. Lara MEB, Garcia MCG, Fatima T, Ehneß R, Lee TK, Proels R, Tanner W, Roitsch T. 2004. Extracellular invertase is an essential component of cytokinin-mediated delay of senescence. Plant Cell 16: 1276-1287. Letham DS. 1963. Zeatin, a factor inducing cell division from Zea mays. Life Sci. 8: 569-573. 39 Letham DS. 1978. Cytokinins. In: Phytohormones and related compounds: a comprehensive Ttreatise. Letham DS, Goodwin PB, Higgins THV (eds), Elsevier, Amsterdam, The Netherlands, pp. 205-263. Letham DS. 1994. Cytokinins as phytohormones – site of biosynthesis, translocation, and function of translocated cytokinin. In: Cytokinins: Chemistry, Activity, and Function. D. W. S. Mok and M. C. Mok (eds), CRC Press, Boca Raton, FL. pp. 57-80. Letham DS, Palni LMS. 1983. The biosynthesis and metabolism of cytokinins. Annu Rev Plant Physiol 34: 163-197. Lichtenthaler HK. 1999. The 1-deoxy-D-xylulose-5-phosphate pathway of isoprenoid biosynthesis in plants. Annu Rev Plant Physiol Mol Biol 50: 47-65. Lohrmann J, Harter K. 2002. Plant two-component signalling systems and the role of response regulators. Plant Physiol 128: 363-369. Lu JL, Ertl J, Chen CM. 1990. Cytokinin enhancement of the light induction of nitrate reductase transcript levels in etiolated barley leaves. Plant Mol Biol 14: 585-594. Martin RC, Cloud KA, Mok MC, Mok DWS. 2000. Substrate specificity and domain analyses of zeatin O-glycosyltransferases. Plant Growth Regul 32: 289-293. Martin RC, Mok DWS, Smets R, Van Onckelen HA, Mok MC. 2001a. Development of transgenic tobacco harbouring a zeatin O-glucosyltransferase gene from Phaseolus. In Vitro Cell Dev Biol Plant 37: 354-360. Martin RC, Mok MC, Habben JE, Mok DWS. 2001b. A maize gene encoding an Oglucosyltransferase specific to cis-zeatin. Proc Natl Acad Sci USA 98: 5922-5926. Martin RC, Mok MC, Shaw G, Mok DWS. 1989. An enzyme mediating the conversion of zeatin to dihydrozeatin in Phaseolus embryos. Plant Physiol 90: 1630-1635. Martin-Tanguy J. 1985. The occurence and possible function of hydroxycinnamoyl acid amides in plants. Plant Growth Reg 3: 381-399. Martin-Tanguy J. 2001. Metabolism and function of polyamines in plants: recent development (new approaches). Plant Growth Regul 34: 135-148. Masclaux-Daubresse C, Valadier MH, Carrayol E, Reisdorf-Cren M, Hirel B. 2002. Diurnal changes in the expression of glutamate dehydrogenase and nitrate reductase are involved in the C/N balance of tobacco source leaves. Plant Cell Environ 25: 1451-1462. Mansfield TA, McAinsh MR. 1995. Hormones as regulators of water balance. In: Plant hormones. Physiology, biochemistry and molecular biology, Davies PJ (ed), Kluwer Academic Publishers, Dordrecht, The Netherlands, pp. 598-616. 40 McGaw BA, Heald JK, Horgan R. 1984. Dihydrozeatin in radish seedlings. Phytochemistry 23: 1373-1377. Miller CO, Skoog F, Okumura FS, von Saltza MH, Strong FM. 1955. Structure and synthesis of kinetin. J. Am. Chem. Soc. 78: 2662-2663. Miller CO, Skoog F, von Saltza MH, Strong FM. 1955. Kinetin, a cell division factor from deoxyribonucleic acid. J. Am. Chem. Soc. 77: 1392. Miyawaki K, Matsumoto-Kitano M, Kakimoto T. 2004. Expression of cytokinin biosynthetic isopentenyltransferase genes in Arabidopsis: tissue specificity and regulation by auxin, cytokinin, and nitrate. Plant J 37: 128-138. Miyawaki K, Tarkowski P, Matsumoto-Kitano M, Kato T, Sato S, Tarkowská D, Tabata S, Sandberg G, Kakimoto T. 2006. Roles of Arabidopsis ATP/ADP isopentenyltransferases and tRNA isopentenyltransferases in cytokinin biosynthesis. Proc Natl Acad Sci USA 103: 16598-16603. Mizoguchi T, Niinuma K, Yoshida R. 2007. Day-neutral response of photoperiodic flowering in tomatoes: possible implications based on recent molecular genetics of Arabidopsis and rice. Plant Biotech 24: 83-86. Mizoguchi T, Wright L, Fujiwara S, Cremer F, Lee K, Onouchi H, Mouradov A, Fowler S, Kamada H, Putterill J, Coupland G. 2005. Distinct roles of GIGANTEA in promoting flowering and regulating circadian rhythms in Arabidopsis. Plant Cell 17: 2255–2270. Mok DWS, Mok MC. 2001. Cytokinin metabolism and action. Annu Rev Plant Physiol Plant Mol Biol 52: 89-118. Mok DWS, Martin RC, Shan X, Mok MC. 2000. Genes encoding zeatin Oglycosyltransferases. Plant Growth Regul 32: 285-287. Morris RO. 1995. Genes specifying auxin and cytokinin biosynthesis in prokaryotes. In: Plant Hormones. Physiology, Biochemistry and Molecular Biology. Davies PJ (ed), Kluwer Academic Publishers, Dordrecht, pp. 318-339. Morris RO, Bilyeu KD, Laskey JG, Cheikh N. 1999. Isolation of a gene encoding a glycosylated cytokinin oxidase from maize. Biochem Biophys Res Comm 255: 328-333. Motyka V, Faiss M, Strnad M, Kaminek M, Schmülling T. 1996. Changes in cytokinin content and cytokinin oxidase activity in response to derepression of ipt gene transcription in trangenis tobacco calli and plants. Plant Physiol 112: 1035-1043. Motyka V, Vaňková R, Čapková V, Petrášek J, Kamínek M, Schmülling T. 2003. Cytokinininduced upregulation of cytokinin oxidase activity in tobacco includes changes in enzyme glycosylation and secretion. Plant Physiol 117: 11-21. 41 Murai N. 1994. Cytokinin biosynthesis in tRNA and cytokinin incorporation into plant RNA. In: Mok DWS, Mok MC (eds) Cytokinin: Chemistry, Activity and Function, CRC, Boca Raton, pp 87-99. Nakamichi N, Kusano M, Fukushima A, Kita M, Ito S, Yamashino T, Saito K, Sakakibara H, Mizuno T. 2009. Transcript Profiling of an Arabidopsis PSEUDO RESPONSE REGULATOR Arrhythmic Triple Mutant Reveals a Role for the Circadian Clock in Cold Stress Response. Plant Cell Physiol 50: 447 – 462. Noodén LD. 1988. The phenomena of senescence and aging. In: Senescence and aging in plants. L. D. Noodén and A. C. Leopold (eds.), Academie Russ Inc. San Diego, Calif, pp. 2-50. Nordström A, Tarkowski P, Tarkowska D, Norbaek R, Åstot C, Dolezal K, Sandberg G. 2004. Auxin regulation of cytokinin biosynthesis in Arabidopsis thaliana: a factor of potential importance for auxin-cytokinin-regulated development. Proc Natl Acad Sci USA 101: 8039–8044. Nováková M, Motyka V, Dobrev PI, Malbeck J, Gaudinová A, Vaňková R. 2005. Diurnal variation of cytokinin, auxin and abscisic acid levels in tobacco leaves. J Exp Bot 56: 2877-2883. Pačes V, Werstink E, Hall RH. 1971. Conversion of N6-(Δ2-isopentenyl)adenosine to adenosine by enzyme activity in tobacco tissue. Plant Physiol 48: 775-778. Parker CW, Badenoch-Jones J, Letham DS. 1989. Radioimmunoassay for quantifying the cytokinins cis-zeatin and cis-zetin riboside and its aplication to xylem sap samples. J Plant Growth Regul 8: 93-105. Peng J, Peng FT, Zhu CF, Wei SC. 2008. Molecular cloning of a putative gene encoding isopentenyltransferase from pingyitiancha (Malus hupehensis) and characterization of its response to nitrate. Tree Physiol 28: 899-904. Pfosser M, Konigshofer H, Kandeler R. 1990. Free, conjugated, and bound polyamines during the cell cycle of synchronized cell suspension cultures of Nicotiana tabacum. J Plant Physiol 136: 574-579. Quirino BF, Normanly J, Amasino RM. 1999. Diverse range of gene activity during Arabidopsis thaliana leaf senescence includes pathogen-independent induction of defenserelated genes. Plant Mol Biol 40: 267-278. Rashotte AM, Carson SDB, To JPC, Kieber JJ. 2003. Expression profiling of cytokinin action in Arabidopsis. Plant Physiol 132: 1998-2011. 42 Riccardi F, Gazeau P, Jacquemot MP, Vincent D, Zivy M. 2004. Deciphering genetic variations of proteome responses to water deficit in maize leaves. Plant Physiol Biochem 42: 1003-1011. Riou-Khamlichi C, Huntly R, Jacqmard A, Murray JAH. 1999. Cytokinin activation of Arabidopsis cell division through a D-type cyclin. Science 283: 1541-1544. Rohmer M. 1999. The discovery of a mevalonate-independent pathway for isoprenoid biosynthesis in bacteria, algae and higher plants. Nat Prod Rep 16: 565-574. Rohmer M. 2003. Mevalonate-independent methylerythritol phosphate pathway for isoprenoid biosynthesis: elucidation and distribution. Pure Appl Chem 75: 375-387. Rufty TW, Kerr PS, Huber SC. 1983. Characterization of diurnal changes in activities of enzymes involved in sucrose biosynthesis. Plant Physiol 73: 428-433. Sakamoto T, Sakakibara H, Kojima M, Yamamoto Y, Nagasaki H, Inukai Y, Sato Y, Matsuoka M. 2006. Ectopic expression of KNOTTED 1-like homeobox protein induces expression of cytokinin biosynthesis genes in rice. Plant Physiol 142: 54-62. Sakano Y, Okada Y, Matsunaga A, Suwama T, Kaneko T, Ito K, Noguchi H, Abe I. 2004. Molecular cloning, expression, and characterization of adenylate isopentenyltransferase from hop (Humulus lupulus L.). Phytochemistry 65: 2439-2446. Samuelson ME, Larsson C-M. 1993. Nitrate regulation of zeatin riboside levels in barley roots: effects of inhibitors of N assimilation and comparison with ammonium. Plant Sci 93: 77-84. Schmülling T, Werner T, Riefler M, Krupková E, Bartrina y Manns I. 2003. Structure and function of cytokinin oxidase/dehydrogenase genes of maize, rice, Arabidopsis and other species. J Plant Res 116: 241-252. Shantz EM, Steward FC. 1955. The identification of compound A from coconut milk as 1,3diphenylurea. J. Am. Chem. Soc. 77: 6351-6353. Shelp BJ, Bown AW, McLean MD. 1999. Metabolism and functions of gamma-aminobutyric acid. Trends Plant Sci 4: 446-452. Shukla A, Sawhney VK. 1997. Cytokinin metabolism and cytokinin oxidase and adenine phosporibosyltransferase activity in male sterile Brassica napus leaves. Phytochemistry 44: 377-381. Singh S, Letham DS, Palni LMS. 1992. Cytokinin biochemistry in relation to leaf senescence: VII. Endogenous cytokinin levels and exogenous applications of cytokinins in relation to sequestial leaf senescence of tobacco. Physiol Plant 86: 388-397. 43 Slocum RD. 1991a. Polyamine biosynthesis in plants. In: Biochemistry and physiology of polyamines in plants. Slocum RD, Flores HE (eds), CRC Press, Boca Raton, pp. 22-40. Slocum RD. 1991b. Tissue and subcellular localisation of polyamine metabolism. In: Biochemistry and physiology of polyamines in plants. Slocum RD, Flores HE (eds), CRC Press, Boca Raton, pp. 93-103. Smith TA, Marshall JHA. 1988. The di and polyamine oxidases of plants. In: Progress in polyamine research (Advances in experimental biology and medicine, 250), Plenum Press, New York, pp. 573-587. Soni R, Carmichael JP, Shah ZH, Murray JAH. 1995. A family of cyclin D homologs from plants differentially controlled by growth regulators and containing the conserved retinoblastoma protein interaction motif. Plant Cell 7: 85-103. Spichal L, Rakova NY, Riefler M, Mizuno T, Romanov GA, Strnad M, Schülling T. 2004. Two cytokinin receptors of Arabidopsis thaliana, CRE1/AHK4 and AHK3, differ in their ligand specificity in a bacterial assay. Plant Cell Physiol 45: 1299-1305. Stetler DA, Laetsch WM. 1965. Kinetin-induced chloroplast maturation in cultures of tobacco tissue. Science 149: 1387-1388. Stoynova EZ, Iliev LK, Georgiev GT. 1996. Structural and functional alternations in radish plants induced by the phenylurea cytokinin 4-PU-30. Biol Plant 38: 237-244. Strayer C, Oyama T, Schultz TF, Raman R, Somers DE, Mas P, Pauda S, Kreps JA, Kay SA. 2000. Cloning of the Arabidopsis clock gene TOC1, an autoregulatory response regulator homolog. Science 289: 768-771. Strnad M. 1997. The aromatic cytokinins. Plant Physiol 101: 674-688. Suttle JC, Banowetz GM. 2000. Changes in cis-zeatin and cis-zeatin riboside levels and biological activity during tuber dormancy. Physiol Plant 109: 68-74. Synková H, Schnablová R, Polanská L, Hušák M, Šiffel P, Vácha F, Malbeck J, Macháčková I, Nebesářová J. 2006. Three-dimensional reconstruction of anomalous chloroplasts in transgenic ipt tobacco. Planta 223: 659-671. Synková H, Van Loven K, Pospíšilová J, Valcke R. 1999. Photosynthesis of transgenic pssuipt tobacco. J Plant Physiol 155: 173-182. Takagi M, Yokota T, Murofushi N, Saka H, Takahashi N. 1989. Quantitative changes of free base, riboside, robotide and glucoside cytokinins in developing rice grains. Plant Growth Regul 8: 349-364. 44 Takei K, Sakakibara H, Sugiyama T. 2001a. Identification of genes encoding adenylate isopentenyltransferase, a cytokinin biosynthesis enzyme, in Arabidopsis thaliana. J Biol Chem 276: 26405-26410. Takei K, Sakakibara H, Taniguchi M, Sugiyama T. 2001. Nitrogen-dependent accumulation of cytokinins in root and the translocation to leaf: implication of cytokinin species that induces gene expression of maize response regulator. Plant Cell Physiol 42: 85-93. Takei K, Takahashi T, Sugiyama T, Yamaya T, Sakakibara H. 2002. Multiple routes communicating nitrogen availibility from roots to shoots: a signal transduction pathway mediated by cytokinin. J Exp Bot 53: 971-977. Takei K, Ueda N, Aoki K, Kuromori T, Hirayama T, Shinozaki K, Yamaya T, Sakakibara H. 2004. AtIPT3 is a key determinant of nitrate-dependent cytokinin biosynthesis in Arabidopsis. Plant Cell Physiol 45: 1053-1062. Tallman G. 2004. Are diurnal patterns of stomatal movement the result of alternating metabolism of endogenous guard cell ABA and accumulation of ABA delivered to the apoplast around guard cells by transpiration? J Exp Bot 55: 1963-1976. Tanaka Y, Suzuki T, Yamashino T, Mizuno T. 2004. Comparative studies of the AHP histidine-containing phosphotransmitters implicated in His-to-Asp phosphorelay in Arabidopsis thaliana. Biosci Biotechnol Biochem 68: 462-465. Tanaka M, Takei K, Kojima M, Sakakibara H, Mori H. 2006. Auxin controls local cytokinin biosynthesis in the nodal stem in apical dominance. Plant J 45: 1028–1036. Taya Y, Tanaka Y, Mishimura S. 1978. 5´-AMP is a direct precursor of cytokinin in Dictyostelium discoideum. Nature 27: 545-547. Theiss C, Bohley P, Voigt J. 2002. Regulation by polyamines of ornithine decarboxylase activity and cell division in the unicellular green alga Chlamydomonas reinhadtii. Plant Physiol 128: 1470-1479. Thomas T, Thomas TJ. 2001. Polyamine in cell growth and cell death: Molecular mechanisms and therapeutic aplications. Cell Mol Life Sci 58: 244-248. Tiburcio AF, Altabella T, Borrell A, Masgrau C. 1997. Polyamine metabolism and its regulation. Phys Plant 100: 664-674. Torrigiani P, Serafini-Fracassini D, Biondi S, Bagni N. 1986. evidence for the subcellular localization of polyamines and their biosynthetic enzymes in plant cells. J Plant Physiol 124: 23-29. Tucker DE, Allen DJ, Ort DR. 2004. Control of nitrate reductase by circadian and diurnal rhytms in tomato. Planta 219: 277-285. 45 Turner JE, Mok DWS, Mok MC, Shaw G. 1987. Isolation and partial purification of an enzyme catalyzing the formation of O-xylosylteatin in Phaseolus vulgaris embryos. Proc Natl Acad Sci USA 84: 3714-3717. Urao T, Yamaguchi-Shinozaki K, Shinozaki K. 2000a. Two-component systems in plant signal transduction. Trends Plant Sci 5: 67-74. Von Schwartzenberg K, Kruse S, Reski R, Moffatt BA, Laloue M. 1998. Cloning and characterization of an adenosine kinase from Physcomitrella involved in cytokinin metabolism. Plant J 13: 249-257. Walters D. 2003. Resistence to plant pathogens: possible roles for free polyamines and polyamine catabolism. New Phytol 159: 109-115. Wang TL, Thompson AG, Horgan R. 1977. A cytokinin glucoside from the leaves of Phaseolus vulgaris L. Planta 135: 285-288. Weaver LM, Gan S, Qurino B, Amasino RM. 1998. A comparison of the expression patterns of several senescence-associated genes in response to stress and hormone treatment. Plant Mol Biol 37: 455-469. Werner T, Motyka V, Laucou V, Smets R, Van Onckelen H, Schmülling T. 2003. Cytokinindeficient transgenic Arabidopsis plants show multiple developmental alternations indicating opposite functions of cytokinins in the regulation of shoot and root meristem activity. Plant Cell 15: 2532-2550. Werner T, Motyka V, Strnad M, Schmülling T. 2001. Regulation of plant growth by cytokinin. Proc Nat Acad Sci, USA 98: 10487-10492. Werner T, Schmülling T. 2009. Cytokinin action in plant development. Curr Opinion Plant Biol 12: 527-538. West AH, Stock AM. 2001. Histidine kinases and response regulator proteins in twocomponent signalling systems. Trends Biochem Sci: 26: 369-376. Whitty CD, Hall RH. 1974. A cytokinin oxidase in Zea mays. Can J Biochem 52: 787-799. Wilhelmová N, Kutík J. 1995. Influence of exogenously applied 6-benzylaminopurine on the structure of chloroplasts and arrangement of their membranes. Photosynthetica 31: 559570. Yamada H, Suzuki T, Terada K, Takei K, Ishikawa K, Miwa K, Yamashino T, Mizuno T. 2001. The Arabidopsis AHK4 histidine kinase is a cytokinin-binding receptor that transduces cytokinin signals across the membrane. Plant Cell Physiol 42: 1017-1023. 46 Ye CJ, Wu SW, Kong FN, Zhou CJ, Yang QK, Sun Y, Wang B. 2006. Identification and characterization of an isopentenyltransferase (IPT) gene in soybean (Glycine max L.). Plant Sci 170: 542-550. Yevdakova NA, von Schwartzenberg K. 2007. Characterisation of a prokaryote-type tRNAisopentenyltransferase gene from the moss Physcomitrella patens. Planta 226: 683-695. Zhang K, Letham DS, John PCL. 1996. Cytokinin controls the cell cycle at mitosis by stimulating the tyrosine dephosphorylation and activation of p34cdc2-like H1 histone kinase. Planta 200: 2-12. Zubko E, Adams CJ, Macháčková I, Malbeck J, Scollan C, Meyer P. 2002. Activation tagging identifies a gene from Petunia hybrida responsible for the production of active cytokinins in plants. Plant J 29: 797-808. 47 Journal of Experimental Botany, Vol. 56, No. 421, pp. 2877–2883, November 2005 doi:10.1093/jxb/eri282 Advance Access publication 12 September, 2005 This paper is available online free of all access charges (see http://jxb.oxfordjournals.org/open_access.html for further details) RESEARCH PAPER Diurnal variation of cytokinin, auxin and abscisic acid levels in tobacco leaves Marie Nováková1,2, Václav Motyka1, Petre I. Dobrev1, Jiřı́ Malbeck1, Alena Gaudinová1 and Radomı́ra Vanková1,* 1 Institute of Experimental Botany AS CR, Rozvojova 135, 165 02 Prague 6, Czech Republic 2 Department of Biochemistry, Faculty of Science, Charles University, Hlavova 2030/8, 128 43 Prague 2, Czech Republic Received 15 April 2005; Accepted 5 August 2005 Abstract As many processes are regulated by both light and plant hormones, evaluation of diurnal variations of their levels may contribute to the elucidation of the complex network of light and hormone signal transduction pathways. Diurnal variation of cytokinin, auxin, and abscisic acid levels was tested in tobacco leaves (Nicotiana tabacum L. cv. Wisconsin 38) grown under a 16/8 h photoperiod. The main peak of physiologically active cytokinins (cytokinin bases and ribosides) was found after 9 h of light, i.e. 1 h after the middle of the light period. This peak coincided with the major auxin peak and was closely followed by a minor peak of abscisic acid. Free abscisic acid started to increase at the light/dark transition and reached its maximum 3 h after dark initiation. The content of total cytokinins (mainly N-glucosides, followed by cis-zeatin derivatives and nucleotides) exhibited the main peak after 9 h of light and the minor peak after the transition to darkness. The main, midday peak of active cytokinins was preceded by a period of minimal metabolic conversion of tritiated trans-zeatin (less than 30%). The major cytokinin-degrading enzyme, cytokinin oxidase/ dehydrogenase (EC 1.5.99.12), exhibited maximal activity after the dark/light transition and during the diminishing of the midday cytokinin peak. The former peak might be connected with the elimination of the long-distance cytokinin signal. These cytokinin oxidase/dehydrogenase peaks were accompanied by increased activity of b-glucosidase (EC 3.2.1.21), which might be involved in the hydrolysis of cytokinin O-glucosides and/or in fine-tuning of active cytokinin levels at their midday peak. The achieved data indicate that cytokinin metabolism is tightly regulated by the circadian clock. Key words: Abscisic acid, auxin, cytokinin, cytokinin oxidase/ dehydrogenase, diurnal rhythm, b-glucosidase, tobacco. Introduction It is assumed that plant growth and development are regulated by the interactions between the environment and endogenous factors, especially hormones. Light is an important environmental signal that activates and modifies various growth and developmental processes in plants, for example, seed germination, flower induction, leaf movements, stem elongation, and de-etiolation. All of these processes are regulated by plant hormones as well. Effects of light and phytohormones can be synergistic (e.g. chloroplast development and cytokinins [CKs]), additive (e.g. inhibition of hypocotyls elongation and CKs), or antagonistic (e.g. germination and abscisic acid [ABA]). There are numerous reports that light itself may influence the levels of different hormones. Light was found to regulate directly the biosynthesis of active gibberellins (Foster and Morgan, 1995) and ethylene (Jasoni et al., 2002) as well as to promote the degradation of ABA (Kraepiel et al., 1994). The effect of light on auxin levels is less transparent and may be realized by influencing auxin * To whom correspondence should be addressed. Fax: +420 220 390 446. E-mail: [email protected] Abbreviations: ABA, abscisic acid; CK, cytokinin; CKX, cytokinin oxidase/dehydrogenase; DHZ, dihydrozeatin; DHZR, dihydrozeatin riboside; IAA, b-indolylacetic acid; iP, N 6-(D2-isopentenyl)adenine; iPR, N 6-(D2-isopentenyl)adenosine; iP7G, N6-(D2-isopentenyl)adenine 7-glucoside; iP9G, N6-(D2-isopentenyl) adenine 9-glucoside; Z, trans-zeatin; ZR, trans-zeatin riboside; ZOG, trans-zeatin O-glucoside; ZROG, trans-zeatin riboside O-glucoside; Z7G, trans-zeatin 7-glucoside; Z9G, trans-zeatin 9-glucoside. ª The Author [2005]. Published by Oxford University Press [on behalf of the Society for Experimental Biology]. All rights reserved. The online version of this article has been published under an Open Access model. Users are entitled to use, reproduce, disseminate, or display the Open Access version of this article for non-commercial purposes provided that: the original authorship is properly and fully attributed; the Journal and the Society for Experimental Biology are attributed as the original place of publication with the correct citation details given; if an article is subsequently reproduced or disseminated not in its entirety but only in part or as a derivative work this must be clearly indicated. For commercial re-use, please contact: [email protected] 48 2878 Nováková et al. polar transport (Tian and Reed, 2001) or by the modulation of cell sensitivity towards auxin (Jones et al., 1991). The data on the effect of light on the metabolism of cytokinins are rather scarce. The assumption of a positive effect of light on CK accumulation has been based on the decreased level of trans-zeatin in phytochrome mutants (Kraepiel et al., 1995). Diurnal variation of CK levels was followed in barley (Kurapov et al., 2000) and carrot (Paasch et al., 1997), but these studies were limited only to the determination of CK bases and ribosides. The recent development of analytical methods (especially of mass spectrometry) has enabled the study of the diurnal variation of a broad array of CK metabolites. Diurnal changes in the levels of 27 CK metabolites were studied in tobacco leaves. The potential contribution of the cleavage of CK O-glucoconjugates to the fast changes of the level of physiologically active CKs (CK bases and ribosides) was estimated via the activity of b-glucosidase (EC 3.2.1.21). In order to study CK catabolism, the activity of the main CK-degrading enzyme, cytokinin oxidase/ dehydrogenase (CKX, EC 1.5.99.12) was determined. As CK levels are also affected by other hormones, the diurnal variation of auxin and ABA levels was followed. USA) and Oasis MCX mixed mode (cation exchange and reversephase) columns (150 mg, Waters, USA) and evaporated. HPLC of radiolabelled trans-zeatin and its metabolic products Evaporated samples were resuspended in 200 mm3 20% (v/v) acetonitrile. Radiolabelled metabolites of [3H]Z were analysed according to Blagoeva et al. (2004) using a Series 200 Quaternary HPLC Pump (Perkin Elmer, Wellesley, MA, USA) coupled to a 235C Diode Array Detector (Perkin Elmer, USA) and a RAMONA 2000 flow-through radioactivity detector (Raytest, Straubenhardt, Germany), column: Luna C18(2), 150 mm/4.6 mm/3 lm (Phenomenex, Torrance, CA, USA). The radioactive metabolites were identified on the basis of coincidence of retention times with authentic standards. HPLC of auxin and abscisic acid Indolyl acetic acid and ABA were determined using two-dimensional HPLC according to Dobrev et al. (2005). The instrumental set-up consisted of a series 200 autosampler (Perkin Elmer, Norwalk, CT, USA), two HPLC gradient pump systems (first pump: ConstaMetric 3500 and 3200 with 500 ll mixer, TSP, Riviera Beach, FL, USA; second pump: Series 200 Quaternary Pump, Perkin Elmer), two columns (first column: ACE-3CN, 15034.6 mm, 3 mm, ACT, Aberdeen, Scotland, UK; second column: Luna C18(2), 15034.6 mm, 3 mm, Phenomenex, Torrance, CA, USA), one 2-position, fluid processor SelectPRO with 1 ml loop (Alltech, Deerfield, IL, USA). The segment containing IAA and ABA obtained in the first dimension was collected in the loop of the fluid processor and redirected to the second HPLC dimension. IAA was quantified using fluorescence detector LC 240 (Perkin Elmer). Quantification of ABA was performed on the basis of UV detection using diode array-detector 235C (Perkin Elmer). Materials and methods Plant material Tobacco plants (Nicotiana tabacum L. cv. Wisconsin 38) were grown in soil in a growth chamber for 6 weeks at a 16 h photoperiod and 130 lmol mÿ2 sÿ1, a day/night temperature of 25/23 8C, and a relative humidity of c. 80%. Mixed samples of leaves (3rd–5th of 7, i.e. with the exception of the youngest and oldest ones) from three plants were collected every hour during a 24 h period. After removal of the main vein the samples were immediately frozen in liquid nitrogen. HPLC/mass spectrometry LC-MS analysis was performed as described by Lexa et al. (2003) using a Rheos 2000 HPLC gradient pump (Flux Instruments, Switzerland) and HTS PAL autosampler (CTC Analytics, Switzerland) coupled to an Ion Trap Mass Spectrometer Finnigan MAT LCQ-MSn equipped with an electrospray interface. Detection and quantification were carried out using a Finnigan LCQ operated in the positive ion, full-scan MS/MS mode using a multilevel calibration graph with deuterated cytokinins as internal standards. The levels of 27 different cytokinin derivatives were measured. The detection limit was calculated for each compound as 3.3 r/S, where r is the standard deviation of the response and S the slope of the calibration curve. Three independent experiments were done. Each sample was injected at least twice. Incubation with tritiated trans-zeatin Tip parts of the 3rd or 4th leaf (c. 1 g FW) were cut at 3 h intervals and their cut surface was submerged into aqueous tritiated trans-zeatin ([3H]Z, 33104 Bq, specific activity 0.6 TBq mmolÿ1, prepared by Dr Jan Hanuš, Isotope Laboratory, Institute of Experimental Botany AS CR, Prague, Czech Republic). After 1 h incubation (under the same light regime as the intact plants) leaf tips were frozen in liquid nitrogen. The uptake of [3H]Z by the vascular tissue was efficient (35–50% of the applied radioactivity). Determination of CKX activity The enzyme preparations were extracted and partially purified using the method of Chatfield and Armstrong (1986) as modified by Motyka et al. (2003). The CKX activity was determined by in vitro assays based on the conversion of [2-3H]isopentenyladenine (prepared by Dr Jan Hanuš) to [3H]adenine. The assay mixture (50 mm3 final volume) included 100 mM TAPS-NaOH buffer containing 75 lM 2,6-dichloroindophenol (pH 8.5), 2 lM substrate ([2-3H] isopentenyladenine, 7.4 Bq molÿ1) and the enzyme preparation equivalent to 250 mg tissue FW (corresponding to 0.02–0.26 mg protein gÿ1 FW). After incubation at 37 8C the reaction was terminated by the addition of 10 mm3 Na4EDTA (200 mM) and 120 mm3 95% (v/v) ethanol. Separation of the substrate from the product of the enzyme reaction was achieved by the same HPLC system as described for tritiated trans-zeatin metabolites according to Blagoeva et al. (2004). The CKX activity was determined in two independent experiments. As the data from both experiments showed Plant hormone extraction and purification Phytohormones were extracted and purified according to Dobrev and Kamı́nek (2002). Frozen leaf samples (c. 1 g FW) were ground in liquid nitrogen and extracted overnight with 10 cm3 methanol/water/ formic acid (15/4/1, by vol., pH;2.5, ÿ20 8C). For analyses of endogenous CKs, 50 pmol of each of the following 12 deuteriumlabelled standards were added: [2H5]Z, [2H5]ZR, [2H5]Z7G, [2H5]Z9G, [2H5]ZOG, [2H5]ZROG, [2H6]iP, [2H6]iPR, [2H6]iP7G, [2H6]iP9G, [2H5]DHZ, [2H5]DHZR (Apex Organics, Honington, UK). Tritiated internal standards were used for the determination of auxin (3[5(n)-3H] indolyl acetic acid, Amersham, UK, specific activity 74 GBq mmolÿ1, 53103 Bq) and ABA (Amersham, UK, specific activity 1.74 TBq mmolÿ1, 53103 Bq). The extracts were purified using Si-C18 columns (SepPak Plus, Waters, Milford, MC, 49 Cytokinin, auxin, and ABA diurnal rhythms 2879 the same tendencies (but different absolute values), the results of one representative experiment (mean value of three parallel assays for each of three replicate protein preparations) are presented. Determination of b-glucosidase activity Tobacco leaves (0.2 g FW) were ground in liquid nitrogen and transferred into 0.4 cm3 ice-cold extraction buffer [25 mM TRIS-Cl, pH 7.5, 5 mM MgCl2, 1 mM EDTA, and protease Complete EDTAfree cocktail tablet (Roche)]. Samples were centrifuged at 13 000 g for 15 min (4 8C) and 0.02 cm3 supernatant were added into 0.28 cm3 citrate-phosphate buffer (0.02 M citric acid; 0.06 M NaH2PO4; pH 5.5). The reaction was initiated by 0.1 cm3 20 mM p-nitrophenyl a-D-glucopyranoside. After 120 min at 30 8C the reaction was stopped by adding 0.6 cm3 2 M Na2CO3. The released p-nitrophenol was measured at 405 nm (UNICAM 5625 UV/VIS spectrometer). Results Cytokinins Physiologically active cytokinins, i.e. trans-zeatin, isopentenyladenine, dihydrozeatin, and their corresponding ribosides, exhibited in tobacco leaves a sharp maximum after 9 h of light, i.e. 1 h after the middle of the light period (Fig. 1). Another, considerably lower, peak was detected after the transition to darkness and the lowest one was found at the beginning of the light phase. The content of cis-zeatin and its riboside exhibited a minor peak coinciding with the maximum of active CKs. Their maximum followed the second peak of active CKs after the light/ dark transition (with 2 h delay). The content of total CKs (the sum of physiologically active CKs together with CK phosphates, cis-zeatin derivatives, N- and O-glucosides) have an increasing trend during the light period, up to a maximum 9 h after the light was switched on (i.e. 1 h after midday). The second half of the light period was characterized by a sharp decrease in CKs. After the light/dark transition another transient CK peak occurred. During the course of the dark period, the total CK level gradually declined. (The levels of individual CKs are given in a table as supplementary material that can be found at JXB online.) The concentration of the CK deactivation metabolites (N-glucosides) increased at the end of the dark period and remained at an elevated level up to 6 h of light (Fig. 2). The main maximum of CK N-glucosides coincided with (or closely followed) the peak of active CKs after the middle of the light period. Then, a sharp decrease took place. The second main peak, nearly as large as the first one, was found after the light was switched off, coinciding with a minor peak of physiologically active cytokinins. The level of cis-zeatin derivatives followed a trend similar to the N-glucosides, with the exception of the second maximum after the light/dark transition, which was, in the case of cis-zeatin derivatives, delayed and much less profound. The levels of CK phosphates exhibited a transient peak around 4 h of light. Their maximum was found after the middle of the light period. Fig. 1. Endogenous levels of total cytokinins, physiologically active cytokinins (trans-zeatin, isopentenyladenine, dihydrozeatin, and the corresponding ribosides) and cis-zeatin (riboside) in tobacco leaves during a 24 h period. The photoperiod is represented by the black and white box on the abscissa. Means 6SE of three independent experiments are shown. Fig. 2. Endogenous levels of CK N-glucosides, CK phosphates, and ciszeatin derivates in tobacco leaves during a 24 h period. Other details as for Fig. 1. Cytokinin O-glucosides and activity of b-glucosidase The level of CK storage forms (O-glucosides) was found to increase around the dark/light transition. Another peak of O-glucosides was detected after 7 h of light, preceding the peak of active CKs as well as that of CK deactivation metabolites. However, this CK O-glucoside peak was significantly lower than those of the deactivation products. Minor peaks of CK O-glucosides were detected after 10 h of light, when the level of active CKs was decreasing, and 2 h after the light/dark transition. The activity of b-glucosidase, the enzyme that catalyses the cleavage of glucosyl moieties from O-glucosides releasing, in the case of CK O-glucosides, active CKs, exhibited the first maximum after 1 h of light, which correlated with the decrease of the peak of CK O-glucosides found around the dark/light transition (Fig. 3). The major peak of b-glucosidase activity was found 7 h after the light was switched on, preceding the midday peak of active CKs. 50 2880 Nováková et al. Cytokinin oxidase/dehydrogenase adenosine, products of CK breakdown, was found close to the dark/light transition and a minor one around 9 h of light, i.e. at the periods preceding the detected maxima of CKX activity. The peak of non-metabolized trans-zeatin, corresponding to the period of the diminished trans-zeatin breakdown, preceded the midday peak of active CKs. The maximum of the non-metabolized labelled trans-zeatin correlated with the minimum level of the ribosylated and phosphoribosylated forms. A comparison of the diurnal variations in CKX activity and in the levels of CKs, which are the substrates of this enzyme, i.e. isopentenyladenine, trans-zeatin, and their corresponding ribosides (but not dihydrozeatin or its derivatives), is shown in Fig. 4. The main CKX maximum was found 3 h after the transition from dark to light. The second maximum followed the midday peak of active CKs. The third, least profound peak of CKX activity accompanied the minor peak of CKs after the light/dark transition (with a 1 h delay). Indolyl acetic acid and abscisic acid The midday CK peak coincided with the maximum level of indolyl acetic acid and was accompanied (or closely followed) by a minor peak of ABA (Fig. 6). No other profound peak of indolyl acetic acid was detected during the diurnal cycle. At the end of the light period, a gradual increase in ABA level was found. This enhancement was accelerated after the light was switched off, the maximum Metabolic conversion of radiolabelled trans-zeatin Diurnal changes in the levels of endogenous CKs and the activity of CK metabolic enzymes were also evaluated following the metabolic conversion of radiolabelled trans-zeatin (Fig. 5). A high peak of labelled adenine and Fig. 3. Endogenous levels of CK O-glucosides and activity of b-glucosidase in tobacco leaves during a 24 h period. Other details as for Fig. 1. Fig. 5. Metabolism of radiolabelled trans-zeatin in tobacco leaves during a 24 h period (incubation time 1 h). Other details as for Fig. 1. Fig. 4. The activity of cytokinin oxidase/dehydrogenase (CKX) and endogenous levels of cytokinins, which are the substrates of this enzyme (trans-zeatin, isopentenyladenine, and the corresponding ribosides), in tobacco leaves during a 24 h period. Other details as for Fig. 1. Fig. 6. Endogenous levels of indole-3-acetic acid (IAA), abscisic acid (ABA), and active cytokinins in tobacco leaves during a 24 h period. Other details as for Fig. 1. 51 Cytokinin, auxin, and ABA diurnal rhythms being reached after 3 h of dark. The high level of ABA was only transient. A low, sharp peak of ABA was also detected immediately after the light was switched on. Discussion Timing of the main CK peak in tobacco leaves shortly (1 h) after the middle of the light period is in accordance with the diurnal profile in barley leaves (main peaks of CK bases and ribosides around 14.00 h; Kurapov et al., 2000) and slightly differs from that in carrot plants kept under continuous light (maximum of CK bases and ribosides around 16.00 h; Paasch et al., 1997). As both CKs and auxin are necessary for the progression of cell division, this finding of the coincidence of the main CK peak with the maximum of IAA indicates that cell division may be initiated at this time. The close connection between CK increase and growth stimulation was reported in barley (at the stage of the third leaf), when the CK peak at 14.00 h was shown to precede the period of maximal growth from 15.00 h to 18.00 h (Kurapov et al., 2000). The minimal growth was determined before dawn (at 04.00–05.00 h). A correlation between growth and the hormone level was also found in the case of auxin, the link between the oscillation in the level of IAA in the rosette leaves and the stem growth rate in Arabidopsis was reported by Jouve et al. (1998, 1999). The results on the midday peak of IAA are in accordance with the data of Pavlova and Krekule (1984), who reported a peak of IAA content after the middle of the light phase in Chenopodium rubrum plants grown under very long day (20 h). In shorter periods (16 h, 12 h, 8 h), connected with flower induction, this peak was shifted towards the dark phase. It should be kept in mind that these results reflect the situation in still developing but not the youngest tobacco leaves, in other tissues and/or species hormones may exhibit a different diurnal pattern. The effect of developmental stage was observed in barley. Leaves at the stage of the third leaf had the main CK peak at 14.00 h, while at the stage of tillering the peak was shifted to 17.00 h (Kurapov et al., 2000). Another point is that the experiments were aimed at the determination of the diurnal profile of isoprenoid CKs. Due to the differences in physiological function, diurnal rhythms of isoprenoid CKs are likely to differ from those of their aromatic counterparts. The peaks of aromatic CKs m-methoxytopolin and its riboside were detected at 09.30 h in poplar leaves (Tarkowska et al., 2003), which corresponded approximately to 3 h after the light period initiation, i.e. 6 h earlier than for isoprenoid CKs in tobacco. The finding of the ABA maximum at the beginning of the dark phase is in accordance with the survey of diurnal changes in ABA metabolism summarized by Tallman (2004), 2881 who reported a stimulation of ABA biosynthesis at dark, in order to maintain the stomata in the closed position during the night to ensure maximal tissue rehydration. The sharp decrease in ABA concentration, which was found shortly after the beginning of the light period, seems to be connected with the depletion of endogenous ABA and the activation of its catabolism. According to Tallman (2004), after irradiation ABA levels drop due to the activation of the xanthophyll cycle (conversion of ABA precursors to zeaxanthin), isomerization into physiologically inactive trans-forms, as well as to ABA degradation by cytochrome P450 mono-oxygenase. Phytochrome-mediated activation of ABA degradation was also reported by Kraepiel et al. (1994). The decrease in ABA concentration in the guard cells and surrounding apoplast in the morning enables stomatal opening, which positively affects plant photosynthesis. Later, at the midday phase, elevated apoplastic ABA around guard cells, delivered by transpiration, is necessary for the regulation of the stomata aperture in order to balance the rate of photosynthesis and plant turgor maintenance (Tallman, 2004). The minor peak of ABA, which was found 1 h after the middle of the light phase, might be related to this ‘after-midday’ regulation or, more probably, may serve to counterbalance the midday peak of CKs, as under temperate conditions stomatal aperture is regulated by the CK/ABA ratio (Fusseder et al., 1992). Fusseder et al. (1992) reported that the increase in xylem-delivered CKs preceded the increase in leaf conductance. They found the peak of xylem CKs in the morning and their rapid decrease in the afternoon in desert-grown almond trees. Beck and Wagner (1994) suggested that the high concentration of xylem CKs in the morning originated from the steady rate of CK biosynthesis in the roots and of their exudation into the xylem sap during the night when the transpiration rate is low. In the morning, the flow of xylem sap sharply increases, delivering accumulated CKs throughout the plant. Due to the increased volume of the transpiration stream during the day and the steady rate of CK biosynthesis in roots, CK concentration in the xylem sap gradually decreases. The importance of the velocity of the transpiration stream on the translocation of CKs produced in roots to other parts of plant was recently demonstrated by Aloni et al. (2005). In our experiments xylem transport of CKs was not measured. Indirect evidence of the existence of such translocation may be given by the position of the main maximum of CKX activity that was detected early in the light period. The other peaks of CKX activity (after midday and following the light/dark transition) correlated well with the decrease of endogenous CKs in leaf blades, indicating that CK breakdown by CKX is involved in the down-regulation of these endogenous CK peaks. However, the main maximum of CKX activity, which appeared early in the light phase, did not coincide with substantial changes of endogenous CK levels in leaf 52 2882 Nováková et al. Supplementary data blades, suggesting that this CKX peak may serve in the regulation of xylem CKs. To evaluate the role of CK reversible conjugation in the regulation of their levels, the diurnal rhythm of bglucosidase activity was followed. In the first hours of the light period, the increase in b-glucosidase activity created conditions for the cleavage of CK O-glucosides (one of the CK transport forms) necessary for their further degradation with CKX. The b-glucosidase increase could, of course, also result in hydrolysis of other unidentified compounds, for example, ABA-glucosylester. However, at this time-point, when the level of free ABA was decreasing (Tallman, 2004), the function of b-glucosidase activity stimulation did not seem to be ABA-glucoside deconjugation. The occurrence of b-glucosidase peak in the period before and after the main peak of CKs indicates that hydrolysis of CK O-glucosides may be involved in the finetuning of active CK levels at midday, even if this deglucosylation cannot be the only source of the midday peak of physiologically active CKs. As CK translocation from the roots, which is the primary site of CK biosynthesis (Letham, 1963; Aloni et al., 2005), was reported to be high in the morning, declining rapidly during the day (Beck and Wagner 1994), CK biosynthesis seems to be a more probable source of the sharp midday increase of CKs. As only free ABA was determined, hardly any conclusion may be drawn about the role of b-glucosidase in the regulation of free ABA levels at the ‘after-midday’ peak. This will be the subject of further research. Determination of the bulk leaf ABA content represents another methodological limitation, as for regulation of stomata aperture ABA concentration in the guard cells and the surrounding apoplast is crucial, and the bulk ABA concentration includes the ABA stored in mesophyll chloroplasts as well. The physiological significance of CK peaks may be related, apart from the stimulation of cell division in leaves or the regulation of stomatal aperture, to the generation of a potential CK signal to other plant parts (e.g. flower induction, Macháčková et al., 1996) or the induction/ repression of CK responsive genes (Hare and van Staden, 1997). These genes usually display circadian oscillations in their levels of abundance with maximal expression during the light period (Thomas et al., 1997). The circadian clock represents an important adaptation of life to the conditions on the earth and provides the necessary external and internal co-ordination of the metabolic processes. The results here have shown that CKs, auxin, and ABA levels undergo significant variations in leaves during the day and night. These data provide, to our knowledge, the first report of a diurnal rhythm of CK metabolic enzymes and CK metabolites other than CK bases and ribosides. The coincidence of the peak of CKs and auxin indicates that the changes in plant hormone levels have physiological relevance and correlate with the diurnal changes of plant growth and development. Supplementary data for this paper can be found at JXB online. Acknowledgements The authors wish to acknowledge Eva Kobzová and Marie Korecká for skilful technical assistance. This work was supported by grants from the Grant Agency of the Czech Republic, No. 206/03/0369, and from the Ministry of Education, Youth and Sports, LN LN00A081. References Aloni R, Langhans M, Aloni E, Dreieicher E, Ullrich CI. 2005. Root-synthesized cytokinin in Arabidopsis is distributed in the shoot by the transpiration stream. Journal of Experimental Botany 56, 1535–1544. Beck E, Wagner BM. 1994. Quantification of the daily cytokinin transport from the root to the shoot of Urtica dioica L. Botanica Acta 107, 342–348. Blagoeva E, Dobrev PI, Malbeck J, Motyka V, Strnad M, Hanuš J, Vanková R. 2004. Cytokinin N-glucosylation inhibitors suppress deactivation of exogenous cytokinins in radish, but their effect on active endogenous cytokinins is counteracted by other regulatory mechanisms. Physiologia Plantarum 121, 215–222. Chatfield JM, Armstrong DJ. 1986. Regulation of cytokinin oxidase activity in callus tissues of Phaseolus vulgaris L. cv. Great Northern. Plant Physiology 80, 493–499. Dobrev PI, Havlı́ček L, Vágner M, Malbeck J, Kamı́nek M. 2005. Purification and determination of plant hormones auxin and abscisic acid using solid phase extraction and two-dimensional high performance liquid chromatography. Journal of Chromatography A 1075, 159–166. Dobrev P, Kamı́nek M. 2002. Fast and efficient separation of cytokinins from auxin and abscisic acid and their purification using mixed-mode solid-phase extraction. Journal of Chromatography A 950, 21–29. Foster KR, Morgan PW. 1995. Genetic-regulation of development in Sorghum bicolor. 9. The Ma(3)(R) allele disrupts diurnal control of gibberellin biosynthesis. Plant Physiology 108, 337–343. Fusseder A, Wartinger A, Hartung W, Schulze ED, Heilmeier H. 1992. Cytokinins in the xylem sap of desert-grown almond (Prunus dulcis) trees: daily courses and their possible interactions with abcisic acid and leaf conductance. New Phytologist 122, 45–52. Hare PD, Van Staden J. 1997. The molecular basis of cytokinin action. Plant Growth Regulation 23, 41–78. Jasoni RL, Cothren JT, Morgan PW, Sohan DE. 2002. Circadian ethylene production in cotton. Plant Growth Regulation 36, 127–133. Jones AM, Cochran DS, Lamerson PM, Evans ML, Cohen JD. 1991. Red light-regulated growth. 1. Changes in the abundance of indole acetic-acid and 22-kilodalton auxin-binding protein in the maize mesocotyl. Plant Physiology 97, 352–358. Jouve L, Gaspar T, Kevers C, Greppin H, Agosti RD. 1999. Involvement of indole-3-acetic acid in the circadian growth of the first internode of Arabidopsis. Planta 209, 136–142. Jouve L, Greppin H, Agosti RD. 1998. Arabidopsis thaliana floral stem elongation: evidence for an endogenous circadian rhythm. Plant Physiology and Biochemistry 36, 469–472. Kraepiel Y, Marrec K, Sotta B, Caboche M, Miginiac E. 1995. In vitro morphogenic characteristic of phytochrome mutants in 53 Cytokinin, auxin, and ABA diurnal rhythms Nicotiana plumbaginifolia are modified and correlated to high indole-3-acetic acid levels. Planta 197, 142–146. Kraepiel Y, Rousselin P, Sotta B, Kerhoas L, Einhorn J, Caboche M, Miginiac E. 1994. Analysis of phytochrome- and ABAdeficient mutants suggests that ABA degradation is controlled by light in Nicotiana plumbaginifolia. The Plant Journal 6, 665–672. Kurapov PB, Skorobogatova IV, Sal’nikova EI, Sorkina GL, Siusheva AG. 2000. Diurnal courses of endogenous phytohormones in barley. Izvestiya Akademii Nauk Seriya Biologicheskaya 1, 108–114. Letham DS. 1963. Zeatin, a factor inducing cell division from Zea mays. Life Science 8, 569–573. Lexa M, Genkov T, Malbeck J, Macháčková I, Brzohohatý B. 2003. Dynamics of endogenous cytokinins pool in tobacco seedlings: a modelling approach. Annals of Botany 91, 585–597. Macháčková I, Eder J, Motyka V, Hanuš J, Krekule J. 1996. Photoperiodic control of cytokinin transport and metabolism in Chenopodium rubrum. Physiologia Plantarum 98, 564–570. Motyka V, Vaňková R, Čapková V, Petrášek J, Kamı́nek M, Schmülling T. 2003. Cytokinin-induced upregulation of cytokinin oxidase activity in tobacco includes changes in enzyme glycosylation and secretion. Physiologia Plantarum 117, 11–21. 2883 Paasch K, Lein C, Nessiem M, Neumannn HK. 1997. Changes in the concentration of some phytochromes in cultured root explants and in intact carrot plants (Daucus carota L.) during the day. Angewandte Botanik 71, 85–89. Pavlová L, Krekule J. 1984. Fluctuation of free IAA under inductive and non-inductive photoperiods in Chenopodium rubrum. Plant Growth Regulation 2, 91–98. Tallman G. 2004. Are diurnal patterns of stomatal movement the result of alternating metabolism of endogenous guard cell ABA and accumulation of ABA delivered to the apoplast around guard cells by transpiration? Journal of Experimental Botany 55, 1963–1976. Tarkowská D, Doležal K, Tarkowski P, Astot C, Holub J, Fuksová K, Schmülling T, Sandberg G, Strnad M. 2003. Identification of new aromatic cytokinins in Arabidopsis thaliana and Populus3canadensis leaves by LC-(+)ESI-MS and capillary liquid chromatography/frit-fast atom bombardment mass spectrometry. Physiologia Plantarum 117, 579–590. Thomas TH, Hare PD, van Staden J. 1997. Phytochrome and cytokinin responses. Plant Growth Regulation 23, 105–122. Tian Q, Reed JW. 2001. Molecular links between light and auxin signaling pathways. Journal of Plant Growth Regulation 20, 274–280. 54 Journal of Experimental Botany, Vol. 57, No. 6, pp. 1413–1421, 2006 doi:10.1093/jxb/erj121 Advance Access publication 23 March, 2006 This paper is available online free of all access charges (see http://jxb.oxfordjournals.org/open_access.html for further details) RESEARCH PAPER Diurnal changes in polyamine content, arginine and ornithine decarboxylase, and diamine oxidase in tobacco leaves Lenka Gemperlová1,2, Marie Nováková1,3, Radomı́ra Vaňková1, Josef Eder1 and Milena Cvikrová1,* 1 Institute of Experimental Botany, Academy of Sciences of Czech Republic, Rozvojová 135, 165 02 Prague 6, Czech Republic 2 Department of Plant Physiology, Faculty of Sciences, Charles University, Viničná 5, 128 44 Prague 2, Czech Republic 3 Department of Biochemistry, Faculty of Sciences, Charles University, Hlavova 2030/8, 128 44 Prague 2, Czech Republic Received 29 August 2005; Accepted 13 January 2006 Abstract Changes in the contents of polyamines (PAs) in tobacco leaves (Nicotiana tabacum L. cv. Wisconsin 38) grown under 16 h photoperiod were correlated with arginine and ornithine decarboxylase (EC 4.1.1.19 and EC 4.1.1.17) and diamine oxidase (EC 1.4.3.6) activities. The maximum of free and soluble conjugated forms of PAs occurred 1–2 h after the middle of the light period and was followed by two distinct peaks at the end of the light and at the beginning of the dark phase. Putrescine was the most abundant and cadaverine the least abundant PA in both free and PCA-soluble forms. However, cadaverine was predominant in PCAinsoluble conjugates, followed by putrescine, spermidine, and spermine. Both arginine and ornithine decarboxylases are involved in putrescine biosynthesis in tobacco leaves. Light dramatically stimulated the activity of ornithine decarboxylase, while no photoinduction of arginine decarboxylase activity was observed. Ornithine decarboxylase was found mainly in the particulate fraction. Only one peak, just after light induction, occurred in the cytosolic fraction, with 35% of the total ornithine decarboxylase activity. By contrast, the total arginine decarboxylase activity was equally divided between the soluble and pellet fractions. A sharp increase in diamine oxidase activity occurred 1 h after exposure to light, concomitant with the light-induced increase in ornithine decarboxylase activity. After a decline, diamine oxidase activity increased again, together with the rise in the amount of free Put. The roles of both conjugation of PAs with hydroxycinnamic acids and oxidative degradation of putrescine in maintaining free PA levels during the 24 h light/dark cycle are discussed. The presented results have shown that the parameters studied here followed rhythmical changes and were not only affected by light. Key words: Arginine decarboxylase, diamine oxidase, ornithine decarboxylase, polyamines, tobacco. Introduction Polyamines (PAs) are ubiquitous organic polycations that are involved in the regulation of different cellular processes in animals and plants. The biological activity of PAs is attributed to the cationic nature of these molecules, which enables their interactions with anionic biomolecules (e.g. DNA, RNA, proteins, and phospholipids). In plant cells, PAs are usually present in amounts varying from micromolar to more than millimolar, i.e. levels significantly higher than those of plant hormones. Nevertheless, PAs affect many processes regulated by cytokinins and auxins * To whom correspondence should be addressed. E-mail: [email protected] Abbreviations: ADC, arginine decarboxylase; Cad, cadaverine; DAO, diamine oxidase; EDTA, ethylenediaminetetraacetic acid; GABA, c-aminobutyric acid; ODC, ornithine decarboxylase; PA, polyamine; PAL, phenylalanine ammonia-lyase; PCA, perchloric acid; Put, putrescine; SAMDC, S-adenosylmethionine decarboxylase; Spd, spermidine; Spm, spermine. ª The Author [2006]. Published by Oxford University Press [on behalf of the Society for Experimental Biology]. All rights reserved. The online version of this article has been published under an Open Access model. Users are entitled to use, reproduce, disseminate, or display the Open Access version of this article for non-commercial purposes provided that: the original authorship is properly and fully attributed; the Journal and the Society for Experimental Biology are attributed as the original place of publication with the correct citation details given; if an article is subsequently reproduced or disseminated not in its entirety but only in part or as a derivative work this must be clearly indicated. For commercial re-use, please contact: [email protected] Downloaded from http://jxb.oxfordjournals.org at Department of Plant Physiology, Faculty of Science, Charles University on 25 March 2009 55 1414 Gemperlová et al. described in leaves of Pharbitis nil (Yoshida and Hirasawa, 1998). Transcription of the SAMDC gene was shown to be under circadian control (Yoshida et al., 1999). Changes in PA content, ADC and ODC activities during light/ dark phases in maize calluses did not only respond to light, but they were also found to be regulated by a daily rhythm (Bernet et al., 1999). Photoperiodic control of PA accumulation was observed in petal explants of Araujia sericifera (Moysset et al., 2002). The purpose of the current work was to identify the daily variation of PA levels in tobacco plants in order to find the optimal time for sample collection. These data will be used in further studies of the role of PAs in abiotic stress response and PA cross-talk with cytokinins using transgenic tobacco plants over-expressing cytokinin metabolic enzymes under different (constitutive as well as inducible) promoters. Moreover, determination of diurnal variation of the levels of individual PAs might contribute to the understanding of mechanisms involved in the regulation of physiologically active PAs. Data presented here show detailed changes in the contents of free Put, Spd, Spm, and Cad, their soluble and insoluble conjugates in tobacco leaves grown under a long day. The focus was on the evaluation of the effect of light on ADC and ODC activities, as well as on the elucidation of the potential correlation of ADC and/or ODC activities and Put and, subsequently, Spd and Spm accumulation. The role of PA conjugation and oxidative deamination in PA homeostasis during the 24 h cycle was followed, too. and, in co-operation with these plant hormones, modulate various morphogenic processes (Altamura et al., 1993). In plants, PAs participate in the control of such important processes as cell division and growth (Theiss et al., 2002) morphogenesis and differentiation (Paschalidis et al., 2001), programmed cell death (Serafini-Fracassini et al., 2002; Pignatti et al., 2004), stabilization of nucleic acids and membranes (Thomas and Thomas, 2001), and stress responses (Bouchereau et al., 1999). The polyamine content in cells can be regulated by both metabolic and transport processes. In higher plants PAs are biosynthesized by two main pathways. The diamine putrescine (Put) may be formed directly from ornithine by ornithine decarboxylase (ODC; EC 4.1.1.17) or indirectly from arginine by arginine decarboxylase (ADC; EC 4.1.1.19). S-adenosylmethionine decarboxylase (SAMDC; EC 4.1.1.50) is essential for spermidine (Spd) and spermine (Spm) biosynthesis. Spd is formed from Put as well as Spm from Spd by successive addition of aminopropyl groups derived from decarboxylated S-adenosylmethionine generated by the activity of SAMDC (Wallace et al., 2003). A clear relationship between the diamine cadaverine (Cad) and other PAs has not yet been found (Bagni and Tassoni, 2001). The intracellular concentrations of free PAs may be regulated by the rate of biosynthesis, which is controlled at the transcriptional, post-transcriptional, translational, and post-translational levels (Kumar et al., 1997). Apart from biosynthesis, PA levels may be modulated by conjugation either with small molecules, especially hydroxycinnamic acids (soluble conjugated PAs; Martin-Tanguy, 1985; Bagni and Tassoni, 2001), or with high molecular-mass substances like hemicelluloses, lignin, and, to a lesser extent, proteins as well (insoluble conjugated PAs; Creus et al., 1991). Cytoplasmic levels of PAs are also affected via storage in the vacuoles, mitochondria, and chloroplasts and by extracellular transport (Flores, 1991; Beranger-Novat et al., 1997; Bagni and Tassoni, 2001). The mechanisms which regulate polyamine transport in eukaryotes remain largely unknown. Only translocation of free polyamines has been reported (Antognoni et al., 1998). Another important process of PA level regulation is degradation through oxidative deamination. Two catabolic enzymes were found in plants. Diamine oxidase (DAO; EC 1.4.3.6) which primarily catalyses the conversion of Put to ammonia, H2O2 and pyrroline, and polyamine oxidase (PAO; EC 1.5.3.) with high affinity towards Spd and Spm. Data on the relationship between the physiological effects of PAs and light are not very extensive. Alterations in the contents and forms of PAs during photoperiodic flower induction were observed in soybean (Caffaro and Vicente, 1994). Photosynthetically active radiation caused Put and Spd accumulation in soybean (Kramer et al., 1992), which is consistent with the hypothesis that phytochrome can regulate the activity of ADC in chloroplasts (Besford et al., 1993). Photo-induction of ADC was also Materials and methods Plant material Tobacco (Nicotiana tabacum L. cv. Wisconsin 38) plants were grown in soil in a growth chamber for 6 weeks at a 16/8 h photoperiod (130 lmol mÿ2 sÿ1) with a day/night temperature of 25/23 8C and relative humidity c. 80%. Mixed samples of leaves (3rd–5th of 7, i.e. with the exception of the youngest and the oldest ones) from three plants were pooled and collected every h during a 24 h period. The sampling during the dark period was realized in dim light and did not exceed 1 min. The samples were immediately frozen in liquid nitrogen. Polyamine analysis The leaves were ground in liquid nitrogen and extracted overnight at 4 8C with 5% (v/v) perchloric acid (PCA) (100 mg fresh weight tissue cmÿ3 5% PCA). 1,7-Diaminoheptane was added as an internal standard. The extracts were centrifuged at 21 000 g for 15 min, and then PCA-soluble free PAs were analysed in one-half of the supernatant. The remaining supernatant and pellet were acid hydrolysed in 6 M HCl for 18 h at 110 8C to obtain PCA-soluble and PCA-insoluble conjugates of PAs as described by Slocum et al. (1989). Standards (Sigma-Aldrich, St Louis, MO, USA), PCAsoluble free PAs, and acid hydrolysed PA conjugates were benzoylated. HPLC analysis of benzoyl-amines was performed on a Beckman-Video Liquid Chromatograph equipped with a UV detector (detection at 254 nm) and C18 Spherisorb 5 ODS2 column (particle size 5 lm, column length 25034.6 mm) according to the method of Slocum et al. (1989). Downloaded from http://jxb.oxfordjournals.org at Department of Plant Physiology, Faculty of Science, Charles University on 25 March 2009 56 Diurnal changes in polyamine content 1415 Statistical analyses Means 6SE of two independent experiments with two replicates are shown in the figures. Statistical tests were analysed using the Student’s t distribution criteria. Results Polyamine contents Free and bound PAs (PCA-soluble and PCA-insoluble conjugates) were analysed in tobacco leaves each h during a 24 h light/dark cycle. All four PAs (Put, Spd, Spm, and Cad) were present in both free and conjugated forms. Although the content of PAs is usually expressed on a fresh weight basis, it was decided to express the PA contents in nmol gÿ1 protein as well (Fig. 1) because the focus of this work was on the evaluation of the potential correlation between biosynthetic enzyme activities and the contents of PAs. Differences between the graphs are related to changes in protein in the course of the light/dark cycle, which were higher during the light, especially in the first half of the light period. As a consequence, the peaks of PAs per fresh weight at the end of the light and at the beginning of the dark phase (Fig. 1A) became more pronounced when expressed per protein (Fig. 1B). Total contents of PAs (nmol mg-1 protein) Diamine oxidase assay Diamine oxidase (DAO, EC 1.4.3.6) activity was assayed by a spectrophotometric method based on detection of the aldehyde with cis-1,4-diamino-2-butene as the substrate (Peč et al., 1991). Samples were homogenized in 0.1 M TRIS–HCl buffer, pH 8.5, containing 2 mM mercaptoethanol and 1 mM EDTA, and centrifuged at 20 000 g for 15 min at 4 8C. The reaction mixture contained 0.1 M TRIS–HCl buffer, pH 8.5, catalase (25 lg) and 0.01 M cis-1,4-diamino-2-butene. The reaction was started by the addition of 0.2 cm3 supernatant, incubated for 1 h at 37 8C and stopped by adding 1 cm3 of Ehrlich’s reagent. The reaction mixture was incubated at 50 8C for 5 min, and then chilled in an ice bath before reading the absorbance of produced pyrrol at 563 nm. Enzymatic activity is expressed in pkat mgÿ1 protein. Protein content was measured according to Bradford’s method using bovine serum albumin as a standard (Bradford, 1976). Total contents of PAs (nmol g-1 FW) Ornithine decarboxylase and arginine decarboxylase assays Ornithine decarboxylase (ODC; EC 4.1.1.17) and arginine decarboxylase (ADC; EC 4.1.1.19) were determined by a radiochemical method described by Tassoni et al. (2000). Samples were extracted in 3 vols of ice-cold 0.1 M TRIS–HCl buffer, pH 8.5, containing 2 mM b-mercaptoethanol, 1 mM EDTA and 0.1 mM pyridoxal phosphate, and centrifuged at 20 000 g for 30 min at 4 8C. Aliquots (0.1 cm3) of both supernatant (soluble fraction) and resuspended pellet (particulate fraction) were used to determine ODC and ADC activity. Enzyme activity assays were performed by measuring the 14CO2 evolution from 7.4 kBq L-[1-14C]ornithine (1.92 GBq mmolÿ1, Amersham Pharmacia Biotech UK) or 7.4 kBq L-[U-14C]arginine (11.5 GBq mmolÿ1 Amersham Pharmacia Biotech UK), for ODC and ADC, respectively, in the presence of 2 mM unlabelled substrate during a 1.5 h incubation at 37 8C. CO2 was trapped in hyamine hydroxide and the radioactivity was counted by a liquid scintillation analyser (Tri-Carb 2900TR; Packard). 2500 2000 Put Cad Spd Spm A Put Cad Spd Spm B 1500 1000 500 0 500 400 300 200 100 0 0 4 8 12 16 20 Time (h) Fig. 1. Changes in total contents of Put, Spd, Spm, and Cad (represented by the sum of free, PCA-soluble and insoluble conjugates) in tobacco leaves grown under a 16/8 h photoperiod. (A) PA contents expressed as nmol gÿ1 FW; (B) PA contents expressed as nmol mgÿ1 protein. The photoperiod is represented by the black and white boxes on the abscissa. Means 6SE of two independent experiments with two replicates are shown. The ratio among PA forms changed during the 24 h cycle (Fig. 2). The most abundant PA in free and PCA-soluble form was Put and the least abundant was Cad. However, Cad prevailed in PCA-insoluble conjugates, followed by Put, Spm, and Spd (Fig. 3). The endogenous level of free Put showed a transient increase at the second and the third hours of the light period (P <0.05 compared with the value at the end of the dark period) and exhibited a maximum after 9 h of light (P <0.005 compared with the value at 7 h). The largest peak occurred at the light/dark transition (P <0.005 compared with the value at 11 h; Fig. 2A). The contents of PCA-soluble Put conjugates showed a first peak 1 h after the middle of the light period (P <0.005 compared to the value at 7 h), and further transient peaks were detected 2 h before the light was switched off and at the beginning of the dark period (statistical significance of both peaks P <0.005 compared with the value at 13 h). A slight rise was observed at the end of the dark phase (Fig. 2A). The peaks in the free Spd content were less marked than those of Put (statistical significance of the increase at 11 h Downloaded from http://jxb.oxfordjournals.org at Department of Plant Physiology, Faculty of Science, Charles University on 25 March 2009 57 1416 Gemperlová et al. F1 F2 200 A Cad (nmol mg-1 protein) Put (nmol mg-1 protein) 250 200 150 100 50 0 120 F1 F2 B 100 80 60 40 20 Cad (nmol mg-1 protein) Spm (nmol mg-1 protein) 0 50 40 150 100 50 0 Put, Spd, Spm (nmol mg-1 protein) Spd (nmol mg-1 protein) 140 A C F1 F2 30 B Put Spd Spm 20 10 0 0 4 8 12 16 20 Time (h) 30 Fig. 3. Changes in PCA-insoluble conjugates of Cad (A) and Put, Spd, Spm (B) in tobacco leaves grown under a 16/8 h photoperiod. Other details as for Fig. 2. 20 10 peaks, the contents of soluble conjugates of Put and Spd were significantly higher than those of their free forms (PCA-soluble forms represented about 75% of the total content at the midday maximum). The contents of free Spm and Cad and of their soluble conjugates did not change significantly during the diurnal cycle, except that of Spm conjugates, which increased 5 h before the end of the light (P <0.05 compared with the value at 10 h of light) and at the beginning of the dark period (P < 0.005 compared to the value at 10 h of light; Fig. 2C, D). The time-course of the levels of PCA-insoluble conjugates showed different trends from those observed in free and soluble forms (Fig. 3). Three dominant peaks of insoluble Cad conjugates were detected in tobacco leaves grown under the 16/8 h photoperiod. After the decline at the dark/light transition the first maximum occurred at the beginning (at the second and third hours) of the light period (P <0.005 compared with the value at 1 h of light), the second one 4 h after midday (P <0.005 compared with the value at 10 h of light), and the third one at the beginning of the dark period (2 h after the light was switched off, P <0.005 compared with the value at the light/dark transition). Insoluble conjugates of Put, after the transient 0 15 D F1 F2 10 5 0 0 4 8 12 16 20 Time (h) Fig. 2. Changes in free and PCA-soluble conjugates of PAs in tobacco leaves grown under 16/8 h photoperiod. (A) Put contents; (B) Spd contents; (C) Spm contents; (D) Cad contents. F1, free PAs; F2, PCAsoluble conjugates. The photoperiod is represented by the black and white boxes on the abscissa. Means 6SE of two independent experiments with two replicates are shown. For further details, see text. of light was P <0.05 compared with the value at 7 h, which was similar to the peaks at the light/dark transition compared with 13 h). PCA-soluble conjugates of Spd showed a similar trend to those of Put (Fig. 2B). In the Downloaded from http://jxb.oxfordjournals.org at Department of Plant Physiology, Faculty of Science, Charles University on 25 March 2009 58 Diurnal changes in polyamine content 1417 decline at the dark/light transition, reached their high value in the middle of the light period (P <0.005 compared with the value at 4 h of light), and then decreased. Their second peak coincided with that of Cad (2 h after the light was switched off, P <0.005 compared with the value at 14 h of light). Relatively high level of insoluble conjugates of Cad and Put were observed during the dark period. However, the amounts of insoluble Put conjugates were much lower than those of Cad. The content of insoluble Spd conjugates was the highest at the first third of the light period. The contents of Spm conjugates followed a trend similar to Cad but the peaks were much less pronounced. ADC activity (pkat mg-1 protein) 10 ADC and ODC activities DAO activity DAO activity was induced by light in a similar way as the ODC activity. A sharp peak of activity was observed 1 h after the start of light (P <0.005 compared with the value at the dark/light transition), after which the activity declined. The main DAO maximum was found 2 h after midday (P <0.05 compared with the value at 6 h of light) and then the enzyme activity steadily declined. The peak of activity at the light/dark transition (P <0.05 compared with the value at 14 h of light) coincided with the increase in the activities of both biosynthetic enzymes (Fig. 6). 8 6 4 2 0 0 4 8 12 16 20 Time (h) Fig. 4. Time-course of ADC activity found in soluble and particulate fractions in tobacco leaves grown under a 16/8 h photoperiod. Other details as for Fig. 2. ODC activity (pkat mg-1 protein) Biosynthetic enzyme activities were measured both in the soluble and pellet fractions. No increase in ADC activity was observed in light, by contrast, the enzyme activity in the first hour of light decreased. At midday, ADC activity started increasing, reaching a high 1 h later (P <0.005 compared with the value at 7 h of light). After a decline, a second main peak of activity occurred at the light/dark transition (P <0.005 compared with the value at 12 h of light). The enzyme activity in the pellet fraction represented about 45% and 48% of the entire biosynthetic activity in the first and second enzyme peak, respectively (Fig. 4). Light dramatically stimulated the activity of the second PA biosynthetic enzyme, ODC (Fig. 5). The sharp increase at 1 h of light (P <0.005 compared with the value at the dark/light transition) was followed by a transient decrease and then further peaks were observed prior to and after the middle of the light period (statistical significance of both peaks P <0.005 compared with the value at 3 h) and at the beginning (P <0.005 compared with the value at the light/dark transition) and the end of the dark period (P <0.05 compared with the value at 4 h of dark). The ODC enzyme activity was present predominantly in the pellet fraction and was rather low in the soluble fraction except for a peak 1 h after light induction (P <0.005 compared with the value at the dark/light transition), at which time it represented about 35% of total ODC activity. From the third hour of light until the end of the dark period the soluble ODC activity was very low again. soluble particulate 10 soluble particulate 8 6 4 2 0 0 4 8 12 16 20 Time (h) Fig. 5. Time-course of ODC activity found in soluble and particulate fractions in tobacco leaves grown under a 16/8 h photoperiod. Other details as for Fig. 2. Discussion Although the number of studies on PAs has been growing constantly, it is still unclear whether ODC or ADC plays a decisive role in Put biosynthesis in plants. The existence of two alternative routes for Put synthesis could be explained by the necessity of specific regulation of different processes affected by Put during growth and development (Koetje et al., 1993). It has been suggested that the pathway via ADC is generally linked to stress responses and morphogenetic processes, whereas ODC is involved in the regulation of cell division in actively growing plant cell tissue (Koetje et al., 1993; Galston et al., 1997). However, it seems that species and/or tissue specificity should be Downloaded from http://jxb.oxfordjournals.org at Department of Plant Physiology, Faculty of Science, Charles University on 25 March 2009 59 1418 Gemperlová et al. non-photosynthetic tissues of tobacco, the particulate ADC activity appeared to be located mainly in nuclei (Bortolotti et al., 2004). The different compartmentation of biosynthetic enzymes may be related to distinct functions in different cell types (Bortolotti et al., 2004). The results presented here show that, in tobacco leaves, the total ADC activity was equally divided between the soluble and pellet fractions while the ODC activity was mainly localized in the pellet fraction and only very little activity was present in the cytosolic fraction (there was one peak of activity after light induction in which 35% of the total ODC activity was found). A significant peak of free and soluble conjugated PA forms observed in tobacco leaves 1–2 h after the middle of the light period coincided with the high activity of ODC pellet fraction. It is well documented that in many plant systems dividing tissues have high PA levels and activities of their biosynthetic enzymes (Pfosser et al., 1990; Bezold et al., 2003). In vitro experiments have shown that free PAs affect many functions of nucleic acids, including transcription and translation (for a review see Kumar et al., 1997). The observed midday peak in the contents of free Put and Spd in tobacco leaves may be related to high physiological activity of the leaf tissue at this time period. This hypothesis might be supported by the abovementioned fact that ODC expression coincided with early cell divisions observed in hypogenous tobacco tissues (Paschalidis and Roubelkis-Angelakis, 2005). Furthermore, this peak coincided with the main peak of cytokinins and with the maximum of indole-3-acetic acid (IAA) in tobacco leaves cultivated under identical light/ dark conditions (Nováková et al., 2005). Since cytokinins, auxin, and PAs are necessary for the progression through the cell cycle, this result leads to the assumption that cell division occurs at this period. At midday, growth in ADC activity was observed and the main peak of the activity was found at the light/dark transition. This distinct peak of enzymatic activity was accompanied by the accumulation of free PAs and PCAsoluble conjugates. The PA maximum occurring at the light/dark transition coincided with the increase in abscisic acid content determined in tobacco leaves cultivated under the identical light/dark conditions (Nováková et al., 2005). As ABA is a key component of plant defence, the coincidence of its maximum and ADC peak is in a good accordance with the presumption of a potential link of ADC to stress response. The observed midday and dark phase peaks in Put and Spd contents are in agreement with the accumulation pattern of PAs in maize calli during the light/dark phases (Bernet et al., 1999). Put was the most abundant PA in maize callus, as in tobacco leaves, reaching peaks at 8 h and 12 h in the light and then in the middle of the dark period. Large intracellular accumulation of free PAs would have harmful effects on the maintenance of cellular pH, ion DAO activity (pkat mg-1 protein) 5 4 3 2 1 0 0 4 8 12 16 20 Time (h) Fig. 6. Time-course of DAO activity in tobacco leaves grown under a 16/8 h photoperiod. Other details as for Fig. 2. taken into consideration in plants. In apple, the primary biosynthetic pathway for Put is biosynthesis through ADC (Hao et al., 2005), while in the green alga Chlamydomonas reinhardtii the activity of ODC exceeded that of ADC by orders of magnitude, being highly induced by light (Voigt et al., 2000). On the contrary, in leaves of Pharbitis nil a very rapid photoresponse of ADC with the major activity peak 1 h after illumination was observed, while the activity of ODC did not respond to illumination (Yoshida and Hirasawa, 1998). In tobacco plants, ADC seemed to be responsible for Put synthesis in old hypergenous vascular tissues, whereas ODC expression coincided with early cell divisions and catalysed Put synthesis in hypogenous tissues (Paschalidis and Roubelkis-Angelakis, 2005). In tobacco leaves a different effect of light was observed on the induction of PA biosynthetic enzymes. Marked photoinduction of ODC activity 1 h after the exposure to light was found. Further peaks of ODC were observed prior to and after the middle of the light period with a slight increase at the beginning and the end of the dark period. By contrast, ADC activity decreased at the first hour of the light period and started increasing at midday with the main peak at the light/dark transition. There is still a lack of knowledge about the tissue and subcellular localization of the enzymes involved in PA metabolism. In animal tissues, despite extensive research, the exact intracellular localization of ODC remains to be clarified. It varies from an exclusively cytoplasmic to both a cytoplasmic and a nuclear localization (Schipper et al., 2004). The assays performed on leaves of Arabidopsis thaliana showed that high particulate ODC activity was associated with the nuclei and chloroplast-enriched fractions (Tassoni et al., 2003). In photosynthetic tissues of tobacco plants ADC was found mainly in chloroplasts, mitochondria, and cytoplasm (Slocum, 1991), whereas in Downloaded from http://jxb.oxfordjournals.org at Department of Plant Physiology, Faculty of Science, Charles University on 25 March 2009 60 Diurnal changes in polyamine content 1419 homeostasis, and on a number of physiological functions where PAs are implied. The endogenous free PA levels could be regulated, apart from the control of PA biosynthesis, also by conjugation with hydroxycinnamic acids in some plant species (Martin-Tanguy, 1985; Creus et al., 1991; Bagni and Tassoni, 2001; Biondi et al., 2001). The detected high levels of PCA-soluble conjugates of Put and Spd (about 75% of the entire Put and Spd pool in the peak 1–2 h after the midday) seem to be the common phenomena in Nicotiana and have been reported by several authors (Martin-Tanguy, 1985; Pfosser et al., 1990; Gemperlová et al., 2005). It has been shown in alfalfa cell suspension cultures and in oak embryogenic cultures that the rate of PA conjugation with hydroxycinnamic acids influenced the endogenous free PA level and, indirectly, cell division (Cvikrová et al., 1999, 2003). The biological function of conjugated PAs remains unclear. It has been speculated that enzymatic hydrolysis of reversible conjugates could supply the cells with an additional amine reserve (Bonneau et al., 1994). The origin and function of bound PAs are still an open issue. Some reports demonstrated high levels of Put and Spd binding to proteins in actively dividing young tissues. The incorporation of Put, Spd, and Spm into thylakoid and stromal proteins was stimulated by light (Aribaud et al., 1995). Spm was the most efficiently conjugated PA (Della Mea et al., 2004). In tobacco leaves grown under a 16/8 h photoperiod the most abundant PA in PCA-insoluble conjugates was Cad, followed by Put, Spm, and Spd. Cad in animal tissues does not normally participate in any particular metabolic process, but it is completely excreted from cells into the culture medium (Hawel et al., 1994; Hawel and Byus, 2002). Synthesized Cad is, in tobacco leaves, most probably immediately incorporated into the PA-insoluble fraction, as the levels of both free and soluble conjugated forms are very low during the light/dark cycle. In this experiment, the peaks of insoluble Spm conjugates coincided with the peaks of Cad. It may be hypothesized that Cad participates in the protection of chloroplasts from photo damage or stress. Significant light-affected conjugation of PAs to endogenous proteins was observed in isolated chloroplasts from Helianthus tuberosus (Dondini et al., 2003). It should be noted that light quality is also important for PA metabolism. Lettuce explants cultured under blue light contained a higher proportion of PCA-insoluble conjugated PAs compared with explants grown under white or red light, which contained more of free PA and/or soluble conjugates (Hunter and Burritt, 2005). The level of free and PCA-soluble PAs seems to be during the light/dark cycle regulated by the activities of biosynthetic enzymes (in tobacco especially ODC) and the catabolic one, DAO, activity of which was found to be substrate inducible (Srivastava et al., 1977). The sharp increase in DAO activity in tobacco leaves 1 h after the exposure to light closely followed or was even concomitant with the light-induced increase in the activity of ODC. This may be the reason why only a slight increase in free Put content was observed after the photoinduction of ODC activity at the beginning of the light phase. The role of both PA conjugation with hydroxycinnamic acids and oxidation in maintaining free PA levels was documented with leaf explants of Chrysanthemum after inhibition of DAO activity. High levels of PAs, especially soluble PA conjugates accumulated in these explants (Aribaud et al., 1994). Timing of the main peak of DAO activity shortly (1 h) after the middle of the light phase is in agreement with the enzyme maximum observed in isolated etioplasts from Hordeum vulgare after light exposure for 9 h (Andreadakis and Kotzabasis, 1996). The DAO was found to be localized in cell walls in pea and maize tissues (Slocum and Furey, 1991) and imunohistochemical analyses revealed that DAO expression occurs in cells destined to undergo lignification in tobacco tissues (Paschalidis and Roubelakis-Angelakis, 2005). However, the comparable soluble and cell debris-associated DAO activity was measured in tobacco thin layers (Biondi et al., 2001) and the decline in Put content in the early Sphase of the cell cycle in Helianthus tuberosus was probably due to the high DAO activity occurring in the early S-phase (Serafini-Fracassini, 1991). From the published literature it is clear that the catabolic degradation of Put by DAO is not only a means to regulate the cellular content of this diamine (Kumar et al., 1997; Bagni and Tassoni, 2001). The reaction product of DAO, pyrroline, can be further catabolized to c-aminobutyric acid (GABA) and subsequently trans-aminated and finally incorporated into the Krebs cycle (Tiburcio et al., 1997) and/or GABA has been shown to play a key role in signal transduction pathways during stress response of many plants (Shelp et al., 1999). The other reaction product of DAO, hydrogen peroxide, might trigger the hypersensitive response, which is thought to be a form of programmed cell death (Walters, 2003) or might be utilized in lignification both during the normal growth and stress response (Angelini et al., 1993). The results presented here have shown that ADC and ODC activities, as well as putrescine oxidizing activity, correlated well with the alterations in PA content during the day/night period. The data indicate that the parameters studied here were not only light-affected but also appear to be regulated by an internal rhythm. Acknowledgements We are grateful to Dr I Macháčková and Professor M Mok for critical reading of the manuscript. We thank N Hatašová for skilful technical assistance. This work was supported by grants from the Grant Agency of the Czech Republic No. 206/03/0369. Downloaded from http://jxb.oxfordjournals.org at Department of Plant Physiology, Faculty of Science, Charles University on 25 March 2009 61 1420 Gemperlová et al. References Creus JA, Eucuentra A, Gavalda EG, Barcelo J. 1991. Binding of polyamines to different macromolecules in plants. In: Galston AW, Tiburcio AF, eds. Polyamines as modulators of plant development. Madrid: Ediciones Peninsular, 30–34. Cvikrová M, Binarová P, Eder J, Vágner M, Hrubcová M, Zoń J, Macháčková I. 1999. Effect of inhibition of phenylalanine ammonia-lyase activity on growth of alfalfa cell suspension culture: alterations in mitotic index, ethylene production, and contents of phenolics, cytokinins, and polyamines. Physiologia Plantarum 107, 329–337. Cvikrová M, Malá J, Hrubcová M, Eder J, Zoń J, Macháčková I. 2003. Effect of inhibition of biosynthesis of phenylpropanoids on sessile oak somatic embryogenesis. Plant Physiology and Biochemistry 41, 251–259. Della Mea M, Di Sandro A, Dondini L, Del Duca S, Vantini F, Bergamini C, Bassi R, Serafini-Fracassini D. 2004. A Zea mays 39 kDa thylakoid transglutaminase catalyses the modification by polyamines of light-harvesting complex II in a light-dependent way. Planta 219, 754–764. Dondini L, Del Duca S, Dall’Agata L, Bassi R, Gastaldelli M, Della Mea M, Di Sandro A, Claparols I, Serafini-Fracassini D. 2003. Suborganellar localization and effect of light on Helianthus tuberosus chloroplast transglutaminases and their substrates. Planta 217, 84–95. Flores HE. 1991. Catabolic pathway and secondary metabolism of polyamines in plants. In: Galston AW, Tiburcio AF, eds. Polyamines as modulators of plant development. Madrid: Ediciones Peninsular, 23–26. Galston AW, KaurSawhney R, Altabella T, Tiburcio AF. 1997. Plant polyamines in reproductive activity and response to abiotic stress. Botanica Acta 110, 197–207. Gemperlová L, Eder J, Cvikrová M. 2005. Polyamine metabolism during the growth cycle of tobacco BY-2 cells. Plant Physiology and Biochemistry 43, 375–381. Hao YJ, Kitashiba H, Honda C, Nada K, Moriguchi T. 2005. Expression of arginine decarboxylase and ornithine decarboxylase genes in apple cells and stressed shoots. Journal of Experimental Botany 56, 1105–1115. Hawel III L, Byus CV. 2002. A streamlined method for the isolation and quantitation of nanomole levels of exported polyamines in cell culture media. Analytical Biochemistry 311, 127–132. Hawel L, Tjandrawinata RR, Fukumoto GH, Byus CV. 1994. Biosynthesis and selective export of 1,5-diaminopentane (cadaverine) in mycoplasma-free cultured mammalian cells. Journal of Biological Chemistry 269, 7412–7418. Hunter DC, Burritt DJ. 2005. Light quality influences the polyamine content of lettuce (Lactuca sativa L.) cotyledon explants during shoot production in vitro. Plant Growth Regulation 45, 53–61. Koetje DS, Kononowicz H, Hodges TK. 1993. Polyamine metabolism associated with growth and embryogenic potential of rice. Journal of Plant Physiology 141, 215–221. Kramer GF, Krizek DT, Mirecki RM. 1992. Influence of photosynthetically active radiation and spectral quality on UVB-induced polyamine accumulation in soybean. Phytochemistry 31, 1119–1125. Kumar A, Altabella T, Taylor MA, Tiburcio AF. 1997. Recent advances in polyamine research. Trends in Plant Science 2, 124–136. Martin-Tanguy J. 1985. The occurrence and possible function of hydroxycinnamoyl acid amides in plants. Plant Growth Regulation 3, 381–399. Moysset L, Trull O, Santos MA, Simon E, Torne JM. 2002. Effect of end-of-day irradiations on polyamine accumulation in petal cultures of Araujia sericifera. Physiologia Plantarum 114, 135–141. Nováková M, Motyka V, Dobrev PI, Malbeck J, Gaudinova A, Vankova R. 2005. Diurnal variation of cytokinin, auxin, and Altamura MM, Torrigiani P, Falasca G, Rossini P, Bagni N. 1993. Morpho-functional gradients in superficial and deep tissues along tobacco stem: polyamine levels, biosynthesis and oxidation, and organogenesis in vitro. Journal of Plant Physiology 142, 543–551. Andreadakis A, Kotzabasis K. 1996. Changes in the biosynthesis and catabolism of polyamines in isolated plastids during chloroplast photodevelopment. Journal of Photochemistry and Photobiology B-Biology 33, 163–170. Angelini R, Bragaloni M, Federico R, Infantino A, Portapuglia A. 1993. Involvement of polyamines, diamine oxidase and peroxidase in resistance of chickpea to Ascochyta rabiei. Journal of Plant Physiology 142, 704–709. Antognoni F, Fornale S, Grimmer C, Komor E, Bagni N. 1998. Long-distance translocation of polyamines in phloem and xylem of Ricinus communis L. plants. Planta 204, 520–527. Aribaud M, Carre M, Martin-Tanguy J. 1994. Polyamine metabolism and in-vitro cell multiplication and differentiation in leaf explants of Chrysanthemum morifolium Ramat. Plant Growth Regulation 15, 143–155. Aribaud M, Carré M, Martin-Tanguy J. 1995. Transglutaminaselike activity in chrysanthemum leaf explants cultivated in vitro in relation to cell growth and hormone treatment. Plant Growth Regulation 16, 11–17. Bagni N, Tassoni A. 2001. Biosynthesis, oxidation and conjugation of aliphatic polyamines in higher plants. Amino Acids 20, 301–317. Beranger-Novat N, Monin J, Jassey J, Martin-Tanguy J. 1997. Polyamine catabolism in dormant embryos of the spindle tree (Euonymus europaeus L.) and in dormancy break obtained after treatment with giberelic acid. Plant Growth Regulation 21, 65–70. Bernet E, Claparols I, Dondini L, Santos MA, Serafini-Fracassini D, Torné JM. 1999. Changes in polyamine content, arginine and ornithine decarboxylases and transglutaminase activities during light–dark phases (of initial differentiation) in maize calluses and their chloroplasts. Plant Physiology and Biochemistry 37, 899–909. Besford RT, Richardson CM, Campos JL, Tiburcio AF. 1993. Effect of polyamines on stabilization of molecular complexes of thylakoid membranes of osmotically stressed oat leaves. Planta 189, 201–206. Bezold TN, Loy JB, Minocha SC. 2003. Changes in the cellular content of polyamines in different tissues of seed and fruit of a normal and a hull-less seed variety of pumpkin during development. Plant Science 164, 743–752. Biondi S, Scaramagli S, Capitani F, Altamura MM, Torrigiani P. 2001. Methyl jasmonate up-regulates biosynthetic gene expression, oxidation and conjugation of polyamines, and inhibits shoot formation in tobacco thin layers. Journal of Experimental Botany 52, 231–242. Bonneau L, Carre M, Martin-Tanguy J. 1994. Polyamines and related enzymes in rice seeds differing in germination potential. Plant Growth Regulation 15, 75–82. Bortolotti C, Cordeiro A, Alcazar R, Borrel A, Culianez-Macia FA, Tiburcio AF, Altabella T. 2004. Localization of arginine decarboxylase in tobacco plants. Physiologia Plantarum 120, 84–92. Bouchereau A, Aziz A, Larher F, Martin-Tanguy J. 1999. Polyamines and environmental challenges: recent development. Plant Science 140, 103–125. Bradford MM. 1976. A rapid and sensitive method for quantification of microquantities of protein utilizing the principle of protein– dye binding. Analytical Biochemistry 72, 248–254. Caffaro SV, Vincente C. 1994. Polyamine implication during soybean flowering induction and early reproductive transition of vegetative buds. Plant Physiology and Biochemistry 32, 391–397. Downloaded from http://jxb.oxfordjournals.org at Department of Plant Physiology, Faculty of Science, Charles University on 25 March 2009 62 Diurnal changes in polyamine content 1421 abscisic acid levels in tobacco leaves. Journal of Experimental Botany 56, 2877–2883. Paschalidis KA, Aziz A, Geny L, Primikirios NI, RoubelakisAngelakis KA. 2001. Polyamines in grapevine. In: RoubelakisAngelakis KA, ed. Molecular biology and biotechnology of the grapevine. Dordrecht, The Netherlands: Kluwer Academic Publishers, 109–152. Paschalidis KA, Roubelakis-Angelakis KA. 2005. Spatial and temporal distribution of polyamine levels and polyamine anabolism in different organs/tissues of the tobacco plant. Correlations with age, cell division/expansion, and differentiation. Plant Physiology 138, 142–152. Peč P, Chudý J, Macholán L. 1991. Determination of the activity of diamine oxidase from pea with Z-1,4-diamino-2-butene as a substrate. Biológia 46, 665–672. Pfosser M, Konigshofer H, Kandeler R. 1990. Free, conjugated, and bound polyamines during the cell cycle of synchronized cell suspension cultures of Nicotiana tabacum. Journal of Plant Physiology 136, 574–579. Pignatti C, Tantini B, Stefanelli C, Flamigni F. 2004. Signal transduction pathways linking polyamines to apoptosis. Amino Acids 27, 359–365. Schipper RG, Cuijpers VMJI, de Groot LHJK, Thio M, Verhofstad AAJ. 2004. Intracellular localization of ornithine decarboxylase and its regulatory protein, antizyme-1. Journal of Histochemistry and Cytochemistry 52, 1259–1266. Serafini-Fracassini D. 1991. Polyamine biosynthesis and conjugation to macromolecules during the cell cycle of Helianthus tuberosus tuber. In: Galston AW, Tiburcio AF, eds. Polyamines as modulators of plant development. Madrid: Ediciones Peninsular, 40–44. Serafini-Fracassini D, Del Duca S, Monti F, Poli F, Sacchetti G, Bregoli AM, Biondi S, Della Mea M. 2002. Transglutaminase activity during senescence and programmed cell death in the corolla of tobacco (Nicotiana tabacum) flowers. Cell Death and Differentiation 9, 309–321. Shelp BJ, Bown AW, McLean MD. 1999. Metabolism and functions of gamma-aminobutyric acid. Trends in Plant Science 4, 446–452. Slocum RD. 1991. Tissue and subcellular localization of polyamines and enzymes of polyamine metabolism. In: Slocum RD, Flores HE, eds. Biochemistry and physiology of polyamines in plants. Boca Raton: CRC Press, 93–105. Slocum RD, Flores HE, Galston AW, Einstein LH. 1989. Improved method for HPLC analysis of polyamines, agmatine and aromatic monoamines in plant tissue. Plant Physiology 89, 512–517. Slocum RD, Furey MJ. 1991. Electron-microscopic cytochemicallocalization of diamine and polyamine oxidases in pea and maize tissues. Planta 183, 443–450. Srivastava SK, Parkash V, Naik BI. 1977. Regulation of diamine oxidase in germinating pea seedlings. Indian Journal of Experimental Biology 16, 185–187. Tassoni A, Fornale S, Bagni N. 2003. Putative ornithine decarboxylase activity in Arabidopsis thaliana: inhibition and intracellular localization. Plant Physiology and Biochemistry 41, 871–875. Tassoni A, Van Buuren M, Franceschetti M, Fornale S, Bagni N. 2000. Polyamine content and metabolism in Arabidopsis thaliana and effect of spermidine on plant development. Plant Physiology and Biochemistry 38, 383–393. Theiss C, Bohley P, Voigt J. 2002. Regulation by polyamines of ornithine decarboxylase activity and cell division in the unicellular green alga Chlamydomonas reinhardtii. Plant Physiology 128, 1470–1479. Thomas T, Thomas TJ. 2001. Polyamine in cell growth and cell death: molecular mechanisms and therapeutic applications. Cell and Molecular Life Science 58, 244–258. Tiburcio AF, Altabella T, Borrell A, Masgrau C. 1997. Polyamine metabolism and its regulation. Physiologia Plantarum 100, 664–674. Voigt J, Deinert B, Bohley P. 2000. Subcellular localization and light–dark control of ornithine decarboxylase in the unicellular green alga Chlamydomonas reinhardtii. Physiologia Plantarum 108, 353–360. Wallace HM, Fraser AV, Hughes A. 2003. A perspective of polyamine metabolism. Biochemistry 376, 1–14. Walters D. 2003. Resistance to plant pathogens: possible roles for free polyamines and polyamine catabolism. New Phytologist 159, 109–115. Yoshida I, Hirasawa E. 1998. Photoinduction of arginine decarboxylase activity in leaves of Pharbitis nil. Phytochemistry 49, 2255–2259. Yoshida I, Yamagata H, Hirasawa E. 1999. Blue- and red-light regulation and circadian control of gene expression of Sadenosylmethionine decarboxylase in Pharbitis nil. Journal of Experimental Botany 53, 1525–1529. Downloaded from http://jxb.oxfordjournals.org at Department of Plant Physiology, Faculty of Science, Charles University on 25 March 2009 63 Journal of Experimental Botany Advance Access published December 14, 2006 Journal of Experimental Botany, Page 1 of 13 doi:10.1093/jxb/erl235 RESEARCH PAPER Altered cytokinin metabolism affects cytokinin, auxin, and abscisic acid contents in leaves and chloroplasts, and chloroplast ultrastructure in transgenic tobacco Lenka Polanská1,2,*, Anna Vičánková1,2, Marie Nováková1,3, Jiřı́ Malbeck1, Petre I. Dobrev1, Břetislav Brzobohatý4, Radomı́ra Vaňková1 and Ivana Macháčková1 1 Institute of Experimental Botany AS CR, Rozvojová 135, 165 02 Prague 6-Lysolaje, Czech Republic Charles University in Prague, Faculty of Science, Department of Plant Physiology, Viničná 5, 128 44 Prague 2, Czech Republic 2 3 Charles University in Prague, Faculty of Science, Department of Biochemistry, Hlavova 2030/8, 128 43 Prague 2, Czech Republic 4 Institute of Biophysics AS CR, Královopolská 135, 612 65 Brno, Czech Republic Received 2 August 2006; Revised 12 October 2006; Accepted 16 October 2006 Abstract Cytokinins (CKs) are involved in the regulation of plant development including plastid differentiation and function. Partial location of CK biosynthetic pathways in plastids suggests the importance of CKs for chloroplast development. The impact of genetically modified CK metabolism on endogenous CK, indole-3-acetic acid, and abscisic acid contents in leaves and isolated intact chloroplasts of Nicotiana tabacum was determined by liquid chromatography/mass spectrometry and two-dimensional high-performance liquid chromatography, and alterations in chloroplast ultrastructure by electron microscopy. Ectopic expression of Sho, a gene encoding a Petunia hybrida isopentenyltransferase, was employed to raise CK levels. The increase in CK levels was lower in chloroplasts than in leaves. CK levels were reduced in leaves of tobacco harbouring a CK oxidase/dehydrogenase gene, AtCKX3. The total CK content also decreased in chloroplasts, but CK phosphate levels were higher than in the wild type. In a transformant overexpressing a maize b-glucosidase gene, Zm-p60.1, naturally targeted to plastids, a decrease of CK-O-glucosides in chloroplasts was found. In leaves, the changes were not significant. CK-Oglucosides accumulated to very high levels in leaves, but not in chloroplasts, of plants overexpressing a ZOG1 gene, encoding trans-zeatin-O-glucosyltransferase from Phaseolus lunatus. Manipulation of the CK content affected levels of indole-3-acetic and abscisic acid. Chloroplasts of plants constitutively overexpressing Sho displayed ultrastructural alterations including the occasional occurrence of crystalloids and an increased number of plastoglobuli. The other transformants did not exhibit any major differences in chloroplast ultrastructure. The results suggest that plant hormone compartmentation plays an important role in hormone homeostasis and that chloroplasts are rather independent organelles with respect to regulation of CK metabolism. Key words: Abscisic acid, auxin, chloroplast ultrastructure, cytokinin metabolism, cytokinin oxidase/dehydrogenase, bglucosidase, isopentenyltransferase, Nicotiana tabacum, Sho, zeatin-O-glucosyltransferase. Introduction The plant hormones cytokinins (CKs) are known to be involved in many processes related to plastid development and functions. Within 10 years of the discovery of kinetin (6-furfurylaminopurine) as a plant growth regulator CK (Miller et al., 1956) in cultured tobacco tissue, it was * To whom correspondence should be addressed. E-mail: [email protected] Abbreviations: ABA, abscisic acid; BA, benzyladenine; CaMV, cauliflower mosaic virus; CK, cytokinin; DMSO, dimethylsulphoxide; FW, fresh weight; IAA, indole-3-acetic acid; IPT, isopentenyltransferase; LCMS, liquid chromatography/mass spectrometry; WT, wild type. ª The Author [2006]. Published by Oxford University Press [on behalf of the Society for Experimental Biology]. All rights reserved. For Permissions, please e-mail: [email protected] 64 2 of 13 Polanská et al. CKs may be synthesized in plastids by the methylerythritol phosphate pathway and that four isopentenyltransferases (IPTs) are localized to plastids in Arabidopsis cells. Unexpectedly, Sakakibara et al. (2005) have found that Agrobacterium IPT functions in plastids, even though it lacks a typical plastid-targeting sequence, and it prefers a substrate different from that of genuine plant IPTs. It was previously shown that the enhanced content of CKs in chloroplasts of tobacco plants overexpressing the A. tumefaciens ipt gene driven by a light-inducible promoter of a Pisum sativum small subunit of Rubisco (Pssu-ipt) correlated with highly anomalous chloroplast ultrastructure (Synková et al., 2006). Interestingly, although a clear increase in CKs was observed in chloroplasts, it remained about an order of magnitude lower than the levels found in whole leaf extracts. In the present study, the impact of manipulation of CK metabolism on the endogenous CK pool in leaves and isolated intact chloroplasts, and on chloroplast ultrastructure in tobacco (Nicotiana tabacum L.) was evaluated. CK-overproducing plants (expressing the Sho gene coding for a Petunia hybrida IPT), plants with decreased CK levels (overexpressing CK oxidase/dehydrogenase AtCKX3 from Arabidopsis thaliana), and plants with altered CK glucoconjugation (overexpressing maize bglucosidase Zm-p60.1 or zeatin-O-glucosyltransferase ZOG1 from Phaseolus lunatus) were compared. The simplified model of CK metabolism with highlighted enzymes involved in this study is presented in Fig. 1. Taking into account the immense cross-talk among plant hormones, the contents of auxin (IAA) and abscisic acid (ABA) were also determined. shown that kinetin could stimulate chloroplast differentiation from proplastids, including grana formation (Stetler and Laetsch, 1965). Kinetin was also found to almost double chloroplast number per cell in etiolated tobacco leaf discs incubated in the light (Boasson and Laetsch, 1969). It has been demonstrated that exogenously applied CKs induce partial development of chloroplasts from proplastids, amyloplasts, and etioplasts, especially in darkness (Chory et al., 1994; Kusnetsov et al., 1994). For example, benzyladenine (BA) treatment of excised watermelon cotyledons accelerated degradation of reserve material and differentiation of plastids with a more developed inner membrane system, as well as increased levels of plastid pigments and enzymes (Longo et al., 1979). Moreover, applied BA accelerated the redevelopment of grana and stroma in regreening plastids, the increase of reappeared NADPH-protochlorophyllide oxidoreductase, and the contents of chlorophyll and protein in regreening tobacco leaves (Zavaleta-Mancera et al., 1999a, b). The action of CKs in plastids seems to be mediated by the stimulation of the expression of several plastid-related genes, of both nuclear and plastid origin, in particular that of the nuclear gene for chlorophyll a/b-binding polypeptide of the light-harvesting complex and the gene coding for the small subunit of Rubisco in Nicotiana tabacum (Abdelghani et al., 1991), Arabidopsis (Chory et al., 1994), or Dianthus caryophyllus (Winiarska et al., 1994). For extensive reviews, see Schmülling et al. (1997) or Parthier (1989). Recently, Brenner et al. (2005) identified by genomewide expression profiling five rapidly CK-induced plastid transcripts in Arabidopsis seedlings, indicating a fast transfer of the CK signal to plastids or its direct perception there. The CK effect on gene expression may be mediated via the hormone interaction with specific proteins. In the presence of trans-zeatin, a 64 kDa chloroplast zeatinbinding protein was found to activate the chloroplast but not the nuclear genome (Lyukevich et al., 2002). The occurrence of endogenous CKs in plastids has been proven. CKs are present in chloroplast tRNAs (Vreman et al., 1978). Zeatin-type CKs were detected in isolated spinach chloroplasts by paper chromatography and soybean callus bioassay (Davey and Van Staden, 1981). Immunolocalization of CKs revealed low labelling in the stroma of plastids of Tilia cordata embryo cotyledons and roots (Kärkönen and Simola, 1999). Eventually, CK occurrence in chloroplasts was rigorously confirmed by LCMS (liquid chromatography/mass spectrometry) which allows identification of a whole range of CK metabolites with high accuracy (Benková et al., 1999). The importance of CKs for plastid development and function may be deduced from the partial localization of the CK biosynthetic pathway to this compartment. Kasahara et al. (2004) showed that the prenyl group of Materials and methods Plant materials and growth conditions Samples of mature leaves from four different types of transgenic tobacco plants with altered CK metabolism and the corresponding wild types (WTs), all in the vegetative stage, collected imediately after the dark period, were used. 1. Tobacco in vitro lines expressing an IPT homologue Sho (Shooting) gene from P. hybrida under the control of a constitutive cauliflower mosaic virus (CaMV) 35S promoter or construct with four 35S enhancer elements (Zubko et al., 2002) and WT N. tabacum L. cv. Petit Havana SR1 cultivated on solid hormone-free MS medium (Murashige and Skoog, 1962) supplemented with sucrose (30 g l1). To achieve inducible Sho expression, a dexamethasone-inducible gene expression system (pOp6/LhGR; Šámalová et al., 2005) was employed. Sho expression was induced in 8–9-week-old transgenic plants with 20 lM dexamethasone [in 0.05% dimethylsulphoxide (DMSO), 5350 ml per plant watered for 13 d]. As controls, transgenic plants treated with water or 0.05% DMSO and WT SR1 plants treated with water, 0.05% DMSO, or 20 lM dexamethasone in 0.05% DMSO were used. The plants were cultivated in a growth chamber (16/8 h photoperiod at 130 lE m2 s1, day/night temperature of 25/23 C and relative humidity ~80%) for 5 weeks and then transferred, before Sho induction, to a greenhouse (in summer, without additional light supplementation). 65 Cytokinins in chloroplasts of transgenic tobacco 3 of 13 (1000 g; 2 min; 4 C). The pellet was resuspended in medium. All the procedures were done at 4 C. Chlorophyll determination Chlorophyll was extracted into 80% (v/v) acetone. The total chlorophyll content (a+b) was calculated from the absorbance at 652 nm of the clear extract after centrifugation (500 g, 5 min) according to Arnon (1949). Chloroplast integrity The integrity of chloroplasts was determined by the latency of glyceraldehyde-3-phosphate dehydrogenase according to Latzko and Gibbs (1968). Chloroplasts were incubated in a reaction mixture (0.33 M TRIS–HCl, pH 8.5; 17 mM Na2HAsO47H2O; 4 mM cysteine; 20 mM NaF; 40 lM NADP+). The reaction was initiated with 0.02 M glyceraldehyde-3-phosphate. Reduction of NADP+ was followed at 340 nm for 5 min. The same assay was run with chloroplasts disrupted with 0.01 M MgCl2. The percentage of intact chloroplasts was calculated from the difference between the original and disrupted sample. Fig. 1. Simplified model of CK metabolism including known localization of examples of involved enzymes proved by in planta experiments. The enzymes employed in this study are highlighted. AtCKX, Arabidopsis CK oxidase/dehydrogenase (Werner et al., 2003); AtIPT, Arabidopsis isopentenyltransferase (Kasahara et al., 2004); Sho, petunia isopentenyltransferase (Zubko et al., 2002); UGT, Arabidopsis Nglucosyltransferase (Hou et al., 2004); Zm-p60.1, maize b-glucosidase (Kristoffersen et al., 2000); ZOG1, bean trans-zeatin-O-glucosyltransferase (Martin et al., 2001a); cisZOG, maize cis-zeatin-O-glucosyltransferase (Martin et al., 2001b); ZOX1, bean zeatin-O-xylosyltransferase (Martin et al., 1993); ?, unknown or uncertain localization. The interconversions among iP-, transZ-, cisZ-, and DHZ-type CKs and their bases, ribosides, and phosphates, CK transport and tissue localization are not depicted. For more details see Sakakibara (2006). All the other plants were cultivated in a soil substrate in a growth chamber (16/8 h photoperiod at 150 lE m2 s1, 26/20 C and relative humidity ~80%).: 2. Eleven-week-old tobacco plants (35S:AtCKX3) overexpressing a gene for CK oxidase/dehydrogenase from A. thaliana under a constitutive CaMV 35S promoter (Werner et al., 2001) and 9week-old WT N. tabacum L. cv. Samsun NN. 3. Eight- to 9-week-old tobacco plants (35S:P60) overexpressing a maize b-glucosidase Zm-p60.1 naturally targeted to plastids under a CaMV 35S promoter (Kiran et al., 2006) and WT Nicotiana tabacum L. cv. Petit Havana SR1. 4. Eight- to 9-week-old tobacco plants (35S:ZOG1) harbouring a trans-zeatin-O-glucosyltransferase gene from P. lunatus under a constitutive CaMV 35S promoter (Martin et al., 2001a) and WT N. tabacum L. cv. Wisconsin 38. Chloroplast isolation Intact chloroplasts were isolated and purified as described by Kiran et al. (2006). The chloroplast fraction was recovered from the homogenate of deribbed leaves in homogenization medium [0.33 M sorbitol; 50 mM TRIS–HCl, pH 7.8; 0.4 mM KCl; 0.04 mM Na2EDTA; 0.1% (w/v) bovine serum albumin; 1% (w/v) polyvinylpyrrolidone; 5 mM isoascorbic acid] by centrifugation, layered on a Percoll density gradient [40% and 80% (v/v) Percoll solution in resuspension medium (0.33 M sorbitol; 2 mM Na2EDTA; 1 mM MgCl2; 1 mM MnCl2; 50 mM HEPES, pH 7.6)] and centrifuged (1000 g; 15 min; 4 C). Intact chloroplasts were collected at the interface of the gradient, diluted with medium, and centrifuged Extraction and purification of IAA, ABA, and CK The detailed procedure for hormone extraction, purification, and quantification has been described in Kiran et al. (2006). IAA, ABA, and CKs were extracted overnight at –20 C with Bieleski solvent (Bieleski, 1964). [3H]IAA and [3H]ABA (Sigma, USA) and 12 deuterium-labelled CKs ([2H5]t-Z, [2H5]t-ZR, [2H5]t-Z7G, [2H5]tZ9G, [2H5]t-ZOG, [2H5]t-ZROG, [2H3]DHZ, [2H3]DHZR, [2H6]iP, [2H6]iPR, [2H6]iP7G, and [2H6]iP9G; Apex Organics, UK) were added as internal standards. The extracts were purified using Sep-Pak C18 cartridges (Waters Corporation, Milford, MA, USA) and an Oasis MCX mixed mode, cation exchange, reverse-phase column (150 mg, Waters) (Dobrev and Kamı́nek, 2002). After a wash with 1 M HCOOH, IAA and ABA were eluted with 100% MeOH and evaporated to dryness. Further, CK phosphates were eluted with 0.34 M NH4OH in water, and CK bases, ribosides, and glucosides were eluted with 0.34 M NH4OH in 60% (v/v) MeOH. Phosphates were converted to ribosides with alkaline phosphatase. IAA and ABA were separated and quantified by two-dimensional high-performance liquid chromatography (2D-HPLC) according to Dobrev et al. (2005). Purified CK samples were analysed by an LCMS system consisting of an HTS PAL autosampler (CTC Analytics, Switzerland), Rheos 2000 quaternary pump (FLUX, Switzerland) with Csi 6200 Series HPLC oven (Cambridge Scientific Instruments, UK), and an LCQ Ion Trap mass spectrometer (Finnigan, USA) equipped with an electrospray. A 10 ll aliquot of sample was injected onto a C18 column (AQUA, 2 mm3250 mm35 lm, Phenomenex, USA) and eluted with 0.0005% acetic acid (A) and acetonitrile (B). The HPLC gradient profile was as follows: 5 min 10% B, then increasing to 17% for 10 min, and to 46% for a further 10 min at a flow rate of 0.2 ml min1. The column temperature was kept at 30 C. The effluent was introduced into a mass spectrometer being operated in the positive ion, full-scan MS/MS mode. Quantification was performed using a multilevel calibration graph with deuterated CKs as internal standards. Plant hormone content is usually presented in picomoles per gram fresh weight (FW) of plant tissue. Here, hormone concentration is also expressed per unit of chlorophyll, allowing the comparison with hormone levels in isolated chloroplasts. The differences were estimated according to the value of the standard error. Transmission electron microscopy Leaf blade segments (2 mm32 mm3leaf thickness) were fixed for 2 h with 2.5% (v/v) glutaraldehyde in 0.1 M phosphate buffer, 66 4 of 13 Polanská et al. pH 7.2, followed by 2 h in 2% (w/v) osmic acid. After dehydration in a graded ethanol series up to 100% ethanol, samples were embedded via propylene oxide in low viscosity Spurr’s resin. The ultrastructure was evaluated on transverse ultrathin sections of embedded objects contrasted with a saturated solution of uranyl acetate in 70% (v/v) aqueous ethanol, followed by a lead citrate solution treatment according to Reynolds (1963) using a transmission electron microscope (Philips Morgani) at an operating voltage of 70 kV. Results Hormone analyses In order to investigate the role of hormone compartmentation and to study the interactions among plant hormones, the contents of CKs, free IAA, and ABA were determined in mature leaves and in isolated intact chloroplasts of four different types of transgenic tobacco with altered CK metabolism and in the corresponding WTs during the vegetative stage of plant development. The integrity of the analysed chloroplasts immediately after their isolation was >90%, as checked by the latency of the activity of the stromal enzyme glyceraldehyde-3-phosphate dehydrogenase. Plants expressing isopentenyltransferase Sho The total CK content in leaves after Sho induction by dexamethasone was eight times that in the corresponding WT (SR1) (Fig. 2A). The most elevated CK metabolites were phosphates (predominantly iPMP) and N-glucosides (especially iP7G). The level of physiologically active CK bases and ribosides (predominantly of iP and iPR) increased to a lesser extent. No significant difference in the content of CK-O-glucosides was observed. As CK levels did not differ significantly among WT plants treated with water, with dexamethasone in DMSO, or with DMSO solution or non-induced Sho gene-carrying plants, only the CK content of WT plants treated with water is shown. The CK elevation in induced Sho leaves was lower when the CK content was expressed per mg of chlorophyll, as the transgenics had a higher chlorophyll content than the WT (Fig. 2B). The difference between Sho transformants and the WT was much lower in isolated chloroplasts (Fig. 2C) than in leaves. Indeed, except for CK-O-glucosides, the percentage of leaf CKs sequestered in chloroplasts was lower in Sho compared with WT (see supplementary Table 1 available at JXB online). Nevertheless, the CK content of chloroplasts from plants expressing Sho was more than twice as high as that of the WT. The contents of individual CKs in leaves and isolated chloroplasts are given in supplementary Table 2 available at JXB online. On a FW basis, the higher CK content in Sho leaves compared with the WT was accompanied by a slightly higher content of IAA and ABA (Table 1). These differences were not significant when expressed per unit of chlorophyll. Chloroplasts of Sho plants had lower IAA, but higher ABA content than chloroplasts of WT plants. Fig. 2. Cytokinin content (B+R, free bases and ribosides; NG, N-glucosides; OG, O-glucosides; P, phosphates) in leaves (per fresh weight A and per unit of chlorophyll B) and in isolated intact chloroplasts (C) of transgenic dexamethasone-induced Sho (SHO, black) and control SR1 (WT, striped) tobacco. Error bars display the SE. Plants expressing CK oxidase/dehydrogenase AtCKX3 The constitutive overexpression of AtCKX3 significantly delayed development of transgenic plants; consequently, hormone contents and chloroplast ultrastructure were 67 Cytokinins in chloroplasts of transgenic tobacco 5 of 13 Table 1. Endogenous contents of free IAA and ABA in leaves and in isolated intact chloroplasts of transgenic dexamethasoneinduced Sho and WT (SR1) tobacco Values represent the mean of two replicates (individual measurements are shown in parentheses). Sho IAA Leaves (pmol g1 FW) Leaves (pmol mg1 chlorophyll) Chloroplasts (pmol mg1 chlorophyll) ABA Leaves (pmol g1 FW) Leaves (pmol mg1 chlorophyll) Chloroplasts (pmol mg1 chlorophyll) 90.7 (117.9; 63.5) 73.7 (95.8; 51.6) 2.5 (2.8; 2.2) WT (SR1) 71.2 (81.1; 61.3) 74.2 (84.4; 63.9) 3.4 (2.8; 4.0) 453.5 (749.1; 157.8) 368.7 (609.0; 128.3) 360.0 (628.1; 91.9) 375.0 (654.3; 95.7) 16.8 (18.7; 14.9) 12.8 (13.3; 12.2) analysed in transgenic plants that were older than the corresponding WT (SNN) plants. The total CK content was reduced in leaves (Fig. 3A) to about one-half of that in the WT. Particularly altered were the levels of CK-Nglucosides and, to a lesser extent, bases and ribosides. The chlorophyll content was reduced in transgenic plants, so that the differences between transgenic and WT tobacco are reduced when expressed per unit of chlorophyll (Fig. 3B). The CK content was changed to a lesser extent in isolated intact chloroplasts (Fig. 3C). Further, when AtCKX3 and WT were compared, the percentage of leaf CK bases and ribosides remained almost unchanged, whilst those of phosphates sequestered in chloroplasts increased; the values for CK-glucosides dropped significantly (see supplementary Table 1 available at JXB online). The content of IAA was not significantly changed in AtCKX3 leaves (Table 2) compared with WT when expressed per FW, but it was higher when expressed per unit of chlorophyll. The ABA level was lower in leaves of transgenic tobacco. No significant changes of IAA or ABA levels were observed in chloroplasts (Table 2). Plants expressing b-glucosidase Zm-p60.1 Compared with WT (SR1) leaves, a trend towards an increase in CK bases and ribosides and CK-N-glucosides was observed in leaves (Fig. 4A, B) of tobacco overexpressing maize b-glucosidase Zm-p60.1 targeted to plastids. A dramatic decrease in the levels of CK-Oglucosides remained restricted to the CK pool found in chloroplasts (Fig. 4C). Interestingly, a moderate decrease was also observed for CK-N-glucosides in the chloroplast fraction. The trend towards an increase in CK bases and ribosides observed in whole leaf extracts was also seen in the chloroplast fraction. Accordingly, the percentage of leaf CK-O- and N-glucosides sequestered in chloroplasts was significantly lower in the transgenic plants than in WT plants, while the values changed only marginally for Fig. 3. Cytokinin content (B+R, free bases and ribosides; NG, Nglucosides; OG, O-glucosides; P, phosphates) in leaves (per fresh weight A and per unit of chlorophyll B) and in isolated intact chloroplasts (C) of transgenic 35S:AtCKX3 (CKX, black) and control SNN (WT, striped) plants. Error bars display the SE. CK bases and ribosides, and for phosphates (see supplemetary Table 1 available at JXB online). The IAA measurements revealed a significantly lower level of IAA in leaves and in chloroplasts of transgenic plants compared with the WT IAA level (Table 3). The 68 6 of 13 Polanská et al. Table 2. Endogenous contents of free IAA and ABA in leaves and in isolated intact chloroplasts of transgenic 35S:AtCKX3 and WT (SNN) tobacco Values represent the mean of two replicates (individual measurements are shown in parentheses). 35S:AtCKX3 IAA Leaves (pmol g1 FW) Leaves (pmol mg1 chlorophyll) Chloroplasts (pmol mg1 chlorophyll) ABA Leaves (pmol g1 FW) Leaves (pmol mg1 chlorophyll) Chloroplasts (pmol mg1 chlorophyll) 125.0 (177.4; 72.6) 200.5 (279.1; 121.8) 5.2 (4.3; 6.1) WT (SNN) 122.1 (147.3; 96.9) 114.4 (151.6; 77.2) 5.2 (9.9; 0.4) 359.5 (447.7; 271.3) 1102.3 (822.9; 1381.6) 579.9 (704.4; 455.4) 974.1 (847.2; 1100.9) 68.0 (44.0; 92.0) 58.8 (106.6; 10.9) ABA content was reduced in Zm-p60.1 leaves compared with WT (Table 3). In chloroplasts, the difference in ABA level was less pronounced. Plants expressing zeatin-O-glucosyltransferase ZOG1 When a ZOG1 gene was constitutively expressed in tobacco, CK-O-glucosides accumulated in leaves (Fig. 5A). The level of t-ZOG was two orders of magnitude higher in leaves of transformants than in those of WT (W38). No significant difference in the content of CK-Nglucosides, and only marginal reduction of the remaining CK forms was observed. These differences reduced further when the CK content was expressed per unit of chlorophyll (Fig. 5B), while that of CK-O-glucosides was more profound, as the chlorophyll content was lower in ZOG1 than in WT. By contrast, in chloroplasts, very low levels of CK-O-glucosides were found (Fig. 5C). The predominant CK forms were bases and ribosides, the contents of which, as well as that of CK phosphates, were higher in ZOG1 chloroplasts than in WT chloroplasts. Indeed, out of the transgenic plants analysed, only in the plants overexpressing ZOG1 was the percentage of leaf CK bases and ribosides sequestered in chloroplasts significantly higher in the transgenic plants than in WT plants (see supplementary Table 1 available at JXB online). In leaves of transgenic plants, a 2-fold higher IAA content was detected (Table 4). The ABA level was not significantly affected in ZOG1 leaves on a FW basis, but was higher when expressed per unit of chlorophyll. IAA and ABA contents were higher in ZOG1 chloroplasts than in WT chloroplasts. Fig. 4. Cytokinin content (B+R, free bases and ribosides; NG, Nglucosides; OG, O-glucosides; P, phosphates) in leaves (per fresh weight A and per unit of chlorophyll B) and in isolated intact chloroplasts (C) of transgenic 35S:Zm-p60.1 (P60, black) and control SR1 (WT, striped) tobacco. Error bars display the SE. enhancers (Fig. 6A, detailed view in Fig. 6B) as well as the 35S promoter. However, some chloroplasts from these plants did not show apparent changes in the structure compared with the WT (Fig. 6C). Chloroplasts from Shooverexpressing plants contained more plastoglobuli than WT chloroplasts (Fig. 6D). The crystalloids were never present in chloroplasts of WT plants cultivated in vitro. Ultrastructure observations The most striking anomaly in chloroplast ultrastructure was the occasional occurrence of crystalloid structures in chloroplasts of in vitro-cultivated plants harbouring the Sho gene under transcriptional control of four 35S 69 Cytokinins in chloroplasts of transgenic tobacco 7 of 13 Table 3. Endogenous contents of free IAA and ABA in leaves and in isolated intact chloroplasts of transgenic 35S:Zm-p60.1 and WT (SR1) tobacco Values represent the mean of two replicates (individual measurements are shown in parentheses). 35S:Zm-p60.1 IAA Leaves (pmol g1 FW) Leaves (pmol mg1 chlorophyll) Chloroplasts (pmol mg1 chlorophyll) ABA Leaves (pmol g1 FW) Leaves (pmol mg1 chlorophyll) Chloroplasts (pmol mg1 chlorophyll) 85.7 (95.1; 76.2) 60.3 (66.9; 53.6) 0.8 (1.4; 0.2) WT (SR1) 144.0 (146.5; 141.5) 135.2 (137.5; 132.8) 1.2 (2.1; 0.3) 481.6 (486.9; 476.3) 338.7 (342.4; 334.9) 839.0 (785.9; 892.1) 787.5 (737.6; 837.4) 23.7 (21.5; 25.8) 26.3 (40.4; 12.1) After induction of the Sho gene expression, no crystalloids were found, only an increased grana stacking was observed (Fig. 6E). The crystalloids did not appear in chloroplasts of WT plants treated either with water (Fig. 6F), with dexamethasone in DMSO, or with DMSO solution, or in chloroplasts of non-induced plants carrying the Sho gene. Representative chloroplasts from the other types of transgenic tobacco and corresponding WTs are shown in Fig. 7. The most obvious feature observed was the different starch accumulation. The chloroplasts of tobacco overexpressing AtCKX3 (Fig. 7A) and Zm-p60.1 (Fig. 7C) have less starch inclusions than chloroplasts of WT SNN (Fig. 7B) and SR1 (Fig. 7D), respectively. In contrast, the chloroplasts of ZOG1 transformants (Fig. 7E) contained more starch inclusions than WT W38 chloroplasts (Fig. 7F). With increasing starch content, the shape of chloroplasts altered from lens-shaped to more loaf-like. Discussion Hormone analyses In order to elucidate the degree of chloroplast CK autonomy, the impact of altered CK metabolism on the CK pool in isolated intact chloroplasts and in bulk leaf tissue was compared. Bearing in mind hormone cross-talk, the effect of altered CK metabolism on IAA and ABA levels was also followed. The presence of these hormones in chloroplasts was demonstrated in earlier studies through the use of immunocytochemical (Sossountzov et al., 1986; Ohmiya and Hayashi, 1992; Kärkönen and Simola, 1999) or fractionation techniques, followed by HPLC or LCMS analysis (Sandberg et al., 1990; Benková et al., 1999). Chloroplasts were found to be the compartment where most of the ABA in leaf tissue is formed by the methylerythritol phosphate pathway (Milborrow and Lee, 1998). Fig. 5. Cytokinin content (B+R, free bases and ribosides; NG, Nglucosides; OG, O-glucosides; P, phosphates) in leaves (per fresh weight A and per unit of chlorophyll B) and in isolated intact chloroplasts (C) of transgenic 35S:ZOG1 (ZOG, black) and control W38 (WT, striped) tobacco. Error bars display the SE. This pathway also provides most of the prenyl group of CKs (Kasahara et al., 2004). Nordström et al. (2004) demonstrated de novo CK synthesis in Arabidopsis shoots and tobacco leaves, and suggested that the presence of chloroplasts might be a prerequisite for the iPMPindependent pathway of CK biosynthesis. 70 8 of 13 Polanská et al. Plants expressing isopentenyltransferase Sho Numerous studies of plants transformed with the ipt gene, encoding an IPT from Agrobacterium tumefaciens, have demonstrated enhanced biosynthesis of endogenous CKs, especially of the Z-type (Redig et al., 1996; Motyka et al., 2003). Recently, genuine plant IPT genes (AtIPT1– 9) were identified in the A. thaliana genome (Kakimoto, 2001; Takei et al., 2001) and their homologue Sho in P. hybrida (Zubko et al., 2002). The overexpression of the plant homologue AtIPT8 (Sun et al., 2003), as well as Sho (Zubko et al., 2002), in contrast to the bacterial ipt, leads to profound accumulation of the iP-type of CKs. A dexamethasone-inducible gene expression system (pOp6/ LhGR; Šámalová et al., 2005) was used to drive Sho expression in transgenic tobacco. Following Sho induction in leaf extracts, major increases were observed in iP7G and iPMP, which is in line with previous results reported for petunia and tobacco overexpressing Sho constitutively (Zubko et al., 2002). The CK content was also increased in chloroplasts, although to a relatively lower extent. This is in accordance with preliminary results of feeding experiments (data not shown) showing that this IPT preferentially utilizes mevalonate as a substrate, indicating its cytosolic localization. In a recent analysis of Pssu-ipt tobacco (Synková et al., 2006), similarly to Sho plants, Table 4. Endogenous contents of free IAA and ABA in leaves and in isolated intact chloroplasts of transgenic 35S:ZOG1 and WT (W38) tobacco Values represent the mean of two replicates (individual measurements are shown in parentheses). 35S:ZOG1 IAA Leaves (pmol g1 FW) Leaves (pmol mg1 chlorophyll) Chloroplasts (pmol mg1 chlorophyll) ABA Leaves (pmol g1 FW) Leaves (pmol mg1 chlorophyll) Chloroplasts (pmol mg1 chlorophyll) 88.8 (102.1; 75.5) 77.2 (88.8; 65.6) 1.5 (1.8; 1.1) 528.1 (662.4; 393.8) 459.2 (575.9; 342.4) 13.1 (20.9; 5.2) WT (W38) 44.2 (38.4; 50.0) 24.0 (20.8; 27.1) 1.0 (0.5; 1.4) 523.3 (566.9; 479.6) 284.0 (307.7; 260.3) 10.1 (15.0; 5.2) Fig. 6. Transmission electron micrographs of representative chloroplast cross-sections taken from the intact leaves of tobacco expressing the Sho gene and four 35S enhancer elements (A, C) with detail of crystalloid (B) and control SR1 (D) tobacco growing under sterile conditions; and chloroplast cross-section detail taken from the leaf of transgenic dexamethasone-induced Sho (E) and control SR1 (F) tobacco cultivated in a greenhouse. C, crystalloid; CW, cell wall; GT, granal thylakoids; IT, intergranal thylakoids; M, mitochondrion; P, plastoglobuli; S, stroma; SI, starch inclusion. Bar¼2 lm in A, C and D; and 500 nm in B, E and F. 71 Cytokinins in chloroplasts of transgenic tobacco 9 of 13 Fig. 7. Transmission electron micrographs of representative chloroplast cross-sections taken from the intact leaves of transgenic 35S:AtCKX3 (A) and control SNN (B), transgenic 35S:Zm-p60.1 (C) and control SR1 (D), transgenic 35S:ZOG1 (E) and control W38 (F) tobacco. CW, cell wall; GT, granal thylakoids; IT, intergranal thylakoids; M, mitochondrion; P, plastoglobuli; S, stroma; SI, starch inclusion. Bar¼2 lm. CK levels increased in chloroplasts to much lower levels compared with leaf extracts, indicating that the main CK pool accumulates outside chloroplasts in both Z- and iPtype CK-overproducing plants. The content of IAA in leaves of Sho-expressing plants did not differ significantly from that in WT, whereas the chloroplasts contained a lower level of IAA than WT chloroplasts. In ipt tobacco, Eklöf et al. (1997, 2000) found lower levels of IAA and reduced rates of IAA synthesis and turnover. This has led to the suggestion that CKs might down-regulate IAA levels. Recently, a decrease in auxin biosynthesis following an increase in CK levels achieved by induction of ipt was observed in Arabidopsis, however, with a significant delay. Thus, the decrease was hypothesized to be mediated through an altered development rather than through a direct CK–auxin cross-talk (Nordström et al., 2004). A higher ABA content was found in chloroplasts and in leaves (when expressed per FW) of Sho plants compared with WT. After overexpression of bacterial ipt, both a decrease (Synková et al., 1999; Chang et al., 2003) and an increase (Macháčková et al., 1997) in ABA content was reported. Plants expressing CK oxidase/dehydrogenase AtCKX3 The CK oxidase/dehydrogenase selectively cleaves unsaturated N6 side chains of Z, iP, and their corresponding ribosides, while CK phosphates, O-glucosides, and CKs with saturated side chains are not substrates for CK oxidase/dehydrogenase (Armstrong, 1994). It was shown that overexpression of AtCKX genes in tobacco and Arabidopsis plants resulted in a reduced content of endogenous CKs and a strongly altered phenotype, with dwarfed shoot habit, enhanced root growth, and delayed flowering and senescence (Werner et al., 2001, 2003). More abundant Z-derived CKs were more strongly reduced than were iP derivatives in 35S:AtCKX1 and 35S:AtCKX2 Arabidopsis, in contrast to 35S:AtCKX1 and 35S:AtCKX2 tobacco where the more profound changes were in iP-type CKs. To our knowledge, the CK content has never been measured in any 35S:AtCKX3 transgenic plant. These plants are phenotypically identical to 35S:AtCKX1 and the enzyme is ultimately targeted to the same compartment, the vacuole, as proved by in planta experiments exploring an AtCKX–green fluorescent protein (GFP) fusion. Nevertheless, the apparent Km value for 72 10 of 13 Polanská et al. Lower IAA levels were determined in leaves as well as isolated chloroplasts of Zm-p60.1 plants compared with WT. This agrees with previous results, where a steeper fall in IAA gradient from high to low from youngest leaves downward, and a decreased IAA content in the apex and first internodes in Zm-p60.1 tobacco plants was found (Kiran et al., 2006). CK effects might be partly mediated by elevation of ethylene biosynthesis (Genkov et al., 2003). Ethylene treatment was shown to stimulate the oxidative decarboxylation of IAA (Winer et al., 2000). Thus, the elevated contents of active CKs in Zm-p60.1 plants could lead to lower levels of free IAA. ABA measurements revealed a lower ABA content in Zm-p60.1 leaves than in WT. This finding is in contrast to those of a previous study (Kiran et al., 2006), when the tendency of increased ABA accumulation especially in older leaves was observed. AtCKX3 against iP was 14 times higher than that for AtCKX1 (Werner et al., 2003). It has been shown that the ectopic expression of AtCKX3 reduced the CK levels in leaves of adult tobacco plants and also influenced the CK levels in chloroplasts. In leaves, the most strongly reduced CK metabolites were N-glucosides. It is obvious that after a tremendous increase of the activity of an important CK catabolic pathway (side chain cleavage by oxidase/dehydrogenase), the activity of the other CK deactivation pathway (N-glucosylation) was suppressed. In young tobacco seedlings with very low N-glucosylation activity, oxidase/dehydrogenase overexpression had only a marginal effect on the level of N-glucosides (Werner et al., 2001). This discrepancy could be explained by the different developmental stages of the analysed tobacco plants. In 35S:AtCKX1 and 35S:AtCKX2 Arabidopsis seedlings, Z9G and iPG were reduced compared with WT. No significant differences were observed between the IAA levels, expressed on a FW basis, in leaves of adult AtCKX3 and WT tobacco plants. This finding contrasts with the results of Werner et al. (2003), who found reduced levels of IAA in 35S:AtCKX1 and 35S:AtCKX2 Arabidopsis seedlings. However, the authors speculate that CKs need not directly regulate IAA metabolism, but that the different tissue composition, i.e. reduced size of the shoot apical meristem with fewer meristematic cells and reduced cell production in leaves, might lead to a lowering of IAAproducing shoot tissue in transgenic plants. A lower ABA level was found in AtCKX3 leaves compared with WT leaves, which might be linked to prolonged life span and retarded senescence of transgenic plants. Brugière et al. (2003) demonstrated induction of maize CKX1 gene expression in leaf discs by ABA, suggesting a role for this hormone in lowering CK concentrations under different abiotic stresses. Plants expressing trans-zeatin-O-glucosyltransferase ZOG1 Tobacco transformed with the ZOG1 gene from P. lunatus under the control of the CaMV 35S promoter was constructed by Martin et al. (2001a). They found a strongly elevated t-ZOG content with no substantial changes in the level of other CK metabolites. In the present study, an order of magnitude higher accumulation of ZOG in ZOG1 leaves was detected, which may have been caused by the analysis of older material in the present case and different cultivation conditions. The elevated CK-O-glucoside concentrations in bulk leaf tissue did not significantly influence the content of these metabolites in intact chloroplasts isolated from the transgenic plants. This indicates that CK-O-glucosides are accumulated mainly outside chloroplasts, which agrees with the previous observation that transgenic plants overexpressing Zm-p60.1 can accumulate high levels of t-ZOG when grown on a medium supplemented with t-Z despite high Zm-p60.1 b-glucosidase activity in chloroplasts (Kiran et al., 2006). The vacuole has been proposed as the main storage compartment for CK-O-glucosides, based on feeding experiments in Chenopodium rubrum (Fußeder and Ziegler, 1988). A 2-fold higher IAA content was observed in ZOG1 leaves compared with WT. Martin et al. (2001a) reported that IAA does not serve as a substrate for zeatin-Oglucosyltransferase and therefore excluded changes in IAA glucosylation as triggering the observed variations in morphology of transformants (e.g. aerial roots, multiple shoots, more compact stature than WT). Plants expressing b-glucosidase Zm-p60.1 Maize b-glucosidase Zm-p60.1 was the first plant enzyme shown to release active CKs from CK-O- and N3-glucosides in vivo (Brzobohatý et al., 1993). Recently, it was shown that ectopic overexpression of Zm-p60.1 can perturb CK homeostasis in transgenic tobacco (Kiran et al., 2006). Higher levels of CK bases and ribosides in upper leaves and internodes of transgenic plants were found. The present results are consistent with those of the previous report. Moreover, it was found that a tendency to higher accumulation of CK bases and ribosides is also apparent in chloroplasts. The dramatic decrease in CK-Oglucoside levels reported here for chloroplasts isolated from Zm-p60.1 plants corresponded well with the plastid/ chloroplast location of Zm-p60.1 (Kristoffersen et al., 2000) and a strong decrease in apparent Km for t-ZOG when the values obtained for the enzyme purified in vitro and present in isolated chloroplasts were compared (Kiran et al., 2006). Ultrastructure observations Considerable changes were observed in the chloroplast ultrastructure of tobacco constitutively expressing the Sho gene, including an increased number of plastoglobuli and the appearance of crystalloids in plants grown in vitro. 73 Cytokinins in chloroplasts of transgenic tobacco 11 of 13 These crystalloids were not found in other CK metabolism transformants analysed in this work and their corresponding WTs. In another type of transgenic tobacco, with strongly elevated endogenous CKs, Pssu-ipt, anomalies in chloroplast ultrastructure together with crystalloid occurrence were also found, and the distinct crystalline structures were hypothesized to be formed by 2D crystals of light-harvesting complex proteins (Synková et al., 2006). When dexamethasone-induced Sho plants were analysed, increased grana stacking was noticed. Likewise, an increase in grana stacking was observed in chloroplasts of transgenic ipt tobacco (Čatský et al., 1993) and after CK treatment (Wilhelmová and Kutı́k, 1995; Salopek-Sondi et al., 2002). Except for Sho tobacco, marked changes in the ultrastructure of chloroplasts were not found in the other CK transformants analysed. Only a difference in starch accumulation, associated with a change in chloroplast shape, was perceived. This finding is in accordance with data showing that exogenous CK application may increase starch accumulation in chloroplasts (Wilhelmová and Kutı́k, 1995; Stoynova et al., 1996). The involvement of endogenous CKs in starch formation and amyloplast development was proved by Miyazawa et al. (2002) in tobacco BY-2 cells. In conclusion, a highly non-uniform distribution of CKs was found between chloroplasts and bulk leaf tissue in transgenic tobacco lines with several distinct alterations of CK metabolism, indicating that chloroplasts are relatively independent organelles with respect to the regulation of CK metabolism. Thus, it is evident that the estimation of overall CK content in plant tissue is not sufficient to assess its level at a particular site of action and that different compartmentation of individual CK metabolites should be taken into account. CK glucosides do not usually accumulate in chloroplasts, while active forms of CKs, bases, and ribosides are often found in them. In line with previous reports, the present data suggest a complex nature of mutual interactions between CKs, IAA, and ABA that apparently cannot be described by a simple model based on the current state of the art in this field. Thus, Werner et al. (2003) and Nordström et al. (2004) suggested that the effect of CKs on IAA levels might be mediated through altered development. Very variable data can be found in the literature on the interaction of CKs and ABA. CKs are known to counteract many processes induced by ABA (e.g. senescence or stomata closure). Elevated CK content thus might be balanced by higher ABA levels, which is also the case for Sho plants. However, the CK–ABA cross-talk does not seem to be straightforward. It probably strongly depends on the actual physiological conditions of the studied material including environmental factors and/or tissue specificity. Supplementary data Supplementary data available at JXB online are Table 1 showing the percentage of leaf CKs sequestered to chloroplasts and Table 2 listing the CK contents in leaves and isolated chloroplasts. Acknowledgements We are very grateful to Professor Peter Meyer (Sho), Professor Thomas Schmülling (AtCKX3), and Professors Machteld and David Mok (ZOG1) for kind donation of transgenic material, and to Professor David Morris for critical reading of the manuscript. This work was supported by grant nos 206/03/0369 and 206/06/1306 (Grant Agency of the Czech Republic) and IAA600040612 (Grant Agency of the Academy of Sciences of the Czech Republic). References Abdelghani MO, Suty L, Chen JN, Renaudin JP, Teyssendier de la Serve B. 1991. Cytokinins modulate the steady-state levels of light-dependent and light-independent proteins and mRNAs in tobacco cell suspensions. Plant Science 77, 29–40. Armstrong DJ. 1994. Cytokinin oxidase and the regulation of cytokinin degradation. In: Mok DWS, Mok MC, eds. Cytokinins: chemistry, activity, and function. Boca Raton, FL: CRC Press, 139–154. Arnon DI. 1949. Copper enzymes in isolated chloroplasts: polyphenol oxidase in Beta vulgaris. Plant Physiology 24, 1–15. Benková E, Witters E, Van Dongen W, Kolář J, Motyka V, Brzobohatý B, Van Onckelen HA, Macháčková I. 1999. Cytokinins in tobacco and wheat chloroplasts. Occurrence and changes due to light/dark treatment. Plant Physiology 121, 245–251. Bieleski RL. 1964. The problem of halting enzyme action when extracting plant tissues. Analytical Biochemistry 9, 431–442. Boasson R, Laetsch WM. 1969. Chloroplast replication and growth in tobacco. Science 166, 749–751. Brenner WG, Romanov GA, Köllmer I, Bürkle L, Schmülling T. 2005. Immediate-early and delayed cytokinin response genes of Arabidopsis thaliana identified by genome-wide expression profiling reveal novel cytokinin-sensitive processes and suggest cytokinin action through transcriptional cascades. The Plant Journal 44, 314–333. Brugière N, Jiao S, Hantke S, Zinselmeier C, Roessler JA, Niu X, Jones RJ, Habben JE. 2003. Cytokinin oxidase gene expression in maize is localized to the vasculature, and is induced by cytokinins, abscisic acid, and abiotic stress. Plant Physiology 132, 1228–1240. Brzobohatý B, Moore I, Kristoffersen P, Bako L, Campos N, Schell J, Palme K. 1993. Release of active cytokinin by a bglucosidase localized to the maize root meristem. Science 262, 1052–1054. Čatský J, Pospı́šilová J, Macháčková I, Wilhelmová N, Šesták Z. 1993. Photosynthesis and water relations in transgenic tobacco plants with T-DNA carrying gene 4 for cytokinin synthesis. Biologia Plantarum 35, 393–399. Chang H, Jones ML, Banowetz GM, Clark DG. 2003. Overproduction of cytokinins in petunia flowers transformed with PSAG12-IPT delays corolla senescence and decreases sensitivity to ethylene. Plant Physiology 132, 2174–2183. Chory J, Reinecke D, Sim S, Wasburn T, Brenner M. 1994. A role for cytokinins in de-etiolation in Arabidopsis. Plant Physiology 104, 339–347. 74 12 of 13 Polanská et al. Davey JE, Van Staden J. 1981. Cytokinins in spinach chloroplasts. Annals of Botany 48, 243–246. Dobrev PI, Havlı́ček L, Vágner M, Malbeck J, Kamı́nek M. 2005. Purification and determination of plant hormones auxin and abscisic acid using solid phase extraction and two-dimensional high performance liquid chromatography. Journal of Chromatography A 1075, 159–166. Dobrev PI, Kamı́nek M. 2002. Fast and efficient separation of cytokinins from auxin and abscisic acid and their purification using mixed-mode solid-phase extraction. Journal of Chromatography A 950, 21–29. Eklöf S, Åstot C, Blackwell J, Moritz T, Olsson O, Sandberg G. 1997. Auxin–cytokinin interactions in wild-type and transgenic tobacco. Plant and Cell Physiology 38, 225–235. Eklöf S, Åstot C, Sitbon F, Moritz T, Olsson O, Sandberg G. 2000. Transgenic tobacco plants co-expressing Agrobacterium iaa and ipt genes have wild-type hormone levels but display both auxin- and cytokinin-overexpressing phenotypes. The Plant Journal 23, 279–284. Fubeder A, Ziegler P. 1988. Metabolism and compartmentation of dihydrozeatin exogenously supplied to photoautotrophic suspension cultures of Chenopodium rubrum. Planta 173, 104–109. Genkov T, Vágner M, Dubová J, Malbeck J, Moore I, Brzobohatý B. 2003. Increased ethylene production can account for some of the phenotype alterations accompanying activation of a cytokinin biosynthesis gene, during germination and early seedling development in tobacco. In: Vendrell M, Klee H, Pech JC, Romojaro F, eds. Biology and biotechnology of the plant hormone ethylene, III. Amsterdam: IOS Press, 142–146. Hou B, Lim EK, Higgins GS, Bowles DJ. 2004. N-Glucosylation of cytokinins by glycosyltransferases of Arabidopsis thaliana. Journal of Biological Chemistry 279, 47822–47832. Kakimoto T. 2001. Identification of plant cytokinin biosynthesis enzymes as dimethylallyl diphosphate: ATP/ADP isopentenyltransferases. Plant and Cell Physiology 42, 677–685. Kärkönen A, Simola LK. 1999. Localization of cytokinins in somatic and zygotic embryos of Tilia cordata using immunocytochemistry. Physiologia Plantarum 105, 356–366. Kasahara H, Takei K, Ueda N, Hishiyama S, Yamaya T, Kamiya Y, Yamaguchi S, Sakakibara S. 2004. Distinct isoprenoid origins of cis- and trans-zeatin biosyntheses in Arabidopsis. Journal of Biological Chemistry 279, 14049–14054. Kiran NS, Polanská L, Fohlerová R, et al. 2006. Ectopic overexpression of the maize b-glucosidase Zm-p60.1 perturbs cytokinin homeostasis in transgenic tobacco. Journal of Experimental Botany 57, 985–996. Kristoffersen P, Brzobohatý B, Hohfeld I, Bako L, Melkonian M, Palme K. 2000. Developmental regulation of the maize Zm-p60.1 gene encoding a b-glucosidase localized to plastids. Planta 210, 407–415. Kusnetsov VV, Oelmüller R, Sarwatt MI, Porfirova SA, Cherepneva GN, Herrmann RG, Kulaeva ON. 1994. Cytokinins, abscisic acid and light affect accumulation of chloroplast proteins in Lupinus luteus cotyledons without notable effect on steady-state mRNA levels. Planta 194, 318–327. Latzko E, Gibbs M. 1968. Distribution and activity of enzymes of the reductive pentose phosphate cycle in spinach leaves and chloroplasts isolated by different methods. Zeitschrift für Pflanzenphysiologie 59, 184–194. Longo GP, Pedretti M, Rossi G, Longo CP. 1979. Effect of benzyladenine on the development of plastids and microbodies in excised watermelon cotyledons. Planta 145, 209–217. Lyukevich TV, Kusnetsov VV, Karavaiko NN, Kulaeva ON, Selivankina SY. 2002. The involvement of the chloroplast zeatin-binding protein in hormone-dependent transcriptional con- trol of chloroplast genome. Russian Journal of Plant Physiology 49, 92–98. Macháčková I, Sergeeva L, Ondřej M, Zaltsman O, Konstantinova T, Eder J, Ovesná J, Golyanovskaya S, Rakitin Y, Aksenova N. 1997. Growth pattern, tuber formation and hormonal balance in in vitro potato plants carrying ipt gene. Plant Growth Regulation 21, 27–36. Martin RC, Mok MC, Mok DWS. 1993. Cytolocalization of zeatin O-xylosyltransferase in Phaseolus. Proceedings of the National Academy of Sciences, USA 90, 953–957. Martin RC, Mok DWS, Smets R, Van Onckelen HA, Mok MC. 2001a. Development of transgenic tobacco harbouring a zeatin Oglucosyltransferase gene from Phaseolus. In Vitro Cellular and Developmental Biology-Plant 37, 354–360. Martin RC, Mok MC, Habben JE, Mok DWS. 2001b. A maize cytokinin gene encoding an O-glucosyltransferase specific to ciszeatin. Proceedings of the National Academy of Sciences, USA 98, 5922–5926. Milborrow BV, Lee H-S. 1998. Endogenous biosynthetic precursors of (+)-abscisic acid. VI. Carotenoids and ABA are formed by the ‘non-mevalonate’ triose-pyruvate pathway in chloroplasts. Australian Journal of Plant Physiology 25, 507–512. Miller CO, Skoog F, Okumura FS, Von Saltza MH, Strong FM. 1956. Isolation, structure and synthesis of kinetin, a substance promoting cell division. Journal of the American Chemical Society 78, 1375–1380. Miyazawa Y, Kato H, Muranaka T, Yoshida S. 2002. Amyloplast formation in cultured tobacco BY-2 cells requires a high cytokinin content. Plant Cell Physiology 43, 1534–1541. Motyka V, Vaňková R, Čapková V, Petrášek J, Kamı́nek M, Schmülling T. 2003. Cytokinin induced upregulation of cytokinin oxidase activity in tobacco includes changes in enzyme glycosylation and secretion. Physiologia Plantarum 117, 11–21. Murashige T, Skoog F. 1962. A revised medium for rapid growth and bioassays with tobacco tissue culture. Physiologia Plantarum 15, 473–497. Nordström A, Tarkowski P, Tarkowska D, Norbaek R, Åstot C, Doležal K, Sandberg G. 2004. Auxin regulation of cytokinin biosynthesis in Arabidopsis thaliana: a factor of potential importance for auxin–cytokinin-regulated development. Proceedings of the National Academy of Sciences, USA 101, 8039–8044. Ohmiya A, Hayashi T. 1992. Immuno-gold localization of IAA in leaf cells of Prunus persica at different stages of development. Physiologia Plantarum 85, 439–445. Parthier B. 1989. Hormone-induced alterations in plant gene expression. Biochemie und Physiologie der Pflanzen 185, 289–314. Redig P, Schmülling T, Van Onckelen H. 1996. Analysis of cytokinin metabolism in ipt transgenic tobacco by liquid chromatography–tandem mass spectrometry. Plant Physiology 112, 141–148. Reynolds ES. 1963. The use of lead citrate at high pH as an electron-opaque stain in electron microscopy. Journal of Cell Biology 17, 208–212. Sakakibara H, Kasahara H, Ueda N, et al. 2005. Agrobacterium tumefaciens increases cytokinin production in plastids by modifying the biosynthetic pathway in the host plant. Proceedings of the National Academy of Sciences, USA 102, 9972–9977. Sakakibara H. 2006. Cytokinins: activity, biosynthesis, and translocation. Annual Review of Plant Biology 57, 431–449. Salopek-Sondi B, Kovač M, Prebeg T, Magnus V. 2002. Developing fruit direct post-floral morphogenesis in Helleborus niger L. Journal of Experimental Botany 53, 1949–1957. Šámalová M, Brzobohatý B, Moore I. 2005. pOp6/LhGR: a stringently regulated and highly responsive dexamethasone-inducible 75 Cytokinins in chloroplasts of transgenic tobacco 13 of 13 gene expression system for tobacco. The Plant Journal 41, 919–935. Sandberg G, Gardeström P, Sitbon F, Olsson O. 1990. Presence of indole-3-acetic acid in chloroplasts of Nicotiana tabacum and Pinus sylvestris. Planta 180, 562–568. Schmülling T, Schäfer S, Romanov G. 1997. Cytokinins as regulators of gene expression. Physiologia Plantarum 100, 505–519. Sossountzov L, Sotta B, Maldiney R, Sabbagh I, Miginiac E. 1986. Immunoelectron-microscopy localization of abscisic acid with colloidal gold on Lowicryl-embedded tissues of Chenopodium polyspermum L. Planta 168, 471–481. Stetler DA, Laetsch WM. 1965. Kinetin-induced chloroplast maturation in cultures of tobacco tissue. Science 149, 1387–1388. Stoynova EZ, Iliev LK, Georgiev GT. 1996. Structural and functional alterations in radish plants induced by the phenylurea cytokinin 4-PU-30. Biologia Plantarum 38, 237–244. Sun J, Niu Q-W, Tarkowski P, Zheng B, Tarkowska D, Sandberg G, Chua N-H, Zuo J. 2003. The Arabidopsis AtIPT8/ PGA22 gene encodes an isopentenyl transferase that is involved in de novo cytokinin biosynthesis. Plant Physiology 131, 167–176. Synková H, Schnablová R, Polanská L, Hušák M, Šiffel P, Vácha F, Malbeck J, Macháčková I, Nebesářová J. 2006. Three-dimensional reconstruction of anomalous chloroplasts in transgenic ipt tobacco. Planta 223, 659–671. Synková H, Van Loven K, Pospı́šilová J, Valcke R. 1999. Photosynthesis of transgenic Pssu-ipt tobacco. Journal of Plant Physiology 155, 173–182. Takei K, Sakakibara H, Sugiyama T. 2001. Identification of genes encoding adenylate isopentenyltranferase, a cytokinin biosynthesis enzyme in Arabidopsis thaliana. Journal of Biological Chemistry 276, 26405–26410. Vreman HJ, Thomas R, Corse J, Swaminathan S, Murai N. 1978. Cytokinins in tRNA obtained from Spinacia oleracea L. leaves and isolated chloroplasts. Plant Physiology 61, 296–306. Werner T, Motyka V, Laucou V, Smets R, Van Onckelen H, Schmülling T. 2003. Cytokinin-deficient transgenic Arabidopsis plants show multiple developmental alterations indicating opposite functions of cytokinins in the regulation of shoot and root meristem activity. The Plant Cell 15, 2532–2550. Werner T, Motyka V, Strnad M, Schmülling T. 2001. Regulation of plant growth by cytokinin. Proceedings of the National Academy of Sciences, USA 98, 10487–10492. Wilhelmová N, Kutı́k J. 1995. Influence of exogenously applied 6-benzylaminopurine on the structure of chloroplasts and arrangement of their membranes. Photosynthetica 31, 559–570. Winer L, Goren R, Riov J. 2000. Stimulation of the oxidative decarboxylation of indole-3-acetic acid in citrus tissues by ethylene. Plant Growth Regulation 32, 231–237. Winiarska G, Legocka J, Schneider J, Deckert J. 1994. Cytokinin-controlled expression of the apoprotein of the lightharvesting complex of photosystem II in the tissue culture of Dianthus caryophyllus. III. Effect of benzyladenine on the protein synthesis in vivo and on the level of translatable LHCP mRNA. Acta Physiologiae Plantarum 16, 203–210. Zavaleta-Mancera HA, Franklin KA, Ougham HJ, Thomas H, Scott IM. 1999a. Regreening of senescent Nicotiana leaves. I. Reappearance of NADPH-protochlorophyllide oxidoreductase and light-harvesting chlorophyll a/b-binding protein. Journal of Experimental Botany 50, 1677–1682. Zavaleta-Mancera HA, Thomas BJ, Thomas H, Scott IM. 1999b. Regreening of senescent Nicotiana leaves. II. Redifferentiation of plastids. Journal of Experimental Botany 50, 1683–1689. Zubko E, Adams CJ, Macháčková I, Malbeck J, Scollan C, Meyer P. 2002. Activation tagging identifies a gene from Petunia hybrida responsible for the production of active cytokinins in plants. The Plant Journal 29, 797–808. 76 Plant, Cell and Environment (2008) 31, 341–353 doi: 10.1111/j.1365-3040.2007.01766.x The role of cytokinins in responses to water deficit in tobacco plants over-expressing trans-zeatin O-glucosyltransferase gene under 35S or SAG12 promoters MARIE HAVLOVÁ1,2, PETRE I. DOBREV1, VÁCLAV MOTYKA1, HELENA ŠTORCHOVÁ1, JIŘÍ LIBUS1, JANA DOBRÁ1,2, JIŘÍ MALBECK1, ALENA GAUDINOVÁ1 & RADOMIRA VANKOVÁ1 1 Institute of Experimental Botany AS CR, Rozvojova 263, 165 02 Prague 6, Czech Republic and 2Department of Biochemistry, Faculty of Natural Sciences, Charles University, Hlavova 2030/8, 128 43 Prague 2, Czech Republic enable plants to cope with water deficit include fast responses, such as stomatal closure, and longer-term changes in metabolism, growth and development. The latter include the accelerated senescence and abscission of lower leaves, as well as changes in source–sink relationships and the partitioning of dry matter between root and shoot systems (Rodrigues, Pacheco & Chaves 1995). Both fast stress responses and transduction of stress-related signals into changes in specific gene expression involve the mediation by phytohormones. Most studies of the part played by plant hormones in stress physiology have focused on the role of abscisic acid (ABA). An elevation in ABA concentration in plants exposed to various stresses has frequently been reported (see Mansfield & McAinsh 1995). In response to drought, ABA reduces the transpiration rate by inducing stomatal closure. In addition, at least two of five signal transduction pathways known to regulate drought- and cold-inducible genes involve ABA action (Shinozaki & YamaguchiShinozaki 1996; Davies, Kudoyarova & Hartung 2005). Precise control of individual processes involved in responses to stress requires both positive and negative regulation. For example, many ABA-mediated physiological processes induced by water deficit, including closure of the stomata and acceleration of leaf senescence, are counteracted by cytokinins (CKs; for a review, see Pospíšilová & Dodd 2005). CKs increase stomatal aperture and/or delay ABA-induced stomatal closure (Stoll, Loveys & Dry 2000). It has been suggested (Pospíšilová & Dodd 2005) that in longer-term responses to stress, hormones such as ABA and CK may function to regulate the production, metabolism and distribution of metabolites essential for stress survival and recovery. A role for phytohormones in plant stress responses can be inferred from plant reactions to exogenously applied natural hormones or their synthetic analogues. Exogenous CKs have been reported to increase tolerance to mild stress and to speed up recovery (e.g. Itai, Benzioni & Munz 1978). Applied CKs reduce the negative effects of water deficit on chlorophyll and carotenoid content and on photosynthetic ABSTRACT The impact of water deficit progression on cytokinin (CK), auxin and abscisic acid (ABA) levels was followed in upper, middle and lower leaves and roots of Nicotiana tabacum L. cv. Wisconsin 38 plants [wild type (WT)]. ABA content was strongly increased during drought stress, especially in upper leaves. In plants with a uniformly elevated total CK content, expressing constitutively the trans-zeatin O-glucosyltransferase gene (35S::ZOG1), a delay in the increase of ABA was observed; later on, ABA levels were comparable with those of WT. As drought progressed, the bioactive CK content in leaves gradually decreased, being maintained longer in the upper leaves of all tested genotypes. Under severe stress (11 d dehydration), a large stimulation of cytokinin oxidase/ dehydrogenase (CKX) activity was monitored in lower leaves, which correlated well with the decrease in bioactive CK levels. This suggests that a gradient of bioactive CKs in favour of upper leaves is established during drought stress, which might be beneficial for the preferential protection of these leaves. During drought, significant accumulation of CKs occurred in roots, partially because of decreased CKX activity. Simultaneously, auxin increased in roots and lower leaves. This indicates that both CKs and auxin play a role in root response to severe drought, which involves the stimulation of primary root growth and branching inhibition. Key-words: abscisic acid; auxin; cytokinin; drought; tobacco. INTRODUCTION Under the selection pressures resulting from their sessile growth habit in a variable environment, plants have evolved a considerable capacity to cope with adverse environmental conditions. Water availability represents a major limiting factor for plant growth and development. Strategies that Correspondence: R. Vanková. Fax: +420 225 106 446; e-mail: [email protected] © 2008 The Authors Journal compilation © 2008 Blackwell Publishing Ltd 341 77 342 M. Havlová et al. activity (e.g. Metwally, Tsonev & Zeinalov 1997) and improve the recovery of stomatal conductance and net photosynthesis after rehydration (Rulcová & Pospíšilová 2001). Moreover, the transcription of many stress-induced genes is stimulated by CKs (Hare & Van Staden 1997). Modulation of hormone levels may also be achieved by the over-expression of genes coding for their synthesis or metabolism. In the case of CKs, a biosynthetic gene, ipt (coding for the Agrobacterium tumefaciens isopentenyltransferase), has been frequently used. In constitutive transformants, the highly elevated bioactive CK content may itself, unfortunately, lead to substantial changes in plant morphology, which in turn may complicate the interpretation of the potential role of CKs in the response of plants to stress conditions. To minimize this limitation, the ipt gene expression was given under the control of the highly senescence-specific SAG12 promoter (Gan & Amasino 1995). In these transformed plants, the activation of ipt transcription is triggered mainly in senescing leaves, and plants develop normally. SAG12::ipt plants have been used to evaluate the effect of CKs on leaf senescence (Gan & Amasino 1995; Jordi et al. 2000), photosynthesis and nitrogen partitioning (Jordi et al. 2000), flooding tolerance (Zhang et al. 2000; Huynh et al. 2005), seed germination, and response to mild water stress (Cowan et al. 2005; but see further). An alternative approach has been to over-express the gene coding for trans-zeatin O-glucosyltransferase (ZOG1), which results in the elevation of CK O-glucoside content. ZOG1 gene was isolated from Phaseolus lunatus (Martin, Mok & Mok 1999). Tobacco plants transformed with ZOG1 under the control of the constitutive (35S) promoter were prepared by Martin et al. (2001). In 35S::ZOG1 plants, the total CK content (especially of CK O-glucosides) was greatly increased, while the level of bioactive CKs (bases, ribosides) was not changed significantly (Martin et al. 2001). CK O-glucosides represent CK storage forms, as they can be converted to bioactive CKs by the action of b-glucosidases (e.g. Campos et al. 1992). A study of the role of CKs in ontogenetic and stressinduced transitions in the tobacco plants was recently published by Cowan et al. (2005). They proposed that senescence-dependent ontogenetic and water stressinduced transitions in plants may be influenced by CK regulation of sink–source relations. This proposal was based, inter alia, on a comparison of the CK and soluble sugar content and distribution, pigment (chlorophylls and carotenoids) distribution, ABA content, and photosynthetic activity in leaves of different ages in wild-type (WT) plants and plants transformed with the SAG12::ipt gene. However, these authors did not provide data on the CK and soluble sugar composition of water-stressed plants of the two genotypes. Thus, a direct evidence for their hypothesis that CK-mediated, stress-induced changes in sink–source relationships play a role in the plant response to water deficit has been still missing. In the work described here, we sought to clarify the role of CKs in the responses of plants to water deficits.We compared the impact of constitutive and senescence-induced ZOG1 over-expression (using the 35S and SAG12 promoters, respectively) on CK levels, distribution and metabolism, on leaf production and senescence, and on indolyl-3-acetic acid (IAA) and ABA content during the onset and progression of water deficit in tobacco plants and during the early stages of their recovery following rehydration. Measurement of the changes in ABA levels was undertaken because the ABA content is known to correlate well with the degree of stress to which a plant is exposed (Mansfield & McAinsh 1995). Similarly, as auxins and CKs interact in the regulation of many physiological processes, including those likely to be involved in the recovery from stress, the effects of water deficit on IAA levels were also determined. Furthermore, as the transgenic plants accumulate high levels of CK O-glucosides, we also investigated the potential contribution of this readily mobilizable CK reserve to the pool of bioactive CKs following the onset of stress. MATERIALS AND METHODS Plant material and stress application WT tobacco (Nicotiana tabacum L. cv. Wisconsin 38) and transgenic SAG12::ZOG1 and 35S::ZOG1 tobacco plants (for detailed information, see Martin et al. 2001) were sown in sterilized soil (Garden Substrate B; Rašelina, Soběslav, Czech Republic; pot volume 350 mL) without added fertilizer. Plants were raised in a growth chamber (Sanyo MLR 350H; Sanyo, Osaka, Japan) for 6 weeks on a 16 h photoperiod at 1000 mmol m-2 s-1, day/night temperature of 25/23 °C and relative humidity ca. 80%. Water was permanently available to plants. After 6 weeks, half of the plants of each genotype were transferred into another growth chamber under the same light and temperature regimes, but with the relative humidity decreased to 35%. Plants of the same leaf number were selected for the stress experiments in order to exclude the variability within the population (highest in the case of 35S::ZOG1 plants). The surface of the soil was covered with a film impervious to water vapour (Parafilm; Pechiney, Menasha, WI, USA). No further water was applied, and the impact of water deficit on various plant parameters was determined 1 d (‘mild stress’), 6 d (‘moderate stress’) and 11 d (‘severe stress’) after the onset of dehydration. The stressed plants were then watered, and the same parameters were measured 1 d later (‘recovery’). Samples of upper leaves (the two youngest unfolded), lower leaves (the two lowest, still viable) and the middle ones (two in between the upper and lower ones) were collected, cut into pieces and immediately frozen in liquid nitrogen (Linde Gas, Pardubice, Czech Republic) (after removal of the main vein). The roots were shaken to remove the soil, washed briefly with cold tap water (ca. 1 min), dried with filter paper and frozen in liquid nitrogen. For each treatment, a total of three (control) or six (stress) plants were harvested (collected in three independent experiments); the presented results are the means of the three samples. © 2008 The Authors Journal compilation © 2008 Blackwell Publishing Ltd, Plant, Cell and Environment, 31, 341–353 78 Role of cytokinins in responses of tobacco to water deficit 343 (a) RWC was measured using individual cut leaves (upper and middle ones), which were weighed [fresh mass (FM)], saturated with water in the beakers [water saturated mass (SM)] and dried at 88 °C [dry mass (DM)]. RWC was counted as: RWC (%) = 100 - [(SM - FM)/(SM - DM)] ¥ 100 (Salisbury & Ross 1992). 12 10 150 8 6 100 4 50 2 Leaf number 200 Determination of relative water content (RWC) 14 WT RNA extraction and reverse transcription 0 200 Total RNA from the leaves or roots was extracted by means of the RNeasy Plant Kit (Qiagen, Hilden, Germany) and treated with DNAse I on column. One microgram of RNA and random hexamer oligonucleotides (4 mg) or oligo dT primers (500 ng) were heated 5 min at 65 °C, chilled on ice and mixed with buffer, 0.5 mL of Protector RNase Inhibitor (Roche, Mannheim, Germany), 2 mL of 10 mM dNTPs and 10 units of Transcriptor Reverse Transcriptase (Roche). The first strand cDNA was synthesized at 55 °C for 30 min, which was preceded by 10 min at 25 °C if random hexamers were used. 14 SAG:ZOG 12 10 150 8 6 100 4 50 Leaf number (b) 2 0 0 200 (c) 35S:ZOG 14 qRT-PCR 12 First strand cDNA was 10¥ or 20¥ diluted and qRT-PCR was performed using the FastStart DNA MasterPLUS SYBR Green I Kit (Roche) in a final volume of 10 mL with 300 nM of each of the high-performance liquid chromatography (HPLC)-purified primers (Metabion, Plantegg, Martinsried, Germany).The LightCycler 1.2 (Roche) was programmed as follows: 10 min of initial denaturation at 95 °C, then 45 cycles for 10 s at 95 °C, 4 s at the appropriate annealing temperature (Supplementary Table S1; annealing temperature determined by experiment), and 7 s at 72 °C. cDNA derived from calibrator RNA was included into each LighCycler run to correct for run-to-run differences. PCR efficiencies (Supplementary Table S1) were estimated from the calibration curves generated from the serial dilution of cDNAs. The calibrator normalized relative ratio of the relative amount of the target and reference gene was calculated as follows: 10 150 8 6 100 4 50 Leaf number % of plant initial weight 0 2 0 0 1 3 5 7 9 11 1 3 5 7 9 11 Control Drought Time (days) Figure 1. Time course of the water loss and leaf number (•) in control and drought-stressed (a) wild-type (WT), (b) SAG12::ZOG1 and (c) 35S::ZOG1 plants. The plants were cultivated in soil (pot size 350 mL). The age of the plants was 6 weeks at the beginning of the experiment. Results from three independent experiments are presented (mean values ⫾ SE are shown). ET C pT (C )− C pT (S) × ER C pR (S)− C pR (C ) where ET/ER is the efficiency of target/reference amplification, CpT/CpR is the cycle number at the target/reference detection threshold (crossing point), S is the sample and C is the calibrator. ZOG1 transcript levels were normalized against Act9 (Volkov, Panchuk & Schoffl 2003) and 18S rRNA (in case of random primed cDNA). The degree of stress was followed via the decrease in soil moisture (estimated as a loss of pot weight). In the stressed plants, the greatest loss of water occurred during the first 24 h (Fig. 1). The weight of the control pots did not change significantly during the experiments. The time course of fresh weight/dry weight (FW/DW) ratio (mean values) was: control plants – all leaves ca. 10.6, roots 5.6; mild stress – all leaves 9.2, roots 5.3; moderate stress – leaves (upper, middle, lower) 6.8, 6.8, 5.6, roots 4.9; severe stress – leaves 5.3, 5.0, 4.0, roots 4.2; rehydration – leaves 7.9, 8.5, 8.0, roots 4.6. Plant hormone extraction and purification Phytohormones were extracted and purified according to Dobrev & Kaminek (2002). For analyses of endogenous CKs, 50 pmol of each of the following 17 deuteriumlabelled standards were added - [2H5]Z, [2H5]ZR, [2H5]Z7G, [2H5]Z9G, [2H5]ZOG, [2H5]ZROG, [2H6]iP, [2H6]iPR, © 2008 The Authors Journal compilation © 2008 Blackwell Publishing Ltd, Plant, Cell and Environment, 31, 341–353 79 344 M. Havlová et al. [2H6]iP7G, [2H6]iP9G, [2H3]DHZ, [2H3]DHZR, [2H3]DHZ9G, [2H7]DHZOG, [2H5]ZMP, [2H3]DHZMP and [2H6]iPMP (Apex Organics, Honiton, UK). cis-Zeatin derivatives were determined using the retention times and mass spectroscopy (MS) spectra of unlabelled standards and the response ratio of their trans-zeatin counterparts. The term ‘(total) CK pool’ is used for the sum of all isoprenoid CK metabolites. Tritiated internal standards were used for the determination of auxin (3[5(n)-3H] IAA;Amersham Pharmacia Biotech, Little Chalfont, Bucks, UK; specific activity 74 GBq/mmol, 5 ¥ 103 Bq) and ABA (Amersham; specific activity 1.74 TBq/mmol, 5 ¥ 103 Bq). (anova) and Tukey’s post hoc test implemented in XLSTAT (Addinsoft, Deutschland, Andernach, Germany) or t-test as implemented in Microsoft Excel. RESULTS Effects of water deficit on leaf formation and leaf loss In well-watered plants, leaf number increased throughout the experiment, from ca. 8 to 12–13 leaves per plant by day 12. The rate of leaf formation did not differ significantly among the genotypes tested (Fig. 1). Accelerated senescence of the lowest leaves induced by the water stress treatment resulted in a decrease in viable leaf number in all genotypes. This started around day 7, and by the end of the stress period (11 d from start) the water-stressed plants had lost, on average, around three true leaves (Fig. 1). In transformed plants constitutively expressing ZOG1 (35S::ZOG1), leaf loss started about 1 d earlier than in WT and SAG::ZOG1 transformants (Fig. 1). Within 24 h of rehydration, leaf formation resumed and the number of leaves per plant in WT and SAG::ZOG1-transformed plants began to increase (Fig. 1); the most rapid recovery occurring in the SAG::ZOG1 plants. Recovery of leaf formation in water-stressed 35S::ZOG1 plants was barely detectable after 24 h of rehydration. In the stress treatments, the RWC of leaves decreased with the duration of the stress conditions (Fig. 2), with RWC falling by ca. 20% at day 6 (moderate stress) and by ca. 40% at day 11 (severe stress). One day after rehydration, RWC increased in all water-stressed plants, almost to the level observed in the control plants (Fig. 2). HPLC of auxin and ABA IAA and ABA were determined using two-dimensional HPLC according to Dobrev et al. (2005). IAA was quantified using a fluorescence detector LC 240 (Perkin Elmer, Wellesley, MA, USA). Quantification of ABA was performed on the basis of ultraviolet (UV) detection using a diode array detector 235C (Perkin Elmer). The method was verified by comparison with GC-MS. HPLC/MS LC-MS analysis was performed as described by Dobrev et al. (2002) using an Ion Trap Mass Spectrometer Finnigan MAT LCQ-MSn (Thermo Fisher, Waltham, MA, USA) equipped with an electrospray interface. Detection and quantification were carried out using a Finnigan LCQ operated in the positive ion, full-scan MS/MS mode using a multi-level calibration graph with deuterated CKs as internal standards.The levels of 27 different CK derivatives were measured. The detection limit was calculated for each compound as 3.3 s/S, where s is the standard deviation of the response and S is the slope of the calibration curve. Three independent experiments were performed. Each sample was injected at least twice. Quantification of ZOG1 expression in transgenic plants In order to determine the extent of ZOG1 expression in transgenic plants, the amount of ZOG1 mRNA was quantified by real-time RT-PCR in the individual leaves and roots of the control (well watered) and drought-treated plants. Normalization of ZOG1 expression to actin (Act9) and 18S rRNA led to similar results. No specific product was amplified using ZOG1 primers from WT cDNAs. The highest expression levels of ZOG1 were found in SAG::ZOG1 transformants in lower (older) leaves (Table 1). ZOG1 expression increased dramatically in lower leaves following the onset of water deficit, and it continued to increase throughout the stress period. SAG promoter activity sharply decreased after rehydration (Table 1). The impact of moderate stress (6 d) on the pattern of SAG::ZOG1 expression is shown in more detail in Fig. 3. Considerable stimulation of SAG promoter activity started in the fourth leaf (numbered from the base) of the drought-stressed plants and in the fifth leaf of the control plants. Physiologically, these two leaf positions approximately corresponded to each other, as the stressed Determination of CKX activity The enzyme preparations were extracted and partially purified using the method described by Motyka et al. (2003). The CKX activity was determined by in vitro assays based on the conversion of [2-3H]isopentenyladenine (prepared by Dr Jan Hanuš, Institute of Experimental Botany AS CR, Prague, Czech Republic) to [3H]adenine. Separation of the substrate from the product was achieved by HPLC. The CKX activity was determined in two independent experiments. As the data from both experiments showed the same tendencies (but different absolute values), the results of one representative experiment (mean value of three parallel assays for each of three replicate protein preparations) are presented. Statistical analysis The significance of the differences among the measured values was assessed by a four-factor analysis of variance © 2008 The Authors Journal compilation © 2008 Blackwell Publishing Ltd, Plant, Cell and Environment, 31, 341–353 80 Role of cytokinins in responses of tobacco to water deficit 345 RWC (%) Moderate 100 80 80 60 60 40 40 20 20 0 0 100 Severe Recovery 100 80 80 60 60 40 40 20 20 RWC (%) Middle leaves – control Upper leaves – drought Middle leaves – drought Mild 100 Figure 2. Relative water content (RWC) during the drought stress progression in wild-type (WT), SAG12::ZOG1 (SAG) and 35S::ZOG1 (35S) plants. Mild stress: 1 d dehydration; moderate stress: 6 d dehydration; severe stress: 11 d dehydration; recovery: subsequent 1 d rehydration. C, well-watered plants; D, drought-stressed plants. 0 0 C D WT C D SAG C D 35S C D WT C D SAG C D 35S Stress Table 1. SAG::ZOG1 expression after mild, moderate and severe drought stress and subsequent rehydration quantified by real-time RT-PCR Recovery Mild (1 d) Moderate (6 d) ZOG1 expression (relative units) Severe (11 d) 1d Leaves Upper Middle Lower 0.0006 0.0261 0.9792 0.0025 1.2087 2.3478 0.0002 0.0089 0.0041 0.0036 0.0283 1.7981 ZOG1 expression levels were normalized to actin (Act9). Results from two distinct experiments on the basis of triplicate determination were combined. ZOG1, trans-zeatin O-glucosyltransferase gene. ZOG1 expression (relative units) plants had three leaves less than the control ones at the time of measurement because of drought-induced growth cessation. Different levels of ZOG1 expression driven by the constitutive 35S promoter were detected under drought and control conditions (Fig. 3). Analysis of individual leaves of two plants grown under moderate and severe stress and two control plants revealed that ZOG1 transcript level was significantly higher (more than double) in drought-stressed plants than in the watered ones (t-test, P = 0.0001). As no correlation between 35S::ZOG1 expression and leaf position was observed, the mean values ⫾ SE for control and stressed leaves are presented. SAG12::ZOG1 – control SAG12::ZOG1 – drought 35S::ZOG1 – control 35S::ZOG1 – drought 12 10 8 6 4 2 0 1 1 2 2 3 4 5 6 7 8 Leaf position (drought) 3 9 4 5 6 7 8 9 10 11 12 Leaf position (control) CK content Figure 3. Quantitation of SAG12::ZOG1 and 35S::ZOG1 CK content has been expressed on a DW basis in order to eliminate the difference caused by different water contents between control and stressed plants. CK levels in the upper, middle and lower leaves and roots of WT, SAG::ZOG1 and 35S::ZOG1 plants under control and different drought conditions are shown in Fig. 4. Individual CK metabolites were assigned to one of seven groups according to their structure and function. These groups included bioactive CKs, CK expression in plants under control and moderate drought stress conditions (6 d dehydration) by real-time RT-PCR. ZOG1 expression levels were normalized to actin (Act9). As no correlation between 35S::ZOG1 expression and leaf position was observed, mean values ⫾ SE for the control and stressed leaves are presented. Results from two distinct experiments on the basis of triplicate determination were combined. © 2008 The Authors Journal compilation © 2008 Blackwell Publishing Ltd, Plant, Cell and Environment, 31, 341–353 81 346 M. Havlová et al. cis-Z(R) Active CK N-glc cis-N-glc O-glc cis-O-glc Phosphates Mild 1000 800 Moderate cis-Z(R) Active CK N-glc cis-N-glc O-glc cis-O-glc Phosphates 1200 1000 800 600 600 400 400 200 200 0 1200 cis-Z(R) Active CK N-glc cis-N-glc O-glc cis-O-glc Phosphates Severe 1000 800 Recovery 0 1200 cis-Z(R) Active CK N-glc cis-N-glc O-glc cis-O-glc Phosphates 1000 800 600 600 400 400 200 200 0 UML R UML R UML R UML R UML R UML R C D WT C D C SAG UML R UML R UML R UML R UML R UML R D C 35S D C WT D SAG C Total CK content (pmol g–1 DW) Total CK content (pmol g–1 DW) 1200 0 D 35S Figure 4. Endogenous cytokinin (CK) levels in tobacco leaves and roots during the drought stress progression in wild-type (WT), SAG12::ZOG1 (SAG) and 35S::ZOG1 (35S) plants. CK metabolites were summarized into seven groups: (1) cis-zeatin (riboside) [cis-Z(R)]; (2) bioactive CKs (trans-zeatin, isopentenyladenine, dihydrozeatin and the corresponding ribosides); (3) N-glucosides (7N- and 9N-glucosides of trans-zeatin, isopentenyladenine and dihydrozeatin) (N-glc); (4) cis-N-glucosides (cis-zeatin 7N- and 9N-glucoside) (cis-N-glc); (5) O-glucosides [trans-zeatin (riboside) O-glucoside and dihydrozeatin (riboside) O-glucoside] (O-glc); (6) cis-O-glucosides [cis-zeatin (riboside) O-glucoside] (cis-O-glc); and (7) CK phosphates (trans- and cis-zeatin-, isopentenyladenine- and dihydrozeatin riboside phosphates). In the grouped bars, the first bar corresponds to the upper leaves (U), the second one to the middle leaves (M), the third one to the lower leaves (L) and the last one to the roots (R). SE was in the range 5–18%. Other details the same as for Fig. 2. DW, dry weight. deactivation forms (N-glucosides), CK storage forms (O-glucosides) and CK phosphates (CK biosynthesis intermediates). As cis-zeatin and its riboside are recognized by some CK receptors, but lack physiological activity in most CK bioassays, cis-zeatin derivatives have been shown as separate groups (with the exception of cis-zeatin phosphate, which was included with other CK phosphates). As the bioactive CKs represent only a small portion of the total Actie CK content (pmol g–1 DW) iP iPR DHZ DHZR Z ZR Moderate iP iPR DHZ DHZR 25 20 20 15 15 10 10 5 5 0 0 25 Severe Z ZR iP iPR DHZ DHZR Recovery Z ZR iP iPR DHZ DHZR 25 20 20 15 15 10 10 5 5 0 UML R UML R UML R C D WT UML R UML R UML R C D SAG C D 35S UML R UML R C D WT UML R C UML R UML R D SAG UML R C D 35S 0 Actie CK content (pmol g–1 DW) Z ZR 25 Mild CK pool, their contents are shown in detail in Fig. 5. Similarly, the dynamics of cis-zeatin/riboside contents is given in Fig. 6.When expressed on a DW basis, the total CK contents in roots are relatively lower than those in leaves, but per gram FW, they are similar to, or higher than, those in leaves. After mild stress, a significant elevation of bioactive CKs was observed in WT upper leaves (t-test, P = 0.039; Fig. 5). This increase was mainly a result of the elevation of Figure 5. Endogenous bioactive cytokinin (CK) levels (trans-zeatin, isopentenyladenine, dihydrozeatin and the corresponding ribosides) in tobacco leaves and roots during the drought stress progression in wild-type (WT), SAG12::ZOG1 (SAG) and 35S::ZOG1 (35S) plants. Mean values ⫾ SE are given. Other details the same as for Figs 2 and 4. © 2008 The Authors Journal compilation © 2008 Blackwell Publishing Ltd, Plant, Cell and Environment, 31, 341–353 82 Role of cytokinins in responses of tobacco to water deficit 347 Upper leaves Middle leaves Lower leaves Roots Mild Moderate 25 20 20 15 10 10 5 0 0 30 Severe Recovery 30 20 20 10 10 0 C D WT C D SAG C D 35S C D WT C D SAG trans-zeatin/riboside and, to a low extent, also of isopentenyladenine/riboside. The most remarkable increase was monitored in the case of trans-zeatin. No significant changes in bioactive CK levels in upper leaves were observed between stressed and control SAG::ZOG1 or 35S::ZOG1 plants. The levels of cis-zeatin/riboside were not significantly affected in either genotype. Mild stress did not impose any remarkable effect on the size of the total CK pool in either of the tested genotypes. No significant difference was found in the total CKs between WT and SAG::ZOG1 plants. Both control and stressed 35S::ZOG1 plants had considerably higher total CK levels than WT (Tukey’s test, P < 0.0001), mainly as a result of higher levels of O-glucosides, both of the trans-zeatin and the ciszeatin type. In roots, trans-zeatin O-glucosides strongly predominated. Moderate drought stress led to a decreased level of bioactive CKs in the leaves of all stressed plants, the highest effect being imposed on trans-zeatin and its riboside. Nevertheless, a gradient in bioactive CKs in favour of the upper leaves was observed in all tested genotypes, together with a gradient in CK phosphates. The relatively small difference in bioactive CK levels in the middle and lower leaves of stressed SAG::ZOG1 plants might be the consequence of enhanced SAG promoter activity in stressed lower leaves (see Fig. 3). A much higher impact of stimulated SAG promoter activity was, however, observed at the distribution of total CKs in these plants. Besides an increase in total CK level in the upper leaves in response to water deficit, a comparable elevation was found in lower leaves. The increase of total CKs in leaves, caused by both water stress and natural senescence, was found predominantly in ZOG1 transgenics. It was mainly the result of raised levels of the inactive CK forms – CK O- and N-glucosides, especially of the cis-zeatin type. Elevation of free cis-zeatin/ riboside was observed as well. In roots, the levels of CK N- and O-glucosides were generally much lower. However, in the roots of stressed plants, a mild elevation of the bioactive CK content was found. C D cis-Z(R) content (pmol g–1 DW) cis-Z(R) content (pmol g–1 DW) 30 0 35S Figure 6. Levels of cis-zeatin and cis-zeatin riboside [cis-Z(R)] in tobacco leaves and roots during the drought stress progression in wild-type (WT), SAG12::ZOG1 (SAG) and 35S::ZOG1 (35S) plants. Mean values ⫾ SE are given. Other details the same as for Fig. 2. Severely stressed plants were completely wilty. The bioactive CK content decreased in all genotypes, compared with that in moderate stress. Differences between the upper, middle and lower leaves were diminished. The flattening of the CK gradient might indicate that the strength of the stress nearly reached the limits of the plant’s ability to cope with drought. In SAG::ZOG1 plants, SAG promoter activity increased in stressed middle leaves (in comparison with moderate stress), leading to an increase in the total CK content of these leaves comparable with that in the upper ones. In lower leaves, strong SAG expression caused a further elevation of the total CK content (Tukey’s test, P < 0.0001). Besides elevated N-glucoside levels, large increases in cis-zeatin O-glucosides and, to a lesser extent, in trans-zeatin O-glucosides were also detected. In control SAG::ZOG1 plants, the activity of the SAG promoter, driven by natural senescence, resulted in the formation of a gradient of total CKs in favour of the lower leaves. Because of the SAG promoter activity, the levels of bioactive CKs in the lower leaves of control plants were maintained approximately at the levels of the upper ones. In stressed 35S::ZOG1 plants, the levels of bioactive CKs in the upper leaves were higher than those in the middle and lower ones. The large increase in total CKs, especially in the lower leaves, was caused by the increases in both O- and N-glucosides (Tukey’s test, P < 0.0001). In all the genotypes tested, severe stress led to the further accumulation of total CKs in the roots. An increase in bioactive CKs was observed especially in the roots of WT and SAG::ZOG1 plants. Following watering, a decrease in the total CK content was observed in the roots in all tested genotypes. Simultaneously, the level of bioactive CKs in the leaves increased significantly in comparison with severely stressed plants (Tukey’s test, P = 0.007). Considerable elevation was found predominantly in the case of trans-zeatin/riboside. WT plants had an elevated bioactive CK content in the upper and, to a lower extent, also in the middle leaves. New leaf formation commenced within 1 d of renewed water © 2008 The Authors Journal compilation © 2008 Blackwell Publishing Ltd, Plant, Cell and Environment, 31, 341–353 83 348 M. Havlová et al. CKX activity (pmol Ade·mg–1 protein·h–1) 1000 1200 Moderate 1000 800 800 600 600 400 400 200 200 0 1200 Severe 0 1200 Recovery 1000 1000 800 800 600 600 400 400 200 200 0 C D WT C D SAG C D 35S C D WT C D SAG application. SAG::ZOG1 plants exhibited an increase in bioactive CK content in all leaves, and new leaf formation commenced even faster than in WT plants. The level of inactive CKs in SAG::ZOG1 plants fell, which correlated well with the decrease of SAG promoter activity following rehydration (see Table 1); nevertheless, two CK peaks (in upper and lower leaves) were still apparent. 35S::ZOG1 transgenics had relatively high levels of bioactive CKs in all leaves, in which CK O-glucosides had accumulated during stress progression. Considerable decrease of free cis-zeatin/ riboside content was observed after rehydration, especially in ZOG1 transgenics. C D 35S 0 CKX activity (pmol Ade·mg–1 protein·h–1) Upper leaves Middle leaves Lower leaves Roots 1200 Mild Figure 7. The activity of cytokinin oxidase/dehydrogenase (CKX) in tobacco leaves and roots during the drought stress progression in wild-type (WT), SAG12::ZOG1 (SAG) and 35S::ZOG1 (35S) plants. Other details the same as for Fig. 2. control plants. After rehydration, CKX activity rose almost to the level observed in the roots of control plants. ABA The ABA content of the plants was clearly correlated with the degree of applied water stress (Fig. 8). Under mild stress conditions, a slight increase in ABA concentration occurred in the upper and middle leaves of WT plants and in the middle leaves of SAG::ZOG1 plants. ABA levels did not increase in 35S::ZOG1 plants subjected to mild water stress. Under moderate stress, ABA content had risen substantially in all stressed plants, with SAG::ZOG1 plants exhibiting the greatest increase. Severe stress led to a further substantial increase in ABA content (by more than one order of magnitude in comparison with control conditions). The highest concentration of ABA occurred in the upper leaves. Following rehydration, ABA levels began to decrease in all treatments, although after a 1 d recovery period, they remained significantly higher than those observed in control plants. No significant differences in ABA levels between the genotypes were observed during recovery. CKX activity In young, well-watered plants, constitutive ZOG1 expression led to a considerable decrease in CKX activity in leaves in comparison with WT plants (Fig. 7). After moderate stress, an increase in CKX activity in the middle leaves in SAG::ZOG1 plants was observed. In constitutive transformants, a decrease in CKX activity in stressed upper leaves was found. After severe stress, a strong enhancement of CKX activity was observed in the lower leaves of all plants, together with the decrease in CKX activity in the upper leaves, which might contribute to the maintenance of a bioactive CK gradient. After rehydration, a rise in CKX activity in the leaves was observed, presumably to regulate bioactive CK levels after the re-establishment of the transpiration flow. CKX activity was much higher in the roots than in the leaves in all genotypes (Tukey’s test, P < 0.0001). After mild stress, CKX activity in the roots decreased in WT and 35S::ZOG1 plants. After moderate stress, CKX activity fell in the roots of all the tested genotypes.The decreasing trend of CKX activity in the roots was reversed after severe stress, but still, the CKX activity was well below the level in Auxin After 1 d of dehydration, an increased level of free IAA was found in WT lower leaves (Fig. 9). Mild elevation was observed in the roots of both transgenics. In all genotypes, moderate stress resulted in a substantial increase in IAA concentration, both in the lower leaves and roots (Tukey’s test, P = 0.024).After severe stress, the highest IAA concentration was found in the lower leaves of all tested plants (highly significant increase in comparison with the control, Tukey’s test, P < 0.0003). Differences in IAA level between the stressed lower and upper leaves were highly significant © 2008 The Authors Journal compilation © 2008 Blackwell Publishing Ltd, Plant, Cell and Environment, 31, 341–353 84 Role of cytokinins in responses of tobacco to water deficit 349 ABA content (pmol g–1 DW) Moderate 60000 10000 40000 5000 0 80000 20000 Severe 0 15000 Recovery 60000 10000 ABA content (pmol g–1 DW) Upper leaves Middle leaves Lower leaves Roots 15000 Mild Figure 8. Endogenous level of abscisic acid (ABA) in tobacco leaves and roots during the drought stress progression in wild-type (WT), SAG12::ZOG1 (SAG) and 35S::ZOG1 (35S) plants. Mean values ⫾ SE are given. Other details the same as for Fig. 2. 40000 5000 20000 C D C WT D C SAG D C 35S D WT C D SAG Upper leaves Middle leaves Lower leaves Roots IAA content (pmol g–1 DW) 4000 D 0 35S (Tukey’s test, P < 0.0002), between the stressed lower and middle leaves were significant (Tukey’s test, P = 0.002), and between the lower leaves and roots were significant (Tukey’s test, P = 0.008). In all plants, rehydration led to a fall in IAA levels in the lower leaves and roots, and in WT, almost to the control values. In SAG::ZOG1 transgenics, a significantly elevated IAA content was observed in all rehydrated leaves. Natural senescence led to a mild elevation of IAA content in the lower leaves of WT and 35S::ZOG1 plants, with some delay also in SAG::ZOG1 plants. Fast and strong drought treatment (at 35% humidity) led, apart from the elevation of IAA in the roots, to the slight increase of WT root biomass, mainly caused by root elongation and thickening (not shown). Because of the low SAG promoter activity in the roots, only a negligible negative effect on root growth was observed in SAG::ZOG1 plants. The 35S::ZOG1 plants, which already had a longer root 5000 Mild C system under control conditions, exhibited a reduced ability to increase root biomass during the drought; thus, its rootto-shoot ratio was the lowest among the plant types tested (results not shown). DISCUSSION Changes in the distribution of plant hormones in leaves in response to water deficit In plants adequately supplied with water, the concentration of bioactive CKs is normally greatest in the young leaves and declines as the leaves age and start to senesce (e.g. Singh, Letham & Palni 1992). Accordingly, we observed at the beginning of the experiment, when the control plants were juvenile, with no visible signs of senescence, a similar bioactive CK content in all leaves. During the experiment, 5000 Moderate 4000 3000 3000 2000 2000 1000 1000 0 5000 Severe 0 5000 Recovery 4000 4000 3000 3000 2000 2000 1000 1000 0 C D WT C D SAG C D 35S C D WT C D SAG C D IAA content (pmol g–1 DW) 0 0 35S © 2008 The Authors Journal compilation © 2008 Blackwell Publishing Ltd, Plant, Cell and Environment, 31, 341–353 85 Figure 9. Endogenous level of indolyl-3-acetic acid (IAA) in tobacco leaves and roots during the drought stress progression in wild-type (WT), SAG12::ZOG1 (SAG) and 35S::ZOG1 (35S) plants. Mean values ⫾ SE are given. Other details the same as for Fig. 2. 350 M. Havlová et al. simultaneously with the development of the symptoms of senescence (gradual decrease in chlorophyll content of lower leaves), a gradient of bioactive CKs in favour of the upper leaves was observed (Fig. 5). In order to elucidate, at least partially, CK function during the drought stress, we compared the water deficit response of plants with a uniform elevation of CK content before stress initiation and those exhibiting CK elevation localized to senescing tissues. As constitutive overexpression of CK biosynthetic gene results in the severe modification of plant morphology, we decided to manipulate the CK levels in more gentle way using plants overexpressing ZOG1. Constitutive over-expression of ZOG1 had only a minor effect on plant morphology, as well as on the levels of bioactive CKs (see also Martin et al. 2001). However, because of the permanently enhanced CK O-glucosylation, which stimulated the CK turnover, 35S::ZOG1 plants exhibited under control conditions symptoms of CK elevation, i.e. higher transpiration rate and higher net photosynthetic rate (Pospíšilová, personal communication). For the targeted elevation of CK levels, we used the SAG12 promoter. Since the pioneering work of Gan & Amasino (1995) on SAG12::ipt tobacco, this promoter has repeatedly been used for the targeted manipulation of CK content. In accordance with Cowan et al. (2005), who found an increase in trans-zeatin-type CKs in senescing wellwatered SAG12::ipt plants, we found an elevation of CK content in senescing control SAG12::ZOG1 plants. Apart from the expected increase of the level of CK O-glucosides, we also found an elevation of bioactive CKs when SAG promoter activity was stimulated. In order to check the ability of ZOG1 transformants to utilize the elevated CK O-glucoside pool, b-glucosidase activity was compared in WT and ZOG1 expressing plants. Mild elevation of b-glucosidase activity in ZOG1 transformants (results not shown) seemed to prove their capacity to hydrolyse the CK storage pool. Another question arose, whether the stress response of ZOG1 transformants could not be modulated by a massive hydrolysis of the CK O-glucoside pool under the stress conditions. The dynamics of b-glucosidase activity during the stress progression (results not shown), as well as the CK analyses, indicated that CK O-glucoside hydrolysis has been under strict environmental control. In WT plants, the formation of a gradient of bioactive CKs in favour of the upper leaves was observed when mild stress was imposed. The occurrence of this gradient might reflect the necessity for the preferential protection of upper leaves under unfavourable environmental conditions. The importance of CKs in sink–source polarization during mild water stress was recently discussed by Cowan et al. (2005). Constitutive ZOG1 transformants did not exhibit any significant response to mild stress. No increase in ABA content (observed in the WT upper and middle leaves) was detected. Neither the bioactive CK nor the IAA levels were considerably changed. The lack of difference in CK levels between the upper and middle leaves of 35S::ZOG1 plants coincided with the delayed decrease of stomatal aperture and net photosynthetic rate of the middle leaves in response to water deficit (Pospíšilová, personal communication), which is in accordance with the positive effect of CKs on stomatal aperture (e.g. Rulcová & Pospíšilová 2001). Elevated CK content before stress initiation, thus, might postpone the reaction of the plants to adverse environmental conditions. The moderate water deficit (RWC decreased by ca. 20%) had two main effects on the concentrations of bioactive CKs. Firstly, in agreement with Goicoechea et al. (1995), it caused a progressive fall in the concentration of bioactive CKs (especially of trans-zeatin/riboside) in all the leaves of all the genotypes tested. Secondly, the steepness of the apex-to-base gradient in bioactive CKs increased substantially, which was remarkable, especially in 35S::ZOG1 plants. In SAG::ZOG1 plants, moderate stress caused the high stimulation of ZOG1 expression in the lower leaves (Table 1), which resulted in an increase in the total CK levels in these leaves. However, the bioactive CK content in the lower leaves did not differ significantly from that in the middle ones, which indicates that bioactive CK levels are strictly controlled under stress conditions. ABA levels strongly correlated with the degree of stress. This finding agrees with previous reports (e.g. Mansfield & McAinsh 1995; Riccardi et al. 2004). As ABA is an important component of the plant stress defence system, its higher levels, observed in the upper leaves, might reflect their preferential protection. Interestingly, as senescence occurred in lower leaves, these results also indicate that while ABA levels may increase during stress-induced leaf senescence, ABA itself is not the cause of senescence. Under severe stress (RWC decreased by ca. 40%), all the leaves of the tested genotypes were wilty. The bioactive CK gradient was suppressed in the WT and SAG::ZOG1 plants, while it was more profound in the 35S::ZOG1 ones. Nevertheless, the decrease of CKX activity in the upper leaves and, especially, the large increase in CKX activity in the lower leaves seems to indicate a tendency for the gradient of bioactive CKs to be maintained. Large elevation of IAA content, observed in the stressed lower leaves (at mild stress only in WT, at moderate stress in WT and SAG::ZOG1 plants, and at severe stress in all genotypes) and also in the severely stressed middle leaves of 35S::ZOG1 plants, is in agreement with the results of Quirino, Normanly & Amasino (1999), who found an increase in the free auxin levels in senescing Arabidopsis leaves. This increase correlated with the stimulation of the expression of nitrilase 2, an enzyme catalysing the conversion of indole-3-acetonitrile to IAA. Both senescence and pathogen attack have been found to induce nitrilases 2 and 4 in leaves (Bartel & Fink 1994). The function of the increased IAA content of senescing leaves is not clear. As senescence is a strongly regulated process, high IAA levels might be necessary for its control. The flow of auxin to the abscission zone is necessary to prevent leaf abscission. A possible explanation might be that the increased protein breakdown in senescing leaves leads to an increase in the © 2008 The Authors Journal compilation © 2008 Blackwell Publishing Ltd, Plant, Cell and Environment, 31, 341–353 86 Role of cytokinins in responses of tobacco to water deficit 351 content of tryptophan, which may be converted to IAA, in order to facilitate the transport from senescing leaves. Rehydration resulted in an increase in bioactive CK levels (especially of trans-zeatin/riboside) in the leaves of all genotypes, which might be, at least partially, result from the re-establishment of xylem sap flow from the roots. The crucial impact of the transpiration stream on CK transport from the roots was recently nicely demonstrated (Aloni et al. 2006). Mild elevation of CKX activity in leaves might reflect the necessity to control the sudden excess of CKs. The stress indicators (SAG promoter activity and ABA level) exhibited a highly significant decrease. Elevation of IAA levels in the middle leaves (and, in SAG::ZOG1 plants, also in the upper leaves) seems to be related to growth re-establishment. The slightly faster growth initiation in SAG::ZOG1 plants in comparison with WT is in accordance with the known positive effect of CKs on plant recovery (e.g. Itai et al. 1978). The slower recovery of 35S::ZOG1 transgenics indicates that elevated CK content need not be advantageous in case of prolonged, severe drought stress. This conclusion agrees with data on pssu:ipt over-expressing tobacco plants, which had a highly elevated CK content and which were found to be more vulnerable to water deficit (Synková et al. 1999). roots, in contrast to the situation in senescing leaves (Quirino et al. 1999), auxin increase in the stressed roots might be a consequence of auxin polar transport. Increased IAA levels were also observed under desiccation stress in maize roots (Ribaut & Pilet 1994; Xin, Zhou & Pilet 1997). The function of auxin during the drought-stress response of roots remains to be elucidated. Studies on axr1-3 mutants of Arabidopsis revealed that functional auxin signalling has been necessary for the induction of stress-related proteins RD29B and glutamine synthetase in roots, as well as for drought rhizogenesis (Bianchi, Danerval & Vartanian 2002). Increased levels of auxin might cause an additional elevation of CK content, as auxin was found to promote the expression of genes for CK biosynthetic enzymes in roots (namely, AtIPT5 and AtIPT7; Miyawaki, MatsumotoKitano & Kakimoto 2004). An elevated content of both hormones might stimulate root cell division, which resulted, because of the inhibitory effect of CKs on lateral root formation, in primary root elongation. In conclusion, data on SAG12 promoter activity demonstrated that drought strongly accelerated the senescence of tobacco plants, which coincided with the gradient of bioactive CKs in favour of the upper leaves. During the drought stress progression, the level of bioactive CKs decreased in growth-suppressed leaves. Nevertheless, the gradient of bioactive CKs was maintained (even if flattened) by the stimulation of the activity of CKX, the main CK-degrading enzyme, in the lower leaves and by its suppression in the upper leaves. Accumulation of CKs and free auxin in the stressed roots suggests their importance for a droughtinduced change of root morphology (enhanced root elongation). Constitutive 35S::ZOG1 transformants exhibited a delay in the stress-induced elevation of ABA. Thus, the uniform elevation of CK content before stress initiation might have a negative effect during the prolonged, severe drought. Targeted, senescence-induced elevation of ZOG1 expression exhibited a positive effect both on the delay of senescence of the stressed lower leaves and during the recovery. Root growth modification Our observation of the effect of constitutive ZOG1 expression on the morphology of the root system (primary root elongation and diminished branching) is in agreement with the concept of root apical dominance described by Aloni et al. (2006). It also agrees with data on the impact of decreased CK content (in plants over-expressing genes for CKX; Werner et al. 2003) or after the disturbance of the CK signal transduction pathway via mutation of individual CK receptors (Riefler et al. 2006), when the stimulation of the lateral and adventitious root formation was found. However, a functioning CK signal transduction pathway seems to be necessary for root growth, as mutations of all three CK receptors were reported to have strong inhibitory effects (Nishimura et al. 2004). During the stress progression, we observed the accumulation of CKs in the roots of all tested genotypes. As CKs are supposed to be predominantly root-borne phytohormones, which are distributed in the shoot via the xylem stream (Beck & Wagner 1994), CK accumulation in roots might be caused passively by a decrease in the velocity of the transpiration stream. However, the diminished loading of CKs into the xylem observed during the drought (Yang et al. 2002), as well as the large decrease of CKX activity in the stressed roots, suggest the physiological relevance of drought-induced CK accumulation in roots. Limited transport of CKs into shoots seemed to coincide with their growth suppression, while CK accumulation in roots might be of importance for the increase of the root-to-shoot ratio. During the prolonged drought stress, the elevation of IAA concentrations was observed in the lower leaves and roots. As no expression of nitrilase was detected in the ACKNOWLEDGMENTS We are very grateful to Professor Richard Amasino (University of Wisconsin, Madison, WI, USA) for his donation of the SAG12 promoter and Professors Machteld and David Mok and Dr Ruth Martin (Oregon State University, Corvallis, OR, USA) for supplementation of the seeds of the 35S::ZOG1- and SAG12::ZOG1-transformed tobacco. We express our thanks to Professor David Morris (Palacky University, Olomouc, Czech Republic) and Dr Miroslav Kaminek (Institute of Experimental Botany, Prague, Czech Republic) for a critical reading of this manuscript. For technical assistance, we thank Eva Kobzova and Marie Korecka. This work was supported by the Grant Agency of the Czech Republic, Project No. 522/ 04/0549, and by the Ministry of Education, Youth and Sports CR, Project Nos. LC06034 and ME868. © 2008 The Authors Journal compilation © 2008 Blackwell Publishing Ltd, Plant, Cell and Environment, 31, 341–353 87 352 M. Havlová et al. Martin R.C., Mok M.C. & Mok D.W.S. (1999) Isolation of a cytokinin gene, ZOG1, encoding zeatin O-glucosyltransferase from Phaseolus lunatus. Proceedings of the National Academy of Sciences of the United States of America 96, 284–289. Martin R.C., Mok D.W.S., Smets R., Van Onckelen H.A. & Mok M.C. (2001) Development of transgenic tobacco harboring a zeatin O-glucosyltransferase gene from Phaseolus. In Vitro Cellular Development Biology – Plant 37, 354–360. Metwally A., Tsonev T. & Zeinalov Y. (1997) Effect of cytokinins on the photosynthetic apparatus in water-stressed and rehydrated bean plants. Photosynthetica 34, 563–567. Miyawaki K., Matsumoto-Kitano M. & Kakimoto T. (2004) Expression of cytokinin biosynthetic isopentenyltransferase genes in Arabidopsis: tissue specificity and regulation by auxin, cytokinin, and nitrate. The Plant Journal 37, 128–138. Motyka V., Vanková R., Capkova V., Petrasek J., Kaminek M. & Schmulling T. (2003) Cytokinin-induced upregulation of cytokinin oxidase activity in tobacco includes changes in enzyme glycosylation and secretion. Physiologia Plantarum 117, 11– 21. Nishimura C., Ohashi Y., Sato S., Kato T., Tabata S. & Ueguchi C. (2004) Histidine kinase homologs that act as cytokinin receptors possess overlapping functions in the regulation of shoot and root growth in Arabidopsis. The Plant Cell 16, 1365–1377. Pospíšilová J. & Dodd I.C. (2005) Role of plant growth regulators in stomatal limitation to photosynthesis during water stress. In Handbook of Photosynthesis (ed. M. Pessarakli), pp. 811–825. Taylor & Francis, New York, NY, USA. Quirino B.F., Normanly J. & Amasino R.M. (1999) Diverse range of gene activity during Arabidopsis thaliana leaf senescence includes pathogen-independent induction of defense-related genes. Plant Molecular Biology 40, 267–278. Ribaut J.M. & Pilet P.E. (1994) Water-stress and indol-3yl-acetic acid content of maize roots. Planta 193, 502–507. Riccardi F., Gazeau P., Jacquemot M.P., Vincent D. & Zivy M. (2004) Deciphering genetic variations of proteome responses to water deficit in maize leaves. Plant Physiology and Biochemistry 42, 1003–1011. Riefler M., Novak O., Strnad M. & Schmulling T. (2006) Arabidopsis cytokinin receptor mutants reveal functions in shoot growth, leaf senescence, seed size, germination, root development, and cytokinin metabolism. The Plant Cell 18, 40–54. Rodrigues M.L., Pacheco C.M.A. & Chaves M.M. (1995) Soil-plant water relations, root distribution and biomass partitioning in Lupinus albus L. under drought conditions. Journal of Experimental Botany 46, 947–956. Rulcová J. & Pospíšilová J. (2001) Effect of benzylaminopurine on rehydration of bean plants after water stress. Biologia Plantarum 44, 75–81. Salisbury F.B. & Ross C.W. (1992) Plant Physiology, p. 43. Waldsworth Publishing Co., Belmont, CA, USA. Shinozaki K. & Yamaguchi-Shinozaki K. (1996) Molecular responses to drought and cold stress. Current Opinion in Biotechnology 7, 161–167. Singh S., Letham D.S. & Palni L.M.S. (1992) Cytokinin biochemistry in relation to leaf senescence: VII. Endogenous cytokinin levels and exogenous applications of cytokinins in relation to sequential leaf senescence of tobacco. Physiologia Plantarum 86, 388–397. Stoll M., Loveys B. & Dry P. (2000) Hormonal changes induced by partial rootzone drying of irrigated grapevine. Journal of Experimental Botany 51, 1627–1634. Synková H., Van Loven K., Pospíšilová J. & Valcke R. (1999) Photosynthesis of transgenic pssu-ipt tobacco. Journal of Plant Physiology 155, 173–182. REFERENCES Aloni R., Aloni E., Langhans M. & Ullrich C.I. (2006) Role of cytokinin and auxin in shaping root architecture: regulating vascular differentiation, lateral root initiation, root apical dominance and root gravitropism. Annals of Botany 97, 883–893. Bartel B. & Fink G.R. (1994) Differential regulation of an auxinproducing gene family in Arabidopsis thaliana. Proceedings of the National Academy of Sciences of the United States of America 91, 6649–6653. Beck E. & Wagner B.M. (1994) Quantification of the daily cytokinin transport from the root to the shoot of Urtica dioica L. Botanica Acta 107, 342–348. Bianchi M.W., Danerval C. & Vartanian N. (2002) Identification of proteins regulated by cross-talk between drought and hormone pathways in Arabidopsis wild-type and auxin-insensitive mutants, axr1 and axr2. Functional Plant Biology 29, 55–61. Campos N., Bakó L., Feldwisch J., Schell J. & Palme K. (1992) A protein from maize labelled with azido-IAA has novel b-glucosidase activity. The Plant Journal 2, 675–684. Cowan A.K., Freeman M., Björkman P.O., Nicander B., Sitbon F. & Tillberg E. (2005) Effects of senescence-induced alternation in cytokinin metabolism on source-sink relationships and ontogenic and stress-induced transitions in tobacco. Planta 221, 801–814. Davies W.J., Kudoyarova G. & Hartung W. (2005) Long-distance ABA signaling and its relation to other signaling pathways in the detection of soil drying and the mediation of the plant’s response to drought. Journal of Plant Growth Regulation 24, 285–295. Dobrev P.I. & Kaminek M. (2002) Fast and efficient separation of cytokinins from auxin and abscisic acid and their purification using mixed-mode solid-phase extraction. Journal of Chromatography A 950, 21–29. Dobrev P., Motyka V., Gaudinová A., Malbeck J., Travnickova A., Kaminek M. & Vanková R. (2002) Transient accumulation of cis- and trans-zeatin type cytokinins and its relation to cytokinin oxidase activity during cell cycle of synchronized tobacco BY-2 cells. Plant Physiology and Biochemistry 40, 333–337. Dobrev P.I., Havlicek L., Vagner M., Malbeck J. & Kaminek M. (2005) Purification and determination of plant hormones auxin and abscisic acid using solid phase extraction and twodimensional high performance liquid chromatography. Journal of Chromatography A 1075, 159–166. Gan S. & Amasino R.M. (1995) Inhibition of leaf senescence by autoregulated production of cytokinin. Science 70, 1986–1988. Goicoechea N., Dolezal K., Antolin M.C., Strnad M. & SanchezDiaz M. (1995) Influence of mycorrhizae and Rhizobium on cytokinin content in drought-stressed alfalfa. Journal of Experimental Botany 46, 1543–1549. Hare P.D. & Van Staden J. (1997) The molecular basis of cytokinin action. Plant Growth Regulation 23, 41–78. Huynh L.N., Van Toai T., Streeter J. & Banowetz G. (2005) Regulation of flooding tolerance of SAG:ipt Arabidopsis plants by cytokinin. Journal of Experimental Botany 56, 1397–1407. Itai C., Benzioni A. & Munz S. (1978) Heat stress: effects of abscisic acid and kinetin on response and recovery of tobacco leaves. Plant Cell Physiology 19, 453–459. Jordi W., Schapendonk A., Davelaar E., Stoopen G.M., Pot C.S., De Visser R., Van Rhijn J.A., Gan S. & Amasino R.M. (2000) Increased cytokinin levels in transgenic PSAG12-IPT tobacco plants have large direct and indirect effects on leaf senescence, photosynthesis and N partitioning. Plant, Cell & Environment 23, 279–289. Mansfield T.A. & McAinsh M.R. (1995) Hormones as regulators of water balance. In Plant Hormones. Physiology, Biochemistry and Molecular Biology (ed. P.J. Davies), pp. 598–616. Kluwer Academic Publishers, Dordrecht, the Netherlands. © 2008 The Authors Journal compilation © 2008 Blackwell Publishing Ltd, Plant, Cell and Environment, 31, 341–353 88 Role of cytokinins in responses of tobacco to water deficit 353 Volkov R.A., Panchuk I.I. & Schoffl F. (2003) Heat-stressdependency and developmental modulation of gene expression: the potential of house-keeping genes as internal standards in mRNA expression profiling using real-time RT-PCR. Journal of Experimental Botany 54, 2343–2349. Werner T., Motyka V., Laucou V., Smets R., Van Onckelen H. & Schmulling T. (2003) Cytokinin-deficient transgenic Arabidopsis plants show multiple developmental alterations indicating opposite functions of cytokinins in the regulation of shoot and root meristem activity. The Plant Cell 15, 2532–2550. Xin Z.-Y., Zhou X. & Pilet P.E. (1997) Level changes of jasmonic, abscisic and indole-3yl-acetic acids in maize under desiccation stress. Journal of Plant Physiology 151, 120–124. Yang J., Zhang J., Wang Z., Zhu Q. & Liu L. (2002) Abscisic acid and cytokinins in the root exudates and leaves and their relationship to senescence and remobilization of carbon reserves in rice subjected to water stress during grain filling. Planta 215, 645–652. Zhang J., Van Toai T., Huynh L. & Preiszner J. (2000) Development of flooding-tolerant Arabidopsis thaliana by autoregulated cytokinin production. Molecular Breeding 6, 135–144. Received 28 September 2007; received in revised form 20 November 2007; accepted for publication 30 November 2007 SUPPLEMENTARY MATERIAL The following supplementary material is available for this article: Table S1. Primers for Nicotiana tabacum genes used in qRT-PCR. This material is available as part of the online article from http://www.blackwell-synergy.com/doi/abs/10.1111/ j.1365-3040.2007.01766.x (This link will take you to the article abstract) Please note: Blackwell Publishing is not responsible for the content or functionality of any supplementary materials supplied by the authors. Any queries (other than missing material) should be directed to the corresponding author for the article. © 2008 The Authors Journal compilation © 2008 Blackwell Publishing Ltd, Plant, Cell and Environment, 31, 341–353 89
© Copyright 2025 Paperzz