28.10.2014 Views

No.42 - 農業生物資源研究所

No.42 - 農業生物資源研究所

No.42 - 農業生物資源研究所

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

ISSN 0435-1096<br />

Gamma Field Symposia<br />

Gamma Field Symposia<br />

Number 42<br />

PLANT HORMONE RESEARCH AND MUTATION<br />

2 0 0 3<br />

INSTITUTE OF RADIATION BREEDING<br />

NIAS<br />

Ohmiya-machi, Naka-gun, Ibaraki-ken<br />

Japan


PLANTHORMONERESEARCH AND MUTATION<br />

Report of Symposium<br />

held on<br />

July 16-17, 2003<br />

Institute of Radiation Breeding<br />

NIAS<br />

Ohmiya-machi, Naka-gun, Ibaraki-ken 319-2293<br />

Japan


The lecturers and the members of the Symposium Committee<br />

General discussion


List of Participants<br />

(42nd GF Symposium)<br />

ABE, F.<br />

AJIRO, T.<br />

ARAI, T.<br />

ASAUMI, H.<br />

ASHIKARI, M.<br />

DEGI, K.<br />

EHIRA, M.<br />

ENDO, T.<br />

EZURA, H.<br />

FUJITA, M.<br />

FUJITA, Y.<br />

FUKATSU, E.<br />

FUWA, N.<br />

GONAI, T.<br />

GOTO, Y.<br />

GOTO, Y.<br />

HARA, H.<br />

HASEGAWA, H.<br />

HATASHITA, M.<br />

HATTORI, E.<br />

HATTORI, K.<br />

HATTORI, T.<br />

HAYAKAWA, Y.<br />

HAYASHI, Y.<br />

HIGO, K.<br />

HIRAI, T.<br />

HIRATA, Y.<br />

HIRAYAMA, T.<br />

HOBO, T.<br />

HOSOYA, T.<br />

INAGAKI, H.<br />

INOUE, E.<br />

ISHII, T.<br />

ISHIMARU, K.<br />

ITO, A.<br />

ITO, Y.<br />

IWABUCHI, M.<br />

IYOZUMI, H.<br />

KADOTA, N.<br />

KAMIYA, Y.<br />

National Institute of Crop Science<br />

Tokyo University of Agriculture and Technology<br />

Gifu Research Institute for Agricultural Sciences<br />

Ehime Agricultural Experiment Station<br />

Bioscience and Biotechnology Center, Nagoya University<br />

Institute of Radiation Breeding, National Institute of Agrobiological Sciences<br />

Tokyo University of Agriculture and Technology<br />

Miyagi Furukawa Agricultural Experiment Station<br />

University of Tsukuba<br />

RIKEN Tsukuba Institute<br />

Japan International Research Center for Agricultural Sciences<br />

Forest Tree Breeding Center<br />

Snow Brand Seed Co., Ltd.<br />

Ibaraki Plant Biotechnology Institute<br />

Forest Tree Breeding Center<br />

Hitachi High-Technologies<br />

Ibaraki University<br />

University of Shiga Prefecture<br />

The Wakasawan Energy Research Center<br />

Toyota Motor Corporation<br />

Graduated School of Bioagricultural Sciences, Nagoya University<br />

Bioscience and Biotechnology Center, Nagoya University<br />

Fukui Prefectural University<br />

Fukushima Agricultural Experiment Station<br />

National Institute of Agrobiological Sciences<br />

Tokyo University of Agriculture<br />

Tokyo University of Agriculture and Technology<br />

RIKEN Central Institute<br />

RIKEN Laboratory of Plant Molecular Biology<br />

High School of Agriculture, Mito Ibaraki<br />

Shizuoka Agricultural Experiment Station<br />

Ibaraki University<br />

Ibaraki Agricultural Center<br />

University of Tsukuba<br />

National Institute of Fruit Tree Science<br />

Institute of Radiation Breeding, National Institute of Agrobiological Sciences<br />

National Institute of Agrobiological Sciences<br />

Shizuoka Agricultural Experiment Station<br />

Takii & Co., LTD<br />

RIKEN Yokohama Institute


KANEKO, T.<br />

KARITA, E.<br />

KASHIMA, M.<br />

KATAOKA, H.<br />

KATO, K.<br />

KATO, M.<br />

KAWAKATSU, M.<br />

KOBAYASHI, M.<br />

KOBAYASHI, S.<br />

KUBOTA, K.<br />

KUBOYAMA, T.<br />

KUDO, S.<br />

KUJIOKA, H.<br />

KURODA, K.<br />

KUSABA, M.<br />

KUSANO, M.<br />

KUSANO, T.<br />

KUSHIRO, T.<br />

KUZUYA, M.<br />

MATSUBAYASHI, Y.<br />

MATSUI, H.<br />

MIYANO, S.<br />

MORISHITA, T.<br />

MORITA, M.<br />

MORITA, R.<br />

MORITA, R.<br />

MORITA, Y.<br />

MOTODA, J.<br />

NAGATO, Y.<br />

NAGATOMI, S.<br />

NAITO, K.<br />

NAITO, T.<br />

NAITO, Y.<br />

NAKAJIMA, I.<br />

NAKAMURA, S.<br />

NAKANO, T.<br />

NARA, Y.<br />

NARUKAWA, M.<br />

NIKI, T.<br />

NISHIMURA, M.<br />

NISHIMURA, S.<br />

NONAKA, S.<br />

NOZAWA, G.T.<br />

OBARA, N.<br />

Plant Bioengineering Research Laboratories, Sapporo Breweries<br />

Tokyo University of Science<br />

Ibaraki Agricultural Center<br />

University of Tsukuba<br />

Shizuoka Agricultural Experiment Station<br />

Takii Plant Breeding & Experiment Station<br />

Institute of Radiation Breeding, National Institute of Agrobiological Sciences<br />

RIKEN Tsukuba Institute<br />

National Institute of Fruit Tree Science<br />

Nagano Agricultural Experiment Station<br />

Ibaraki University<br />

Graduate school of Agricultural Science, Tohoku University<br />

High School of Agriculture, Mito Ibaraki<br />

National Institute of Agrobiological Sciences<br />

Institute of Radiation Breeding, National Institute of Agrobiological Sciences<br />

Kaisui Chemical Industry Co., LTD.<br />

Graduate School of Life Sciences, Tohoku University<br />

RIKEN Plant Science Center<br />

Ibaraki Agricultural Center<br />

Graduated School of Bioagricultural Sciences, Nagoya University<br />

Kyoto University<br />

Keisei Rose Nurseries, LTD.<br />

Institute of Radiation Breeding, National Institute of Agrobiological Sciences<br />

Kyoto University<br />

Institute of Radiation Breeding, National Institute of Agrobiological Sciences<br />

Tokyo University of Agriculture<br />

The University of Tokyo<br />

Graduated School of Bioagricultural Sciences, Nagoya University<br />

Graduate School of Agricultural and Life Science, The University of Tokyo<br />

Institute of Radiation Breeding, National Institute of Agrobiological Sciences<br />

Institute of Radiation Breeding, National Institute of Agrobiological Sciences<br />

School of Agriculture, Meiji University<br />

Tokyo University of Agriculture<br />

National Institute of Fruit Tree Science<br />

National Institute of Crop Science<br />

RIKEN Central Institute<br />

Tokyo University of Science<br />

Tokyo University of Science<br />

National Institute of Floricultural Science<br />

Institute of Radiation Breeding, National Institute of Agrobiological Sciences<br />

University of Tsukuba<br />

University of Tsukuba<br />

Tokyo University of Agriculture and Technology<br />

VISTA, LTD


OHMIYA, Y.<br />

OHSAWA, K.<br />

OHTA, K.<br />

OHTA, Y.<br />

OHTSUBO, N.<br />

OKA, A.<br />

OKA, S.<br />

OKAMURA, J.<br />

OKAMURA, M.<br />

OKAZAKI, K.<br />

OKUDAIRA, M.<br />

OKUMOTO, Y.<br />

ONO, Y.<br />

ONOZAKI, T.<br />

RIKIISHI, K.<br />

SAISHO, D.<br />

SAITO, K.<br />

SAITO, M.<br />

SANADA, T.<br />

SANO, Y.<br />

SATO, M.<br />

SATO, M.<br />

SATO, T.<br />

SATO, Y.<br />

SEKIGUCHI, F.<br />

SHIMOMURA, S.<br />

SOEJIMA, J.<br />

SUGIYAMA, K.<br />

SUNOHARA, H.<br />

SUZUKI, A.<br />

SUZUKI, Y.<br />

TAJI, T.<br />

TAKABATAKE, R.<br />

TAKAHASHI, H.<br />

TAKAHASHI, M.<br />

TAKAHASHI, T.<br />

TAKANO, T.<br />

TAKASAKI, T.<br />

TAKATSU, Y.<br />

TAKEUCHI, S.<br />

TAKYU, T.<br />

TANAKA, M.<br />

TANIGUCHI, T.<br />

TANISAKA, T.<br />

Forest Tree Breeding Center<br />

Nagano Agricultural Research Center<br />

Tokyo University of Agriculture and Technology<br />

Tokyo University of Agriculture and Technology<br />

Ministry of Education, Culture, Sports, Science and Technology<br />

Institute for Chemical Research, Kyoto University<br />

National Institute of Agrobiological Sciences<br />

Sakata Seed Corp.<br />

Plant Lab., Kirin Brewery Co., Ltd.<br />

Niigata University<br />

Iwate Agricultural Research Center<br />

Kyoto University<br />

Bio-oriented Technology Research Advancement Institution<br />

National Institute of Floricultural Science<br />

Okayama University<br />

Okayama University<br />

Saitama Agriculture and Forestry Research Center<br />

Fukui Agriculture Experiment Station<br />

National Institute of Fruit Tree Science<br />

Graduate School of Agriculture, Hokkaido University<br />

Oita Agricultural Research Center<br />

Japan International Research Center for Agricultural Sciences<br />

Akita Agricultural Experiment Station<br />

Toyota Motor Corporation<br />

Japan Women’s University<br />

National Institute of Agrobiological Sciences<br />

National Institute of Fruit Tree Science<br />

Shizuoka Citrus Experiment Station<br />

Nagoya University<br />

Iwate University<br />

National Institute of Crop Science<br />

RIKEN Laboratory of Plant Molecular Biology<br />

National Institute of Agrobiological Sciences<br />

Tokyo University of Science<br />

Forest Tree Breeding Center<br />

Watanabe Seed Co., LTD.<br />

Institute of Radiation Breeding, National Institute of Agrobiological Sciences<br />

Tochigi Agricultural Experiment Station<br />

Ibaraki Agricultural Center<br />

Keisei Rose Nurseries LTD.<br />

Institute of Radiation Breeding, National Institute of Agrobiological Sciences<br />

Nagasaki Fruit Tree Experiment Station<br />

Forest Tree Breeding Center<br />

Graduate School of Agriculture, Kyoto University


TANO, S.<br />

TSUCHIDA, Y.<br />

TSUTSUMI, N.<br />

UCHIDA, H.<br />

UEKI, T.<br />

WATANABE, H.<br />

WATANABE, M.<br />

YAMAGUCHI, H.<br />

YAMAGUCHI, I.<br />

YAMANOUCHI, H.<br />

YASHIRO, K.<br />

YAZAWA, H.<br />

YOKOYAMA, T.<br />

YOSHIDA, T.<br />

YOSHIKAWA, T.<br />

YOSHIOKA, T.<br />

YUHASHI, K.<br />

ZHOU, T.<br />

National Institute of Fruit Tree Science<br />

Graduate School of Agricultural and Life Science, The University of Tokyo<br />

Hamamatsu Photonics, LTD<br />

High School of Agriculture, Mito Ibaraki<br />

Japan Atomic Energy Research Institute-Takasaki<br />

Iwate University<br />

Institute of Radiation Breeding, National Institute of Agrobiological Sciences<br />

Japan Seed Trade Association<br />

Institute of Radiation Breeding, National Institute of Agrobiological Sciences<br />

Ibaraki Agricultural Center<br />

Yokohama Nursery Co,. LTD.<br />

Saitama Flower & Garden Center<br />

Takii & Co., LTD.<br />

Kyoto University<br />

National Institute of Crop Science<br />

University of Tsukuba<br />

Sapporo Breweries


FOREWORD<br />

Forty years have passed since full-scale mutation breeding research began with the founding<br />

of the Institute of Radiation Breeding (IRB). During this time, many mutation varieties bred in<br />

Japan have been accepted as familiar agricultural products. Mutation breeding is expected to play<br />

an increasingly important role in Japan for improving plants. In this sense, this symposium<br />

provides a vital opportunity for exchanging research information and engaging in much-needed<br />

discussion.<br />

Dramatic progress has been made during the last decade in the study of biosynthesis,<br />

perception, and signal transduction of plant hormones. Due to the importance of recent basic<br />

molecular knowledge of plant hormones to future research in mutation breeding, we have selected<br />

the theme “Plant hormone research and mutation” for the 42nd Gamma Field Symposium. We<br />

have invited eight lectures on this subject, including a special lecture entitled, “Biosyntheses and<br />

regulation of plant hormones” on the physiological action of plant hormones to be given by Dr.<br />

Yuji Kamiya of the Institute of Physical and Chemical Research (RIKEN).<br />

The IRB has economically evaluated mutation breeding in agricultural production and<br />

prepared pamphlets to introduce mutation varieties released in Japan. As of March 2003, 320<br />

mutant varieties have served in developing 50 kinds of crops. The cumulative cultivation of<br />

mutation varieties has occupied 5.5 million hectares, corresponding to 1.1-fold of the total area<br />

under cultivation per year in Japan. The cumulative production value of such agricultural products<br />

reached 7 trillion yen, which corresponds to total annual domestic agricultural production.<br />

Mutagens (agents of mutation induction) used for these 320 mutant varieties included gammarays<br />

at 72%, tissue culture at 11%, chemical substances at 10%, and X-rays and others at 7%.<br />

Among mutation varieties, 145 were used as is, obtained directly from mutagen processing. Of<br />

these 74 varieties used IRB facilities, accounting for 51%. The number of indirectly used varieties<br />

obtained by hybridizing mutants or the progeny of mutants was 175, demonstrating how the use of<br />

mutated genes had advanced.<br />

In closing, we thank the lecturers who have so kindly taken time out from busy schedules to<br />

prepare for this symposium and to all who have provided their unstinting support in making this<br />

symposium a success.


The Symposium Committee<br />

Shigeki NAGATOMI, Chairperson<br />

Yuji ITO<br />

Toshikazu MORISHITA<br />

Yasuo NAGATO<br />

Minoru NISHIMURA<br />

Seibi OKA<br />

Tetsuro SANADA<br />

Yoshio SANO<br />

Takatoshi TANISAKA<br />

Nobuhiro TSUTSUMI


PROGRAM<br />

Opening address : S. NAGATOMI<br />

Congratulatory address : M. IWABUCHI<br />

Special lecture<br />

Chairperson : S. TANO<br />

Biosyntheses and Regulation of Plant Hormones Y. KAMIYA<br />

Session <br />

Chairperson : S. OKA<br />

Molecular Mechanisms for Auxin Responce and Signal Transduction S. SHIMOMURA<br />

Session <br />

Chairperson : Y. NAGATO<br />

Gibberellin Responce and Signal Transduction M. ASHIKARI<br />

Session <br />

Chairperson : T. TANISAKA<br />

Cytokinin Signal Transduction and Two-Component Regulatory System A. OKA<br />

Session <br />

Chairperson : N. TSUTSUMI<br />

Molecular Mechanisms for ABA Responce and Signal Transduction T. HATTORI<br />

Session <br />

Chairperson : T. SANADA<br />

Molecular Mechanisms for Ethylene Perception and Signal Transduction T. HIRAYAMA<br />

Session <br />

Chairperson : H. HASEGAWA<br />

Mechanism of Brassinosteroid Signaling T. NAKANO<br />

Session <br />

Chairperson : K. HATTORI<br />

Peptide Plant Hormone, Phytosulfokine Y. MATSUBAYASHI<br />

Session <br />

Chairperson : Y. SANO<br />

General discussion<br />

Closing address : Y. NAGATO


CONTENTS<br />

Y. KAMIYA Biosyntheses and Regulation of Plant Hormones 1<br />

M. ASHIKARI Gibberellin Responce and Signal Transduction 13<br />

H. ITOH<br />

M. UEGUCHI<br />

M. MATSUOKA<br />

A. OKA Cytokinin Signal Transduction and Two-Component Regulatory System 25<br />

T. HIRAYAMA Molecular Mechanisms for Ethylene Perception and<br />

T. UGAJIN Signal Transduction 41<br />

T. NAKANO Mechanism of Brassinosteroid Signaling 53<br />

S. YOSHIDA<br />

T. ASAMI<br />

Y. MATSUBAYASHI Peptide Plant Hormone, Phytosulfokine 67<br />

General discussion (in Japanese 79


Gamma Field Symposia, No. 42, 2003 Institute of Radiation Breeding<br />

NIAS, Japan<br />

REGULATION OF PLANT HORMONE BIOSYNTHESES<br />

1<br />

BIOSYNTHESES AND REGULATION OF<br />

PLANT HORMONES<br />

Yuji KAMIYA<br />

RIKEN Plant Science Center<br />

1-7-22, Suehiro-cho, Tsurumi-ku, Yokohama, 230-0045<br />

Introduction<br />

Plant hormones are signal molecules, present in trace quantities. Changes in hormone<br />

concentration and tissue sensitivity mediate a whole range of developmental process in plants,<br />

many of which involve interactions with environmental factors. So far seven hormones are<br />

considered as major plant hormones, namely auxin, gibberellins, cytokinins, abscisic acid,<br />

ethylene, brassinosteroids and jasmonic acid. Each of these hormones has its own particular<br />

properties, so the pathways regulating their production and degradation are quite diverse and have<br />

been elucidated by use of chemistry, biochemistry, plant physiology, genetics and molecular<br />

genetics. Recently the genomic sequence of Arabidopsis is available for hormone study and that of<br />

the rice is also in hand. Now it is possible to study hormone biosyntheses and their regulation by<br />

reverse genetic approach. In this symposium I focused on biosyntheses of gibberellins, cytokinins<br />

and absisic acid, which we are now studying in my laboratories. Part of this paper involves some<br />

new results and data, which are now provisionally accepted in some journals. Therefore this paper<br />

is neither a review nor an original paper. However, this paper describes what I talked during the<br />

symposium. In order to give credits to the researchers in the three different topics, major<br />

researchers are listed in parentheses.<br />

Regulation of GA biosynthesis by cold temperature in Arabidopsis germinating seeds<br />

(Yukika YAMAUCHI and Shinjiro YAMAGUCHI)<br />

Gibberellins (GA)s are involved in many processes of plant development, such as seed<br />

germination, stem elongation, leaf expansion, flowering, and seed development (DAVIES 1995).<br />

GAs are synthesized from geranylgeranyl diphosphate (GGDP), which is sequentially converted to<br />

biologically active GAs by terpene cyclases, cytochrome P450 monooxygenases and 2-oxoglutaratedependent<br />

dioxygenases (Fig. 1) (HEDDEN and KAMIYA 1997). Most of the genes encoding GA<br />

biosynthesis and catabolism enzymes have now been identified (OLSZEWSKI et al. 2002).


2<br />

Yuji KAMIYA<br />

Fig. 1. Gibberellin biosynthesis in Arabidopsis.<br />

It has been known that GA promotes seed germination in many plant species. In Arabidopsis,<br />

severe GA deficient-mutants, such as ga1-3 and ga2-1, are defective in seed germination<br />

(KOORNNEEF and VAN DER VEEN 1980), and chemical inhibitors of GA biosynthesis inhibit<br />

germination (NAMBARA et al. 1991). These observations indicate that de novo GA biosynthesis is<br />

necessary for seed germination in Arabidopsis (HEDDEN and KAMIYA 1997).<br />

Light is a critical environmental factor for seed germination in some small-seeded plants such<br />

as lettuce, tomato and Arabidopsis (SHINOMURA, 1997). The effect of light on seed germination is<br />

primarily mediated by phytochromes (BORTHWICK et al. 1952; BUTLER et al. 1959). Genes<br />

encoding GA 3-oxidases, which convert inactive precursor to active GAs, are regulated by<br />

phytochromes in germinating lettuce and Arabidopsis seeds (TOYOMASU et al. 1998; YAMAGUCHI<br />

et al. 1998).<br />

Temperature is another crucial external cue that controls seed germination (BEWLEY and<br />

BLACK 1982). In many plant species, exposure of seeds to low temperature (typically 2-5)<br />

immediately after imbibition is effective to promote germination. (SHROPSHIRE et al. 1961; CONE<br />

and SPRUIT 1983). This treatment is called “stratification”. Although this method is widely used to<br />

improve the frequency and synchronization of germination, the mechanism for the thermoregulation<br />

of seed germination has been unclear.<br />

The effect of cold treatment on GA content has been reported in the 1970s, based on semiquantitative<br />

analysis of endogenous GAs using -amylase bioassay and/or gas chromatographymass<br />

spectrometry (GC-MS) (ROSS and BRADBEER 1971; SINSKA et al. 1973; WILLIAMS et al.<br />

1974). DERKX et al. (1994) analyzed the effect of pre-chilling of Arabidopsis seeds and reported<br />

that bioactive GA4 was detectable only in pre-chilled seeds, but not in dark-imbibed seeds. It has<br />

not been clear whether the effect of pre-chilling is a direct response to low temperature because pre-


REGULATION OF PLANT HORMONE BIOSYNTHESES<br />

3<br />

chilling treatment in published studies also involved a longer imbibition period.<br />

Yukika YAMAUCHI and Shinjiro YAMAGUCHI have worked intensively about the effect of cold<br />

treatment. We first carried out large-scale expression analysis during imbibition of after-ripened<br />

Arabidopsis dry seeds at 4. This investigation indicates that a number of GA-related genes are<br />

differentially expressed between dry and cold-treated seeds. Our GC-MS and reverse transcription-<br />

PCR analyses show that GA biosynthesis is activated in response to low temperature in darkimbibed<br />

seeds, and that the effect of temperature is targeted to particular GA biosynthesis genes.<br />

Using a loss-of-function mutant of the cold-inducible AtGA3ox1 gene, we show that this gene is<br />

required for cold-promoted synthesis of active GAs and seed germination. Our results suggest that<br />

germination of Arabidopsis seeds is stimulated in response to low temperature in part through<br />

modulating GA biosynthesis (Fig. 2). Furthermore, we show that cold treatment increases the<br />

number of cell types accumulating the AtGA3ox1 transcript detectable by in situ hybridization<br />

analysis, suggesting a complex regulatory mechanism by which the spatial distribution of GA<br />

biosynthesis is determined (YAMAUCHI et al. 2004).<br />

Biosynthesis of prenyl side chain of cytokinins<br />

(Hiroyuki KASAHARA, Kentaro TAKEI, Shinjiro YAMAGUCHI and Hitoshi SAKAKIBARA)<br />

Cytokinins (CKs) have many physiological roles in plants, such as promotion of cell division<br />

and shoot formation in the presence of auxin, release of lateral buds from apical dominance,<br />

stimulation of chloroplast development, and delay of senescence. The biological activity, biosynthesis<br />

and metabolism of CKs have been well studied (MOK and MOK, 2001). Recently, a CK receptor<br />

has been identified in Arabidopsis (INOUE et al. 2001; YAMADA et al. 2001). Functional analysis<br />

Fig. 2. In Arabidopsis germinating seeds, GA3ox1 gene is regulated by<br />

light, temperature and negative feed back. Red and blue and<br />

arrows indicate effects of light and cold temperature respectively.


4<br />

Yuji KAMIYA<br />

using the receptor protein expressed in Escherichia coli and yeast have indicated that free CKs, transzeatin<br />

(tZ) and isopentenyl adenine (iP), are active CK species in Arabidopsis (INOUE et al. 2001).<br />

Most of CKs identified from plants are derivatives of N 6 -prenylated adenine. Plants have two<br />

possible biosynthetic pathways for the production of the side chain of CKs, namely mevalonate<br />

(MVA) pathway in the cytosol and methylerythritol phosphate (MEP) pathway in plastids<br />

(LICHITENTHALER 1999; ROHMER 2003). Both pathways supply common precursors, isopentenyl<br />

diphosphate (IPP) and dimethylallyl diphosphate (DMAPP). Although the MVA and MEP<br />

pathways are localized in different subcellular compartments, there is some exchange of common<br />

precursor(s) between the two pathways (KASAHARA et al. 2002). Therefore, application of a<br />

precursor specific to the MEP pathway can suppress the growth inhibition caused by a block in the<br />

MVA pathway, and vice versa (HEMMERLIN et al. 2003).<br />

Hiroyuki KASAHARA and Shinjiro YAMAGUCHI in collaboration with Kentaro TAKEI and<br />

Hitoshi SAKAKIBARA worked about the prenylation of CK side chain.<br />

CK biosynthesis pathway in plants has initially been deduced on the basis of bacterial<br />

enzymes. The formation of N 6 -prenylated adenine from AMP was first demonstrated using cellfree<br />

extracts of a slime mold, Dictyostelium discoideum (TAYA et al. 1978). A CK biosynthesis<br />

gene TMR, which encode AMP:DMAPP-isopentenyltransferase (AMP:DMAPP-IPT), was isolated<br />

from Ti-plasmid of Agrobacterium tumefaciens (BARRY et al. 1984; AKIYOSHI et al. 1984).<br />

Because CK levels were elevated in transgenic plants that overproduce bacterial AMP: DMAPP-<br />

IPT, the formation of isopentenyl adenosine monophosphate (iPRMP) from AMP and DMAPP is<br />

likely to be a committed step in CK biosynthesis. Subsequent formation of iP from iPRMP<br />

requires 5’-nucleotidase and adenosine nucleosidase. The conversion of iP into tZ is catalyzed by<br />

trans-hydroxylase, which is probably a P450 monooxygenase (CHEN and LEISNER 1984).<br />

Recently, a search for Arabidopsis genes that are homologous to bacterial AMP:DMAPP-IPT has<br />

identified nine AtIPT genes (TAKEI et al. 2001; KAKIMOTO 2001). Unlike bacterial IPTs, seven<br />

AtIPTs were able to transfer DMAPP not only to AMP, but also to ADP and ATP, to give<br />

corresponding nucleotide CKs in vitro. Thus, the AMP/ADP/ATP-dependent pathway has been<br />

proposed for the biosyntheses of iP/tZ in plants.<br />

On the other hand, the tRNA-dependent pathway has been also proposed for the biosynthesis<br />

of CKs in plants because tRNAs in bacteria, yeast and plants contain a N 6 -prenylated adenine<br />

moiety, which, by hydrolysis, is capable of forming CKs (MURAI 1994). Among nine IPT-related<br />

sequences in Arabidopsis, two AtIPT genes encode (putative) tRNA-isoprenyltransferases (tRNA-<br />

IPTs). The prenylated adenine element in tRNA is usually consisted of iP and cis-zeatin (cZ), the<br />

cis-isomer of tZ. Therefore, the tRNA-dependent pathway has not been considered as the main<br />

route to tZ. However, the occurrence of cis-trans isomerase activity in Phaseolus vulgaris immature<br />

seeds suggested that the tRNA-dependent pathway might also contribute to the biosynthesis of tZ<br />

through cZ (BASSIL et al. 1993).<br />

The MVA pathway had been considered as the sole route providing DMAPP to CKs until the


REGULATION OF PLANT HORMONE BIOSYNTHESES<br />

5<br />

MEP pathway was uncovered recently. The incorporation of 14 C-labeled MVA into the iP element<br />

of tRNA in vivo has been demonstrated in tobacco pith tissue (CHEN and HALL 1969). Also, 13 C-<br />

labeled MVA was incorporated into iP and trans-zeatin riboside (tZR) in vitro in the endosperm of<br />

Sechium edule (PIAGGESI et al. 1997). In addition, there are some reports that indicate that CK<br />

levels are reduced in plants when the MVA pathway is limited (ÅSTOT et al. 2000). On the other<br />

hand, the contribution of the MEP pathway to CK biosynthesis has never been issued before. It<br />

should be noted that the incorporation of MVA does not exclude a potential role of the MEP<br />

pathway in the biosynthesis of CKs because it has often been observed that isoprene units from<br />

both MVA and MEP pathways are incorporated into a single downstream isoprenoid (KASAHARA<br />

et al. 2002; HEMMERLIN et al. 2003). Thus, a possible contribution of the MEP pathway to the<br />

biosynthesis of CKs needs to be examined to better understand how CKs are synthesized in plant.<br />

In order to selectively label metabolites from the MVA or MEP pathways with 13 C in vivo, we<br />

have previously carried out feeding of [1- 13 C] 1-deoxy-D-xylulose (DX) or [2- 13 C]mevalonolactone<br />

(MVL) to Arabidopsis seedlings. DX is converted into an MEP pathway intermediate 1-deoxy-Dxylulose<br />

5-phosphate (DXP) by phosphorylation. Therefore, exogenous DX is able to complement<br />

the albino phenotype of the cla1-1 mutant (ESTEVEZ et al. 2000), which is defective in DXP in the<br />

MEP pathway. Similarly, the growth inhibition due to a block in the MVA pathway by mevastatin<br />

(an inhibitor of HMG-CoA reductase) is rescued by exogenous application of MVL (KASAHARA<br />

et al. 2002). Efficient 13 C-labeling of metabolites from the MVA or MEP pathways was thus<br />

achieved by feeding 13 C-labeled DX and MVL to the cla1-1 mutant and mevastatin-treated plants,<br />

respectively. These 13 C-labeling systems allowed us to determine contribution of the MVA and<br />

MEP pathways to the biosynthesis of GAs by gas chromatography-mass spectrometry (KASAHARA<br />

et al. 2002).<br />

We studied the biosynthesis route for the prenyl moiety of CKs using the 13 C-labeled tracers in<br />

Arabidopsis seedlings. Our data demonstrate that the prenyl side chain of tZ- and iP-type CKs are<br />

mainly produced through the MEP pathway (Fig. 3), whereas a large fraction of cZ derivatives is<br />

synthesized through the MVA pathway. We also show the subcellular location of AtTPTs produced<br />

as GFP-fusion proteins. Based on these data, we proposed a crucial role of the plastid-localized<br />

MEP pathway in CK biosynthesis (KASAHARA et al. 2004).<br />

Cloning of P450 genes involved in abscisic acid degradation<br />

(Tetsuo KUSHIRO, Masanori OKAMOTO, Kazumi NAKABAYASHI and Eiji NAMBARA)<br />

Abscisic acid (ABA) controls numerous aspects of plant life cycle including seed dormancy,<br />

germination and adaptive responses to environmental stresses (ZEEVAART and CREELMAN, 1988).<br />

ABA-deficient mutants from several plant species show reduced seed dormancy and wilty<br />

phenotype (MCCARTY 1995). ABA content increases during seed development or when a plant is<br />

subjected to various stresses such as osmotic stress, while it rapidly decreases during subsequent


6<br />

Yuji KAMIYA<br />

Fig. 3. Origin of side chains of trans-zeatin and cis-zeatin in Arabidopsis.<br />

cZ, cis-zeatin; cZR, cis-zeatin riboside; cZRMP, cis-zeatin ribosidemonophosphate;<br />

DMAPP, dimethylallyl diphosphate; iP, isopentenyladenine; iPR, isopentenyladenine<br />

riboside; iPRMP, isopentenyladenine riboside monophosphate; MEP, methylerythritol<br />

phosphate; MVA, mevalonate<br />

germination or the recovery from stress. ABA content is determined by the balance between<br />

biosynthesis and catabolism. When endogenous ABA levels is maintained high, both ABA<br />

biosynthesis and catabolism are active (HARRISON and WALTON 1975; ZEEVAART 1980; PIERCE<br />

and RASCHKE, 1981). Constitutive expression of ABA biosynthetic gene in transgenic plants<br />

exhibits a more prominent accumulation of the catabolites compared to a moderate increase in<br />

ABA contents (QIN and ZEEVAART, 2002). Recently, most of ABA biosynthetic genes have been<br />

identified (SCHWARTZ et al. 2003; SEO and KOSHIBA 2002). However, molecular mechanisms<br />

underlying ABA catabolism remain poorly understood.<br />

ABA is catabolized into inactive forms either by oxidation or conjugation (MILBORROW<br />

1969; MILBORROW 1975; WALTON and SONDHEIMER 1972; SONDHEIMER et al. 1974; XU et al.<br />

2002; see review; CUTLER and KROCHKO 1999) (See Fig. 4). The predominant pathway for ABA<br />

catabolism is the oxidative pathway, which is triggered by hydroxylation at C-8’ to produce 8’-<br />

hydroxy ABA. The 8’-hydroxy ABA is subsequently isomerized spontaneously to form phaseic<br />

acid (PA) (MILBORROW et al. 1988). Biological activity of PA is significantly less than that of<br />

ABA, therefore, the major regulatory step in inactivation is likely to be 8’-hydroxylation of ABA<br />

(ARAI et al. 1999). This reaction is known to be catalyzed by a cytochrome P450 monooxygenase<br />

(P450) (GILLARD and WALTON, 1976; KROCHKO et al. 1998). However, the gene encoding ABA<br />

8’-hydroxylase remained elusive. It is necessary to identify this gene to understand the molecular<br />

mechanism controlling the hormonal level of ABA. Tetsuo KUSHIRO, Masanori OKAMOTO (Ph.D.<br />

student from the Tokyo Metropolitan Univ., Prof. Tomokazu KOSHIBA), Kazumi NAKABAYASHI


REGULATION OF PLANT HORMONE BIOSYNTHESES<br />

7<br />

Fig. 4. Abscisic acid catabolism. 8’-Hydroxylation is catalyzed by CYP707A1-A4 in Arabidopsis.<br />

and Eiji NAMBARA have worked intensively about the cloning and characterization of the ABA<br />

catabolic enzyme.<br />

P450s are large family of enzymes that catalyze the oxidation of various low molecular<br />

weight compounds. They have been conserved for billions of years and exist in most of the<br />

organisms on the earth ranging from bacteria to mammals. In mammals, P450s play a major role in<br />

drug metabolism, and have been a center of research in the pharmaceutical field. In plants, P450s<br />

participate in numerous aspects of plant metabolism, which include phytohormones and secondary<br />

metabolites (SCHULER 1996; CHAPPLE 1998).<br />

Completion of the Arabidopsis genome sequencing has revealed that there are at least 272<br />

P450 genes in this single organism. In the following completion of the rice genome sequencing,<br />

nearly 450 P450 genes have been identified (http://drnelson.utmem.edu/rice.html). These numbers<br />

obviously indicate how widely these genes have evolved and deeply rooted in the plant life cycle.<br />

The genome sequencing efforts were successful in accumulating sequence information, however, it<br />

is still a major challenge to identify the function of each gene. This will be a major task in the post<br />

genomic era, and will require novel approaches as well as systematic analysis of the gene. P450s<br />

are especially challenging since in most cases, the substrate of the enzyme cannot be easily<br />

predicted. Furthermore, the number of possible steps where P450 participates along the metabolic<br />

pathways is largely unknown.<br />

In order to identify the P450 gene for ABA 8’-hydroxylase, we have set out to search for the<br />

gene of our interest among hundreds of candidate genes. Once the gene for ABA 8’-hydroxylase is<br />

identified, it would be possible to fine-tune the level of ABA in plants, and thus, would expect to<br />

improve drought tolerance as well as to prevent precocious germination in crops. Therefore, the<br />

identification of this gene would have an enormous impact on the agricultural industry. Our<br />

extensive and careful prediction led to the first successful identification of the members of<br />

CYP707A family as ABA 8’-hydroxylase genes. Expression analysis and genetic analysis<br />

demonstrated that CYP707 genes play a regulatory role in vivo to define the ABA level during<br />

these processes (KUSHIRO et al. 2004).


8<br />

Yuji KAMIYA<br />

References<br />

AKIYOSHI, D.E., KLEE, H., AMASINO, R.M., NESTER, E.W. and GORDON, M.P. (1984) T-DNA of Agrobacterium<br />

tumefaciens encodes an enzyme of cytokinin biosynthesis. Proc. Natl. Acad. Sci. USA 81: 5994-5998.<br />

ARAI, S., TODOROKI, Y., IBARAKI, S., NAOE, Y., HIRAI, N. and OHIGASHI, H. (1999) Synthesis and activity of<br />

3’-chloro, -bromo, and -iodoabscisic acids, and biological activity of 3’-fluoro-8’-hydroxy abscisic acid.<br />

Phytochemistry 52: 1185-1193.<br />

ÅSTOT, C., DOLEZAL, K., NORDSTRÖM, A., WANG, Q., KUNKEL, T., MORITZ, T., CHUA, N.H. and SANDBERG,<br />

G. (2000) An alternative cytokinin biosynthesis pathway. Proc. Natl. Acad. Sci. USA 97: 14778-14783.<br />

BARRY, G.F., ROGERS, S.G., FRALEY, R.T. and BRAND, L. (1984) Identification of a cloned cytokinin<br />

biosynthetic gene. Proc. Natl. Acad. Sci. USA 81: 4776-4780.<br />

BASSIL, N.V., MOK, D. and MOK, M.C. (1993) Partial Purification of a cis-trans-Isomerase of Zeatin from<br />

Immature Seed of Phaseolus vulgaris L. Plant Physiol. 102: 867-872.<br />

BEWLEY, J.D. and BLACK, M. (1982) Physiology and biochemistry of seeds. (Volume 2). Viability, dormancy<br />

and environmental control. (Berlin, Germany: Springer-Verlag).<br />

BORTHWICK, H.A., HENDRICKS, S.B., PARKER, M.W., TOOLE, E.H. and TOOLE, V.K. (1952) A reversible<br />

photoreaction controlling seed germination. Proc. Natl. Acad. Sci. USA 38: 662-666.<br />

BUTLER, W.L., NORRIS, K.H., SIEGELMAN, H.W. and HENDRICKS, S.B. (1959) Detection, assay, and<br />

preliminary purification of the pigment controlling phtoresponsive development of plants. Proc. Natl.<br />

Acad. Sci. USA 45: 1703-1708.<br />

CHAPPLE, C. (1998) Molecular genetics analysis of plant cytochrome P450-dependent monooxygenases.<br />

Annu. Rev. Plant Physiol. Mol. Biol. 49: 311-343.<br />

CHEN, C.M. and HALL, R.H. (1969) Biosynthesis of N 6 -(d2-isopentenyl) adenosine in the transfer ribonucleic<br />

acid of cultured tabacco pith tissue. Phytochemistry 8: 1687-1696.<br />

CHEN, C.M. and LEISNER, S.M. (1984) Modification of cytokinins by cauliflower microsomal enzymes. Plant<br />

Physiol. 75: 442-446.<br />

CONE, J.W. and SPRUIT, C.J.P. (1983) Imbibition conditions and seed dormancy of Arabidopsis thaliana.<br />

Physiol. Plant. 59: 416-420.<br />

CUTLER, A.J. and KROCHKO, J.E. (1999) Formation and breakdown of ABA. Trends Plant Sci. 4: 472-478.<br />

DAVIES, P.J. (1995) Plant Hormones: Physiology, Biochemistry and Molecular Biology. (Dordrecht, The<br />

Netherlands: Kluwer Academic Publishers).<br />

DERKX, M.P.M., VERMEER, E. and KARSSEN, C.M. (1994) Gibberellins in seeds of Arabidopsis thaliana:<br />

Biological activities, identification and effects of light and chilling on endogenous levels. Plant Growth<br />

Regul. 15: 223-234.<br />

ESTEVEZ, J.M., CANTERO, A., ROMERO, C., KAWAIDE, H., JIMENEZ, L.F., KUZUYAMA, T., SETO, H., KAMIYA,<br />

Y. and LEON, P. (2000) Complementation of albino mutant cla1-1 of Arabidopsis thaliana by 1-deoxy-Dxylulose:<br />

A CLA1 gene require for the non-mevalonate pathway. Plant Physiol. 124: 95-104.<br />

GILLARD, D.F. and WALTON, D.C. (1976) Abscisic acid metabolism by a cell-free preparation from<br />

Echinocystis lobata liquid endosperm. Plant Physiol. 58: 790-795.<br />

HARRISON, M. A. and WALTON, D.C. (1975) Abscisic acid metabolism in water-stressed bean leaves. Plant<br />

Physiol. 56: 250-254.<br />

HEDDEN, P. and KAMIYA, Y. (1997) Gibberellin biosynthesis: Enzymes, genes and their regulation. Annu. Rev.<br />

Plant Phys. Plant Mol. Biol. 48: 431-460.<br />

HEMMERLIN, A., HOEFFLER, J. F., MEYER, O., TRITSCH, D., KAGAN, I.A., GROSDEMANGE-BILLIARD, C.,<br />

ROHMER, M. and BACH, T.J. (2003) Cross-talk between the cytosolic mevalonate and the plastidial<br />

methylerythritol phosphate pathways in tobacco Bright Yellow-2 cells. J. Biol. Chem. 278: 26666-26676.


REGULATION OF PLANT HORMONE BIOSYNTHESES<br />

9<br />

HOLTON, T.A. and CORNISH, E.C. (1995) Genetics and biochemistry of anthocyanin biosynthesis. Plant Cell 7:<br />

1071-1083.<br />

INOUE, T., HIGUCHI, M., HASHIMOTO, Y., SEKI, M., KOBAYASHI, M., KATO, T., TABATA, S., SHINOZAKI, K.<br />

and KAKIMOTO, T. (2001) Identification of CRE1 as a cytokinin receptor from Arabidopsis. Narure 409:<br />

1060-1063.<br />

KAKIMOTO, T. (2001) Identification of plant cytokinin biosynthetic enzymes as dimethylallyl diphosphate:<br />

ATP/ADP isopentenyltransferases. Plant Cell Physiol. 42: 677-685.<br />

KASAHARA, H., HANADA, A., KUZUYAMA, T., TAKAGI, M., KAMIYA, Y. and YAMAGUCHI, S. (2002)<br />

Contribution of the mevalonate and methylerythritol phosphate pathways to the biosynthesis of<br />

gibberellins in Arabidopsis. J. Biol. Chem. 277: 45188-94.<br />

KASAHARA, H., TAKEI, K., UEDA, N., HISHIYAMA, S., YAMAYA. T., KAMIYA, Y., YAMAGUCHI, S. and<br />

SAKAKIBARA, H. (2004) Distinct isoprenoid origins of cis-and trans-Zeatin biosyntheses. J. Biol. Chem.<br />

279: 14049-14054.<br />

KOORNNEEF, M. and VAN DER VEEN, J.H. (1980) Induction and analysis of gibberellin sensitive mutants in<br />

Arabidopsis thaliana (L.) Heynh. Theor. Appl. Genet. 58: 257-263.<br />

KROCHKO, J.E., ABRAMS, G.D., LOEWEN, M.K., ABRAMS, S.R. and CUTLER, A.J. (1998) (+)-Abscisic acid 8’-<br />

hydroxylase is a cytochrome P450 monooxygenase. Plant Physiol. 118: 849-860.<br />

KUSHIRO, T., OKAMOTO, M., NAKABAYASHI, K., KITAMURA, S., ASAMI, T., HIRAI, N., KOSHIBA, T.,<br />

KAMIYA, Y. and NAMBARA, E. (2004) The Arabidopsis cytochrome P450 CYP707A encodes ABA8’-<br />

hydoroxylases: Key enzyme in ABA catabolism. The EMBO Jornal 23: 1647-1656.<br />

LICHITENTHALER, H.K. (1999) The 1-deoxy-D-xylulose-5-phosphate pathway of isoprenoid biosynthesis in<br />

plants. Annu. Rev. Plant Physiol. Plant Mol. Biol. 50: 47-65.<br />

MCCARTY (1995) Genetic control and integration of maturation and germination pathways in seed<br />

development. Annu. Rev. Plant Physiol. Plant Mol. Biol. 46: 71-93.<br />

MILBORROW, B.V. (1969) Identification of “Metabolite C” from abscisic acid and a new structure for phaseic<br />

acid. Chem. Commun. 966-967.<br />

MILBORROW, B.V. (1975) The absolute configuration of phaseic and dihrdrophaseic acids. Phytochemistry 14:<br />

1045-1053.<br />

MILBORROW, B.V., CARRINGTON, N.J. and VAUGHAN, G.T. (1988) The cyclization of 8’-hydroxy abscisic acid<br />

to phaseic acid in vivo. Phytochemistry 27: 757-759.<br />

MOK, D.W.S. and MOK, M.C. (2001) Cytokinin metabolism and action. Annu. Rev. Plant Physiol. Plant Mol.<br />

Biol. 52: 89-118.<br />

NAMBARA, E., AKAZAWA, T. and MCCOURT, P. (1991) Effects of the gibberellin biosynthetic inhibitor<br />

uniconazol on mutants of Arabidopsis. Plant Physiol. 97: 736-738.<br />

PIAGGESI, A., PICCIARELLI, P., CECCARELLI, N. and LORENZI, R. (1997) Cytokinin biosynthesis in endosperm<br />

of Sechium edule Sw.. Plant Sci. 129: 131-140.<br />

PIERCE, M. and RASCHKE, K. (1981) Synthesis and metabolism of abscisic acid in detached leaves of<br />

Phaseolus vulgaris L. after loss and recovery of turgor. Planta 153: 156-165.<br />

QIN, X. and ZEEVAART, J.A.D. (1999) The 9-cis-epoxycarotenoid cleavage reaction is the key regulatory step<br />

of abscisic acid biosynthesis in water-stressed bean. Proc. Natl. Acad. Sci. USA 96: 15354-15361.<br />

ROHMER, M. (2003) Mevalonate-independent methylerythritol phosphate pathway for isoprenoid biosynthesis.<br />

Elucidation and distribution. Pure Appl. Chem. 75: 375-387.<br />

ROSS, J.D. and BRADBEER, J.W. (1971) Studies in seed dormancy. V. The content of endogenous gibberellins<br />

in seeds of Corylus avellana L. Planta 100: 288-302.<br />

SCHULER, M.A. (1996) Plant cytochrome P450 monooxygenases. Crit. Rev. Plant Sci. 15: 235-284.<br />

SCHWARTZ, S.H., QIN, X. and ZEEVAART, J.A.D. (2003) Elucidation of the indirect pathway of abscisic acid<br />

biosynthesis by mutants, genes, and enzymes. Plant Physiol. 131: 1591-1601.<br />

SEO, M. and KOSHIBA, T. (2002) Complex regulation of ABA biosynthesis in plants. Trends Plant Sci. 7: 41-48.


10<br />

Yuji KAMIYA<br />

SHINOMURA, T. (1997) Phytochrome regulation of seed germination. J. Plant Res. 110: 151-161.<br />

SHROPSHIRE, J.W., KLEIN, W.H. and ELSTAD, V.B. (1961) Action spectra of photomorphogenic induction and<br />

photoinactivation of germination in Arabidopsis thaliana. Plant Cell Physiol. 2: 63-69.<br />

SINSKA, I., LEWAK, S., GASKIN, P. and MACMILLAN, J. (1973) Reinvestigation of apple-seed gibberellins.<br />

Planta 114: 359-364.<br />

SONDHEIMER, E., GALSON, E.C., TINELLI, E. and WALTON D.C. (1974) The metabolism of hormones during<br />

seed germination and dormancy. Plant Physiol. 54: 803-808.<br />

TAKEI, K., SAKAKIBARA, H. and SUGIYAMA, T. (2001) Identification of genes encoding adenylate<br />

isopentenyltransferase, a cytokinin biosynthesis enzyme, in Arabidopsis thaliana. J. Biol. Chem. 276:<br />

26405-26410.<br />

TAYA, Y., TANAKA, Y. and NISHIMURA, S. (1978) 5’-AMP is a direct precursor of cytokinin in Dictyostelium<br />

discoideum. Nature 271:545-547.<br />

TOYOMASU, T., KAWAIDE, H., MITSUHASHI, W., INOUE, Y. and KAMIYA, Y. (1998) Phytochrome regulates<br />

gibberellin biosynthesis during germination of photoblastic lettuce seeds. Plant Physiol. 118: 1517-1523.<br />

WALTON, D.C., and SONDHEIMER, E. (1972) Metabolism of 2- 14 C-()-abscisic acid in excised bean axes. Plant<br />

Physiol. 49: 285-289.<br />

WILLIAMS, P.M., BRADBEER, J.W., GASKIN, P. and MACMILLAN, J. (1974) Studies in seed dormancy VIII.<br />

The identification and determination of gibberellins A1 and A9 in seed of Corylus avellana L. Planta 117:<br />

101-108.<br />

XU, Z-J., NAKAJIMA, M., SUZUKI, Y. and YAMAGUCHI, I. (2002) Cloning and characterization of the abscisic<br />

acid-specific glucosyltransferase gene from Adzuki bean seedlings. Plant Physiol. 129: 1285-1295.<br />

YAMADA, H., SUZUKI, T., TERADA, K., TAKEI, K., ISHIKAWA, K., MIWA, K., YAMASHINO, T. and MIZUNO, T.<br />

(2001) The Arabidopsis AHK4 histidine kinase is a cytokinin-binding receptor that transduces cytokinin<br />

signals across the membrane. Plant Cell Physiol. 42: 1017-1023.<br />

YAMAGUCHI, S., SMITH, M.W., BROWN, R.G., KAMIYA, Y. and SUN, T.-P. (1998) Phytochrome regulation and<br />

differential expression of gibberellin 3b-hydroxylase genes in germinating Arabidopsis seeds. Plant Cell<br />

10: 2115-2126.<br />

YAMAUCHI, Y., OGAWA, M., KUWAHARA, A., HANADA, A., KAMIYA, Y. and YAMAGUCHI, S. (2004)<br />

Activation of gibberellin biosynthesis and response pathways by low temperature during imbibition of<br />

Arabidopsis thaliana seeds. Plant Cell 16: 367-378.<br />

ZEEVAART, J.A.D. (1980) Changes in the levels of abscisic acid and its metabolites in excised leaf blades of<br />

Xanthium strumarium during and after water stress. Plant Physiol. 66: 672-678.<br />

ZEEVAART, J.A.D. and CREELMAN, R.A. (1988) Metabolism and physiology of abscisic acid. Annu. Rev. Plant<br />

Physiol. Plant Mol. Biol. 39: 439-473.


REGULATION OF PLANT HORMONE BIOSYNTHESES<br />

11<br />

<br />

<br />

<br />

230-0045 1-7-22<br />

<br />

<br />

<br />

<br />

<br />

Stratification<br />

2 GA3 <br />

AtGA3ox1 AtGA3ox2 GC-<br />

MS <br />

AtGA3ox1 GA4 <br />

<br />

AtGA3ox1 T-DNA <br />

AtGA3ox1 <br />

<br />

MVA <br />

MEP <br />

cla1 <br />

MEP <br />

<br />

MEP <br />

trans- <br />

cis- trans- MEP<br />

cis- MVA <br />

<br />

<br />

<br />

<br />

P450


12<br />

Yuji KAMIYA<br />

8’ 8’ <br />

<br />

273 P450 <br />

<br />

4 NADPH <br />

<br />

CYP707A 8’


Gamma Field Symposia, No. 42, 2003 Institute of Radiation Breeding<br />

NIAS, Japan<br />

GA SIGNAL TRANSDUCTION<br />

13<br />

GIBBERELLIN RESPONSE AND SIGNAL TRANSDUCTION<br />

Motoyuki ASHIKARI, Hironori ITOH, Miyako UEGUCHI and Makoto MATSUOKA<br />

Bioscience and Biotechnology Center, Nagoya University<br />

Furoucho, Chikusaku, Nagoya, 464-8601<br />

E-mail : ashi@agr.nagoya-u.ac.jp<br />

Introduction<br />

GAs are a large family of tetracyclic diterpenoid plant growth regulators and have been<br />

reported to be associated with a number of plant growth and development processes such as seed<br />

germination, stem elongation, flowering and fruit development (Reid 1993; Hooley 1994; Ross et<br />

al. 1997). GA-related mutants in plants show dwarf or elongated phenotypes, and these mutants are<br />

crucial for elucidating the regulatory mechanisms governing the GA biosynthetic and signal<br />

transduction pathways. Because dwarf characteristics are favored in plant breeding, the study of<br />

these characteristics has applications not only for understanding basic plant biology but also for<br />

molecular breeding. Many GA-related mutants have been isolated from numerous plant species<br />

(Reid 1993; Hooley 1994; Ross et al. 1997) and can be roughly classified into 2 categories: GAsensitive<br />

and GA-insensitive. A GA-sensitive mutant responds to exogenous GA because it cannot<br />

produce GA, or it produces insufficient GA due to a deficiency in genes encoding GA catalytic<br />

enzymes. On the other hand, a GA-insensitive mutant does not respond to exogenous GA, and gene<br />

related to GA-insensitivity may be associated with GA signal transduction (Hedden and Phillips<br />

2000; Olszewski et al. 2002). Here shows the rice GA insensitive mutants and their gene functions.<br />

slender rice 1 (slr1) mutant<br />

The slender rice 1 (slr1) mutant show a slender phenotype with an elongated stem and leaf<br />

and reduced root number and length, which is similar to that of rice plants treated with GA3 (Fig. 1)<br />

(Ikeda et al. 2001; Itoh et al. 2002). The slr1 mutant was first identified on the basis of its<br />

abnormal elongation phenotype at the seedling stage, which is similar to the appearance of wildtype<br />

rice plants infected by “Bakanae-disease”. In fact, it is difficult to distinguish between slr1 and<br />

“Bakanae- disease” plants. The slr1 phenotype seems to be the result of saturation with GAs,<br />

however, the levels of endogenous GAs (GA19, GA20 and GA1) in slr1 are actually lower than in the<br />

wild-type. Also, GA-inducible -amylase (Ramy1A) is produced in the aleurone cells in the<br />

absence of GA application. However, the GA-saturation phenotype of slr1 is not affected by


14<br />

Motoyuki ASHIKARI, Hironori ITOH, Miyako UEGUCHI and Makoto MATSUOKA<br />

Fig. 1 Gross morphology of slender1 (slr1) and domain structure of SLR1<br />

treatment with uniconazole, a GA biosynthesis inhibitor (Ikeda et al. 2001). These results indicate<br />

that slr1 is a constitutive GA response mutant and that the SLR1 protein may be associated with<br />

GA signal transduction as a negative regulator (Ikeda and et al. 2001; Itoh et al. 2002).<br />

The SLR1 gene has been isolated by linkage analyses between a rice gene homologous to<br />

Arabidopsis GAI and the slender phenotype. Some slr1 alleles contain a nucleotide substitution or<br />

deletion that disrupts the open reading frame, and therefore these are considered to be loss-of-function<br />

alleles. Actually, the introduction of the wild-type SLR1 gene complements the slender mutation<br />

(Ikeda et al. 2001). On the basis of these findings, the SLR1 gene is regarded to be homologous to<br />

Arabidopsis GAI, which encodes a putative repressor protein for the GA signaling pathway.<br />

The SLR1 protein shares high amino acid identity with Arabidopsis GAI (47.2 %), RGA<br />

(41.2 %), wheat RHT-D1a (77.2 %) and d8 (80.3 %). The SLR1 gene is located on the long arm of<br />

rice chromosome 3, a region which shows the genome synteny with the wheat Rht locus of<br />

chromosome 4 and maize D8 locus of chromosome 1, confirming that these genes of grass species<br />

are orthologous (Peng et al. 1999; Ikeda et al. 2001).<br />

The deduced SLR1 protein has 625 amino acid residues and contains the DELLA, TVHYNP<br />

domain (called regions I and II in GAI) in the N-terminal region which is conserved among<br />

Arabidopsis GAI and RGA, wheat RHT and maize d8 (Peng et al. 1999) (Fig. 2). SLR1 also contains<br />

other consensus domains at the C-terminal region, such as a leucine heptad repeat, NLS, VHIID,<br />

PFYRF and SAW, which belong to the GRAS family (Pysh et al. 1999). Since proteins in the GRAS<br />

family, including Arabidopsis SCR (Laurenzio et al. 1996), are considered to function as<br />

transcriptional factors, SLR1 may have a similar role. Biochemical analyzes of SLR1, namely nuclear<br />

localization and transcriptional activity, support this idea (Itoh et al. 2002; Ogawa et al. 2000).<br />

To investigate the function of SLR1 in plants, we have generated transgenic rice plants that


GA SIGNAL TRANSDUCTION<br />

15<br />

Fig. 2 Schematic structure of SLR1<br />

constitutively produce the SLR1-GFP protein under the control of the rice Actin1 promoter. These<br />

transgenic plants show the dwarf phenotype, supporting the idea that SLR1 functions as a negative<br />

regulator of GA signaling (Itoh et al. 2002). The GFP signal is localized in the nucleus but<br />

disappears following treatment with GA3; this effect is accompanied by leaf and stem elongation.<br />

The disappearance of SLR1 in response to GA3 treatment has been confirmed by immunoblot<br />

analysis using an anti-SLR1 antibody (Itoh et al. 2002). Based on these results, we have proposed a<br />

model for SLR1 function whereby, in the absence of a GA signal, the SLR1 protein localized in the<br />

nucleus suppresses GA activity as a transcriptional regulator, but SLR1 rapidly degrades in<br />

response to a GA signal, thereby releasing the suppression of GA action (Itoh et al. 2002). Similar<br />

findings have also been reported for SLR1 homologous proteins: the Arabidopsis RGA protein and<br />

barley SLN protein are localized in the nuclei (Dill and Sun 2001; Silverstone et al. 2001; Gubler<br />

et al. 2002) and RGA and SLN disappear following the application of GA3 (Dill and Sun 2001;<br />

Silverstone et al. 2001; Fu et al. 2002; Gubler et al. 2002). This suggests that the suppressive<br />

action of SLR1, SLN1, and RGA in rice, barley, and Arabidopsis, respectively, is similar in the<br />

regulation of GA signaling.<br />

Unlike SLR1, RGA and SLN1 proteins, the GAI and RGL1 (RGA-like1) proteins in<br />

Arabidopsis are not degraded by the GA treatment (Fleck and Harberd 2002; Wen and Chang<br />

2002). There are two classes of the SLR1 orthologous proteins in Arabidopsis, one of which<br />

(RGA) disappears from the nucleus in response to GA-treatment, the other (GAI and RGL1) does<br />

not (Fleck and Harberd 2002).<br />

Dominant alleles in the Arabidopsis gai, wheat Rht-B1/Rht-D1, and maize D8 loci confer GAinsensitive<br />

mutants with the dwarf phenotype (Koornneef et al. 1985; Peng et al. 1993; Peng et al.<br />

1997; Harberd and Freeling 1989; Winkler and Freeling 1994). Molecular cloning of Arabidopsis<br />

GAI has demonstrated that the in-frame deletion of its N-terminal domain, DELLA (region I),<br />

induces the gai mutant (Peng et al. 1997). Similarly, wheat Rht-B1/Rht-D1 and maize D8 have<br />

mutations in their N-terminal domains, DELLA (region I) and TVHYNP (region II), as in GAI<br />

(Peng et al. 1999). Transgenic plants that overproduce a SLR1 protein truncated in the DELLA<br />

domain have a dominant dwarf phenotype similar to Arabidopsis gai (Ikeda et al. 2001; Itoh et al.<br />

2002). Interestingly, all of these mutants and transgenic plants that overproduce the truncated form<br />

of SLR1 show GA-insensitive characteristics. These results suggest that the N-terminal region<br />

involving the DELLA and TVHYNP domains may function as a receptor for upstream GA signals.


16<br />

Motoyuki ASHIKARI, Hironori ITOH, Miyako UEGUCHI and Makoto MATSUOKA<br />

To examine the function of the conserved domain of SLR1, we have performed a domain<br />

analysis of SLR1 using transgenic plants that overproduce various truncated SLR1 proteins.<br />

Transformants overproducing DELLA or TVHYNP show a severe dwarf phenotype and lack<br />

GA-responsiveness. Correspondingly, the DELLA and TVHYNP proteins do not degrade<br />

following GA treatment. These results strongly suggest that these N-terminal domains are involved<br />

in the perception of GA signals. In contrast to the N-terminal proteins, the C-terminal region<br />

containing the VHIID, PFYRE, and SAW domains is involved in the suppressive function of<br />

SLR1. This is supported by the finding that the null alleles of slr1 often contain nucleotide<br />

substitutions or deletions in the C-terminal region. Domain analysis has also revealed that there are<br />

an additional two functional domains in SLR1, that is, a dimer formation domain and a regulatory<br />

domain. As its name suggests, the dimer domain is important for formation of a dimer of SLR1,<br />

and proteins lacking this domain (LZ) do not retain their repressive function. Conversely, if a<br />

truncated SLR1 protein containing the dimer domain, but not the suppressive domain (C-ter<br />

protein), is overproduced in the wild-type, the transformants show the slender phenotype,<br />

demonstrating the dominant negative function of the truncated SLR1 containing the dimer domain.<br />

The regulatory domain, which is rich in serine/threonine residues, may be involved in the<br />

regulation of SLR1 repression activity (Itoh et al. 2002). In fact, it has been proposed that the<br />

activity or stability of SLR1 is regulated by O-GluNAcylation or phosphorylation via the action of<br />

the SPINDLY protein (Thornton et al. 1999) or kinase, with the serine/threonine residues as the<br />

target site<br />

Dill et al. (2001) have also performed a domain analysis of RGA in Arabidopsis using<br />

transgenic plants overproducing truncated RGA proteins. Transgenic plants with DELLA show<br />

the GA-insensitive severe dwarf phenotype and the protein is resistant to degradation following GA<br />

treatment. This also demonstrates that the DELLA motif is essential for GA-induced RGA degradation.<br />

Why do the loss-of-function alleles of RGA or GAI show an almost normal phenotype, even<br />

though rice slr1 and barley sln1 show the GA-constitutive response phenotype? For example, gait6,<br />

the loss-of-function allele of gai has wild-type features but has slightly increased resistance to<br />

paclobutrazol (PAC), an inhibitor of GA biosynthesis. This has been explained by a functional<br />

redundancy of GAI, RGA and other orthologous proteins. Indeed, RGA has a highly similar<br />

structure to that of GAI, and also works as a negative regulator of GA signaling (Silverstone et al.<br />

1998). Consequently, the loss-of-function of RGA does not result in a typical constitutive GA<br />

response phenotype but rather a partial suppression of the dwarf phenotype conferred by the GAdeficient<br />

mutation, ga1-3 (Silverstone et al. 1997; Silverstone et al. 1998). Double mutants gai/gai,<br />

rga/rga do not show the slender phenotype, but it slightly higher than wild. It is probably due to the<br />

presence of redundant genes, RGLs (RGL1, RGL2 and RGL3) (Gill and Sun 2001). However,<br />

recently it is reported that RGL1 and RGL2 play a larger role in seed germination than does GAI or<br />

RGA which are mainly associate with stem elongation (Wen and Chang 2002; Lee et al. 2002). In<br />

contrast, the barley sln1 mutant has the slender phenotype (Foster 1977) and induces -amylase


GA SIGNAL TRANSDUCTION<br />

17<br />

expression without GA treatment, as is the case in rice (Chandler 1988; Lanahan and Ho 1988;<br />

Croker et al. 1990). The rice and barley genomes have only one gene encoding an orthologous<br />

protein to GAI/RGA (Chandler et al. 2002). Such non-redundancy of GA-related genes in rice<br />

should provide an advantage for studying the GA signal transduction pathway.<br />

gid2 mutant<br />

The gid2 mutant lines show a severe dwarf phenotype with wide leaf blades and dark green<br />

leaves (Fig. 3), which are features of GA-related mutants such as d1 and d18 (Ashikari et al. 1999;<br />

Itoh et al. 2001). gid2 does not show any GA-responsiveness when measured against the three<br />

criteria as follows, second leaf sheath elongation, -amylase induction in aleurone, and feed-back<br />

expression of GA20 oxidase. Moreover, even though the gid2 mutants have severe dwarfism, they<br />

accumulate more than 150 times the level of bioactive GA1 than that in wild-type plants. Given the<br />

GA-insensitivity of the gid2 mutant, we expect that the GID2 gene encodes a positive regulator of<br />

GA signaling.<br />

To elucidate the molecular function of GID2, the gene has been isolated by positional cloning.<br />

Genetic analysis enabled us to narrow-down the gid2 mutation to a 13kb region on rice<br />

chromosome 2. A comparison of the nucleotide sequence of this region between gid2 and the wildtype<br />

revealed that all three gid2 alleles have nucleotide substitutions or deletions in one putative<br />

gene that introduce novel stop codons, suggesting that these are null alleles. Introduction of a wild<br />

DNA fragment spanning the entire region of the candidate gene into the gid2 mutant rescues the<br />

Fig. 3 Gross morphology of gibberellin-insensitive<br />

dwarf 1 and 2 mutants (gid1 and gid2)<br />

Left: wild-type, Right: gid2


18<br />

Motoyuki ASHIKARI, Hironori ITOH, Miyako UEGUCHI and Makoto MATSUOKA<br />

gid2 phenotype to normal. The GID2 gene encodes a 636bp open reading frame, capable of<br />

producing a polypeptide of 212 amino acid residues. The deduced amino acid sequence of GID2<br />

contains an F-box domain, which is a conserved motif of F-box proteins that form a component of<br />

an E3 ubiquitin-ligase complex. The F-box sequence in GID2 is well conserved in other F-box<br />

proteins from Arabidopsis, yeast, mold, and humans. Many F-box proteins contain a proteinprotein<br />

interaction domain, such as leucine-rich repeat (LRR) or WD-40 repeat sequences outside<br />

the F-box (Dashaies 1999; Yang et al. 1999; Li and Jonston 1997; Skowyra et al. 1997; Winston et<br />

al. 1999). However, we have not found any conserved motifs outside the F-box in the GID2<br />

structure, but the structure of GID2 is similar to that of Arabidopsis SLY1 protein which is<br />

considered to be a positive regulator of GA signaling in Arabidopsis (McGinnis et al. 2003) It is<br />

very likely that the rice GID2 and Arabidopsis SLY1 are orthologous proteins.<br />

As described above, the SLR1 protein functions as a repressor of GA signaling in rice and its<br />

degradation is essential for the downstream action of GA. Since the GID2 gene encodes a F-box<br />

protein, which is a component of a SCF complex (E3 ubiquitin-ligase complex), we thought that<br />

the SLR1 protein might be targeted for degradation by the SCF complex in a GA-dependent<br />

manner. Immunoblot analysis with an anti-SLR1 antibody has revealed that the SLR1 protein<br />

accumulates at a high level in the gid2 mutant, whereas it is only present at low levels in the wildtype.<br />

The immunoreactive SLR1 protein in the wild-type is degraded following GA3 treatment, but<br />

this does not occur in the gid2 mutant. These findings indicate that the GA-dependent degradation<br />

of SLR1 is defective in gid2 and therefore SCF GID2 may directly target the SLR1 protein for<br />

degradation through ubiquitination.<br />

Interestingly, there are two immunoreactive bands with different mobilities on SDS-PAGE in<br />

the gid2 mutant whereas only one band is detected in the wild-type. (Sasaki et al. 2003). In gid2,<br />

the band with higher mobility (Form I) has the same mobility as the protein synthesized in E. coli,<br />

indicating that this band corresponds to the nascent protein of SLR1. We suspect that the band with<br />

lower mobility (Form II) may be an intermediate in the SLR1 degradation process (Sasaki et al.<br />

2003). Actually, the band with higher mobility is not detected under natural SCF GID2 functional<br />

conditions. The appearance of a band with higher mobility has also been noted in the barley sln1d<br />

mutant (Gubler and others 2002), and therefore may be a common part of the degradation process<br />

of the SLR1/ RGA/ SLN1 proteins.<br />

Treatment of a crude extract of gid2 with calf intestine alkaline phosphatase (CIP) prior to<br />

immunoblotting leads to the disappearance of Form II SLR1.This suggests that Form II is a<br />

phosphorylated form of the SLR1 protein. Phosphorylation of SLR1 has also been examined by in<br />

vivo labeling with radioactive phosphate, 32 - PO4 (Sasaki et al. 2003). When the wild-type plants<br />

were treated with 32 - PO4 , we detected one faint radioactive SLR1 band, which disappeared<br />

following GA3 treatment. In contrast, one strong radioactive band was observed when the gid2<br />

plants were treated with 32 - PO4 , and its intensity was increased by GA3 treatment. These results<br />

suggest to us that GA increases the phosphorylated form of SLR1 and leads to its degradation by


GA SIGNAL TRANSDUCTION<br />

19<br />

interacting with the SCF GID2 complex. In gid2, the phosphorylation of SLR1 also occurs following<br />

GA3 treatment but the degradation does not occur due to the loss-of-function of the GID2 protein,<br />

and consequently the SLR1 protein is accumulated (Sasaki et al. 2003). This model is consistent<br />

with previous findings in yeast, mammals and plant, that is, phosphorylation of a target protein<br />

triggers the degradation process (Deshaies 1999). A recent publication describes the inhibition of<br />

barley SLN1 protein degradation by a proteasome inhibitor (Fu et al. 2002). This supports the<br />

notion that the SLR1 protein is degraded through the proteasome.<br />

Conclusions and Prospects<br />

Based on the results described in this review, we conclude that SLR1 functions as a molecular<br />

switch in GA signaling in rice plants. Actually, whether GA activity occurs or not is readily<br />

determined by the absence or presence, respectively, of the functional SLR1 protein in the nucleus.<br />

GID2 encodes an F-box protein that may be a component of an SCF ubiquitin-ligase complex.<br />

The fact that GID2 encodes an F-box protein and SLR1 is highly accumulated in the gid2 mutant<br />

led us to speculate that GA-dependent degradation of SLR1 is mediated by the SCF GID2 complex.<br />

This is supported by the finding that a phosphorylated form of the SLR1 protein is also<br />

accumulated in gid2. So far, there are previous reports that phosphorylation of target proteins<br />

triggers SCF-mediated degradation, our results also indicate that GA-dependent phosphorylation of<br />

SLR1 triggers the ubiquitin-mediated degradation (Fig. 3), in a similar manner to the SCFmediated<br />

pathway in plant, yeast and animals. On the other hand, the mechanism by which SLR1<br />

perceives the GA signal is still unknown. It is possible that the other GA-insensitive dwarf gene,<br />

GID1, modifies the molecular structure of the SLR1 protein.<br />

Unlike other plant hormones, the GA receptor has not yet been identified. Identification of<br />

new mutants associated with GA signaling will be important for elucidating the mechanism of the<br />

GA signal transduction pathway, including identification of the GA receptor. As in the case of<br />

Fig. 4 Putative model for the GA signal transduction pathway in rice.


20<br />

Motoyuki ASHIKARI, Hironori ITOH, Miyako UEGUCHI and Makoto MATSUOKA<br />

SLR1, there is a tendency for the rice genome to have a single gene associated with GA signaling.<br />

This non-redundant relationship of GA signal-related genes in rice plants should facilitate the<br />

study of the GA signal transduction pathway.<br />

References<br />

1. Ashikari, M., Wu, J., Yano, M., Sasaki, T. and Yoshimura, A. (1999) Rice gibberellin-insensitive dwarf<br />

mutant gene Dwarf 1 encodes the alpha-subunit of GTP-binding protein. PNAS 96: 10284-10289.<br />

2. Chandler, P. M. (1988) Hormonal regulation of gene expression in the “slender” mutant of barley<br />

(Hordeum vulgare L.). Planta 175: 115-120.<br />

3. Chandler, P. M., Marion, P. A, Ellis, M. and Gubler, F. (2002) Mutants at the Slender1 locus of barley cv<br />

Himalaya. Molecular and physiological characterization. Plant Physiol 129: 181-190.<br />

4. Croker, S. J., Hedden, P., Lenton, J. R. and Stoddart, J. (1990) Comparison of gibberellin in normal and<br />

slender barley seedlings. Plant Physiol 94: 194-200.<br />

5. Dashaies, R. J. (1999) SCF and Cullin/Ring H2-based ubiquitin ligases. Annu. Rev. Cell Dev. Biol. 15:<br />

435-467.<br />

6. Dill, A., Jung, H. S., Sun, T. P. (2001) The DELLA motif is essential for gibberellin-induced degradation<br />

of RGA. PNAS 98: 14162-14167.<br />

7. Fleck, B. and Harberd, N. P. (2002) Evidence that the Arabidopsis nuclear gibberellin signaling protein<br />

GAI is not destabilized by gibberellin. Plant Journal. 32: 935-947.<br />

8. Foster, C. A. (1977) Slender: an accelerated extension growth mutant of barley. Barley Genetics<br />

Newsletter 7: 24-27.<br />

9. Fu, X., Richards, D. E., Ait-Ali, T., Hynes, L. W., Ougham, H., Peng, J. and Harberd, N. P. (2002)<br />

Gibberellin-mediated proteasome-dependent degradation of the barley DELLA protein SLN1 repressor.<br />

Plant Cell 14: 3191-3200.<br />

10. Gubler, F., Chandler, P. M., White, R. G., Llewellyn, D. J. and Jacobsen, J. V. (2002) Gibberellin<br />

signaling in barley aleurone cells. Control of SLN1 and GAMYB expression. Plant Physiol 129: 191-200.<br />

11. Harberd, N. P. and Freeling, M. (1989) Genetics of dominant gibberellin-insensitive dwarfism in maize.<br />

Genetics 121: 827-838.<br />

12. Hedden, P. and Phillips, A. L. (2000) Gibberellin metabolism: new insights revealed by the genes. Trend<br />

Plant Sci 5: 523-530.<br />

13. Hooley, R. (1994) Gibberellin: perception, transduction and responses. Plant Mol Bio 26: 1529-1555.<br />

14. Ikeda, A., Ueguchi-Tanaka, M., Sonoda, Y., Kitano, H., Koshioka, M., Futsuhara, Y., Matsuoka, M. and<br />

Yamaguchi, J. (2001) slender rice, a constitutive gibberellin response mutant, is caused by a null mutation<br />

of the SLR1 gene, an ortholog of the height-regulating gene GAI/RGA/RHT/D8. Plant Cell 13: 999-1010.<br />

15. Itoh, H., Ueguchi-Tanaka, M., Sentoku, N., Kitano, H., Matsuoka, M. and Kobayashi, M. (2001) Cloning<br />

and functional analysis of two gibberellin 3 beta -hydroxylase genes that are differently expressed during<br />

the growth of rice. PNAS 98: 8909-8914.<br />

16. Itoh, H., Ueguchi-Tanaka, M., Sato, Y., Ashikari, M. and Matsuoka, M. (2002) The gibberellin signaling<br />

pathway is regulated by the appearance and disappearance of SLENDER RICE1 in nuclei. Plant Cell 14:<br />

57-70.<br />

17. Koornneef, M., Elgersma, A., Hanhart, C. J., van Loenen-Martinet, E. P., van Rign, L., and Zeevaart, J. A.<br />

D. (1985) A gibberellin insensitive mutant of Arabidopsis thaliana. Physiol Plant 65: 33-39.<br />

18. Lanahan, M. B. and Ho, T. H. D. (1988) Slender barley: A constitutive gibberellin-response mutant.<br />

Planta 175: 107-114.<br />

19. Laurenzio, D. L., Wysocka-Diller, J., Malamy, J. E., Pysh, L., Helariutta, Y., Freshour, G., Hahn, M. G.,<br />

Feldmann, K. A. and Benfey, P. N. (1996) The SCARECROW gene regulates an asymmetric cell division


GA SIGNAL TRANSDUCTION<br />

21<br />

that is essential for generating the radial organization of the Arabidopsis root. Cell. 86:423-33.<br />

20. Lee, S., Cheng, H., King, K. E., Wang, W., He, Y., Hussain, A., Lo, J., Harverd, N. P. and Peng J. (2002)<br />

Gibberellin regulates Arabidopsis seed germination via RGL2, a GAI/RGA-like gene whose expression is<br />

up-regulated following imbibition. Gene Develop. 16 : 646-658.<br />

21. Li, F. N. and Johnston, M. (1997) Grr1 of Saccharomyces cerevisiae is connected to the ubiquitin<br />

proteolysis machinery through Skp1: coupling glucose sensing to gene expression and the cell cycle.<br />

EMBO J. 16: 5629-5638.<br />

22. McGinnis, K. M., Thomas, S. G., Soule, J. D., Strader, L. C., Zale, J. M., Sun, T. P. and Steber, C. M.<br />

(2003) The Arabidopsis SLEEPY1 gene encodes a putative F-BOX subunit of an E3 ubiquitin ligase.<br />

Plant Cell 15: 1120-1130.<br />

23. Ogawa, M., Kusano, T., Katsumi, M. and Sano, H. (2000) Rice giberellin-insensitive gene homolog, OsGAI,<br />

encodes a nuclear-localized protein capable of gene activation at transcriptional level. Gene 245: 21-29.<br />

24. Olszewski, N., Sun, T. P. and Gubler, F. (2002) Gibberellin signaling: biosynthesis, catabolism, and<br />

response pathway. Plant Cell 14: 61-80.<br />

25. Peng, J. and Harberd, N. P. (1993) Derivative alleles of the Arabidopsis gibberellin-insensitive (gai)<br />

mutation confer a wild-type phenotype. Plant Cell 5: 351-360.<br />

26. Peng, J., Carol, P., Richards, D. E., King, K. E., Cowling, R. J., Murphy, G. P. and Harberd, N. P. (1997)<br />

The Arabidopsis GAI gene defines a signaling pathway that negatively regulates gibberellin responses.<br />

Gene Develop. 11: 3194-3205.<br />

27. Peng, J., Richards, D. E., Hartley, N. M., Murphy, G. P., Devos, K. M., Flintham, J. E., Beales, J., Fish, L.<br />

J., Worland, A. J., Pelica, F., Sudhakar, D., Christou, P., Snape, J. W., Gale, M. D. and Harberd, N. P.<br />

(1999) ‘Green revolution’ genes encode mutant gibberellin response modulators. Nature 400: 256-261.<br />

28. Pysh, L. D., Wysocka-Diller, J. W., Camilleri, C., Bouchez, D. and Benfey, P. N. (1999) The GRAS gene<br />

family in Arabidopsis: sequence characterization and basic expression analysis of the SCARECROW-<br />

LIKE genes. Plant J. 18: 111-119.<br />

29. Reid, J. B. (1993) Plant hormone mutants. J Plant Growth Regul. 12: 207-226.<br />

30. Ross, J. J., Murfet, I. C. and Reid, J. B. (1997) Gibberellin mutants. Physiol. Plant 100: 550-560.<br />

31. Sasaki, A., Itoh, H., Gomi, K., Ueguchi-Tanaka, M., Ishiyama, K., Kobayashi, M., Jeong, D-H., An, G.,<br />

Kitano, H., Ashikari, M. and Matsuoka, M. (2003) Accumulation of phosphorylated repressor for<br />

gibberellin signaling in an F-box mutant. Science 299: 1896-1898.<br />

32. Silverstone, A. L., Mak, P. Y., Martinez, E. C. and Sun, T. P. (1997) The new RGA locus encodes a<br />

negative regulator of gibberellin response in Arabidopsis thaliana. Genetics. 146: 1087-1099.<br />

33. Silverstone, A. L, Ciampaglio, C. N. and Sun, T. S. (1998) The Arabidopsis RGA gene encodes a<br />

transcriptional regulator repressing the gibberellin signal transduction pathway. Plant Cell 10: 155-169.<br />

34. Silverstone, A. L, Jung, H. S., Dill, A., Kawaide, H., Kamiya, Y. and Sun, T. S. (2001) Repressing a<br />

repressor: Gibberellin-induced rapid reduction of the RGA protein in Arabidopsis. Plant Cell 13: 1555-1566.<br />

35. Skowyra, D., Craig, K. L., Tyers, M., Elledge, S. J. and Harper, J. W. (1997) F-box proteins are receptors<br />

that recruit phosphorylated substrates to the SCF ubiquitin-ligase complex. Cell 91: 209-219.<br />

36. Thornton, T. M., Swain, S. M. and Olszewski, N. E. (1999) Gibberellin signal transduction presents...the<br />

SPY who O-GlcNAc’d me. Trends Plant Sci. 4: 424-428.<br />

37. Wen, C. K. and Chang, C. (2002) Arabidopsis RGL1 encodes a negative regulator of gibberellin response.<br />

Plant Cell. 14:87-100.<br />

38. Winkler, R. G. and Freeling, M. (1994) Physiological genetics of the dominant gibberellin-nonresponsive<br />

maize dwarf, Dwarf8 and Dwarf9. Planta 193: 341-348.<br />

39. Winston, J. T., Strack, P., Beer-Romero, P., Chu, C. Y., Elledge, S. J. and Harper, J. W. (1999) The SCF-<br />

TRCP-ubiquitin ligase complex associates specifically with phosphorylated destruction motifs in IB and -<br />

catenin and stimulates IB ubiquitination in vitro. Genes Dev. 13: 270-283.<br />

40. Yang, M., Hu, Y., Lodhi, M., McCombie, W. R. and Ma, H. (1999) The Arabidopsis SKP1-LIKE1 gene is<br />

essential for male meiosis and may control homologue separation. PNAS 96: 11416-11421.


22 Motoyuki ASHIKARI, Hironori ITOH, Miyako UEGUCHI and Makoto MATSUOKA<br />

<br />

<br />

<br />

464-8601<br />

GA: gibberellin<br />

GA <br />

<br />

<br />

GA <br />

GA GA<br />

GA <br />

GA GA <br />

<br />

GA GA GA <br />

GA <br />

<br />

GA <br />

GA <br />

<br />

GA DELLA GA <br />

GA <br />

DELLA GA GA <br />

<br />

DELLA <br />

<br />

slr1 <br />

<br />

<br />

slender1


GA SIGNAL TRANSDUCTION<br />

23<br />

<br />

<br />

<br />

gid1 slr1 <br />

<br />

gid1 slender slender gid2 <br />

slender slender slender gid1 gid2 <br />

<br />

<br />

D1G <br />

<br />

D1 D1 <br />

- <br />

- <br />

D1 - <br />

100 <br />

<br />

G <br />

<br />

<br />

D1 <br />

100 <br />

D1 <br />

<br />

D1 <br />

<br />

<br />

slr1 <br />

<br />

gid2 <br />

slender <br />

gid2 slender <br />

<br />

<br />

<br />

<br />

gid1 gid2 D1 DM 2


24 Motoyuki ASHIKARI, Hironori ITOH, Miyako UEGUCHI and Makoto MATSUOKA<br />

<br />

gid2 <br />

10 <br />

D1 <br />

<br />

<br />

gid1 slender <br />

gid2 <br />

<br />

<br />

gid2 <br />

<br />

<br />

gid2gid2 <br />

5 10


Gamma Field Symposia, No. 42, 2003 Institute of Radiation Breeding<br />

NIAS, Japan<br />

CYTOKININ SIGNAL TRANSDUCTION AND TWO-COMPONENT REGULATORY SYSTEM<br />

25<br />

CYTOKININ SIGNAL TRANSDUCTION AND<br />

TWO-COMPONENT REGULATORY SYSTEM<br />

Atsuhiro OKA<br />

Institute for Chemical Research, Kyoto University<br />

Uji, Kyoto 611-0011, Japan<br />

Introduction<br />

Expression of bacterial genes is frequently modulated in response to specific environmental<br />

stimuli (e.g., the genes involved in Escherichia coli chemotaxis, Rhizobium nitrogen fixation, and<br />

Agrobacterium pathogenicity). In most such cases, several functionally related genes are<br />

collectively governed by a single regulatory system, thus constituting a regulon. Many of these<br />

regulons have a common regulatory feature, termed the two-component regulatory system (STOCK<br />

et al. 2000). This type of system consists of two signal transducers, a sensor protein histidine<br />

kinase anchored at the cell membrane and a response regulator present in the cytoplasm. The twocomponent<br />

regulatory system appears to be a powerful device for a wide variety of adaptive<br />

responses in bacterial cells, as the genome of each bacterial species encodes mostly 30-60 different<br />

pairs of sensor histidine kinases and response regulators. Although this regulatory system was<br />

initially thought to be specific to prokaryotes, many instances have since been uncovered in diverse<br />

eukaryotic species, including higher plants, yeast, fungi, and slime molds.<br />

Cytokinins are a class of plant hormones that are responsible for a variety of physiological<br />

events. The first cytokinin identified, kinetin, was purified from an autoclaved DNA sample<br />

through tracing its ability to promote cell division of tobacco parenchymal cells, and its chemical<br />

structure was identified as 6-furfurylaminopurine (for a history of cytokinin research, see OKA<br />

2003). Both natural and synthetic cytokinins are now known, and chemically they are either N 6 -<br />

substituted aminopurines (e.g., trans-zeatin, isopentenyladenine, kinetin, 6-benzylaminopurine) or<br />

diphenylurea derivatives (e.g., thidiazuron). Along with identification of these compounds, a great<br />

deal of physiological information on cytokinins has been accumulated. In addition to promoting<br />

cell division, cytokinins induce chloroplast development, seed germination, sink nutritional<br />

enhancement, release of lateral bad inhibition, inhibition of root elongation, formation of vascular<br />

bundles, increase of pollination efficiency, stomatal opening, cotyledon and leaf development,<br />

delay of leaf senescence, shoot formation from calli, etc. (MOK and MOK 2001). For more than two<br />

decades, a number of attempts were made to elucidate how cytokinins are recognized by plant cells


26<br />

Atsuhiro OKA<br />

and how their signals are transduced intracellularly. However, no accurate answers to these<br />

questions were obtained until 2001. Using both forward and reverse genetics approaches, two<br />

protein species, each resembling the bacterial histidine kinases and response regulators, were found<br />

to sense and to be activated by cytokinins, respectively (INOUE et al. 2001; SAKAI et al. 2001). This<br />

review will briefly summarize the characteristics of the prokaryotic two-component regulatory<br />

system, and then present the results of recent research on cytokinin signaling in Arabidopsis,<br />

focusing on the similarities and dissimilarities between the plant and prokaryotic two-component<br />

regulatory systems. For further information regarding the observations and experimental results<br />

presented without citations in the text to save space, the reader should refer to the appropriate<br />

reviews together with references therein (AOYAMA and OKA 2003; HABERER and KIEBER 2002;<br />

HWANG et al. 2002; OKA et al. 2002; STOCK et al. 2000; WURGLER-MURPHY and SAITO 1997).<br />

Phosphorelay signaling by bacterial two-component regulatory systems<br />

The two-component regulatory system was so named because of a set of two protein components<br />

that collectively control expression of the member genes of a regulon (Fig. 1). One component, a<br />

sensor protein histidine kinase, is generally composed of an individual N-terminal periplasmic<br />

domain with membrane-anchored regions, and a common C-terminal transmitter/kinase (HK)<br />

domain, which extends into the cytoplasm. The N-terminal domain together with the neighboring<br />

Fig. 1. Phosphorelay and the domain architecture of protein components involved in the two-component<br />

regulatory system. The arrow indicates transfer of the phosphoryl group after autophosphorylation<br />

of sensor histidine kinase. “H” within the HK (brown) and HPt (blue) domains and “D” within<br />

the RR domain (yellow) are histidine and aspartate residues that are targeted by the phosphoryl<br />

group. (a) represents the most popular phosphorelay example of the bacterial two-component<br />

regulatory system. Instances corresponding to (b), (c), and (d) are the Agrobacterium VirA-VirG<br />

pathogenicity system, the yeast Sln1-Ypd1-Ssk1 osmosis system, and the B. subtilis KinB-<br />

Spo0F-Spo0B-Spo0A sporulation system, respectively.


CYTOKININ SIGNAL TRANSDUCTION AND TWO-COMPONENT REGULATORY SYSTEM<br />

27<br />

regions are thought to be involved in monitoring environmental stimuli either directly or indirectly,<br />

whereas the HK domain phosphorylates its own specific histidine residue and then transfers the<br />

phosphoryl group to the other component, a cytoplasmic response regulator. It also consists of two<br />

domains. The common N-terminal region is called the signal receiver (RR) domain that contains,<br />

as a hallmark, two aspartates and one lysine residue (D-D-K) separated by invariant distances. The<br />

central aspartate acquires the phosphoryl group from the phospho-histidine on the HK domain. The<br />

RR domain is followed by an individual output domain, which is involved mainly in binding to DNA<br />

and activating transcription. Some sensor kinases, such as VirA (Agrobacterium phenolic compound<br />

sensor) and ArcB (E. coli anaerobic sensor), are called hybrid-type sensor kinases, in which an extra<br />

region resembling the RR domain of the cognate response regulator follows the HK domain.<br />

Bacterial two-component regulatory systems usually involve no additional component for<br />

signal transduction. However, the phosphoryl group is sometimes transferred through a bridge<br />

component that carries the histidine-containing phosphotransfer (HPt) domain (e.g., Bacillus<br />

subtilis Spo0B and E. coli ArcB), which exists either alone or as a portion of the sensor kinase.<br />

Another bridge component is the polypeptide molecule that harbors the RR domain without any<br />

obvious output domain (e.g., B. subtilis Spo0F). The RR domain on hybrid-type sensor kinases<br />

might be included in this category. Phosphotransfer always occurs as either His-to-Asp or Asp-to-<br />

His (Fig. 1). Thus, the two-component regulatory system is also called the His-Asp phosphorelay<br />

signal transduction system.<br />

A brief history of the early research on plant two-component regulatory systems<br />

By 1993, it was evident that a considerable number of bacterial two-component regulatory<br />

systems function in adaptive responses, and that the underlying molecular mechanism involves a<br />

unique His-Asp phosphorelay. At about the same time, one of the first eukaryotic sensor histidine<br />

kinases, Sln1, was discovered in the budding yeast Saccharomyces cerevisiae (OTA and<br />

VARSHAVSKY 1993). Sln1 is an osmosensor, the architecture of which closely resembles that of the<br />

bacterial hybrid-type kinase. Subsequently, two downstream components, Ypd1 of an HPt protein<br />

and Ssk1 of a response regulator, were identified. Phosphorelay in this system occurs in the order<br />

Sln1 (His)>Sln1 (Asp)>Ypd1 (His)>Ssk1 (Asp). The Sln1 RR domain and the Ypd1 HPt protein<br />

act as bridge components between Sln1 HK and Ssk1 RR. An interesting feature is that the Ssk1<br />

response regulator is not located at the end of the signal flow, further modulating the downstream<br />

Hog1 MAP kinase cascade, the underlying mechanism of which is not phosphorelay but consecutive<br />

protein-protein interactions. Exposure to high osmolarity inhibits Sln1 autophosphorylation, and<br />

the resulting accumulation of non-phosphorylated molecules of Ssk1 activates the Ssk2 MAPKKK<br />

(POSAS and SAITO 1998). Bacterial sensor kinases are generally active under conditions that are not<br />

conducive to survival, but the opposite is true in the case of Sln1. Furthermore, the genome<br />

sequence indicates that S. cerevisiae harbors no other typical two-component regulatory system.


28<br />

Atsuhiro OKA<br />

At almost the same time as the identification of Sln1, the Arabidopsis ETR1 gene encoding a<br />

hybrid-type histidine kinase was identified as the causative gene of a dominant ethylene-insensitive<br />

mutant (CHANG et al. 1993). It was later shown that the ETR1 protein actually binds to ethylene,<br />

and acts as an ethylene receptor (SCHALLER and BLEECKER 1995). Another histidine kinase gene,<br />

CKI1, was identified using an activation-tagging procedure with hypocotyl explants (KAKIMOTO<br />

1996). Overexpression of CKI1 promoted greening and shoot formation from calli in the absence<br />

of exogenous cytokinin, which is usually required for wild-type explants to generate green shoots.<br />

Therefore, the CKI1 protein was presumed to be a cytokinin sensor, although its involvement in<br />

cytokinin signaling is still not clear. The view was thus established in 1996 that the plant histidine<br />

kinases participate in the first step of some phytohormone signaling pathways, as in the case of<br />

bacterial adaptive responses. However, information regarding their expected partners, plant<br />

response regulators, was not available from studies with phytohormone-insensitive mutants.<br />

In 1996, we began in silico screening of higher plants for bacterial-like response regulators,<br />

and found that a considerable number of response regulator genes are distributed along the<br />

Arabidopsis and rice genomes. Furthermore, these loci encode proteins that can be classified into<br />

two groups. The A-type proteins are relatively small molecules consisting of only the RR domain<br />

with an extremely short stretch at the N- and/or C-termini, whereas the B-type proteins contain<br />

additional functional domains downstream of the RR domain (IMAMURA et al. 1998; SAKAI et al.<br />

1998). As the entire genome sequence of Arabidopsis is now available (The Arabidopsis Genome<br />

Initiative 2000), we know that Arabidopsis carries eleven genes for each of the A-type and B-type<br />

response regulators (ARR1 to ARR22), eleven sensor histidine kinase genes, and five HPt genes<br />

(AHP1 to AHP5). Moreover, there are five genes encoding phytochromes for light sensing (PHYA<br />

to PHYE), which show extremely weak homology to the HK domain (SCHNEIDER-POETSCH et al.<br />

1991), and seven genes for pseudo-response regulators (APRR), in which the aspartate<br />

phosphorylation target is substituted by glutamate. The architectures of all these components, with<br />

the exception of APRRs that are not relevant to this review, are illustrated schematically in Fig. 2.<br />

The CRE1 histidine kinase is a cytokinin sensor<br />

In early 2001, mutants were isolated that showed impaired generation of green calli from<br />

hypocotyl explants grown in the presence of exogenous cytokinin (INOUE et al. 2001). Another<br />

mutant was also reported, in which root elongation was less inhibited by cytokinin than in wildtype<br />

plants (UEGUCHI et al. 2001). Their causative genes were identified independently and found<br />

to be the same (CRE1/AHK4). This gene encodes a hybrid-type histidine kinase with an extra,<br />

atypical, RR segment (Fig. 2). Two additional genes, AHK2 and AHK3, resembling CRE1 were<br />

also found. These gene products actually function as cytokinin sensors, as demonstrated by<br />

functional complementation with heterologous systems (INOUE et al. 2001; SUZUKI et al. 2001;<br />

UEGUCHI et al. 2001). In S. cerevisiae cells, CRE1 functions as a substitute for the Sln1


CYTOKININ SIGNAL TRANSDUCTION AND TWO-COMPONENT REGULATORY SYSTEM<br />

29<br />

Fig. 2. Architecture of Arabidopsis proteins belonging to the family of two-component<br />

regulatory systems. The relative sizes of proteins and functional domains are<br />

drawn on an approximate scale. Vertical bars show putative membrane-spanning<br />

regions. Each functional domain is colored as in Fig. 1.<br />

osmosensor only when cytokinin is present in medium. Similar replacements are possible for the<br />

Schizosaccharomyces pombe Phk1/Phk2/Phk3 osmosensors and the E. coli RcsC regulator for<br />

extracellular polysaccharide synthesis. The signal flow in these systems appears to occur through<br />

the His-Asp phosphorelay, as (i) mutations in the putative phosphorylation sites of CRE1 (His-459<br />

and Asp-973) abolish the ability of CRE1 to complement the defect of Sln1, and (ii) the<br />

heterologous downstream components (Ypd1 and Ssk1) are absolutely required for complementation.<br />

Furthermore, CRE1 has the ability to bind to cytokinin chemicals, as shown using the membrane<br />

fractions derived from S. pombe synthesizing CRE1 (YAMADA et al. 2001). Therefore, CRE1 is<br />

able to deliver a signal to a non-cognate HPt protein, dependent on cytokinin, as a substitute for the<br />

respective sensor kinases in hetero-complementation analyses. As both N 6 -substituted aminopurines<br />

and diphenylurea derivatives are effective in this regard, CRE1 is a universal receptor for<br />

cytokinins. The wooden leg (wol) mutant, which is defective in generation of phloem and cambium<br />

cells in the root vasculature, carries a missense mutation in the CRE1 gene. This mutant protein<br />

cannot bind cytokinin. Cytokinin detected by CRE1 is thus implicated in asymmetrical cell


30<br />

Atsuhiro OKA<br />

division for xylem and phloem development during early embryogenesis (MÄHÖNEN et al. 2000).<br />

The visible wol phenotype indicates limited functional redundancy among CRE1, AHK2, and<br />

AHK3, although their expression patterns macroscopically overlap in the roots and other adult<br />

tissues. Thus, it is obvious that CRE1, AHK2, and AHK3 recognize cytokinin in a partly<br />

overlapping manner followed by initiation of His-Asp phosphorelay to HPt factors.<br />

The B-type response regulators are transcription factors activated by cytokinin<br />

The B-type response regulators have been identified utilizing sequence similarity to known<br />

RR domains (SAKAI et al. 1998). The ARR1 gene was the first to be cloned and has been examined<br />

extensively. Its translation product has an architecture similar to that of the typical bacterial<br />

response regulators. The RR domain is located at the N-terminal end, and is followed by the<br />

ARRM (or GARP) and Q domains (Fig. 2). The ARRM domain, which faintly resembles the DNA<br />

binding domain of mammalian Myb, has the ability to bind double-stranded DNA in a sequencespecific<br />

manner in vitro (5’-AGATT-3’). NMR spectroscopy indicated that it contains three -helices,<br />

of which the latter two constitute a helix-turn-helix motif, and the most C-terminal -helix<br />

together with the N-terminal flexible arm are involved in base-pair recognition (HOSODA et al.<br />

2002). The Q domain is rich in glutamine and proline residues, and is capable of activating<br />

transcription (SAKAI et al. 2000). In addition, there is a typical nuclear localization signal between<br />

RR and ARRM, together with a few additional elements along the ARR1 molecule. Therefore,<br />

ARR1 is equipped with all the functional domains essential for transcription factors. In fact, ARR1<br />

and its truncated version lacking the RR domain are localized almost constantly in nuclei, and<br />

overexpression of ARR1 leads to transcriptional activation of a reporter gene preceded by the<br />

ARR1 target sequence in its promoter region. Transactivation exerted by the truncated ARR1 is<br />

much higher than that by the full-length ARR1, indicating that the RR domain masks the ability of<br />

ARR1 to activate transcription (SAKAI et al. 2000). These structural and biochemical characteristics<br />

are common to ARR2, ARR10, and ARR11, and all of the eleven B-type ARRs may be<br />

functionally redundant.<br />

Most of the B-type ARR genes are probably expressed in all adult tissues at different levels,<br />

and at significantly higher levels in roots than in other tissues, at least in the cases of ARR1, ARR2,<br />

ARR10, and ARR11. Overexpression of ARR1 does not significantly affect plant morphology<br />

except for hypertrophic cotyledons, longer cotyledonary petioles, and shorter roots than wild-type<br />

plants. Removing a DNA region corresponding to the RR domain from this ARR1 transgene<br />

however results in severe phenotypic changes, such as occasional formation of ectopic shoots on<br />

the adaxial surface of cotyledons and growth inhibition with concomitant disordered cell<br />

proliferation around the shoot apex, depending on the expression level of the transgene. These<br />

serious phenotypes probably result from the constitutive transactivating function of the truncated<br />

ARR1, independent of a signal from an upstream component. The phenomena, such as shorter


CYTOKININ SIGNAL TRANSDUCTION AND TWO-COMPONENT REGULATORY SYSTEM<br />

31<br />

roots, ectopic shoot formation, and disordered cell division, suggest the occurrence of<br />

hypercytokinin responses. This has been verified by characterization of an ARR1 mutant and by<br />

demonstration that the cytokinin sensitivity of plants is correlated to the expression level of ARR1<br />

(SAKAI et al. 2001). Overexpression of ARR2 shows a delayed leaf senescence phenotype (HWANG<br />

and SHEEN 2001). Definitive evidence for the involvement of B-type ARRs in cytokinin signaling<br />

has been provided by using, as molecular markers, the A-type ARR genes, expression of which is<br />

induced by cytokinin without protein synthesis (see below). Higher levels of expression of ARR1<br />

lead to higher levels of induction of the A-type genes. Furthermore, artificial activation of ARR1<br />

without de novo protein synthesis elevates the levels of A-type gene transcripts (SAKAI et al. 2001).<br />

Thus, ARR1 is a transcriptional activator for the A-type genes. All of these observations are<br />

consistent with the view that ARR1 and most B-type ARRs are activated by the cytokinin signal,<br />

presumably derived through the CRE1 sensor.<br />

The A-type response regulator genes respond to cytokinin<br />

The A-type response regulators were identified both by their sequence similarity to known RR<br />

domains and by analysis of genes up-regulated upon cytokinin treatment, although no such upregulation<br />

was found for the B-type ARR genes. Subsequent studies revealed that exogenous<br />

cytokinin up-regulates the transcript levels of all the A-type genes to a greater or lesser degree<br />

depending on the gene (D’AGOSTINO et al. 2000). This up-regulation occurs without de novo<br />

protein synthesis, indicating that A-type genes are the genes primarily responsive to cytokinin.<br />

Therefore, the A-type ARRs appear to be effectors acting downstream from cytokinin signal<br />

transduction.<br />

Among the A-type genes, ARR4 has been studied most extensively. This gene product can be<br />

associated with the AtDBP1 and AtDBP2 DNA-binding proteins, and also with PhyB. The binding<br />

of phospho-ARR4 to the PhyB N-terminal portion stabilizes its active Pfr form, leading to<br />

elevation of red light sensitivity (SWEERE et al. 2001). ARR4 thus appears to modulate light<br />

signaling by PhyB, as an effector of the cytokinin signal. The expression patterns of the A-type<br />

genes significantly differ from one another, although all of these genes are expressed in all tissues.<br />

Their intracellular localization patterns are also different: some occur exclusively in the nucleus,<br />

while others have been detected in both the cytoplasm and the nucleus. These variations in the<br />

expression and intracellular localization patterns of the A-type ARRs may reflect their possible<br />

diverse functions as effectors of cytokinin signal transduction.<br />

The AHP (HPt) proteins bridge the cytokinin signal from CRE1 to ARR1<br />

The B-type ARRs are transcription factors that are activated by cytokinin in plant cells, while<br />

CRE1 together with its relatives are cytokinin sensors. Expression of the truncated ARR1 missing


32<br />

Atsuhiro OKA<br />

the RR domain in the wol mutant suppresses the wol phenotype, implying that ARR1 and probably<br />

also the other B-type members are actually located downstream of CRE1 in the cytokinin signaling<br />

cascade (Fig. 3). CRE1 is presumably anchored at the plasma membrane, whereas ARR1 is always<br />

localized in the nucleus, thereby preventing any direct association. Their functional relationship<br />

seems to be mediated by the HPt proteins encoded by the five AHP genes (SUZUKI et al. 2000), as<br />

(i) AHPs can potentially associate directly with several B-type and A-type ARRs, including ARR1,<br />

and some histidine kinases, such as ETR1 and CKI1, and (ii) hetero-complementation by CRE1 in<br />

yeast and eubacteria requires the respective HPt components. Although AHP molecules are small<br />

enough to diffuse through the nuclear pore complex, they are localized mainly in the cytoplasmic<br />

compartment. Some fractions might be translocated to the nucleus transiently upon cytokinin<br />

treatment. The five AHP genes have different expression patterns, which are not influenced by<br />

exogenous cytokinin. These gene products actually function as HPt factors in phosphorelay<br />

because (i) they act as substitutes for Ypd1, but interfere with the ability of CRE1 to complement<br />

the Sln1 defect in the yeast hetero-complementation system, and (ii) they have the potential to<br />

deliver the phosphoryl group to both B- and A-type ARRs under specific in vitro conditions. Based<br />

on this circumstantial evidence and analogies with bacterial systems, it is reasonable to assume that<br />

Fig. 3. Framework of intracellular cytokinin signal transduction pathway in Arabidopsis.<br />

The hybrid-type sensor histidine kinases, CRE1, AHK2, and AHK3, perceive<br />

cytokinin and phosphorylate their own conserved histidine residues. The phosphoryl<br />

group is transferred to HPt factors (AHPs) via the C-terminal RR domain of the<br />

sensor histidine kinases. AHPs carrying the phosphoryl group move into the nucleus<br />

and transfer it to ARR1 and other B-type response regulators. ARR1 transactivates<br />

cytokinin-responsive genes including the A-type response regulator genes.


CYTOKININ SIGNAL TRANSDUCTION AND TWO-COMPONENT REGULATORY SYSTEM<br />

33<br />

the cytokinin signal flows from CRE1 to ARR1 via AHP as the following phosphotransfer:<br />

cytokinin stimulus>CRE1 (His)>CRE1 (Asp)>AHP (His)>B-type ARR (Asp)>transactivation of A-<br />

type ARRs>cytokinin responses. It is unclear whether the A-type ARRs are directly involved in the<br />

cytokinin-dependent phosphorelay from CRE1, although they may compete with the B-type ARRs,<br />

negatively regulating the phosphorelay from CRE1 to B-type ARRs.<br />

Two-component regulatory systems for ethylene, osmosis, and light signaling<br />

Ethylene is a well-known phytohormone that modulates a wide range of physiological actions,<br />

including apoptosis of leaves and flowers, fruit ripening, germination, and defense responses.<br />

Many ethylene-insensitive mutants have been isolated by screening for alterations in these<br />

attributes. Among the causative agents, the ETR1, ETR2, and EIN4 genes, encode the hybrid-type<br />

histidine kinase, whereas their two homologs ERS1 and ERS2, which were identified based on<br />

similarities to ETR1 and EIN4, encode prototypal histidine kinases. From the Arabidopsis genome<br />

sequence, we know that there is no additional gene encoding ethylene-related histidine kinases. As<br />

ETR1 binds to ethylene, these kinases function as ethylene receptors. Although the individual<br />

abolition of each of these five genes generally results in slight phenotypic changes, simultaneous<br />

alterations in three or more of the five genes result in constitutive ethylene responses (HUA and<br />

MEYEROWITZ 1998). Therefore, these gene products overlap functionally, and negatively control<br />

the downstream pathway for ethylene responses. CTR1, which interacts directly with ETR1 and<br />

resembles Raf1 (a member of the MAPKKK family), plays a leading role in negative regulation,<br />

because CTR1 mutants show constitutive ethylene responses. CTR1 is activated by ETR1 in the<br />

absence of ethylene to block the downstream pathway for ethylene responses, whereas CTR1<br />

remains inactive in the presence of ethylene, thereby relieving the negative regulation of the<br />

ethylene responses. This regulatory process is similar to that of Sln1 in modulating Ssk2<br />

MAPKKK through Ypd1 and Ssk1. No biochemical or genetic connections have been established<br />

between the ethylene sensors and the downstream HPt proteins and response regulators. As the<br />

ethylene sensors, at least ETR1, bear the enzyme activity of histidine kinase (probably in the<br />

absence of ethylene) and thus may also deliver the phosphoryl group to HPts and then to ARRs,<br />

they seem to be able to attend cytokinin signaling through the pool of five AHPs and eleven B-type<br />

ARRs (and possibly eleven A-type ARRs), generating cross-talk between ethylene and cytokinin<br />

signaling (Fig. 4). This cross-talk appears consistent with the observation that cytokinin and<br />

ethylene have roughly opposite physiological roles. Cytokinin, as a hormone involved in vitality,<br />

stimulates cell division, whereas ethylene provokes maturation and senescence and represents a<br />

hormone for aging and apoptosis.<br />

Among the eleven sensor histidine kinases of Arabidopsis, AHK1 was initially considered an<br />

osmosensor, based on both its relatively high degree of homology to the yeast osmosensor Sln1<br />

and its potential to substitute for Sln1 in yeast cells. However, definitive evidence regarding the


34<br />

Atsuhiro OKA<br />

Fig. 4. Cross-talk between cytokinin, ethylene, and red light signal transduction systems<br />

through the pool of response regulators and HPt proteins in plant cells (green<br />

ellipse). Cytokinin signaling occurs synergistically with light signaling, but is<br />

counterbalanced by ethylene signaling. Ligand binding to AHK1, CKI1, and<br />

CKI2 might also initiate phosphorelay to AHPs and participate in the cross-talk.<br />

role of AHK1 is currently lacking. As described above, CKI1 together with CKI2 missing the<br />

membrane-spanning regions were identified by altering cytokinin-dependence through activation<br />

tagging. Although there is no direct evidence that CKI1 and CKI2 participate in cytokinin<br />

recognition, they may contribute to cytokinin sensing in certain tissues and cells, e.g., CKI1 is<br />

expressed specifically in female gametes. There is no cognate response regulator corresponding to<br />

AHK1, CKI2, and CKI2, but they seem to have the histidine kinase activity, perhaps triggering<br />

phosphorelay to the pool of AHKs and ARRs in plant cells. Therefore, these three histidine kinases<br />

might also affect cytokinin signaling similarly to the ethylene sensors.<br />

In addition to these eleven sensor histidine kinases, five phytochromes (PHYA to PHYE)<br />

belong, in a broad sense, to this category. However, the histidine kinase-like structure located in the<br />

C-terminus of the phytochromes is considerably diverged from the typical histidine kinase, and the<br />

histidine residue that corresponds to the phosphorylation site is missing. As mentioned above,<br />

ARR4 associates specifically with PhyB to preserve the active Pfr form for longer periods. Light<br />

signaling acts synergistically with cytokinin to produce multifarious physiological effects, such as<br />

chloroplast development. As the expression of ARR4, as well as the other A-type ARR genes, is<br />

induced by cytokinin through the CRE1-ARR1 phosphorelay, it is likely that cytokinins indirectly<br />

enhance red light signaling. Therefore, light signaling is another participant in the cross-talk that<br />

occurs between the cytokinin and ethylene signal transduction systems (Fig. 4).<br />

Summary<br />

This review has presented a framework for the Arabidopsis two-component regulatory<br />

systems that are involved in the intracellular transduction of signals from cytokinin, ethylene, and<br />

light. With regard to cytokinin signal transduction, cytokinin recognition by CRE1, AHK2, and<br />

AHK3 is followed by a phosphorelay signal transfer that involves AHPs and B-type ARRs (Fig. 3).<br />

The organization of this system, the structural characteristics of the component proteins, and their


CYTOKININ SIGNAL TRANSDUCTION AND TWO-COMPONENT REGULATORY SYSTEM<br />

35<br />

molecular activities resemble those of bacterial systems, suggesting that the His-Asp phosphorelay<br />

signal transduction system is a powerful device in plat cells for a variety of environmental stress<br />

responses. Meanwhile, characteristics that are not seen in prokaryotes have been elucidated for<br />

plants. Three sensor kinases, five AHPs, and multiple B-type ARRs (probably almost all of the B-<br />

type ARRs) are involved in cytokinin signal transduction and transactivate multiple A-type ARR<br />

genes. In prokaryotes, a histidine kinase generally partners with a cognate response regulator in a<br />

specific one-to-one manner, and phosphorelay swapping among multiple molecular species is rare.<br />

Furthermore, there is no equivalent of ETR1 in prokaryotes, for which a partner response regulator<br />

is absent. Although the elements of the two-component regulatory systems of plants do not differ<br />

markedly from those of bacteria, the plant systems collectively employ complex cross-talk<br />

networks to govern different adaptive responses.<br />

References<br />

AOYAMA, T. and OKA, A. (2003). Cytokinin signal transduction in plant cells. J. Plant Res. 116: 221-223.<br />

CHANG, C., KWOK, S.F., BLEECKER, A.B. and MEYEROWITZ, E.M. (1993). Arabidopsis ethylene-response<br />

gene ETR1: similarity of product to two-component regulators. Science 262: 539-544.<br />

D’AGOSTINO, I.B., DERUERE, J. and KIEBER, J.J. (2000). Characterization of the response of the Arabidopsis<br />

response regulator gene family to cytokinin. Plant Physiol. 124: 1706-1717.<br />

HABERER, G. and KIEBER, J.J. (2002). Cytokinins. New insights into a classic phytohormone. Plant Physiol.<br />

128: 354-362.<br />

HOSODA, K., IMAMURA, A., KATOH, E., HATTA, T., TACHIKI, M., YAMADA, H., MIZUNO, T. and YAMAZAKI,<br />

T. (2002). Molecular structure of the GARP family of plant Myb-related DNA binding motifs of the<br />

Arabidopsis response regulators. Plant Cell 14: 2015-2029.<br />

HUA, J. and MEYEROWITZ, E.M. (1998). Ethylene responses are negatively regulated by a receptor gene family<br />

in Arabidopsis thaliana. Cell 94: 261-271.<br />

HWANG, I., CHEN, H.C. and SHEEN, J. (2002). Two-component signal transduction pathways in Arabidopsis.<br />

Plant Physiol. 129: 500-515.<br />

HWANG, I. and SHEEN, J. (2001). Two-component circuitry in Arabidopsis cytokinin signal transduction.<br />

Nature 413: 383-389.<br />

IMAMURA, A., HANAKI, N., UMEDA, H., NAKAMURA, A., SUZUKI, T., UEGUCHI, C. and MIZUNO, T. (1998).<br />

Response regulators implicated in His-to-Asp phosphotransfer signaling in Arabidopsis. Proc. Natl. Acad.<br />

Sci. USA 95: 2691-2696.<br />

INOUE, T., HIGUCHI, M., HASHIMOTO, Y., SEKI, M., KOBAYASHI, M., KATO, T., TABATA, S., SHINOZAKI, K.<br />

and KAKIMOTO, T. (2001). Identification of CRE1 as a cytokinin receptor from Arabidopsis. Nature 409:<br />

1060-1063.<br />

KAKIMOTO, T. (1996). CKI1, a histidine kinase homolog implicated in cytokinin signal transduction. Science<br />

274: 982-985.<br />

MÄHÖNEN, A.P., BONKE, M., KAUPPINEN, L., RIIKONEN, M., BENFEY, P.N. and HELARIUTTA, Y. (2000). A<br />

novel two-component hybrid molecule regulates vascular morphogenesis of the Arabidopsis root. Genes<br />

Dev. 14: 2938-2943.<br />

MOK, D.W.S. and MOK, M.C. (2001). Cytokinin metabolism and action. Annu. Rev. Plant Physiol. Plant Mol.<br />

Biol. 52: 89-118.<br />

OKA, A. (2003). New insights into cytokinins. J. Plant Res. 116: 217-220.


36<br />

Atsuhiro OKA<br />

OKA, A., SAKAI, H. and IWAKOSHI, S. (2002). His-Asp phosphorelay signal transduction in higher plants:<br />

receptors and response regulators for cytokinin signaling in Arabidopsis thaliana. Genes Genet. Sys. 77:<br />

383-391.<br />

OTA, I.M. and VARSHAVSKY, A. (1993). A yeast protein similar to bacterial two-component regulators. Science<br />

262: 566-569.<br />

POSAS, F. and SAITO, H. (1998). Activation of the yeast SSK2 MAP kinase kinase kinase by the SSK1 twocomponent<br />

response regulator. EMBO J. 17: 1385-1394.<br />

SAKAI, H., AOYAMA, T., BONO, H. and OKA, A. (1998). Two-component response regulators from Arabidopsis<br />

thaliana contain a putative DNA-binding motif. Plant Cell Physiol. 39: 1232-1239.<br />

SAKAI, H., AOYAMA, T. and OKA, A. (2000). Arabidopsis ARR1 and ARR2 response regulators operate as<br />

transcriptional activators. Plant J. 24: 703-711.<br />

SAKAI, H., HONMA, T., AOYAMA, T., SATO, S., KATO, T., TABATA, S. and OKA, A. (2001). ARR1, a<br />

transcription factor for genes immediately responsive to cytokinins. Science 294: 1519-1521.<br />

SCHALLER, G.E. and BLEECKER, A.B. (1995). Ethylene-binding sites generated in yeast expressing the<br />

Arabidopsis ETR1 gene. Science 270: 1809-1811.<br />

SCHNEIDER-POETSCH, H.A.W. (1992). Signal transduction by phytochrome: phytochromes have a module<br />

related to the transmitter modules of bacterial sensor proteins. Photochem. Photobiol. 56: 839-846.<br />

STOCK, A.M., ROBINSON, V.L. and GOUDREAU, P.N. (2000). Two-component signal transduction. Annu. Rev.<br />

Biochem. 69: 183-215.<br />

SUZUKI, T., MIWA, K., ISHIKAWA, K., YAMADA, H., AIBA, H. and MIZUNO, T. (2001). The Arabidopsis sensor<br />

His-kinase, AHK4, can respond to cytokinins. Plant Cell Physiol. 42: 107-113.<br />

SUZUKI, T., SAKURAI, K., IMAMURA, A., NAKAMURA, A., UEGUCHI, C. and MIZUNO, T. (2000). Compilation<br />

and characterization of histidine-containing phosphotransmitters implicated in His-to-Asp phosphorelay<br />

in plants: AHP signal transducers of Arabidopsis thaliana. Biosci. Biotechnol. Biochem. 64: 2486-2489.<br />

SWEERE, U., EICHENBERG, K., LOHRMANN, J., MIRA-RODADO, V., BÄURLE, I., KUDLA, J., NAGY, F.,<br />

SCHÄFER, E. and HARTER, K. (2001). Interaction of the response regulator ARR4 with phytochrome B in<br />

modulating red light signaling. Science 294: 1108-1111.<br />

THE ARABIDOPSIS GENOME INITIATIVE (2000). Analysis of the genome sequence of the flowering<br />

plant Arabidopsis thaliana. Nature 408: 796-815.<br />

UEGUCHI, C., SATO, S., KATO, T. and TABATA, S. (2001). The AHK4 gene involved in the cytokinin-signaling<br />

pathway as a direct receptor molecule in Arabidopsis thaliana. Plant Cell Physiol. 42: 751-755.<br />

WURGLER-MURPHY, S.M. and SAITO, H. (1997). Two-component signal transducers and MAPK cascades.<br />

Trends Biochem. Sci. 22: 172-176.<br />

YAMADA, H., SUZUKI, T., TERADA, K., TAKEI, K., ISHIKAWA, K., MIWA, K., YAMASHINO, T. and MIZUNO, T.<br />

(2001). The Arabidopsis AHK4 histidine kinase is a cytokinin-binding receptor that transduces cytokinin<br />

signals across the membrane. Plant Cell Physiol. 42: 1017-1023.


CYTOKININ SIGNAL TRANSDUCTION AND TWO-COMPONENT REGULATORY SYSTEM<br />

37<br />

<br />

<br />

<br />

611-0011<br />

http://molbio.dyndns.org/<br />

22 ARR1-<br />

ARR22<br />

2 ARR1 <br />

DNA ARR1 <br />

RR ARR1 RR ARR1 RR<br />

ARR1 ARR1 ARR1 RR <br />

ARR1 ARR1 <br />

ARR1 11 <br />

ARR <br />

ARR4 11 ARR ARR <br />

<br />

HK<br />

11 CRE1AHK2 AHK3 3 <br />

iCRE1 ii Sln1<br />

HK CRE1 <br />

iiiCRE1 <br />

CRE1 ARR1 RR <br />

ARR <br />

CRE1 <br />

ARR1 <br />

5 AHP HPt ARR1 <br />

ARR HK <br />

CRE1<br />

AHK2AHK3>AHPs> ARRs<br />

ARR ARR <br />

<br />

11 ARR


38 Atsuhiro OKA<br />

<br />

ARR4 PHYB <br />

PHYBPfr ARR4 <br />

<br />

<br />

ETR1 5 HK <br />

3 HK CKI1 CKI2 <br />

AHK1 <br />

ETR1 HK <br />

CKI1CKI2AHK1 <br />

HK CRE1 5 AHP <br />

11 ARR


CYTOKININ SIGNAL TRANSDUCTION AND TWO-COMPONENT REGULATORY SYSTEM<br />

39<br />

<br />

30 <br />

<br />

30 <br />

1 1<br />

2 <br />

3 11 11 <br />

6<br />

A


Gamma Field Symposia, No. 42, 2003 Institute of Radiation Breeding<br />

NIAS, Japan<br />

ETHYLENE-SIGNALING PATHWAY<br />

41<br />

MOLECULAR MECHANISMS FOR ETHYLENE PERCEPTION<br />

AND SIGNAL TRANSDUCTION<br />

Takashi HIRAYAMA 1), 2) and Tsutomu UGAJIN 2)<br />

1)<br />

Laboratory of Plant Molecular Biology, Yokohama Institute, RIKEN<br />

2)<br />

Graduate School of Integrated Science, Yokohama City Univ.<br />

1-7-29, Suehiro, Tsurumi, Yokohama, Kanagawa, 230-0045, Japan<br />

Introduction<br />

The gaseous plant hormone ethylene is involved in a variety of growth and developmental<br />

processes including germination, cell elongation, flower and leaf senescence, sex determination<br />

and fruit ripening. To gain insight into the molecular mechanisms of ethylene action, a molecular<br />

genetic approach has been applied using the ethylene-evoked triple response phenotype of<br />

Arabidopsis seedlings. Analysis of those ethylene related mutants allowed us to draw the overall<br />

structure of the ethylene signal transduction pathway, from ethylene perception to nuclear events.<br />

Based on the similarity of the sensor proteins, the ethylene signal transduction pathway has been<br />

thought to be similar to the yeast osmosensing signaling pathway. However, recent studies have<br />

revealed that the ethylene pathway is unique and quite complicated. Although the ethylenesignaling<br />

pathway is most understood among the signal transduction pathways for plant hormones,<br />

there are still many matters to be clarified. Here, we are going to describe the recent studies on the<br />

ethylene-signaling components and discuss the possible model for the ethylene-signaling pathway.<br />

Isolation of ethylene related mutants<br />

A gaseous hormone ethylene is involved in diverse developmental and physiological<br />

processes of plants. Treatment of etiolated seedlings with ethylene evokes dramatic morphological<br />

changes referred to as the “triple response” that includes exaggerated apical hook, radial swelling<br />

of hypocotyl and inhibition of hypocotyl and root elongation in Arabidopsis. These morphological<br />

changes are highly specific for ethylene. A genetic approach that relies on the triple response<br />

phenotype as a morphological marker has allowed the identification of several classes of mutants<br />

with impaired responses to ethylene (BLEECKER et al. 1988; GUZMAN and ECKER 1990). These<br />

mutants can be classified into several groups; ethylene insensitive mutant, etr1, etr2, ein2, ein3,<br />

ein4, ein5/ain1 and ein6; constitutive ethylene response mutant, eto1, 2, 3 and ctr1; and tissue-


42<br />

Takashi HIRAYAMA and Tsutomu UGAJIN<br />

specific ethylene response mutant hls1 and eir1; enhanced ethylene sensitive mutant, eer1; finally,<br />

altered ethylene recognition mutant, ran1. Based on the results from extensive genetic studies with<br />

these mutants, a model has been drawn for the ethylene-signaling pathway, in which identified<br />

components act in a linear pathway (Figure 1) (GUZMAN and ECKER 1990). Isolation of the<br />

corresponding genes and molecular analysis of encoded proteins have greatly facilitated our<br />

understanding of the ethylene-signaling pathway at the molecular level.<br />

Ethylene receptors<br />

Based on the genomic sequence, Arabidopsis has five ethylene receptors, namely, ETR1,<br />

ERS1, ETR2, EIN4 and ERS2. Structure of ethylene receptor protein is reminiscent of membranespanning<br />

histidine kinase. The three N-terminal membrane-spanning domains are necessary and<br />

sufficient for ethylene binding. This region is highly conserved among ethylene receptors. Analysis<br />

of mutants such as etr1-1 and knock-out mutants have shown that ethylene receptors negatively<br />

regulate down stream components (HUA and MEYEROWITZ 1998). In the absent of ethylene<br />

molecule, ethylene receptors are active and inhibit the ethylene response. On the other hands, in<br />

the presence of ethylene, ethylene receptors become inactive and let downstream ethylene-<br />

Figure 1 The ethylene-signaling pathway drawn based on the results obtained from genetic<br />

analyses. Genetically identified components involved in the ethylene response are<br />

shown. The allow heads indicate only the direction of the signal. RAN1 and EER1<br />

function as modifier for ethylene receptors and CTR1, respectively (see text).


ETHYLENE-SIGNALING PATHWAY<br />

43<br />

signaling pathway turns on. However, as discussed below, the biochemical property of active<br />

ethylene receptor has not been elucidated yet.<br />

As ethylene is a tiny olefin molecule, it has been proposed that transition metals such as<br />

copper or zinc ion are required for ethylene recognition. Bleecker’s group has demonstrated that<br />

copper (I) ion is required for ethylene binding, and that Cys65 and His69 residues in the second<br />

membrane spanning domain are required for copper coordination and ethylene binding activities<br />

using recombinant ETR1 proteins expressed in the budding yeast cell (RODRIGUEZ et al. 1999).<br />

Based on these results, they proposed a model for ethylene recognition in which two Cys65<br />

residues and two His69 residues of ETR1 dimer coordinate with one copper (I) ion that is able to<br />

bind one ethylene molecule (Figure 2). These results, however, did not offer any idea on the<br />

linkage between copper requirement and ethylene receptor activity. It could be speculated at this<br />

time that ethylene receptor is active without copper ion since the conversion Cys65 to Tyr (etr1-1<br />

mutation) confers an ethylene insensitive phenotype.<br />

It is the analysis of the ran1 mutants that offered in planta functional evidence for the copper<br />

requirement for both ethylene perception and the proper conformation of ethylene receptors<br />

(HIRAYAMA et al. 1999). The ran1 mutants were isolated in a screen for mutants with altered<br />

ethylene-recognition specificities aiming to gain insight into the mechanism of ethylene<br />

Figure 2 A proposed model for ethylene recognition. Two ethylene receptor<br />

molecules constitute a functional ethylene receptor complex. Cys65<br />

and His69 on the second membrane-spanning domain coordinate<br />

with one copper (I) ion that binds one ethylene molecule.


44<br />

Takashi HIRAYAMA and Tsutomu UGAJIN<br />

perception. The ran1 mutants show the ethylene phenotype in response to treatment with transcyclooctene,<br />

a potent ethylene binding inhibitor, which normally inhibits the ethylene response.<br />

The ran1 mutants seem to have a relaxed ligand-specificity since they respond normally to<br />

ethylene. The RAN1 gene has been identified by map based cloning and shown to encode a P-type<br />

copper transporter similar to human Menkes or Wilson disease protein, or yeast Ccc2p. These<br />

proteins have been demonstrated to deliver copper ions to the post Golgi compartment where<br />

copper requiring proteins are modified and sorted subsequently to the plasma membrane or outside<br />

of the cell, etc. These finding lead us the idea that RAN1 delivers copper ions to ethylene<br />

receptors. Several RAN1 co-suppressed transgenic lines and the null-type ran1 mutant (ran1-3)<br />

exhibit strong ethylene constitutive responding phenotypes (HIRAYAMA et al. 1999; WOESTE and<br />

KIEBER 2000). In the ran1-3 mutant, ethylene receptors are expressed at the same level as wild<br />

type (ZHAO et al. 2002). These results strongly suggest that ethylene receptors cannot function<br />

without copper delivery.<br />

However, this conclusion seems inconsistent with the etr1-1 phenotype. As discussed above<br />

the conversion Cys65 to Tyr confers an ethylene insensitive phenotype, while the defect in RAN1<br />

results in the constitutive activation of the ethylene response. One possible and plausible<br />

explanation for this discrepancy is that Cys65 is required not only for copper coordination but also<br />

for the proper conformation of ethylene receptor. Presumably, ethylene receptor is locked at active<br />

state without Cys65.<br />

Since both of the ran1-1 and ran1-2 mutations are missense mutations, these mutated genes<br />

presumably express mutated copper transporters. These mutations cause the conversion of<br />

important amino acid residue for copper transporting activity to another residue, indicating the<br />

mutated copper transporters have reduced activities. Actually a recombinant ran1-1 protein has a<br />

reduced copper transporting activity in the yeast cells (HIRAYAMA et al. 1999). If one copper ion is<br />

incorporated in a functional ethylene receptor as Bleecker’s group proposed, the ran1-1 or ran1-2<br />

plant would have just two types of ethylene receptors, receptor with copper or without copper.<br />

However, having these two types of ethylene receptors cannot explain the relaxed ligand specificity<br />

of the ran1-1 and ran1-2 mutants. Determination of the fine structure of ethylene recognition<br />

domain is necessary.<br />

Although all the ethylene receptors have a His-kinase like domain, the amino acid sequences are<br />

different among them. ETR1 and ERS1 have all the motifs that are required for His-kinase activity,<br />

namely H, N, G1, F and G2 motifs. By contrast, ETR2, EIN4 and ERS2 lack some or all of them,<br />

indicating these His-kinase like domains do not have His-kinase activity. Recently, Wang et al.<br />

reported that the physiological roles of His-kinase accompanied receptors, ETR1, ERS1, are different<br />

from those of ETR2, EIN4 and ERS2, and that their His-kinase activities are not required for the<br />

ethylene response (WANG et al. 2003). Furthermore, NTHK1, a tobacco ethylene receptor, was<br />

shown to have Ser/Thr kinase activity (XIE et al. 2003). Based on these reports, it is more unlikely<br />

that the ethylene-signaling pathway belongs to His-Asp phospho-relay signaling pathway.


ETHYLENE-SIGNALING PATHWAY<br />

45<br />

The fact that the ethylene-signaling pathway does not require the His-kinase activity of<br />

ethylene receptors raises another question why ETR1 and ERS1 possess His-kinase activity. It is<br />

possible that the His-kinase activities of those ethylene receptors are required for functions other<br />

than the ethylene-signaling pathway. Arabidopsis has several His-kinases although their<br />

physiological roles are not fully elucidated. A cytokinin receptor, CRE1, is one of the histidine<br />

kinase signal transducers whose physiological functions have been assigned. Recent studies have<br />

demonstrated that the cytokinin signal was transduced from CRE1 to nucleus by phospho-relay<br />

components (SAKAI et al. 2001). Some of those components have the ability to interact with ETR1<br />

(URAO et al. 2000). It might be possible that the His-kinase activity of ETR1 or ERS1 is involved<br />

in the crosstalk between the signal transduction pathways for ethylene and other stimuli or plant<br />

hormones such as cytokinin.<br />

Since ethylene molecules can pass freely through the lipid bilayer membrane, ethylene<br />

receptors do not need to localize to the plasma membrane. Schaller’s group showed that ethylene<br />

receptors localized to the membrane of endoplasmic reticulum (CHEN et al. 2002). By contrast,<br />

Xie et al. demonstrated that a tobacco ethylene receptor, NTHK1, fused to green fluorescent<br />

protein localized to the plasma membrane (XIE et al. 2003). Further analysis is required for the<br />

determination of the subcellular localization of ethylene receptors.<br />

Signaling components from receptor to nucleus<br />

How do the ethylene receptors transduce the signal downstream? The genetically identified<br />

component is CTR1, a Raf-type Ser/Thr kinase. Loss-of-function mutations of CTR1 cause<br />

recessive ethylene constitutive response phenotypes, suggesting CTR1 functions as a negative<br />

regulator like ethylene receptors. Thus ethylene receptors presumably activate CTR1. Two-hybrid<br />

analysis using budding yeast revealed that CTR1 interacted with ETR1 and ERS1 (CLARK et al.<br />

1998). Recently, one of the mutant alleles, ctr1-8, turned out to be defective in the association with<br />

ETR1, providing in vivo evidence for the requirement of physical interaction between them for the<br />

CTR1 function (HUANG et al. 2003). Furthermore, a recent study showed that ETR1 and CTR1 colocalized<br />

to ER membrane, supporting this idea (GAO et al. 2003).<br />

Raf-1 in mammal cells functions as a MAP-kinase-kinase-kinase. It has been postulated that<br />

CTR1 is one of the components of a MAP kinase cascade that is responsible for the ethylenesignaling<br />

pathway. Ouaked et al. reported that a sort of MAP-Kinase-kinase and MAP-kinase were<br />

activated upon ethylene treatment in a CTR1-dependent manner (OUAKED et al. 2003). The<br />

authors reported overexpression of this MAPKK induced the triple response phenotype in the<br />

absence of ethylene. These kinases might constitute a MAP kinase cascade with CTR1. However<br />

these kinases are activated by ethylene while CTR1 is inactivated. Such a regulation has not been<br />

reported in other MAP kinase systems so far.<br />

Although the ctr1 mutants exhibit constitutive ethylene phenotypes, treatment with ethylene


46<br />

Takashi HIRAYAMA and Tsutomu UGAJIN<br />

enhances ethylene phenotypes. This observation suggests the existence of other CTR1 like<br />

components. There are several CTR1 like genes on the Arabidopsis genome. Although the<br />

physiological functions of those are not known yet, these kinases might function in the ethylenesignaling<br />

pathway.<br />

Recently, Larsen and Chang reported another ethylene related mutant, eer1. This mutant<br />

shows enhanced ethylene response phenotypes (LARSEN and CHANG 2001). The ctr1 eer1 double<br />

mutant exhibits stronger constitutive ethylene response phenotypes. The EER1 gene has been<br />

cloned and shown to be identical to RCN1 that encodes an A subunit of protein phosphatase 2A<br />

(PP2A) (LARSEN and CANCEL 2003). RCN1 has been reported to be involved in the responses to<br />

auxin and abscisic acid in root and in guard cell, respectively. In mammal cells, PP2A positively<br />

regulates Raf-1 and shown to interact with Raf-1 kinase directly (ABRAHAM et al. 2000). Actually,<br />

not RCN1 but a C subunit of PP2A, PP2A-1C, can interact with the N-terminal domain of CTR1 in<br />

vitro. It might be possible that RCN1/EER1 and/or other PP2As are involved in the regulation of<br />

CTR1 activity.<br />

The ein2 mutants have a semi-dominant strong ethylene insensitive phenotype, suggesting its<br />

important function in the ethylene-signaling pathway. However, little is known about EIN2. The<br />

EIN2 gene encodes a novel membrane-spanning protein (ALONSO et al. 1999). EIN2 has twelve<br />

putative membrane-spanning domains in its N-terminal half. This region has a significant<br />

similarity to Nramp divalent cation transporters. However, there is no evidence for the transporter<br />

activity of EIN2. In addition, two residues that have been shown to be required for the transporting<br />

activity of yeast Smf1p are not conserved in EIN2, suggesting that EIN2 does not have such an iontransporting<br />

activity. Although EIN2 must localize to membrane structures because of these<br />

membrane-spanning domains, the subcellular localization of EIN2 has not been determined yet.<br />

The C-terminal half seems to be a cytoplasmic domain. The function of this region is also not<br />

clarified yet since this region does not have any known motifs. However, overexpression of the C-<br />

terminal half confers constitutive ethylene response phenotypes in the absence of ethylene,<br />

suggesting that this region has a pivotal role in the activation of the down stream signal transducer.<br />

The mutation sites of dozens of the ein2 mutants have been determined. All of them except one<br />

(ein2-9) are non-sense mutations or frame-sift mutations. Given that the C-terminal domain is<br />

required for the EIN2 function, these mutant EIN2 proteins cannot activate the downstream<br />

component(s) and result in the ethylene insensitive phenotype. ein2-9 is a missense mutation that<br />

causes an amino acid conversion His1143 to Pro in the C-terminal domain. Since this residue is<br />

not conserved in the rice EIN2 homologue, it is postulated that this His residue does not have a<br />

specific function but the conversion from this His residue to Pro might disturb the functional<br />

structure of the C-terminal region.<br />

The relationship between EIN2 and Nramp seems similar to that of yeast glucose transporters<br />

(HXTs) and glucose sensors (Snf3p and Rgt2p). Snf3p and Rgt2p have a structure similar to<br />

HXTs at their N-terminal regions and a unique C-terminal cytoplasmic domain. At the beginning,


ETHYLENE-SIGNALING PATHWAY<br />

47<br />

Snf3p and Rgt2p had been thought to be a kind of glucose transporter because of the similarity in<br />

their N-terminal domain. Detailed studies of those proteins have revealed that these proteins<br />

function as not glucose transporters but glucose sensors, and that the C-terminal domains of those<br />

proteins have the ability to function as a signal transducer. Overexpression of the C-terminal<br />

domain of those proteins in yeast cells activates the glucose response. Analogy to this relationship,<br />

it can be postulated that EIN2 might be a sensor for something, for example a divalent cation. So<br />

far, no missense mutation in the N-terminal region has been reported. To address this possibility,<br />

we generated dozens of the mutated ein2 genes that have an insertion of five-amino-acid in the N-<br />

terminal region. Preliminary results indicate that the N-terminal region of EIN2 is also required<br />

for the ethylene response (Ugajin and Hirayama, unpublished data). Detailed analysis of this<br />

region will reveal the EIN2 function in the ethylene response. The identification of the EIN2<br />

function is necessary for the understanding of the ethylene-signaling pathway.<br />

Nuclear events<br />

Ethylene treatment induces the expression of a lot of genes. The inducible expression of such<br />

genes evokes the ethylene response. The known downstream component of EIN2 is EIN3. Defect<br />

in the EIN3 gene confers a weak ethylene insensitive phenotype. The EIN3 gene encodes a novel<br />

protein. Although it does not have any known motifs, the amino acid composition of EIN3 implies<br />

its nuclear function (CHAO et al. 1997). Actually, the EIN3-GUS fusion protein expressed<br />

transiently in Arabidopsis cells localized to nucleus. It has been reported that ethylene inducible<br />

genes has a cis-element called GCC-box in their promoter regions. A screen for Arabidopsis<br />

proteins that bind GCC-box sequence identified EREBP, an AP2 transcriptional factor, functioning<br />

in the gene activation by ethylene (OHME-TAKAGI and SHINSHI 1995). EIN3 does not have an AP2-<br />

like structure and the ability to bind GCC-box sequence. The level of EIN3 mRNA is not affected<br />

by ethylene treatment. Therefore, Solano et al. thought that early ethylene inducible genes could<br />

be candidates for EIN3 target, and tried to find such genes. They found that a gene encoding an<br />

EREBP-like protein, ERF1, was induced very quickly by ethylene treatment in an EIN3-dependent<br />

manner (SOLANO et al. 1998). In addition, a recombinant EIN3 protein had the ability to bind the<br />

promoter region of ERF1. Based on these results, they concluded that EIN3 activates the ERF1<br />

gene and produced ERF1 in turn activates ethylene inducible genes through GCC-box.<br />

Overexpression of ERF1 in the ein3 mutant induced the ethylene constitutive phenotypes,<br />

confirming their idea. The mechanisms for EIN3 activation have not been clarified yet.<br />

Arabidopsis has several EIN3-like genes. Among them, at least EIL1 and EIL2 have similar<br />

function since overexpression of EIL1 or EIL2 suppresses the ein3 mutant phenotype. The eil1<br />

mutant exhibits a very weak ethylene insensitive phenotype. Interestingly, the ein3 eil1 double<br />

mutant shows a strong ethylene insensitive phenotype similar to ein2, suggesting that the functions<br />

of EIN3 and EIL1 in the ethylene-signaling pathway are largely overlapping (ALONSO et al. 2003).


48<br />

Takashi HIRAYAMA and Tsutomu UGAJIN<br />

This idea is consistent with the weak ethylene insensitive phenotype of the ein3 mutants.<br />

Future perspective<br />

As described above, the over all structure of the ethylene-signaling pathway, from the<br />

perception to the gene expression, has been elucidated through the studies using Arabidopsis<br />

(Figure 3). Orthologues for Arabidopsis ethylene-signaling components have been found in other<br />

plant systems, including tomato, maize and rice. Therefore the ethylene-signaling pathway<br />

described here presumably is common in large parts in plant kingdom.<br />

Figure 3 Schematic representation of a model for the ethylene-signaling pathway. Ethylene receptor<br />

binds a copper ion presumably during protein modification in the Golgi apparatus. Copper<br />

ion is delivered by RAN1 copper transporter. Ethylene receptor seems to localize to the ER<br />

membrane although this is still controversial. The subcellular localization of EIN2 is not<br />

elucidated yet. EIN2 might sense some signals although there is no evidence. EIN3 functions<br />

in the nuclear although it is not clear if EIN3 localizes to the nuclear in the absence of<br />

ethylene or not. In the absence of ethylene, ethylene receptor and CTR1 are active and<br />

presumably inhibit EIN2. When ethylene receptor recognizes ethylene molecule, it turns off<br />

and consequently allows EIN2 to activate EIN3 somehow. Activated EIN3 induces the<br />

expressions of the ERF1 gene and some ethylene responsive genes (not illustrated). Produced<br />

ERF1, a transcriptional factor, in turn activates ethylene responsive genes.


ETHYLENE-SIGNALING PATHWAY<br />

49<br />

In spite of amount of efforts made by many researchers world wide, there are still many<br />

unsolved issues. For example, how do ethylene receptors activate down stream signaling<br />

molecules?, what is the CTR1 substrate?, what is the EIN2 function?, how is EIN3 regulated?, how<br />

is EIN5 or EIN6 involved in the ethylene-signaling pathway?, etc. Additional genetic studies, such<br />

as the isolation of suppressor mutants for existing mutants, may be required to identify new<br />

components. Furthermore, describing the biochemical properties, cellular functions and threedimensional<br />

structures of known components is necessary for understanding the ethylene-signaling<br />

pathway.<br />

Reference<br />

1. ABRAHAM, D., PODAR, K., PACHER, M., KUBICEK, M., WELZEL, N., HEMMINGS, B.A., DILWORTH, S.M.,<br />

MISCHAK, H., KOLCH, W. and BACCARINI, M. (2000). Raf-1-associated protein phosphatase 2A as a<br />

positive regulator of kinase activation. J Biol Chem. 275: 22300-22304.<br />

2. ALONSO, J.M., HIRAYAMA, T., ROMAN, G., NOURIZADEH, S. and ECKER, J.R. (1999). EIN2, a<br />

bifunctional transducer of ethylene and stress responses in Arabidopsis. Science. 284: 2148-2152.<br />

3. ALONSO, J.M., STEPANOVA, A.N., SOLANO, R., WISMAN, E., FERRARI, S., AUSUBEL, F.M. and ECKER,<br />

J.R. (2003). Five components of the ethylene-response pathway identified in a screen for weak ethyleneinsensitive<br />

mutants in Arabidopsis. Proc Natl Acad Sci U S A. 100: 2992-2297.<br />

4. BLEECKER, A.B., ESTELLE, M.A., SOMERVILLE, C. and KENDE, H. (1988). Insensitivity to ethylene<br />

conferred by a dominant mutation in Arabidopsis thaliana. Science. 241: 1086-1089.<br />

5. CHAO, Q., ROTHENBERG, M., SOLAN, R., ROMAN, G., TERZAGHI, W. and ECKER, J.R. (1997). Activation<br />

of the ethylene gas response pathway in Arabidopsis by the nuclear protein ETHYLENE-INSENSITIVE3<br />

and related proteins. Cell. 89: 1133-1144.<br />

6. CHEN, Y.F., RANDLETT, M.D., FINDELL, J.L. and SCHALLER, G.E. (2002). Localization of the ethylene<br />

receptor ETR1 to the endoplasmic reticulum of Arabidopsis. Journal of Biological Chemistry. 277: 19861-<br />

19866.<br />

7. CLARK, K.L., LARSEN, P.B., WANG, X. and CHANG, C. (1998). Association of the Arabidopsis CTR1 Raflike<br />

kinase with the ETR1 and ERS ethylene receptors. Proc Natl Acad Sci U S A. 95: 5401-5406.<br />

8. GAO, Z., CHEN, Y.F., RANDLETT, M.D., ZHAO, X.C., FINDELL, J.L., KIEBER, J.J. and SCHALLER, G.E.<br />

(2003). Localization of the Raf-like Kinase CTR1 to the Endoplasmic Reticulum of Arabidopsis through<br />

Participation in Ethylene Receptor Signaling Complexes. J Biol Chem. 278: 34725-34732.<br />

9. GUZMAN, P. and ECKER, J.R. (1990). Exploiting the triple response of Arabidopsis to identify ethylenerelated<br />

mutants. Plant Cell. 2: 513-523.<br />

10. HIRAYAMA, T., KIEBER, J.J., HIRAYAMA, N., KOGAN, M., GUZMAN, P., NOURIZADEH, S., ALONSO, J.M.,<br />

DAILEY, W.P., DANCIS, A. and ECKER, J.R. (1999). RESPONSIVE-TO-ANTAGONIST1, a Menkes/Wilson<br />

disease-related copper transporter, is required for ethylene signaling in Arabidopsis. Cell. 97: 383-393.<br />

11. HUA, J. and MEYEROWITZ, E.M. (1998). Ethylene responses are negatively regulated by a receptor gene<br />

family in Arabidopsis thaliana. Cell. 94: 261-271.<br />

12. HUANG, Y., LI, H., HUTCHISON, C.E., LASKEY, J. and KIEBER, J.J. (2003). Biochemical and functional<br />

analysis of CTR1, a protein kinase that negatively regulates ethylene signaling in Arabidopsis. Plant<br />

Journal. 33: 221-233.<br />

13. LARSEN, P.B. and CANCEL, J.D. (2003). Enhanced ethylene responsiveness in the Arabidopsis eer1<br />

mutant results from a loss-of-function mutation in the protein phosphatase 2A A regulatory subunit,<br />

RCN1. Plant Journal. 34: 709-718.


50<br />

Takashi HIRAYAMA and Tsutomu UGAJIN<br />

14. LARSEN, P.B. and CHANG, C. (2001). The Arabidopsis eer1 mutant has enhanced ethylene responses in<br />

the hypocotyl and stem. Plant Physiology. 125: 1061-1073.<br />

15. OHME-TAKAGI, M. and SHINSHI, H. (1995). Ethylene-inducible DNA-binding proteins that interact with<br />

an ethylene-responsive element. Plant Cell. 7: 173-182.<br />

16. OUAKED, F., ROZHON, W., LECOURIEUX, D. and HIRT, H. (2003). A MAPK pathway mediates ethylene<br />

signaling. The EMBO Journal. 22: 1282-1288.<br />

17. RODRIGUEZ, F.I., ESCH, J.J., HALL, A.E., BINDER, B.M., SCHALLER, G.E. and BLEECKER, A.B. (1999).<br />

A copper cofactor for the ethylene receptor ETR1 from Arabidopsis. Science. 283: 996-998.<br />

18. SAKAI, H., HONMA, T., AOYAMA, T., SATO, S., KATO, T., TABATA, S. and OKA, A. (2001). ARR1, a<br />

transcription factor for genes immediately responsive to cytokinins. Science. 294: 1519-1521.<br />

19. SOLANO, R., STEPANOVA, A., CHAO, Q.M. and ECKER, J.R. (1998). Nuclear events in ethylene signaling:<br />

a transcriptional cascade mediated by ETHYLENE-INSENSITIVE3 and ETHYLENE-RESPONSE-<br />

FACTOR1. Genes & Development. 12: 3703-3714.<br />

20. URAO, T., MIYATA, S., YAMAGUCHI-SHINOZAKI, K. and SHINOZAKI, K. (2000). Possible His to Asp<br />

phosphorelay signaling in an Arabidopsis two-component system. FEBS Letters. 478: 227-232.<br />

21. WANG, W., HALL, A.E., O’MALLEY, R. and BLEECKER, A.B. (2003). Canonical histidine kinase activity<br />

of the transmitter domain of the ETR1 ethylene receptor from Arabidopsis is not required for signal<br />

transmission. Proc Natl Acad Sci USA. 100: 352-357.<br />

22. WOESTE, K.E. and KIEBER, J.J. (2000). A strong loss-of-function mutation in RAN1 results in constitutive<br />

activation of the ethylene response pathway as well as a rosette-lethal phenotype. Plant Cell. 12: 443-455.<br />

23. XIE, C., ZHANG, J.S., ZHOU, H.L., LI, J., ZHANG, Z.G., WANG, D.W. and CHEN, S.Y. (2003).<br />

Serine/threonine kinase activity in the putative histidine kinase-like ethylene receptor NTHK1 from<br />

tobacco. Plant Journal. 33: 385-393.<br />

24. ZHAO, X.-C., QU, X., MATHEWS, D.E. and SCHALLER, G.E. (2002). Effect of ethylene pathway mutations<br />

upon expression of the ethylene receptor ETR1 from Arabidopsis. Plant Physiology. 130: 1983-1991.


ETHYLENE-SIGNALING PATHWAY<br />

51<br />

<br />

<br />

1<br />

<br />

2<br />

<br />

230-0045 1-7-29<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

ETR1 <br />

<br />

Raf-1 <br />

CTR1 EIN2 EIN3 <br />

<br />

CTR1 <br />

EIN2 <br />

<br />

His-Asp <br />

CTR1 EIN2 <br />

<br />

<br />

<br />

CTR1 EIN2 EIN3 <br />

EIN5, EIN6<br />

<br />

<br />

ETR1 CTR1 <br />

EIN2 EIN3


52 Takashi HIRAYAMA and Tsutomu UGAJIN<br />

RAN1 <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

2 <br />

RAN1 <br />

GFP <br />

RAN1 <br />

<br />

<br />

<br />

<br />

<br />

<br />

etr1<br />

<br />

EIN3 EIN2


Gamma Field Symposia, No. 42, 2003 Institute of Radiation Breeding<br />

NIAS, Japan<br />

MECHANISM OF BRASSINOSTEROID SIGNALING<br />

53<br />

MECHANISM OF BRASSINOSTEROID SIGNALING<br />

Takeshi NAKANO, Shigeo YOSHIDA and Tadao ASAMI<br />

Plant Functions Lab., RIKEN<br />

2-1 Hirosawa, Wako 351-0198<br />

Introduction<br />

Steroid hormones, compounds with a signature tetracyclic structure, are synthesized<br />

downstream of the isoprenoid pathway in many organisms. In animals, steroid hormones regulate<br />

embryonic and adult cell development, and also act on neurons, heart tissue, sperm, oocytes and<br />

other organs and cells. In insects, the steroid hormone ecdysone also controls metamorphosis and<br />

reproduction. Brassinolide is the most bioactive form in all plant steroids and the first purification<br />

and determination of the structure was done in using of chemicals from bee-collected rape pollen.<br />

Brassinosteroids, plant steroid hormones that are widely distributed throughout the plant kingdom,<br />

have biological effects on many plant growth processes, such as stem and pollen tube elongation,<br />

leaf development, and xylem development. Brassinosteroids can also regulate chloroplast that is<br />

closely associated with the unique plant organelle. As these steroids can regulate each organism<br />

specific organ development, steroids are interesting in the aspect of molecular evolution (Clouse,<br />

2001, Fujioka and Yokota, 2003).<br />

Brassinosteroid biosynthesis<br />

In the past decade, Arabidopsis brassinosteroid biosynthetic mutants such as det2 (Li et al.,<br />

1996), dwf4 (Choe et al., 1998) and cpd (Szekeres et al., 1996) had been identified and<br />

characterized (Fig. 1). Before the screening of these mutants, the importance of brassinosteroid for<br />

plant growth had not been confirmed yet. Brassinosteroid feeding for plant could not cause so<br />

drastic change on plant phenotype and some visible effects might be thought as similar to other<br />

known phytohormones. But, these brassinosteroid-deficient mutants have a pleiotropic dwarf<br />

phenotype with very short stem, shorten and waving leaves, and dark greened leaves. Gibberellin<br />

deficient mutants were known to similar to the dwarf phenotype, but these brassinosteroid-deficient<br />

mutants did not showed the germination inhibition and the late flowering that were observed in<br />

gibberellin deficient mutants. Then, brassinosteroids could be identified as special and unique<br />

bioactivity in comparison with another known phytohormones, and had participated in actual


54<br />

Takeshi NAKANO, Shigeo YOSHIDA and Tadao ASAMI<br />

Fig. 1 Brassinosteroid biosynthesis pathway.<br />

Arabidopsis mutants such as det2, dwf4 and cpd clarified the detail<br />

brassinosteroid biosynthetic pathway. One possible target of Brz is the<br />

cytochrome P450 enzyme encoded by DWF4.<br />

phytohormone group.<br />

Brz, Brassinosteroid biosynthesis inhibitor, revealed brassinosteroid functions in plant growth<br />

The recently-synthesized compound brassinazole (Brz or Brz220) was the first specific<br />

inhibitor of brassinosteroid biosynthesis to be discovered. One possible target of Brz220 is the<br />

cytochrome P450 enzyme encoded by DWF4 (Asami et al., 2001) (Fig. 1). Brz220 treatment of


MECHANISM OF BRASSINOSTEROID SIGNALING<br />

55<br />

plants under dark conditions causes phenotypes similar to those displayed by brassinosteroiddeficient<br />

mutants, such as shortened hypocotyl, cotyledon opening, and de-etiolation. In the light,<br />

plants treated with the Brz displayed dwarf and highly greened leaves (Asami et al., 2000, Asami<br />

and Yoshida, 1999). In the stem tissues, development of the vascular bundle cells was inhibited by<br />

Brz. (Nagata et al., 2001)<br />

The Brz also revealed the regulation of chloroplast development by brassinosteroid. In the<br />

cotyledon of dark-germinated plant, the mRNA of chloroplast genes; i.e., cab and rbcS, did not<br />

expressed. But, in the cotyledon of dark-germinated plant with Brz, these chloroplast genes<br />

expression was detected (Asami et al., 2000, Nagata et al., 2000). After 2 hour light emission to<br />

these dark-germinated plant, the Brz-treated plant harbored very quick developed chloroplast that<br />

contained more numbered thylakoid membrane layer that Brz-untreated plant. These result<br />

revealed that plastid development was regulated brassinosteroid. (Fig. 2)<br />

Brassinosteroid receper BRI1<br />

The Arabidopsis bri1 mutant was also identified by its dwarf phenotype, but brassinosteroid<br />

treatment did not recover the dwarfism to wild-type phenotype and did not inhibit the elongation of<br />

the roots of this mutant (Clouse et al., 1996, Li and Chory, 1997) (Fig. 3). Study of bri1 revealed<br />

that BRI1 is a critical component in brassinosteroid signaling, and that mutation in its BRI1 gene<br />

Fig. 2 Brz effect to the plant.<br />

Brz treatment of plants under dark conditions causes photomorphogenesis such as shortened<br />

hypocotyl, cotyledon opening. In the light, plants treated with the Brz displayed dwarf and highly<br />

greened leaves. The Brz also revealed the regulation of chloroplast development by brassinosteroid.


56<br />

Takeshi NAKANO, Shigeo YOSHIDA and Tadao ASAMI<br />

Fig. 3 Brassinosteroid receptor mutant bri1.<br />

The Arabidopsis bri1 mutant was identified by its dwarf phenotype.<br />

BRI1 is a member of the leucine-rich repeat (LRR) receptor kinase family.<br />

causes a deficiency in brassinosteroids. BRI1 is a member of the leucine-rich repeat (LRR)<br />

receptor kinase family, and brassinolide binds strongly to a plasma membrane fraction purified<br />

using an anti-BRI1 antibody (Wang et al., 2001). In animal cells, steroid hormones are perceived<br />

through nuclear-localized steroid-binding proteins, but plants can perceive steroid hormones at the<br />

cell surface by the BRI1 component (Schumacher and Chory, 2000). How this signal is transduced<br />

to regulate plant nuclear gene expression is unknown.<br />

Application of Brz (brassinosteroid biosyhthesis inhibitor)<br />

for screening of brassinosteroid signaling mutants<br />

Many research on molecular biological mechanism for plant growth has been performed using<br />

genetic methods in Arabidopsis. As initial steps in the study of the molecular genetics of a plant<br />

hormone, screens are conducted to identify phytohormone-deficient and phytohormone-insensitive<br />

mutants. These trials can identify a number of genes, but these genes are likely not all of the<br />

players in the regulation of plant growth by the phytohormone (Kende and Zeevaart, 1997). The<br />

next step to consider is a screen to identify suppressor mutants that repress phytohormone<br />

deficiency symptoms, as these mutants may be permanently activated in phytohormone signaling.<br />

In the gibberellin research field, rga was identified as a suppressor mutant of the ga1-3 gibberellin<br />

biosynthesis mutant , and the mutated gene was found to belong to the VHIID family (Silverstone<br />

et al., 2001). In addition, the spy mutant was identified on the basis of its resistance to the


MECHANISM OF BRASSINOSTEROID SIGNALING<br />

57<br />

gibberellin biosynthesis inhibitor paclobutrazol, and the gene was found to encode a homolog of N-<br />

acetylglucosamine transferase (Jacobsen et al., 1996). Research in the brassinosteroid signaling<br />

field is now proceeding to the second strategy for Arabidopsis mutant screening, using<br />

brassinosteroid-deficient mutants.<br />

In order to analyze in detail the mechanisms of brassinosteroid biosynthesis and signal<br />

transduction, we performed a screen for mutants with altered responses to Brz220 treatment in<br />

darkness in the germination stage (Fig. 4). A screen of 140,000 Arabidopsis seeds that had been<br />

subjected to EMS and fast neutron mutagenesis revealed several mutants that had significantly<br />

longer hypocotyls than the wild type when grown in the dark and treated with Brz220. These<br />

plants were designated bil mutants (Brz-insensitive-long hypocotyl).<br />

bil1, Brz-insensitive-long hypocotyl1<br />

Initially, we identified a dominant mutant, bil1-1D, from the EMS-treated lines. When grown<br />

in medium containing 3 µM Brz, wild-type plants had quite short hypocotyls, but bil1 mutants had<br />

hypocotyls as long as those of wild-type plants grown on unsupplemented medium (Fig. 5). In<br />

parallel, bzr1-1D and bes1-1D were identified as Brz-resistant and bri1-suppressor mutants,<br />

respectively (Fig. 5). Gene sequencing revealed that the bzr1-1D gene is the same gene as bil1-1D,<br />

even containing the same mutation (Wang et al., 2002). These genes are 88% identical to BES1,<br />

Fig. 4 Strategy for new brassinosteroid signaling mutants by using of Brz.<br />

Hopeful bil (Brz-insensitive-long hypocotyl) mutant can overcome the dwarf<br />

and shorten hypocotyl that are caused by brassinosteroid deficiency with Brz.


58<br />

Takeshi NAKANO, Shigeo YOSHIDA and Tadao ASAMI<br />

Fig. 5 Screened bil (Brz-insensitive-long hypocotyl) mutants of Arabidopsis.<br />

bil1 mutants with 3 M Brz had hypocotyls as long as those of wild-type plants<br />

grown on unsupplemented medium.<br />

and the bes1 mutant has the same nucleotide substitution (Yin et al., 2002). The plant-specific<br />

gene family encompassing BZR1, BES1 and BIL1 encodes novel phosphoproteins containing a<br />

putative nuclear localization signal. The BIL1/BZR1:CFP fusion protein localizes mainly to the<br />

cytoplasm and also to the nucleus at low levels, but treatment with brassinosteroids results in a<br />

significant increase of BIL1/BZR1:CFP levels in the nucleus within thirty minutes. In contrast, a<br />

BZR1:CFP protein containing the mutation localizes continuously to the nucleus (Wang et al.,<br />

2002). Then, the mutants felt that brassinosteroid signaling is on in every time, and these can<br />

showed wild-type like phenotype in brassinosteroid deficient condition. These results suggest that<br />

BIL1/BZR1 and BES1 is a key component in brassinosteroid signaling from the cell surface to the<br />

nucleus (Fig. 6).<br />

The phenotypes of the bzr1/bil1 and bes1 Brz-resistance and bri1-suppressor mutants,<br />

respectively, are very strong, with the resistant mutant just like wild-type plants in appearance,<br />

even though it is severely deficient in brassinosteroids. These mutants, the mutant alleles of which<br />

are both dominants, resulted from the substitution of just one amino acid as compared to the wild<br />

type. Overexpression of the BZR1 or BES1 wild-type genes via the CaMV 35S promoter resulted<br />

in only weak resistance against brassinosteroid deficiency (Wang et al., 2002, Yin et al., 2002).<br />

These results suggest that bzr1/bil1 and bes1 mutants would be difficult to identify from activationtagged<br />

pools of plants in the background of a mutant deficient in brassinosteroids, such as det2 or<br />

dwf4. An alternative method is to induce point mutations by chemical treatment of brassinosteroiddeficient<br />

mutants. The most widely-used method to identify point mutations is genomic walking,<br />

performed using backcrosses with another ecotype. The screening is best performed on<br />

recombinant F2 plants, to allow identification of the brassinosteroid-deficiency mutation as a


MECHANISM OF BRASSINOSTEROID SIGNALING<br />

59<br />

Fig. 6 Mechanism of brassinosteroid signaling conducted by BIL1/BZR1/BES1 protein.<br />

Treatment with brassinosteroids results in movement of BIL1/BZR1/BES1 from cytosol to<br />

nucleus. BIL1/BZR1/BES1 protein containing the mutation localizes continuously to the nucleus.<br />

homozygous det2/det2 plant, and not as det2/DET2 or DET2/DET2 plants. This is because<br />

distinguishing between putative brassinosteroid-signaling mutants with a det2/det2 homozygous<br />

background and det2/DET2 or DET2/DET2 plants with no mutation is very difficult, since they<br />

would all have a phenotype of resistance against brassinosteroid deficiency. Identifying a<br />

suppressor mutant resulting from a point mutation might also be challenging. These predictions<br />

suggest that the combination of Brz and a simple point mutation in the wild-type Colombia<br />

ecotype can allow rapid identification of crucial signaling proteins. This role for Brz will be a<br />

great contribution to plant science.<br />

bil5, Brz-insensitive-long hypocotyl5<br />

We have identified a recessive mutant, bil5, from seeds that received fast neutron treatment.<br />

The length of the hypocotyls of this mutant on medium containing Brz is less than that of bil1-D,<br />

but at least twice that of the wild type (Fig. 5). Interestingly, adult bil5 plants have pale green, thin<br />

stems, thin leaves and a shortened stem length. The dwarf-like phenotype is distinct from the<br />

brassinosteroid-deficient dwarf phenotype, and the pale-green leaves of bil5 are in contrast to the<br />

dark-green leaves of brassinosteroid-deficient mutants. A preliminary analysis has shown that<br />

chloroplast gene expression is lower in bil5 than in the wild type. At least in darkness,<br />

brassinosteroid is a negative regulator of chloroplast development. BIL5 may be a key protein in<br />

the brassinosteroid regulation of chloroplasts. In addition, the bil5 shortened stem phenotype can<br />

be rescued by humidity of 85% or above and light conditions of less than 25 µE, and its stomata<br />

have lowered responses to ABA and tend to stay opened. This observation suggests that the


60<br />

Takeshi NAKANO, Shigeo YOSHIDA and Tadao ASAMI<br />

mutant can be rescued by less stressful conditions, which may relate to the cross-talk between<br />

brassinosteroids and ABA. The bil5 mutation maps to the lower arm of chromosome I. Isolation<br />

of the bil5 mutant gene, by sequencing and complementation, is in progress.<br />

To screen for more bil mutants, we are starting from an activation-tagged line, in collaboration<br />

with the groups of Dr. Shinozaki and Dr. Matsui. The genes identified by analysis of these mutants<br />

may also be members of the brassinosteroid signaling cascade.<br />

Additional new players for brassinosteroid signaling<br />

brs1 and bak1 were identified as bri1-5 dwarf suppressor mutants by activation tagging.<br />

BRS1 encodes a carboxypeptidase, and its role in BR signalling has not been defined (Li et al.,<br />

2001) (Fig. 7). BAK1, however, encodes a leucine-rich-repeat type receptor-like kinase that could<br />

interact directly with BRI1 (Li et al., 2002, Nam and Li, 2002) (Fig. 7). On the basis of their<br />

phenotypes, bak1-1D and brs1-1D mutants could potentially be Brz-insensitive mutants. Several<br />

Fig. 7 Current key players on brassinosteroid signaling.<br />

BIL1: Brz-insensitive-long hypocotyl, BRI1: brassinosteroid insentive1, BAK1:bri1<br />

associated kinase, BIN: brassinosteorid insenstive, BRS1: bri1 suppressor. Brz:<br />

Brassinosteroid biosynthesis inhibitor.


MECHANISM OF BRASSINOSTEROID SIGNALING<br />

61<br />

brassinosteroid-insensitive dwarf mutants, bin2 and bin3/bin5, have also been identified. BIN2<br />

encodes a cytosolic GSK-kinase (He et al., 2002), whereas BIN3 and BIN5 encode proteins of the<br />

topoisomerase family (Yin et al., 2002) (Fig. 7). These proteins could be related to brassinosteroid<br />

signaling. This idea could be investigated by examining the phenotypes of transformed plants in<br />

which expression of these genes has been modified, such as by overexpression, and monitoring the<br />

plants for a Brz-insensitive phenotype. Rop2 is a type of GTPase, and transformants in which this<br />

protein is constitutively active show hypersensitivity to BRs. (Li et al., 2001) (Fig. 7). As a<br />

dominant negative Rop2 transformant does not display BR insensitivity, the actual relationship of<br />

this gene with BRs is not yet clear. The det3 mutant, with a lesion in a gene encoding a vacuolelocalized<br />

ATPase, is less sensitive to BRs (Schumacher et al., 1999). Future studies using this<br />

mutant should help reveal the currently unknown role of the vacuole in BR signaling. These two<br />

genes have possible roles in brassinosteroid signal transduction, and transformed plants with altered<br />

expression of these genes also could be Brz-insensitive. The combined analysis of the above<br />

mutants, gene-modified plants, and Brz should give further insights into brassinosteroid signaling.<br />

In another approach toward the understanding of brassinosteroid signaling, Drs. Joanne Chory<br />

and Detlef Weigel have mapped quantitative trait loci (QTL) responsible for natural variations in<br />

hormone and light responses (Borevitz et al., 2002). They first collected 141 Arabidopsis thaliana<br />

accessions from the Northern hemisphere and analyzed the lengths of their hypocotyls in different<br />

hormone and light conditions. From these accessions, an Arabidopsis recombinant inbred line<br />

(RIL) resulting from a cross of the Cape Verde Islands (Cvi) and Landsberg erecta (Ler) accessions<br />

was chosen for detailed analysis with Brz treatment. The resulting QTL map predicted at least<br />

three strong loci that confer Brz insensitivity and long hypocotyls in darkness, and five weaker loci<br />

were also identified. As these strong Brz-insensitivity loci do not map near the already confirmed<br />

or potential Brz-insensitivity genes, a more detailed QTL analysis and more genetic screening for<br />

BR signaling mutants will be needed to clarify the mysterious mechanisms of plant growth<br />

regulation by brassinosteroids.<br />

Recently, gene chip methods have been used to predict genes induced by brassinosteroids<br />

(Mussig et al., 2002, Goda et al., 2002). However, it is difficult to determine which genes are<br />

actually involved in brassinosteroid signaling in plants under normal conditions, because these<br />

methods are based on artificial situations such as chemical stimulation. Reverse genetics<br />

approaches that study the phenotypes of transgenic plants in which brassinosteroid or Brz synthesis<br />

genes are overexpressed or suppressed should be more useful in determining which genes are truly<br />

involved in brassinosteroid signaling.<br />

Acknowledgements<br />

We are grateful to Drs. J. Chory, Y. Yin, and S. Mora-Garcia, at Plant Biology Laboratory,<br />

Salk Institute, CA, USA for their kind help and coopolation in this research. We also thank to Ms.


62<br />

Takeshi NAKANO, Shigeo YOSHIDA and Tadao ASAMI<br />

T. Sato, S. Kanda, A. Yamagami, R. Kiuchi, and Mrs. T. Matsuyama, K. Sekimata, T. Suzuki,<br />

T.Komatsu, at Plant Functions Laboratory, RIKEN, for their valuable technical assistance. This<br />

research was supported in part by Grants-in-Aid from the Ministry of Education, Culture, Sports,<br />

Science and Technology of Japan (No. 15031229 to T.N.), and a grant from Bioarchitect Research<br />

Project of RIKEN (to T.N.).<br />

References<br />

1. ASAMI, T., MIN, Y.K., NAGATA, N., YAMAGISHI, K., TAKATSUTO, S., FUJIOKA, S., MUROFUSHI, N.,<br />

YAMAGUCHI, I. and YOSHIDA, S. (2000) Characterization of brassinazole, a triazole-type brassinosteroid<br />

biosynthesis inhibitor. Plant Physiology 123: 93-99.<br />

2. ASAMI, T., MIZUTANI, M., FUJIOKA, S., GODA, H., MIN, Y.K., SHIMADA, Y., NAKANO, T., TAKATSUTO,<br />

S., MATSUYAMA, T., NAGATA, N., SAKATA, K. and YOSHIDA, S. (2001) Selective interaction of triazole<br />

derivatives with DWF4, a cytochrome P450 monooxygenase of the brassinosteroid biosynthetic pathway,<br />

correlates with brassinosteroid deficiency in Planta. Journal of Biological Chemistry 276: 25687-25691.<br />

3. ASAMI, T. and YOSHIDA, S. (1999) Brassinosteroid biosynthesis inhibitors. Trends in Plant Science 4:<br />

348-353.<br />

4. BOREVITZ, J.O., MALOOF, J.N., LUTES, J., DABI, T., REDFERN, J.L., TRAINER, G.T., WERNER, J.D.,<br />

ASAMI, T., BERRY, C.C., WEIGEL, D. and CHORY, J. (2002) Quantitative trait loci controlling light and<br />

hormone response in two accessions of Arabidopsis thaliana. Genetics 160: 683-696.<br />

5. CHOE, S.W., DILKES, B.P., FUJIOKA, S., TAKATSUTO, S., SAKURAI, A. and FELDMANN, K.A. (1998) The<br />

DWF4 gene of Arabidopsis encodes a cytochrome P450 that mediates multiple 22 alpha-hydroxylation<br />

steps in brassinosteroid biosynthesis. Plant Cell 10: 231-243.<br />

6. CLOUSE, S.D. (2001) Integration of light and brassinosteroid signals in etiolated seedling growth. Trends<br />

in Plant Science 6: 443-445.<br />

7. CLOUSE, S.D., LANGFORD, M. and MCMORRIS, T.C. (1996) A brassinosteroid-insensitive mutant in<br />

Arabidopsis thaliana exhibits multiple defects in growth and development. Plant Physiology 111: 671-678.<br />

8. FUJIOKA, S. and YOKOTA, T. (2003) Biosynthesis and metabolism of brassinosteroid. Annual Review of<br />

Plant Biology 54: 137-164.<br />

9. GODA, H., SHIMADA, Y., ASAMI, T., FUJIOKA, S. and YOSHIDA, S. (2002) Microarray analysis of<br />

brassinosteroid-regulated genes in Arabidopsis. Plant Physiology 130: 1319-1334.<br />

10. HE, J.X., GENDRON, J.M., YANG, Y.L., LI, J.M. and WANG, Z.Y. (2002) The GSK3-like kinase BIN2<br />

phosphorylates and destabilizes BZR1, a positive regulator of the brassinosteroid signaling pathway in<br />

Arabidopsis. Proceedings of the National Academy of Sciences of the United States of America 99:<br />

10185-10190.<br />

11. JACOBSEN, S.E., BINKOWSKI, K.A. and OLSZEWSKI, N.E. (1996) SPINDLY, a tetratricopeptide repeat<br />

protein invlovled in gibberellin signal transduction in Arabidopsis. Proceedings of the National Academy<br />

of Sciences of the United States of America 93: 9292-9296.<br />

12. KENDE, H. and ZEEVAART, J.A.D. (1997) The five “classical” plant hormones. Plant Cell 9: 1197-1210.<br />

13. LI, H., SHEN, J.J., ZHENG, Z.L., LIN, Y. and YANG, Z. (2001) The Rop GTPase switch controls multiple<br />

developmental processes in Arabidopsis. Plant Physiology 126: 670-684.<br />

14. LI, J., LEASE, K.A., TAX, F.E. and WALKER, J.C. (2001) BRS1, a serine carboxypeptidase, regulates BRI1<br />

signaling in Arabidopsis thaliana. Proceedings of the National Academy of Sciences of the United States<br />

of America 98: 5916-5921.<br />

15. LI, J., WEN, J.Q., LEASE, K.A., DOKE, J.T., TAX, F.E. and WALKER, J.C. (2002) BAK1, an Arabidopsis


MECHANISM OF BRASSINOSTEROID SIGNALING<br />

63<br />

LRR receptor-like protein kinase, interacts with BRI1 and modulates brassinosteroid signaling. Cell 110:<br />

213-222.<br />

16. LI, J.M. and CHORY, J. (1997) A putative leucine-rich repeat receptor kinase involved in brassinosteroid<br />

signal transduction. Cell 90: 929-938.<br />

17. LI, J.M., NAGPAL, P., VITART, V., MCMORRIS, T.C. and CHORY, J. (1996) A role for brassinosteroids in<br />

light-dependent development of Arabidopsis. Science 272: 398-401.<br />

18. MUSSIG, C., FISCHER, S. and ALTMANN, T. (2002) Brassinosteroid-regulated gene expression. Plant<br />

Physiology 129: 1241-1251.<br />

19. NAGATA, N., ASAMI, T. and YOSHIDA, S. (2001) Brassinazole, an inhibitor of brassinosteroid<br />

biosynthesis, inhibits development of secondary xylem in cress plants (Lepidium sativum). Plant and Cell<br />

Physiology 42: 1006-1011.<br />

20. NAGATA, N., MIN, Y.K., NAKANO, T., ASAMI, T. and YOSHIDA, S. (2000) Treatment of dark-grown<br />

Arabidopsis thaliana with a brassinosteroid-biosynthesis inhibitor, brassinazole, induces some<br />

characteristics of light-grown plants. Planta 211: 781-790.<br />

21. NAM, K.H. and LI, J.M. (2002) BRI1/BAK1, a receptor kinase pair mediating brassinosteroid signaling.<br />

Cell 110: 203-212.<br />

22. SCHUMACHER, K. and CHORY, J. (2000) Brassinosteroid signal transduction: still casting the actors.<br />

Current Opinion in Plant Biology 3: 79-84.<br />

23. SCHUMACHER, K., VAFEADOS, D., MCCARTHY, M., SZE, H., WILKINS, T. and CHORY, J. (1999) The<br />

Arabidopsis det3 mutant reveals a central role for the vacuolar H+-ATPase in plant growth and<br />

development. Genes & Development 13: 3259-3270.<br />

24. SILVERSTONE, A.L., JUNG, H.S., DILL, A., KAWAIDE, H., KAMIYA, Y. and SUN, T.P. (2001) Repressing a<br />

repressor: Gibberellin-induced rapid reduction of the RGA protein in Arabidopsis. Plant Cell 13: 1555-<br />

1565.<br />

25. SZEKERES, M., NEMETH, K., KONCZKALMAN, Z., MATHUR, J., KAUSCHMANN, A., ALTMANN, T., REDEI,<br />

G.P., NAGY, F., SCHELL, J. and KONCZ, C. (1996) Brassinosteroids rescue the deficiency of CYP90, a<br />

cytochrome P450, controlling cell elongation and de-etiolation in Arabidopsis. Cell 85: 171-182.<br />

26. WANG, Z.Y., NAKANO, T., GENDRON, J., HE, J.X., CHEN, M., VAFEADOS, D., YANG, Y.L., FUJIOKA, S.,<br />

YOSHIDA, S., ASAMI, T. and CHORY, J. (2002) Nuclear-localized BZR1 mediates brassinosteroid-induced<br />

growth and feedback suppression of brassinosteroid biosynthesis. Developmental Cell 2: 505-513.<br />

27. WANG, Z.Y., SETO, H., FUJIOKA, S., YOSHIDA, S. and CHORY, J. (2001) BRI1 is a critical component of<br />

a plasma-membrane receptor for plant steroids. Nature 410: 380-383.<br />

28. YIN, Y.H., CHEONG, H., FRIEDRICHSEN, D., ZHAO, Y.D., HU, J.P., MORA-GARCIA, S. and CHORY, J.<br />

(2002) A crucial role for the putative Arabidopsis topoisomerase VI in plant growth and development.<br />

Proceedings of the National Academy of Sciences of the United States of America 99: 10191-10196.<br />

29. YIN, Y.H., WANG, Z.Y., MORA-GARCIA, S., LI, J.M., YOSHIDA, S., ASAMI, T. and CHORY, J. (2002)<br />

BES1 accumulates in the nucleus in response to brassinosteroids to regulate gene expression and promote<br />

stem elongation. Cell 109: 181-191.


64 Takeshi NAKANO, Shigeo YOSHIDA and Tadao ASAMI<br />

<br />

<br />

<br />

351-0198 2-1<br />

<br />

<br />

<br />

1979 <br />

<br />

40 <br />

10 <br />

<br />

<br />

<br />

<br />

<br />

bri1 <br />

LRR<br />

<br />

BRI1 GFP <br />

GFP BRI1 <br />

<br />

BRI1 <br />

<br />

1999 Brz <br />

Brz<br />

Arabidopsis <br />

Brz <br />

<br />

Brz bil1Brz-insensitive-long hypocotyl1 Arabidopsis EMS<br />

bil1 <br />

BIL1CFP


MECHANISM OF BRASSINOSTEROID SIGNALING<br />

65<br />

BIl1/BZR1 mRNA<br />

30 CFP <br />

BIL1 <br />

GSK3/Shaggy kinase BIN2 <br />

BIL1/BZR1 <br />

BIL1/BZR <br />

BIN2 kinase


Gamma Field Symposia, No. 42, 2003 Institute of Radiation Breeding<br />

NIAS, Japan<br />

PEPTIDE PLANT HORMONE, PHYTOSULFOKINE<br />

67<br />

PEPTIDE PLANT HORMONE, PHYTOSULFOKINE<br />

Yoshikatsu MATSUBAYASHI<br />

Graduate School of Bio-Agricultural Sciences, Nagoya University<br />

Chikusa, Nagoya, 464-8601, Japan<br />

e-mail: matsu@agr.nagoya-u.ac.jp<br />

Key words: phytosulfokine, peptide hormone, receptor kinase, dedifferentiation, proliferation<br />

Abstract<br />

Plant cells retain features characteristic of totipotent stem cells. That is, they have the<br />

potential to dedifferentiate, re-differentiate, and give rise to all the organs of a new plant. However,<br />

relative rates of these cellular processes are strictly dependent on initial cell density, suggesting<br />

that cell-to-cell communication is necessary for these processes. Recent biochemical purification<br />

studies have demonstrated that phytosulfokine (PSK), a small sulfated peptide, acts as an<br />

extracellular ligand involved in the initial step of cellular dedifferentiation, proliferation and redifferentiation.<br />

Furthermore, a 120-kD leucine-rich repeat receptor kinase, specifically interacting<br />

with PSK, has been purified from plasma membranes using ligand-based affinity chromatography.<br />

Lines of evidence suggest that this ligand-receptor pair confers competence for dedifferentiation<br />

and re-differentiation to individual cells, rather than directly determining cellular fate. In this<br />

article, I review what is known about PSK signaling.<br />

Introduction<br />

A high proportion of plant cells, even at a fully differentiated stage, can dedifferentiate and<br />

proliferate in vitro as totipotent stem cells, forming a structure called a callus, after treatment with<br />

plant hormones such as auxin and cytokinin. (Skoog F & Miller CO 1967) Callus cells can<br />

differentiate into various organs and give rise to a new plant, indicating that plant cells from<br />

specific adult tissues are capable of differentiating into cells of all tissues. However, relative rates<br />

of cellular dedifferentiation and growth in vitro generally depend on initial cell density. Cellular<br />

dedifferentiation and growth progress efficiently under high cell density, but are significantly<br />

suppressed under low cell density. To promote cellular growth at low cell density, several<br />

researchers have used specialized culture techniques such as nurse cultures, in which target cells


68<br />

Yoshikatsu MATSUBAYASHI<br />

are grown close to but physically separated from nurse cells (Fig. 1). (Muir WH et al. 1954; Torrey<br />

JG 1957; Raveh et al. 1973) Interestingly, growth suppression of low-density cells is also negated<br />

by addition of conditioned medium in which cells have previously been grown. (Stuart R & Street<br />

HE 1969; Somers DA et al. 1985) These observations strongly indicate that plant cells secrete an<br />

autocrine-type factor responsible for cellular dedifferentiation and growth.<br />

Discovery of phytosulfokine<br />

We tested several cell types for use in a high-throughput bioassay system for purification of<br />

the conditioning factor. Among these, the best results were obtained with a primary culture system<br />

of mechanically dispersed mesophyll cells prepared from asparagus cladodes. Dedifferentiation<br />

and proliferation of dispersed asparagus cells was completely suppressed under low cell density,<br />

but was significantly promoted by addition of conditioned medium derived from asparagus cell culture.<br />

This bioassay system clearly responds to a small amount of crude conditioned medium, and provides<br />

reliable results within one week. (Matsubayashi Y & Sakagami Y. 1996) The activity of the conditioning<br />

factor was completely lost by the protease treatments, indicating that this factor is a peptide.<br />

Using this bioassay system, in 1996, we succeeded in isolating a growth factor from<br />

conditioned medium derived from asparagus suspension culture. (Matsubayashi Y & Sakagami Y.<br />

1996) A purification of approximately 10 7 -fold was achieved, with recovery of activity of about<br />

15%. This factor is a sulfated peptide composed of only 5 amino acids (Fig. 2), and, due to the<br />

presence of sulfate esters, it was named phytosulfokine (PSK). PSK induces cellular dedifferentiation<br />

and proliferation of dispersed plant cells at concentrations as low as 1.0 nM. Sulfated tyrosines are<br />

Fig. 1. Density effect in plant cell culture. Cellular dedifferentiation and<br />

growth are significantly suppressed under the low-density. This growth<br />

suppression is negated by the use of nurse cultures in which target cells<br />

are grown close to but physically separated from nurse cells. These<br />

phenomena strongly indicate that plant cells secrete an autocrine-type<br />

growth factor responsible for cellular dedifferentiation and growth.


PEPTIDE PLANT HORMONE, PHYTOSULFOKINE<br />

69<br />

Fig. 2. Chemical structure of the purified growth factor. This<br />

factor is a sulfated peptide composed of only five amino<br />

acids, and due to the presence of sulfate esters, it was<br />

named phytosulfokine (PSK).<br />

often found in secreted peptides in animals, (Huttner WB. 1982, Niehrs C et al. 1993) but, to date,<br />

PSK is the only example of post-translational sulfation of tyrosine in plants. PSK is present, with<br />

identical structure, in conditioned medium derived from cell lines of many plants, including<br />

dicotyledons and monocotyledons, indicating that it is widely distributed among higher plants.<br />

Possible function of PSK<br />

Initial cell density is a determining factor in in vitro cellular re-differentiation such as tracheary<br />

element (TE) formation, which occur at high frequency under high cell density but are significantly<br />

suppressed below threshold densities, suggesting that intercellular signaling is involved in<br />

initiation and/or subsequent progress in cellular re-differentiation. (Fukuda H, & Komamine A.<br />

1980) Interestingly, PSK triggers TE differentiation of Zinnia mesophyll cells at nanomolar<br />

concentrations.(Matsubayashi Y. et al. 1999) This phenomenon is not a secondary effect caused by<br />

increased cell density, because a high proportion of TEs differentiate directly from dispersed<br />

mesophyll cells without intervening cell division. It has also been demonstrated that PSK signaling<br />

is involved in somatic embryogenesis in a carrot system (Kobayashi T. et al. 1999, Hanai H. et al.<br />

2000), and in adventitious bud formation on callus of Antirrhinum majus. (Yang G. et al. 1999)<br />

However, neither cellular re-differentiation nor dedifferentiation can be induced by PSK alone;<br />

they require certain ratios and concentrations of auxin and cytokinin in addition to PSK. (Matsubayashi<br />

Y. et al. 1999) Whereas PSK triggers TE differentiation of Zinnia mesophyll cells without intervening<br />

cell division in a cytokinin-rich medium, it induces cellular dedifferentiation and proliferation in an<br />

auxin-rich medium. Neither dedifferentiation nor re-differentiation occurs without PSK.<br />

In this context, it is possible that PSK first confers competence to individual cell plants, and<br />

that auxin/cytokinin then determines cell fate (Fig. 3). PSK also promotes adventitious root<br />

formation on cucumber hypocotyls (Yamakawa S. et al. 1998); it is likely that this organ<br />

development is determined by the endogenous auxin/cytokinin level.<br />

Genes for PSK precursor proteins<br />

Five paralogous genes encoding ≈ 80-amino-acid precursors of PSK have been identified in


70<br />

Yoshikatsu MATSUBAYASHI<br />

Fig. 3. Possible mode of action of PSK. PSK confers a competence for<br />

dedifferentiation and/or re-differentiation on fully differentiated cells.<br />

In this step, no morphological change is observed. After the acquisition<br />

of competence, auxin, cytokinin, and other factors determine cell fate.<br />

Arabidopsis. (The Arabidopsis Genome Initiative. 2000, Yang H. et al. 2001) Each predicted<br />

protein has a probable secretion signal at the N-terminus and a single PSK sequence close to the C-<br />

terminus. In addition, there are dibasic amino acid residues immediately upstream from the PSK<br />

domain. (Fig. 4) Peptide-hormones and other biologically active peptides are generally synthesized<br />

as inactive higher-molecular-weight precursors that must undergo a variety of post-translational<br />

processing steps to yield the active peptides. (Harris RB. 1989)<br />

PSK mRNAs are detected not only in dedifferentiated callus cells but also in leaves and roots<br />

Fig. 4. Alignment of the deduced amino acid sequences of PSK precursor proteins in Arabidopsis. Red box<br />

indicate the five-amino-acid PSK domain. Predicted amino terminal signal sequences are underlined<br />

and putative processing site immediately upstream from PSK domain are boxed by yellow. Aspartic<br />

acid residues on the amino-terminal side of the first tyrosine of PSK domain, one of the most<br />

important determinant of the sulfation of PSK precursor, are boxed by blue. Identical amino acid<br />

residues are indicated by an asterisk, and similar amino acid residues are indicated by a colon.


PEPTIDE PLANT HORMONE, PHYTOSULFOKINE<br />

71<br />

of intact plants, indicating that PSK expression is not limited to regions in which cells actively<br />

divide and differentiate. (Yang H. et al. 2001) To study the function of PSK in plants, a large<br />

number of transformation experiments have been performed using Arabidopsis. However, loss-offunction<br />

techniques have not produced visible, directly informative phenotypes, suggesting<br />

functional redundancy between these 5 PSK genes in Arabidopsis. A definite picture will only<br />

emerge when combinations of all five knockouts are available. Overexpression of PSK slightly<br />

promotes callus formation in vitro in the presence of auxin/cytokinin, (Yang H. et al. 2001) but<br />

does not affect the growth of seedlings.<br />

Structure-activity relationships of PSK<br />

Derivatization of peptide hormones with biochemical tags such as photoactivatable groups has<br />

been used in characterization and purification of hormone receptors. (Hazum E. 1983) A key factor<br />

in the use of such functional groups is the ability to derivatize peptides without loss of binding<br />

activity or biological activity.<br />

To identify the active core of PSK, we synthesized several PSK analogs by solid phase peptide<br />

synthesis and direct sulfation of the peptide-resin using dimethylformamide-sulfurtrioxide<br />

complex. (Matsubayashi Y. et al. 1996) As shown in Fig. 5, N-terminal tetrapeptide and tripeptide<br />

of PSK retained 8% and 20% of the activity of the parent pentapeptide, respectively, but N-<br />

terminal dipeptide showed no activity. Deletion of the sulfate groups of Tyr 1 and Tyr 3 resulted in<br />

compounds with 0.6% and 4% of the activity of PSK, respectively, indicating that the sulfate group<br />

of Tyr 1 is more important than that of Tyr 3 for activity. In contrast, the N-terminal-truncated analog<br />

and an unsulfated analog exhibited no activity. Thus, the N-terminal tripeptide fragment Tyr(SO3H)-<br />

Ile-Tyr(SO3H) has been identified as the active core of PSK.<br />

A popular method for covalently linking functional groups to a peptide involves the use of<br />

activated esters of the functional groups, which react with primary amines to form amide bonds.<br />

However, modification of the N-terminal amino group of PSK by addition of Gly strongly<br />

decreases its biological activity. (Matsubayashi Y. et al. 1996) Thus, functional derivatization of<br />

PSK requires incorporation of an additional primary amino group at the C-terminal region, which<br />

is less involved in PSK activity than the N-terminal. To fulfill these requirements, several Alasubstituted<br />

PSK analogs were tested for activity, and the analog [Ala 5 ]PSK and [Lys 5 ]PSK was<br />

found to possess binding activity equal to that of PSK (Fig. 5). (Matsubayashi Y. et al. 1999)<br />

Interestingly, [Lys 5 ]PSK retained significant activity after derivatization of the side chain of Lys 5 by<br />

biotin, even when a very long spacer chain was inserted between the amino group of Lys 5 and<br />

the carboxyl group of biotin. This finding provided the breakthrough in a series of experiments<br />

conducted with the aim of visualization and purification of PSK receptors.


72<br />

Yoshikatsu MATSUBAYASHI<br />

Fig. 5. Structure-activity relationships of PSK. PSK analogs were prepared by solid phase<br />

peptide synthesis and direct sulfation of the peptide-resin. Relative activity of each<br />

analog was determined by the bioassay using asparagus mesophyll cells or the<br />

competitive receptor binding assay (asterisk). ND=not determined. Among these,<br />

[ 125 I]-[N -(4-azidosalicyl)Lys 5 ]PSK was used for photoaffinity labeling experiments,<br />

and [Lys 5 ]PSK-Sepharose was used for purification of the PSK receptor.<br />

PSK receptor<br />

Because of the presence of highly hydrophilic sulfate groups in PSK molecules, it is unlikely<br />

that they pass through plasma membranes and directly interact with target molecules inside cells.<br />

To determine whether a cell surface receptor for PSK exists, radiolabeled PSK was synthesized by<br />

coupling [ 35 S]sulfuric acid with the phenolic groups of tyrosine. (Matsubayashi Y. et al. 1997)<br />

Binding of [ 35 S]PSK was detected on the surface of suspension cultured rice cells and in the<br />

plasma-membrane-enriched fractions. The binding is reversible and saturable, and only PSK<br />

analogs that possess biological activity can effectively displace the radioligand.<br />

To further characterize the PSK receptor, [ 3 H]PSK, which has higher specific radioactivity,<br />

was synthesized by catalytic reduction of a PSK analog containing tetradehydroisoleucine.<br />

(Matsubayashi Y & Sakagami Y. 1999) Ligand saturation analysis using [ 3 H]PSK revealed the<br />

existence of a high-affinity binding site in microsomal fractions derived from rice, maize,


PEPTIDE PLANT HORMONE, PHYTOSULFOKINE<br />

73<br />

asparagus and carrot. Among these, the greatest abundance of binding sites was found in the carrot<br />

membranes: approximately 150 fmol per mg microsomal proteins, with an estimated dissociation<br />

constant (Kd) of 4.2 nM.<br />

This PSK receptor protein was visualized by photoaffinity labeling of carrot plasma<br />

membrane fractions using the photoactivable 125 I-labeled PSK analog [N -(4-azidosalicyl) Lys 5 ]PSK<br />

(Fig. 5), the binding activity of which are ≈ 10% of unmodified PSK.(Matsubayashi Y & Sakagami<br />

Y. 2000) SDS-PAGE analysis of the labeled proteins indicated that a 120-kD protein and a minor<br />

150-kD protein specifically interact with PSK. Both proteins contain approximately 10 kD of N-<br />

linked oligosaccharide chains that can be cleaved by treatment with peptide N-glycosidase F.<br />

We purified these PSK-binding proteins from microsomal fractions of carrot cells by Triton X-<br />

100 solubilization and specific ligand-based affinity chromatography using a [Lys 5 ]PSK-Sepharose<br />

column containing a long spacer chain between ligand and matrix. (Matsubayashi Y. et al. 2002)<br />

SDS-PAGE of the proteins in the fractions eluted by PSK showed specific recovery of a major 120-<br />

kD protein and a minor 150-kD protein. Both proteins were absent in the fractions eluted by an<br />

inactive PSK analogs; this indicates that binding of the proteins is specific.<br />

Several independent purifications were performed, yielding 50 µg of the major 120-kD<br />

protein with 96,000-fold purification from 4,800 mg of microsomal proteins (corresponds to 24 L<br />

of suspension-cultured cells), with an overall recovery rate of 40%. (Matsubayashi Y. et al. 2002)<br />

Based on the internal sequence of this protein, a 3.5-kb cDNA clone was isolated from the carrot<br />

cDNA library. The cDNA encoded a 1021-amino-acid protein with a deduced molecular mass of<br />

112 kD, with features found in several hormone receptors in plants and animals (Fig. 5). It<br />

contained an N-terminal hydrophobic signal sequence, extracellular leucine-rich repeats (LRRs), a<br />

transmembrane domain, and a cytoplasmic kinase domain. Northern blot analysis showed that<br />

mRNA of this protein accumulated ubiquitously in leaf, apical meristem, hypocotyl and root of<br />

carrot seedlings, although its expression level was far lower than in cultured carrot cells. The major<br />

extracellular domain of this protein contained 21 tandem copies of a 24-amino-acid LRR; it has<br />

been suggested that this string of LRRs plays a key role in protein-protein interactions. (Kobe B &<br />

Deisenhofer J. 1994) In addition, a 36-amino-acid island was detected in the 18 th LRR. An island<br />

domain has also been found among the extracellular LRRs of the brassinosteroid receptor BRI1,<br />

and has been shown to be critical for its function. (Li J & Chory J. 1997)<br />

Transgenic carrot cells overexpressing the cDNA of the major 120-kD protein showed a<br />

significant increase in PSK binding sites in the membrane fractions and accelerated growth in<br />

response to PSK, compared with control cells. Photoaffinity cross-linking and immunoprecipitation<br />

analysis of the membrane proteins derived from the transformants revealed expression of the 150-<br />

kD protein in addition to the 120-kD protein, indicating that both proteins are encoded by a single<br />

gene. In contrast, expression of antisense mRNA of the major 120-kD protein significantly<br />

inhibited callus growth. These findings strongly suggest that this receptor kinase is a component of<br />

a functional PSK receptor that directly interacts with PSK.


74<br />

Yoshikatsu MATSUBAYASHI<br />

Fig. 6. Schematic of the 120-kD PSK receptor. The diagram shows the<br />

signal peptide (SP, red), extracellular leucine-rich repeats (LRRs,<br />

yellow), a 36-amino-acid island, a transmembrane domain (TM,<br />

blue), and a cytoplasmic kinase domain (green).<br />

Future perspectives<br />

Now that in vitro function of PSK and the molecular basis of ligand-receptor interaction in<br />

PSK signaling have been established, the next phase of research is characterization of the in vivo<br />

role of PSK and its downstream signaling pathway in plants. The carrot PSK receptor exhibits a<br />

high percentage of amino acid identity with several LRR receptor-like kinases found in<br />

Arabidopsis. The sequencing of the Arabidopsis genome is now complete, and large collections of<br />

gene-disruption lines are available. Once PSK-binding activities of these LRR-RLKs are<br />

confirmed, direct clues to in vivo function of PSK will be provided by phenotypes of knockout<br />

mutants.<br />

References<br />

1. Beisswanger R, Corbeil D, Vannier C, Thiele C, Dohrmann U, Kellner R, Ashman K, Niehrs C, Huttner<br />

WB. (1998) Existence of distinct tyrosylprotein sulfotransferase genes: molecular characterization of<br />

tyrosylprotein sulfotransferase-2. Proc Natl Acad Sci USA 95: 11134-11139.<br />

2. Bundgaard JR, Vuust J, Rehfeld JF. (1997) New consensus features for tyrosine O-sulfation determined<br />

by mutational analysis. J Biol Chem 272: 21700-21705.<br />

3. Fukuda H, Komamine A. (1980) Establishment of an experimental system for the tracheary element<br />

differentiation from single cells isolated from the mesophyll of Zinnia elegans. Plant Physiol 65: 57-60.<br />

4. Hanai H, Matsuno T, Yamamoto M, Matsubayashi Y, Kobayashi T, Kamada H, Sakagami Y. (2000) A<br />

secreted peptide growth factor, phytosulfokine, acting as a stimulatory factor of carrot somatic embryo<br />

formation. Plant Cell Physiol 41: 27-32.<br />

5. Hanai H, Nakayama D, Yang H, Matsubayashi Y, Hirota Y, Sakagami Y. (2000) Existence of a plant<br />

tyrosylprotein sulfotransferase: novel plant enzyme catalyzing tyrosine O-sulfation of preprophytosulfokine<br />

variants in vitro. FEBS Lett 470: 97-101.<br />

6. Harris RB. (1989) Processing of pro-hormone precursor proteins. Arch Biochem Biophys 275: 315-333.<br />

7. Hazum E. (1983) Photoaffinity labeling of peptide hormone receptors. Endocr Rev 4: 352-362.<br />

8. Huttner WB. (1982) Sulphation of tyrosine residues, a widespread modification of proteins. Nature 299:<br />

273-276.<br />

9. Huttner WB. (1988) Tyrosine sulfation and the secretory pathway. Annu Rev Physiol 50: 363-376.<br />

10. Kobayashi T, Eun C-H, Hanai H, Matsubayashi Y, Sakagami Y, Kamada H. (1999) Phytosulfokine-, a<br />

peptidyl plant growth factor, stimulates somatic embryogenesis in carrot. J Exp Botany 50:1123-1128.<br />

11. Kobe B, Deisenhofer J. (1994) The leucine-rich repeat: a versatile binding motif. Trends Biochem Sci 19:<br />

415-421.


PEPTIDE PLANT HORMONE, PHYTOSULFOKINE<br />

75<br />

12. Li J, Chory J. (1997) A putative leucine-rich repeat receptor kinase involved in brassinosteroid signal<br />

transduction. Cell 90: 929-938.<br />

13. Matsubayashi Y, Hanai H, Hara O, Sakagami Y. (1996) Active fragments and analogs of the plant growth<br />

factor, phytosulfokine: structure-activity relationships. Biochem Biophys Res Commun 225: 209-214.<br />

14. Matsubayashi Y, Sakagami Y. (1996) Phytosulfokine, sulfated peptides that induce the proliferation of<br />

single mesophyll cells of Asparagus officinalis L. Proc Natl Acad Sci USA 93: 7623-7627.<br />

15. Matsubayashi Y, Takagi L, Sakagami Y. (1997) Phytosulfokine-, a sulfated pentapeptide, stimulates the<br />

proliferation of rice cells by means of specific high- and low-affinity binding sites. Proc Natl Acad Sci<br />

USA 94: 13357-13362.<br />

16. Matsubayashi Y, Takagi L, Omura N, Morita A, Sakagami Y. (1999) The endogenous sulfated<br />

pentapeptide phytosulfokine- stimulates tracheary element differentiation of isolated mesophyll cells of<br />

Zinnia. Plant Physiol 120: 1043-1048.<br />

17. Matsubayashi Y, Morita A, Matsunaga E, Furuya A, Hanai N, Sakagami Y. (1999) Physiological<br />

relationships between auxin, cytokinin, and a peptide growth factor, phytosulfokine-, in stimulation of<br />

asparagus cell proliferation. Planta 207: 559-565.<br />

18. Matsubayashi Y, Takahata Y, Morita A, Atsumi K, Sakagami Y. (1999) Preparation and characterization<br />

of fully active biotinylated analogs of phytosulfokine-. Biosci Biotechnol Biochem 63: 1847-1849.<br />

19. Matsubayashi Y, Sakagami Y. (1999) Characterization of specific binding sites for a mitogenic sulfated<br />

peptide, phytosulfokine-, in the plasma-membrane fraction derived from Oryza sativa L. Eur J Biochem<br />

262: 666-671.<br />

20. Matsubayashi Y, Sakagami Y. (2000) 120- and 160-kDa receptors for endogenous mitogenic peptide,<br />

phytosulfokine-, in rice plasma membranes. J Biol Chem 275: 15520-15525.<br />

21. Matsubayashi Y, Ogawa M, Morita A, Sakagami Y. (2002) An LRR receptor kinase involved in<br />

perception of a peptide plant hormone, phytosulfokine. Science 296: 1470-1472.<br />

22. Muir WH, Hildebrandt AC, Riker AJ. (1954) Plant tissue cultures produced from single isolated cells.<br />

Science 119: 877-878.<br />

23. Niehrs C, Beisswanger R, Huttner WB. (1993) Protein tyrosine sulfation, an update. (1994) Chem Biol<br />

Interact 92: 257-271.<br />

24. Ouyang Y, Lane WS, Moore KL. (1998) Tyrosylprotein sulfotransferase: purification and molecular<br />

cloning of an enzyme that catalyzes tyrosine O-sulfation, a common posttranslational modification of<br />

eukaryotic proteins. Proc Natl Acad Sci USA 95: 2896-2901.<br />

25. Raveh D, Hubermann E, Galun E. (1973) In vitro culture of tobacco protoplasts: use of feeder techniques<br />

to support division of cells plated at low densities. In Vitro 9: 216-222.<br />

26. Skoog F, Miller CO. (1967) Chemical regulation of growth and organ formation in plant tissues cultured<br />

in vitro. Symp Soc Exp Biol. 11: 118-140.<br />

27. Somers DA, Birnberg PR, Petersen WL, Brenner ML. (1985) The effect of conditioned medium on<br />

colony formation from Black Mexican sweet corn protoplasts. Plant Cell Rep 4: 155-157.<br />

28. Stuart R, Street HE. (1969) Studies on the growth in culture of plant cells. IV. The initiation of division in<br />

suspensions of stationary phase cells of Acer pseudoplatanus L. J Exp Bot 20: 556-571.<br />

29. The Arabidopsis Genome Initiative. (2000) Analysis of the genome sequence of the flowering plant<br />

Arabidopsis thaliana. Nature ;408: 796-815.<br />

30. Torrey JG. (1957) Cell division in single isolated cells in vitro. Proc Natl Acad Sci USA 43: 887-891.<br />

31. Yamakawa S, Sakuta C, Matsubayashi Y, Sakagami Y, Kamada H, Satoh S. (1998) The promotive effects<br />

of a peptidyl plant growth factor, phytosulfokine-, on the formation of adventitious roots and expression<br />

of a gene for a root-specific cystatin in cucumber hypocotyls. J Plant Res 111: 453-458.<br />

32. Yang G, Shen S, Kobayashi T, Matsubayashi Y, Sakagami Y, Kamada H. (1999) Stimulatory effects of a<br />

novel peptidyl plant growth factor, phytosulfokine-, on the adventitious bud formation from callus of<br />

Antirrhinum majus. Plant Biotechnol 16: 231-234.


76<br />

Yoshikatsu MATSUBAYASHI<br />

33. Yang H, Matsubayashi Y, Hanai H, Nakamura K, Sakagami Y. (2000) Molecular cloning and<br />

characterization of OsPSK, a gene encoding a precursor for phytosulfokine-, required for rice cell proliferation.<br />

Plant Mol Biol 635: 635-647.<br />

34. Yang H, Matsubayashi Y, Nakamura K, Sakagami Y. (2001) Diversity of Arabidopsis genes encoding<br />

precursors for phytosulfokine, a peptide growth factor. Plant Physiol 127: 842-851.


PEPTIDE PLANT HORMONE, PHYTOSULFOKINE<br />

77<br />

<br />

<br />

<br />

464-8601<br />

<br />

<br />

<br />

in vitro <br />

<br />

<br />

<br />

<br />

<br />

PSK<br />

PSK <br />

<br />

<br />

PSK 120 kD LRR<br />

PSK <br />

PSK <br />

<br />

<br />

PSK3 APPSK3


78 Yoshikatsu MATSUBAYASHI<br />

GUS <br />

<br />

<br />

<br />

<br />

<br />

<br />

GUS PSK <br />

<br />

<br />

<br />

BRI BRI GUS


79


80<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

BRI1 <br />

<br />

<br />

<br />

<br />

1 100 <br />

5 5 <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

10


81<br />

RHT3 RHT <br />

10 <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

10 <br />

slender rice <br />

<br />

<br />

<br />

slender<br />

rice RHT sd1<br />

6 <br />

2 <br />

sd1 gain-of-function <br />

sd1 <br />

10 <br />

<br />

<br />

<br />

sd1 <br />

<br />

BRI1


82<br />

<br />

<br />

<br />

BRI 10


83<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

SD1 C23 <br />

<br />

D35 <br />

KO <br />

2 <br />

SD1 <br />

slender <br />

<br />

XA1 XA21


84<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

100 <br />

<br />

1970 <br />

ent- <br />

AMO1618 ent- <br />

<br />

P450 <br />

P450


85<br />

2- <br />

3 <br />

<br />

<br />

ABA <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

PSK <br />

<br />

<br />

PSK


86<br />

<br />

PSK <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

23 <br />

<br />

<br />

<br />

BIL1/CFP <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

1 <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

2


87<br />

<br />

<br />

2 <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

2


87<br />

<br />

<br />

2 <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

2

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!