Open Access
How to translate text using browser tools
4 April 2002 Phytochrome Signaling Mechanism
Haiyang Wang, Xing Wang Deng
Author Affiliations +

* This chapter was updated on July 6, 2004. The original version has been archived and is available at  http://dx.doi.org/10.1199/tab.0074

INTRODUCTION

Because of their sessile nature, plants have adapted a high degree of developmental plasticity to optimize their growth and reproduction in response to their ambient environments, such as light, temperature, humidity, and salinity. Plants utilize a wide range of sensory systems to perceive and transduce specific incoming environmental signals. Light is one of the major environmental signals that influences plant growth and development. In addition to being the primary energy source for plants, light also provides them with positional information to modulate many developmental processes, including seed germination, seedling de-etiolation, leaf expansion, stem elongation, phototropism, stomata and chloroplast movement, shade avoidance, circadian rhythms, and flowering time (Deng and Quail, 1999; Wang and Deng, 2003).

The phenotypic changes associated with the seedling photomorphogenic development are among the most dramatic events mediated by light. Dark-grown seedlings undergo skotomorphogenesis (etiolation) and are characterized by long hypocotyls, closed cotyledons and apical hooks, and development of the proplastids into etioplasts. Light-grown seedlings undergo photomorphogenesis (de-etiolation) and are characterized by short hypocotyls, open and expanded cotyledons, and development of the pro-plastids into green mature chloroplasts (McNellis and Deng, 1995, Figure 1).

Plants can detect almost all facets of light such as direction, duration, quantity, and wavelength by using three major classes of photoreceptors: The cryptochromes and phototropins monitor the blue/ultraviolet-A (B/UV-A) (320–500 nm) region of the spectrum, whereas the phytochromes monitor primarily the red (R) and far-red (FR) wavelengths (600–750 nm) (Kendrick and Kronenberg, 1994; Briggs and Olney, 2001). These photoreceptors perceive, interpret, and transduce light signals, via distinct intracellular signaling pathways, to modulate photoresponsive nuclear gene expression, and ultimately leading to adaptive changes at the cell and whole organism levels.

This chapter aims to highlight some of the most recent progress in elucidating the molecular, cellular and biochemical mechanisms of phytochrome signaling in Arabidopsis. The interested readers are referred to the accompanying reviews on other related subjects, such as photomorphogenesis (reviewed by Jennifer Nemhauser and Joanne Chory), circadian rhythms (reviewed by C Robertson McClung, Patrice A. Salomé, and Todd P. Michael), phototropism (reviewed by Emmanuel Liscum), and chapters of phytohormones (such as brassinosteroids and ethylene).

THE DISCOVERY AND ACTION MODES OF PHYTOCHROMES

Phytochrome, stand for “plant color”, was originally coined to describe the proteinous pigment that controls photoperiod detection and floral induction of certain short-day plants (such as cocklebur and soybean, Garner and Allard, 1920) and the reversible seed germination of lettuce (c.v. Grand rapids) by red (R) and far-red (FR) light (Borthwick et al., 1952). Red light promotes seed germination, whereas subsequent far-red light treatment abolishes red light induction of seed germination. The germination response of lettuce seeds repeatedly treated with R/FR cycles is determined by the last light treatment. Thus R/FR reversibility is a characteristic feature of this response. In addition, the law of reciprocity applies to this response, i.e., the response is dependent on the total amount of photons received irrespective of the duration of light treatment. This type of responses is now defined as the low fluence responses (LFRs, Mancinelli, 1994). LFRs also induce other transient responses, such as changes in ion flux, leaf movement, chloroplast rotation and changes in gene expression (Haupt and Hader, 1994; Roux, 1994; Vince-Prue, 1994). Besides the R/FR reversible LFRs, there are two other modes of phytochrome action. The very-low-fluence responses (VLFRs) are activated by extremely low light intensities of different wavelengths (far-red, red and blue), such as light-induced expression of the LHCB gene and light induction of seed germination, and the high-irradiance responses (HIRs) which depend on prolonged exposure to relatively high light intensities. HIRs are primarily responsible for the control of seedling de-etiolation (e.g. inhibition of hypocotyl elongation and promotion of cotyledon expansion) under all light qualities (Mustilli and Bowler, 1997; Casal et al., 1998; Table 1).

THE PHYTOCHROME GENE FAMILY AND THE CHROMOPHORE

Two Reversible Forms of Phytochromes

The enrichment of phytochrome in dark-grown seedlings and its predicted photoreversibility facilitated its initial purification (Butler et al., 1959, 1964). The purified phytochrome from dark-grown plants is blue in color and absorbs red light (termed Pr, absorbs maximally at 660 nm). On exposure to red light, the Pr form converts to the far-red light absorption form (Pfr) which is olive-green in color and absorbs maximally at 730 nm (Quail, 1997, Figure 2). The Pfr form is considered as the biologically active form. The Pfr form converts back to the Pr form on absorption of far-red light or undergoes slow dark reversion in the absence of light. It should be noted that in addition to their maximal absorptions of red and far-red wavelengths, phytochromes also weakly absorb blue light (Furuya and Song, 1994; Figure 2).

Classification of Phytochromes and The Gene Family

Purified phytochrome from etiolated seedlings exists as a soluble homodimer, with each apoprotein monomer attaching a covalently linked chromophore. The molecular mass of the phytochrome apoprotein is approximately 125 kDa. The chromophore is attached via a thioether linkage to an invariant cysteine in a well-conserved domain among all phytochromes. Previous spectrophotometric studies indicated that there are at least two distinct pools of phytochromes, Type I (light labile) and type II (light stable). The light-labile pool degrades fairly rapidly upon exposure to red or white light, whereas Type II phytochromes are stable in these light conditions. In Arabidopsis, there are five phytochromes (termed phyA-phyE) encoded by five distinct members of the phytochrome gene family (Sharrock and Quail, 1989). phyA is a type I phytochrome, and phyB-phyE belong to type II phytochromes. Sequence analysis suggests that these phytochromes can be clustered into three subfamilies: phyA/phyC, phyB/phyD, and phyE (Figure 3A; Table 2). Counterparts of Arabidopsis PHY genes are present in most, if not all, higher plants (Clack et al., 1994; Sharrock and Quail, 1989; Mathews and Sharrock, 1997).

All five Arabidopsis phytochromes are expressed throughout the plant. Although the differences in their expression patterns are minor, their abundance and stability differ dramatically (Somers and Quail 1995; Goosey et al., 1997). phyA is most abundant in dark-grown seedlings and its level drop up to 100-fold after exposure to light, as a result of combinatorial regulation at both the transcriptional and post-transcriptional levels. Light down-regulates PHYA gene transcription (Somers and Quail, 1995; Canton and Quail, 1999) and causes degradation of phyA polypeptide (Clough et al., 1999; Hennig et al., 1999). In light-grown plants, phyB is the most abundant phytochrome, whereas phyC-phyE are the less abundant type II phytochromes (Clark et al., 1994; Hirschfeld et al., 1998).

General Structure of Phytochromes

Studies with biochemically purified phyA holoproteins indicated that the phytochrome molecule consists of two structural domains connected via a flexible hinge region. The N-terminal domain is responsible for chromophore-binding, light absorption and photoreversibility (∼ 70 kDa). The C-terminal domain (∼ 55 kDa) contains several conserved subdomains/motifs including the regulatory core sequence (Quail box), two dimerization motifs, and the histidine kinase-related domain (HKRD). A pair of Per-Arnt-Sim (PAS) motif overlaps with the Quail box (Figure 3B). PAS domains can be used either as platforms for protein-protein interaction or as response modules to small ligands or changes in light conditions, oxygen levels, and redox potentials (Quail, 1997; Neff et al., 2000). A number of point mutations in the C-terminal domain of both phyA and phyB do not affect photoreversibility but eliminate the biological activity (Quail et al., 1995; Quail, 1997), suggesting that the C-terminal domain is essential for proper downstream signaling. A domain swapping and deletion analysis suggested that the N terminus of phytochrome is essential for its specific photosensory properties and that the C termini of phyA and phyB are the output domains and interchangeable (Wagner et al., 1996). However, this notion was recently confronted by the finding that the N-terminal domain of phyB, when dimerized and localized in the nucleus, confers much higher photosensitivity than the full-length phyB. These results suggest that the N-terminal domain of phyB transduces the light signal to downstream targets, and the C-terminal domain of phyB attenuates the activity of phyB (Matsushita et al., 2003). At present, it is not known whether other phytochrome species share the same structure-function relationship as phyB.

Chromophore Biosynthesis

As mentioned above, functional phytochrome holoproteins require the covalent attachment of a chromophore to each phytochrome apoprotein monomer. Using degradation approaches and proton NMR spectroscopy, the structure of the phytochrome chromophore was determined to be a linear tetrapyrrole, phytochromobilin (PΦB), which ligates to a cysteine residue located in the N-terminal half apoprotein of PHYA through its A-ring (Lagarias and Rapoport, 1980).

The synthesis of PΦB is accomplished by a series of enzymatic reactions in the plastid that begins with 5-aminolevulinic acid. The early steps in the PΦB pathway are shared with chlorophyll and heme biosynthesis. The committed step is the oxidative cleavage of heme by a ferredoxin-dependent heme oxygenase (HO) to form billiverdin IX (BV). BV is then converted into 3Z-PΦB by the ferredoxin-dependent bilin reductase PFB synthase. Both 3Z-PΦB and its isomerized form 3E-PΦB can serve as functional precursors of the phytochrome chromophore. PΦB is then exported to the cytoplasm where it binds to the newly synthesized apo-PHYs (Terry et al., 1997, Figure 4 A and B). Absorption of red light triggers a “Z” to ‘E” isomerization in the C-15 double bond between the C and D rings of the linear tetrapyrrole, resulting in the far-red light absorbing form Pfr. This Pr-to-Pfr transition is associated with rearrangement of the protein backbone. Pfr can be converted to Pr either by a slow non-photoinduced reaction (dark reversion) or much faster upon absorption of farred light (Quail, 1997; Fankhauser, 2001; Figure 4C).

Arabidopsis mutants defective in the PΦB-synthetic pathway have been isolated. These mutants (hy1 and hy2) have dramatically reduced levels of PΦB and consequently functional phytochromes, and thus exhibit severely impaired photomorphogenesis (Parks and Quail, 1991). The Arabidopsis HY1 locus encodes a HO (designated AtHO1) responsible for much of PΦB synthesis in Arabidopsis (Davis et al., 1999; Muramoto et al., 1999; Table 2). Three additional HO genes (AtHO2-4) exist in the Arabidopsis genome, and these additional HOs may provide alternative pathways for making BV (Davis et al., 2001). The Arabidopsis HY2 locus, likely a unique gene in the Arabidopsis genome, encodes the phytochromobilin synthase (Kohchi et al., 2001; Table 2).

It is generally assumed that all phys have the same chromophore. The finding that phyB purified from transgenic Arabidopsis shows spectral properties similar to phyA supported this notion (Elich and Chory, 1997). It should be pointed out that besides PΦB, phycocyanobilin (PCB), the chromophore of the light-harvesting pigment phycocyanin, also can bind phytochrome resulting in Pr and Pfr spectra that are slightly blue shifted compared with the PΦB adducts (Lagarias and Rapoport, 1980, Figure 4B). This finding allowed the constitution of photoreversible phytochromes by expressing recombinant phytochrome proteins in yeast and assembling them in vitro. Analysis of reconstituted recombinant phyA, phyB, phyC and phyE revealed that they have similar but not identical spectral properties (Kunkel et al., 1996; Remberg et al., 1998; Eichenberg et al., 2000). Further, overexpression of the mammalian billiverdin reductase in Arabidopsis was found to cause the loss of multiple phytochrome activities by degrading phytochromobilin in vivo and constituted a new class of chromophore mutants which is phenotypically stronger than the hy1 or hy2 mutants, supporting the notion that phytochromobilin could serve as the chromophore for multiple phytochromes (Lagarias et al., 1997).

ACTIONS AND INTERACTIONS OF PHYTOCHROME FAMILY MEMBERS

Phytochromes regulate a variety of developmental processes throughout the life cycle of plants. In most instances, the roles of individual phytochromes are studied in the context of specific responses and/or developmental stages. Loss-of-function analyses combined with gain-of-function analyses and higher order mutants of various combinations are revealing the roles of individual phytochrome in regulating different aspects of plant development. It is clear now that different phytochromes play both distinct and overlapping roles within the spectrum of plant photomorphogenesis (Quail, 1998; Table 3).

Seed Germination and Seedlings De-etiolation

At least three phytochromes have been documented to be involved in control of Arabidopsis seed germination: phyA, phyB and phyE. phyA is responsible for the photoirre-versible VLFR responses triggered by a broad spectrum of irradiations (ultraviolet, visible and far-red light), while phyB controls the R/FR photoreversible effects of low fluence response (Reed et al., 1994; Botto et al., 1996; Shinomura et al., 1996). Recently, it was found that phyE also plays a role in the control of seed germination by FR light. Either phyE is directly involved in the photoperception of FR for this response or the action of phyA in mediating seed germination requires the presence of phyE (Hennig et al., 2002).

Following seed germination, light acts to inhibit hypocotyl elongation and promote cotyledon expansion. Under far-red light, phyA is probably the only active photoreceptor (Nagatani et al., 1993; Whitelam et al., 1993). In white or red light, although phyB plays a major role, the long hypocotyl and reduced cotyledon expansion phenotype of its null mutants is enhanced by mutations in phyA or phyD, suggesting that multiple photoreceptors sense white light during de-etiolation (Neff & Van Volkenburgh, 1994; Johnson et al., 1994; Aukerman et al., 1997). The contribution of phyE to seedling de-etiolation seems negligible (Devlin et al., 1998). Interestingly, a recent high-resolution kinetic analysis of the growth of Arabidopsis seedlings revealed that the red light inhibition of hypocotyl elongation is controlled by a sequential and coordinated action of phyA and phyB. phyA contributes to the initial hypocotyl growth inhibition (first 3 hr of irradiation), while phyB functions in the later phase (Parks and Spalding, 1999).

Previous studies with Arabidopsis phyC overexpressors suggested that phyC likely plays a role in primary leaf expansion (Qin et al., 1997). A more accurate assessment of phyC function became possible until recent identification of phyC loss-of-function mutants. phyC mutant seedlings exhibited a partial loss of sensitivity to red light, with longer hypocotyls and smaller cotyledons than wild type seedlings. Further, hypocotyl of phyAphyC double mutant is significantly longer than that of phyC single mutant, suggesting that phyA and phyC acts redundantly to inhibit hypocotyl elongation under red light (Franklin et al., 2003a, Monte et al., 2003). Consistent with this, phyAphyBphyDphyE quadruple mutant still retains some weak responses to red light, suggesting that phyC has some roles in mediating red light signaling (Franklin et al., 2003b). However, phyBphyC double mutants de-etiolated similarly to phyB single mutants, suggesting that phyC function is dependent on phyB (Monte et al., 2003).

Vegetative Development

The incident light signal embedded in ambient light environments usually activates multiple light signaling pathways and the ultimate cellular responses are determined by the combinatorial effects of these signaling pathways. For example, under a leaf canopy, plants sense the low R/FR ratio of light and initiate the shade-avoidance response by increasing the elongation growth of petioles and stems, the length-to-width ratio of leaves, and accelerating flowering (Smith and Whitelam, 1997). The ability of plants to monitor their light environments and change their architecture provides them with a competitive strategy to survive and finish their life cycle in dense stands.

phyB-deficient plants have a constitutive elongated-petiole and early-flowering phenotype, suggesting that phyB plays a major role in the perception of low R/FR signals (Devlin et al., 1996). Although the monogenic phyD mutant plants have no obvious abnormal phenotype, plants impaired in both the PHYB and the PHYD genes display significantly longer hypocotyls under either R or white light and flower earlier than the phyB monogenic mutants suggesting that phyD performs a similar role of phyB (Aukerman et al., 1997; Devlin et al., 1999).

The shade avoidance responses elicited by low R/FR ratios can be effectively phenocopied by end-of-day farred (EOD-FR) treatments. phyAphyBphyD triple mutant still retains the ability to respond to EOD-FR treatments by developing elongated rosette internodes and accelerated flowering responses (Smith and Whitelam, 1997; Whitelam and Devlin, 1997), implicating the actions of phyC and/or phyE in these responses. The isolation of the phyE mutant confirmed this hypothesis. The phyE mutants show no phenotypic alteration unless it is in the phyB mutant background and the phyBphyE double mutants flower much earlier than the phyB monogenic mutants (Devlin et al., 1998; Franklin et al., 2003b). These studies show that phyB, phyD and phyE control shade avoidance responses and inhibit flowering in a conditional redundant manner.

Structural and Molecular Basis for the Differential Functions of Phytochromes

The underlying mechanism for the observed functional differences between different phytochromes and their action modes has begun to be unraveled. Previous experiments by deleting regions, random mutagenesis of the full-length clone, or domain swapping between phyA and phyB have provided some useful information regarding the function of various parts of phytochromes. For example, the N-terminal 70 amino acids (the 6-kD domain) of oat phyA is required for correct chromophore/apoprotein interactions and undergoes a substantial conformational change upon photoconversion of Pr to Pfr. In this region, one subdomain (residues 13–62) is necessary for conformational stability and another one (residues 6–12) is involved in attenuating phytochrome responses (Jordan et al., 1996, 1997). An oat phyA deletion mutant lacking amino acids 7 through 69 is inactive in transgenic tobacco (Nicotiana tabacum, Cherry et al., 1992), whereas deletion of amino acids 1–52 of oat phyA causes a dominant negative interference in Arabidopsis (Boylan et al., 1994). Clough et al (1999) showed that both the N-terminal and C-terminal halves of phyA are essential for Pfr degradation. The N-terminal region provides recognition signals for ubiquitin conjugation, and an intact C-terminal domain is essential for phyA breakdown.

A recent study demonstrated that continuous FR light treatment could be replaced by intermittent FR light pulses to induce the FR-HIR responses. Analysis of these action spectra suggests that neither phyA in its Pr form synthesized in the dark nor in its photoconverted Pfr form is active in inducing the signal. Instead the short-lived signal was produced during phototransformation from Pfr to Pr (Shinomura et al., 2000). This is in sharp contrast with the case of phyB. Alternative irradiation with R and FR light photoreversibly switches on or off the phyB responses, indicating that Pfr is the active form from which the R-induced signal is derived.

At the molecular level, it has been demonstrated that expression of the nuclear photosynthetic gene LHCB in response to red light depends on both phyA and phyB (Reed et al., 1994; Cerdan et al., 1997). However, phyA and phyB respond to light of different wavelengths and fluences (phyA is responsible for VLFRs, and phyB is responsible for LFRs, Hamazato et al., 1997). Further, it has been shown that phyA mediates the activity of the LHCB promoter in response to VLFRs and HIRs by targeting distinct regions of the same LHCB1*2 promoter (Cerdan et al., 2000), suggesting that these different action modes of phytochromes may entail distinct signaling pathways. Consistent with this notion, VLF1 and VLF2 are distinct components of the VLFR pathway of phyA response (Yanovsky et al., 1997), whereas FHY3 primarily acts in the HIR response pathway of phyA (Yanovsky et al., 2000).

PHYTOCHROME SIGNALING AND THE CIRCADIAN CLOCK

Plants use an array of photoreceptors (including phytochromes) to gather information about the light environment for setting the clock to oscillate with a period of about 24 hours. It has been shown that mutations of photoreceptor genes PHYA and PHYB, cause the circadian rhythm of CAB2 promoter activity to oscillate at a pace slower (with a longer period length) than that of the wild type under various light conditions (Somers et al., 1998). This study revealed that in the regulation of the Arabidopsis circadian clock, phyA acts in low intensities of red light and blue light, while phyB functions in high-intensity red light. Recently it was shown that phyD and phyE also act as photoreceptors in red light input to the clock, and that phyA and phyB act additively in red light input to the clock (Devlin and Kay, 2000). On the other hand, the circadian clock affects the outputs (such as floral induction) of light signaling mediated by these photoreceptors, a feedback regulatory mechanism referred to as gating (Millar and Kay, 1996; Anderson et al., 1997). For example, the mRNA abundance and transcriptional levels of PHYA, PHYB and PHYC have been shown to be regulated by the clock with robust rhythmicity, although PHYD and PHYE expression is weakly rhythmic in Arabidopsis (Bognar et al., 1999; McClung, 2001).

Besides the photoreceptors themselves, several Arabidopsis flowering-time genes have been recently isolated and shown to be associated with the function of the circadian clock. Interestingly, these clock-related genes affect both flowering time and hypocotyl elongation (Dowson-Day and Miller, 1999; Somer, 1999), and they include ELF3, ELF4, CO, TOC1, CCA1, LHY, ADO1/ZTL/LKP1, ADO2/LKP2, ADO3/FKF1, and GI (Hicks et al., 1996; Wang and Tobin, 1998; Schaffer et al., 1998; Fowler et al., 1999; Park et al., 1999; Nelson, et al., 2000; Somers et al., 2000; Strayer et al., 2000; Putterill et al., 1995; Doyle et al., 2002; Table 2). CO is a zinc finger type transcription factor (Putterill et al., 1995), whereas CCA1 and LHY are both MYB-type transcription factors (Wang and Tobin, 1998; Schaffer et al., 1998). ELF3 encodes a nuclear localized putative transcriptional regulator (Liu et al., 2001) and ELF4 encodes a novel protein with no significant homology to proteins of known function (Doyle et al., 2002; Khanna et al., 2003). TOC1 encodes a nuclear protein containing an atypical response regulator receiver domain and two motifs related to transcriptional regulation: a basic motif conserved within the CO family of transcription factors and an acidic domain (Strayer et al., 2000). GI encodes a novel protein localized to the nucleus (Fowler et al., 1999; Huq et al., 2000a). ADO1/ADO2/ADO3 are products of a three-gene family and they all contain an aminoterminus PAS domain, multiple kelch repeats, and an F-box (Jarillo et al., 2001).

It has been proposed that ELF3 and GI are most likely components of the light input pathway to the circadian clock, whereas TOC1, CCA1 and LHY are likely to be components of the central oscillator. Like phyB mutants, both the gi and the elf3 mutants display elongated hypocotyls in red light. However, the gi mutants are late flowering, which is in contrast with the early flowering phenotype of the phyB and elf3 mutants. This clearly suggests that ELF3 and GI play different roles or use different mechanisms in controlling hypocotyl elongation and flowering responses. Intriguingly, both ELF3 and ADO1/ZTL/LKP1 are capable of directly interacting with phyB (Liu et al., 2001; Jarillo et al., 2001). It has been suggested that ELF3 and ADO1 could modulate phyB signaling through such interactions (Matsushita et al., 2003), however; the mechanism of such regulation is still unknown. It was recently shown that phyB and ADO1/ZTL do not seem to affect each other's protein levels (Somers et al., 2004). It has been shown that CCA1 binds in a circadian fashion to a short element of the LHCB1*1 (CAB2) promoter, which is sufficient to confer phytochrome responsiveness and circadian transcription (Wang et al., 1997). Interestingly, the expression of CCA1 and LHY itself is under the control of phytochrome signaling (Martinez-Garcia et al., 2000), suggesting a mechanism for phytochrome control of light input into the clock.

PHYTOCHROMES AS LIGHT-REGULATED KINASES

How do phytochromes initiate their signal transduction upon photo-activation? A long-standing, but much disputed hypothesis is that phytochromes act as light-regulated kinases (Wong et al., 1986; Kim et al., 1989). The recent discovery of phytochrome-like photoreceptors in bacteria, collectively called bacteriophytochromes (BphPs), has generated supporting evidence for such a view (Fankhauser, 2000; Vierstra and Davis, 2000). Phytochrome-like sequences were identified in the cyanobacteria Fremeyella diplosiphon, Synechocystis sp. PCC6803 (Hughes et al., 1997; Hughes and Lamparter, 1999; Kehoe and Grossman, 1996; Wu and Lagarias, 2000), the purple photosynthetic bacterium Rhodospirillum and non-photosynthetic bacteria such as Deinococcus radiodurans, Pseudomononas putida and Pseudomonas aeruginosa (Jiang et al., 1999; Davis et al., 1999). Some of them, such as Cph1 of Synechocystis sp. PCC6803, can bind to the plant phytochrome chromophore (phytochromobilin PΦB or phycocyanobilin PCB) autocatalytically and display red/far-red absorption spectra similar to plant phytochromes. Further, Cph1 was shown to be a light-regulated histidine kinase. Both autophosphorylation of Cph1 and transphosphorylation of Rcp1 (the response regulator for Cph1) are inhibited by red light and stimulated by far-red light (Yeh et al., 1997; Lamparter et al., 1997), suggesting that the phosphorylation/dephosphorylation event of Rcp1 might be physiologically relevant.

Higher Plants Phytochromes

Higher plants phytochromes share limited sequence similarity with BphPs at their C-termini (Schneider-Poetsch, 1992, Figure 5A). However, plant phytochromes have two additional domains: a serine-rich N-terminal extension domain (NTE) and a PAS repeat domain (PRD) located between the CBD and the HKRD (Figure 5). In addition, mutating several critical residues required for bacterial His kinase activity does not affect the activity of plant phytochromes, suggesting that plant phytochromes are not active His kinases (Vierstra and Davis, 2000). Recently, Clark Lagarias's group developed a recombinant system to express and purify plant phytochromes in yeast. Purified oat phyA expressed in S. cerevisiae and the green alga Mesotaenium caldariorum phytochrome expressed in P. pastoris were assembled with PCB in vitro. These phytochromes display the expected spectroscopic properties and protein kinase activity in a light and chromophore-regulated manner. In addition, the purified oat phyA can phosphorylate histone H1 and Rcp1, the response regulator substrate of Cph1. However, unlike their cyanobacterial counterparts, they auto-phosphorylate on Ser/Thr rather than His/Asp. These results provided strong evidence for the kinase property of plant phytochromes (Yeh and Lagarias, 1998).

Physiological Roles of Phytochrome Kinase Activity

The claim that higher plant phytochromes function as protein kinases invite many questions such as what is the biological role of the kinase activity? Which form is more active? Pr or Pfr? Answers to these questions have just begun to be unraveled. For example, recombinant oat phyA is a light and chromophore-modulated protein kinase with Pfr being more active than Pr (Yeh and Lagarias, 1998). A number of phosphorylation sites have been mapped for oat phyA. The Ser-7 is phosphorylated in vivo in both the Pr and Pfr forms, and mutagenesis studies suggest that this residue is implicated in down-regulation of phyA signaling (Stockhaus et al., 1992; Figure 5). The Ser-17 is phosphorylated by protein kinase A in vitro only in the Pr form. Mutation of the first 10 Ser of phyA to Ala (all contained within the first 20 aa) or deletion of this region results in a mutant that is hypersensitive to light, suggesting that these Ser residues are involved in desensitization of phyA (Stockhaus et al., 1992; Jordan et al., 1996, 1997).

Another serine residue, Ser-598, is preferentially phosphorylated in the Pfr form in vivo, and a S598K mutant loses light-regulated kinase activity (Fankhauser et al., 1999). Further, when an oat PHYA cDNA with a Ser-598 to Ala substitution was expressed in the phyA mutants, the transgenic plants exhibited hypersensitivity to far-red light, suggesting that the Ser-598 phosphorylation may also serve as a desensitizing mechanism, possibly by disrupting the interactions between phytochromes with their downstream signaling partners (Park et al., 2000, Figure 5). In addition, a recent report showed that phyA is a substrate of a phosphatase, named FyPP (for flower-specific, phytochrome-associated protein phosphatase). FyPP dephosphorylates phyA in a light-dependent manner and modulates phyA control of flowering time (Kim et al., 2002). Taken together, these results suggest that phosphorylation and dephosphorylation play important roles in regulating phytochrome in vivo function.

It should also be noted that there is no evidence for the kinase activity of phyB-phyE and how it may affect light signaling. A recent study showed that although point mutations in the HKRD region of phyB cause strong phenotypes in hypocotyl length and flowering time, indicating that this domain is important for phyB signaling, deletion of this domain resulted in a milder phenotype, suggesting that this domain is dispensable (Krall and Reed, 2000).

PHYTOCHROME-INTERACTING SIGNALING PARTNERS

Phytochrome-interacting Partners (PIFs)

Protein-protein interactions are necessary for many signal transduction cascades. Both general screenings for phytochrome interacting partners and targeted protein-protein interaction studies have identified a number of phytochrome-interacting factors (PIFs). Those include PIF3 (Ni et al., 1998), PKS1 (Fankhauser et al., 1999), NDPK2 (Choi et al; 1999), cryptochromes (both CRY1 and CRY2) and the AUX/IAA proteins (Ahmad et al., 1998; Colón-Carmona et al., 2000; Mas et al., 2000; Figure 6; Table 2). The physiological roles for some of these factors in phytochrome signaling have been substantiated by recent molecular genetic studies. PIF3 is a nuclear localized basic helix-loop-helix (bHLH) protein and was previously thought as a positive regulator of phyB signaling, as transgenic Arabidopsis seedlings with antisense-imposed reduction in PIF3 levels exhibited strongly reduced responsiveness to red light signals (Ni et al., 1998), whereas a T-DNA tagged pif3 mutant (designated poc1) exhibits enhanced responsiveness to red light. This exaggerated response of the poc1 mutants to red light was originally interpreted as being associated with PIF3 overexpression (Halliday et al., 1999). However, recent re-evaluation of PIF3 function through characterization of T-DNA insertion loss-of-function mutants as well as overexpression studies with full-length PIF3 showed that PIF3 negatively regulates phyB-mediated hypocotyl elongation and cotyledon opening, but promotes both phyA-and phyB-mediated CHS induction (Kim et al., 2003). These studies suggested that phenotypes associated with antisense expression of genes belonging to gene family members (such as PIF3) should be interpreted with extreme cautions, since multiple homologous genes could be simultaneously affected. The exaggerated red light response of the poc1 mutants might be caused by other complications of T-DNA mutagenesis (such as altered expression pattern) rather than PIF3 overexpression as was claimed originally (Halliday et al., 1999). PKS1 is a basic, soluble, cytoplasmic protein and has been proven to be a substrate for light-regulated phytochrome serine/threonine kinase activity. PKS1 overexpressing plants display less sensitivity to red light, suggesting that it acts as an inhibitor of phyB signaling (Fankhauser et al., 1999). Recent studies with loss-of-function mutants of PKS1 and its closest homolog PKS2 suggested that they are primarily involved in the phyA-mediated VLFR responses. Null mutants of PKS1 or PKS2 show enhanced phyA-mediated VLFR responses (with enhanced cotyledon opening, inhibition of hypocotyl elongation, and blocking of greening responses to brief pulses of FR but not to continuous FR or hourly pulses of R), whereas a pks1pks2 double mutant has no phenotype. In addition, overexpression of either PKS1 or PKS2 causes the same phenotype as the pks1 or pks2 single null mutant. These results suggest that PKS1 and PKS2 form a regulatory loop controlling phyA signaling, possibly through their direct physical interaction with each other (Lariguet et al., 2003). NDPK2 (nucleoside diphosphate kinase 2) appears to be a positive regulator of both phyA and phyB signaling. Although hypocotyl elongation is not obviously affected by this locus, its loss of function alleles have a small but significant reduction in cotyledon greening and opening of the hypocotyl/cotyledon hook during de-etiolation (Choi et al; 1999).

Both cryptochromes and the AUX/IAA proteins were identified as in vitro substrates of phytochrome kinase activity (Ahmad et al., 1998; Colón-Carmona et al., 2000). Although the phosphorylation of CRY1 is not light-dependent in an in vitro experiment, in vivo analysis shows that cry1 phosphorylation is stimulated by red light. Moreover, the identification of the shy2 mutant as a suppressor of phyB and of phytochrome chromophore mutants and the gene encodes IAA3, one of the early auxin-inducible genes (Tian and Reed, 1999), suggesting that the interactions between phytochromes and the AUX/IAA proteins and phosphorylation events are likely to be biologically relevant.

A response regulator from Arabidopsis, ARR4, specifically interacts with the extreme N-terminus of phyB both in vivo and in vitro. Interaction of ARR4 with phyB results in the stabilization of the active Pfr form of the photoreceptor. Accordingly, transgenic Arabidopsis plants overexpressing ARR4 display hypersensitivity to red light with shorter hypocotyls and delayed flowering. Further, transgenic plants overexpressing a mutated form of ARR4 in which the Asp involved in phosphotransfer was substituted by an Asn residue revealed a hyposensitive phenotype regarding all phyB-dependent light responses. These data indicate that phyB is the target of a novel two-component phosphorelay system that modulates red-light-dependent signaling by direct interaction of its response regulator (ARR4) with the photoreceptor (Sweere et al., 2001; Table 2). In addition, ARR4 expression is rapidly induced by cytokinin (Brandstatter and Kieber, 1998), suggesting a possible link between cytokinin signaling and red light response mediated by phyB.

GENETICALLY IDENTIFIED EARLY INTERMEDIATES OF PHYTOCHROME SIGNALING

Genetic screening for Arabidopsis mutants potentially defective in signaling intermediates either specific to phyA or phyB, or shared by both pathways has identified a number of candidate loci (Hudson, 2000; Figure 7). There are both positively and negatively acting components in each pathway, and many of them have been characterized at the molecular level (for reviews, see Quail, 2002; Wang and Deng, 2003, Table 2).

Phytochrome A-specific Signaling Components

Mutants affected in phyA-specific signaling process were screened under a continuous far-red light (FRc) condition and a number of potential signaling components specific to this pathway have been identified, including FHY1, FHY3 (Whitelam et al., 1993), FIN2 (Soh et al., 1998), FIN5 (Cho et al., 2003); SPA1 (Hoecker et al., 1998), FAR1 (Hudson et al., 1999), FIN219 (Hsieh et al., 2000), PAT1 (Bolle et al., 2000), EID1 (Buche et al., 2000), HFR1/RSF1/REP1 (Fairchild et al., 2000; Fankhauser and Chory, 2000; Soh et al., 2000), LAF1 (Ballesteros et al. 2001) and LAF3 (Hares et al., 2003). The fhy1, fhy3, fin2, fin5, fin219, far1, laf1, laf3, laf6, and hfr1/rep1/rsf1 mutants show less sensitivity in continuous far-red light, indicating that their respective genes encode positive regulators of the phyA signaling pathway. On the other hand, mutations in the SPA1 and EID1 genes cause increased sensitivity to the FR light signal, and it is most likely that they act as negative regulators of the signaling cascade.

Among the positive regulators of phyA signaling identified, their loss-of-function mutants (mostly null mutation alleles) only exhibit partial defects with different spectra and strength in phyA signaling, suggesting that phyA signaling involves multiple branches or parallel pathways controlling overlapping yet distinctive sets of far-red light responses (hypocotyl growth, apical hook and cotyledon opening, anthocyanin accumulation, far-red light pre-conditioned blocking of greening, gravitropic response, light-regulated gene expression etc., Barnes et al., 1996; Hudson, 2000; Wang and Deng, 2002). Moreover, the finding that the double mutants fhy3-1/far1-2, fhy3-1/fhy1-1 and far1-2/fhy1-1 all display an additive effect, whereas the fhy3-1/spa1-3 double mutant has an intermediate hypocotyl-length phenotype (Wang and Deng, 2002), also indicates that there is no simple downstream/upstream relationship among these phyA signaling components. Further, complex genetic relationships such as non-allelic non-complementation between fin2 and fhy3-1 as well as between fin219 and fhy1 have been reported (Soh et al., 1998; Hsieh et al., 2000), suggesting that their gene products may directly interact or engage in extensive crosstalk.

Several phyA signaling intermediates have been characterized at the molecular level (Table 2). LAF6 is a plastid-localized ATP-binding-cassette protein involved in coordinating intercompartmental communication between plastids and the nucleus (Møller et al., 2000). PAT1 is a new member of the GRAS family (Bolle et al., 2000), whereas FIN219 is a GH3-like protein whose expression is rapidly inducible by auxin (Hsieh et al., 2000). Both PAT1 and FIN219 are cytoplasmically localized proteins. LAF3 encodes two isoforms that differ only at their N termini due to the use of two alternative transcription initiation sites. Both isoforms of LAF3 protein are predicted to contain a transmembrane helical region and have been shown to localize to the perinuclear region, suggesting that LAF3 might regulate nucleo-cytoplasmic trafficking of an inter-mediate(s) involved in phyA signal transduction (Hare et al., 2003). FHY1, FAR1, FHY3, SPA1, HFR1, LAF1 and EID1 are all nuclear localized factors. Interestingly, FAR1 and FHY3 encode two closely related proteins that constitute one branch of a small gene family (Hudson et al., 1999; Wang and Deng, 2002). FHY1 is a novel light regulated protein that accumulates in dark-grown but not farred grown hypocotyl cells (Desnos et al., 2001). HFR1 is an atypical bHLH transcription factor closely related to PIF3 (Fairchild et al., 2000) and LAF1 is a MYB type transcription activator (Ballesteros et al. 2001). SPA1 possesses a C-terminal WD-40 repeat domain that is most closely related to that of COP1 (Hoecker et al., 1999), whereas EID1 is a novel F-box protein most probably involved in ubiquitin-dependent proteolysis (Dieterle et al., 2001). The biochemical functions of FHY1, PAT1, LAF3, FIN219, FAR1, FHY3 and SPA1 remain largely unknown. Determining the protein-protein interactions among these phyA signaling intermediates and identifying their novel interactive partners should enhance our understanding of the phyA signaling pathway.

Interestingly, a jasmonate response locus, JAR1, was mapped to the same locus as fin219 (Hsieh et al., 2000, Staswick et al., 2002). Unlike fin219 mutants, several jar1 alleles are not defective in hypocotyl elongation under farred light, and unlike the jar1 alleles, fin219 shows no increase in resistance to MeJA (Staswick et al., 2002). This discrepancy could be due to the epigenetic nature of fin219 mutation (Hsieh et al., 2000). Further studies are required to clarify this issue.

Phytochrome B-specific Signaling Components

Putative phyB-specific signaling mutants have also been identified, including RED1, PEF2, PEF3, GI, ELF3, ELF4, and SRR1 (Huq et al., 200094; Liu et al., 2001; Staiger et al., 2003; Doyle et al., 2002; Khanna et al., 2003; Table 2). They share a number of features with phyB mutants, such as a long hypocotyl phenotype specifically under red light, early flowering in short days, and elongated petioles, suggesting that these loci positively regulate phyB signaling (Reed et al., 1993; Ahmad and Cashmore, 1996; Wagner et al., 1997). On the other hand, the srl1 and srl2 mutants show enhanced responsiveness to red light, suggesting that SRL1 and SRL2 are negatively acting components specific to phyB signaling (Huq et al., 2000b). SRL2 was shown to encode a PIF3-related protein named PIF4 (Huq and Quail, 2002). ELF3 encodes a nuclear-localized putative transcriptional regulator that interacts with phyB (Liu et al., 2001). GI is a novel protein that is localized to the nucleus (Fowler et al., 1999; Huq et al., 2000a). SRR1 codes for a nuclear/cytoplasmic protein that is conserved in most eukaryotes, however; their biochemical activity is not apparent due to the lack of any recognizable domains (Staiger et al., 2003). The ELF4 gene encodes a novel 111-amino acid protein with no significant homology to proteins of known function (Doyle et al., 2002; Khanna et al., 2003). red1 is allelic to sur2 and atr4 and its wild type gene encodes a cytochrome P450, CYP83B1. The enzyme catalyzes the N-hydroxylation of indole-3-acetaldoxime (IAOx), which is the first committed step in the biosynthesis of indole glucosinolates. IAOx also serves as a precursor for the biosynthesis of IAA. In red1/sur2/atr4 mutants, synthesis of indole glucosinolates is blocked, leading to hyperaccumulation of auxin and reduced responsiveness to red light. CYP83B1 transcript levels are induced specifically by continuous red light (Rc) treatment, suggesting that Rc-induction of the CYP83B1 transcript and auxin homeostasis are necessary for normal seedling de-etiolation under red light (Hoecker et al., 2004). The molecular identities of SRL1, PEF2 and PEF3 are currently unknown.

Signaling Components Shared by Both phyA and phyB

Two mutants, pef1 and psi2, affect responses from multiple photoreceptors. The pef1 mutants show attenuated red and far-red responses, whereas the psi2 mutant is hypersensitive to red and far-red light, and has necrotic lesions in light-grown plants (Ahmad and Cashmore 1996; Genoud et al., 1998), suggesting that these two loci are shared by both the phyA and phyB signaling pathways. The identities of these genes have not been reported yet.

Recently, COG1, PFT1, PRR7 were reported as signaling components shared by both phyA and phyB (Park et al., 2003; Cerdan and Chory, 2003; Kaczorowski and Quail, 2003; Table 2). COG1 encodes a Dof transcription factor and its overexpression causes hyposensitive responses to red and far-red light whereas transgenic lines expressing antisense COG1 were hypersensitive to red and far-red light, suggesting that COG1 functions as a negative regulator of both phyA and phyB signaling (Park et al., 2003). pft1 loss-of-function mutants were hyposensitive to far-red and hypersensitive to red light, and these altered responses require functional phyA and phyB respectively. The opposite effects of pft1 mutation on phyA and phyB signaling suggest that PFT1 may function at a node where phyA and phyB signaling converge (Cerdan and Chory, 2003). In addition, pft1 mutation causes late-flowering under long-day conditions (16h light/8h dark) and suppresses the early flowering phenotype of phyB in the pft1phyB double mutant, suggesting that PFT1 acts downstream of phyB to regulate flowering time. PFT1 encodes a putative protein of 836 amino acids, with a predicted cWF-A (von Willebrand factor type A) domain in the N-terminus and a glutamine-rich region in the C-terminus. Its ability to activate transcription in yeast suggests that PFT1 could function as a transcriptional co-activator (Cerdan and Chory, 2003). prr7 loss-of-function mutants exhibit reduced sensitivity to both red and far-red light, suggesting that PRR7 is a positive regulator of both phyA and phyB signaling. PRR7 is a member of the Arabidopsis pseudo-response regulator gene family and its gene product is nuclear localized. PRR7 is also involved in the regulated expression of CCA1 and LHY, whose gene products are essential components of the circadian clock. Consistent with this, prr7 mutants also display a defect in clock regulation of gene expression including PRR7 itself (Kaczorowski and Quail, 2003).

LIGHT-REGULATED CELLULAR LOCALIZATION OF PHYTOCHROMES

Being synthesized within the cytosol, where phytochrome apoproteins assemble autocatalytically with the plastid-derived chromophore. Early studies using an immunohistological approach and cell fractionation assay suggested that phytochromes are predominantly localized outside the nucleus (Pratt, 1994). Recently, it was shown that upon photoconversion of Pr to Pfr, both phyA and phyB tagged with GUS or green fluorescent protein (GFP) can translocate from cytoplasm into the nucleus where they form intranuclear bodies (NBs) (Sakamoto and Nagatani, 1996; Yamaguchi et al. 1999; Kircher et al., 1999; Nagy et al., 2000). For example, phyB:GFP is localized in the cytosol in etiolated seedlings. phyB:GFP translocates into the nucleus and forms NBs upon treatments with continuous red light treatment, or multiple red light pulses, and this process can be reversed by subsequent far-red treatment, indicating that the nuclear import of phyB is mediated by a low-fluence response (LFR) of phytochrome. On the other hand, nuclear import and NBs formation of phyA:GFP is very rapid (one magnitude of order faster than that of phyB) and can be induced by brief irradiation with red, farred, or blue light. In addition, nuclear translocation of phyA can also be induced by continuous far-red but not by continuous red light treatment. These observations suggest that the nuclear translocation of phyA is mediated by both the very low fluence response (VLFR) and the far-red high irradiance response (HIR). Further, phyA nuclear translocation is preceded by an even faster cytosolic spot formation of the fusion protein, a phenomenon mimicking SAP (sequestered area of phytochrome) formation, and that the transport seems to be not complete, i.e. a significant portion of phyA:GFP remains cytosolic (Nagy et al., 2000).

A recent report described the steady-state dynamics of phyB localization in response to different light conditions using a 35S-phyB:GFP line. Four different phyB subcellular localization patterns were defined: diffuse nuclear localization, small and numerous NBs only, both small and large NBs, and large NBs only. The amounts of phyB accumulates in NBs is mainly determined by the percentage of active Pfr conformer among the total amount of phyB protein, with large phyB NBs correlate with strong phyB responses (Chen et al., 2003). However, the function of phytochrome NBs is still not clear. They could be sites of sequestration for downregulating phy signaling or active transcription complexes in which phy physically interact with their downstream partners to regulate gene expression.

GFP fusions of full-length phyC-E revealed that the nucleocytoplasmic partitioning of these phytochromes seems not to be regulated by light. They are detected in both the cytosol and the nucleus in etiolated seedlings. However, the formation of NBs for these fusion proteins is light dependent, similar to phyA or phyB:GFP (Nagy et al., 2000). Interestingly, the nuclear import of all phytochrome species is regulated by the circadian clock and displays oscillations under constant conditions, suggesting a new regulatory mechanism modulating phytochrome signaling by the circadian clock (Kircher et al., 2002).

It was until recent that the biological significance of the nuclear translocation events for phytochromes was substantiated by experimental data. Huq et al. (2003) used a glucocorticoid receptor (GR)-based fusion protein system to control the nucleocytoplasmic partitioning independent of the light signal. The phyB:GR fusion protein rescued the seedling etiolation phenotype of phyB mutant, only in the presence of the steroid dexamethasone (Dex) under continuous red light, suggesting that the Dex-dependent nuclear translocation of phyB:GR is necessary for phyB function (Huq et al., 2003).

At the current stage, very little is known about the molecular machinery and factors modulating the nucleocytoplasmic partitioning and NB formation of phytochromes. It has been speculated that the conformation change to Pfr, their autophosphorylation patterns and their interaction with other proteins may all attribute to the control of their nuclear import (Nagy et al., 2000). In addition, the various phytochrome signaling intermediates described above and other signaling cascades (such as phytohormones) could also affect the intracellular distribution of phytochromes. It is interesting to note that a genetic screen to identify factors required for phyB NB formation resulted in several phyB mutants or mutants defective in chromophore biosynthesis, suggesting that functional phytochrome themselves are essential for NB formation (Chen et al., 2003). Further studies aimed to map the possible nuclear localization signal (NLS) and nuclear export signal (NES), and elucidation of the molecular composition of NBs should shed lights on the control of phytochrome localization.

COP/DET/FUS PROTEINS AND REGULATED PROTEOLYSIS IN PHYTOCHROME SIGNALING

Aside from these photoreceptor-specific signaling intermediates, genetic screens have also identified eleven pleiotropic negative regulators of photomorphogenesis (the COP/DET/FUS loci) and one positive regulator (HY5), which function downstream of both phytochromes and cryptochromes (Wei and Deng 1996; Figure 8A and B). HY5 encodes a bZIP transcription factor whose abundance is correlated with the extent of photomorphogenesis. HY5 activity is primarily regulated at the level of protein stability, which requires the COP/DET/FUS group of constitutive repressors (Oyama et al., 1997; Osterlund et al., 2000). The COP/DET/FUS group of proteins defines four biochemical entities: COP1, COP10, DET1, and the COP9 signalosome [CSN]). Among the COP/DET/FUS proteins, DET1 is distinct and likely involved in controlling chromatin remodeling and gene expression (Pepper et al., 1994; Benvenuto et al., 2002; Schroeder et al., 2002). The other three groups of COP/DET/FUS proteins are all likely involved in regulated proteolysis. COP10 is an E2 ubiquitin conjugating enzyme variant (Suzuki et al., 2002). The COP9 signalosome (CSN) is composed of eight COP/DET/FUS proteins (CSN1-CSN8) and it shares subunit-by-subunit similarity to the lid subcomplex of the 26S proteasome (Kwok et al., 1999; Peng et al., 2001a, 2001b; Serino et al., 1999, 2002; Castle and Meinke, 1994; Karinol et al., 1999; Wei et al., 1994; Schwechheimer and Deng, 2000).

Of these constitutive photomorphogenic repressors, COP1 is capable of directly interacting with HY5 in the nucleus through its WD40 repeat domain and targets HY5 for proteasome-mediated degradation (Ang et al., 1998; Osterlund et al., 2000). Besides a C-terminal WD40 repeat domain, COP1 also contains a RING finger domain in its N-terminal region, followed by a coiled coil in the middle (Deng et al., 1992). RING-finger domains are commonly conserved in a subclass of ubiquitin-protein ligases (E3s, Callis and Vierstra, 2000). It has been recently demonstrated that COP1 could act as an E3 ubiquitin ligase for targeted degradation of HY5, LAF1 (a myb transcription factor) and possibly HYH (a homolog of HY5) (Seo et al., 2003; Osterlund et al., 2000; Holm et al., 2002; Saijo et al., 2003; Figure 9A). Through targeted destabilization of HY5 and other photomorphogenesis-promoting factors, COP1 represses photomorphogenesis. Further, it was recently demonstrated that COP1 could also act as an E3 ligase for the photoreceptor phyA, leading to its elimination, and subsequent termination of phyA signaling (Seo et al., 2004).

Light-induced photomorphogenic development requires the inactivation of these COP/DET/FUS proteins. However, little is known as to how the light-activated photoreceptors (including phytochromes) regulate the activities of those downstream COP/DET/FUS proteins. At least one of the mechanisms is suggested to be the light triggered nuclear depletion of COP1 (von Arnim and Deng, 1994; Osterlund et al., 1999; Osterlund and Deng, 1998; Figure 9B). However, the slow kinetics of COP1 nuclear depletion (could take up to 24 hours, von Arnim and Deng, 1994) suggests an additional, unknown regulatory event (s) is required to account for the early and rapid inactivation of COP1 by light. Blue light signaling has been reported to involve a direct protein interaction of cryptochromes with COP1 (Wang et al., 2001; Yang et al., 2001). However, little information is available regarding the molecular link between phytochrome activation and COP1 inactivation.

Recently, three independent studies showed that SPA1, a negative regulator of phytochrome A (phyA) signaling, physically interacts with COP1 in vitro and in vivo, and that this interaction is reduced by light (Hoecker and Quail, 2001, Saijo et al., 2003; Seo et al., 2003). SPA1 encodes a 114 kDa nuclear protein with several structural motifs: a protein kinase-like domain at its N terminus, a coiled coil region in the middle, and a C-terminal WD40 domain that is highly related in structure to that of COP1 (Hoecker et al., 1999). It was demonstrated that SPA1 modulates the E3 ubiquitin ligase activity of COP1 on HY5 and LAF1 in vitro, leading to their degradation by the 26S proteasome and desensitized light signaling (Saijo et al., 2003; Seo et al., 2003). It is conceivable that SPA1, through direct physical interaction with COP1 and direct interaction with substrate proteins, may help to capture the substrate proteins and bring them to the proximity of COP1. This might be particularly important for light-grown seedlings, as upon light illumination, the abundance of nuclear COP1 is reduced (von Arnim and Deng, 1994) and its efficiency of ubiquitinating substrate proteins in the nucleus will become weakened. Further, the observation that spa1-3 mutation enhances cop1 mutant phenotypes in the dark (Saijo et al., 2003) suggests SPA1 has some functional role in darkness as well. SPA1 might synergize with COP1 in the dark to deplete transcription factors, like HY5, HYH and LAF1, more rapidly to prepare plants for upcoming light signals. This model is in good agreement with the results of a physiological study which showed spa1 mutant was more persistent in both the very-low-fluence response (VLFR) and continuous far-red light induced high-irradiance responses (HIRs) (Baumgardt et al., 2002), possibly due to the delayed turnover of these photomorphogenesis-promoting transcription factors.

The Arabidopsis genome contains three additional SPA1-related genes, named SPA2, SPA3 and SPA4 (Table 2). Seedlings with mutations in SPA3 or SPA4 also exhibit enhanced photomorphogenesis in the continuous far-red, red and blue light, but show no phenotype in the dark, suggesting that SPA3 and SPA4 are also two negative regulators of light signaling. In addition, SPA3 and SPA4 can interact with COP1 in a yeast two-hybrid assay and in vitro interaction assay, suggesting that they may share a conserved mechanism with SPA1 in fine tuning light signaling (Laubinger and Hoecker, 2003).

The finding that EID1, another negative regulator of phyA signaling, encodes an F-box protein and a likely component of a SCFEID1 ubiquitin ligase complex (Dieterle et al., 2001) reinforce a possible role of targeted protein degradation in phyA signaling. It is conceivable that phyA signaling may utilize the SCFEID1 ligase to mediate regulated proteolysis of various signaling components. Since the level and degradation of phyA are not altered in the eid1 mutants, FHY3, FAR1, LAF1, PIF3 and HFR1 might be the more likely candidates for the targeted degradation. A defect in the degradation of these positive regulators of phyA signaling could well underlie the observed hypersensitivity of the spa1 and eid1 mutants to FR light. It is also possible that SPA1 and the SCFEID1 complex may function together with COP1 and the COP9 signalosome to control the timing or the amplitude of the response by timely and precisely regulating the abundance of these signaling molecules.

PHYTOCHROME REGULATION OF NUCLEAR GENE EXPRESSION

The photoregulation of gene expression in higher plants has been extensively studied during the past two decades and a number of photoregulated genes have been identified from several plant species. For example, the transcript level of both the LHCB genes for light-harvesting chlorophyll a/b-binding protein of photosystem II and the RBCS genes for ribulose-1, 5-biphosphate carboxylase/oxygenase small subunit increases by light illumination, and this light-inducible gene expression is mediated by phytochromes. Consequently, the promoter region of LHCB and RBCS genes have been isolated and analyzed to identify various cis-regulatory elements responsible for light- or phytochrome regulation, such as the GT-1 boxes of RBCS, G-box of tomato RBCS-3A, GATA motif, 3AF binding sites and AT-1 binding site. In addition, trans-acting factors which bind to light-responsive cis-regulatory elements (LREs) such as GT-1, GBF, GAF1, 3AF-1 and AT-1 were isolated, and some of which have been shown to be involved in phytochrome-mediated light responsiveness (von Arnim and Deng, 1996; Terzaghi and Cashmore, 1995; Menkens et al., 1995; Kuno et al., 2000; Kuno and Furuya, 2000). Further, it was suggested that the combinatorial interaction of multiple LREs is the key determinant for mediating light control of promoter activity (Puente et al., 1996). However, these traditional studies only revealed limited information on the role of individual phytochromes in the photoregulation of gene expression.

Recently, the newly developed gene chip technology has been applied to study light regulation of gene expression and to define the roles of individual phytochromes (Ma et al., 2001). It was revealed that a large number of genes, possibly over one third of the genome, are coordinately regulated by various light signals. Utilization of the phyA and phyB null mutants under specific light conditions confirmed that phyA seems to be the primary photoreceptor for mediating far-red light regulation of gene expression, whereas phyB is only one of the phytochromes mediating red light regulation of genome expression (Ma et al., 2001).

The genome profiling study revealed an interesting feature of light-regulated gene expression: many cellular metabolic and regulatory pathways are found to be coordinately regulated by light. Some of them (such as all photo-synthetic genes, glycolysis and the TCA cycle etc.) are activated by light, whereas others (such as cell wall-loosening enzymes and water channel protein aquaporins) are repressed by light (Ma et al., 2001). Similar conclusions have also been drawn from studies on circadian clock control of gene expression (Harmer et al., 2000; Schaffer et al., 2001) and genomic analyses of shade avoidance response in Arabidopsis (Devlin et al., 2003).

The microarray technology has also been used to examine the functional relationship of various genetically identified phyA signaling intermediates. A core group of far-red light regulated genes was defined and used as the basis for a comparative analysis of the genome expression profiles of various phyA signaling mutants. The study suggested that FHY1, FHY3 and FAR1 control large numbers of FR-regulated gene expression and likely represent upstream branch components in the phyA signaling network, whereas FIN219, SPA1 and HFR1 control the expression of smaller sets of genes and likely act more downstream in the network (Wang et al., 2003).

AN EMERGING MODEL OF PHYTOCHROME SIGNALING

A Transcription Hierarchy and Roles of a Class of bHLH Proteins

The recent findings that upon photoconversion of Pr to Pfr, phytochromes can translocate from cytoplasm into the nucleus (Kircher et al., 1999; Nagy et al., 2000) and that the Pfr form of phytochromes is capable of physically interacting with a light-responsive element (G box)-bound bHLH protein, PIF3, suggest that phytochromes can function as transcriptional regulators, which would allow plants to change light-regulated gene expression rapidly (Ni et al., 1999; Martínez-Garcia et al., 2000). Further, Dr. Quail's group found that nearly half of the genes responding to the far-red light signal within the first hour are predicted to encode multiple classes of transcriptional regulators (such as the MYB transcription factors CCA1 and LHY, Tepperman et al., 2001). These results suggest a transcriptional cascade may operate in phytochrome control of gene expression. phyA may directly target the light signal to the promoters of a group of early phytochrome-responsive transcriptional regulators (such as PIF3 and HFR1), which in turn regulates the expression of numerous late phytochrome-responsive transcription factors (such as CCA1 and LHY), leading to sequential expression of downstream target genes constituting the phyA-regulated transcriptional profile and the various ultimate cellular responses (Figure 10).

In addition to PIF3, Quail's group found two additional bHLH proteins involved in phytochrome signaling. One of them, PIF4, was isolated as a PIF3-interacting protein and it also interacts specifically with the Pfr form of phyB, but not phyA. PIF4 could form homodimer with itself or heterodimer with PIF3, and all dimeric forms can bind to the PIF3-binding promoter element. Missense mutants of phyB that are impaired in signaling show reduced binding to PIF4, suggesting a biologically relevant interaction. Overexpressing PIF4 in transgenic Arabidopsis causes a hyposensitive phenotype specific to red light. Conversely, antisense PIF4 plants are hypersensitive to red light. Further, a T-DNA knockout mutant of PIF4, srl2, was independently identified from a genetic screen as a hypersensitive mutant under red light (Huq and Quail, 2002; Toledo-Ortiz et al., 2003). Together, these studies suggest that PIF4 acts as a negative regulator of phyB signaling, similar to PIF3. The apparent phenotypes of pif3 and pif4 loss-of-function suggest that PIF3 and PIF4 have distinct functions. The construction and analysis of pif3/pif4 double mutant should reveal more information about their functional relationship in regulating phyB signaling.

On the other hand, HFR1, another bHLH protein, is originally identified as a positive regulator specific to phyA signaling. It has also been demonstrated that HFR1 is capable of forming a homodimer as well as a heterodimer with PIF3 (Fairchild et al., 2000). However, in contrast to PIF3, HFR1 does not bind to either phyA or phyB, although the HFR1/PIF3 complex can bind preferentially to the Pfr form of both phyA and phyB. Thus, HFR1 may function to modulate phyA signaling via heterodimerization with PIF3. An intriguing question is how HFR1 achieves its specificity in phyA signaling, given that PIF3 binds both phyA and phyB. Further, PIF3 only affects phyB- but not phyA-induced inhibition of hypocotyl elongation, although it affects both phyA- and phyB-induced cotyledon expansion, cotyledon opening, and CHS expression (Kim et al., 2003), suggesting that HFR1 might have a PIF3-indenpendent function (s) in regulating seedling photomorphogenesis, depending on the types of light responses and light conditions. Consistent with this, a recent study showed that HFR1 is a positive regulator of cry1-mediated blue light signaling (Duek and Fankhauser, 2003). Further, HFR1 mRNA is more abundant in far-red and blue light than in continuous red light, suggesting a potential mechanistic basis for the specificity of HFR1 in regulating signaling (Fairchild et al., 2000; Duek and Fankhauser, 2003).

Furthermore, another bHLH protein, named PIL1 (for PIF3-like 1), was recently shown to be required for the rapid shade avoidance responses (such as elongation of internodes and petioles, reduced leaf growth and enhanced apical dominance). PIL1 transcript level increases rapidly up to 35 folds in response to low R/FR ratio (within an hour), and the increase can be detected as early as 8 minutes of lowR/FR. Interestingly, the change of PIL1 transcript levels is reversible upon transferring from low R/FR to high R/FR, and is gated by the circadian clock. It has been proposed that in response to low R/FR, plants sense a change in Pfr level of phyB, D and E, and initiate a rapid increase in PIL1 transcript and protein levels, followed by expression of target genes required for differential elongation growth (Salter et al., 2003). It will be very interesting to define the transcriptional regulators responsible for the rapid increase of PIL1 levels in response to low R/FR ratio.

The Arabidopsis genome contains a large gene family for bHLH proteins, with up to 162 members (Toledo-Ortiz et al., 2003; Bailey et al., 2003; Arabidopsis Genome Initiative, 2000). Due to the known nature of bHLH proteins to form homodimers and/or heterodimers, other bHLH protein could also be involved in regulating phytochrome signaling. It is imaginable that combinatorial heterodimeric interactions of different members of the bHLH family could potentially generate a vast array of transcriptional activities that are essential for plants to respond to the incidental light signal (Quail, 2000, 2002). It should be noted that how phytochromes regulate the transcriptional activities of these bHLH proteins is currently unknown. By interacting with PIF3, phytochromes could regulate transcription either by functioning as a transcriptional co-activator or co-repressor or by altering the biochemical/allosterical nature of PIF3 (Quail, 2000). With the knowledge that the activities of a number of transcription factors are regulated by phosphorylation (Hardtke et al., 2000) and that phytochrome is a light-regulated kinase, it will be interesting to test whether PIF3 is a substrate of phytochrome kinase activity and how it may modulate the function of PIF3.

Cytoplasmic Signaling Events

Besides nuclear signaling events, phytochrome signaling likely entails cytoplasmic events. First, the nuclear import of some phytochrome species (such as phyB) takes hours (Nagy et al., 2000) and their nuclear actions obviously could not explain some rapid phytochrome responses (occur within minutes of irradiation), such as the change in hypocotyl growth rate (Parks and Spalding, 1999). Second, phytochromes interact with a number of cytoplasmic proteins identified so far (such as PKS1 and NDPK2, Fankhauser et al., 1999; Choi et al., 1999). Third, genetic studies have identified cytoplasmic proteins as signaling intermediates (such as PAT1 and FIN219 for phyA signaling, Bolle et al., 2000; Hsieh et al., 2000).

Previous microinjection and pharmacological studies have suggested the involvement of G-proteins, cGMP and Ca2+/calmodulin in phytochrome signaling (Bowler et al., 1994; Neuhaus et al., 1997). Particularly, heterotrimeric G-proteins have been implicated in several processes during plant growth and development, and they transduce extra-cellular signals to the cell. In general, heterotrimeric G-proteins consist of three subunits: α, β, and γ. Analysis of the complete genome sequence of Arabidopsis indicated that the Arabidopsis genome contains only a single canonical Gαgene, previously designated ATGPA1 (Ma et al., 1990), and a single Gβgene, designated AGB1 (Weiss et al., 1994) and possibly one Gγ subunit (Mason and Botella, 2001; Table 2).

It was reported that transgenic Arabidopsis plants conditionally overexpressing the Gα subunit of the heterotrimeric G-protein under the control of a glucocorticoid-inducible promoter exhibited a light-dependent hypersensitive response as a result of reduced hypocotyl cell elongation (Okamoto et al., 2001). However, a separate study reported that loss-of-function gpa1 mutants display partial de-etiolation in the dark, with short hypocotyls and open apical hooks typical of light-irradiated seedlings. The short hypocotyl of gpa1 seedlings is reportedly due to a defect in cell division, not cell elongation (Ullah et al., 2001). To resolve these discrepancies regarding GPA1 function in photomorphogenesis, Jones et al. (2003) re-evaluated the roles of the heterotrimeric G protein employing multiple alleles of both gpa1 and gpb1 (in both Col and WS backgrounds) and their double mutants. They concluded that these mutants have wild-type sensitivity to red and far-red light. In addition, there is no apparent alternation in R or FR sensitivity in transgenic plant overexpressing GPA1 (wild-type or a constitutively active form) or GPB1. The observed shorter hypocotyls and partially open cotyledons of gpa1 and gpb1 mutants grown in the dark are mainly caused by a defect in cell division, rather than cell elongation (Jones et al., 2003).

Genetic studies have also provided supporting evidence for the involvement of Ca2+ in phytochrome signaling. The sub1 mutant exhibits hypersensitive responses to both farred light and blue light. The SUB1 gene was found to encode a Ca2+-binding protein. Genetic interaction studies suggest that SUB1 is a component of a cryptochrome signaling pathway and is a modulator of the phyA signaling pathway. Further, SUB1 negatively regulates HY5, a bZIP transcription factor and a positive regulator of photomorphogenesis (Guo et al., 2001; Table 2).

Some early studies also suggested that phytochrome could exert its effects by rapidly altering the permeability of the plasma membrane to ions. This might be true for certain light responses, such as the bud induction process of the moss Physcomitrella patens and the unrolling of the primary leaf wrapped within the oat coleoptiles (Kendrick and Bossem, 1987; Ermolayeva et al., 1997; Viner et al., 1988). However, there is no report for changes in cytoplasmic ion (such as Ca2+) concentration or ion fluxes in Arabidopsis hypocotyl cells in response to light (Parks and Spalding, 1999; Spalding, 2000). Thus, the role of ion flux and membrane depolarization in phytochrome control of Arabidopsis photomorphogenesis remains elusive.

POTENTIAL APLICATIONS IN AGROBIOTECHNOLOGY

Light is one of the major environmental signals governing many aspects of plant growth and development. Over the past decade or so, we have seen rapid progress in our understanding of the light signaling mechanisms controlling plant development, largely due to the utilization of Arabidopsis thaliana as the model experimental system. As in dicot plants, light controls many aspects of development in monocot plants. For example, seed germination, elongation of coleoptiles and mesocotyls, unrolling of the leaf blade, chloroplast differentiation, pigment synthesis and control of flowering time are all regulated by light in cereals such as oats, maize, barley and rice (Kendrick and Kronenberg 1994; Yano et al., 2001). Light regulation of monocots gene expression, such as that of LHCB (Iwasaki et al., 1997), RBCS (Kyozuka et al., 1993) and PHYA (Kay et al., 1989), has also been reported. The monocots and in particular the grasses, are by far the most agronomically important group of plants. Rice, maize, wheat and sorghum are staple food crops throughout the world and have been under intense human selection pressure for at least 10,000 years. Agronomically important traits such as high yield, early flowering and increased stand densities are all influenced by light (Smith, 1994). Nevertheless, the genetic characterization of light signaling pathways in monocots has just begun. Accumulating evidence suggests that monocots and dicots might share conserved signaling components/mechanisms. For example, mutations in phytochrome genes or function resulted in early flowering in barley, sorghum, rice, and maize (Hanumappa et al., 1999; Childs et al., 1997; Izawa et al., 2000; Basu et al., 2000; Sawers et al., 2002). Homologous proteins for the downstream COP/DET/FUS proteins, a group of general repressors of photomorphogenesis in Arabidopsis, have also been identified in monocots. For example, the rice COP1 is 73% identical to the Arabidopsis COP1 and it can functionally complement an Arabidopsis cop1 mutant (Tsuge et al., 2001). In addition, attempts have been made to genetically modify some horticultural and agricultural crops and promising results have been achieved (Robson et al., 1996; Robson and Smith, 1997; Halliday et al., 1997; Thiele et al., 1999). For example, overexpressing oat phyA in transgenic tobacco and hybrid aspen alleviates the shade-avoidance response, thus increasing dwarfism and harvest index (Robson et al., 1996; Robson and Smith, 1997; Olsen et al., 1997). Overexpressing Arabidopsis phyB in potato increases potential tuber yield both in greenhouse and under field conditions, particularly at high densities (Jackson et al., 1996; Thiele et al., 1999; Boccalandro et al., 2003). In the future, it should be possible to modify agricultural crop plants with genetically improved phytochrome or their downstream signaling molecules to develop transgenic crop species with more fine tuned light-responsive developmental traits to better fit their ambient light environments.

Acknowledgments

We apologize to the authors whose work might have been overlooked here. We thank Thomas Brutnell and Enamul Huq for reading and commenting on the manuscript. The research in Xing Wang Deng's laboratory is supported by grants from NIH and NSF. Research in Haiyang Wang's laboratory is supported by Boyce Thompson Institute and Triad Foundation.

REFERENCES

1.

M. Ahmad and A. R. Cashmore . 1996. The pef mutants of Arabidopsis thaliana define lesions early in the phytochrome signaling pathway. Plant J 10:211103–1110. Google Scholar

2.

M. Ahmad , J. A. Jarillo , O. Smirnova , and A. R. Cashmore . 1998. The CRY1 blue light photoreceptor of Arabidopsis interacts with phytochrome A in vitro. Mol. Cell 1:21939–948. Google Scholar

3.

S. L. Anderson , D. E. Somers , A. J. Miller , K. Hanson , J. Chory , and S. A. Kay . 1997. Attenuation of phytochrome A and B signaling pathways by the Arabidopsis circadian clock. Plant Cell 9:211727–1743. Google Scholar

4.

L-H. Ang , S. Chattopadhyay , N. Wei , T. Oyama , K. Okada , A. Batschauer , and X-W. Deng . 1998. Molecular interaction between COP1 and HY5 defines a regulatory switch for light control of Arabidopsis development. Mol. Cell 1:21213–222. Google Scholar

5.

Arabidopsis Genome Initiative (AGI) 2000. Analysis of the genome sequence of the flowering plant Arabidopsis thaliana. Nature 408:21796–815. Google Scholar

6.

M. J. Aukerman , M. Hirschfeld , L. Wester , M. Weaver , T. Clack , R. M. Amasino , and R. A. Sharrock . 1997. A deletion in the PHYD gene of the Arabidopsis Wassilewskija ecotype defines a role for phytochrome D in red/far-red light sensing. Plant Cell 9:211317–1326. Google Scholar

7.

P. C. Bailey , C. Martin , G. Toledo-Ortiz , P. H. Quail , E. Huq , M. A. Heim , M. Jakoby , M. Werber , and B. Weisshaar . 2003. Update on the basic helix-loop-helix transcription factor gene family in Arabidopsis thaliana. Plant Cell 15:212497–2502. Google Scholar

8.

M. Ballesteros , C. Bolle , L. M. Lois , J. M. Moore , J-P. Vielle-Calzada , U. Grossniklaus , and N-H. Chua . 2001. LAF1, a MYB transcription activator for phytochrome A signaling. Genes Dev 15:212613–2625. Google Scholar

9.

S. A. Barnes , R. B. Quaggio , G. C. Whitelam , and N-H. Chua . 1996. fhy1 defines a branch point in phytochrome A signal transduction pathways for gene expression. Plant J 10:211155–1161. Google Scholar

10.

D. Basu , K. Dehesh , H. J. Schneider-Poetsch , S. E. Harrington , S. R. McCouch , and P. H. Quail . 2000. Rice PhyC gene: structure, expression, map-position and evolution. Plant Mol. Biol 44:2127–42. Google Scholar

11.

R. L. Baumgardt , K. A. Oliverio , J. J. Casal , and U. Hoecker . 2002. SPA1, a component of phytochrome A signal transduction, regulates the light signaling current. Planta 215:21745–753. Google Scholar

12.

G. Benvenuto , F. Formiggini , P. Laflamme , M. Malakhov , and C. Bowler . 2002. The photomorphogenesis regulator DET1 binds the amino-terminal tail of histone H2B in a nucleosome context. Curr. Biol 12:211529–34. Google Scholar

13.

H. E. Boccalandro , E. L. Ploschuk , M. J. Yanovsky , R. A. Sánchez , C. Gatz , and J. J. Casal . 2003. Increased phytochrome B alleviates density effects on tuber yield of field potato crops. Plant Physiol 133:211539–1546. Google Scholar

14.

L. K. Bognar , A. Hall , E. Adam , S. C. Thain , F. Nagy , and A. J. Millar . 1999. The circadian clock controls the expression pattern of the circadian input photoreceptor, phytochrome B. Proc. Natl. Acad. Sci. USA 96:2114652–14657. Google Scholar

15.

C. Bolle , C. Koncz , and N-H. Chua . 2000. PAT1, a member of the GRAS family, is involved in phytochrome A signal transduction. Genes Dev 14:211269–1278. Google Scholar

16.

H. A. Borthwick , S. B. Hendricks , M. W. Parker , E. H. Toole , and V. K. Toole . 1952. A reversible photoreaction controlling seed germination. Proc. Natl. Acad. Sci 38:21662–666. Google Scholar

17.

J. F. Botto , R. A. Sanchez , G. C. Whitelam , and J. J. Casal . 1996. Phytochrome A mediates the promotion of seed germination by very low fluences of light and canopy shade light in Arabidopsis. Plant Physiol 110:21439–444. Google Scholar

18.

C. Bowler , G. Neuhaus , H. Yamagata , and N-H. Chua . 1994. Cyclic GMP and calcium mediate phytochrome phototransduction. Cell 77:2173–81. Google Scholar

19.

M. Boylan , N. Douglas , and P. H. Quail . 1994. Dominant negative suppression of Arabidopsis photoresponses by mutant phytochrome A sequences identifies spatially discrete regulatory domains in the photoreceptor. Plant Cell 6:21449–460. Google Scholar

20.

I. Brandstatter and J. Kieber . 1998. Two genes with similarity to bacterial response regulators are rapidly and specifically induced by cytokinin in Arabidopsis. Plant Cell 10:211009–1020. Google Scholar

21.

W. R. Briggs and M. A. Olney . 2001. Photoreceptors in Plant Photomorphogenesis to Date. Five Phytochromes, Two Cryptochromes, One Phototropin, and One Superchrome. Plant Physiol 125:2185–88. Google Scholar

22.

C. Büche , C. Poppe , E. Schäfer , and T. Kretsch . 2000. eid1: A new Arabidopsis mutant hypersensitive in phytochrome A-dependent high-irradiance responses. Plant Cell 12:21547–558. Google Scholar

23.

W. L. Butler , K. H. Norris , H. W. Siegelman , and S. B. Hendricks . 1959. Detection, assay, and preliminary purification of the pigment controlling photoresponsive development of plants. Proc. Natl. Acad. Sci. USA 45:211703–1708. Google Scholar

24.

W. L. Butler , S. B. Hendricks , and H. W. Siegelman . 1964. Action spectra of phytochrome in vitro. Photochem. Photobiol 3:21521–528. Google Scholar

25.

J. Callis and R. D. Vierstra . 2000. Protein degradation in signaling. Curr. Opin. Plant Biol 3:21381–386. Google Scholar

26.

F. R. Canton and P. H. Quail . 1999. Both phyA and phyB mediate light-imposed repression of PHYA gene expression in Arabidopsis. Plant Physiol 121:211207–1215. Google Scholar

27.

J. J. Casal , R. A. Sánchez , and J. F. Botto . 1998. Modes of action of phytochromes. J. Exp. Bot 49:21127–138. Google Scholar

28.

L. A. Castle and D. W. Meinke . 1994. A FUSCA gene of Arabidopsis encodes a novel protein essential for plant development. Plant Cell 6:2125–41. Google Scholar

29.

P. D. Cerdán , R. J. Staneloni , J. J. Casal , and R. A. Sánchez . 1997. A 146 bp fragment of the tobacco Lhcb1*2 promoter confers very-low-fluence, low-fluence and high-irradiance responses of phytochrome to a minimal CaMV 35S promoter. Plant Mol. Biol 33:21245–255. Google Scholar

30.

P. D. Cerdán , R. J. Staneloni , J. Ortega , M. M. Bunge , M. J. Rodriguez-Batiller , R. A. Sánchez , and J. J. Casal . 2000. Sustained but not transient phytochrome A signaling targets a region of an Lhcb1*2 promoter not necessary for phytochrome action. Plant Cell 12:211203–1211. Google Scholar

31.

P. D. Cerdán and J. Chory . 2003. Regulation of flowering time by light quality. Nature 423:21881–885. Google Scholar

32.

M. Chen , R. Schwab , and J. Chory . 2003. Characterization of the requirements for localization of phytochrome B to nuclear bodies. Proc. Natl. Acad. Sci. USA 100:2114493–14498. Google Scholar

33.

J. R. Cherry , D. Hondred , J. M. Walker , and R. D. Vierstra . 1992. Phytochrome requires the 6-kDa N-terminal domain for full biological activity. Proc. Natl. Acad. Sci. USA 89:215039–5043. Google Scholar

34.

K. L. Childs , F. R. Miller , M. M. Cordonnier-Pratt , L. H. Pratt , P. W. Morgan , and J. E. Mullet . 1997. The sorghum photoperiod sensitivity gene, Ma3, encodes a phytochrome B. Plant Physiol 113:21611–619. Google Scholar

35.

D-S. Cho , S-H. Hong , H-G. Nam , and M-S. Soh . 2003. FIN5 positively regulates far-red light responses in Arabidopsis Thaliana. Plant Cell Physiol 44:21565–572. Google Scholar

36.

G. Choi , H. Yi , J. Lee , Y. K. Kwon , M. S. Soh , B. Shin , S. Luka , T. R. Hahn , and P. S. Song . 1999. Phytochrome signaling is mediated through nucleoside diphosphate kinase 2. Nature 401:21610–613. Google Scholar

37.

T. Clack , S. Mathews , and R. A. Sharrock . 1994. The phytochrome apoprotein family in Arabidopsis is encoded by five genes-The sequences and expression of PHYD and PHYE. Plant Mol. Biol 25:21413–427. Google Scholar

38.

R. C. Clough , E. T. Jordan-Beebe , K. N. Lohman , J. M. Marita , J. M. Walker , C. Gatz , and R. D. Vierstra . 1999. Sequences within the N- and C-terminal domains of phytochrome A are required for PFR ubiquitination and degradation. Plant J 17:21155–167. Google Scholar

39.

A. Colón-Carmona , D. L. Chen , K-C. Yeh , and S. Abel . 2000. Aux/IAA proteins are phosphorylated by phytochrome in vitro. Plant Physiol 124:211728–1738. Google Scholar

40.

S. J. Davis , J. Kurepa , and R. D. Vierstra . 1999. The Arabidopsis thaliana HY1 locus, required for phytochrome-chromophore biosynthesis, encodes a protein related to heme oxygenases. Proc. Natl. Acad. Sci 96:216541–6546. Google Scholar

41.

S. J. Davis , A. V. Vener , and R. D. Vierstra . 1999. Bacteriophytochromes: Phytochrome-like photoreceptors from nonphotosynthetic eubacteria. Science 286:212517–2520. Google Scholar

42.

S. J. Davis , S. H. Bhoo , A. M. Durski , J. M. Walker , and R. D. Vierstra . 2001. The hemeoxygenase family required for phytochrome chromophore biosynthesis is necessary for proper photomorphogenesis in higher plants. Plant Physiol 126:21656–669. Google Scholar

43.

X. W. Deng , M. Matsui , N. Wei , D. Wagner , A. M. Chu , K. A. Feldmann , and P. H. Quail . 1992. COP1, an Arabidopsis regulatory gene, encodes a protein with both a zinc-binding motif and a Gβ homologous domain. Cell 71:21791–801. Google Scholar

44.

X-W. Deng and P. H. Quail . 1999. Signaling in light-controlled development. Semin. Cell Dev. Biol 10:21121–129. Google Scholar

45.

P. F. Devlin , K. J. Halliday , N. P. Harberd , and G. C. Whitelam . 1996. The rosette habit of Arabidopsis thaliana is dependent upon phytochrome action: novel phytochromes control internode elongation and flowering time. Plant J 10:211127–1134. Google Scholar

46.

P. F. Devlin , S. R. Patel , and G. C. Whitelam . 1998. Phytochrome E influences internode elongation and flowering time in Arabidopsis. Plant Cell 10:211479–1487. Google Scholar

47.

P. F. Devlin , P. R. Robson , S. R. Patel , L. Goosey , R. A. Sharrock , and G. C. Whitelam . 1999. Phytochrome D acts in the shade-avoidance syndrome in Arabidopsis by controlling elongation growth and flowering time. Plant Physiol 119:21909–915. Google Scholar

48.

P. F. Devlin and S. A. Kay . 2000. Cryptochromes are required for phytochrome signaling to the circadian clock but not for rhythmicity. Plant Cell 12:212499–2509. Google Scholar

49.

P. F. Devlin , M. J. Yanovsky , and S. A. Kay . 2003. A genomic analysis of the shade avoidance response in Arabidopsis. Plant Physiol 133:211617–1629. Google Scholar

50.

M. Dieterle , Y-C. Zhou , E. Schäfer , M. Funk , and T. Kretsch . 2001. EID1, an F-box protein involved in phytochrome A-specific light signaling. Genes Dev 15:21939–944. Google Scholar

51.

M. J. Dowson-Day and A. J. Millar . 1999. Circadian dysfunction causes aberrant hypocotyl elongation patterns in Arabidopsis. Plant J 17:2163–71. Google Scholar

52.

M. R. Doyle , S. J. Davis , R. M. Bastow , H. G. McWatters , L. Kozma-Bognar , F. Nagy , A. J. Millar , and R. M. Amasino . 2002. The ELF4 gene controls circadian rhythms and flowering time in Arabidopsis thaliana. Nature 419:2174–77. Google Scholar

53.

P. D. Duek and C. Fankhauser . 2003. HFR1, a putative bHLH transcription factor, mediates both phytochrome A and cryptochrome signaling. Plant J 34:21827–836. Google Scholar

54.

K. Eichenberg , I. Baurle , N. Paulo , R. A. Sharrock , W. Rudiger , and E. Schafer . 2000. Arabidopsis phytochromes C and E have different spectral characteristics from those of phytochromes A and B. FEBS Lett 470:21107–112. Google Scholar

55.

T. D. Elich and J. Chory . 1997. Biochemical characterization of Arabidopsis wild-type and mutant phytochrome B holoproteins. Plant Cell 9:212271–2280. Google Scholar

56.

E. Ermolayeva , D. Sanders , and E. Johannes . 1997. Ionic mechanism and role of phytochrome-mediated membrane depolarization in caulonemal side branch initial formation in the moss Physcomitrella patens. Planta 201:21109–118. Google Scholar

57.

C. D. Fairchild , M. A. Schumaker , and P. H. Quail . 2000. HFR1 encodes an atypical bHLH protein that acts in phytochrome A signal transduction. Genes Dev 14:212377–2391. Google Scholar

58.

C. Fankhauser 2000. Phytochromes as light-regulated protein kinases. Semin. Cell Dev. Biol 11:21467–473. Google Scholar

59.

C. Fankhauser 2001. The phytochromes, a family of red/farred absorbing photoreceptors. J. Biol. Chem 276:2111453–11456. Google Scholar

60.

C. Fankhauser and J. Chory . 2000. RSF1, an Arabidopsis locus implicated in phytochrome A signaling. Plant Physiol 124:2139–45. Google Scholar

61.

C. Fankhauser , K. C. Yeh , J. C. Lagarias , H. Zhang , T. D. Elich , and J. Chory . 1999. PKS1, a substrate phosphory-lated by phytochrome that modulates light signaling in Arabidopsis. Science 284:211539–1541. Google Scholar

62.

S. Fowler , K. Lee , H. Onouchi , A. Samach , K. Richardson , B. Morris , G. Coupland , and J. Putterill . 1999. GIGANTEA: a circadian clock-controlled gene that regulates photoperiodic flowering in Arabidopsis and encodes a protein with several possible membrane-spanning domains. EMBO J 18:214679–4688. Google Scholar

63.

K. A. Franklin , S. J. Davis , W. M. Stoddart , R. D. Vierstra , and G. C. Whitelam . 2003a. Mutant analyses define multiple roles for phytochrome C in Arabidopsis photomorphogenesis. Plant Cell 15:211981–1989. Google Scholar

64.

K. A. Franklin , U. Praekelt , W. M. Stoddart , O. E. Billingham , K. J. Halliday , and G. C. Whitelam . 2003b. Phytochromes B, D, and E act redundantly to control multiple physiological responses in Arabidopsis. Plant Physiol 131:211340–1346. Google Scholar

65.

M. Furuya and P-S. Song . 1994. Assembly and properties of holophytochrome. In Photomorphogenesis in Plants, R.E. Kendrick, and G.H.M. Kronenberg, eds, (Dordrecht, The Netherlands: Kluwer Academic Publishers), pp. 21105–140. Google Scholar

66.

W. W. Garner and H. A. Allard . 1920. Effect of the relative length of day and night and other factors of the environment on growth and reproduction in plants. J. Agric. Res 18:21553–606. Google Scholar

67.

T. Genoud , A. J. Millar , N. Nishizawa , S. A. Kay , E. Schäfer , A. Nagatani , and N. H. Chua . 1998. An Arabidopsis mutant hypersensitive to red and far-red light signals. Plant Cell 10:21889–904. Google Scholar

68.

L. Goosey , L. Palecanda , and R. A. Sharrock . 1997. Differential patterns of expression of the Arabidopsis PHYB, PHYD, and PHYE phytochrome genes. Plant Physiol 115:21959–969. Google Scholar

69.

H. Guo , T. Mockler , H. Duong , and C. Lin . 2001. SUB1, an Arabidopsis Ca2+-binding protein involved in cryptochrome and phytochrome coaction. Science 291:21487–490. Google Scholar

70.

K. J. Halliday , B. Thomas , and G. C. Whitelam . 1997. Expression of heterologous phytochromes A, B or C in transgenic tobacco plants alters vegetative development and flowering time. Plant J 12:211079–1090. Google Scholar

71.

K. J. Halliday , M. Hudson , M. Ni , M. Qin , and P. H. Quail . 1999. poc1: an Arabidopsis mutant perturbed in phytochrome signaling because of a T-DNA insertion in the promoter of PIF3, a gene encoding a phytochrome-interacting, bHLH protein. Proc. Natl. Acad. Sci. USA 96:215832–5837. Google Scholar

72.

F. Hamazato , T. Shinomura , H. Hanzawa , J. Chory , and M. Furuya . 1997. Fluence and wavelength requirements for Arabidopsis CAB gene induction by different phytochromes. Plant Physiol 115:211533–1540. Google Scholar

73.

M. Hanumappa , L. H. Pratt , M. M. Cordonnier-Pratt , and G. F. Deitzer . 1999. A photoperiod-insensitive barley line contains a light-labile phytochrome B. Plant Physiol 119:211033–1040. Google Scholar

74.

C. S. Hardtke , K. Gohda , M. T. Osterlund , T. Oyama , K. Okada , and X. W. Deng . 2000. HY5 stability and activity in Arabidopsis is regulated by phosphorylation in its COP1 binding domain. EMBO J 19:214997–5006. Google Scholar

75.

P. D. Hare , S. G. Moller , L-F. Huang , and N. H. Chua . 2003. LAF3, a novel factor required for normal phytochrome A signaling. Plant Physiol 133:211592–1604. Google Scholar

76.

S. L. Harmer , J. B. Hogenesch , M. Straume , H-S. Chang , B. Han , X. Wang , J. A. Kreps , and S. A. Kay . 2000. Orchestrated transcription of key pathways in Arabidopsis by the circadian clock. Science 290:212110–2113. Google Scholar

77.

W. Haupt and D. P. Häder . 1994. Photomovement. In Photomorphogenesis in plants, R.E. Kendrick, and G.H.M. Kronenberg, eds, (Dordrecht, The Netherlands: Kluwer Academic Publishers), pp. 21707–732. Google Scholar

78.

L. Hennig , C. Büche , K. Eichenberg , and E. Schäfer . 1999. Dynamic properties of endogenous phytochrome A in Arabidopsis seedlings. Plant Physiol 121:21571–577. Google Scholar

79.

L. Hennig , W. M. Stoddart , M. Dieterle , G. C. Whitelam , and E. Schafer . 2002. Phytochrome E controls light-induced germination of Arabidopsis. Plant Physiol 128:21194–200. Google Scholar

80.

K. A. Hicks , A. J. Millar , I. A. Carre , D. E. Somers , M. Straume , D. R. Meeks-Wagner , and S. A. Kay . 1996. Conditional circadian dysfunction of the Arabidopsis early-flowering 3 mutant. Science 274:21790–792. Google Scholar

81.

M. Hirschfeld , J. M. Tepperman , T. Clack , P. H. Quail , and R. A. Sharrock . 1998. Coordination of phytochrome levels in phyB mutants of Arabidopsis as revealed by apoprotein-specific monoclonal antibodies. Genetics 149:21523–535. Google Scholar

82.

U. Hoecker , Y. Xu , and P. H. Quail . 1998. SPA1: A new genetic locus involved in phytochrome A-specific signal transduction. Plant Cell 10:2119–33. Google Scholar

83.

U. Hoecker , J. M. Tepperman , and P. H. Quail . 1999. SPA1, a WD-repeat protein specific to phytochrome A signal transduction. Science 284:21496–499. Google Scholar

84.

U. Hoecker and P. H. Quail . 2001. The phytochrome A-specific signaling intermediate SPA1 interacts directly with COP1, a constitutive repressor of light signaling in Arabidopsis. J. Biol. Chem 276:2138173–38178. Google Scholar

85.

U. Hoecker , G. Toledo-Ortiz1 , J. Bender , and P. H. Quail . 2004. The photomorphogenesis-related mutant red1 is defective in CYP83B1, a red light-induced gene encoding a cytochrome P450 required for normal auxin homeostasis. Planta (Published online: February 12: 2004). Google Scholar

86.

M. Holm , C. S. Hardtke , R. Gaudet , and X-W. Deng . 2001. Identification of a structural motif that confers specific interaction with the WD40 repeat domain of Arabidopsis COP1. EMBO J 20:21118–127. Google Scholar

87.

M. Holm , L-G. Ma , L-J. Qu , and X. W. Deng . 2002. Two interacting bZIP proteins are direct targets of COP1-mediated control of light-dependent gene expression in Arabidopsis. Genes Dev 16:211247–1259. Google Scholar

88.

H-L. Hsieh , H. Okamoto , M. Wang , L-H. Ang , M. Matsui , H. Goodman , and X-W. Deng . 2000. FIN219, an auxin-regulated gene, defines a link between phytochrome A and the downstream regulator COP1 in light control of Arabidopsis development. Genes Dev 14:211958–1970. Google Scholar

89.

M. Hudson , C. Ringli , M. T. Boylan , and P. H. Quail . 1999. The FAR1 locus encodes a novel nuclear protein specific to phytochrome A signaling. Genes Dev 13:212017–2027. Google Scholar

90.

M. E. Hudson 2000. The genetics of phytochrome signalling in Arabidopsis. Semi. Cell Dev. Biol 11:21475–483. Google Scholar

91.

J. Hughes , T. Lamparter , F. Mittmann , E. Hartmann , W. Gärtner , A. Wilde , and T. Börner . 1997. A prokaryotic phytochrome. Nature 386:21663–664. Google Scholar

92.

J. Hughes and T. Lamparter . 1999. Prokaryotes and phytochrome. The connection to chromophores and signaling. Plant Physiol 121:211059–1068. Google Scholar

93.

E. Huq , J. M. Tepperman , and P. H. Quail . 2000a. GIGAN-TEA is a nuclear protein involved in phytochrome signaling in Arabidopsis. Proc. Natl. Acad. Sci. USA 97:219789–9794. Google Scholar

94.

E. Huq , Y. Kang , K. J. Halliday , M. Qin , and P. H. Quail . 2000b. SRL1: a new locus specific to the phyB-signaling pathway in Arabidopsis. Plant J 23:21461–470. Google Scholar

95.

E. Huq and P. H. Quail . 2002. PIF4, a phytochrome-interacting bHLH factor, functions as a negative regulator of phytochrome B signaling in Arabidopsis. EMBO J 21:212441–2450. Google Scholar

96.

E. Huq , B. Al-Sady , and P. H. Quail . 2003. Nuclear translocation of the photoreceptor phytochrome B is necessary for its biological function in seedling photomorphogenesis. Plant J 35:21660–664. Google Scholar

97.

T. Izawa , T. Oikawa , S. Tokutomi , K. Okuno , and K. Shimamoto . 2000. Phytochromes confer the photoperiodic control of flowering in rice (a short-day plant). Plant J 22:21391–399. Google Scholar

98.

T. Iwasaki , Y. Saito , E. Harada , M. Kasai , K. Shoji , M. Miyao , and N. Yamamoto . 1997. Cloning of cDNA encoding the rice 22 kda protein of photosystem II (PS-II) and analysis of light-induced expression of the gene. Gene 185:21223–229. Google Scholar

99.

S. D. Jackson , A. Heyer , J. Dietze , and S. Prat . 1996. Phytochrome B mediates the photoperiodic control of tuber formation in potato. Plant J 9:21159–166. Google Scholar

100.

J. A. Jarillo , J. Capel , R-H. Tang , H-Q. Yang , J. M. Alonso , J. R. Ecker , and A. R. Cashmore . 2001. An Arabidopsis circadian clock component interacts with both CRY1 and phyB. Nature 410:21487–490. Google Scholar

101.

Z. Jiang , L. R. Swem , B. G. Rushing , S. Devanathan , G. Tollin , and C. E. Bauer . 1999. Bacterial photoreceptor with similarity to photoactive yellow protein and plant phytochromes. Sciences 285:21406–409. Google Scholar

102.

E. Johnson , M. Bradley , N. P. Harberd , and G. C. Whitelam . 1994. Photoresponses of light-grown phyA mutants of Arabidopsis. Plant Physiol 105:21141–149. Google Scholar

103.

A. M. Jones , J. R. Ecker , and J-G. Chen . 2003. A reevaluation of the role of the heterotrimeric G protein in coupling light responses in Arabidopsis. Plant Physiol 131:211623–1627. Google Scholar

104.

E. T. Jordan , J. R. Cherry , J. M. Walker , and R. D. Vierstra . 1996. The amino-terminus of phytochrome A contains two distinct functional domains. Plant J 9:21243–257. Google Scholar

105.

E. T. Jordan , J. M. Marita , R. C. Clough , and R. D. Vierstra . 1997. Characterization of regions within the N-terminal 6-kilodalton domain of phytochrome A that modulate its biological activity. Plant Physiol 115:21693–704. Google Scholar

106.

K. A. Kaczorowski and P. H. Quail . 2003. Arabidopsis PSEUDO-RESPONSE REGULATOR7 as a signaling intermediate in phytochrome-regulated seedling deetiolation and phasing of the circadian clock. Plant Cell 15:212654–2665. Google Scholar

107.

S. A. Kay , B. Keith , K. Shinozaki , and N. H. Chua . 1989. The sequence of the rice phytochrome gene. Nucleic Acids Res 17:212865–2866. Google Scholar

108.

D. M. Kehoe and A. R. Grossman . 1996. Sensor of chromatic adaptation is similar to phytochrome and ethylene receptors. Science 273:211409–1412. Google Scholar

109.

B. Karniol , P. Malec , and D. A. Chamovitz . 1999. Arabidopsis FUSCA5 encodes a novel phosphoprotein that is a component of the COP9 complex. Plant Cell 11:21839–848. Google Scholar

110.

R. E. Kendrick and M. E. Bossen . 1987. Photocontrol of ion fluxes and membrane properties in plants. In Phytochrome and photoregulation in plants, M. Furuya, ed, (Tokyo, Japan: Academic Press), pp. 21215–224. Google Scholar

111.

R. E. Kendrick and G. H. M. Kronenberg . 1994. Photomorphogenesis in Plants. (Dordrecht, The Netherlands: Kluwer Academic Publishers). Google Scholar

112.

R. Khanna , E. A. Kikis , and P. H. Quail . 2003. EARLY FLOWERING 4 functions in phytochrome B-regulated seedling de-etiolation. Plant Physiol 133:211530–1538. Google Scholar

113.

I. S. Kim , U. Bai , and P. S. Song . 1989. A purified 124-kDa oat phytochrome does not possess a protein kinase activity. Photochem. Photobiol 49:21319–323. Google Scholar

114.

D-H. Kim , J-G. Kang , S-S. Yang , K-S. Chung , P-S. Song , and C-M. Park . 2002. A phytochrome-associated protein phosphatase 2A modulates light signals in flowering time control in Arabidopsis. Plant Cell 14:213043–3056. Google Scholar

115.

J. Kim , H. Yi , G. Choi , B. Shin , P-S. Song , and G. Choi . 2003. Functional characterization of phytochrome interacting factor 3 in phytochrome-mediated light signal transduction. Plant Cell 15:212399–2407. Google Scholar

116.

S. Kircher , L. Kozma-Bogner , L. Kim , E. Adam , K. Harter , E. Schäfer , and F. Nagy . 1999. Light quality-dependent nuclear import of the plant photoreceptors phytochrome A and B. Plant Cell 11:211445–1456. Google Scholar

117.

S. Kircher , P. Gil , L. Kozma-Bognár , E. Fejes , V. Speth , T. Husselstein-Muller , D. Bauer , E. Ádám , E. Schäfer , and F. Nagy . 2002. Nucleocytoplasmic partitioning of the plant photoreceptors phytochrome A, B, C, D, and E is regulated differentially by light and exhibits a diurnal rhythm. Plant Cell 14:211541–1555. Google Scholar

118.

T. Kohchi , K. Mukougawa , N. Frankenberg , M. Masuda , A. Yokota , and J. C. Lagarias . 2001. The Arabidopsis HY2 gene encodes phytochromobilin synthase, a ferredoxin-dependent biliverdin reductase. Plant Cell 13:21425–436. Google Scholar

119.

L. Krall and J. W. Reed . 2000. The histidine kinase-related domain participates in phytochrome B function but is dispensable. Proc. Natl. Acad. Sci. USA 97:218169–8174. Google Scholar

120.

S. F. Kwok , J. M. Staub , and X. W. Deng . 1999. Characterization of two subunits of Arabidopsis 19S proteasome regulatory complex and its possible interaction with the COP9 complex. J. Mol. Biol 285:2185–95. Google Scholar

121.

T. Kunkel , G. Neuhaus , A. Batschauer , N-H. Chua , and E. Schäfer . 1996. Functional analysis of yeast-derived phytochrome A and B phycocyanobilin adducts. Plant J 10:21625–636. Google Scholar

122.

N. Kuno , T. Muramatsu , F. Hamazato , and M. Furuya . 2000. Identification by large-scale screening of phytochrome-regulated genes in etiolated seedlings of Arabidopsis using a fluorescent differential display technique. Plant Physiol 122:2115–24. Google Scholar

123.

N. Kuno and M. Furuya . 2000. Phytochrome regulation of nuclear gene expression in plants. Semi. Cell Dev. Biol 11:21485–493. Google Scholar

124.

J. Kyozuka , D. McElroy , T. Hayakawa , Y. Xie , R. Wu , and K. Shimamoto . 1993. Light-regulated and cell-specific expression of tomato rbcS-gusA and rice rbcS-gusA fusion genes in transgenic rice. Plant Physiol 102:21991–1000. Google Scholar

125.

J. C. Lagarias and H. Rapoport . 1980. Chromopeptides from phytochrome. The structure and linkage of the Pr form of the phytochrome chromophore. J. Am. Chem. Soc 102:214821–4828. Google Scholar

126.

D. M. Lagarias , M. W. Crepeau , M. D. Maines , and J. C. Lagarias . 1997. Regulation of photomorphogenesis by expression of mammalian biliverdin reductase in transgenic Arabidopsis plants. Plant Cell 9:21675–688. Google Scholar

127.

T. Lamparter , F. Mittmann , W. Gartner , T. Borner , E. Hartmann , and J. Hughes . 1997. Characterization of recombinant phytochrome from the cyanobacterium Synechocystis. Proc. Natl. Acad. Sci. USA 94:2111792–11797. Google Scholar

128.

P. Lariguet , H. E. Boccalandro , J. M. Alonso , J. R. Ecker , J. Chory , J. J. Casal , and C. Fankhauser . 2003. A growth regulatory loop that provides homeostasis to phytochrome A signaling. Plant Cell 15:211966–1978. Google Scholar

129.

S. Laubinger and U. Hoecker . 2003. The SPA1-like proteins SPA3 and SPA4 repress photomorphogenesis in the light. Plant J 35:21373–385. Google Scholar

130.

X. L. Liu , M. F. Covington , C. Fankhauser , J. Chory , and D. Ry Wagner . 2001. ELF3 encodes a circadian clock-regulated nuclear protein that functions in an Arabidopsis PHYB signal transduction pathway. Plant Cell 13:211293–1304. Google Scholar

131.

H. Ma , M. F. Yanofsky , and E. Meyerowitz . 1990. Molecular cloning and characterization of GPA1, a G protein subunit gene of Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 87:213821–3825. Google Scholar

132.

L. Ma , J. Li , L. Qu , J. Hager , Z. Chen , H. Zhao , and X. W. Deng . 2001. Light control of Arabidopsis development entails coordinated regulation of genome expression and cellular pathways. Plant Cell 13:212589–2607. Google Scholar

133.

A. L. Mancinelli 1994. The physiology of phytochrome action. In Photomorphogenesis in Plants, R.E. Kendrick, and G.H.M. Kronenberg, eds, (Dordrecht, The Netherlands: Kluwer Academic Publishers), pp. 21211–269. Google Scholar

134.

J. F. Martinez-Garcia , E. Huq , and P. H. Quail . 2000. Direct targeting of light signals to a promoter element-bound transcription factor. Science 288:21859–863. Google Scholar

135.

P. Mas , P. F. Devlin , S. Panda , and S. A. Kay . 2000. Functional interaction of phytochrome B and cryptochrome 2. Nature 408:21207–211. Google Scholar

136.

M. G. Mason and J. Botella . 2001. Completing the heterotrimer: Isolation and characterization of an Arabidopsis thaliana G protein γ-subunit cDNA. Proc. Natl. Acad. Sci. USA 97:2114784–14788. Google Scholar

137.

S. Mathews and R. A. Sharrock . 1997. Phytochrome gene diversity. Plant Cell Environ 20:21666–671. Google Scholar

138.

T. Matsushita , N. Mochizuki , and A. Nagatani . 2003. Dimers of the N-terminal domain of phytochrome B are functional in the nucleus. Nature 424:21571–574. Google Scholar

139.

C. R. McClung 2001. Circadian rhythms in plants. Annu. Rev. Plant Physiol. Plant Mol. Biol 52:21139–162. Google Scholar

140.

T. McNellis and X. W. Deng . 1995. Light control of seedling morphogenetic pattern. Plant Cell 7:211749–1761. Google Scholar

141.

A. E. Menkens , U. Schindler , and A. R. Cashmore . 1995. The G-box: A ubiquitous regulatory DNA element in plants bound by the GBF family of bZIP proteins. TIBS 20:21506–510. Google Scholar

142.

A. J. Millar and S. A. Kay . 1996. Integration of circadian and phototransduction pathways in the network controlling CAB gene transcription in Arabidopsis. Proc. Natl. Acad. Sci. USA 93:2115491–15496. Google Scholar

143.

E. Monte , J. M. Alonso , J. R. Ecker , Y. Zhang , X. Li , J. Young , S. Austin-Phillips , and P. H. Quail . 2003. Isolation and characterization of phyC mutants in Arabidopsis reveals complex crosstalk between phytochrome signaling pathways. Plant Cell 15:211962–1980. Google Scholar

144.

T. Muramoto , T. Kohchi , A. Yokota , I. Hwang , and H. M. Goodman . 1999. The Arabidopsis photomor-phogenic mutant hy1 is deficient in phytochrome chromophore biosynthesis as a result of a mutation in a plastid heme oxygenase. Plant Cell 11:21335–348. Google Scholar

145.

A. C. Mustilli and C. Bowler . 1997. Tuning in to the signals controlling photoregulated gene expression in plants. EMBO J 16:215801–5806. Google Scholar

146.

S. G. Møller , T. Kunkel , and N-H. Chua . 2001. A plastidic ABC protein involved in intercompartmental communication of light signaling. Genes Dev 15:2190–103. Google Scholar

147.

A. Nagatani , J. W. Reed , and J. Chory . 1993. Isolation and initial characterization of Arabidopsis mutants that are deficient in phytochrome A. Plant Physiol 102:21269–277. Google Scholar

148.

F. Nagy , S. Kircher , and E. Schäfer . 2000. Nucleo-cytoplasmic partitioning of the plant photoreceptors phytochromes. Semi. Cell Dev. Biol 11:21505–510. Google Scholar

149.

M. M. Neff and E. Van Volkenburgh . 1994. Light-stimulated cotyledon expansion in Arabidopsis seedlings. Plant Physiol 104:211027–1032. Google Scholar

150.

M. M. Neff , C. Fankhauser , and J. Chory . 2000. Light: An indicator of time and place. Genes Dev 14:21257–271. Google Scholar

151.

D. C. Nelson , J. Lasswell , L. E. Rogg , M. A. Cohen , and B. Bartel . 2000. FKF1, a clock-controlled gene that regulates the transition to flowering in Arabidopsis. Cell 101:21331–340. Google Scholar

152.

G. Neuhaus , C. Bowler , K. Hiratsuka , H. Yamagata , and N-H. Chua . 1997. Phytochrome-regulated repression of gene expression requires calcium and cGMP. EMBO J 16:212554–2564. Google Scholar

153.

M. Ni , J. M. Tepperman , and P. H. Quail . 1998. PIF3, a phytochrome-interacting factor necessary for the normal photoinduced signal transduction, is a novel basic helix-loop-helix protein. Cell 95:211–20. Google Scholar

154.

M. Ni , J. M. Tepperman , and P. H. Quail . 1999. Binding of phytochrome B to its nuclear signaling partner PIF3 is reversibly induced by light. Nature 400:21781–784. Google Scholar

155.

H. Okamoto , M. Matsui , and X. W. Deng . 2001. Overexpression of the heterotrimeric G-protein α-subunit enhances phytochrome-mediated inhibition of hypocotyl elongation in Arabidopsis. Plant Cell 13:211639–1652. Google Scholar

156.

J. E. Olsen , O. Junttila , J. Nielsen , M. E. Eriksson , I. Martinussen , O. Olsson , G. Sandberg , and T. Moritz . 1997. Ectopic expression of oat phytochrome A in hybrid aspen changes critical daylength for growth and prevents cold acclimatization. Plant J 12:211339–1350. Google Scholar

157.

M. T. Osterlund and X. W. Deng . 1998. Multiple photoreceptors mediate the light-induced reduction of GUS-COP1 from Arabidopsis hypocotyl nuclei. Plant J 16:21201–208. Google Scholar

158.

M. T. Osterlund , L. H. Ang , and X. W. Deng . 1999. The role of COP1 in repression of Arabidopsis photomorphogenic development. Trends Cell Biol 9:21113–118. Google Scholar

159.

M. T. Osterlund , C. S. Hardtke , N. Wei , and X-W. Deng . 2000. Targetted destabilization of HY5 during light-regulated development of Arabidopsis. Nature 405:21462–466. Google Scholar

160.

T. Oyama , Y. Shimura , and K. Okada . 1997. The Arabidopsis HY5 gene encodes a bZIP protein that regulates stimulus-induced development of root and hypocotyl. Genes Dev 11:212983–2995. Google Scholar

161.

D. H. Park , D. E. Somers , Y. S. Kim , Y. H. Choy , H. K. Lim , M. S. Soh , H. J. Kim , S. A. Kay , and H. G. Nam . 1999. Control of circadian rhythms and photoperiodic flowering by the Arabidopsis GIGANTEA gene. Science 285:211579–1582. Google Scholar

162.

C-M. Park , S-H. Bhoo , and P-S. Song . 2000. Interdomain crosstalk in the phytochromes molecules. Semin. Cell Dev. Biol 11:21449–456. Google Scholar

163.

D. H. Park , P. O. Lim , J. S. Kim , D. S. Cho , S. H. Hong , and H. G. Nam . 2003. The Arabidopsis COG1 gene encodes a Dof domain transcription factor and negatively regulates phytochrome signaling. Plant J 34:21161–171. Google Scholar

164.

B. M. Parks and P. H. Quail . 1991. Phytochrome-deficient hy1 and hy2 long hypocotyl mutants of Arabidopsis are defective in phytochrome chromophore biosynthesis. Plant Cell 3:211177–1186. Google Scholar

165.

B. M. Parks and E. P. Spalding . 1999. Sequential and coordinated action of phytochromes A and B during Arabidopsis stem growth revealed by kinetic analysis. Proc. Natl. Acad. Sci. USA 96:2114142–14146. Google Scholar

166.

Z. Peng , G. Serino , and X-W. Deng . 2001a. A role of Arabidopsis COP9 signalosome in multifaceted developmental processes revealed by the characterization of its subunit 3. Development 128:214277–4288. Google Scholar

167.

Z. Peng , G. Serino , and X. W. Deng . 2001b. Molecular characterization of subunit 6 of the COP9 signalosome and its role in multifaceted developmental processes in Arabidopsis. Plant Cell 13:212393–2407. Google Scholar

168.

A. Pepper , T. Delaney , T. Washburn , D. Poole , and J. Chory . 1994. DET1, a negative regulator of light-mediated development and gene expression in Arabidopsis, encodes a novel nuclear localized protein. Cell 78:21109–116. Google Scholar

169.

L. H. Pratt 1994. Distribution and localization of phytochrome within the plant. In Photomorphogenesis in Plants, R.E. Kendrick, and G.H.M. Kronenberg, eds, (Dordrecht, The Netherlands: Kluwer Academic Publishers), pp. 21163–185. Google Scholar

170.

P. Puente , N. Wei , and X. W. Deng . 1996. Combinatorial interplay of promoter elements constitutes minimal determinants for light and developmental control of gene expression in Arabidopsis. EMBO J 15:213732–3743. Google Scholar

171.

J. Putterill , F. Robson , K. Lee , R. Simon , and G. Coupland . 1995. The CONSTANS gene of Arabidopsis promotes flowering and encodes a protein showing similarities to zinc finger transcription factors. Cell 80:21847–857. Google Scholar

172.

M. Qin , R. Kuhn , and P. H. Quail . 1997. Overexpressed phytochrome C has similar photosensory specificity to phytochrome B but a distinctive capacity to enhance primary leaf expansion. Plant J 12:211163–1172. Google Scholar

173.

P. H. Quail 1998. The phytochrome family: Dissection of functional roles and signaling pathways among family members. Philos. Trans. R. Soc. Lond. B Biol. Sci 353:211399–1403. Google Scholar

174.

P. H. Quail 1997. An emerging molecular map of the phytochromes. Plant Cell Environ 20:21657–665. Google Scholar

175.

P. H. Quail , M. T. Boylan , B. M. Parks , T. W. Short , Y. Xu , and D. Wagner . 1995. Phytochromes: Photosensory perception and signal transduction. Science 268:21675–680. Google Scholar

176.

P. H. Quail 2000. Phytochrome-interacting factors. Semi. Cell Dev. Biol 11:21457–466. Google Scholar

177.

P. H. Quail 2002. Phytochrome photosensory signalling networks. Nat. Rev. Mol. Cell Biol 3:2185–93. Google Scholar

178.

J. W. Reed , P. Nagpal , D. S. Poole , M. Furuya , and J. Chory . 1993. Mutations in the gene for the red/far-red light receptor phytochrome B alter cell elongation and physiological responses throughout Arabidopsis development. Plant Cell 5:21147–157. Google Scholar

179.

J. W. Reed , A. Nagatani , T. D. Elich , M. Fagan , and J. Chory . 1994. Phytochrome A and phytochrome B have overlapping but distinct functions in Arabidopsis development. Plant Physiol 104:211139–1149. Google Scholar

180.

A. Remberg , A. Ruddat , S. E. Braslavsky , W. Gartner , and K. Schaffner . 1998. Chromophore incorporation, Pr to Pfr kinetics, and Pfr thermal reversion of recombinant N-terminal fragments of phytochrome A and B chromoproteins. Biochemistry 37:219983–9990. Google Scholar

181.

P. R. H. Robson and H. Smith . 1997. Fundamental and biotechnological applications of phytochrome transgenes. Plant Cell Environ 20:21831–839. Google Scholar

182.

P. R. H. Robson , A. C. McCormac , A. S. Irvine , and H. Smith . 1996. Genetic engineering of harvest index in tobacco through overexpression of a phytochrome gene. Nature Biotechnol 14:21995–998. Google Scholar

183.

S. J. Roux 1994. Signal transduction in phytochrome responses. In Photomorphogenesis in plants, R.E. Kendrick, and G.H.M. Kronenberg, eds, (Dordrecht, The Netherlands: Kluwer Academic Publishers), pp. 21187–209. Google Scholar

184.

Y. Saijo , J. A. Sullivan , H. Wang , J. Yang , Y. Shen , V. Rubio , L. Ma , U. Hoecker , and X. W. Deng . 2003. The COP1-SPA1 interaction defines a critical step in phytochorme A-mediated regulation of HY5 activity. Genes Dev 17:212642–2647. Google Scholar

185.

K. Sakamoto and A. Nagatani . 1996. Nuclear localization activity of phytochrome B. Plant J 10:21859–868. Google Scholar

186.

M. G. Salter , K. A. Franklin , and G. C. Whitelam . 2003. Gating of the rapid shade-avoidance response by the circadian clock in plants. Nature 426:21680–683. Google Scholar

187.

R. J. H. Sawers , P. J. Linley , P. R. Farmer , N. P. Hanley , D. E. Costich , M. J. Terry , and T. P. Brutnell . 2002. elongated mesocotyl1, a phytochrome-deficient mutant of maize. Plant Physiol 130:21155–163. Google Scholar

188.

R. Schaffer , N. Ramsay , A. Samach , S. Corden , F. Putterill , I. A. Carrè , and G. Coupland . 1998. The late elongated hypocotyl mutation of Arabidopsis disrupts circadian rhythms and the photoperiodic control of flowering. Cell 93:211219–1229. Google Scholar

189.

R. Schaffer , J. Landgraf , M. Accerbi , V. Simon , M. Larson , and E. Wisman . 2001. Microarray analysis of diurnal and circadian-regulated genes in Arabidopsis. Plant Cell 13:21113–123. Google Scholar

190.

H. A. W. Schneider-Poetsch 1992. Signal transduction by phytochrome: phytochromes have a module related to the transmitter modules of bacterial sensor proteins. Photochem. Photobiol 56:21839–846. Google Scholar

191.

D. F. Schroeder , M. Gahrtz , B. B. Maxwell , R. K. Cook , J. M. Kan , J. M. Alonso , J. R. Ecker , and J. Chory . 2002. De-etiolated 1 and damaged DNA binding protein 1 interact to regulate Arabidopsis photomorphogenesis. Curr. Biol 12:211462–1472. Google Scholar

192.

C. Schwechheimer and X. W. Deng . 2000. The COP/DET/FUS proteins-regulators of eukaryotic growth and development. Semi. Cell Dev. Biol 11:21495–503. Google Scholar

193.

H. S. Seo , J-Y. Yang , M. Ishikawa , C. Bolle , M. Ballesteros , and N-H. Chua . 2003. LAF1 ubiquitination by COP1 controls photomorphogenesis and is stimulated by SPA1. Nature 423:21995–999. Google Scholar

194.

H. S. Seo , E. Watanabe , S. Tokutomi , A. Nagatani , and N-H. Chua . 2004. Photoreceptor ubiquitination by COP1 E3 ligase desensitizes phytochrome A signaling. Genes Dev 18:21617–622. Google Scholar

195.

G. Serino , T. Tsuge , S. Kwok , M. Matsui , N. Wei , and X. W. Deng . 1999. Arabidopsis cop8 and fus4 mutations define the same gene that encodes subunit 4 of the COP9 signalosome. Plant Cell 11:211967–1980. Google Scholar

196.

G. Serino , H. Su , Z. Peng , T. Tsuge , N. Wei , H. Gu , and X. W. Deng . 2002. Characterization of the last subunit of the Arabidopsis COP9 signalosome: implications for the overall structure and origin of the complex. Plant Cell 15:21719–731. Google Scholar

197.

R. A. Sharrock and P. H. Quail . 1989. Novel phytochrome sequences in Arabidopsis thaliana: Structure, evolution, and differential expression of a plant regulatory photoreceptor family. Genes Dev 3:211745–1757. Google Scholar

198.

T. Shinomura , A. Nagatani , H. Hanzawa , M. Kubota , M. Watanabe , and M. Furuya . 1996. Action spectra for phytochrome A- and B-specific photoinhibition of seed germination in Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 93:218129–8133. Google Scholar

199.

T. Shinomura , K. Uchida , and M. Furuya . 2000. Elementary processes of photoperception by phytochrome A for high-irradiance response of hypocotyl elongation in Arabidopsis. Plant Physiol 122:21147–156. Google Scholar

200.

H. Smith 1995. Physiological and ecological function within the phytochrome family. Annu. Rev. Plant Physiol. Plant Mol. Biol 46:21289–315. Google Scholar

201.

H. Smith and G. C. Whitelam . 1997. The shade avoidance syndrome: multiple responses mediated by multiple phytochromes. Plant Cell Environ 20:21840–844. Google Scholar

202.

M. S. Soh , S. H. Hong , H. Hanzawa , M. Furuya , and H. G. Nam . 1998. Genetic identification of FIN2, a far red light-specific signaling component of Arabidopsis thaliana. Plant J 16:21411–419. Google Scholar

203.

M. S. Soh , Y-M. Kim , S-J. Han , and P-S. Song . 2000. REP1, a basic helix-loop-helix protein, is required for a branch pathway of phytochrome A signaling in Arabidopsis. Plant Cell 12:212061–2073. Google Scholar

204.

D. E. Somers and P. H. Quail . 1995. Temporal and spatial expression patterns of PHYA and PHYB genes in Arabidopsis. Plant J 7:21413–427. Google Scholar

205.

D. E. Somers 1999. The physiology and molecular bases of the plant circadian clock. Plant Physiol 121:219–20. Google Scholar

206.

D. E. Somers , P. F. Devlin , and S. A. Kay . 1998. Phytochromes and cryptochromes in the entrainment of the Arabidopsis circadian clock. Science 282:211488–1490. Google Scholar

207.

D. E. Somers , T. F. Schultz , M. Milnamow , and S. A. Kay . 2000. ZEITLUPE encodes a novel clock-associated PAS protein from Arabidopsis. Cell 101:21319–329. Google Scholar

208.

D. E. Somers , W. Y. Kim , and R. Geng . 2004. The F-Box protein ZEITLUPE confers dosage-dependent control on the circadian clock, photomorphogenesis, and flowering time. Plant Cell 16:21769–782. Google Scholar

209.

E. P. Spalding 2000. Ion channels and the transduction of light signals. Plant Cell Environ 23:21665–674. Google Scholar

210.

D. Staiger , L. Allenbach , N. Salathia , V. Fiechter , S. J. Davis , A. J. Millar , J. Chory , and C. Fankhauser . 2003. The Arabidopsis SRR1 gene mediates phyB signaling and is required for normal circadian clock function. Genes Dev 17:21256–268. Google Scholar

211.

P. E. Staswick , I. Tiryaki , and M. L. Rowe . 2002. Jasmonate response locus JAR1 and several related Arabidopsis genes encode enzymes of the firefly luciferase superfamily that show activity on jasmonic, salicylic, and indole-3-acetic acids in an assay for adenylation. Plant Cell 14:211405–1415. Google Scholar

212.

J. Stockhaus , A. Nagatani , U. Halfter , S. Kay , M. Furuya , and N-H. Chua . 1992. Serine-to-alanine substitutions at the amino-terminal region of phytochrome A result in an increase in biological activity. Genes Dev 6:212364–2372. Google Scholar

213.

C. Strayer , T. Oyama , T. F. Schultz , R. Raman , D. E. Somers , P. Mas , S. Panda , J. A. Kreps , and S. A. Kay . 2000. Cloning of the Arabidopsis clock gene TOC1, an autoregulatory response regulator homolog. Science 289:21768–771. Google Scholar

214.

G. Suzuki , Y. Yanagawa , S. F. Kwok , M. Matsui , and X. W. Deng . 2002. Arabidopsis COP10 is a ubiquitin-conjugating enzyme variant that acts together with COP1 and the COP9 signalosome in repressing photomorphogenesis. Genes Dev 16:21554–559. Google Scholar

215.

U. Sweere , K. Eichenberg , J. Lohrmann , V. Mira-Rodado , I. Bäurle , J. Kudla , F. Nagy , E. Schafer , and K. Harter . 2001. Interaction of the response regulator ARR4 with phytochrome B in modulating red light signaling. Science 294:211108–1111. Google Scholar

216.

J. M. Tepperman , T. Zhu , H. S. Chang , X. Wang , and P. H. Quail . 2001. Multiple transcription-factor genes are early targets of phytochrome A signaling. Proc. Natl. Acad. Sci. USA 98:219437–9442. Google Scholar

217.

W. B. Terzaghi and A. R. Cashmore . 1995. Light-regulated transcription. Annu. Rev. Plant Physiol. Plant Mol. Biol 46:21445–474. Google Scholar

218.

M. J. Terry 1997. Phytochrome chromophore-deficient mutants. Plant Cell Environ 20:21740–745. Google Scholar

219.

A. Thiele , M. Herold , I. Lenk , P. H. Quail , and C. Gatz . 1999. Heterologous expression of Arabidopsis phytochrome B in transgenic potato influences photosynthetic performance and tuber development. Plant Physiol 120:2173–82. Google Scholar

220.

Q. Tian and J. W. Reed . 1999. Control of auxin-regulated root development by the Arabidopsis thaliana SHY2/IAA3 gene. Development 126:21711–721. Google Scholar

221.

G. Toledo-Ortiz , E. Huq , and P. H. Quail . 2003. The Arabidopsis basic/helix-loop-helix transcription factor family. Plant Cell 15:211749–1770. Google Scholar

222.

T. Tsuge , N. Iwagaki , T. Yoshizumi , H. Shimada , T. Kawamoto , R. Natsuki , N. Yamamoto , and M. Matsui . 2001. Phytochrome-mediated control of COP1 gene in rice plants. Mol. Genet. Genomics 26:2143–50. Google Scholar

223.

H. Ullah , J-G. Chen , J. C. Young , K-H. Im , M. R. Sussman , and A. M. Jones . 2001. Modulation of cell proliferation by heterotrimeric G protein in Arabidopsis. Science 292:212066–2069. Google Scholar

224.

R. D. Vierstra and S. J. Davis . 2000. Bacteriophytochromes: new tools for understanding phytochrome signal transduction. Semin. Cell Dev. Biol 11:21511–521. Google Scholar

225.

D. Vince-Prue 1994. The duration of light and photoperiodic responses. In Photomorphogenesis in plants, R.E. Kendrick, and G.H.M. Kronenberg, eds, (Dordrecht, The Netherlands: Kluwer Academic Publishers), pp. 21447–490. Google Scholar

226.

N. Viner , G. Whitelam , and H. Smith . 1988. Calcium and phytochrome control of leaf unrolling in dark-grown barley seedlings. Planta 175:21209–213. Google Scholar

227.

A. G. von Arnim and X. W. Deng . 1994. Light inactivation of Arabidopsis photomorphogenic repressor COP1 involves a cell-specific regulation of its nucleocytoplasmic partitioning. Cell 79:211035–1045. Google Scholar

228.

A. Von Arnim and X. W. Deng . 1996. A role for transcriptional repression during light control of plant development. Bioessays 18:21905–910. Google Scholar

229.

D. Wagner , C. D. Fairchild , R. M. Kuhn , and P. H. Quail . 1996. Chromophore-bearing NH2-terminal domains of phytochromes A and B determine their photosensory specificity and differential light lability. Proc. Natl. Acad. Sci. USA 93:214011–4015. Google Scholar

230.

D. Wagner , U. Hoecker , and P. H. Quail . 1997. RED1 is necessary for phytochrome B-mediated red light-specific signal transduction in Arabidopsis. Plant Cell 9:21731–743. Google Scholar

231.

H. Wang , L-G. Ma , J-M. Li , H-Y. Zhao , and X. W. Deng . 2001. Direct interaction of Arabidopsis cryptochromes with COP1 in light control development. Science 294:21154–158. Google Scholar

232.

H. Wang and X. W. Deng . 2002. Arabidopsis FHY3 defines a key phytochrome A signaling component directly interacting with its homologous partner FAR1. EMBO J 21:211339–1349. Google Scholar

233.

H. Wang , L-G. Ma , J-M. Li , H-Y. Zhao , and X. W. Deng . 2003. Analysis of far-red light regulated genome expression profiles of phytochrome A pathway mutants in Arabidopsis. The Plant J 32:21723–733. Google Scholar

234.

H. Wang and X. W. Deng . 2003. Dissecting phytochrome A dependent signaling network in higher plants. Trends in Plant Sci 8:21172–178. Google Scholar

235.

Z-Y. Wang and E. M. Tobin . 1998. Constitutive expression of the CIRCADIAN CLOCK ASSOCIATED 1 (CCA1) gene disrupts circadian rhythms and suppresses its own expression. Cell 93:211207–1217. Google Scholar

236.

Z-Y. Wang , D. Kenigbuch , L. Sun , E. Harel , M. S. Ong , and E. M. Tobin . 1997. A myb-related transcription factor is involved in the phytochrome regulation of an Arabidopsis Lhcb gene. Plant Cell 9:21491–507. Google Scholar

237.

N. Wei , D. A. Chamovitz , and X. W. Deng . 1994. Arabidopsis COP9 is a component of a novel signaling complex mediating light control of development. Cell 78:21117–124. Google Scholar

238.

N. Wei and X. W. Deng . 1996. The role of the COP/DET/FUS genes in light control of Arabidopsis seedling development. Plant Physiol 112:21871–878. Google Scholar

239.

C. A. Weiss , C. W. Garnaat , K. Mukai , Y. Hu , and H. Ma . 1994. Isolation of cDNAs encoding guanine nucleotide-binding protein β-subunit homologues from maize (ZGB1) and Arabidopsis (AGB1). Proc. Natl. Acad. Sci. USA 91:219554–9558. Google Scholar

240.

G. C. Whitelam and P. F. Devlin . 1997. Roles of different phytochromes in Arabidopsis photomorphogenesis. Plant Cell Environ 20:21752–758. Google Scholar

241.

G. C. Whitelam , E. Johnson , J. Peng , P. Carol , M. L. Anderson , J. S. Cowl , and N. P. Harberd . 1993. Phytochrome A null mutants of Arabidopsis display a wild-type phenotype in white light. Plant Cell 5:21757–768. Google Scholar

242.

Y. S. Wong , H. C. Cheng , D. A. Walsh , and J. C. Lagarias . 1986. Phosphorylation of Avena phytochrome in vitro as a probe of light-induced conformational changes. J. Biol. Chem 261:2112089–12097. Google Scholar

243.

S. J. Wu and J. C. Lagarias . 2000. Defining the bilin lyase domain: lessens from the extended phytochrome superfamily. Biochemistry 39:2113487–13495. Google Scholar

244.

R. Yamaguchi , M. Nakamura , N. Mochizuki , S. A. Kay , and A. Nagatani . 1999. Light-dependent translocation of a phytochrome B-GFP fusion protein to the nucleus in transgenic Arabidopsis. J. Cell Biol 3:21437–445. Google Scholar

245.

H-Q. Yang , R-H. Tang , and A. R. Cashmore . 2001. The signaling mechanism of Arabidopsis CRY1 involves direct interaction with COP1. Plant Cell 13:212573–2587. Google Scholar

246.

M. Yano , S. Kojima , Y. Takahashi , H. Lin , and T. Sasaki . 2001. Genetic control of flowering time in rice, a short-day plant. Plant Physiol 127:211425–1429. Google Scholar

247.

M. J. Yanovsky , J. J. Casal , and J. P. Luppi . 1997. The VLF loci, polymorphic between ecotypes Landsberg erecta and Columbia, dissect two branches of phytochrome A signal transduction that correspond to very-low-fluence and high-irradiance responses. Plant J 12:21659–667. Google Scholar

248.

M. J. Yanovsky , G. C. Whitelam , and J. J. Casal . 2000. fhy3-1 retains inductive responses of phytochrome A. Plant Physiol 123:21235–242. Google Scholar

249.

K. C. Yeh and J. C. Lagarias . 1998. Eukaryotic phytochromes: Light-regulated serine/threonine protein kinases with histidine kinase ancestry. Proc. Natl. Acad. Sci. USA 95:2113976–13981. Google Scholar

250.

K-C. Yeh , S-H. Wu , J. T. Murphy , and J. C. Lagarias . 1997. A cyanobacterial phytochrome two-component light sensory system. Science 277:211505–1508. Google Scholar

Figure 1.

The contrasting phenotypes of dark- vs. light-grown Arabidopsis seedlings. Dark-grown seedlings undergo a skotomorphogenic development program (etiolation), which is characterized by elongated hypocotyls, closed cotyledons and apical hooks. Light-grown seedlings undergo photomorphogenesis and are characterized by short hypocotyls, open and expanded green cotyledons.

i1543-8120-18-1-1-f01.jpg

Figure 2.

Absorption spectra of phytochromes. Absorption spectra of the two forms (Pr and Pfr) of phytochromes. The Pr form absorbs maximally at 660 nm, while the Pfr form absorbs maximally at 730 nm.

i1543-8120-18-1-1-f02.jpg

Figure 3.

The Arabidopsis phytochrome family and the domain structure map of a generic phytochrome molecule. (A) The phylogenetic distance tree of the five phytochrome species from Arabidopsis thaliana. PHYB and PHYD share ∼80% amino acid sequence identity, and constitute a branch of the gene family. PHYE itself, PHYA and PHYC form two other branches of the family evolution tree (adapted from Clack et al., 1994).(B) The domain structure map of a generic phytochrome molecule. The coordinates indicate positions of the consensus sequence derived from the alignment of multiple full-length phytochrome polypeptide sequences by Mathews et al. (1995). The N-terminal photosensory domain (CBD, for chromophore binding domain) and the C-terminal regulatory domain are joined by a flexible hinge region (H). The chromophore binding site (C374) is located in the N-terminal photosensory domain. The C-terminal domain contains several conserved motifs, including the regulatory core sequence (Quail box), two dimerization motifs (D1 and D2), two PAS domains (P1 and P2), and the histidine kinase-related domain (HKRD). The positions for the junction site and these individual motifs are indicated (adapted from Quail, 1997).

i1543-8120-18-1-1-f03.gif

Figure 4.

Arabidopsis phytochrome chromophore. (A) The biosynthesis pathway of Arabidopsis phytochrome chromophore (adapted from Kohchi et al., 2001).(B) Chemical structures of heme, BV, PΦB, PCB, and PEB. Heme oxygenase converts heme to BV by an oxidative cleavage between rings A and D at the position marked (arrow).(C) Red light (R) triggers a “Z” to “E” isomerization in the C-15 double bond between the C and D rings of the linear tetrapyrrole, which is accompanied by rearrangement of the apoprotein backbone. This results in the photoconversion of phytochromes from the Pr form to the Pfr form. Far-red (FR) light converts the Pfr form back to the Pr form.

i1543-8120-18-1-1-f04.gif

Figure 5.

Arabidopsis phytochromes function as light-regulated kinases. Domain comparison map between the bacterial phytochrome Cph1 and Arabidopsis phytochromes. The conserved cysteine residue for chromophore binding is indicated (*). CBD: chromophore-binding domain; HKD: histidine kinase domain; PRD: PAS related domain; HKRD: histidine kinase related domain. PIFs: phytochrome-interacting factors. (H) highlights the conserved histidine on the HKD domain of Cph1. The percent amino acid identities between the HKD of Cph1 and both PRD and HKRD of Arabidopsis phytochromes are indicated. Also note that Arabidopsis phytochromes have distinct N-terminal extensions (NTE). Autophosphorylation of phytochromes could affect their stability (in case of phyA), their subcellular localization, and their abilities to interact with PIFs. Alternatively, phytochromes could phosphorylate PIFs and modulate their activities. The amino acid numbers correspond to the oat phyA (adapted from Neff et al., 2000).

i1543-8120-18-1-1-f05.gif

Figure 6.

A summary of identified phytochrome-interacting partners (PIFs). Some PIFs interact with both phyA and phyB, whereas others specifically interact with either phyA only or phyB only. Note that most of these physical interactions were detected with the yeast two-hybrid assay and/or in vitro binding assay.

i1543-8120-18-1-1-f06.gif

Figure 7.

A simplified genetic model for phytochrome-mediated signaling pathways. phyA and phyB signaling entails specific as well as shared components. These loci presumably act upstream of the COP/DET/FUS genes, thus controlling HY5, HYH, LAF1 and the degree of photomorphogenic development. Genes have been characterized at the molecular level are capitalized. The mutant loci have not been cloned are lower case and italicized. Negative regulators of the signaling pathways are underlined. Arrows indicate a positive action, and the bars indicate a repressive effect.

i1543-8120-18-1-1-f07.gif

Figure 8.

The phenotype of cop1 (constitutive photomorphogenic) mutants and the proposed roles of the COP/DET/FUS proteins in photomorphogenesis. (A) Dark-grown cop1 mutant seedlings phenotypically mimic light-grown wild-type seedlings.(B) A total of eleven pleiotropic COP/DET/FUS loci function as repressors of photomorphogenesis. Light signals perceived by multiple photoreceptors are transduced to inactivate these COP/DET/FUS proteins, and to turn on photomorphogenic development.

i1543-8120-18-1-1-f08.jpg

Figure 9.

The cellular basis and proposed biochemical mode of COP1 function. (A) A model depicting COP1 as an E3 ubiquitin ligase. COP1 mediates the ubiquitination of HY5 and its subsequent degradation via the proteasome by recruiting an E2 and HY5 through distinct interacting domains (RING-finger and WD-40 repeat domain, respectively, adapted from Osterlund et al., 2000).(B) Light-regulated nucleocytoplasmic partitioning of COP1. In darkness, COP1 is enriched in the nucleus to suppress photomorphogenic development. Light signals (perceived by phytochromes and other photoreceptors) trigger the nuclear depletion of COP1, thus abrogating the repressive effect of COP1. Note that the COP9 signalosome may contribute to the control of nuclear localization of COP1 or the stability of COP1 in the nucleus. The roles of DET1 and COP10 in this cellular process have not been determined.

i1543-8120-18-1-1-f09.gif

Figure 10.

A molecular model depicting phyA control of gene expression and photomorphogenesis. Upon photoconversion to the Pfr form, phyA translocates into the nucleus, where it regulates the transcription activities of a group of early phytochrome-responsive transcription factors (such as the phyA interactive PIF3 and other genetically defined signaling components HFR1, HY5, LAF1, and COG1 etc.). Many of the encoded early target gene products are transcriptional regulators such as CCA1 and LHY, and they are responsible for orchestrating the expression of somewhat more downstream target genes, thus generating a transcriptional network required for various cellular responses constituting photomorphogenesis. Other nuclear localized phyA signaling intermediates (such as FHY1, FHY3, FAR1 and PRR7) could act as transcriptional co-activators or co-repressors for the various transcription factors. COP1, the COP9 signalosome (CSN), COP10, DET1 and the phyA signaling specific repressors SPA1 and EID1 could interact (directly or indirectly) with these transcription factors and other signaling intermediates and regulate their abundance through a light-regulated proteolysis process. Those cytoplasmically localized phyA signaling molecules, including LAF6, PAT1, FIN219, PKS1, SUB1 and LAF3, could also participate in phyA regulation of gene expression through different mechanisms. Some possibilities include nuclear translocation of phyA and other phyA signaling molecules (such as LAF3 and PKS1), membrane permeability, ion flux and plastid to nucleus signaling (such as LAF6). Activated phyA and subsequent phyA signaling events inactivate COP/DET/FUS gene activities, and abrogate their repressive effect. Conversely, COP/DET/FUS proteins could attenuate phyA signaling through targeted degradation of the phyA photoreceptor itself. R: red light; FR: far-red light; Pr: inactive, Pr conformer of phytochromes; Pfr: active, Pfr conformer of phytochromes; TXN: transcription factor. Arrows indicate a positive action, and the bars indicate a repressive effect. The “?” marks denote undemonstrated mechanisms.

i1543-8120-18-1-1-f10.gif

Table 1.

Diagnosis Features of Different Phytochrome Action Modes

i1543-8120-18-1-1-t01.gif

Table 2.

Molecularly Identified Arabidopsis Genes Cited in This Chapter

i1543-8120-18-1-1-t02.gif

Table 3.

Different Roles of Phytochrome Family Members in Seedling and Early Vegetative Development

i1543-8120-18-1-1-t03.gif

[1] * This chapter was updated on July 6, 2004. The original version has been archived and is available at  http://dx.doi.org/10.1199/tab.0074

The American Society of Plant Biologists
Haiyang Wang and Xing Wang Deng "Phytochrome Signaling Mechanism," The Arabidopsis Book 2004(3), (4 April 2002). https://doi.org/10.1199/tab.0074.1
Published: 4 April 2002
Back to Top