Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

The ATP-Mediated Regulation of KaiB-KaiC Interaction in the Cyanobacterial Circadian Clock

  • Risa Mutoh ,

    Contributed equally to this work with: Risa Mutoh, Atsuhito Nishimura

    Current address: Institute for Protein Research, Osaka University, Yamadaoka, Suita, Osaka, Japan

    Affiliation Center for Gene Research, Nagoya University, Nagoya, Aichi, Japan

  • Atsuhito Nishimura ,

    Contributed equally to this work with: Risa Mutoh, Atsuhito Nishimura

    Affiliations Center for Gene Research, Nagoya University, Nagoya, Aichi, Japan, Division of Biological Science, Graduate School of Science, Nagoya University, Nagoya, Aichi, Japan

  • So Yasui,

    Affiliations Center for Gene Research, Nagoya University, Nagoya, Aichi, Japan, Division of Biological Science, Graduate School of Science, Nagoya University, Nagoya, Aichi, Japan

  • Kiyoshi Onai,

    Affiliation Center for Gene Research, Nagoya University, Nagoya, Aichi, Japan

  • Masahiro Ishiura

    ishiura@gene.nagoya-u.ac.jp

    Affiliations Center for Gene Research, Nagoya University, Nagoya, Aichi, Japan, Division of Biological Science, Graduate School of Science, Nagoya University, Nagoya, Aichi, Japan

Abstract

The cyanobacterial circadian clock oscillator is composed of three clock proteins—KaiA, KaiB, and KaiC, and interactions among the three Kai proteins generate clock oscillation in vitro. However, the regulation of these interactions remains to be solved. Here, we demonstrated that ATP regulates formation of the KaiB-KaiC complex. In the absence of ATP, KaiC was monomeric (KaiC1mer) and formed a complex with KaiB. The addition of ATP plus Mg2+ (Mg-ATP), but not that of ATP only, to the KaiB-KaiC1mer complex induced the hexamerization of KaiC and the concomitant release of KaiB from the KaiB-KaiC1mer complex, indicating that Mg-ATP and KaiB compete each other for KaiC. In the presence of ATP and Mg2+ (Mg-ATP), KaiC became a homohexameric ATPase (KaiC6mer) with bound Mg-ATP and formed a complex with KaiB, but KaiC hexamerized by unhydrolyzable substrates such as ATP and Mg-ATP analogs, did not. A KaiC N-terminal domain protein, but not its C-terminal one, formed a complex with KaiB, indicating that KaiC associates with KaiB via its N-terminal domain. A mutant KaiC6mer lacking N-terminal ATPase activity did not form a complex with KaiB whereas a mutant lacking C-terminal ATPase activity did. Thus, the N-terminal domain of KaiC is responsible for formation of the KaiB-KaiC complex, and the hydrolysis of the ATP bound to N-terminal ATPase motifs on KaiC6mer is required for formation of the KaiB-KaiC6mer complex. KaiC6mer that had been hexamerized with ADP plus aluminum fluoride, which are considered to mimic ADP-Pi state, formed a complex with KaiB, suggesting that KaiB is able to associate with KaiC6mer with bound ADP-Pi.

Introduction

Circadian rhythms—oscillations that regulate metabolic and behavioral activity in approximately 24-h periods—are observed in almost all organisms from prokaryotes to eukaryotes. Cyanobacteria are the simplest organisms to exhibit circadian rhythms [1]. The gene cluster kaiABC that is essential for the generation of circadian rhythms has been cloned and analyzed in the cyanobacterium Synechococcus sp. strain PCC 7942 (hereafter Synechococcus) [2]. The cyanobacterial clock oscillator is composed of only three clock proteins, KaiA, KaiB, and KaiC. Circadian oscillations are generated at the phosphorylation level [3] and ATPase activity [4] of KaiC and the formation of complexes between the proteins [5,6]. KaiC is a homohexameric ATP-binding protein with autokinase activity [7-9], very weak temperature-independent (temperature-compensated) ATPase activity [4,10], and autophosphatase activity [11,12]. The KaiC subunit has a duplicated structure composed of N-terminal and C-terminal domains, and each has a series of ATPase motifs (a Walker's motif A, a Walker's motif B, and a catalytic glutamate (CatE; We named the N-terminal CatE and C-terminal CatE as CatE1 and CatE2, respectively.)) (Figure 1A) [2,13]. KaiC has two phosphorylation sites, Ser431 and Thr432, in its C-terminal domain (Figure 1A). A mutant KaiC with aspartate substitutions at the two phosphorylation sites, KaiCS431D&T432D (KaiCDD), is considered to mimic the fully phosphorylated state of KaiC [14], and a mutant KaiC with alanine substitutions at the two phosphorylation sites, KaiCS431A&T432A (KaiCAA), is considered to mimic fully unphosphorylated state of KaiC [15]. KaiCDD showed higher affinity with KaiB than KaiCAA [5,15-18]. KaiA is a homodimeric protein [9,19,20] that enhances the phosphorylation level [9,21,22] and ATPase activity [4,10] of KaiC. KaiB is a homotetrameric protein [23-25], and it has a positively charged cleft flanked by two negatively charged ridges [24]. KaiB suppresses the autokinase activity [22,26] and ATPase activity [4] of KaiC. KaiC associates with KaiA [8,9] and KaiB [25,27], and KaiB also associates directly with KaiA [16,28,29]. These interactions are essential for the generation of clock oscillation. On interaction with hexameric KaiC (KaiC6mer), tetrameric KaiB (KaiB4mer) likely dissociates into two dimeric KaiB (KaiB2mer) and forms a complex comprising one molecule of KaiC6mer and two of KaiB2mer [17]. Previously, we have shown that KaiB1-94-KaiCDD6mer complex (350 ± 20 kDa) consists of two molecules of KaiB1-94 (dimeric) and one molecule of KaiCDD6mer and also have shown data suggesting that KaiBWT-KaiCDD6mer complex (366 ± 20 kDa) consists of two molecules of dimerized KaiBWT (KaiBWT2mer) and one molecule of KaiCDD6mer [17]. The histidine kinase SasA, which is the first component of the main output pathway that transduces clock oscillation to genome-wide transcription cycles [30], competes with KaiB to form a complex with KaiC [17], and SasA and KaiC associate through their N-terminal domains [14]. How the KaiB-KaiC interaction is regulated, however, remains unknown.

thumbnail
Figure 1. The sequence motifs of KaiC and KaiB-KaiC complex formation.

A. A diagram of the sequence motifs of KaiC. B. Time courses of KaiB-KaiCDD6mer complex formation. KaiBWT (1 μM) and KaiB1-94 (2 μM) were separately incubated with 1 μM KaiCDD6mer in reaction buffer containing Mg-ATP at 4 °C or 40 °C for the various periods indicated, and then aliquots of the reaction mixtures were subjected to native PAGE on 10 % gels. After staining the gels with CBB to visualize proteins, we estimated the amounts of KaiBs-KaiCDD6mer complex by densitometry. We used the maximum value obtained at a time of 6 h as the maximum value. KaiBs added and temperature conditions: open circles, KaiBWT (40 °C); closed circles, KaiB1-94 (40 °C); open squares, KaiB1-94 (4 °C). C. A native PAGE gel of the reaction products of KaiCDD1mer with KaiB1-94. Reaction mixtures containing 15 μM KaiB1-94 and 5 μM KaiCDD1mer in reaction buffer were incubated at 4 °C for 6 h and then subjected to native PAGE. Proteins were visualized by CBB staining of the gel. Because KaiB1-94 has an isoelectric point of 9.7, it moved in the opposite direction and could not detect by native PAGE. D. A native PAGE gel of the reaction products of KaiCDD1mer with KaiBWT. Reaction mixtures containing 7.5 μM KaiBWT and 5 μM KaiCDD1mer in reaction buffer were incubated at 4 °C for 6 h and then subjected to native-PAGE. E. Time course of KaiB1-94-KaiCDD1mer complex formation. Reaction mixtures containing 5 μM KaiB1-94 and 5 μM KaiCDD1mer in reaction buffer were incubated at 4 °C for the periods indicated and then subjected to native PAGE. We used the value obtained at a time of 9 h as the maximum value. A typical experimental data is shown. F. KaiB1-94 concentration-dependence of the KaiB1-94-KaiCDD1mer complex formation. Reaction mixtures containing 1 μM KaiCDD1mer and various amounts (0.5, 0.75, 1.0, 1.25, 1.5, 2.0, 2.5, 3.0, and 4.0 μM) of KaiB1-94 in reaction buffer were incubated at 4 °C for 6 h and then subjected to native PAGE. We used the value obtained at a KaiB1-94 concentration of 4 μM as the maximum value. A typical experimental data is shown.

https://doi.org/10.1371/journal.pone.0080200.g001

Here, we investigated the interaction of KaiC with KaiB in vitro and found that KaiC associates with KaiB via its N-terminal domain and that ATP regulates the KaiB-KaiC interaction. We propose a model for ATP regulation of KaiB-KaiC interaction.

Materials and Methods

Preparation of Kai protein

We produced recombinant Kai proteins derived from the thermophylic cyanobacterium Thermocynechococcus elongatus BP-1 using Escherichia coli as a host, as described previously, with modifications [13]. The plasmids expressing KaiA whose subunit consists of 283 amino acid residues, wild-type KaiB (KaiBWT; pTekaiBWT), whose subunit consists of 108 amino acid residues, a mutant KaiB with a C-terminal deletion of residues 95 to 108 (KaiB1-94; pTekaiB1-94), wild-type KaiC (KaiCWT; pTekaiCWT), whose subunit consists of 518 amino acid residues, and mutant KaiCs— a mutant KaiC with aspartate substitutions at the two phosphorylation sites that is considered to mimic the fully phosphorylated state of KaiC, KaiCS431D&T432D (KaiCDD; pTekaiCDD), a mutant KaiC with alanine substitutions at the two phosphorylation sites that is considered to mimic fully unphosphorylated state of KaiC, KaiCS431A&T432A (KaiCAA; pTekaiCAA), a mutant KaiC with glutamine substitutions at the two deduced catalytic glutamate residues of the C-terminal ATPase motifs that lacks the C-terminal ATPase activity of KaiC, KaiCCatE2- (pTekaiCCatE2-), KaiCCatE2-/S431A&S432A (KaiCCatE2-/AA; pTekaiCCatE2-/AA), an N-terminal domain mutant protein of residues 1 to 268 (KaiCN; pTekaiCN), and a C-terminal domain mutant protein of residues 269 to 518 with a mutation with aspartate substitutions at the two KaiC phosphorylation sites (KaiCC/DD; pTekaiCC/DD)—have been described previously [8,28]. We constructed plasmids for the production of KaiCK53H/ DD (pTekaiCK53H/DD), which is deficient in the N-terminal ATP binding site (N-terminal ATPase motifs) and carries aspartate substitutions at the two phosphorylation sites on the C-terminal domain (DD mutation), and KaiCCatE1-/DD (pTekaiCCatE1-/DD) , which is deficient in the N-terminal ATPase and carries the DD mutation, by replacing a 1.25-kb BamHI-EcoRI fragment carrying the 3'-region of each kaiC gene in pTekaiCK53H and pTekaiCCatE1-, respectively, with that of pTekaiCDD. We constructed a plasmid for KaiCK294H/DD (pTekaiCK294H/DD), which is deficient in the N-terminal ATP binding site and carries the DD mutation as descried previously [13]. We purified KaiA, KaiB, and KaiC as described previously [28] and stored them at -85 °C until used. KaiA is a dimer [9,19,20], KaiBWT tetramer [23-25], and KaiB1-94 dimer [17]. Thus unless otherwise stated, we expressed the concentrations of KaiA, KaiBWT, KaiB1-94, KaiC1mer, and KaiC6mer as those of a dimer, tetramer, dimer, monomer, and hexamer, respectively.

Preparation of KaiC6mer

To allow the hexamerization of KaiC, we incubated KaiC1mer with 1 mM ATP or 1 mM ATP analogs, 5’-adenylylimidodiphosphate (AMPPNP) and adenosine 5'-O-(3-thio) triphosphate (ATPγS), in the presence of 5 mM MgCl2 in 20 mM Tris-HCl buffer (pH 7.5) containing 150 mM NaCl at 20 °C for 20 min. We then subjected the reaction mixtures to gel filtration chromatography on a Superdex 200/HR 10/30 column (GE Healthcare) equilibrated with 20 mM HEPES-NaOH buffer (pH 7.5) containing 150 mM NaCl (reaction buffer), 0.1 mM ATP (or ATP analogs) and 5 mM MgCl2 at 4 °C, and we collected fractions containing KaiC6mer. We similarly prepared KaiC6mer in the absence of MgCl2. Unless otherwise stated, KaiC6mer means KaiC hexamerized with 1 mM ATP plus 5 mM MgCl2 (Mg-ATP). The phosphorylation level of the KaiCWT preparations used here was about 30 %.

Assay for the complex formation of monomeric KaiC (KaiC1mer) with KaiBs by native polyacrylamide gel electrophoresis (native PAGE)

We incubated reaction mixtures containing 15 μM KaiB1-94 or 7.5 μM KaiBWT and 5 μM KaiC1mer or 2.5 μM KaiC6mer at 4 °C in reaction buffer and subjected 20-μl aliquots to native PAGE on 10 % gels (acrylamide: bisacrylamide = 37.5: 1) and stained the gels with Coomassie Brilliant Blue (CBB). We estimated the amount of KaiB-KaiC1mer complex by densitometry using a Lane Analyzer (ATTO, Tokyo, Japan) and a CS Analyzer (ATTO).

Assay for formation of the KaiB1-94-KaiC6mer complex by native-PAGE and 2-dimensional sodium dodecyl sulfate (SDS)-PAGE

We incubated reaction mixtures containing 10 μM KaiB1-94 and 6 μM KaiC6mer at 4, 25, or 40 °C in reaction buffer and subjected aliquots to native PAGE. We confirmed the presence of KaiB1-94 and KaiC in the complex bands by native PAGE followed by SDS-PAGE on 18 % gels, as described previously [8].

Assay for formation of the KaiB1-94-KaiC1mer complex and KaiB1-94-KaiC6mer complex by gel filtration chromatography

To assay for formation of the KaiB1-94-KaiC1mer complex and the KaiB1-94-KaiC6mer complex, we incubated reaction mixtures containing 36 μM KaiB1-94 and 12 μM KaiC1mer in reaction buffer and those containing 7.5 μM KaiB1-94 and 2.5 μM KaiC6mer (KaiCWT or mutant KaiCs) in the buffer with (for KaiC6mer) or without (for KaiC1mer) Mg-ATP at 4 °C for 6 h. We then analyzed the mixtures by gel filtration chromatography on a Superdex 200/HR 10/30 column equilibrated with reaction buffer with (for KaiC6mer) or without (for KaiC1mer) 0.1 mM ATP and 5 mM MgCl2 at 4 °C. We subjected all peak fractions to SDS-PAGE and visualized by staining with CBB.

Immunoblot analysis

Because we could not easily detect the KaiB1-94 contained in a putative complex between KaiB1-94-KaiCN6mer by CBB staining of SDS-PAGE gels, we subjected the gels to immunoblotting. We incubated reaction mixtures containing KaiB1-94 and/or KaiCN6mer at 4 °C for 6 h as described above, and then subjected them to gel filtration chromatography on a Superdex 200/HR 10/30 column equilibrated with reaction buffer containing 0.5 mM ATP and 5 mM MgCl2 at 4 °C. We subjected fractions containing a putative KaiB1-94-KaiCN6mer complex to SDS-PAGE, blotted the proteins onto Immobilon-P Transfer Membrane (Millipore), and visualized them using the ECL Western Blotting Analysis System (GE Healthcare) with a rabbit anti-KaiB antiserum (diluted to 1/2000) as a primary antibody and a donkey anti-rabbit Ig antibody (GE Healthcare) as a secondary antibody, as described previously [8].

Assay for the time course of formation of the KaiB1-94-KaiCDD1mer complex

We incubated KaiCDD1mer (5 μM) with 5 μM KaiB1-94 in reaction buffer at 4 °C for various periods and subjected aliquots of the reaction mixtures to native PAGE on 10 % gels, staining the gels with CBB as described above. We estimated the amount of KaiB1-94-KaiCDD1mer complex from the intensity of the band by densitometry as described above.

KaiB1-94-concentration dependence of KaiB-KaiC1mer complex formation

We incubated 1 μM KaiCDD1mer with various amounts of KaiB1-94 (0.5, 0.75, 1.0, 1.25, 1.5, 2.0, 2.5, 3.0, and 4.0 μM) in reaction buffer at 4 °C for 6 h. Other conditions were the same as described above.

Assay for Mg-ATP-induced dissociation of the KaiB1-94-KaiC1mer complex

We assayed the ATP-induced hexamerization of KaiCDD1mer in the KaiB1-94-KaiC1mer complex and the concomitant release of KaiB1-94. To obtain the KaiB1-94-KaiCDD1mer complex, we incubated reaction mixtures containing 24 μM KaiB1-94 and 24 μM KaiCDD1mer in reaction buffer at 4 °C for 16 h and then subjected the mixtures to gel filtration chromatography on a Superdex 75/HR 10/30 column (GE Healthcare) equilibrated with reaction buffer at 4 °C. We added 1 mM ATP with or without 5 mM MgCl2 to the obtained KaiB1-94-KaiCDD1mer complex and then incubated the reaction mixtures in reaction buffer at 4 °C for 6 h. Then, we subjected the reaction mixtures to gel filtration chromatography on a Superdex 200/HR 10/30 column equilibrated with reaction buffer at 4 °C. We analyzed all peak fractions by SDS-PAGE on 18 % gels, and then stained the gels with CBB.

Preparation of KaiC6mer formed with ADP and aluminum fluoride (KaiC6mer (ADP-AlFX)) and assay for the KaiB1-94-KaiC6mer (ADP-AlFX) complex

We incubated KaiC1mer (20 μM) in reaction buffer containing 6 mM ADP, 30 mM MgCl2, 2.5 mM NaF, and 2.5 mM AlCl3 at 25 °C for 2 h to form KaiC6mer (ADP-AlFX) [31]. We then mixed the reacion mixtures with an equal volume of 0 or 60 μM KaiB1-94 in the same buffer and incubated them further at 25 °C for 6 h. We subjected the reaction mixtures to gel filtration chromatography on a Superdex 200/HR 10/30 column equilibrated with reaction buffer at 4 °C. We subjected peak fractions to SDS-PAGE and visualized the proteins by staining with CBB.

Assay for the KaiA-enhanced autophosphorylation of KaiC6mer

We incubated reaction mixtures containing 0.5 μM KaiC6mer, 0.5 μM KaiA, and Mg-ATP in reaction buffer at 40 °C for various periods. We then subjected aliquots of the reaction mixtures to SDS-PAGE on 12.5 % gels (acrylamide: bisacrylamide = 144: 1), and stained the gels with CBB. We estimated the amount of the protein from the intensities of bands by densitometry as described above.

Assay for ATPase activity

We measured the ATPase activity of KaiC6mer using BIOMOL GREEN (Enzo Life Science International, Inc., Farmingdale, New York, USA) according to the supplier's instruction manual. We incubated KaiC6mer (1 μM) in reaction buffer containing Mg-ATP at 4 °C for 6 h, and then incubated 50-μl aliquots of the reaction mixtures with 50 μl BIOMOL GREEN at 25 °C for 20 min. We measured the absorbance of the samples at 620 nm (A620) using an ARVO X4 plate reader (PerkinElmer Inc., Waltham, MA, USA), calculated the amount of Pi released from ATP from triplicate experiments, and expressed the ATPase activity as mol Pi released per mol KaiC6mer per h at 4 °C. We showed the values after subtracting the buffer background (0.024 ± 0.03 ATP molecules/h).

Results

Association of KaiC1mer with KaiB1-94

At 40 °C, KaiBWT formed a complex with KaiCDD6mer (Figure 1B) but the complex formation took more than 9 h to reach a plateau (Figure 1B). KaiB1-94-KaiCDD6mer complex formation, on the other hand, reached a plateau within 6 h with a time (t1/2) of 1.2 h where the half maximal KaiB1-94-KaiC6mer complex formation occurred, and even at 4 °C showed no lag (Figure 1B). Thus, to analyze the formation of the KaiB-KaiC complex in detail, we used KaiB1-94 and KaiCDD at 4 °C.

First, we examined KaiCDD1mer complex formation with KaiB1-94 and KaiBWT. When we incubated 5 μM KaiCDD1mer with 15 μM KaiB1-94 or 7.5 μM KaiBWT at 4 °C for 6 h, KaiCDD1mer formed a complex with KaiB1-94 (Figure 1C and Table S1) as KaiCDD6mer (Figure 1B), whereas it showed only weak complex formation with KaiBWT (Figure 1D). The complex formation occurred without delay and reached a plateau at 6 h (Figure 1E). The time (t1/2) where the half maximal KaiB1-94-KaiC1mer complex formation occurred at 4 °C was about 2.5 h (Figure 1E). The concentration of KaiB1-94 where the half maximal KaiB1-94-KaiC1mer complex formation occurred was about 0.7 μM (Figure 1F). This value falls within the concentration range at which clock oscillations occur in the in vitro KaiABC clock system [5].

KaiB-interacting domain of KaiC

The KaiC subunit is composed of an N-terminal domain (KaiCN) and a C-terminal domain (KaiCC) [32]. When we incubated 5 μM of KaiCDD monomeric domain protein (KaiCN1mer or KaiCC/DD1mer) with 30 μM KaiB1-94 at 4 °C for 6 h, only KaiCN1mer formed a complex with KaiB1-94 (Figures 2A and B, and Table S1). Even when the reaction mixtures were incubated at 25 °C, we did not detect any KaiB1-94-KaiCC/DD1mer complex (Figure 2B). Thus, the KaiB-interacting site of KaiC was located on the N-terminal domain of KaiC.

thumbnail
Figure 2. Native PAGE gels, gel filtration chromatography elution profiles, and the immunoblots of the KaiB1-94-KaiC complexes.

A, B. Native PAGE gels. Reaction mixtures containing 15 μM KaiB1-94 and 5 μM KaiCN1mer (A) or KaiCC/DD1mer (B) were incubated at 4 °C for the periods indicated (hereafter, unless otherwise stated, 0 h shows data for samples taken at the onset of incubation) and subjected to native PAGE. Other conditions were the same as for Figure 1C legend. C. Gel filtration chromatography elution profiles and an immunoblot. Reaction mixtures containing 7.5 μM KaiB1-94, 2.5 μM KaiCN6mer, and 1 mM ATP plus 5 mM MgCl2 (Mg-ATP) were incubated at 4 °C for 6 h and then subjected to gel filtration chromatography on a Superdex 200/HR 10/30 column equilibrated with reaction buffer containing 0.5 mM ATP and 5 mM MgCl2 at 4 °C. We also separately analyzed KaiB1-94 and KaiCN6mer by gel filtration chromatography as controls. The peak fraction samples were subjected to SDS-PAGE, blotted to PVDF membranes, and reacted with an anti-KaiB antiserum. Other conditions were the same as for Figure 1C legend. Black solid line, KaiB1-94 + KaiCN6mer; gray solid line, KaiCN6mer; black broken line, KaiB1-94.

https://doi.org/10.1371/journal.pone.0080200.g002

When we incubated 2.5 μM KaiCN6mer with 7.5 μM KaiB1-94 in the presence of Mg-ATP, KaiCN6mer showed a weak association with KaiB1-94 that was detected by gel filtration chromatography but not by Native PAGE (Figure 2C and Table S2). Unexpectedly, most KaiCN6mer became monomeric when it formed a complex with KaiB1-94 (Figure 2C). KaiCC/DD6mer, on the other hand, did not form a complex with KaiB1-94 (Table S2). KaiC6mer, therefore, also associates with KaiB1-94 via its N-terminal domain.

Mg-ATP-induced dissociation of the KaiB1-94-KaiC1mer complex

Addition of ATP as well as Mg-ATP was able to hexamerize KaiCDD1mer and KaiCN1mer (Figure 3A) as described previously [33]. When the KaiB1-94-KaiCDD1mer complex was incubated in the presence of 1 mM ATP or 1 mM ATP plus 5 mM MgCl2 (Mg-ATP) at 4 °C for 6 h, Mg-ATP, but not ATP, induced the hexamerization of KaiCDD (Figure 3B) and the concomitant release of KaiB1-94 (Figures 3B and 3C). We also detected a small amount of the KaiB1-94-KaiCDD6mer complex (Figure 3C), which was likely formed from the KaiCDD6mer hexamerized by Mg-ATP, and the KaiB1-94 released from the KaiB1-94-KaiCDD1mer complex during incubation. These results suggest that Mg-ATP but not ATP reduced the affinity of KaiC1mer for KaiB1-94 to dissociate the KaiB1-94-KaiCDD1mer complex. When we examined the ATP- and Mg-ATP-induced oligomerization of KaiCN in the KaiB1-94-KaiCN1mer complex, we obtained essentially the same results; Mg-ATP, to a much greater extent than ATP, induced oligomerization of KaiCN in the KaiB1-94-KaiCN1mer complex (Figure 3B). These results suggest that Mg-ATP (and ATP) inhibited the association of KaiB1-94 with the N-terminal domain of KaiC.

thumbnail
Figure 3. Mg-ATP-induced dissociation of the KaiB1-94-KaiC1mer complex.

A. Native PAGE gels showing the ATP-induced hexamerization of KaiCDD1mer and KaiCN1mer in the presence or absence of MgCl2. KaiCs1mer (5 μM) were hexamerized by incubation with 1 mM ATP with (Mg-ATP) or without (ATP) 5 mM MgCl2 at 4 °C for the periods indicated. Other conditions were the same as described for Figure 1C legend. B. Native PAGE gels of the KaiB1-94-KaiC1mer complex after incubation with ATP or Mg-ATP. Reaction mixtures containing 15 μM KaiB1-94 and 5 μM KaiCDD1mer in reaction buffer were incubated at 4 °C for 16 h to allow the formation of the KaiB1-94-KaiC1mer complex. After addition of 1 mM ATP with (Mg-ATP) or without (ATP) 5 mM MgCl2 to the complex, the reaction mixtures were further incubated in reaction buffer at 4 °C for 6 h. Other conditions were the same as described for Figure 1C. C. Gel filtration chromatography elution profiles of the KaiB1-94-KaiCDD1mer complex incubated with or without Mg-ATP. Reaction mixtures containing 24 μM KaiB1-94 and 24 μM KaiCDD1mer in reaction buffer were incubated at 4 °C for 16 h and then subjected to gel filtration chromatography on a Superdex 75/HR 10/30 column equilibrated with reaction buffer at 4 °C, and KaiB1-94-KaiCDD1mer complex fractions were collected. With (gray) or without (black) addition of Mg-ATP to the complex, the reaction mixtures were further incubated in reaction buffer at 4 °C for 6 h and then subjected to gel filtration chromatography on a Superdex 200/HR 10/30 column equilibrated with reaction buffer. The peak fractions were subjected to SDS-PAGE. Other conditions were the same as described for Figure 1C. Left and right gels, the 1st and 3rd peak fractions of the reaction products with addition of Mg-ATP, respectively; middle gels, the peak fraction products without addition of Mg-ATP corresponding to the 2nd peak fraction of the reaction products with addition of Mg-ATP.

https://doi.org/10.1371/journal.pone.0080200.g003

Effects of mutations in the ATPase motifs of KaiC on KaiB1-94-KaiC complex formation

Both the N- and C-terminal domains of the KaiC subunit have a series of ATPase motifs (a Walker's motif A, a Walker's motif B, and a CatE [2]). When we examined the effects of mutations in those motifs on formation of KaiB1-94-KaiC1mer complexes—using KaiCs with K53H and K294H mutations in Walker’s motif A and CatE1- and CatE2- mutations in CatEs [32]—we found that all the mutants we examined (KaiCK53H/DD, KaiCCatE1-/DD, KaiCK294H/DD, and KaiCCatE2-/DD) formed complexes with KaiB1-94 (Figure 4A and Table S1). Thus, none of the mutations in the ATPase motifs affected formation of the KaiB1-94-KaiCDD1mer complex that occurs in the absence of ATP.

thumbnail
Figure 4. Effects of mutations in the ATPase motifs and phosphorylation sites of KaiC on formation of the KaiB1-94-KaiC6mer complex.

A. Native PAGE gels of the reaction products of KaiC1mer with KaiB1-94. Reaction mixtures containing 15 μM KaiB1-94 and 5 μM KaiC1mer were incubated at 4 °C for the periods indicated. Other conditions were the same as described for Figure 1. B. Native PAGE gels of the reaction products of KaiC6mer with KaiB1-94. Reaction mixtures containing 5 μM KaiB1-94 and 1 μM KaiC6mer were incubated in the presence of Mg-ATP at 4 °C for the periods indicated. Other conditions were the same as described for Figure 1C. C. A typical 2D SDS-PAGE gel from the native-PAGE gel shown in Figure 4B. The protein bands were excised, and the proteins were extracted from them and subjected to SDS-PAGE. Other conditions were the same as described for Figure 1C. The bands 1 to 3 were the KaiCCatE2-/DD6mer band (control), the upper band of the reaction products of KaiCCatE2-/DD6mer with KaiB1-94 incubated at 4 °C for 6 h, and the lower band of the reaction products of KaiCCatE2-/DD6mer with KaiB1-94 incubated similarly. D. Native PAGE gels of the reaction products of KaiC6mer with KaiB1-94. Reaction mixtures were incubated at 4 °C for the periods indicated. Other conditions were the same as described for Figure 4B expect that unphosphorylatable mutant KaiCs were used. E. SDS-PAGE gels showing no phosphorylation of KaiCAA6mer, KaiCCatE2-6mer, and KaiCK294H6mer. KaiCs6mer (0.5 μM) were incubated with Mg-ATP in the presence of 0.5 μM KaiA at 40 °C for the periods indicated and then subjected to SDS-PAGE. p-KaiC, the phosphorylated forms of KaiC; np-KaiC, the unphosphorylated form of KaiC. Other conditions were the same as described for Figure 3.

https://doi.org/10.1371/journal.pone.0080200.g004

When we examined KaiB1-94-KaiC6mer complex formation in the presence of Mg-ATP, both KaiCK294H/DD6mer and KaiCCatE2-/DD6mer, which have a C-terminal ATPase motif mutation, as well as KaiCDD control, formed complexes with KaiB1-94 whereas KaiCK53H/DD6mer and KaiCCatE1-/DD6mer, which have an N-terminal ATPase motif mutation, did not (Figure 4B). Native PAGE followed by SDS-PAGE revealed that all the candidate complex bands examined contained both KaiB1-94 and KaiC (a typical example is shown in Figure 4C). These results indicate that KaiC’s N-terminal ATPase motifs were responsible for formation of the KaiB1-94-KaiC6mer complex that occurred in the presence of Mg-ATP. This observation is consistent with the finding described above that KaiC associates with KaiB1-94 via its N-terminal domain (Figure 2) and that Mg-ATP (and ATP) inhibits the association (Table S2).

Effects of the ATP, Mg-ATP, Mg-AMPPNP, and Mg-ATPγS used for the hexamerization of KaiC on KaiB1-94-KaiC6mer complex formation

The observations described above suggest that ATP hydrolysis followed by the release of ADP from the N-terminal ATPase motifs are involved in KaiC6mer-KaiB1-94 complex formation. We therefore examined whether the KaiCs6mer hexamerized by unhydrolyzable substrate analogs of ATPases, ATP (KaiC6mer (ATP)), Mg-AMPPNP (KaiC6mer (Mg-AMPPNP)), and Mg-ATPγS (KaiC6mer (Mg-ATPγS)) formed a complex with KaiB1-94 and found that only KaiC6mer hexamerized with Mg-ATP (KaiC6mer (Mg-ATP)), which is a hydrolyzable substrate for ATPases, did (Figure 5A). These results support our hypothesis that hydrolysis of ATP in the N-terminal ATPase motifs is required for KaiB1-94-KaiC6mer complex formation because ATP and the unhydrolyzable substrates analogs examined inhibited KaiB1-94-KaiC6mer complex formation.

thumbnail
Figure 5. Effects of the ATP and ATP analogs used for the hexamerization of KaiC on formation of the KaiB1-94-KaiCDD6mer complex assayed by native PAGE and gel filtration chromatography.

A. Native PAGE gels. Reaction mixtures were incubated at 4 °C for the periods indicated. Other conditions were the same as described for Figure 4B except that KaiCDD6mer (ATP), KaiCDD6mer (Mg-AMPPNP), and KaiCDD6mer (Mg-ATPγS) were used in the presence of 1 mM ATP, Mg-AMPPNP, and Mg-ATPγS, respectively. B. Gel filtration chromatography elution profiles of KaiCDD6mer (Mg-ADP-AlFX) (dotted line) and a KaiB1-94-KaiCDD6mer (Mg-ADP-AlFX) complex (solid line). KaiCDD1mer (20 μM) was incubated in reaction buffer containing 6 mM ADP, 30 mM MgCl2, 2.5 mM NaF, and 2.5 mM AlCl3 at 25 °C for 2 h, and the reaction mixtures were mixed with an equal volume of 0 or 60 μM KaiB1-94 in the same buffer and then further incubated at 25 °C for 6 h. The reaction mixtures were then subjected to gel filtration chromatography on a Superdex 200/HR 10/30 column equilibrated with reaction buffer at 4 °C. Other conditions were the same as described for Figure 2.

https://doi.org/10.1371/journal.pone.0080200.g005

KaiC6mer (Mg-ADP-AlFX) forms a complex with KaiB1-94

Because ADP-AlFX mimics ADP-Pi [34], we used gel filtration chromatography to determine whether Mg-ADP-AlFX hexamerizes KaiCDD1mer and confirmed that it did (Figure 5B). Next, we examined the possible complexes formed by KaiCDD6mer (Mg-ADP-AlFx) with KaiB1-94. KaiCDD6mer (Mg-ADP-AlFX) formed a complex with KaiB1-94 (Figure 5B), but we also detected a KaiB1-94-KaiCDD1mer complex under conditions wherein a substantial amount of KaiCDD1mer was not hexamerized (Figure 5B). This observation supports our above conclusion that ATP hydrolysis is required for formation of the KaiB1-94-KaiCDD6mer complex. Therefore, the conformation of KaiCDD6mer (Mg-ADP-Pi) that allows complex formation with KaiB1-94 may differ from that of KaiCDD6mer (Mg-ATP) without ATP hydrolysis and that of KaiC6mer (ATP).

Effects of KaiC phosphorylation-site mutations and autophosphorylation mutations on KaiB1-94-KaiC6mer complex formation

We examined the effects of mutations in the two phosphorylation sites of KaiC on KaiB1-94-KaiC6mer complex formation. While KaiCWT6mer (Figure 4D) and KiaCDD6mer (Figure 5A) formed a complex with KaiB1-94, KaiCAA6mer did not form such a complex (Figure 4D). These results are consistent with previous reports showing that phosphorylated KaiC (KaiCDD and KaiCDE, a mutant KaiC similar to KaiCDD) but not unphosphorylated KaiC (KaiCAA) formed a complex with KaiBWT [5,16-18]. When KaiCWT6mer was incubated with Mg-ATP in the presence of KaiA, it was highly phosphorylated (Figure 4E) [8,9,13,21,27,35]. The two upper bands correspond to the phosphorylated forms of KaiC whereas the lowest band corresponds to the unphosphorylated form of KaiC (Figure 4E) [3,13]. KaiCAA6mer, KaiCCatE2-6mer, and KaiCK294H6mer did not show any phosphorylated bands even in the presence of KaiA (Figure 4E).

Next, we examined the effects of KaiC autophosphorylation mutations on KaiB1-94-KaiC6mer complex formation. In spite of lacking the autophosphorylation [13], both KaiCK294H6mer and KaiCCatE2-6mer formed complexes with KaiB1-94 (Figure 4D). Thus, phosphorylation of KaiC per se was not essential for KaiB1-94-KaiC6mer complex formation, although KaiC’s phosphorylation state might indirectly modulate complex formation.

Effects of KaiC phosphorylation-site mutations on the N-terminal ATPase activity of KaiC

To examine the effects of mutations in the two phosphorylation sites of KaiC on the N-terminal ATPase activity of KaiC, we compared the ATPase activities of KaiCCatE2-/AA6mer, KaiCCatE2-/DD6mer, and KaiCCatE2-6mer because KaiCCatE2- lacks the C-terminal ATPase activity of KaiC [13]. The ATPase activities of KaiCCatE2-/DD6mer and KaiCCatE2-6mer reflect the N-terminal ATPase activity while that of KaiCDD6mer and KaiCWT6mer reflect the total ATPase activity. The former activities were approximately half of the latter activities (Figure 6 and Table S4). These four KaiCs6mer formed a complex with KaiB1-94 (Tables S2 and S3). On the other hand, the ATPase activity of KaiCAA6mer, which did not form a complex with KaiB1-94 (Table S3), showed 6 times higher ATPase activity than KaiCDD6mer and KaiCWT6mer (Figure 6). In consistent with this, KaiCAA has been reported to show 2.5 times higher ATPase activity than KaiCDE, which is a mutant KaiC similar to KaiCDD [4]. The ATPase activity of KaiCCatE2-/AA, which formed a complex with KaiB1-94 that could only be detected by silver staining (Table S3), was more than 3 times as high as those of KaiCCatE2-/DD6mer and KaiCCatE2-6mer. The ATPase activities of KaiCCatE2-/AA6mer, KaiCCatE2-/DD6mer, and KaiCCatE2-6mer all reflect the N-terminal ATPase activity of KaiC. Thus, although the N-terminal ATPase activity of KaiC is probably required for KaiC6mer to form a complex with KaiB1-94, its excessively high activity (KaiCAA6mer and KaiCCatE2-/AA6mer) may inhibit complex formation. The ATPase activity of KaiCCatE1-/DD6mer reflecting the C-terminal ATPase activity of KaiC was almost the same as that of KaiCC/DD6mer and approximately half of that of KaiCDD6mer (Figure 6 and Table S4), as described previously [10].

thumbnail
Figure 6. Effects of mutations in the ATPase motifs and phosphorylation sites of KaiC on KaiC’s ATPase activity.

We incubated 1 μM KaiCs6mer with Mg-ATP in reaction buffer at 4 °C for 6 h and then measured ATPase activities. Values are means ± SD from triplicate assay. KaiCs6mer: WT, KaiCWT6mer; DD, KaiCDD6mer; AA, KaiCAA6mer; N, KaiCN6mer; C/DD, KaiCC/DD6mer; K53H/DD, KaiCK53H/DD6mer; CatE1-/DD, KaiCCatE1-/DD6mer; CatE2-, KaiCCatE2-6mer; CatE2-/AA, KaiCCatE2-/AA6mer; CatE2-/DD, KaiCCatE2-/DD6mer; K294H/DD, KaiCK294H/DD6mer.

https://doi.org/10.1371/journal.pone.0080200.g006

Discussion

Since KaiC1mer formed a complex with KaiB1-94 (Figure 1C), the KaiC subunit per se is able to form a complex with KaiB. Recently, NMR analysis revealed that KaiC1mer forms a complex with KaiB via its N-terminal domain [29], and here KaiC (KaiC6mer as well as KaiC1mer) formed a complex with KaiB1-94 via its N-terminal domain (Figure 2). ATP, Mg-ATP, Mg-AMPPNP, and Mg-ATPγS hexamerize KaiC, probably by binding subunits [13,32,36,37]. KaiC6mer (ATP), KaiC6mer (Mg-AMPPNP), and KaiC6mer (Mg-ATPγS) did not form a complex with KaiB1-94 (Figure 5A). This indicates that when KaiC-bound ATP (or ATP analogs) is unhydrolyzable, KaiC does not form a complex with KaiB. KaiCDD6mer (Mg-ADP-AlFX), which mimics the Mg-ADP-Pi state of KaiCDD6mer, formed a complex with KaiB1-94 (Figure 5B). It is likely, therefore, that ATP regulates KaiB-KaiC6mer complex formation by hindering complex formation by KaiC6mer-bound ATP, and its hydrolysis is required for complex formation.

That mutations in the N-terminal but not its C-terminal ATPase motifs affected the complex formation of KaiC6mer (Mg-ATP) with KaiB1-94 (Figure 4B) indicates that ATP hydrolysis by the N-terminal KaiC’s ATPase motifs is responsible for KaiB-KaiC6mer complex formation. These results are consistent with a recent report [38]. Because the KaiCN6mer-KaiB1-94 complex rapidly dissociated into a KaiB1-94-KaiCN1mer complex (Figure 2C), the partial dissociation (or relaxation) of the N-terminal domains of KaiC6mer probably occurred on the interaction of KaiC6mer with KaiB. This partial dissociation of the KaiC6mer N-terminal domains is likely required for formation of KaiB-KaiC6mer complex via KaiC’s N-terminal domain. The relaxation of the KaiC6mer N-terminal domains on interaction of KaiB has been revealed recently by NMR analysis [29].

The KaiB molecule, which is a homotetramer organized as a dimer of dimers (KaiB4mer) [24], probably dissociates into two dimers (KaiB2mer) on interaction with KaiC6mer and forms a complex comprising one molecule of KaiC6mer and two molecules of KaiB2mer [17], as suggested by cryo-electron microscopy analysis [25]. We have proposed that the positively charged cleft (PC) of the KaiB4mer molecule, where the functionally important KaiB residues are concentrated, is an active site(s) required for interaction with KaiA and KaiC [24,28,39]. The PC of KaiB4mer, which is located on the dimer-dimer interface [24], is probably exposed by dissociation of KaiB4mer into dimers to interact with KaiC6mer [28,39], whereas the corresponding region of KaiB1-94, a dimeric mutant of KaiB, is always exposed [17]. Two areas on KaiC6mer molecule, on the other hand, are highly negatively charged—one around and inside the pore of KaiC6mer N-terminal domains and the other around the inter-subunit interface of one of two adjacent KaiC6mer N-terminal domains (Figures S1A and S1B) [37]. Interestingly, the ATP bound to the N-terminal ATPase motifs (namely, ATP-binding sites) is located adjacent to the latter area of KaiC (Figure S1B) [37]. Electrostatic interaction between the PC on KaiB and the aforementioned area of KaiC may allow sequestration of KaiB4mer (also KaiB2mer such as KaiB1-94) and induce dissociation into dimers (temporal weak association). Then, the dissociation of two adjacent N-terminal domains in KaiC6mer resulting from the hydrolysis of ATP bound to the N-terminal ATPase motifs on one of the two adjacent subunits (Figures S1C and S1D), which pastes the two N-terminal domains each other [13,32,36,37], may expose the latter area of KaiC—a possible KaiB-interacting surface—to KaiB2mer, and electrostatic interaction between the PC on KaiB and the latter area of KaiC may result in the tight association of KaiB2mer with KaiC6mer. The ATP bound to the N-terminal ATPase motifs inhibits the association of KaiB2mer with KaiC6mer via KaiC N-terminal domains, as demonstrated in KaiB1-94-KaiC6mer complex formation (Figure 3C). Thus, we calculated the surface potentials of KaiC6mer without ATP (Figure S1C) and with ATP (Figure S1D) and found them to be almost the same and unlikely to affect the interaction of KaiC with KaiB. While KaiBWT4mer formed a complex formation with KaiC6mer (Figure 1B), it did so only slightly with KaiC1mer (Figure 1D). KaiB1-94, in contrast, formed a complex with both KaiC6mer and KaiC1mer (Figures 1B, 1C and 5A). Therefore, interaction with KaiC6mer but not with KaiC1mer likely enhanced KaiB4mer dimerization, suggesting the possibility that enhancement requires the hexameric structure of KaiC6mer N-terminal domains.

It has been previously proposed that the phosphorylation state of KaiC was involved in its forming a complex with KaiB [18]. However, our data described here showing that KaiCK294H6mer and KaiCCatE2-6mer, which lack the autokinase activity (Figure 4E) [13], formed complexes with KaiB1-94 (Figure 4D) indicated that the phosphorylated state of KaiC is not essential for KaiB-KaiC6mer complex formation. Our results are consistent with the recently reported results that KaiCCatE2-6mer formed a complex with KaiBWT [38]. The phosphorylation state of KaiC, therefore, is not directly involved in and essential for complex formation.

KaiCAA6mer and KaiCCatE2-/AA6mer, which did not form a complex with KaiB1-94, showed much higher ATPase activity than any other KaiC ATPase motif mutants we examined (Figure 6 and Table S4). The excessively high ATPase activity of the N-terminal ATPase motifs of KaiCAA6mer, which bounces in and out of the ATP bound to the N-terminal ATPase motifs, may hinder formation of the KaiB-KaiC6mer complex. In KaiC6mer, the phosphorylation state of the C-terminal domain could affect its association with KaiB via the N-terminal domain through modulating the N-terminal ATPase activity. However, we cannot exclude a possibility that KaiCAA6mer, which is likely not a perfect mimic for the fully unphosphorylated form of KaiC, might have a changed structure, which might enhance its ATPase activity but might reduce its association with KaiB. We propose the following model for ATP regulation of KaiB-KaiC interaction. The KaiC subunit is able to form a complex with KaiB (Figure 7A). The N-terminal domains of KaiC6mer are partially dissociated (relaxed) when the ATP bound to the N-terminal ATPase motifs that pastes adjacent N-terminal domains each other in KaiC6mer is hydrolyzed (Figure 7B), which allows KaiB to associate with the KaiC6mer N-terminal domains. Then, KaiC6mer-associated KaiB suppresses the ATPase activity of KaiC6mer [4] by inhibiting ATP binding to KaiC6mer N-terminal domains. KaiCAA6mer and KaiCCatE2-/AA6mer, which seem to mimic the unphosphorylation state of KaiC, have excessively high N-terminal ATPase activity (Figure 6 and Table S4), and that may cause rapid interconversion of the ridged (ATP-bound) and relaxed (ATP-hydrolyzed; ADP-bound or unbound) conformations of the N-terminal domains in KaiC6mer. We propose here that this rapid interconversion inhibits KaiB-KaiC6mer complex formation though we cannot explain this inhibiting mechanism at present (Figure 7C).

thumbnail
Figure 7. ATP-mediated regulation model for KaiB-KaiC interaction.

A. Interaction between KaiB and KaiC1mer. B. Partial dissociation (relaxation) of the N-terminal domain of KaiC6mer and complex formation of one KaiC6mer molecule with 2 KaiB2mer molecules. For simplification, we express Mg-chelated ATP and ADP as ATP and ADP. The nucleotide state of the C-terminal ATP-binding site (ATPase motifs) of KaiC6mer is not known. C. Rapid interconversion between the rigid ATP-bound and relaxed ATP-hydrolyzed form (ADP-bound or unbound) conformations in the N-terminal domains of KaiC6mer.

https://doi.org/10.1371/journal.pone.0080200.g007

The C-terminal ATPase motifs of KaiC are involved in the hexamerization of KaiC C-terminal domains [13] as well as their inter-subunit autophosphorylation [8] and probably autodephosphorylation [12]. Although the KaiC N-terminal ATPase motifs are involved in the hexamerization of KaiC N-terminal domains [13], and the affinity of the N-terminal ATPase motifs for ATP is higher than that of its C-terminal ATPase motifs [13], and therefore, the N-terminal domains are likely to be more tightly connected than the C-terminal domains [13,16,29], the function of the N-terminal ATPase motifs remains unknown. In this investigation, we have succeeded in revealing that the nucleotide state of the N-terminal ATPase motifs regulates KaiB-KaiC interaction. Because KaiB and SasA competitively associate with KaiC via KaiC N-terminal domains [14,17], the nucleotide state also can regulate KaiC-SasA interaction via the KaiB-KaiC interaction. ATP acts not only as a biological fuel, but also as a physiological regulator. There are some examples for ATP regulation of the physiological function. Many different cell types release ATP in response to mechanical or biochemical stimulation, and the released ATP modulates cell function by activating nearby purinoceptors, such as ion channel P2X receptors and G-protein-coupled P2Y receptors [40-42].

Supporting Information

Figure S1.

Electrostatic surface potential of KaiC6mer and KaiC1mer of Synechococcus KaiC (PDB code: 2GBL). We calculated electrostatic surface representations of KaiC6mer (A, B) and KaiC1mer (C, D) using the PyMOL plug-in APBS [43]. A. Top view of the N-terminal domain of KaiC6mer with ATP. B. Side view of KaiC6mer with ATP. Interface of KaiC without ATP (C) and with ATP (D). The saturation thresholds were -5 and +5. For electrostatic surface potential: blue, positive; red, negative. Arrows indicated the negatively charged areas.

https://doi.org/10.1371/journal.pone.0080200.s001

(TIF)

Table S1.

Complex formation of KaiCs1mer with KaiB1-94.

https://doi.org/10.1371/journal.pone.0080200.s002

(DOC)

Table S2.

Formation of KaiB1-94-KaiCs6mer complex.

https://doi.org/10.1371/journal.pone.0080200.s003

(DOC)

Table S3.

Complex formation of unphosphorylatable KaiCs6mer with KaiB1-94.

https://doi.org/10.1371/journal.pone.0080200.s004

(DOC)

Table S4.

Effects of mutations in the ATPase motifs and phosphorylation sites on KaiC ATPase activity.

https://doi.org/10.1371/journal.pone.0080200.s005

(DOC)

Acknowledgments

We thank Satoko Ogawa and Kumiko Tanaka for technical support and Miriam Bloom (SciWrite Biomedical Writing & Editing Services) for professional editing. We also thank Drs. Reiko Murakami and Tomoko Miyata for technical advice.

Author Contributions

Conceived and designed the experiments: RM MI. Performed the experiments: RM AN SY KO. Analyzed the data: RM MI. Contributed reagents/materials/analysis tools: RM MI. Wrote the manuscript: RM.

References

  1. 1. Grobbelaar N, Huang TC, Lin HY, Chow TJ (1986) Dinitrogen-fixing endogenous rhythm in Synechococcus RF-1. FEMS Microbiol Lett 37: 173–177. doi:https://doi.org/10.1111/j.1574-6968.1986.tb01788.x.
  2. 2. Ishiura M, Kutsuna S, Aoki S, Iwasaki H, Andersson CR et al. (1998) Expression of a gene cluster kaiABC as a circadian feedback process in cyanobacteria. Science 281: 1519–1523. doi:https://doi.org/10.1126/science.281.5382.1519. PubMed: 9727980.
  3. 3. Nakajima M, Imai K, Ito H, Nishiwaki T, Murayama Y et al. (2005) Reconstitution of circadian oscillation of cyanobacterial KaiC phosphorylation in vitro. Science 308: 414–415. doi:https://doi.org/10.1126/science.1108451. PubMed: 15831759.
  4. 4. Terauchi K, Kitayama Y, Nishiwaki T, Miwa K, Murayama Y et al. (2007) ATPase activity of KaiC determines the basic timing for circadian clock of cyanobacteria. Proc Natl Acad Sci U_S_A 104: 16377–16381. doi:https://doi.org/10.1073/pnas.0706292104. PubMed: 17901204.
  5. 5. Kageyama H, Nishiwaki T, Nakajima M, Iwasaki H, Oyama T et al. (2006) Cyanobacterial circadian pacemaker: Kai protein complex dynamics in the KaiC phosphorylation cycle in vitro. Mol Cell 23: 161–171. doi:https://doi.org/10.1016/j.molcel.2006.05.039. PubMed: 16857583.
  6. 6. Akiyama S, Nohara A, Ito K, Maéda Y (2008) Assembly and disassembly dynamics of the cyanobacterial periodosome. Mol Cell 29: 703-716. doi:https://doi.org/10.1016/j.molcel.2008.01.015. PubMed: 18342562.
  7. 7. Nishiwaki T, Iwasaki H, Ishiura M, Kondo T (2000) Nucleotide binding and autophosphorylation of the clock protein KaiC as a circadian timing process of cyanobacteria. Proc Natl Acad Sci U_S_A 97: 495–499. doi:https://doi.org/10.1073/pnas.97.1.495. PubMed: 10618446.
  8. 8. Hayashi F, Ito H, Fujita M, Iwase R, Uzumaki T et al. (2004) Stoichiometric interactions between cyanobacterial clock proteins KaiA and KaiC. Biochem Biophys Res Commun 316: 195–202. doi:https://doi.org/10.1016/j.bbrc.2004.02.034. PubMed: 15003530.
  9. 9. Uzumaki T, Fujita M, Nakatsu T, Hayashi F, Shibata H et al. (2004) Crystal structure of the C-terminal clock-oscillator domain circadian clock of cyanobacteria. Nat Struct Mol Biol 104: 16377–16381.
  10. 10. Murakami R, Miyake A, Iwase R, Hayashi F, Uzumaki T et al. (2008) ATPase activity and its temperature compensation of the cyanobacterial clock protein KaiC. Genes Cells 13: 387-395. doi:https://doi.org/10.1111/j.1365-2443.2008.01174.x. PubMed: 18363969.
  11. 11. Egli M, Mori T, Pattanayek R, Xu Y, Qin X et al. (2012) Dephosphorylation of the core clock protein KaiC in the cyanobacterial KaiABC circadian oscillator proceeds via an ATP synthase mechanism. Biochemistry 51: 1547-1558. doi:https://doi.org/10.1021/bi201525n. PubMed: 22304631.
  12. 12. Nishiwaki T, Kondo T (2012) Circadian autodephosphorylation of cyanobacterial clock protein KaiC occurs via formation of ATP as intermediate. J Biol Chem 287: 18030-18035. doi:https://doi.org/10.1074/jbc.M112.350660. PubMed: 22493509.
  13. 13. Hayashi F, Itoh N, Uzumaki T, Iwase R, Tsuchiya Y et al. (2004) Roles of two ATPase-motif-containing domains in cyanobacterial circadian clock protein KaiC. J Biol Chem 279: 52331–52337. doi:https://doi.org/10.1074/jbc.M406604200. PubMed: 15377674.
  14. 14. Valencia SJ, Bitou K, Ishii K, Murakami R, Morishita M et al. (2012) Phase-dependent generation and transmission of time information by the KaiABC circadian clock oscillator through SasA-KaiC interaction in cyanobacteria. Genes Cells 17: 398-419. doi:https://doi.org/10.1111/j.1365-2443.2012.01597.x. PubMed: 22512339.
  15. 15. Nishiwaki T, Satomi Y, Nakajima M, Lee C, Kiyohara R et al. (2004) Role of KaiC phosphorylation in the circadian clock system of Synechococcus elongatus PCC 7942. Proc Natl Acad Sci U_S_A 101: 13927-13932. doi:https://doi.org/10.1073/pnas.0403906101. PubMed: 15347812.
  16. 16. Chang YG, Kuo NW, Tseng R, LiWang A (2011) Flexibility of the C-terminal, or CII, ring of KaiC governs the rhythm of the circadian clock of cyanobacteria. Proc Natl Acad Sci U_S_A 108: 14431-14436. doi:https://doi.org/10.1073/pnas.1104221108. PubMed: 21788479.
  17. 17. Murakami R, Mutoh R, Iwase R, Furukawa Y, Imada K et al. (2012) The roles of the dimeric and tetrameric structures of the clock protein KaiB in the generation of circadian oscillations in cyanobacteria. J Biol Chem 287: 29506-29515. doi:https://doi.org/10.1074/jbc.M112.349092. PubMed: 22722936.
  18. 18. Nishiwaki T, Satomi Y, Kitayama Y, Terauchi K, Kiyohara R et al. (2007) A sequential program of dual phosphorylation of KaiC as a basis for circadian rhythm in cyanobacteria. EMBO J 26: 4029-4037. doi:https://doi.org/10.1038/sj.emboj.7601832. PubMed: 17717528.
  19. 19. Garces RG, Wu N, Gillon W, Pai EF (2004) Anabaena circadian clock proteins KaiA and KaiB reveal a potential common binding site to their partner KaiC. EMBO J 23: 1688-1698. doi:https://doi.org/10.1038/sj.emboj.7600190. PubMed: 15071498.
  20. 20. Ye S, Vakonakis I, Ioerger TR, LiWang AC, Sacchettini JC (2004) Crystal structure of circadian clock protein KaiA from Synechococcus elongatus. J Biol Chem 279: 20511-20518. doi:https://doi.org/10.1074/jbc.M400077200. PubMed: 15007067.
  21. 21. Iwasaki H, Nishiwaki T, Kitayama Y, Nakajima M, Kondo T (2002) KaiA-stimulated KaiC phosphorylation in circadian timing loops in cyanobacteria. Proc Natl Acad Sci U_S_A 99: 15788–15793. doi:https://doi.org/10.1073/pnas.222467299. PubMed: 12391300.
  22. 22. Williams SB, Vakonakis I, Golden SS, LiWang AC (2002) Structure and function from the circadian clock protein KaiA of Synechococcus elongatus: a potential clock input mechanism. Proc Natl Acad Sci U_S_A 99: 15357–15362. doi:https://doi.org/10.1073/pnas.232517099. PubMed: 12438647.
  23. 23. Hitomi K, Oyama T, Han SG, Arvai AS, Getzoff ED (2005) Tetrameric architecture of the circadian clock protein KaiB - A novel interface for intermolecular interactions and its impact on the circadian rhythm. J Biol Chem 280: 19127-19135. doi:https://doi.org/10.1074/jbc.M411284200. PubMed: 15716274.
  24. 24. Iwase R, Imada K, Hayashi F, Uzumaki T, Morishita M et al. (2005) Functionally important substructures of circadian clock protein KaiB in a unique tetramer complex. J Biol Chem 280: 43141-43149. doi:https://doi.org/10.1074/jbc.M503360200. PubMed: 16227211.
  25. 25. Pattanayek R, Williams DR, Pattanayek S, Mori T, Johnson CH et al. (2008) Structural model of the circadian clock KaiB-KaiC complex and mechanism for modulation of KaiC phosphorylation. EMBO J 27: 1767–1778. doi:https://doi.org/10.1038/emboj.2008.104. PubMed: 18497745.
  26. 26. Kitayama Y, Iwasaki H, Nishiwaki T, Kondo T (2003) KaiB functions as an attenuator of KaiC phosphorylation in the cyanobacterial circadian clock system. EMBO J 22: 2127-2134. doi:https://doi.org/10.1093/emboj/cdg212. PubMed: 12727879.
  27. 27. Pattanayek R, Williams DR, Rossi G, Weigand S, Mori T et al. (2011) Combined SAXS/EM based models of the S. elongatus post-translational circadian oscillator and its interactions with the output His-kinase SasA. PLOS ONE 6(8): e23697. doi:https://doi.org/10.1371/journal.pone.0023697. PubMed: 21887298.
  28. 28. Mutoh R, Mino H, Murakami R, Uzumaki T, Takabayashi A et al. (2010) Direct interaction between KaiA and KaiB revealed by a site-directed spin labeling electron spin resonance analysis. Genes Cells 15: 269–280. doi:https://doi.org/10.1111/j.1365-2443.2009.01377.x. PubMed: 20113360.
  29. 29. Chang YG, Tseng R, Kuo NW, LiWang A (2012) Rhythmic ring-ring stacking drives the circadian oscillator clockwise. Proc Natl Acad Sci U_S_A 109: 16847-16851. doi:https://doi.org/10.1073/pnas.1211508109. PubMed: 22967510.
  30. 30. Iwasaki H, Williams SB, Kitayama Y, Ishiura M, Golden SS et al. (2000) A kaiC-interacting sensory histidine kinase, SasA, necessary to sustain robust circadian oscillation in cyanobacteria. Cell 101: 223-233. doi:https://doi.org/10.1016/S0092-8674(00)80832-6. PubMed: 10786837.
  31. 31. Galkin VE, Wu Y, Zhang XP, Qian X, He Y et al. (2006) The Rad51/RadA N-terminal domain activates nucleoprotein filament ATPase activity. Structure 14: 983-992. doi:https://doi.org/10.1016/j.str.2006.04.001. PubMed: 16765891.
  32. 32. Hayashi F, Suzuki H, Iwase R, Uzumaki T, Miyake T et al. (2003) ATP-induced hexameric ring structure of the cyanobacterial circadian clock protein KaiC. Genes Cells 8: 287–296. doi:https://doi.org/10.1046/j.1365-2443.2003.00633.x. PubMed: 12622725.
  33. 33. Hayashi F, Iwase R, Uzumaki T, Ishiura M (2006) Hexamerization by the N-terminal domain and intersubunit phosphorylation by the C-terminal domain of cyanobacterial circadian clock protein KaiC. Biochem Biophys Res Commun 348: 864-872. doi:https://doi.org/10.1016/j.bbrc.2006.07.143. PubMed: 16901465.
  34. 34. Lunardi J, Dupuis A, Garin J, Issartel JP, Michel L et al. (1988) Inhibition of H+-transporting ATPase by formation of a tight nucleoside diphosphate-fluoroluminate complex at the catalytic site. Proc Natl Acad Sci U_S_A 85: 8958-8962. doi:https://doi.org/10.1073/pnas.85.23.8958. PubMed: 2904148.
  35. 35. Xu Y, Mori T, Johnson CH (2003) Cyanobacterial circadian clock: roles of KaiA, KaiB and the kaiBC promoter in regulating KaiC. EMBO J 22: 2117-2126. doi:https://doi.org/10.1093/emboj/cdg168. PubMed: 12727878.
  36. 36. Mori T, Saveliev SV, Xu Y, Stafford WF, Cox MM et al. (2002) Circadian clock protein KaiC forms ATP-dependent hexameric rings and binds DNA. Proc Natl Acad Sci U_S_A 99: 17203–17208. doi:https://doi.org/10.1073/pnas.262578499. PubMed: 12477935.
  37. 37. Pattanayek R, Wang J, Mori T, Xu Y, Johnson CH et al. (2004) Visualizing a circadian clock protein: crystal structure of KaiC and functional insights. Mol Cell 15: 375-388. doi:https://doi.org/10.1016/j.molcel.2004.07.013. PubMed: 15304218.
  38. 38. Phong C, Markson JS, Wilhoite CM, Rust MJ (2013) Robust and tunable circadian rhythms from differentially sensitive catalytic domains. Proc Natl Acad Sci U_S_A 110: 1124-1129. doi:https://doi.org/10.1073/pnas.1212113110. PubMed: 23277568.
  39. 39. Mutoh R, Mino H, Murakami R, Uzumaki T, Ishiura M (2011) Thermodynamically induced conformational changes of the cyanobacterial circadian clock protein KaiB. Appl Magn Reson 40: 525-534. doi:https://doi.org/10.1007/s00723-011-0228-2.
  40. 40. Khakh BS, Burnstock G (2009) The double life of ATP. Sci Am 301: 84-92. doi:https://doi.org/10.1038/scientificamerican0909-84. PubMed: 20058644.
  41. 41. Milner P, Kirkpatrick KA, Ralevic V, Toothill V, Pearson J et al. (1990) Endothelial cells cultured from human umbilical vein release ATP, substance P and acetylcholine in response to increased flow. Proc Biol Sic 241. pp. 245-248. PubMed: 1701555.
  42. 42. Yegutkin GG (2008) Nucleotide- and nucleoside-converting ectoenzymes: important modulators of purinergic signaling cascade. Biochim Biophys Acta 1783: 673-694. doi:https://doi.org/10.1016/j.bbamcr.2008.01.024. PubMed: 18302942.
  43. 43. Baker NA, Sept D, Joseph S, Holst MJ, McCammon JA (2001) Electrostatics of nanosystems: application to microtubules and the ribosome. Proc Natl Sci USA 98: 10037-10041. PubMed: 11517324.