Next Article in Journal
The Conjugates of Indolo[2,3-b]quinoline as Anti-Pancreatic Cancer Agents: Design, Synthesis, Molecular Docking and Biological Evaluations
Previous Article in Journal
Monolacunary Wells-Dawson Polyoxometalate as a Novel Contrast Agent for Computed Tomography: A Comprehensive Study on In Vivo Toxicity and Biodistribution
Previous Article in Special Issue
Improved Solubility and Stability of a Thermostable Carbonic Anhydrase via Fusion with Marine-Derived Intrinsically Disordered Solubility Enhancers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Function, Regulation, and Mechanism of Protein Turnover in Circadian Systems in Neurospora and Other Species

by
Haoran Zhang
,
Zengxuan Zhou
and
Jinhu Guo
*
MOE Key Laboratory of Gene Function and Regulation, School of Life Sciences, Sun Yat-sen University, Guangzhou 510275, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(5), 2574; https://doi.org/10.3390/ijms25052574
Submission received: 27 December 2023 / Revised: 18 February 2024 / Accepted: 20 February 2024 / Published: 22 February 2024
(This article belongs to the Special Issue Protein Stability Research)

Abstract

:
Circadian clocks drive a large array of physiological and behavioral activities. At the molecular level, circadian clocks are composed of positive and negative elements that form core oscillators generating the basic circadian rhythms. Over the course of the circadian period, circadian negative proteins undergo progressive hyperphosphorylation and eventually degrade, and their stability is finely controlled by complex post-translational pathways, including protein modifications, genetic codon preference, protein–protein interactions, chaperon-dependent conformation maintenance, degradation, etc. The effects of phosphorylation on the stability of circadian clock proteins are crucial for precisely determining protein function and turnover, and it has been proposed that the phosphorylation of core circadian clock proteins is tightly correlated with the circadian period. Nonetheless, recent studies have challenged this view. In this review, we summarize the research progress regarding the function, regulation, and mechanism of protein stability in the circadian clock systems of multiple model organisms, with an emphasis on Neurospora crassa, in which circadian mechanisms have been extensively investigated. Elucidation of the highly complex and dynamic regulation of protein stability in circadian clock networks would greatly benefit the integrated understanding of the function, regulation, and mechanism of protein stability in a wide spectrum of other biological processes.

1. Introduction

Living organisms have adapted to the daily rotation of the Earth on its axis. By means of cell-autonomous circadian clocks that can be synchronized to the daily and seasonal changes in external time cues, most notably light and temperature, organisms anticipate environmental transitions and perform activities during the day in sync with cycling environmental cues. Circadian rhythms are ubiquitous phenomena that recur daily in a self-sustaining, entrainable, and oscillatory manner with periods of approximately 24 h. Circadian clocks are critical timekeeping systems that control a wide spectrum of physiological and behavioral rhythms in most eukaryotic and prokaryotic organisms [1,2,3]. Circadian clocks regulate the expression of approximately 20–80% of protein-encoding genes in different species [4,5,6,7,8,9].
The rhythmic change in the abundance of core circadian clock proteins defines a critical nodal point for negative feedback and cause-and-effect oscillation within the circadian clock mechanism [1,10]; therefore, controlling the stability of circadian proteins is essential for maintaining robust and self-sustained oscillations.

2. The Molecular Network of the Circadian Clock

The Circadian clock controls the circadian rhythms of molecular, metabolic, physiological, and behavioral parameters within a period of approximately 24 h in most eukaryotes and certain prokaryotes [11]. “Circadian” means “about a day”, which derives from the Latin words “circa” and “diem”, respectively [12]. Circadian rhythms have several fundamental characteristic properties, which include the following: (1) persistability with a period of ~24 h under constant conditions without cycling external time cues, e.g., light or temperature; (2) resetability by external or internal cycling cues; and (3) temperature compensation which renders clock output largely insensitive to environmental temperature fluctuations [13].
At the molecular level, circadian clocks are controlled by a set of core circadian genes and many associated factors. The core circadian components constitute interlocked transcriptional and translational feedback loops, called transcriptional–translational negative feedback loops (TTFLs), which oscillate and drive the circadian expression of clock-controlled genes [10,14,15]. TTFL regulatory networks are present in all the identified kingdoms that possess circadian clocks, from prokaryotes to mammals (Figure 1).
Cyanobacteria emerged on Earth approximately 3–4 billion years ago, and some cyanobacterial species possess endogenous circadian clocks that control important processes, e.g., photosynthesis and nitrogen fixation [16]. The cyanobacterial circadian clock harbors both protein phosphorylation and transcription–translation cycles to produce circadian rhythms. In the unicellular cyanobacterium Synechococcus elongatus PCC 7942, KaiA, KaiB, and KaiC (KaiABC) are the core circadian components that constitute the pacemaker. KaiC has versatile activities, including autokinase, autophosphatase, and ATPase. KaiA binds to the A-loop of KaiC and promotes the phophosphorylation of S431 and T432 on KaiC, forming a post-translational oscillator (PTO) [17,18]. The circadian system in S. elongatus also contains TTFL, in which the transcripts from the kaiBC operon and the protein KaiC oscillate in a circadian manner in constant light. Overexpression of KaiC abolishes the expression of kaiBC [19]. The cyanobacterial PTO system can run independently of TTFL; however, the latter is necessary to stabilize and enhance the resetting of the Kai-based PTO in Synechococcus [20,21,22,23].
N. crassa is a filamentous fungus that has long served as an important circadian model organism in circadian research. The PAS domain-containing factors WHITE COLLAR 1 (WC-1) and WC-2 are circadian positive elements that form the heterodimer WHITE COLLAR complex (WCC). The WCC functions to activate the transcription of the frequency (frq) gene, which encodes the negative element of the circadian clock [24]. The FRQ-interacting RNA helicase (FRH) binds to FRQ and assists in the formation of the correct structure, proper phosphorylation, degradation, and association with other circadian and noncircadian factors [25,26,27]. The WCC progressively activates frq transcription before subjective dawn. The synthesis of FRQ is maximally repressed at subjective dusk, and the FRQ level decreases during the subjective night [28]. FRQ feeds back to repress the function of WCC as a transcription activator, thus ceasing its own transcription of frq. FRQ binds with casein kinase 1 (CSNK1, CK1) and other protein kinases, and hyperphosphorylated FRQ proteins undergo gradual degradation, after which WCC becomes free again and resumes its function in facilitating frq transcription. Neurospora TTFL directly controls the rhythmic expression of downstream clock-controlled genes (ccgs) by regulating the cis-elements located in the promoters of these genes, which are called Clock boxes (C-boxes) [10].
Figure 1. Circadian oscillators in different species. The circadian regulation in five important model organisms is shown, including cyanobacteria (S. elongatus PCC 7942) (A) [18,29], fungi (N. crassa) (B) [30], plants (A. thaliana) (C) [31,32,33], insects (D. melanogaster) (D) [11], and mammals (M. musculus) (E) [34]. Note that only the basic regulatory circuits are shown. Lines in orange highlight translation.
Figure 1. Circadian oscillators in different species. The circadian regulation in five important model organisms is shown, including cyanobacteria (S. elongatus PCC 7942) (A) [18,29], fungi (N. crassa) (B) [30], plants (A. thaliana) (C) [31,32,33], insects (D. melanogaster) (D) [11], and mammals (M. musculus) (E) [34]. Note that only the basic regulatory circuits are shown. Lines in orange highlight translation.
Ijms 25 02574 g001
In D. melanogaster, the core oscillator proteins include the positive elements Clock (Clk) and Cycle (Cyc) and the negative elements Period (Per) and Timeless (Tim). Clk and Cyc function as transcriptional activators to control the rhythmic transcription of per and tim, which encode the negative circadian elements PER and Tim. From early to midnight, Per and Tim are predominantly localized in the cytoplasm and form heterodimers. When it is late at night, the Per/Tim complex enters the nucleus and represses the transcription activity of Clk/Cyc, which subsequently decreases the expression of per and tim. A series of post-translational modifications (PTMs) finely control the pace of accumulation and degradation of the Per and Tim proteins. The gradual degradation of TIM and then of PER removes the repression activity on Clk/Cyc and restarts the transcription of per and tim, and Clk and Cyc form heterodimers to activate the transcription of per and tim and other clock-controlled genes containing the cis element of the enhancer box (E-box) [35,36]. Furthermore, Vri and Pdp1 are leucine zipper motif-containing transcription factors, which regulate clock transcription through V/P-box; on the other hand, Clk and Cyc also control the expression of vrille (vri) and par domain protein 1 (pdp1) through E-box elements in their promoters [37,38].
The mammalian circadian pacemaker is the suprachiasmatic nucleus (SCN) in the anterior of the hypothalamus [39]. The circadian rhythm is regulated by the positive transcription factors CLOCK and brain and muscle Arnt-like protein-1 (BMAL1) (also termed Arnt-like protein 1, ARNTL1), which induce the expression of the negative clock elements PER1/2 and CRY1/2, forming an important negative-feedback loop. PER2 is significantly more unstable than PER1, and the progressive phosphorylation of PER1 occurs significantly faster than that of PER2 [40]. The orphan nuclear receptors REV-ERBs (−α and −β) and retinoic acid receptor (RAR)-related orphan receptor (ROR) (−α, −β, and −γ) act in interconnected ancillary circuits that act as ancillary loops to regulate Bmail1 expression. In this circuit, REV-ERB proteins inhibit, while ROR proteins activate, the transcription of Bmal1 [41].
Plant circadian clocks endow plants with the advantageous ability to adapt to environmental alternations by controlling key essential processes, such as photosynthetic activity, hypocotyl elongation, and the floral transition [42,43]. In A. thaliana, the basic TTFL is composed of a pair of dawn-expressed MYB-related transcription factors, LATE ELONGATED HYPOCOTYL (LHY), and CIRCADIAN CLOCK ASSOCIATED 1 (CCA1), which are expressed around dawn. CCA1 and LHY homo- or heterodimerize and bind to the evening element (EE) motifs on the promoters of other clock components, including PSEUDO RESPONSE REGULATOR (PRR), TIMING OF CAB EXPRESSION 1 (TOC1/PRR1), GI, and LUX ARRHYTHMO (LUX), EARLY FLOWERING 3 (ELF3), and ELF4, which form the ELF3-ELF4-LUX evening complex (EC) and downstream genes, to repress the expression of those evening-phased genes.
In addition, LHY and CCA1 mutually repress the expression of these genes [32]. PRR9, PRR7, PRR5, PRR3, and TOC1 are sequentially expressed and repress their own transcription, as well as that of CCA1 and LHY, and the loss of these proteins alters the period. Moreover, the transcription of PRR9 and PRR7 is negatively regulated by the EC complex. The transcriptional coactivators LIGHT-REGULATED WD1 (LWD1) and LWD2 interact with the TEOSINTE BRANCHED 1-CYCLOIDEA-PCF20 (TCP20) and TCP22 to promote the expression of CCA1, PRR9, PRR7, and TOC1 in the morning. In the afternoon, transcriptional activation is mediated by the NIGHT LIGHT-INDUCIBLE AND CLOCK-REGULATED proteins (LNK). RVE8 and RVE4 recruit transcriptional coactivators, and the RVE-LNK complex promotes the transcription of PRR9, PRR5, TOC1, GI, LUX, and ELF4. FHY3, FAR1, and HY5 are also implicated in the transcriptional activation of ELF4. In the evening, TOC1 represses all the daytime components, as well as GI, LUX, and ELF4. LUX and ELF4, together with ELF3, form the evening complex (EC), which is a transcriptional repressor of GIGANTEA (GI), PRR9, and PRR7. Moreover, GI and an EC variant containing BOA (NOX) function to activate the transcription of CCA1 and LHY [31,32,33]. In addition to regulating CO transcription through the TTFL pathway, PRR proteins interact with and stabilize CO and facilitate its accumulation under long days, which increases the binding of CO to the promoter of FLOWERING LOCUS (FT) and consequently enhances FT transcription and early flowering [44].
Despite the evolutionary distance between clock genes in different species, the architectures of molecular clockworks are highly similar across kingdoms, and the regulation of the circadian clock is an example of convergent evolution [11,45]. In eukaryotes, all circadian systems involve progressive increases in the global phosphorylation of certain clock components until highly phosphorylated forms are targeted for rapid degradation [10,46]. Moreover, the regulators of post-translational modifications and degradation of circadian clock proteins are evolutionally conserved. For instance, CK1 in fungi and animals and CK2 in plants are crucial for the phosphorylation of circadian clock proteins. In eukaryotes, the ubiquitin–proteasome system and autophagy are involved in the modification and degradation of clock proteins [47,48].

3. The Regulation of Circadian Clock Protein Stability in the Neurospora Circadian Clock

Neurospora is an important model for circadian studies and has greatly propelled advances in the investigation of a variety of circadian mechanisms at the molecular level [30]. It has been proposed that fungal circadian clock systems are evolutionarily closely related to animal clocks [10].

3.1. Post-Translational Regulation of FRQ Protein Stability in Neurospora

FRQ is dynamically controlled by a large number of time-of-day-specific phosphorylation sites. The phosphorylation of FRQ determines its activities and binding partners and eventually leads to its inactivation (Figure 2) [49,50,51,52]. To measure the degradation rate of a circadian clock protein, cycloheximide (CHX) is usually added to inhibit translation, after which the cells/tissues are harvested at a series of time points and subsequently subjected to Western blot analysis and densimetric quantification. In Neurospora, the FRQ decay rate can also be observed in vivo by transferring Neurospora from LL to DD, which results in steady-state phosphorylation and degradation [50,52,53].
FRQ contains multiple degrons that are critical for PTM and FRQ turnover. Different blocks of the FRQ protein have different impacts on its own turnover [51,54]. Strains with mutations within the N’ 689 aa region of FRQ show decreased turnover rates of the FRQ protein and longer periods, but the strains with mutations within the region from aa 690–989 show increased FRQ turnover and shorter periods [50,51,55]. In some of the mutants harboring mutations at certain residues, e.g., FRQ1 (with the FRQG482S mutation), FRQ3 (with the FRQE364K mutation), FRQ7 (with the FRQG259D mutation), FRQ2 (with the FRQ2895T mutation), and circadian periods are altered [56]. Among these mutants, FRQ7 is more stable than the wild-type control, while FRQ1 is less stable, which correlates with the circadian phenotypes [57].
FRQ has a large form of 989 aa (large FRQ, l-FRQ) and a small form of 890 aa (small FRQ, s-FRQ). l-FRQ has an extra 99 aa at the N’ terminus that is absent in s-FRQ. l-FRQ contains 132 serine and 56 threonine residues, 103 of which have been found to be phosphorylated in vivo or in vitro, and most of these sites undergo phosphorylation in a circadian fashion. Some of these sites are differentially phosphorylated between FRQ isoforms (Figure 2A) [50,51,52]. Baker et al. reported that an early-phase phosphorylated region can increase period length partially by stabilizing FRQ, while later phosphorylation of a physically separate region can decrease period length by decreasing FRQ stability [50]. In this scenario, the phosphorylation pattern of FRQ is similar to that of a drum in a music box, which has many convex dots, each presenting a note and touching the comb in a certain order to produce music (Figure 2B,C).
A strain exclusively expressing l-FRQ has a shorter period than a wild-type control; in contrast, a strain exclusively expressing s-FRQ has a longer circadian period [58]. The overall structure of l-FRQ is looser and more unstable than that of s-FRQ [52]. Similarly, loss of Cry1 in mice causes a shorter circadian period, while deletion of Cry2 causes a longer period [59].
CK1a is a major kinase involved in FRQ phosphorylation, but other kinases are also implicated in its phosphorylation. Neurospora FRQ harbors two CK1 binding regions, the FRQ-CK1 binding domain (FCD1, 319 aa–326 aa) and FCD2 (488 aa–495 aa), which may be rather close in the tertiary structure [55,60,61]. Deletion or mutation of the FCD1 and FCD2 regions abolishes the FRQ-CK1a interaction and CK1a-mediated FRQ phosphorylation [62]. Conidiation rhythmicity is abolished in a null strain of pkac-1, which encodes one of the catalytic subunits of cyclic AMP-dependent protein kinase A (PKA), and in a strain with a 19-bp deletion in the essential pkar gene, which encodes the regulatory subunit of PKA. PKA regulates the positive elements of the circadian clock by inhibiting the activities of WC proteins by serving as a priming kinase for casein kinases. In addition, PKA regulates FRQ phosphorylation and stabilizes FRQ [63]. These findings suggest that the crosstalk between different kinases plays a critical role in regulating FRQ phosphorylation and function.
Figure 2. The regulation of FRQ phosphorylation. (A) FRQ contains more than 100 phosphorylation sites, a substantial set of which are phosphorylated in a circadian fashion [50,61]. (B,C) Internal structure of a music box (B) and the detailed structure of the comb and drum (C).
Figure 2. The regulation of FRQ phosphorylation. (A) FRQ contains more than 100 phosphorylation sites, a substantial set of which are phosphorylated in a circadian fashion [50,61]. (B,C) Internal structure of a music box (B) and the detailed structure of the comb and drum (C).
Ijms 25 02574 g002
FRQ is an intrinsically disordered protein (IDP) that likely contains only a small amount of helical structure. FRH is a homolog of Saccharomyces cerevisiae Mtr4 [26,64]. FRH acts as a “nanny” molecule to stabilize the FRQ protein, although its RNA helicase activity is not necessary for circadian function. Disruption of the FRQ-FRH complex (FFC) interaction results in increased degradation of FRQ through the F-box and WD40 repeat-containing protein 1 (FWD-1)-dependent pathway and another unknown turnover pathway [25,27,65,66]. The highly unstructured N-terminal region (aa 1–150) of FRH is critical for FRQ-FRH association and FRQ stability [25,65].

3.2. Post-Translational Regulation of WC-1 and WC-2 Stability in Neurospora

The GATA-type zinc finger-containing WC-1 and WC-2 proteins are two PAS domain-containing transcription factors, of which WC-1 is a blue-light photoreceptor. WC-2 exhibits a constant expression pattern while WC-1 abundance oscillates with a low amplitude and a phase delayed by ∼6 h relative to that of FRQ [67,68]. WC-1 is less abundant than WC-2 and is the limiting factor of transcription factor capacity in the WCC [69]. WC-1 and WC-2 form heterodimeric complexes that activate frq transcription. The shuttling and activity of WCC are modulated by FRQ in a circadian fashion [28,70,71]. Multiple dispersed regions on FRQ, especially certain Gln/Glu residues in some of these regions, are necessary for its association with WCC. Deletion of these regions abolishes the interaction between FFC and WCC and decreases WCC levels. Mutations in some of these residues result in dramatically repressed FRQ levels owing to destabilization [49]. The WC-1/2 interaction is essential for maintaining the steady-state level of WC-1 and the function of WC proteins in the circadian clock and light responses [51].
The phosphorylation of WC proteins inhibits WCC as a transcription factor [55]. Despite their constant protein levels, the phosphorylation of WC-1/2 oscillates in a circadian manner [72,73]. WC-1 is phosphorylated through both FRQ-dependent and FRQ-independent pathways; the former is achieved by CK1 and CK2 recruited by FRQ, and the latter is achieved by additional kinases, e.g., PKA [74]. Phosphorylation of WC-1 at S885 and S887 by CK1 contributes to its accumulation, which is dependent on FRQ. FRQ promotes WC-2 expression at the transcriptional level [67,68,73]. Moreover, the abolition of the FRQ-FRH interaction leads to a decrease in WCC and increased WC-1 phosphorylation [27,75]. The abundance of nuclear FRQ is far less than that of nuclear WC-1; to close the negative feedback loop, the binding of FRQ to WC-1 in the nucleus leads to rapid phosphorylation and degradation of WC-1, which results in the inhibition of the transcriptional activity of WCC [76]. Negative feedback occurs in the nucleus, whereas positive support of WC-1 occurs in the cytosol, and positive and negative feedback loops are interlocked through FRQ-dependent phosphorylation of the WCC [28,73].
Mass spectrometry analysis revealed five in vivo phosphorylation sites in the WC-1 protein, including S988, S990, S992, S994, and S995. The degradation rate of the WC-1 protein did not significantly differ between a strain in which all five of these sites were mutated to alanine and the wild-type control, and mutations of S992 and S995 led to disruption of conidial circadian rhythms [77]. Sancar et al. identified S433 as the phosphorylation site on WC-2. In total, 80 and 15 phosphorylation sites were identified on WC1 and WC-2, respectively [78,79].
The mutual positive regulatory effects of FRQ, WC-1, and WC-2 are crucial for ensuring the robustness of circadian oscillation. Cytosolic FRQ proteins support WC-1 and WC-2 levels through different positive loops [28,70,71]. WC-1 is an inhibitor of wc-2 transcription, and WC-1 is stable only in complex with WC-2; FRQ likely supports the accumulation of wc-2 transcripts by repressing the activity of the WCC and may act indirectly as an activator of wc-2 [28,67,68]. Likewise, such mechanisms function in the mammalian circuits of mutual regulation between CLOCK/BMAL1 and Rev-erbα and in the regulation of vri by Clk/Cyc in Drosophila [80].
FRH plays multiple roles in regulating the properties of FRQ [25,26,27,64,65,75,81]. The FRHR806H mutation abolishes the FRQ-FRH interaction and circadian arrhythmicity; moreover, the protein levels of both WC-1 and WC-2 are repressed, and the degradation rate of WC-1 increases [75]. The loss of normal conformation and function due to the abolishment of the FRQ-FRH interaction may contribute to these alterations.
The CCR4–NOT complex is a conserved multisubunit functionally diverse machine that regulates gene expression [82]. NOT1 is a scaffold protein in this complex, and knockdown of NOT1 expression results in dramatically slowed growth and abolishment of circadian rhythms owing to decreased levels, hypophosphorylation, and increased degradation of WC proteins [74]. It is possible that NOT1 regulates WC degradation through association with other CCR4–NOT subunits, e.g., the E3 ubiquitin ligase CCR4 [82].

4. The Regulators of Clock Protein Stability in Different Organisms

PTMs are covalent, enzymatic, or nonenzymatic attachments of specific chemical groups to amino acid side chains that broadly control numerous important biological processes [83]. The fine control of protein stability is crucial for maintaining homeostasis and avoiding disease. The stability of core circadian clock components closely correlates with circadian homeostasis and determines circadian parameters. To date, a variety of different PTMs have been implicated in the control of the stability of circadian clock proteins [84].

4.1. Protein Degradation Pathways Implicated in the Circadian Clock

4.1.1. Subsubsection

Regulated proteolysis is essential for normal physiology. In general, cellular proteins are degraded by two distinct but connected pathways: the ubiquitin–proteasome system (UPS), which eliminates short-lived proteins and soluble misfolded proteins, and the lysosome system, which eliminates long-lived proteins, insoluble protein aggregates, entire organelles, macromolecular compounds, and intracellular parasites via endocytosis, phagocytosis, or autophagy pathways [85].
Ubiquitin ligases, including Ub activating enzymes (E1), E2, and E3, attach 76-aa ubiquitin proteins to target proteins sequentially [85]. F-box proteins are key elements of Skp-cullin-F-box protein (SCF) complexes, a class of E3 ubiquitin ligase complexes that facilitate the ubiquitination of proteins for proteasomal degradation [86]. In Drosophila, Slimb, an F-box/WD40 repeat-containing protein, is a component of the SCF (Skp1/Cullin/F-box protein) ubiquitin ligase (E3), which mediates the ubiquitylation of the phosphorylated Per protein [87,88]. PerS47 phosphorylated by double-time (Dbt) can be recognized and bound to Slimb, which decreases Per stability and negatively correlates with circadian period length [89].
Neurospora FWD-1 is a homolog of Slimb that interacts with and ubiquitinates phosphorylated FRQ. FWD-1 inactivation leads to hyperphosphorylated, stabilized FRQ, and arrhythmic conidiation [66,77]. Nonetheless, the rapid decrease in FRQ due to the abolishment of FRQ-FRH interaction is independent of FWD-1, suggesting that the degradation of misfolded FRQ proteins may be mediated by a different pathway [27].
The COP9 signalosome (CSN) is a conserved protein complex composed of eight subunits (designated CSN1-8) that functions in the ubiquitin–proteasome pathway. COP9 mediates the deneddylation of the cullin subunit, which is required for recycling of the SCFFWD−1 complex [90]. Moreover, the conidiation of the Neurospora knockout strains csn-1, csn-2, csn-4, csn-5, csn-6, and csn-7 is irregular. All these knockout strains showed dramatically lower expression of FWD-1 and several other E3 ubiquitin ligases [66,91]. In the csn-5 knockout strain, FRQ protein is stabilized by repressing the FWD-1 level [92]. The stability of the SCF complex is regulated by the COP9 signalosome, which, therefore, has an indirect yet important function in the circadian system [77].
Plant cellular protein degradation is modulated by proteasome activity on ubiquitinated proteins in the cytosol, the autophagy of cytosolic complexes and organelles through vacuole-based protein degradation machinery, and a network of organelle-localized proteases [93]. TOC1 and PRR5 are potential proteolytic substrates of the E3 ubiquitin ligase SCFZTL (ZEITLUPE, ZTL) within the PRR family. The circadian photoreceptor and F-box protein ZTL associates with Skip and Cullin to form the Skip–Cullin–F-Box (SCF) E3 ligase complex, which ubiquitinates and targets TOC1, PRR5, and CCA1 HIKING EXPEDITION (CHE) for degradation by the 26S proteasome [94,95]. ZTL contains an N-terminal light–oxygen–voltage sensing domain (LOV) for sensing blue light, in addition to an F-box domain and a Kelch-repeat domain. The F-box domain endows ZTL with activity as an E3 ubiquitin ligase [96,97]. GI is a plant-specific nuclear protein with diverse functions in multiple physiological processes, including flowering time regulation, light signaling, hypocotyl elongation, and circadian rhythm control. A protein–protein interaction (PPI) between GI and ZTL stabilizes ZTL, and this interaction can be enhanced by light through the amino-terminal flavin-binding LOV domain of ZTL [98]. Two additional F-boxes containing ZTL family members, flavin binding, Kelch Repeat, F-Box1 (FKF1), and LOV Kelch Protein 2 (LKP2), are also involved in the regulation of the plant circadian clock [99]. Moreover, EARLY FLOWERING 3 (ELF3) interacts with MUT9-LIKE KINASE4 (MLK4), which is a granulin repeat family cysteine protease, and an F-box family protein (AT2G16365), which may be implicated in regulating ELF4 turnover [100].
CONSTITUTIVE PHOTOMORPHOGENIC 1 (COP1) is a RING-type E3 ubiquitin ligase that represses flowering by promoting CO degradation at night. ELF3 interacts with and recruits COP1 to increase the degradation of GI, which controls developmental responses, including the timing of flowering and seasonal changes in day length [101]. Moreover, the photoreceptor phytochrome (phy) A is also recognized and ubiquitinated by COP1 for degradation [102].
Ubiquitination and its counteracting process, deubiquitination, together contribute to the control of the circadian clock. Ubiquitin-specific peptidases (BIPs) (also known as ubiquitin-binding peptidases, UBPs) are the largest class of deubiquitylating enzymes DUBs in plants. UBP12 and UBP13, two members regulated by the circadian clock, are associated with ZTL, and reduced expression of UBP12 and UBP13 results in a longer period of expression [99]. Mutations in ubiquitin carboxyl Hydrolases 1, 2, and 3 (UCH 1/2/3), which are also members of DUBs, lead to prolonged periods of LHY and GI expression at high temperatures, probably by mediating their degradation [103].
In mammals, the F-Box protein family comprises approximately 40 members that direct ubiquitination and proteasome-mediated degradation of target proteins through the Spk/Cullin/F-Box protein (SCF) E3 ubiquitin ligase complex [104,105,106]. β-Transducin repeat-containing protein (βTrCP), also known as Fbxw1 or Fbxw11, is a homolog of Slimb and targets PER for degradation in a manner dependent on CK1-dependent PER phosphorylation. Phosphorylation of PER2S478 by CK1 recruits β-TrCP and facilitates PER2 degradation. Moreover, PER2S478A knock-in mice show accumulated PER2 protein and longer locomotion periods [55,107,108]. In addition to SCFβ-TrCP, the ubiquitin ligase mouse double minute 2 homolog (MDM2) is involved in the control of PER2 degradation independent of PER2 phosphorylation [109].
Ubiquitin-binding enzyme 2C (UBE2C) facilitates PER1 degradation by promoting its ubiquitination via the E3 ligase S-phase kinase-associated protein 1 (SKP1)-cullin 1 (CUL1) [110]. The CRY1 and CRY2 proteins are ubiquitinated and degraded via the SCFFbxl3 ubiquitin ligase complex, which is a prerequisite for the reactivation of CLOCK-BMAL1 and the consequent expression of Per1 and Per2 [105]. Mutations in FBXL3, e.g., C358S and I364T, stabilize CRY1 by blocking its proteasomal degradation and lengthen the circadian period [104,105,106].
Two competing E3 ubiquitin ligases are engaged in the control of the circadian period by degrading CRY. Loss of FBXL3 function leads to period lengthening, while mutation of F-box and leucine rich repeat protein 21 (Fbxl21) causes period shortening. FBXL21 plays dual roles: it protects CRY from FBXL3 degradation in the nucleus and promotes CRY degradation within the cytoplasm [111]. In the nucleus, FBXL21 counteracts the destabilizing effects of FBXL3 by ubiquitinating mouse CRY proteins and protecting them from FBXL3-dependent degradation [112].
REV-ERBα can be ubiquitinated and subsequently degraded by the F-box protein FBXW7; therefore, FBXW7 enhances clock gene transcription by repressing the inhibitory effects of REV-ERBα. Disruption of FBXW7 in the liver results in altered expression of core clock genes and perturbation of whole-body lipid and glucose metabolism. Phosphorylation of REV-ERBαT275A by cyclin-dependent kinase 1 (CKD1) is necessary for its degradation by FBXW7. CDK1-mediated phosphorylation of REV-ERBα oscillates and contributes to rhythmic changes in REV-ERBα protein levels [113].
REGγ/PA28γ (encoded by the PSME3 gene) binds and activates the proteasome to promote ubiquitin- and ATP-independent cleavage of intact proteins and mediates protein degradation through a pathway different from the canonical ubiquitin–proteasome system [114]. REGγ KO mice exhibited shorter locomotion periods in constant darkness. Silencing or deficiency of REGγ caused elevated mRNA levels of Per1, Per2, Cry1, Clock, Bmal1, and Rorα and elevated protein levels of PER2 and BMAL1. REGγ directly promotes the proteasome-dependent degradation of BMAL1 protein, and GSK-3β can repress this process [115].

4.1.2. Degradation of Circadian Clock Proteins Mediated by Chaperone-Mediated Autophagy

Cell homeostasis is dictated by the balance of the opposing biosynthesis and turnover. The lysosome is a cellular organelle for the degradation of target proteins with a wide array of resident acid hydrolases. In chaperone-mediated autophagy (CMA), cargo containing a peptide signal (KFERQ) binds with a cytosolic chaperone HSC70 and docks on the lysosomal surface receptor LAMP2A. HSCs facilitate the translocation of cargo into the lysosomal lumen [116].
CMA contributes to the rhythmic removal of clock machinery proteins, and disruption of lysosome-associated membrane protein type 2A (L2A) results in a phase shift and amplitude change in molecular rhythms. Conversely, circadian disruption abolishes the rhythms of CMA component expression and function. In addition to the protein turnover of clock components, the crosstalk between CMA and the circadian clock governs the oscillation of cellular proteome [117].
Macroautophagy is a highly conserved catabolic pathway responsible for quality control and degrades cytoplasmic contents in lysosomes. The mammalian circadian proteins BMAL1, CLOCK, REV-ERBα, and CRY1 are lysosomal targets, and macroautophagy affects the circadian clock by selectively degrading CRY1. CRY1 interacts with light chain 3 (LC3) through its LC3-interacting regions (LIRs) in autophagy. Two of the LIRs (285SLYGQL290 and 492SRYRGL497) are implicated in the regulation of CRY1 degradation and circadian control of hepatic gluconeogenesis [118]. BMAL1 inhibits ferroptosis through repressing Egln2 transcription and consequently activating HIF1A [119]. CMA mediates the direct transport of substrate proteins containing the signaling peptide KFERQ to lysosomes for degradation [120]. A number of circadian proteins, e.g., BMAL1, CRY1, and PER1, contain the KFERQ signal and are degraded by CMA in lysosomes during the light or dark cycle, which is dependent on the canonical CMA pathway [117]. These findings demonstrate that several pathways are involved in the efficient removal of circadian proteins, which may function in a coordinated manner [116].
ClockΔ19 mice bearing a deletion of 51 amino acids in the CLOCK protein exhibit a very long periodicity of locomotion activity and a low amplitude under a constant environment [121]. The autophagy adaptor protein p62/SQSTM1 regulates CLOCK-induced BMAL1 degradation through autophagy, and CLOCKΔ19 mutation leads to stabilization of BMAL1 through attenuation of the proteasomal and autophagic degradation pathways [122].
In bacteria, the ubiquitin–proteasome system and lysosome are present and participate in the regulation of multiple aspects of infection and host defense in certain orders of bacteria [123,124,125]. However, the functions of these factors in circadian clocks have barely been investigated.

4.2. Dynamic Regulation of the Phosphorylation State of Circadian Clock Components

Dynamic phosphorylation of circadian clock proteins on a timescale is crucial for circadian timekeeping and regulates circadian rhythms through versatile roles ranging from regulating degradation to complex formation, subcellular localization, shuttling dynamics between the cytoplasm and nucleus, and activity [47,70,76,126,127,128].
All of the known circadian clock proteins are phosphoproteins [127]. Among the post-translational regulators, a set of protein kinases and phosphatases including CK1, CK2 and a calcium/calmodulin-dependent kinase (CAMK-1), PKA, checkpoint kinase 2 (Chk2), SHAGGY/GSK-3, cyclin-dependent kinase 1 (CDK1), and the phosphatases PP1, PP2, and PP4, are involved in the regulation of circadian clock [113,126,129,130,131,132,133,134,135].
CK1 and CK2 are also designated CKI and CKII, respectively, in some studies [28,60,136,137]. The CK1 family comprises a large number of conserved kinases in fungi, algae, and animals [138]. Mammals have six CK1 genes that encode CK1α, δ, ε, ɣ1, ɣ2, and ɣ3, along with several splice variants. CK1a forms homodimers, and its activity is regulated in multiple ways, including through inhibition of autophosphorylation and subcellular distribution [54,139,140].
The Drosophila core clock components, e.g., PER, are progressively phosphorylated and eventually degraded in a circadian manner to close the feedback loop [141,142]. PER has 25–30 phosphorylation sites, many of which are distributed in clusters [89,143]. The PerLong (PerL) mutant bears a V243D mutation with a period of 29 h; the PerShort (perS) fly mutant, which bears an S589N mutation, shows a ~19 h short period of locomotor rhythm. The Per0 mutant bears a nonsense mutation at Q464, which is arrhythmic [144,145,146].
Similarly, a number of Drosophila kinases, including Dbt [141,147,148], Casein kinase 2 (Ck2) [149,150,151], Shaggy (Sgg/GSK3β) [152,153], and Nemo-like kinase (Nlk) [154,155], have been shown to modulate PER phosphorylation. Drosophila Dbt, an ortholog of human CK1ε, phosphorylates Per, which reduces its stability and the level of Per proteins. The dbtS mutants exhibited an 18.0 h short period, while the dbtL mutants exhibited a period of 26.8 h in constant darkness. The overt circadian rhythms are eliminated in dbtP mutants in which the expression of Dbt is abolished [141,147]. Dbt stably binds and phosphorylates PER and facilitates Per proteasome-dependent degradation in both the cytoplasm and nucleus, and phosphorylation by Dbt also controls the nuclear entry of Per [156,157,158]. The K38R mutation of Dbt still interacts with Per, but it lacks kinase activity, which causes an increased Per stability and a long circadian period [159].
In Neurospora, treatment with 6-dimethylaminopurine (6-DMAP), a general kinase inhibitor, blocks FRQ phosphorylation and stabilizes FRQ, suggesting the general effects of phosphorylation on the Neurospora circadian clock [160]. Neurospora Ck1a, a homolog of mammalian CK1ε/δ and Drosophila Dbt, is essential for viability and has a function in the circadian clock [10,108,161]. The FRQ-CK1a interaction is dynamic, and the affinity between these proteins decreases with increasing FRQ phosphorylation [55,81]. FRQ proteins are phosphorylated immediately upon synthesis [55]. The regulation of CK1 is also important for the control of the circadian clock. PERIOD 2 (PRD-2) participates in the regulation of the Neurospora circadian clock by stabilizing the ck-1a transcript, and the mutation of prd-2 caused recessive inheritance of a long-period phenotype of ~26 h [162].
In Arabidopsis, the CASEIN KINASE 1 LIKE family (CKL) contains 13 CK1-like kinases. The functions of CKL kinases in regulating the circadian clock have not been fully elucidated compared to those of CK1 kinases in fungi and animals. Inhibition or knockdown of CKL lengthens the circadian period, likely through stabilizing PRR5 and TOC1 [138]. In addition, Arabidopsis has four MUT9-LIKE KINASEs (MLK, also known as PHOTOREGULATORY PROTEIN KINASE, PPK), which are close homologs of CKL members and interact with clock components [163].
CKIε is the mammalian homolog of Drosophila Dbt [164]. The semidominant autosomal tau mutation (CK1εR178C) is a gain-of-function mutation in Syrian hamsters that causes significant shortening of activity cycles to 20 h in both SCN and peripheral tissues through the specific destabilization of nuclear and cytoplasmic PER proteins [46,165,166,167]. The CK1εtau mutation promotes the degradation of both nuclear and cytoplasmic PERIOD but not of CRYPTOCHROME [168]. Nucleocytoplasmic partitioning of core clock components is essential for the proper orchestration and function of the circadian system [94]. In mice expressing mutated Per2 that cannot interact with CK1, the phosphorylation of CLOCK is not induced. In these mice, PER proteins are stabilized and arrhythmically expressed, and the negative feedback loop is impaired [169].
The eukaryotic CK2 holoenzyme is an α2β2 heterotetramer [170], and cka is the only gene encoding the α subunit of CK2 in Neurospora. In the Neurospora ckaRIP strain, in which CK2α is inactivated, FRQ is mostly hypophosphorylated, and its level is higher than that of the wild type. This strain shows dramatically slow growth and abolished rhythmicities in conidiation and frq/FRQ expression [137]. The phosphorylation is decreased, and FRQ degradation is decreased in this strain [171]. CK2 also contributes to the inactivation of the WCC and the regulation of temperature compensation [55,60,172,173].
Mammalian CK2 activity is involved in enhancing circadian amplitude and controlling the circadian period. CK2 binds with and phosphorylates PER2, and collaborates with CK1ε to promote PER2 degradation [174]. CK2α phosphorylates BMAL1 by maintaining the nuclear portion of BMAL1 [175]. CK2 subunits, including the regulatory subunits CKB3 and CKB4, play a critical role in the Arabidopsis circadian clock, and overexpression of these proteins leads to shorter circadian periods [176,177,178]. Arabidopsis CK2 interacts with LHY, CCA1, and phosphorylates CCA1 [179]. Arabidopsis CKB3 is associated with CCA1, and CKB3 overexpression leads to a shorter period [177].
In Drosophila, shaggy (sgg) encodes glycogen synthase kinase-3 (GSK-3), which promotes the nuclear transfer of PER/TIM complex through phosphorylating TIM. Elevation of sgg expression results in a shortened period length; in contrast, inhibition of GSK-3β leads to a shorter period in cultured mammalian cells [152,180].
Mammalian GSK-3β plays different roles in controlling the expression and function of clock proteins through the control of phosphorylation and degradation. GSK3βphosphorylates BMAL1 specifically on S17 and T21 and primes it for ubiquitylation. Inhibition of GSK3β activity enhances BMAL1 stability in the nucleus [181]. GSK3β phosphorylates BMAL1 specifically on S17 and T21, which primes it for ubiquitylation and proteasomal degradation, and inhibition of GSK3β leads to a shortened period length in mammalian cells [180,181]. Nucleocytoplasmic shuttling of BMAL1 plays an essential role in facilitating the accumulation of CLOCK in the nucleus and the degradation of the CLOCK/BMAL1 heterodimer. BMAL1 and CLOCK reciprocally induce each other’s turnover, while the binding of CRY stabilizes the CLOCK/BMAL1 complex. Transcriptional activation of CLOCK/BMAL1 is tightly associated with its turnover [182]. In addition, GSK3β promotes CRY2 degradation and the stabilization of Rev-erbα [183,184].
Spengler et al. identified a conserved phosphodegron containing a cluster of phosphorylation sites in CLOCK, ranging from aa 425 to 461. S431 in this region is a prerequisite phosphorylation site for the generation of BMAL-dependent phosphor-primed CLOCK and for potential GSK-3 phosphorylation at S427. Phosphorylation of these sites destabilized CLOCK proteins [185].
Piwi-like RNA-mediated gene silencing 1 (PIWIL1) and PIWIL2 participate in stem cell self-renewal, spermatogenesis, and RNA silencing. PIWIL1 and PIWIL2 play similar roles in activating the PI3K-AKT signaling pathway to phosphorylate and inactivate GSK3β and promote the stability of CLOCK and BMAL1 by repressing GSK3β-induced phosphorylation and ubiquitination of these two proteins in cancer cells. Mammalian GSK-3β is expressed in the suprachiasmatic nucleus and liver of mice and in NIH3T3 cells, and GSK-3β phosphorylation results in robust circadian oscillation [51,186].
Hyperphosphorylated PER has a more open structure and is more amenable to further global phosphorylation and degradation. In Drosophila, Dbt-dependent phosphorylation of PER at the N’ terminus promotes its binding to the F-box protein Slimb (β-TrCP), resulting in proteasome-dependent PER degradation [89].
In addition to its impact on the function of the circadian oscillator, the phosphorylation of clock proteins also modulates the expression of circadian output genes. The mTOR effector kinase, ribosomal S6 protein kinase 1 (S6K1), an important regulator of translation, rhythmically phosphorylates BMAL1. S6K1 phosphorylates BMAL1 at S42, and the phosphorylation at BMAL1S42G stabilizes BMAL1 and increases its cytosolic distribution of BMAL1. S6K1 phosphorylation at BMAL1 contributes to the association with the translational machinery and stimulates protein translation of a set of genes [187].
The mammalian DNA-dependent protein kinase catalytic subunit (DNA-PKcs) interacts with CRY. Both repression of this kinase and phosphorylation at CRY1S588 by DNA-PKcs result in circadian rhythms with abnormally long periods, possibly through stabilization of the nuclear abundance and prevention of the FBXL3-mediated degradation of CRY1 proteins [188]. A set of phosphorylation sites in different circadian clock components and their associated circadian phenotypes in various organisms are listed in a review article [189].
In Neurospora, disruption of ckb1 (encoding the catalytic subunit gene of CK2) results in hypophosphorylation and increased levels of the FRQ protein by decreasing its stability. The circadian rhythms of conidiation and the oscillation of FRQ occur over a long period with a decreased amplitude [137,171].
In addition to CK1, other protein kinases can be activated in response to external or internal cues and are implicated in the control of the circadian clock, although their effects on the circadian period are limited compared to CK1 [55]. In Neurospora and mammals, Chk2/PRD-4 interact and phosphorylate FRQ or PER1, respectively, in response to DNA damage and translational stress, triggering phosphorylation and destabilization [190,191]. Mammalian nutrient-responsive adenosine monophosphate-activated protein kinase (AMPK) phosphorylates and destabilizes CRY1. Treatment with the AMPK agonist aminoimidazole carboxamide ribonucleotide (AICAR) lengthened the circadian period and repressed the amplitude [192]. Other protein kinases are also involved in the regulation of circadian rhythms; however, their direct influences on the stability of circadian clock components remain unclear. In Neurospora, Ca/CaM-dependent kinase 1 (CAMK-1) and protein kinase A (PKA) may also regulate the circadian clock, although whether they phosphorylate FRQ in vivo remains undetermined [63,193]. It has been proposed that the phosphorylation of FRQ and WCC by kinases other than CK1 is not essential for the principal functions of the core oscillator [55]. In mice, cyclin-dependent kinase 5 (CDK5) phosphorylates CLOCK at residues T451 and T461, which accounts for the regulation of CLOCK stability and the cytoplasmic–nuclear distribution of CLOCK [194].
As a counteracting process, protein phosphatases also regulate the phosphorylation of circadian proteins. In Neurospora, the eukaryotic serine/threonine protein phosphatases PP1 (protein phosphatase 1), PP2A, and PP4 are involved in the circadian clock. PP1 and PP2A can dephosphorylate FRQ in vitro. The Neurospora ppp-1 (catalytic subunit of PP1) mutant shows an advanced phase and a short period, and FRQ protein is unstable in this strain. In the rgb-1 (a regulatory subunit of PP2A) mutant, FRQ stability is not affected, but the frq mRNA and FRQ protein levels are very low. This mutant has a low amplitude and a long period [195], suggesting that PP2A may regulate the circadian period through a pathway without direct impacts on FRQ stability.
Disruption of PP4 in Neurospora leads to a decreased stability of FRQ, a low amplitude, and a short period. Moreover, PP4 is required for the nuclear enrichment of WCC [47]. Consistently, in mice and flies, silencing of PP4R2, the regulatory subunit of PP4, shortens the circadian period, while overexpression of PP4R2 lengthens the circadian period [196]. Drosophila PERS47 is phosphorylated and destabilized by CKIδ and DBT in vitro, whereas PP2A (and/or PP1) stabilizes PER at least in part by decreasing the phospho-occupancy of PERS47 [89].

4.3. Preference for Genetic Codon Usage

Different organisms display different preferences for genetic codon usage; in one species, different genes exhibit genetic codon bias [197]. Intriguingly, the core circadian genes in the circadian systems exhibit overt codon bias in cyanobacteria, Neurospora, and Drosophila. The biased codon usage in circadian genes may be associated with specifically controlled translation rates [198,199,200,201,202].
Neurospora frq and Drosophila Per show dramatic preference for low-use codons, and their extensive optimization severely compromises or abolishes the circadian rhythms. Codon optimization causes less robust functions in promoting WC-1 and WC-2 levels in the circadian positive limb and changes in the stability and accessibility of the structure of the FRQ protein [199,202]. Similarly, the optimized codon usage of Drosophila Per leads to changes in the conformation, stability, and nuclear localization of Per in the pigment dispersing factor (PDF)-positive neurons in the brain [200]. Importantly, as many key circadian components across species are IDPs or contain large parts of IDP regions, regulation by phase separation may be an essential and conserved characteristic of circadian systems [25,203,204,205].

4.4. Additional Post-Translational Control Factors

Circadian clock components undergo complex PTMs; for instance, the expression and function of BMAL1 are mediated by a multitude of PTMs, including acetylation, phosphorylation, and protein instability by kinases, sumoylation, and ubiquitylation [115]. Many additional PTMs, e.g., acetylation and glycosylation, also regulate the stability of circadian clock proteins. The K560R, K576R, and K591R on CRY2 mutations abolish the acetylation of these three sites, and the mutated proteins are less stable. In contrast, in a mutant mimicking constant acetylation at these sites, the CRY protein showed an increased stability [206].
Covalent O-GlcNAcylation (O-GlcNAc) modifies the hydroxyl groups of Ser/Thr residues on proteins in addition to phosphorylation and exerts diverse functions, including protein–protein interactions, protein turnover, and subcellular localization [207]. Drosophila PER contains at least 6 O-GlcNAcylation sites, which undergo daily changes in O-GlcNAcylation. Furthermore, O-GlcNAcylation of PER leads to its stabilization [208,209]. O-GlcNAcylation at PERS942 attenuates the interaction between PER and CLK [209]. Inhibition of the O-GlcNAcylation pathway increases the period length in mammalian cells. The positive elements, BMAL1 and CLOCK, are rhythmically O-GlcNAcylated, which stabilizes BMAL1 and CLOCK by inhibiting their ubiquitination [210]. Phosphorylation and O-GlcNAcylation may compete on the same sites; for instance, the O-GlcNAc transferase and GSK3β have reciprocal regulations [211,212].
Overt circadian rhythms are absent in the plateau black-lipped pika. Epas1/HIF-2α is a transcription factor involved in the induction of genes regulated by oxygen. Epas-1 binds with BMAL1 at the E-box locus to initiate transcriptional activation. Pl-Pika Epas1 has insertion mutations between exon 14 and exon 15, which disrupts an RNA splicing site. Under simulated hypoxic conditions, Pl-Pika-specific L-Epas1 proteins are significantly more stable than S-Epas1 present in other species, resulting in interference with circadian circuits and arrhythmic phenotype. L-Epas1-induced loss of circadian rhythms and enhancement against heart damage may be due to adaptations to the plateau environment [213].
Importantly, PPI between circadian components contributes to protein turnover. Mammalian PER3 interacts with PER1/2 and promotes nuclear translocation and stability [214,215]. Drosophila CLK acts to stabilize CYC, and loss of CLK results in low CYC levels [216]. In Neurospora, binding with WC-2 can stabilize WC-1 [69]. In Neurospora, the activated WCC as a transcription factor and LOV domain-containing protein VIVID (VVD) competes with the dimerization of the WCC, thereby reducing its activity and stabilizing WC proteins in photoadaptation [217]. The Drosophila scaffold protein, suppressor of Ras (SUR-8), interacts with PP1-87B to stabilize PER, and depletion of SUR-8 causes lengthening of the circadian period by approximately 2 h and delays TIM nuclear entry [218]. Moreover, the circadian clock regulates the stability of downstream proteins. For instance, the clock protein REV-ERBα directly interacts with O-GlcNAc transferase (OGT) and stabilizes OGT in different cellular compartments as the cellular localization of REV-ERBα oscillates [219].
Light exposure affects PTMs and the degradation of circadian proteins. In Drosophila, the newly translated PER proteins are required to maintain their stability and that of TIM proteins. Moreover, exposure to light increases TIM degradation [220,221,222,223]. Similarly, light exposure induces the accumulation of hyperphosphorylated WC-1 in an FRQ-independent manner in Neurospora [224]. In Arabidopsis, nuclear CO protein is degraded in the morning or in the dark, but light exposure stabilizes CO in the evening [225]. Arabidopsis CRYPTOCHROME 2 (CRY2) regulates light regulation of seedling development and photoperiodic flowering. CRY2 remains unphosphorylated, inactive, and stable in the absence of light; in contrast, blue light induces the phosphorylation of CRY2, photomorphogenic responses, and eventually degradation [226]. The Arabidopsis core clock protein GI associates with ZTL in a blue light-dependent manner, and the two proteins reciprocally stabilize each other. Their subsequent dissociation in darkness provides precise input to the photoperiod [98]. Ubiquitination also regulates light-induced responses through the photoreceptors CRYPTOCHROME 1 (CRY1) and CRY2. Upon light exposure, CRY1/2 proteins bind and activate CONSTITUTIVE PHOTOMORPHOGENESIS 1 (COP1), a RING domain E3 ligase. COP1, in conjunction with ELF3, destabilizes GI and tags it for degradation [227]. The day length also affects the stability of clock proteins (e.g., PRR5 degrades faster at LL19 and more slowly at LL7) [228].
Among the regulators of circadian clock protein stability, circadian regulation is involved in the expression, subcellular localization, or function of some of the regulators. For instance, mammalian CKIϵ and CKIδ and their interactions with circadian components show remarkable circadian rhythms in the nucleus despite the total protein levels being constant [148,229]. Consistently, the subcellular localization of Drosophila DBT in the nucleus but not the total DBT level oscillates in photoreceptor cells and lateral neurons [148]. The expression of plant CK2 is also under circadian control [178]. In the mouse liver, the activity and nuclear localization of AMPK oscillates, which is inversely correlated with CRY1 nuclear protein abundance [192]. The expression and phosphorylation of GSK-3β are rhythmically controlled robust circadian oscillations [129,230].
The expression of factors directly associated with protein decay, including some E3 ligases and the activity of the CMA-autophagy complex, displays overt circadian rhythms in a tissue-specific manner [117,160]. The expression of the Arabidopsis F-box protein ZTL is rhythmic, which controls the proteasome-dependent degradation of a central clock protein, TIMING OF CAB EXPRESSION 1 (TOC1), and is necessary for sustaining the circadian cycles [98,231,232]. The with-no-lysine (K) protein kinase (WNK) family members WNK1, WNK2, WNK4, and WNK6 are expressed in plants [233]. In the rat SCN, WNK3 binds with PER1 to regulate the circadian rhythms [234]. Drosophila TWINS (TWS), a PP2A regulatory subunit, is under circadian control [235].
Changes in clock protein stability and function have been studied only marginally in prokaryotes. In cyanobacteria, both the translation and degradation of KaiC are under circadian control. The synthesis rate of the KaiC protein reaches a peak around 12 h after exposure to LL (LL12), with a 6 h delay compared to the total KaiC level; the trough of degradation of KaiC occurs approximately 16 h after exposure to LL [236].

5. The Correlation of Clock Protein Decay with the Circadian Parameters

5.1. Fine Control of the Circadian Period

Accumulating evidence has demonstrated that PTMs, especially phosphorylation, play a very important role in controlling the circadian period. A set of protein kinases, including CK1, CK2, and protein kinase A, phosphorylate circadian components to regulate the stability of circadian elements and circadian period length [107,161,170,237]. It has long been proposed that the stability of FRQ protein is tightly correlated with the length of the circadian period [170,238]. Supportively, the mutant FRQ in the frq7 mutant was degraded approximately twice as slowly as the corresponding wild-type control [237]. l-FRQ shows a faster degradation rate than s-FRQ, and consistently, the degradation period of the l-frq strain is shorter than that of the s-frq strain [52,58,239,240]. The Neurospora ck-1aL strain carrying a mutation equivalent to that of the Drosophila dbtL strain exhibited a ~32 h period due to increased FRQ stability. In this strain, the FRQ, WC-1, and WC-2 proteins undergo hypophosphorylation [60]. 6-DMAP treatment blocks FRQ phosphorylation, results in FRQ stabilization, and prolongs the period of the clock in a dose-dependent manner. The FRQS513 mutation leads to a reduction in the FRQ degradation rate and a very long period (>30 h) [170].
Mutations in circadian negative components and CK1 in fungi and animals stabilizing these components lead to longer periods, and contrarily, those mutations destabilizing the circadian negative components lead to shorter periods [241]. The phosphotimer model has been proposed to explain the circadian clocks in multiple model organisms, in which a sequential series of phosphorylation events are regulated by multiple kinases at different sites over the course of a day. These phosphorylated sites interact and synergistically determine the stability of clock proteins and the length of the circadian period [107,143,242]. Phosphorylation regulates period length mainly through its effects on stability, together with other post-translational modifications. Protein degradation-associated factors are also determinants of circadian period length. For instance, ubiquitin ligase mouse double minute 2 (MDM2)-mediated PER2 turnover negatively correlates with the circadian period length in mammalian cells [109].
However, there are a variety of exceptions to the phosphorylation-stability-period model. Mutation of TIMELESSR1081 to a stop codon leads to stabilization of TIM, and mice bearing this mutant show a phase advance of sleep-wake behavior but no significant change in period length [243]. Different phosphorylation modifications multidirectionally contribute to the circadian period length, some of which accelerate the circadian pace, some deaccelerate, and some of which show no overt effects on the circadian pace. Phosphorylation on different regions on FRQ leads to opposing effects on circadian rhythms and FRQ stability, of which mutations of the phosphorylation sites at the N-terminal part result in lengthened periods through increasing FRQ stability although mutations of the C-terminal sites result in shorter periods due to decreased FRQ stability [50,51]. The FRQ stability does not always correlate with the circadian period length. The FRQT781A mutation confers a slower degradation of FRQ; however, there is no detectable change in the period length of circadian conidiation [52]. In addition, in contrast to most phosphorylation of FRQ by CK1 and CK2, FRQ phosphorylation by PKA increases FRQ stability [63].
In Drosophila, a number of hyperphosphorylation modifications are unrelated to PER stability, e.g., S48 modulates PER stability in a manner independent of phosphorylation at this site [89]. The PERS942A mutant degraded at a similar rate as wild-type control in the presence of O-GlcNAc transferase, demonstrating that the S942A mutation has no significant effect on PER stability [209].
In circadian oscillators, phosphorylation has functions in addition to regulating the stability of circadian clock components, and these functions are closely interconnected [51]. For instance, the degradation of GSK-3β, a known promoter for BMAL1 degradation, exacerbates the effects of REGγ depletion in addition to its impacts on BMAL1 phosphorylation and control of its stability [115]. Dbt-dependent phosphorylation of Drosophila PERS826,S828, two sites located in a putative nuclear localization signal (NLS), can be rapidly induced by light and blocked by TIM. However, mutation of these two sites causes only mild changes in behavioral rhythms with a slightly longer period (by 1-h), suggesting that phosphorylation of these two sites may play an unknown role rather than controlling the circadian period [244]. Notably, some of the phosphorylation events show no impact on the circadian period, likely because some phosphodegrons only function as entire phosphorylated domains rather than as partially modified amino acids [53].
In Neurospora fwd-1 and csn-2 mutants, the circadian oscillator and output are decoupled, although the TTFL circuit is still functional, and the total clearance of the negative element is not an essential step. Despite the lack of rhythmicity in the bulk FRQ proteins and conidiation, robust rhythmicity of FRQ-LUC expression, which represents protein synthesis, persists in these strains. These findings suggest that FRQ stability is a correlative measurement but not a determinant of period length [53]. In some Neurospora strains, the FRQ-CK1 interaction, rather than FRQ stability, is correlated with the circadian period. A Neurospora strain with mutations in the Pro/Glu, (Asp)/Ser/Thr-rich sequence 1 (PEST-1) domain (M10) exhibits an extra long circadian period (32.1 h), although the turnover rate of FRQ is comparable to the wild-type control. Mutants with mutations in the FCD region, including FRQQ325N, FRQQ494N, FRQV320I, and FRQL488V, show overt associations between period length and the FRQ-CK1a interaction in these mutants [242]. These findings suggest that the FRQ-CK1 interaction plays more roles than just controlling the phosphorylation and stability of FRQ and that it modulates the circadian period through an undetermined mechanism rather than through a phosphorylation-stability-period circuit. Given that the FRQ-CK1 association is necessary for the hyperphosphorylation and inactivation of the WCC, it is likely that the altered WCC function may also account for the period length in these strains [60].
The stability of clock proteins closely correlates with the period length only within a certain range, and it may be decoupled. Based on the available knowledge, we infer that the stability of circadian proteins regulated by CK1 is the predominant determinant of circadian period length; however, the CK1-FRQ interaction also induces changes in the functions of other circadian clock components, which collectively participate in the determination of period length (Figure 3). For instance, it is possible that the PER2S478A mutation may exert multiple functions more than the removal of phosphorylation at this site and influence the stability of PER2, e.g., affected subcellular distribution of PER1, CRY1, and CRY2 [107], which leads to superimposed effects on the circadian period.
Figure 3. Numerous factors are involved in controlling the phosphorylation state, the stability of circadian components, and the length of circadian period at multiple layers. Protein homeostasis is one of the key regulatory mechanisms. There are many determinants of period length, among which CK1 is the master regulatory kinase of circadian clocks in different species [241]. The ellipses denote unlisted and unidentified factors or functions. PPI, nucleus–cytoplasm shuttling, and other regulatory mechanisms may also affect period length by modulating the PTMs and the turnover of clock proteins. The dashed line denotes the possible regulatory circuit to be confirmed.
Figure 3. Numerous factors are involved in controlling the phosphorylation state, the stability of circadian components, and the length of circadian period at multiple layers. Protein homeostasis is one of the key regulatory mechanisms. There are many determinants of period length, among which CK1 is the master regulatory kinase of circadian clocks in different species [241]. The ellipses denote unlisted and unidentified factors or functions. PPI, nucleus–cytoplasm shuttling, and other regulatory mechanisms may also affect period length by modulating the PTMs and the turnover of clock proteins. The dashed line denotes the possible regulatory circuit to be confirmed.
Ijms 25 02574 g003
Cyanobacterial core circadian components, KaiABC, dynamically interact with each other and can generate sustainable rhythms of KaiC phosphorylation in vitro [23]. Interestingly, the dynamics of the interaction of KaiA, KaiB, and KaiC determines the circadian period and amplitude of the in vitro oscillation [20], which demonstrates the importance of PPIs in determining the circadian period in prokaryotes.

5.2. Regulation of Circadian Temperature Compensation

Temperature compensation is a property of circadian rhythms that buffers the circadian period against the ambient temperature changes. An altered period length is usually associated with an abnormality in temperature compensation [238]. Hastings and Sweeney proposed that circadian temperature compensation may be a network mechanism arising from multiple counteracting reactions in response to temperature changes that cancel out their opposing effects on the period length [155,245].
The Neurospora mutants of ck2β1 and ck2α exhibit abnormal temperature compensation features compared to those of the wild-type strain. Mutations in the CK2 targeting sites in FRQ phenocopied the deficient temperature compensation in the ck2β1 and ck2α mutants. These data demonstrate that CK2 is involved in temperature compensation [172]. The Neurospora strain with a mutation in prd-4/chk2 has a short period and defective temperature compensation, but the turnover rate of FRQ has not been determined [246]. In Neurospora, prediction with the Goodwin model suggested that the stability of FRQ should determine the length of the circadian period as well as temperature compensation [57]. Moreover, the CK1-FRQ interaction not only determines the circadian period but also contributes to temperature compensation [62].
The degradation rate of PER2 regulated by CK1ε/δ-dependent phosphorylation is temperature-insensitive in living clock cells but sensitive to chemical perturbations [247]. Zhou et al. proposed that some phosphorylation sites in PER2 may compete with each other to promote or repress PER2 stability. This regulation is temperature sensitive, i.e., PER2 degradation decreases at higher temperatures and increases at lower temperatures [248].
The circadian period length closely correlates with its phase [249], and some examples are presented here. Inhibition of CK1ε by PF-670462 treatment in mammalian cells results in a lengthened period and a delay in the circadian phase [250]. CK1 phosphorylation of PER2S478 recruits the ubiquitin E3 ligase β-TrCP, which leads to accelerated PER2 degradation. Moreover, PER2S478A mutant mice exhibited a longer period and a delayed phase in the liver [107]. CAMK-1 also affects the circadian phase in Neurospora, leading to a delayed phase but with no significant effects on the circadian period [193]. The Neurospora ck-1aL strain also exhibited a significant phase delay in conidial rhythms and the FRQ progressive phosphorylation as well [60]. The mammalian protein kinase C (PKC) family is involved in regulating circadian phases. The neuron-specific protein PKCγ participates in food entrainment of circadian rhythms by partially stabilizing and reducing the ubiquitylation of BMAL1 [251]. GSK-3β overexpression advances the phase of clock gene expression; in contrast, inhibition of GSK-3β causes a delay in the circadian phase [129]. Among the properties of circadian rhythms, we emphasize the circadian period, while the circadian phase and temperature compensation are only briefly mentioned due to space limitations.

6. Abnormal Turnover of Circadian Clock Proteins and Disorders

Desynchronization of the circadian clock compromises fitness to environmental cues [2,252], and in humans, circadian misalignment has been associated with numerous disorders, e.g., sleep and mood disorders [253]. Misalignment in the circadian clock is usually correlated with sleep disturbance, and it has been demonstrated that some of the core clock genes (PER2, PER3, Casein kinase Iδ, CRY2, DEC2) affect the timing of sleep-wake behavior or the sleep duration in humans. Patients with familial advanced sleep phase syndrome (FASPS) exhibit a dramatic early sleep phase that occurs 4–5 h earlier than in unaffected individuals. The PER2S659G and PER2S662G mutations are associated with FASPS, which promotes the degradation of PER2. The accelerated degradation of PER2 caused by PER2S569G and PER2S662A mutations resulted in shorter periods [254,255,256,257], which is consistent with the positive relationship between circadian period length and phase [258]. The PER3P415A/H417R variant, which is associated with FASPS and seasonal mood traits, destabilizes PER3 and PER1/2 [215]. Flavin adenine dinucleotide (FAD) is a mediator of the circadian clock through stabilizing CRY proteins, and CRY2A260T mutation, a missense mutation localized in the (FAD) binding pocket, causes an advanced sleep phase in humans. The turnover rate of CRY2A260T is significantly faster in both the nucleus and the cytosol [259,260].
The expression of a substantial set of genes associated with disease or drug development is under circadian control [8]. Given the important roles of protein kinases in the circadian clock, it is not surprising that pharmacological attempts to manipulate circadian rhythms have identified many molecules targeting these kinases. Hirota et al. identified a small molecule named longdaysin, which is effective at lengthening the circadian period in a dose-dependent manner. Longdaysin targets several protein kinases that are crucial for the circadian clock, including CKIδ, CKIα, and ERK2. Longdaysin treatment inhibits PER1 degradation, which causes a longer period through the combined effects of these kinases [261]. N-(3-(9H-carbazol-9-yl)-2-hydroxypropyl)-N-(furan-2-ylmethyl)methanesulfonamide (KL001), a small CRY-interacting molecule, can prevent ubiquitin-dependent degradation of CRY1 and CRY2 and thus elicits a lengthened circadian period [262]. In a large-scale screening of small compounds, three compounds that inhibit CK1ε activity and lengthen the circadian period were identified [263]. Isojima et al. identified pharmacologically active compounds that can inhibit the phosphorylation of PER2 by CKIε or CKIδ, stabilize PER2, and lengthen the circadian period as a consequence [247].
Abnormalities in circadian rhythms correlate with bipolar disorder (BD), seasonal affective disorder (SAD), and major depressive disorder (MDD) [264,265,266,267]. As a mood stabilizer, lithium is an effective treatment for BD and major depressive disorder [267,268]. Lithium treatment improves sleep and circadian rhythms. Lithium is a GSK-3 inhibitor that can alter the stability and/or nuclear translocation of clock proteins, including CLOCK, CRY2, PER2, and REV-ERBα, and treatment with lithium causes phase delay and/or period lengthening of biochemical and physiological variables [268,269,270]. The effects of lithium on circadian rhythms are well conserved in a variety of species; for instance, in Neurospora, treatment with lithium led to increased FRQ protein stability, prolonged period lengths in a dose-dependent manner, and partial loss of temperature compensation [271].

7. Prospects

The circadian period and temperature compensation are controlled at multiple post-translational layers, similar to a nesting doll. It is necessary to carry out systemic profiling experiments to identify additional PTMs controlling the stability of circadian clock components, which would help elucidate their functions and mechanisms. The correlation between CK1 and the circadian negative component and period length will be important for validating the findings in Neurospora in other species.
To date, many kinases and kinase regulators have been revealed to be potential targets for adjusting circadian rhythms, and an increasing understanding of the detailed regulatory mechanisms of clock protein stability will highlight the implications for the development of new therapeutics. As many of the regulators of protein stability have pleiotropic effects in addition to affecting the circadian clock, in the future, to be effective and practicable, identifying and developing small molecules or drugs to specifically mitigate circadian disorders without affecting other important biological processes is crucial.

Author Contributions

Conceptualization, writing—draft and reviewing, J.G.; writing—original draft preparation, H.Z., Z.Z. and J.G.; revision, J.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (No. 32171162), the National Key R&D Program of China (2022YFA1604504), the Open Fund of the National Key Laboratory of Human Factors Engineering in the China Astronaut Research and Training Center (SYFD062008K), and the Joint Grant of Lingang Laboratory and National Key Laboratory of Human Factors Engineering of Chinese Astronaut Research and Training Center (LG-TKN-202203-01).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Bell-Pedersen, D.; Cassone, V.M.; Earnest, D.J.; Golden, S.S.; Hardin, P.E.; Thomas, T.L.; Zoran, M.J. Circadian Rhythms from Multiple Oscillators: Lessons from Diverse Organisms. Nat. Rev. Genet. 2005, 6, 544–556. [Google Scholar] [CrossRef] [PubMed]
  2. Vaze, K.M.; Sharma, V.K. On the Adaptive Significance of Circadian Clocks for Their Owners. Chronobiol. Int. 2013, 30, 413–433. [Google Scholar] [CrossRef] [PubMed]
  3. Luo, R.; Huang, Y.; Ma, H.; Guo, J. How to Live on Mars with a Proper Circadian Clock? Front. Astron. Space Sci. 2022, 8, 796943. [Google Scholar] [CrossRef]
  4. Harmer, S.L.; Hogenesch, J.B.; Straume, M.; Chang, H.-S.; Han, B.; Zhu, T.; Wang, X.; Kreps, J.A.; Kay, S.A. Orchestrated Transcription of Key Pathways in Arabidopsis by the Circadian Clock. Science 2000, 290, 2110–2113. [Google Scholar] [CrossRef] [PubMed]
  5. Vijayan, V.; Zuzow, R.; O’Shea, E.K. Oscillations in Supercoiling Drive Circadian Gene Expression in Cyanobacteria. Proc. Natl. Acad. Sci. USA 2009, 106, 22564–22568. [Google Scholar] [CrossRef]
  6. Díaz, R.D.; Larrondo, L.F. A Circadian Clock in Neurospora Crassa Functions during Plant Cell Wall Deconstruction. Fungal Biol. 2020, 124, 501–508. [Google Scholar] [CrossRef] [PubMed]
  7. Ito, H.; Mutsuda, M.; Murayama, Y.; Tomita, J.; Hosokawa, N.; Terauchi, K.; Sugita, C.; Sugita, M.; Kondo, T.; Iwasaki, H. Cyanobacterial Daily Life with Kai-Based Circadian and Diurnal Genome-Wide Transcriptional Control in Synechococcus Elongatus. Proc. Natl. Acad. Sci. USA 2009, 106, 14168–14173. [Google Scholar] [CrossRef] [PubMed]
  8. Zhang, R.; Lahens, N.F.; Ballance, H.I.; Hughes, M.E.; Hogenesch, J.B. A Circadian Gene Expression Atlas in Mammals: Implications for Biology and Medicine. Proc. Natl. Acad. Sci. USA 2014, 111, 16219–16224. [Google Scholar] [CrossRef]
  9. Mure, L.S.; Le, H.D.; Benegiamo, G.; Chang, M.W.; Rios, L.; Jillani, N.; Ngotho, M.; Kariuki, T.; Dkhissi-Benyahya, O.; Cooper, H.M.; et al. Diurnal Transcriptome Atlas of a Primate across Major Neural and Peripheral Tissues. Science 2018, 359, eaao0318. [Google Scholar] [CrossRef]
  10. Loros, J.J. Principles of the Animal Molecular Clock Learned from Neurospora. Eur. J. Neurosci. 2020, 51, 19–33. [Google Scholar] [CrossRef]
  11. Patke, A.; Young, M.W.; Axelrod, S. Molecular Mechanisms and Physiological Importance of Circadian Rhythms. Nat. Rev. Mol. Cell Biol. 2020, 21, 67–84. [Google Scholar] [CrossRef]
  12. Halberg, F. Physiologic 24-hour periodicity; general and procedural considerations with reference to the adrenal cycle. Int. Z. Vitaminforschung Beih. 1959, 10, 225–296. [Google Scholar]
  13. Edery, I. Circadian Rhythms in a Nutshell. Physiol. Genom. 2000, 3, 59–74. [Google Scholar] [CrossRef] [PubMed]
  14. Glossop, N.R.J.; Lyons, L.C.; Hardin, P.E. Interlocked Feedback Loops within the Drosophila Circadian Oscillator. Science 1999, 286, 766–768. [Google Scholar] [CrossRef] [PubMed]
  15. Shearman, L.P.; Sriram, S.; Weaver, D.R.; Maywood, E.S.; Chaves, I.; Zheng, B.; Kume, K.; Lee, C.C.; Horst, G.T.J.v.d.; Hastings, M.H.; et al. Interacting Molecular Loops in the Mammalian Circadian Clock. Science 2000, 288, 1013–1019. [Google Scholar] [CrossRef] [PubMed]
  16. Berman-Frank, I.; Lundgren, P.; Falkowski, P. Nitrogen Fixation and Photosynthetic Oxygen Evolution in Cyanobacteria. Res. Microbiol. 2003, 154, 157–164. [Google Scholar] [CrossRef] [PubMed]
  17. Johnson, C.H.; Stewart, P.L.; Egli, M. The Cyanobacterial Circadian System: From Biophysics to Bioevolution. Annu. Rev. Biophys. 2011, 40, 143–167. [Google Scholar] [CrossRef]
  18. Swan, J.A.; Golden, S.S.; LiWang, A.; Partch, C.L. Structure, Function, and Mechanism of the Core Circadian Clock in Cyanobacteria. J. Biol. Chem. 2018, 293, 5026–5034. [Google Scholar] [CrossRef]
  19. Zwicker, D.; Lubensky, D.K.; ten Wolde, P.R. Robust Circadian Clocks from Coupled Protein-Modification and Transcription-Translation Cycles. Proc. Natl. Acad. Sci. USA 2010, 107, 22540–22545. [Google Scholar] [CrossRef]
  20. Qin, X.; Byrne, M.; Mori, T.; Zou, P.; Williams, D.R.; Mchaourab, H.; Johnson, C.H. Intermolecular Associations Determine the Dynamics of the Circadian KaiABC Oscillator. Proc. Natl. Acad. Sci. USA 2010, 107, 14805–14810. [Google Scholar] [CrossRef]
  21. Tomita, J.; Nakajima, M.; Kondo, T.; Iwasaki, H. No Transcription-Translation Feedback in Circadian Rhythm of KaiC Phosphorylation. Science 2005, 307, 251–254. [Google Scholar] [CrossRef] [PubMed]
  22. Hosokawa, N.; Kushige, H.; Iwasaki, H. Attenuation of the Posttranslational Oscillator via Transcription–Translation Feedback Enhances Circadian-Phase Shifts in Synechococcus. Proc. Natl. Acad. Sci. USA 2013, 110, 14486–14491. [Google Scholar] [CrossRef] [PubMed]
  23. Nakajima, M.; Imai, K.; Ito, H.; Nishiwaki, T.; Murayama, Y.; Iwasaki, H.; Oyama, T.; Kondo, T. Reconstitution of Circadian Oscillation of Cyanobacterial KaiC Phosphorylation in Vitro. Science 2005, 308, 414–415. [Google Scholar] [CrossRef] [PubMed]
  24. Dunlap, J.C.; Feldman, J.F. On the Role of Protein Synthesis in the Circadian Clock of Neurospora Crassa. Proc. Natl. Acad. Sci. USA 1988, 85, 1096–1100. [Google Scholar] [CrossRef] [PubMed]
  25. Hurley, J.M.; Larrondo, L.F.; Loros, J.J.; Dunlap, J.C. Conserved RNA Helicase FRH Acts Nonenzymatically to Support the Intrinsically Disordered Neurospora Clock Protein FRQ. Mol. Cell 2013, 52, 832–843. [Google Scholar] [CrossRef] [PubMed]
  26. Cheng, P.; He, Q.; He, Q.; Wang, L.; Liu, Y. Regulation of the Neurospora Circadian Clock by an RNA Helicase. Genes Dev. 2005, 19, 234–241. [Google Scholar] [CrossRef]
  27. Guo, J.; Cheng, P.; Liu, Y. Functional Significance of FRH in Regulating the Phosphorylation and Stability of Neurospora Circadian Clock Protein FRQ. J. Biol. Chem. 2010, 285, 11508–11515. [Google Scholar] [CrossRef]
  28. Brunner, M.; Káldi, K. Interlocked Feedback Loops of the Circadian Clock of Neurospora Crassa. Mol. Microbiol. 2008, 68, 255–262. [Google Scholar] [CrossRef]
  29. Johnson, C.H.; Zhao, C.; Xu, Y.; Mori, T. Timing the Day: What Makes Bacterial Clocks Tick? Nat. Rev. Microbiol. 2017, 15, 232–242. [Google Scholar] [CrossRef]
  30. Dunlap, J.C.; Loros, J.J. Making Time: Conservation of Biological Clocks from Fungi to Animals. Microbiol. Spectr. 2017, 5, 10–1128. [Google Scholar] [CrossRef]
  31. Xu, X.; Yuan, L.; Yang, X.; Zhang, X.; Wang, L.; Xie, Q. Circadian Clock in Plants: Linking Timing to Fitness. J. Integr. Plant Biol. 2022, 64, 792–811. [Google Scholar] [CrossRef] [PubMed]
  32. Nohales, M.A.; Kay, S.A. Molecular Mechanisms at the Core of the Plant Circadian Oscillator. Nat. Struct. Mol. Biol. 2016, 23, 1061–1069. [Google Scholar] [CrossRef] [PubMed]
  33. McClung, C.R. The Plant Circadian Oscillator. Biology 2019, 8, 14. [Google Scholar] [CrossRef] [PubMed]
  34. Chen, S.; Yang, J.; Zhang, Y.; Duan, C.; Liu, Q.; Huang, Z.; Xu, Y.; Zhou, L.; Xu, G. Ubiquitin-Conjugating Enzyme UBE2O Regulates Cellular Clock Function by Promoting the Degradation of the Transcription Factor BMAL1. J. Biol. Chem. 2018, 293, 11296–11309. [Google Scholar] [CrossRef] [PubMed]
  35. Yu, W.; Hardin, P.E. Circadian Oscillators of Drosophila and Mammals. J. Cell Sci. 2006, 119, 4793–4795. [Google Scholar] [CrossRef] [PubMed]
  36. McDonald, M.J.; Rosbash, M.; Emery, P. Wild-Type Circadian Rhythmicity Is Dependent on Closely Spaced E Boxes in the Drosophila Timeless Promoter. Mol. Cell. Biol. 2001, 21, 1207–1217. [Google Scholar] [CrossRef] [PubMed]
  37. Brown, S.A.; Kowalska, E.; Dallmann, R. (Re)Inventing the Circadian Feedback Loop. Dev. Cell 2012, 22, 477–487. [Google Scholar] [CrossRef] [PubMed]
  38. Lim, C.; Allada, R. Emerging Roles for Post-Transcriptional Regulation in Circadian Clocks. Nat. Neurosci. 2013, 16, 1544–1550. [Google Scholar] [CrossRef]
  39. Hastings, M.H.; Maywood, E.S.; Brancaccio, M. Generation of Circadian Rhythms in the Suprachiasmatic Nucleus. Nat. Rev. Neurosci. 2018, 19, 453–469. [Google Scholar] [CrossRef]
  40. Park, J.; Lee, K.; Kim, H.; Shin, H.; Lee, C. Endogenous Circadian Reporters Reveal Functional Differences of PERIOD Paralogs and the Significance of PERIOD:CK1 Stable Interaction. Proc. Natl. Acad. Sci. USA 2023, 120, e2212255120. [Google Scholar] [CrossRef]
  41. Takahashi, J.S. Transcriptional Architecture of the Mammalian Circadian Clock. Nat. Rev. Genet. 2017, 18, 164–179. [Google Scholar] [CrossRef]
  42. Seo, P.J.; Mas, P. Multiple Layers of Posttranslational Regulation Refine Circadian Clock Activity in Arabidopsis. Plant Cell 2014, 26, 79–87. [Google Scholar] [CrossRef]
  43. Xu, X.; Yuan, L.; Xie, Q. The Circadian Clock Ticks in Plant Stress Responses. Stress Biol. 2022, 2, 15. [Google Scholar] [CrossRef] [PubMed]
  44. Hayama, R.; Sarid-Krebs, L.; Richter, R.; Fernández, V.; Jang, S.; Coupland, G. PSEUDO RESPONSE REGULATORs Stabilize CONSTANS Protein to Promote Flowering in Response to Day Length. EMBO J. 2017, 36, 904–918. [Google Scholar] [CrossRef]
  45. Jabbur, M.L.; Johnson, C.H. Spectres of Clock Evolution: Past, Present, and Yet to Come. Front. Physiol. 2021, 12, 815847. [Google Scholar] [CrossRef] [PubMed]
  46. Lowrey, P.L.; Shimomura, K.; Antoch, M.P.; Yamazaki, S.; Zemenides, P.D.; Ralph, M.R.; Menaker, M.; Takahashi, J.S. Positional Syntenic Cloning and Functional Characterization of the Mammalian Circadian Mutation Tau. Science 2000, 288, 483–491. [Google Scholar] [CrossRef] [PubMed]
  47. Cha, J.; Chang, S.-S.; Huang, G.; Cheng, P.; Liu, Y. Control of WHITE COLLAR Localization by Phosphorylation Is a Critical Step in the Circadian Negative Feedback Process. EMBO J. 2008, 27, 3246–3255. [Google Scholar] [CrossRef]
  48. Marzoll, D.; Serrano, F.E.; Diernfellner, A.C.R.; Brunner, M. Neurospora Casein Kinase 1a Recruits the Circadian Clock Protein FRQ via the C-terminal Lobe of Its Kinase Domain. FEBS Lett. 2022, 596, 1881–1891. [Google Scholar] [CrossRef] [PubMed]
  49. Wang, B.; Stevenson, E.-L.; Dunlap, J.C. Functional Analysis of 110 Phosphorylation Sites on the Circadian Clock Protein FRQ Identifies Clusters Determining Period Length and Temperature Compensation. G3 2023, 13, jkac334. [Google Scholar] [CrossRef]
  50. Baker, C.L.; Kettenbach, A.N.; Loros, J.J.; Gerber, S.A.; Dunlap, J.C. Quantitative Proteomics Reveals a Dynamic Interactome and Phase-Specific Phosphorylation in the Neurospora Circadian Clock. Mol. Cell 2009, 34, 354–363. [Google Scholar] [CrossRef]
  51. Tang, C.-T.; Li, S.; Long, C.; Cha, J.; Huang, G.; Li, L.; Chen, S.; Liu, Y. Setting the Pace of the Neurospora Circadian Clock by Multiple Independent FRQ Phosphorylation Events. Proc. Natl. Acad. Sci. USA 2009, 106, 10722–10727. [Google Scholar] [CrossRef]
  52. Chen, X.; Liu, X.; Gan, X.; Li, S.; Ma, H.; Zhang, L.; Wang, P.; Li, Y.; Huang, T.; Yang, X.; et al. Differential Regulation of Phosphorylation, Structure, and Stability of Circadian Clock Protein FRQ Isoforms. J. Biol. Chem. 2023, 299, 104597. [Google Scholar] [CrossRef] [PubMed]
  53. Larrondo, L.F.; Olivares-Yañez, C.; Baker, C.L.; Loros, J.J.; Dunlap, J.C. Circadian Rhythms. Decoupling Circadian Clock Protein Turnover from Circadian Period Determination. Science 2015, 347, 1257277. [Google Scholar] [CrossRef] [PubMed]
  54. Querfurth, C.; Diernfellner, A.; Heise, F.; Lauinger, L.; Neiss, A.; Tataroglu, Ö.; Brunner, M.; Schafmeier, T. Posttranslational Regulation of Neurospora Circadian Clock by CK1a-Dependent Phosphorylation. Cold Spring Harb. Symp. Quant. Biol. 2007, 72, 177–183. [Google Scholar] [CrossRef]
  55. Diernfellner, A.C.R.; Brunner, M. Phosphorylation Timers in the Neurospora Crassa Circadian Clock. J. Mol. Biol. 2020, 432, 3449–3465. [Google Scholar] [CrossRef] [PubMed]
  56. Aronson, B.D.; Johnson, K.A.; Dunlap, J.C. Circadian Clock Locus Frequency: Protein Encoded by a Single Open Reading Frame Defines Period Length and Temperature Compensation. Proc. Natl. Acad. Sci. USA 1994, 91, 7683–7687. [Google Scholar] [CrossRef] [PubMed]
  57. Ruoff, P.; Loros, J.J.; Dunlap, J.C. The Relationship between FRQ-Protein Stability and Temperature Compensation in the Neurospora Circadian Clock. Proc. Natl. Acad. Sci. USA 2005, 102, 17681–17686. [Google Scholar] [CrossRef] [PubMed]
  58. Liu, Y.; Garceau, N.Y.; Loros, J.J.; Dunlap, J.C. Thermally Regulated Translational Control of FRQ Mediates Aspects of Temperature Responses in the Neurospora Circadian Clock. Cell 1997, 89, 477–486. [Google Scholar] [CrossRef]
  59. Anand, S.N.; Maywood, E.S.; Chesham, J.E.; Joynson, G.; Banks, G.T.; Hastings, M.H.; Nolan, P.M. Distinct and Separable Roles for Endogenous CRY1 and CRY2 within the Circadian Molecular Clockwork of the Suprachiasmatic Nucleus, as Revealed by the Fbxl3Afh Mutation. J. Neurosci. 2013, 33, 7145–7153. [Google Scholar] [CrossRef]
  60. He, Q.; Cha, J.; He, Q.; Lee, H.-C.; Yang, Y.; Liu, Y. CKI and CKII Mediate the FREQUENCY-Dependent Phosphorylation of the WHITE COLLAR Complex to Close the Neurospora Circadian Negative Feedback Loop. Genes Dev. 2006, 20, 2552–2565. [Google Scholar] [CrossRef]
  61. Querfurth, C.; Diernfellner, A.C.R.; Gin, E.; Malzahn, E.; Höfer, T.; Brunner, M. Circadian Conformational Change of the Neurospora Clock Protein FREQUENCY Triggered by Clustered Hyperphosphorylation of a Basic Domain. Mol. Cell 2011, 43, 713–722. [Google Scholar] [CrossRef]
  62. Hu, Y.; Liu, X.; Lu, Q.; Yang, Y.; He, Q.; Liu, Y.; Liu, X. FRQ-CK1 Interaction Underlies Temperature Compensation of the Neurospora Circadian Clock. mBio 2021, 12, e0142521. [Google Scholar] [CrossRef] [PubMed]
  63. Huang, G.; Chen, S.; Li, S.; Cha, J.; Long, C.; Li, L.; He, Q.; Liu, Y. Protein Kinase A and Casein Kinases Mediate Sequential Phosphorylation Events in the Circadian Negative Feedback Loop. Genes Dev. 2007, 21, 3283–3295. [Google Scholar] [CrossRef] [PubMed]
  64. Morales, Y.; Olsen, K.J.; Bulcher, J.M.; Johnson, S.J. Structure of Frequency-Interacting RNA Helicase from Neurospora Crassa Reveals High Flexibility in a Domain Critical for Circadian Rhythm and RNA Surveillance. PLoS ONE 2018, 13, e0196642. [Google Scholar] [CrossRef]
  65. Conrad, K.S.; Hurley, J.M.; Widom, J.; Ringelberg, C.S.; Loros, J.J.; Dunlap, J.C.; Crane, B.R. Structure of the Frequency-interacting RNA Helicase: A Protein Interaction Hub for the Circadian Clock. EMBO J. 2016, 35, 1707–1719. [Google Scholar] [CrossRef] [PubMed]
  66. He, Q. FWD1-Mediated Degradation of FREQUENCY in Neurospora Establishes a Conserved Mechanism for Circadian Clock Regulation. EMBO J. 2003, 22, 4421–4430. [Google Scholar] [CrossRef]
  67. Cheng, P.; Yang, Y.; Liu, Y. Interlocked Feedback Loops Contribute to the Robustness of the Neurospora Circadian Clock. Proc. Natl. Acad. Sci. USA 2001, 98, 7408–7413. [Google Scholar] [CrossRef]
  68. Lee, K.; Loros, J.J.; Dunlap, J.C. Interconnected Feedback Loops in the Neurospora Circadian System. Science 2000, 289, 107–110. [Google Scholar] [CrossRef]
  69. Cheng, P.; Yang, Y.; Gardner, K.H.; Liu, Y. PAS Domain-Mediated WC-1/WC-2 Interaction Is Essential for Maintaining the Steady-State Level of WC-1 and the Function of Both Proteins in Circadian Clock and Light Responses of Neurospora. Mol. Cell. Biol. 2002, 22, 517–524. [Google Scholar] [CrossRef]
  70. Schafmeier, T.; Diernfellner, A.; Schäfer, A.; Dintsis, O.; Neiss, A.; Brunner, M. Circadian Activity and Abundance Rhythms of the Neurospora Clock Transcription Factor WCC Associated with Rapid Nucleo–Cytoplasmic Shuttling. Genes Dev. 2008, 22, 3397–3402. [Google Scholar] [CrossRef]
  71. Denault, D.L.; Loros, J.J.; Dunlap, J.C. WC-2 Mediates WC-1-FRQ Interaction within the PAS Protein-Linked Circadian Feedback Loop of Neurospora. EMBO J. 2001, 20, 109–117. [Google Scholar] [CrossRef] [PubMed]
  72. Schafmeier, T.; Haase, A.; Káldi, K.; Scholz, J.; Fuchs, M.; Brunner, M. Transcriptional Feedback of Neurospora Circadian Clock Gene by Phosphorylation-Dependent Inactivation of Its Transcription Factor. Cell 2005, 122, 235–246. [Google Scholar] [CrossRef] [PubMed]
  73. Schafmeier, T.; Káldi, K.; Diernfellner, A.; Mohr, C.; Brunner, M. Phosphorylation-Dependent Maturation of Neurospora Circadian Clock Protein from a Nuclear Repressor toward a Cytoplasmic Activator. Genes Dev. 2006, 20, 297–306. [Google Scholar] [CrossRef] [PubMed]
  74. Huang, G.; He, Q.; Guo, J.; Cha, J.; Liu, Y. The Ccr4-Not Protein Complex Regulates the Phase of the Neurospora Circadian Clock by Controlling WHITE COLLAR Protein Stability and Activity. J. Biol. Chem. 2013, 288, 31002–31009. [Google Scholar] [CrossRef] [PubMed]
  75. Shi, M.; Collett, M.; Loros, J.J.; Dunlap, J.C. FRQ-Interacting RNA Helicase Mediates Negative and Positive Feedback in the Neurospora Circadian Clock. Genetics 2010, 184, 351–361. [Google Scholar] [CrossRef] [PubMed]
  76. Hong, C.I.; Ruoff, P.; Loros, J.J.; Dunlap, J.C. Closing the Circadian Negative Feedback Loop: FRQ-Dependent Clearance of WC-1 from the Nucleus. Genes Dev. 2008, 22, 3196–3204. [Google Scholar] [CrossRef] [PubMed]
  77. He, Q.; Shu, H.; Cheng, P.; Chen, S.; Wang, L.; Liu, Y. Light-Independent Phosphorylation of WHITE COLLAR-1 Regulates Its Function in the Neurospora Circadian Negative Feedback Loop. J. Biol. Chem. 2005, 280, 17526–17532. [Google Scholar] [CrossRef]
  78. Wang, B.; Kettenbach, A.N.; Zhou, X.; Loros, J.J.; Dunlap, J.C. The Phospho-Code Determining Circadian Feedback Loop Closure and Output in Neurospora. Mol. Cell 2019, 74, 771–784.e3. [Google Scholar] [CrossRef]
  79. Sancar, G.; Sancar, C.; Brunner, M.; Schafmeier, T. Activity of the Circadian Transcription Factor White Collar Complex Is Modulated by Phosphorylation of SP-motifs. FEBS Lett. 2009, 583, 1833–1840. [Google Scholar] [CrossRef]
  80. Harms, E.; Kivimäe, S.; Young, M.W.; Saez, L. Posttranscriptional and Posttranslational Regulation of Clock Genes. J. Biol. Rhythm. 2004, 19, 361–373. [Google Scholar] [CrossRef]
  81. Lauinger, L.; Diernfellner, A.; Falk, S.; Brunner, M. The RNA Helicase FRH Is an ATP-Dependent Regulator of CK1a in the Circadian Clock of Neurospora Crassa. Nat. Commun. 2014, 5, 3598. [Google Scholar] [CrossRef] [PubMed]
  82. Collart, M.A.; Panasenko, O.O. The Ccr4–Not Complex. Gene 2012, 492, 42–53. [Google Scholar] [CrossRef] [PubMed]
  83. Lee, J.M.; Hammarén, H.M.; Savitski, M.M.; Baek, S.H. Control of Protein Stability by Post-Translational Modifications. Nat. Commun. 2023, 14, 201. [Google Scholar] [CrossRef] [PubMed]
  84. Miller, S.; Hirota, T. Pharmacological Interventions to Circadian Clocks and Their Molecular Bases. J. Mol. Biol. 2020, 432, 3498–3514. [Google Scholar] [CrossRef] [PubMed]
  85. Dikic, I. Proteasomal and Autophagic Degradation Systems. Annu. Rev. Biochem. 2017, 86, 193–224. [Google Scholar] [CrossRef] [PubMed]
  86. Deshaies, R.J. SCF and Cullin/Ring H2-Based Ubiquitin Ligases. Annu. Rev. Cell Dev. Biol. 1999, 15, 435–467. [Google Scholar] [CrossRef] [PubMed]
  87. Ko, H.W.; Jiang, J.; Edery, I. Role for Slimb in the Degradation of Drosophila Period Protein Phosphorylated by Doubletime. Nature 2002, 420, 673–678. [Google Scholar] [CrossRef]
  88. Grima, B.; Lamouroux, A.; Chélot, E.; Papin, C.; Limbourg-Bouchon, B.; Rouyer, F. The F-Box Protein Slimb Controls the Levels of Clock Proteins Period and Timeless. Nature 2002, 420, 178–182. [Google Scholar] [CrossRef]
  89. Chiu, J.C.; Vanselow, J.T.; Kramer, A.; Edery, I. The Phospho-Occupancy of an Atypical SLIMB-Binding Site on PERIOD That Is Phosphorylated by DOUBLETIME Controls the Pace of the Clock. Genes Dev. 2008, 22, 1758–1772. [Google Scholar] [CrossRef]
  90. Wei, N.; Serino, G.; Deng, X. The COP9 Signalosome: More than a Protease. Trends Biochem. Sci. 2008, 33, 592–600. [Google Scholar] [CrossRef]
  91. Wang, J.; Hu, Q.; Chen, H.; Zhou, Z.; Li, W.; Wang, Y.; Li, S.; He, Q. Role of Individual Subunits of the Neurospora Crassa CSN Complex in Regulation of Deneddylation and Stability of Cullin Proteins. PLoS Genet. 2010, 6, e1001232. [Google Scholar] [CrossRef]
  92. Zhou, Z.; Wang, Y.; Cai, G.; He, Q. Neurospora COP9 Signalosome Integrity Plays Major Roles for Hyphal Growth, Conidial Development, and Circadian Function. PLoS Genet. 2012, 8, e1002712. [Google Scholar] [CrossRef]
  93. Li, L.; Nelson, C.J.; Trösch, J.; Castleden, I.; Huang, S.; Millar, A.H. Protein Degradation Rate in Arabidopsis thaliana Leaf Growth and Development. Plant Cell 2017, 29, 207–228. [Google Scholar] [CrossRef] [PubMed]
  94. Kim, J.; Geng, R.; Gallenstein, R.A.; Somers, D.E. The F-Box Protein ZEITLUPE Controls Stability and Nucleocytoplasmic Partitioning of GIGANTEA. Dev. Camb. Engl. 2013, 140, 4060–4069. [Google Scholar] [CrossRef] [PubMed]
  95. Lee, C.-M.; Feke, A.; Li, M.-W.; Adamchek, C.; Webb, K.; Pruneda-Paz, J.; Bennett, E.J.; Kay, S.A.; Gendron, J.M. Decoys Untangle Complicated Redundancy and Reveal Targets of Circadian Clock F-Box Proteins. Plant Physiol. 2018, 177, 1170–1186. [Google Scholar] [CrossRef] [PubMed]
  96. Más, P.; Kim, W.-Y.; Somers, D.E.; Kay, S.A. Targeted Degradation of TOC1 by ZTL Modulates Circadian Function in Arabidopsis thaliana. Nature 2003, 426, 567–570. [Google Scholar] [CrossRef] [PubMed]
  97. Kiba, T.; Henriques, R.; Sakakibara, H.; Chua, N.-H. Targeted Degradation of PSEUDO-RESPONSE REGULATOR5 by an SCFZTL Complex Regulates Clock Function and Photomorphogenesis in Arabidopsis thaliana. Plant Cell 2007, 19, 2516–2530. [Google Scholar] [CrossRef] [PubMed]
  98. Kim, W.-Y.; Fujiwara, S.; Suh, S.-S.; Kim, J.; Kim, Y.; Han, L.; David, K.; Putterill, J.; Nam, H.G.; Somers, D.E. ZEITLUPE Is a Circadian Photoreceptor Stabilized by GIGANTEA in Blue Light. Nature 2007, 449, 356–360. [Google Scholar] [CrossRef]
  99. Lee, C.-M.; Li, M.-W.; Feke, A.; Liu, W.; Saffer, A.M.; Gendron, J.M. GIGANTEA Recruits the UBP12 and UBP13 Deubiquitylases to Regulate Accumulation of the ZTL Photoreceptor Complex. Nat. Commun. 2019, 10, 3750. [Google Scholar] [CrossRef]
  100. Huang, H.; Alvarez, S.; Bindbeutel, R.; Shen, Z.; Naldrett, M.J.; Evans, B.S.; Briggs, S.P.; Hicks, L.M.; Kay, S.A.; Nusinow, D.A. Identification of Evening Complex Associated Proteins in Arabidopsis by Affinity Purification and Mass Spectrometry. Mol. Cell. Proteom. 2016, 15, 201–217. [Google Scholar] [CrossRef]
  101. Yu, J.-W.; Rubio, V.; Lee, N.-Y.; Bai, S.; Lee, S.-Y.; Kim, S.-S.; Liu, L.; Zhang, Y.; Irigoyen, M.L.; Sullivan, J.A.; et al. COP1 and ELF3 Control Circadian Function and Photoperiodic Flowering by Regulating GI Stability. Mol. Cell 2008, 32, 617–630. [Google Scholar] [CrossRef] [PubMed]
  102. Seo, H.S.; Watanabe, E.; Tokutomi, S.; Nagatani, A.; Chua, N.-H. Photoreceptor Ubiquitination by COP1 E3 Ligase Desensitizes Phytochrome A Signaling. Genes Dev. 2004, 18, 617–622. [Google Scholar] [CrossRef] [PubMed]
  103. Hayama, R.; Yang, P.; Valverde, F.; Mizoguchi, T.; Furutani-Hayama, I.; Vierstra, R.D.; Coupland, G. Ubiquitin Carboxyl-Terminal Hydrolases Are Required for Period Maintenance of the Circadian Clock at High Temperature in Arabidopsis. Sci. Rep. 2019, 9, 17030. [Google Scholar] [CrossRef] [PubMed]
  104. Siepka, S.M.; Yoo, S.-H.; Park, J.; Song, W.; Kumar, V.; Hu, Y.; Lee, C.; Takahashi, J.S. Circadian Mutant Overtime Reveals F-Box Protein FBXL3 Regulation of Cryptochrome and Period Gene Expression. Cell 2007, 129, 1011–1023. [Google Scholar] [CrossRef]
  105. Busino, L.; Bassermann, F.; Maiolica, A.; Lee, C.; Nolan, P.M.; Godinho, S.I.H.; Draetta, G.F.; Pagano, M. SCFFbxl3 Controls the Oscillation of the Circadian Clock by Directing the Degradation of Cryptochrome Proteins. Science 2007, 316, 900–904. [Google Scholar] [CrossRef]
  106. Godinho, S.I.H.; Maywood, E.S.; Shaw, L.; Tucci, V.; Barnard, A.R.; Busino, L.; Pagano, M.; Kendall, R.; Quwailid, M.M.; Romero, M.R.; et al. The After-Hours Mutant Reveals a Role for Fbxl3 in Determining Mammalian Circadian Period. Science 2007, 316, 897–900. [Google Scholar] [CrossRef] [PubMed]
  107. Masuda, S.; Narasimamurthy, R.; Yoshitane, H.; Kim, J.K.; Fukada, Y.; Virshup, D.M. Mutation of a PER2 Phosphodegron Perturbs the Circadian Phosphoswitch. Proc. Natl. Acad. Sci. USA 2020, 117, 10888–10896. [Google Scholar] [CrossRef] [PubMed]
  108. Gallego, M.; Virshup, D.M. Post-Translational Modifications Regulate the Ticking of the Circadian Clock. Nat. Rev. Mol. Cell Biol. 2007, 8, 139–148. [Google Scholar] [CrossRef]
  109. Liu, J.; Zou, X.; Gotoh, T.; Brown, A.M.; Jiang, L.; Wisdom, E.L.; Kim, J.K.; Finkielstein, C.V. Distinct Control of PERIOD2 Degradation and Circadian Rhythms by the Oncoprotein and Ubiquitin Ligase MDM2. Sci. Signal. 2018, 11, eaau0715. [Google Scholar] [CrossRef] [PubMed]
  110. Jiang, H.; Zheng, S.; Qian, Y.; Zhou, Y.; Dai, H.; Liang, Y.; He, Y.; Gao, R.; Lv, H.; Zhang, J.; et al. Restored UBE2C Expression in Islets Promotes β-Cell Regeneration in Mice by Ubiquitinating PER1. Cell. Mol. Life Sci. 2023, 80, 226. [Google Scholar] [CrossRef]
  111. Yoo, S.-H.; Mohawk, J.A.; Siepka, S.M.; Shan, Y.; Huh, S.K.; Hong, H.-K.; Kornblum, I.; Kumar, V.; Koike, N.; Xu, M.; et al. Competing E3 Ubiquitin Ligases Govern Circadian Periodicity by Degradation of CRY in Nucleus and Cytoplasm. Cell 2013, 152, 1091–1105. [Google Scholar] [CrossRef] [PubMed]
  112. Hirano, A.; Yumimoto, K.; Tsunematsu, R.; Matsumoto, M.; Oyama, M.; Kozuka-Hata, H.; Nakagawa, T.; Lanjakornsiripan, D.; Nakayama, K.I.; Fukada, Y. FBXL21 Regulates Oscillation of the Circadian Clock through Ubiquitination and Stabilization of Cryptochromes. Cell 2013, 152, 1106–1118. [Google Scholar] [CrossRef] [PubMed]
  113. Zhao, X.; Hirota, T.; Han, X.; Cho, H.; Chong, L.-W.; Lamia, K.; Liu, S.; Atkins, A.R.; Banayo, E.; Liddle, C.; et al. Circadian Amplitude Regulation via FBXW7-Targeted REV-ERBα Degradation. Cell 2016, 165, 1644–1657. [Google Scholar] [CrossRef] [PubMed]
  114. Lecker, S.H.; Goldberg, A.L.; Mitch, W.E. Protein Degradation by the Ubiquitin-Proteasome Pathway in Normal and Disease States. J. Am. Soc. Nephrol. 2006, 17, 1807–1819. [Google Scholar] [CrossRef] [PubMed]
  115. Kubra, S.; Zhang, H.; Si, Y.; Gao, X.; Wang, T.; Pan, L.; Li, L.; Zhong, N.; Fu, J.; Zhang, B.; et al. REGγ Regulates Circadian Clock by Modulating BMAL1 Protein Stability. Cell Death Discov. 2021, 7, 335. [Google Scholar] [CrossRef] [PubMed]
  116. He, C. Chaperone-Mediated Autophagy on the Clock. Nat. Cell Biol. 2021, 23, 1220–1221. [Google Scholar] [CrossRef] [PubMed]
  117. Juste, Y.R.; Kaushik, S.; Bourdenx, M.; Aflakpui, R.; Bandyopadhyay, S.; Garcia, F.; Diaz, A.; Lindenau, K.; Tu, V.; Krause, G.J.; et al. Reciprocal Regulation of Chaperone-Mediated Autophagy and the Circadian Clock. Nat. Cell Biol. 2021, 23, 1255–1270. [Google Scholar] [CrossRef]
  118. Toledo, M.; Batista-Gonzalez, A.; Merheb, E.; Aoun, M.L.; Tarabra, E.; Feng, D.; Sarparanta, J.; Merlo, P.; Botrè, F.; Schwartz, G.J.; et al. Autophagy Regulates the Liver Clock and Glucose Metabolism by Degrading CRY1. Cell Metab. 2018, 28, 268–281.e4. [Google Scholar] [CrossRef]
  119. Yang, M.; Chen, P.; Liu, J.; Zhu, S.; Kroemer, G.; Klionsky, D.J.; Lotze, M.T.; Zeh, H.J.; Kang, R.; Tang, D. Clockophagy Is a Novel Selective Autophagy Process Favoring Ferroptosis. Sci. Adv. 2019, 5, eaaw2238. [Google Scholar] [CrossRef]
  120. He, C.; Klionsky, D.J. Regulation Mechanisms and Signaling Pathways of Autophagy. Annu. Rev. Genet. 2009, 43, 67–93. [Google Scholar] [CrossRef]
  121. Vitaterna, M.H.; King, D.P.; Chang, A.-M.; Kornhauser, J.M.; Lowrey, P.L.; McDonald, J.D.; Dove, W.F.; Pinto, L.H.; Turek, F.W.; Takahashi, J.S. Mutagenesis and Mapping of a Mouse Gene, Clock, Essential for Circadian Behavior. Science 1994, 264, 719–725. [Google Scholar] [CrossRef] [PubMed]
  122. Jeong, K.; He, B.; Nohara, K.; Park, N.; Shin, Y.; Kim, S.; Shimomura, K.; Koike, N.; Yoo, S.-H.; Chen, Z. Dual Attenuation of Proteasomal and Autophagic BMAL1 Degradation in ClockΔ19/+ Mice Contributes to Improved Glucose Homeostasis. Sci. Rep. 2015, 5, 12801. [Google Scholar] [CrossRef] [PubMed]
  123. Höltje, J.V. Bacterial Lysozymes. EXS 1996, 75, 65–74. [Google Scholar] [CrossRef] [PubMed]
  124. Becker, S.H.; Darwin, K.H. Bacterial Proteasomes: Mechanistic and Functional Insights. Microbiol. Mol. Biol. Rev. 2017, 81, e00036-16. [Google Scholar] [CrossRef] [PubMed]
  125. Ashida, H.; Sasakawa, C. Bacterial E3 Ligase Effectors Exploit Host Ubiquitin Systems. Curr. Opin. Microbiol. 2017, 35, 16–22. [Google Scholar] [CrossRef] [PubMed]
  126. Reischl, S.; Kramer, A. Kinases and Phosphatases in the Mammalian Circadian Clock. FEBS Lett. 2011, 585, 1393–1399. [Google Scholar] [CrossRef]
  127. Tataroğlu, Ö.; Schafmeier, T. Of Switches and Hourglasses: Regulation of Subcellular Traffic in Circadian Clocks by Phosphorylation. EMBO Rep. 2010, 11, 927–935. [Google Scholar] [CrossRef]
  128. Diernfellner, A.C.R.; Querfurth, C.; Salazar, C.; Höfer, T.; Brunner, M. Phosphorylation Modulates Rapid Nucleocytoplasmic Shuttling and Cytoplasmic Accumulation of Neurospora Clock Protein FRQ on a Circadian Time Scale. Genes Dev. 2009, 23, 2192–2200. [Google Scholar] [CrossRef]
  129. Iitaka, C.; Miyazaki, K.; Akaike, T.; Ishida, N. A Role for Glycogen Synthase Kinase-3β in the Mammalian Circadian Clock. J. Biol. Chem. 2005, 280, 29397–29402. [Google Scholar] [CrossRef]
  130. Kinoshita, C.; Miyazaki, K.; Ishida, N. Chronic Stress Affects PERIOD2 Expression through Glycogen Synthase Kinase-3β Phosphorylation in the Central Clock. NeuroReport 2012, 23, 98–102. [Google Scholar] [CrossRef]
  131. Diernfellner, A.C.R.; Schafmeier, T. Phosphorylations: Making the Neurospora crassa Circadian Clock Tick. FEBS Lett. 2011, 585, 1461–1466. [Google Scholar] [CrossRef] [PubMed]
  132. Kweon, S.H.; Lee, J.; Lim, C.; Choe, J. High-Amplitude Circadian Rhythms in Drosophila Driven by Calcineurin-Mediated Post-Translational Control of Sarah. Genet. 2018, 209, 815–828. [Google Scholar] [CrossRef] [PubMed]
  133. Agrawal, P.; Hardin, P.E. An RNAi Screen To Identify Protein Phosphatases That Function Within the Drosophila Circadian Clock. G3 Genes Genomes Genet. 2016, 6, 4227–4238. [Google Scholar] [CrossRef] [PubMed]
  134. Tataroglu, O.; Emery, P. The Molecular Ticks of the Drosophila Circadian Clock. Curr. Opin. Insect Sci. 2015, 7, 51–57. [Google Scholar] [CrossRef] [PubMed]
  135. Baruah, D.; Marak, C.N.K.; Roy, A.; Gohain, D.; Kumar, A.; Das, P.; Borkovich, K.A.; Tamuli, R. Multiple Calcium Signaling Genes Play a Role in the Circadian Period of Neurospora Crassa. FEMS Microbiol. Lett. 2023, 370, fnad044. [Google Scholar] [CrossRef]
  136. Marzoll, D.; Serrano, F.E.; Shostak, A.; Schunke, C.; Diernfellner, A.C.R.; Brunner, M. Casein Kinase 1 and Disordered Clock Proteins Form Functionally Equivalent, Phospho-Based Circadian Modules in Fungi and Mammals. Proc. Natl. Acad. Sci. USA 2022, 119, e2118286119. [Google Scholar] [CrossRef] [PubMed]
  137. Yang, Y.; Cheng, P.; Liu, Y. Regulation of the Neurospora Circadian Clock by Casein Kinase II. Genes Dev. 2002, 16, 994–1006. [Google Scholar] [CrossRef] [PubMed]
  138. Uehara, T.N.; Mizutani, Y.; Kuwata, K.; Hirota, T.; Sato, A.; Mizoi, J.; Takao, S.; Matsuo, H.; Suzuki, T.; Ito, S.; et al. Casein Kinase 1 Family Regulates PRR5 and TOC1 in the Arabidopsis Circadian Clock. Proc. Natl. Acad. Sci. USA 2019, 116, 11528–11536. [Google Scholar] [CrossRef]
  139. Knippschild, U.; Gocht, A.; Wolff, S.; Huber, N.; Löhler, J.; Stöter, M. The Casein Kinase 1 Family: Participation in Multiple Cellular Processes in Eukaryotes. Cell. Signal. 2005, 17, 675–689. [Google Scholar] [CrossRef]
  140. Li, S.-S.; Dong, Y.-H.; Liu, Z.-P. Recent Advances in the Development of Casein Kinase 1 Inhibitors. Curr. Med. Chem. 2021, 28, 1585–1604. [Google Scholar] [CrossRef]
  141. Price, J.L.; Blau, J.; Rothenfluh, A.; Abodeely, M.; Kloss, B.; Young, M.W. Double-Time Is a Novel Drosophila Clock Gene That Regulates PERIOD Protein Accumulation. Cell 1998, 94, 83–95. [Google Scholar] [CrossRef] [PubMed]
  142. Edery, I.; Zwiebel, L.J.; Dembinska, M.E.; Rosbash, M. Temporal Phosphorylation of the Drosophila Period Protein. Proc. Natl. Acad. Sci. USA 1994, 91, 2260–2264. [Google Scholar] [CrossRef] [PubMed]
  143. Duvall, L.B.; Taghert, P.H. Circadian Rhythms: Biological Clocks Work in Phospho-Time. Curr. Biol. 2011, 21, R305–R307. [Google Scholar] [CrossRef] [PubMed]
  144. Yu, Q.; Jacquier, A.C.; Citri, Y.; Hamblen, M.; Hall, J.C.; Rosbash, M. Molecular Mapping of Point Mutations in the Period Gene That Stop or Speed up Biological Clocks in Drosophila melanogaster. Proc. Natl. Acad. Sci. USA 1987, 84, 784–788. [Google Scholar] [CrossRef] [PubMed]
  145. Konopka, R.J.; Benzer, S. Clock Mutants of Drosophila melanogaster. Proc. Natl. Acad. Sci. USA 1971, 68, 2112–2116. [Google Scholar] [CrossRef] [PubMed]
  146. Baylies, M.K.; Bargiello, T.A.; Jackson, F.R.; Young, M.W. Changes in Abundance or Structure of the per Gene Product Can Alter Periodicity of the Drosophila Clock. Nature 1987, 326, 390–392. [Google Scholar] [CrossRef] [PubMed]
  147. Kloss, B.; Price, J.L.; Saez, L.; Blau, J.; Rothenfluh, A.; Wesley, C.S.; Young, M.W. The Drosophila Clock Gene Double-Time Encodes a Protein Closely Related to Human Casein Kinase Iε. Cell 1998, 94, 97–107. [Google Scholar] [CrossRef]
  148. Kloss, B.; Rothenfluh, A.; Young, M.W.; Saez, L. Phosphorylation of PERIOD Is Influenced by Cycling Physical Associations of DOUBLE-TIME, PERIOD, and TIMELESS in the Drosophila Clock. Neuron 2001, 30, 699–706. [Google Scholar] [CrossRef]
  149. Top, D.; Harms, E.; Syed, S.; Adams, E.L.; Saez, L. GSK-3 and CK2 Kinases Converge on Timeless to Regulate the Master Clock. Cell Rep. 2016, 16, 357–367. [Google Scholar] [CrossRef]
  150. Szabó, Á.; Papin, C.; Zorn, D.; Ponien, P.; Weber, F.; Raabe, T.; Rouyer, F. The CK2 Kinase Stabilizes CLOCK and Represses Its Activity in the Drosophila Circadian Oscillator. PLoS Biol. 2013, 11, e1001645. [Google Scholar] [CrossRef]
  151. Smith, E.M.; Lin, J.-M.; Meissner, R.-A.; Allada, R. Dominant-Negative CK2alpha Induces Potent Effects on Circadian Rhythmicity. PLoS Genet. 2008, 4, e12. [Google Scholar] [CrossRef]
  152. Martinek, S.; Inonog, S.; Manoukian, A.S.; Young, M.W. A Role for the Segment Polarity Gene Shaggy/GSK-3 in the Drosophila Circadian Clock. Cell 2001, 105, 769–779. [Google Scholar] [CrossRef] [PubMed]
  153. Ko, H.W.; Kim, E.Y.; Chiu, J.; Vanselow, J.T.; Kramer, A.; Edery, I. A Hierarchical Phosphorylation Cascade That Regulates the Timing of PERIOD Nuclear Entry Reveals Novel Roles for Proline-Directed Kinases and GSK-3β/SGG in Circadian Clocks. J. Neurosci. 2010, 30, 12664–12675. [Google Scholar] [CrossRef]
  154. Yu, W.; Houl, J.H.; Hardin, P.E. NEMO Kinase Contributes to Core Period Determination by Slowing the Pace of the Drosophila Circadian Oscillator. Curr. Biol. 2011, 21, 756–761. [Google Scholar] [CrossRef]
  155. Chiu, J.C.; Ko, H.W.; Edery, I. NEMO/NLK Phosphorylates PERIOD to Initiate a Time-Delay Phosphorylation Circuit That Sets Circadian Clock Speed. Cell 2011, 145, 357–370. [Google Scholar] [CrossRef] [PubMed]
  156. Kivimäe, S.; Saez, L.; Young, M.W. Activating PER Repressor through a DBT-Directed Phosphorylation Switch. PLoS Biol. 2008, 6, e183. [Google Scholar] [CrossRef] [PubMed]
  157. Cyran, S.A.; Yiannoulos, G.; Buchsbaum, A.M.; Saez, L.; Young, M.W.; Blau, J. The Double-Time Protein Kinase Regulates the Subcellular Localization of the Drosophila Clock Protein Period. J. Neurosci. 2005, 25, 5430–5437. [Google Scholar] [CrossRef]
  158. Nawathean, P.; Rosbash, M. The Doubletime and CKII Kinases Collaborate to Potentiate Drosophila PER Transcriptional Repressor Activity. Mol. Cell 2004, 13, 213–223. [Google Scholar] [CrossRef]
  159. Muskus, M.J.; Preuss, F.; Fan, J.-Y.; Bjes, E.S.; Price, J.L. Drosophila DBT Lacking Protein Kinase Activity Produces Long-Period and Arrhythmic Circadian Behavioral and Molecular Rhythms. Mol. Cell. Biol. 2007, 27, 8049–8064. [Google Scholar] [CrossRef]
  160. Hurley, J.M.; Jankowski, M.S.; Santos, H.D.L.; Crowell, A.M.; Fordyce, S.B.; Zucker, J.D.; Kumar, N.; Purvine, S.O.; Robinson, E.W.; Shukla, A.; et al. Circadian Proteomic Analysis Uncovers Mechanisms of Post-Transcriptional Regulation in Metabolic Pathways. Cell Syst. 2018, 7, 613–626.e5. [Google Scholar] [CrossRef]
  161. Gorl, M. A PEST-like Element in FREQUENCY Determines the Length of the Circadian Period in Neurospora Crassa. EMBO J. 2001, 20, 7074–7084. [Google Scholar] [CrossRef] [PubMed]
  162. Kelliher, C.M.; Lambreghts, R.; Xiang, Q.; Baker, C.L.; Loros, J.J.; Dunlap, J.C. PRD-2 Directly Regulates Casein Kinase I and Counteracts Nonsense-Mediated Decay in the Neurospora Circadian Clock. eLife 2020, 9, e64007. [Google Scholar] [CrossRef] [PubMed]
  163. Ni, W.; Xu, S.-L.; González-Grandío, E.; Chalkley, R.J.; Huhmer, A.F.R.; Burlingame, A.L.; Wang, Z.-Y.; Quail, P.H. PPKs Mediate Direct Signal Transfer from Phytochrome Photoreceptors to Transcription Factor PIF3. Nat. Commun. 2017, 8, 15236. [Google Scholar] [CrossRef] [PubMed]
  164. Fish, K.J.; Cegielska, A.; Getman, M.E.; Landes, G.M.; Virshup, D.M. Isolation and Characterization of Human Casein Kinase I∊ (CKI), a Novel Member of the CKI Gene Family. J. Biol. Chem. 1995, 270, 14875–14883. [Google Scholar] [CrossRef] [PubMed]
  165. Dey, J.; Carr, A.-J.F.; Cagampang, F.R.A.; Semikhodskii, A.S.; Loudon, A.S.I.; Hastings, M.H.; Maywood, E.S. The Tau Mutation in the Syrian Hamster Differentially Reprograms the Circadian Clock in the SCN and Peripheral Tissues. J. Biol. Rhythm. 2005, 20, 99–110. [Google Scholar] [CrossRef] [PubMed]
  166. Ralph, M.R.; Menaker, M. A Mutation of the Circadian System in Golden Hamsters. Science 1988, 241, 1225–1227. [Google Scholar] [CrossRef] [PubMed]
  167. Meng, Q.-J.; Logunova, L.; Maywood, E.S.; Gallego, M.; Lebiecki, J.; Brown, T.M.; Sládek, M.; Semikhodskii, A.S.; Glossop, N.R.J.; Piggins, D.; et al. Setting Clock Speed in Mammals: The CK1ε Tau Mutation in Mice Accelerates the Circadian Pacemaker by Selectively Destabilizing PERIOD Proteins. Neuron 2013, 58, 78–88. [Google Scholar] [CrossRef]
  168. Loudon, A.S.I.; Meng, Q.J.; Maywood, E.S.; Bechtold, D.A.; Boot-Handford, R.P.; Hastings, M.H. The Biology of the Circadian Ck1ε Tau Mutation in Mice and Syrian Hamsters: A Tale of Two Species. Cold Spring Harb. Symp. Quant. Biol. 2007, 72, 261–271. [Google Scholar] [CrossRef]
  169. An, Y.; Yuan, B.; Xie, P.; Gu, Y.; Liu, Z.; Wang, T.; Li, Z.; Xu, Y.; Liu, Y. Decoupling PER Phosphorylation, Stability and Rhythmic Expression from Circadian Clock Function by Abolishing PER-CK1 Interaction. Nat. Commun. 2022, 13, 3991. [Google Scholar] [CrossRef]
  170. Liu, Y.; Loros, J.; Dunlap, J.C. Phosphorylation of the Neurospora Clock Protein FREQUENCY Determines Its Degradation Rate and Strongly Influences the Period Length of the Circadian Clock. Proc. Natl. Acad. Sci. USA 2000, 97, 234–239. [Google Scholar] [CrossRef]
  171. Yang, Y.; Cheng, P.; He, Q.; Wang, L.; Liu, Y. Phosphorylation of FREQUENCY Protein by Casein Kinase II Is Necessary for the Function of the Neurospora Circadian Clock. Mol. Cell. Biol. 2003, 23, 6221–6228. [Google Scholar] [CrossRef] [PubMed]
  172. Mehra, A.; Shi, M.; Baker, C.L.; Colot, H.V.; Loros, J.J.; Dunlap, J.C. A Role for Casein Kinase 2 in the Mechanism Underlying Circadian Temperature Compensation. Cell 2009, 137, 749–760. [Google Scholar] [CrossRef] [PubMed]
  173. Yang, L.-M.; Rinke, R.; Korbmacher, C. Stimulation of the Epithelial Sodium Channel (ENaC) by cAMP Involves Putative ERK Phosphorylation Sites in the C Termini of the Channel’s β- and γ-Subunit. J. Biol. Chem. 2006, 281, 9859–9868. [Google Scholar] [CrossRef] [PubMed]
  174. Maier, B.; Wendt, S.; Vanselow, J.T.; Wallach, T.; Reischl, S.; Oehmke, S.; Schlosser, A.; Kramer, A. A Large-Scale Functional RNAi Screen Reveals a Role for CK2 in the Mammalian Circadian Clock. Genes Dev. 2009, 23, 708–718. [Google Scholar] [CrossRef] [PubMed]
  175. Tamaru, T.; Hirayama, J.; Isojima, Y.; Nagai, K.; Norioka, S.; Takamatsu, K.; Sassone-Corsi, P. CK2α Phosphorylates BMAL1 to Regulate the Mammalian Clock. Nat. Struct. Mol. Biol. 2009, 16, 446–448. [Google Scholar] [CrossRef] [PubMed]
  176. Portolés, S.; Más, P. Altered Oscillator Function Affects Clock Resonance and Is Responsible for the Reduced Day-length Sensitivity of CKB4 Overexpressing Plants. Plant J. 2007, 51, 966–977. [Google Scholar] [CrossRef] [PubMed]
  177. Sugano, S.; Andronis, C.; Green, R.M.; Wang, Z.-Y.; Tobin, E.M. Protein Kinase CK2 Interacts with and Phosphorylates the Arabidopsis Circadian Clock-Associated 1 Protein. Proc. Natl. Acad. Sci. USA 1998, 95, 11020–11025. [Google Scholar] [CrossRef]
  178. Perales, M.; Portolés, S.; Más, P. The Proteasome-dependent Degradation of CKB4 Is Regulated by the Arabidopsis Biological Clock. Plant J. 2006, 46, 849–860. [Google Scholar] [CrossRef]
  179. Daniel, X.; Sugano, S.; Tobin, E.M. CK2 Phosphorylation of CCA1 Is Necessary for Its Circadian Oscillator Function in Arabidopsis. Proc. Natl. Acad. Sci. USA 2004, 101, 3292–3297. [Google Scholar] [CrossRef]
  180. Hirota, T.; Lewis, W.G.; Liu, A.C.; Lee, J.W.; Schultz, P.G.; Kay, S.A. A Chemical Biology Approach Reveals Period Shortening of the Mammalian Circadian Clock by Specific Inhibition of GSK-3β. Proc. Natl. Acad. Sci. USA 2008, 105, 20746–20751. [Google Scholar] [CrossRef]
  181. Sahar, S.; Zocchi, L.; Kinoshita, C.; Borrelli, E.; Sassone-Corsi, P. Regulation of BMAL1 Protein Stability and Circadian Function by GSK3β-Mediated Phosphorylation. PLoS ONE 2010, 5, e8561. [Google Scholar] [CrossRef] [PubMed]
  182. Kwon, I.; Lee, J.; Chang, S.H.; Jung, N.C.; Lee, B.J.; Son, G.H.; Kim, K.; Lee, K.H. BMAL1 Shuttling Controls Transactivation and Degradation of the CLOCK/BMAL1 Heterodimer. Mol. Cell. Biol. 2006, 26, 7318–7330. [Google Scholar] [CrossRef] [PubMed]
  183. Harada, Y.; Sakai, M.; Kurabayashi, N.; Hirota, T.; Fukada, Y. Ser-557-Phosphorylated mCRY2 Is Degraded upon Synergistic Phosphorylation by Glycogen Synthase Kinase-3β. J. Biol. Chem. 2005, 280, 31714–31721. [Google Scholar] [CrossRef] [PubMed]
  184. Yin, L.; Wang, J.; Klein, P.S.; Lazar, M.A. Nuclear Receptor Rev-Erbalpha Is a Critical Lithium-Sensitive Component of the Circadian Clock. Science 2006, 311, 1002–1005. [Google Scholar] [CrossRef] [PubMed]
  185. Spengler, M.L.; Kuropatwinski, K.K.; Schumer, M.; Antoch, M. A Serine Cluster Mediates BMAL1-Dependent CLOCK Phosphorylation and Degradation. Cell Cycle 2009, 8, 4138–4146. [Google Scholar] [CrossRef] [PubMed]
  186. Lu, Y.; Zheng, X.; Hu, W.; Bian, S.; Zhang, Z.; Tao, D.; Liu, Y.; Ma, Y. Cancer/Testis Antigen PIWIL2 Suppresses Circadian Rhythms by Regulating the Stability and Activity of BMAL1 and CLOCK. Oncotarget 2017, 8, 54913–54924. [Google Scholar] [CrossRef] [PubMed]
  187. Lipton, J.O.; Yuan, E.D.; Boyle, L.M.; Ebrahimi-Fakhari, D.; Kwiatkowski, E.; Nathan, A.; Güttler, T.; Davis, F.; Asara, J.M.; Sahin, M. The Circadian Protein BMAL1 Regulates Translation in Response to S6K1-Mediated Phosphorylation. Cell 2015, 161, 1138–1151. [Google Scholar] [CrossRef]
  188. Gao, P.; Yoo, S.-H.; Lee, K.-J.; Rosensweig, C.; Takahashi, J.S.; Chen, B.P.; Green, C.B. Phosphorylation of the Cryptochrome 1 C-Terminal Tail Regulates Circadian Period Length. J. Biol. Chem. 2013, 288, 35277–35286. [Google Scholar] [CrossRef]
  189. Brenna, A.; Albrecht, U. Phosphorylation and Circadian Molecular Timing. Front. Physiol. 2020, 11, 612510. [Google Scholar] [CrossRef]
  190. Gery, S.; Komatsu, N.; Baldjyan, L.; Yu, A.; Koo, D.; Koeffler, H.P. The Circadian Gene Per1 Plays an Important Role in Cell Growth and DNA Damage Control in Human Cancer Cells. Mol. Cell 2006, 22, 375–382. [Google Scholar] [CrossRef]
  191. Pregueiro, A.M.; Liu, Q.; Baker, C.L.; Dunlap, J.C.; Loros, J.J. The Neurospora Checkpoint Kinase 2: A Regulatory Link between the Circadian and Cell Cycles. Science 2006, 313, 644–649. [Google Scholar] [CrossRef] [PubMed]
  192. Lamia, K.A.; Sachdeva, U.M.; DiTacchio, L.; Williams, E.C.; Alvarez, J.G.; Egan, D.F.; Vasquez, D.S.; Juguilon, H.; Panda, S.; Shaw, R.J.; et al. AMPK Regulates the Circadian Clock by Cryptochrome Phosphorylation and Degradation. Science 2009, 326, 437–440. [Google Scholar] [CrossRef] [PubMed]
  193. Yang, Y.; Cheng, P.; Zhi, G.; Liu, Y. Identification of a Calcium/Calmodulin-Dependent Protein Kinase That Phosphorylates the Neurospora Circadian Clock Protein FREQUENCY. J. Biol. Chem. 2001, 276, 41064–41072. [Google Scholar] [CrossRef] [PubMed]
  194. Kwak, Y.; Jeong, J.; Lee, S.; Park, Y.-U.; Lee, S.-A.; Han, D.-H.; Kim, J.-H.; Ohshima, T.; Mikoshiba, K.; Suh, Y.-H.; et al. Cyclin-Dependent Kinase 5 (Cdk5) Regulates the Function of CLOCK Protein by Direct Phosphorylation. J. Biol. Chem. 2013, 288, 36878–36889. [Google Scholar] [CrossRef] [PubMed]
  195. Yang, Y.; He, Q.; Cheng, P.; Wrage, P.; Yarden, O.; Liu, Y. Distinct Roles for PP1 and PP2A in the Neurospora Circadian Clock. Genes Dev. 2004, 18, 255–260. [Google Scholar] [CrossRef] [PubMed]
  196. Klemz, S.; Wallach, T.; Korge, S.; Rosing, M.; Klemz, R.; Maier, B.; Fiorenza, N.C.; Kaymak, I.; Fritzsche, A.K.; Herzog, E.D.; et al. Protein Phosphatase 4 Controls Circadian Clock Dynamics by Modulating CLOCK/BMAL1 Activity. Genes Dev. 2021, 35, 1161–1174. [Google Scholar] [CrossRef] [PubMed]
  197. Parvathy, S.T.; Udayasuriyan, V.; Bhadana, V. Codon Usage Bias. Mol. Biol. Rep. 2022, 49, 539–565. [Google Scholar] [CrossRef]
  198. Hurley, J.M.; Dunlap, J.C. A Fable of Too Much Too Fast. Nature 2013, 495, 57–58. [Google Scholar] [CrossRef]
  199. Zhou, M.; Guo, J.; Cha, J.; Chae, M.; Chen, S.; Barral, J.M.; Sachs, M.S.; Liu, Y. Non-Optimal Codon Usage Affects Expression, Structure and Function of Clock Protein FRQ. Nature 2013, 495, 111–115. [Google Scholar] [CrossRef]
  200. Fu, J.; Murphy, K.A.; Zhou, M.; Li, Y.H.; Lam, V.H.; Tabuloc, C.A.; Chiu, J.C.; Liu, Y. Codon Usage Affects the Structure and Function of the Drosophila Circadian Clock Protein PERIOD. Genes Dev. 2016, 30, 1761–1775. [Google Scholar] [CrossRef]
  201. Xu, Y.; Ma, P.; Shah, P.; Rokas, A.; Liu, Y.; Johnson, C.H. Non-Optimal Codon Usage Is a Mechanism to Achieve Circadian Clock Conditionality. Nature 2013, 495, 116–120. [Google Scholar] [CrossRef] [PubMed]
  202. Zhou, M.; Wang, T.; Fu, J.; Xiao, G.; Liu, Y. Nonoptimal Codon Usage Influences Protein Structure in Intrinsically Disordered Regions. Mol. Microbiol. 2015, 97, 974–987. [Google Scholar] [CrossRef] [PubMed]
  203. Zhuang, Y.; Li, Z.; Xiong, S.; Sun, C.; Li, B.; Wu, S.A.; Lyu, J.; Shi, X.; Yang, L.; Chen, Y.; et al. Circadian Clocks Are Modulated by Compartmentalized Oscillating Translation. Cell 2023, 186, 3245–3260.e23. [Google Scholar] [CrossRef] [PubMed]
  204. Pelham, J.F.; Dunlap, J.C.; Hurley, J.M. Intrinsic Disorder Is an Essential Characteristic of Components in the Conserved Circadian Circuit. Cell Commun. Signal. 2020, 18, 181. [Google Scholar] [CrossRef] [PubMed]
  205. Zhu, K.; Celwyn, I.J.; Guan, D.; Xiao, Y.; Wang, X.; Hu, W.; Jiang, C.; Cheng, L.; Casellas, R.; Lazar, M.A. An Intrinsically Disordered Region Controlling Condensation of a Circadian Clock Component and Rhythmic Transcription in the Liver. Mol. Cell 2023, 83, 3457–3469.e7. [Google Scholar] [CrossRef] [PubMed]
  206. Xia, K.; Li, S.; Yang, Y.; Shi, X.; Zhao, B.; Lv, L.; Xin, Z.; Kang, J.; Ren, P.; Wu, H. Cryptochrome 2 Acetylation Attenuates Its Antiproliferative Effect in Breast Cancer. Cell Death Dis. 2023, 14, 250. [Google Scholar] [CrossRef] [PubMed]
  207. Hanover, J.A.; Krause, M.W.; Love, D.C. The Hexosamine Signaling Pathway: O-GlcNAc Cycling in Feast or Famine. Biochim. Biophys. Acta 2010, 1800, 80–95. [Google Scholar] [CrossRef]
  208. Kim, E.Y.; Jeong, E.H.; Park, S.; Jeong, H.-J.; Edery, I.; Cho, J.W. A Role for O -GlcNAcylation in Setting Circadian Clock Speed. Genes Dev. 2012, 26, 490–502. [Google Scholar] [CrossRef]
  209. Li, Y.H.; Liu, X.; Vanselow, J.T.; Zheng, H.; Schlosser, A.; Chiu, J.C. O-GlcNAcylation of PERIOD Regulates Its Interaction with CLOCK and Timing of Circadian Transcriptional Repression. PLOS Genet. 2019, 15, e1007953. [Google Scholar] [CrossRef]
  210. Li, M.-D.; Ruan, H.-B.; Hughes, M.E.; Lee, J.-S.; Singh, J.P.; Jones, S.P.; Nitabach, M.N.; Yang, X. O-GlcNAc Signaling Entrains the Circadian Clock by Inhibiting BMAL1/CLOCK Ubiquitination. Cell Metab. 2013, 17, 303–310. [Google Scholar] [CrossRef]
  211. Kaasik, K.; Kivimäe, S.; Allen, J.J.; Chalkley, R.J.; Huang, Y.; Baer, K.; Kissel, H.; Burlingame, A.L.; Shokat, K.M.; Ptáček, L.J.; et al. Glucose Sensor O-GlcNAcylation Coordinates with Phosphorylation to Regulate Circadian Clock. Cell Metab. 2013, 17, 291–302. [Google Scholar] [CrossRef] [PubMed]
  212. Hart, G.W. How Sugar Tunes Your Clock. Cell Metab. 2013, 17, 155–156. [Google Scholar] [CrossRef] [PubMed]
  213. Liu, N.; Tian, H.; Yu, Z.; Zhao, H.; Li, W.; Sang, D.; Lin, K.; Cui, Y.; Liao, M.; Xu, Z.; et al. A Highland-Adaptation Mutation of the Epas1 Protein Increases Its Stability and Disrupts the Circadian Clock in the Plateau Pika. Cell Rep. 2022, 39, 110816. [Google Scholar] [CrossRef] [PubMed]
  214. Yagita, K.; Yamaguchi, S.; Tamanini, F.; Horst, G.T.J.V.D.; Hoeijmakers, J.H.J.; Yasui, A.; Loros, J.J.; Dunlap, J.C.; Okamura, H. Dimerization and Nuclear Entry of mPER Proteins in Mammalian Cells. Genes Dev. 2000, 14, 1353–1363. [Google Scholar] [CrossRef] [PubMed]
  215. Zhang, L.; Hirano, A.; Hsu, P.-K.; Jones, C.R.; Sakai, N.; Okuro, M.; McMahon, T.; Yamazaki, M.; Xu, Y.; Saigoh, N.; et al. A PERIOD3 Variant Causes a Circadian Phenotype and Is Associated with a Seasonal Mood Trait. Proc. Natl. Acad. Sci. USA 2016, 113, E1536–E1544. [Google Scholar] [CrossRef] [PubMed]
  216. Liu, T.; Mahesh, G.; Yu, W.; Hardin, P.E. CLOCK Stabilizes CYCLE to Initiate Clock Function in Drosophila. Proc. Natl. Acad. Sci. USA 2017, 114, 10972–10977. [Google Scholar] [CrossRef] [PubMed]
  217. Malzahn, E.; Ciprianidis, S.; Káldi, K.; Schafmeier, T.; Brunner, M. Photoadaptation in Neurospora by Competitive Interaction of Activating and Inhibitory LOV Domains. Cell 2010, 142, 762–772. [Google Scholar] [CrossRef] [PubMed]
  218. Xue, Y.; Chiu, J.C.; Zhang, Y. SUR-8 Interacts with PP1-87B to Stabilize PERIOD and Regulate Circadian Rhythms in Drosophila. PLoS Genet. 2019, 15, e1008475. [Google Scholar] [CrossRef] [PubMed]
  219. Berthier, A.; Vinod, M.; Porez, G.; Steenackers, A.; Alexandre, J.; Yamakawa, N.; Gheeraert, C.; Ploton, M.; Maréchal, X.; Dubois-Chevalier, J.; et al. Combinatorial Regulation of Hepatic Cytoplasmic Signaling and Nuclear Transcriptional Events by the OGT/REV-ERBα Complex. Proc. Natl. Acad. Sci. USA 2018, 115, E11033–E11042. [Google Scholar] [CrossRef]
  220. Hardin, P.E. Molecular Genetic Analysis of Circadian Timekeeping in Drosophila. Adv. Genet. 2011, 74, 141–173. [Google Scholar] [CrossRef]
  221. Hunter-Ensor, M.; Ousley, A.; Sehgal, A. Regulation of the Drosophila Protein Timeless Suggests a Mechanism for Resetting the Circadian Clock by Light. Cell 1996, 84, 677–685. [Google Scholar] [CrossRef] [PubMed]
  222. Peschel, N.; Chen, K.F.; Szabo, G.; Stanewsky, R. Light-Dependent Interactions between the Drosophila Circadian Clock Factors Cryptochrome, Jetlag, and Timeless. Curr. Biol. 2009, 19, 241–247. [Google Scholar] [CrossRef] [PubMed]
  223. Hara, T.; Koh, K.; Combs, D.J.; Sehgal, A. Post-Translational Regulation and Nuclear Entry of TIMELESS and PERIOD Are Affected in New Timeless Mutant. J. Neurosci. 2011, 31, 9982–9990. [Google Scholar] [CrossRef] [PubMed]
  224. Talora, C.; Franchi, L.; Linden, H.; Ballario, P.; Macino, G. Role of a White Collar-1-White Collar-2 Complex in Blue-Light Signal Transduction. EMBO J. 1999, 18, 4961–4968. [Google Scholar] [CrossRef] [PubMed]
  225. Valverde, F.; Mouradov, A.; Soppe, W.; Ravenscroft, D.; Samach, A.; Coupland, G. Photoreceptor Regulation of CONSTANS Protein in Photoperiodic Flowering. Science 2004, 303, 1003–1006. [Google Scholar] [CrossRef] [PubMed]
  226. Shalitin, D.; Yang, H.; Mockler, T.C.; Maymon, M.; Guo, H.; Whitelam, G.C.; Lin, C. Regulation of Arabidopsis Cryptochrome 2 by Blue-Light-Dependent Phosphorylation. Nature 2002, 417, 763–767. [Google Scholar] [CrossRef] [PubMed]
  227. Srikanta, S.B.; Cermakian, N. To Ub or Not to Ub: Regulation of Circadian Clocks by Ubiquitination and Deubiquitination. J. Neurochem. 2021, 157, 11–30. [Google Scholar] [CrossRef]
  228. Wang, X.; Ma, L. Unraveling the Circadian Clock in Arabidopsis. Plant Signal Behav. 2013, 8, e23014. [Google Scholar] [CrossRef]
  229. Lee, C.; Etchegaray, J.-P.; Cagampang, F.R.A.; Loudon, A.S.I.; Reppert, S.M. Posttranslational Mechanisms Regulate the Mammalian Circadian Clock. Cell 2001, 107, 855–867. [Google Scholar] [CrossRef]
  230. Besing, R.C.; Paul, J.R.; Hablitz, L.M.; Rogers, C.O.; Johnson, R.L.; Young, M.E.; Gamble, K.L. Circadian Rhythmicity of Active GSK3 Isoforms Modulates Molecular Clock Gene Rhythms in the Suprachiasmatic Nucleus. J. Biol. Rhythms 2015, 30, 155–160. [Google Scholar] [CrossRef]
  231. Fujiwara, S.; Wang, L.; Han, L.; Suh, S.-S.; Salomé, P.A.; McClung, C.R.; Somers, D.E. Post-Translational Regulation of the Arabidopsis Circadian Clock through Selective Proteolysis and Phosphorylation of Pseudo-Response Regulator Proteins. J. Biol. Chem. 2008, 283, 23073–23083. [Google Scholar] [CrossRef]
  232. Kim, W.-Y.; Geng, R.; Somers, D.E. Circadian Phase-Specific Degradation of the F-Box Protein ZTL Is Mediated by the Proteasome. Proc. Natl. Acad. Sci. USA 2003, 100, 4933–4938. [Google Scholar] [CrossRef] [PubMed]
  233. Murakami-Kojima, M.; Nakamichi, N.; Yamashino, T.; Mizuno, T. The APRR3 Component of the Clock-Associated APRR1/TOC1 Quintet Is Phosphorylated by a Novel Protein Kinase Belonging to the WNK Family, the Gene for Which Is Also Transcribed Rhythmically in Arabidopsis thaliana. Plant Cell Physiol. 2002, 43, 675–683. [Google Scholar] [CrossRef] [PubMed]
  234. Zhang, Z.-H.; Xiong, J.-M.; Zhu, Y.-Y.; Zhang, X.-D.; Wu, W.-J.; Zhou, L.; Zhuang, J.-H.; Xu, X.-H. WNK3-PER1 Interactions Regulate the Circadian Rhythm in the Suprachiasmatic Nucleus in Rats. Am. J. Transl. Res. 2022, 14, 1001–1009. [Google Scholar] [PubMed]
  235. Sathyanarayanan, S.; Zheng, X.; Xiao, R.; Sehgal, A. Posttranslational Regulation of Drosophila PERIOD Protein by Protein Phosphatase 2A. Cell 2004, 116, 603–615. [Google Scholar] [CrossRef] [PubMed]
  236. Imai, K.; Nishiwaki, T.; Kondo, T.; Iwasaki, H. Circadian Rhythms in the Synthesis and Degradation of a Master Clock Protein KaiC in Cyanobacteria. J. Biol. Chem. 2004, 279, 36534–36539. [Google Scholar] [CrossRef] [PubMed]
  237. Ruoff, P.; Vinsjevik, M.; Monnerjahn, C.; Rensing, L. The Goodwin Oscillator: On the Importance of Degradation Reactions in the Circadian Clock. J. Biol. Rhythms 1999, 14, 469–479. [Google Scholar] [CrossRef]
  238. Dunlap, J.C.; Loros, J.J.; Colot, H.V.; Mehra, A.; Belden, W.J.; Shi, M.; Hong, C.I.; Larrondo, L.F.; Baker, C.L.; Chen, C.-H.; et al. A Circadian Clock in Neurospora: How Genes and Proteins Cooperate to Produce a Sustained, Entrainable, and Compensated Biological Oscillator with a Period of about a Day. Cold Spring Harb. Symp. Quant. Biol. 2007, 72, 57–68. [Google Scholar] [CrossRef]
  239. Colot, H.V.; Loros, J.J.; Dunlap, J.C. Temperature-Modulated Alternative Splicing and Promoter Use in the Circadian Clock Gene Frequency. Mol. Biol. Cell 2005, 16, 5563–5571. [Google Scholar] [CrossRef]
  240. Diernfellner, A.C.R.; Schafmeier, T.; Merrow, M.W.; Brunner, M. Molecular Mechanism of Temperature Sensing by the Circadian Clock of Neurospora crassa. Genes Dev. 2005, 19, 1968–1973. [Google Scholar] [CrossRef]
  241. Narasimamurthy, R.; Virshup, D.M. The Phosphorylation Switch That Regulates Ticking of the Circadian Clock. Mol. Cell 2021, 81, 1133–1146. [Google Scholar] [CrossRef] [PubMed]
  242. Liu, X.; Chen, A.; Caicedo-Casso, A.; Cui, G.; Du, M.; He, Q.; Lim, S.; Kim, H.J.; Hong, C.I.; Liu, Y. FRQ-CK1 Interaction Determines the Period of Circadian Rhythms in Neurospora. Nat. Commun. 2019, 10, 4352. [Google Scholar] [CrossRef] [PubMed]
  243. Kurien, P.; Hsu, P.-K.; Leon, J.; Wu, D.; McMahon, T.; Shi, G.; Xu, Y.; Lipzen, A.; Pennacchio, L.A.; Jones, C.R.; et al. TIMELESS Mutation Alters Phase Responsiveness and Causes Advanced Sleep Phase. Proc. Natl. Acad. Sci. USA 2019, 116, 12045–12053. [Google Scholar] [CrossRef]
  244. Yildirim, E.; Chiu, J.C.; Edery, I. Identification of Light-Sensitive Phosphorylation Sites on PERIOD That Regulate the Pace of Circadian Rhythms in Drosophila. Mol. Cell. Biol. 2016, 36, 855–870. [Google Scholar] [CrossRef] [PubMed]
  245. Hastings, J.W.; Sweeney, B.M. On the Mechanism of Temperature Independence in a Biological Clock. Proc. Natl. Acad. Sci. USA 1957, 43, 804–811. [Google Scholar] [CrossRef] [PubMed]
  246. Gardner, G.F.; Feldman, J.F. Temperature Compensation of Circadian Period Length in Clock Mutants of Neurospora crassa. Plant Physiol. 1981, 68, 1244–1248. [Google Scholar] [CrossRef] [PubMed]
  247. Isojima, Y.; Nakajima, M.; Ukai, H.; Fujishima, H.; Yamada, R.G.; Masumoto, K.; Kiuchi, R.; Ishida, M.; Ukai-Tadenuma, M.; Minami, Y.; et al. CKIepsilon/Delta-Dependent Phosphorylation Is a Temperature-Insensitive, Period-Determining Process in the Mammalian Circadian Clock. Proc. Natl. Acad. Sci. USA 2009, 106, 15744–15749. [Google Scholar] [CrossRef] [PubMed]
  248. Zhou, M.; Kim, J.K.; Eng, G.W.L.; Forger, D.B.; Virshup, D.M. A Period2 Phosphoswitch Regulates and Temperature Compensates Circadian Period. Mol. Cell 2015, 60, 77–88. [Google Scholar] [CrossRef]
  249. Duffy, J.F.; Czeisler, C.A. Age-Related Change in the Relationship between Circadian Period, Circadian Phase, and Diurnal Preference in Humans. Neurosci. Lett. 2002, 318, 117–120. [Google Scholar] [CrossRef]
  250. Badura, L.; Swanson, T.; Adamowicz, W.; Adams, J.; Cianfrogna, J.; Fisher, K.; Holland, J.; Kleiman, R.; Nelson, F.; Reynolds, L.; et al. An Inhibitor of Casein Kinase I Epsilon Induces Phase Delays in Circadian Rhythms under Free-Running and Entrained Conditions. J. Pharmacol. Exp. Ther. 2007, 322, 730–738. [Google Scholar] [CrossRef]
  251. Zhang, L.; Abraham, D.; Lin, S.-T.; Oster, H.; Eichele, G.; Fu, Y.-H.; Ptáček, L.J. PKCγ Participates in Food Entrainment by Regulating BMAL1. Proc. Natl. Acad. Sci. USA 2012, 109, 20679–20684. [Google Scholar] [CrossRef] [PubMed]
  252. Ma, H.; Li, L.; Yan, J.; Zhang, Y.; Ma, X.; Li, Y.; Yuan, Y.; Yang, X.; Yang, L.; Guo, J. The Resonance and Adaptation of Neurospora Crassa Circadian and Conidiation Rhythms to Short Light-Dark Cycles. J. Fungi 2021, 8, 27. [Google Scholar] [CrossRef] [PubMed]
  253. Bechtold, D.A.; Gibbs, J.E.; Loudon, A.S.I. Circadian Dysfunction in Disease. Trends Pharmacol. Sci. 2010, 31, 191–198. [Google Scholar] [CrossRef] [PubMed]
  254. Philpott, J.M.; Freeberg, A.M.; Park, J.; Lee, K.; Ricci, C.G.; Hunt, S.R.; Narasimamurthy, R.; Segal, D.H.; Robles, R.; Cai, Y.; et al. PERIOD Phosphorylation Leads to Feedback Inhibition of CK1 Activity to Control Circadian Period. Mol. Cell 2023, 83, 1677–1692.e8. [Google Scholar] [CrossRef] [PubMed]
  255. Toh, K.L.; Jones, C.R.; He, Y.; Eide, E.J.; Hinz, W.A.; Virshup, D.M.; Ptácek, L.J.; Fu, Y.H. An hPer2 Phosphorylation Site Mutation in Familial Advanced Sleep Phase Syndrome. Science 2001, 291, 1040–1043. [Google Scholar] [CrossRef] [PubMed]
  256. Shanware, N.P.; Hutchinson, J.A.; Kim, S.H.; Zhan, L.; Bowler, M.J.; Tibbetts, R.S. Casein Kinase 1-Dependent Phosphorylation of Familial Advanced Sleep Phase Syndrome-Associated Residues Controls PERIOD 2 Stability. J. Biol. Chem. 2011, 286, 12766–12774. [Google Scholar] [CrossRef] [PubMed]
  257. Vanselow, K.; Vanselow, J.T.; Westermark, P.O.; Reischl, S.; Maier, B.; Korte, T.; Herrmann, A.; Herzel, H.; Schlosser, A.; Kramer, A. Differential Effects of PER2 Phosphorylation: Molecular Basis for the Human Familial Advanced Sleep Phase Syndrome (FASPS). Genes Dev. 2006, 20, 2660–2672. [Google Scholar] [CrossRef]
  258. Moore-Ede, M.C. Physiology of the Circadian Timing System: Predictive versus Reactive Homeostasis. Am. J. Physiol. 1986, 250, R737–R752. [Google Scholar] [CrossRef]
  259. Hirano, A.; Braas, D.; Fu, Y.; Ptáček, L.J. FAD Regulates CRYPTOCHROME Protein Stability and Circadian Clock in Mice. Cell Rep. 2017, 19, 255–266. [Google Scholar] [CrossRef]
  260. Hirano, A.; Shi, G.; Jones, C.R.; Lipzen, A.; Pennacchio, L.A.; Xu, Y.; Hallows, W.C.; McMahon, T.; Yamazaki, M.; Ptáček, L.J.; et al. A Cryptochrome 2 Mutation Yields Advanced Sleep Phase in Humans. eLife 2016, 5, e16695. [Google Scholar] [CrossRef]
  261. Hirota, T.; Lee, J.W.; Lewis, W.G.; Zhang, E.E.; Breton, G.; Liu, X.; Garcia, M.; Peters, E.C.; Etchegaray, J.-P.; Traver, D.; et al. High-Throughput Chemical Screen Identifies a Novel Potent Modulator of Cellular Circadian Rhythms and Reveals CKIα as a Clock Regulatory Kinase. PLoS Biol. 2010, 8, e1000559. [Google Scholar] [CrossRef] [PubMed]
  262. Hirota, T.; Lee, J.W.; St. John, P.C.; Sawa, M.; Iwaisako, K.; Noguchi, T.; Pongsawakul, P.Y.; Sonntag, T.; Welsh, D.K.; Brenner, D.A.; et al. Identification of Small Molecule Activators of Cryptochrome. Science 2012, 337, 1094–1097. [Google Scholar] [CrossRef]
  263. Chen, Z.; Yoo, S.-H.; Park, Y.-S.; Kim, K.-H.; Wei, S.; Buhr, E.; Ye, Z.-Y.; Pan, H.-L.; Takahashi, J.S. Identification of Diverse Modulators of Central and Peripheral Circadian Clocks by High-Throughput Chemical Screening. Proc. Natl. Acad. Sci. USA 2012, 109, 101–106. [Google Scholar] [CrossRef] [PubMed]
  264. Menculini, G.; Verdolini, N.; Murru, A.; Pacchiarotti, I.; Volpe, U.; Cervino, A.; Steardo, L.; Moretti, P.; Vieta, E.; Tortorella, A. Depressive Mood and Circadian Rhythms Disturbances as Outcomes of Seasonal Affective Disorder Treatment: A Systematic Review. J. Affect. Disord. 2018, 241, 608–626. [Google Scholar] [CrossRef] [PubMed]
  265. Schnell, A.; Sandrelli, F.; Ranc, V.; Ripperger, J.A.; Brai, E.; Alberi, L.; Rainer, G.; Albrecht, U. Mice Lacking Circadian Clock Components Display Different Mood-Related Behaviors and Do Not Respond Uniformly to Chronic Lithium Treatment. Chronobiol. Int. 2015, 32, 1075–1089. [Google Scholar] [CrossRef] [PubMed]
  266. Melo, M.C.A.; Abreu, R.L.C.; Linhares Neto, V.B.; De Bruin, P.F.C.; De Bruin, V.M.S. Chronotype and Circadian Rhythm in Bipolar Disorder: A Systematic Review. Sleep Med. Rev. 2017, 34, 46–58. [Google Scholar] [CrossRef] [PubMed]
  267. Undurraga, J.; Sim, K.; Tondo, L.; Gorodischer, A.; Azua, E.; Tay, K.H.; Tan, D.; Baldessarini, R.J. Lithium Treatment for Unipolar Major Depressive Disorder: Systematic Review. J. Psychopharmacol. 2019, 33, 167–176. [Google Scholar] [CrossRef] [PubMed]
  268. Rohr, K.E.; McCarthy, M.J. The Impact of Lithium on Circadian Rhythms and Implications for Bipolar Disorder Pharmacotherapy. Neurosci. Lett. 2022, 786, 136772. [Google Scholar] [CrossRef]
  269. Snitow, M.E.; Bhansali, R.S.; Klein, P.S. Lithium and Therapeutic Targeting of GSK-3. Cells 2021, 10, 255. [Google Scholar] [CrossRef]
  270. Li, J.; Lu, W.-Q.; Beesley, S.; Loudon, A.S.I.; Meng, Q.-J. Lithium Impacts on the Amplitude and Period of the Molecular Circadian Clockwork. PLoS ONE 2012, 7, e33292. [Google Scholar] [CrossRef]
  271. Jolma, I.W.; Falkeid, G.; Bamerni, M.; Ruoff, P. Lithium Leads to an Increased FRQ Protein Stability and to a Partial Loss of Temperature Compensation in the Neurospora Circadian Clock. J. Biol. Rhythm. 2006, 21, 327–334. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, H.; Zhou, Z.; Guo, J. The Function, Regulation, and Mechanism of Protein Turnover in Circadian Systems in Neurospora and Other Species. Int. J. Mol. Sci. 2024, 25, 2574. https://doi.org/10.3390/ijms25052574

AMA Style

Zhang H, Zhou Z, Guo J. The Function, Regulation, and Mechanism of Protein Turnover in Circadian Systems in Neurospora and Other Species. International Journal of Molecular Sciences. 2024; 25(5):2574. https://doi.org/10.3390/ijms25052574

Chicago/Turabian Style

Zhang, Haoran, Zengxuan Zhou, and Jinhu Guo. 2024. "The Function, Regulation, and Mechanism of Protein Turnover in Circadian Systems in Neurospora and Other Species" International Journal of Molecular Sciences 25, no. 5: 2574. https://doi.org/10.3390/ijms25052574

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop