Quick viewing(Text Mode)

Cellular Timekeeping: It's Redox O'clock

Cellular Timekeeping: It's Redox O'clock

Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

Cellular Timekeeping: It’s Redox o’

Nikolay B. Milev,1 Sue-Goo Rhee,2 and Akhilesh B. Reddy1,3

1The Francis Crick Institute, London NW1 1AT, United Kingdom 2Yonsei University College of Medicine, Seodaemun-gu, Seoul 120-752, Korea 3UCL Institute of Neurology, London WC1N 3BG, United Kingdom Correspondence: [email protected]

Mounting evidence in recent supports the extensive interaction between the circadian and redox systems. The existence of such a relationship is not surprising because most organisms, be they diurnal or nocturnal, display daily oscillations in energy intake, locomo- tor activity,and exposure to exogenous and internally generated oxidants. The transcriptional clock controls the levels of many antioxidant proteins and redox-active cofactors, and, conversely, the cellular redox poise has been shown to feed back to the transcriptional oscillator via redox-sensitive transcription factors and enzymes. However, the circadian cycles in the S-sulfinylation of the peroxiredoxin (PRDX) proteins constituted the first example of an autonomous circadian redox oscillation, which occurred independently of the transcriptional clock. Importantly, the high phylogenetic conservation of these rhythms suggests that they might predate the evolution of the transcriptional oscillator, and therefore could be a part of a primordial circadian redox/metabolic oscillator. This discovery forced the reappraisal of the dogmatic transcription-centered view of the clockwork and opened a new avenue of research. Indeed, the investigation into the links between the circadian and redox systems is still in its infancy, and many important questions remain to be addressed.

NOTHING NEW UNDER THE SUN: THE was based on observations of overt OF THE RESEARCH IN rhythms in a variety of organisms, but the un- NONTRANSCRIPTIONAL OSCILLATORS derlying fundamental mechanisms were still largely unknown. It was clear that because uni- he concept that the central mechanism that cellular organisms show daily oscillations, the Tdrives the circadian clock might be based on timekeeping mechanism must be essentially a a nontranscriptional process is not a recent one. cellular phenomenon. What was also evident In fact, it dates all the way back to the 1960 was that the clockwork must consist of a feed- when the first chronobiology meeting took back system capable of generating 24-h cycles in place, as part of the Cold Spring Harbor Sym- various cellular parameters. Three main models posia on Quantitative Biology. At the , most were built around this concept, providing dif- of the evidence for the existence of circadian ferent interpretations on the same theme. The

Editors: Paolo Sassone-Corsi, Michael W. Young, and Akhilesh B. Reddy Additional Perspectives on Circadian Rhythms available at www.cshperspectives.org Copyright # 2018 Cold Spring Harbor Laboratory Press; all rights reserved; doi: 10.1101/cshperspect.a027698 Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698

1 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

N.B. Milev et al.

first envisioned the core oscillator as an inter- showed the existence of nontranscriptional os- connected network of coupled biochemical cillations. This organism consists of a giant cell events occurring in fixed succession, such that (1–10 cm at maturity) that contains a single they generate a closed self-sustainable loop (the nucleus residing in a basal root-like structure, network hypothesis) (Pavlidis 1969). The sec- called rhizoid (Mandoli 1998). Interestingly, it ond model used the same basic principle but was shown that Acetabularia remains viable for favored transcription as the central mechanism several after amputation of the nucleus- (the chronon model) (Ehret and Trucco 1967). bearing rhizoid, which made the alga an impor- The third interpretation was somewhat different tant model system for studying the nuclear–cy- and considered the primary oscillator to be a toplasmic interactions (Mandoli 1998). Taking feedback system involving ion gradients and advantage of this peculiarity, it was shown that ion-transport channels (the Njus–Sulzman– enucleated Acetabularia major plants show en- Hastings membrane model) (Njus et al. 1974). dogenous rhythms in photosynthesis and chlo- The next 30 years were marked by exciting roplast shape in constant conditions (Sweeney findings that provided supporting evidence for and Haxo 1961; Driessche 1966). These results all three models (Fig. 1). However, because of were later confirmed in another two Acetabu- the discovery of numerous “clock genes” in laria species—A. medeternea (Mergenhagen diverse organisms in the 1990s, the research and Schweiger 1975a) and A. crenulata (Ter- in nontranscriptional oscillators lost its mo- borgh and McLeod 1967). Although early graft- mentum. The great efforts in finding new ing experiments (whereby the nucleus-contain- “clock genes” culminated in the development of ing rhizoid of one plant is transplanted onto the transcriptional/translation feedback loop another) indicated that the nucleus is capable (TTFL) model, which seemed to largely account of imposing control over the phase of the overt for the generation of circadian rhythmicity in rhythms, these studies lacked sufficient sam- most organisms (Hardin et al. 1990). The over- pling resolution to make firm conclusions whelming success of this model led many chro- (Schweiger et al. 1964; Vanden Driessche nobiologists to adopt an almost dogmatic view 1966). Indeed, later it was shown that when a of the core clock mechanism, ruling out any nucleus entrained in the opposite phase is rein- other possibilities. Nevertheless, some apparent troduced into the cell, the cytoplasmic oscillator anomalies to the TTFL (Lakin-Thomas 2006), dominates over the nuclear to maintain the and the seemingly divergent architecture of the phase the photosynthetic rhythm (Woolum transcriptional clock in different organisms 1991). Consistently, experiments using phar- (Young and Kay 2001) preserved the interest macological inhibitors of transcription showed in alternative cellular timekeeping models that the plants remain rhythmic for the initial long enough to give rise to the wave of 14 days of treatment, thus suggesting that tran- discoveries in the field of nontranscriptional os- scription is not required to generate rhythmic- cillators (Reddy and O’Neill 2011; van Ooijen ity, but is rather involved in maintaining the and Millar 2012; Reddy and Rey 2014). levels of key components of the oscillator (Mer- Indeed, the biggest obstacle that chronobi- genhagen and Schweiger 1975b). ologists faced in their attempt to investigate A few years after the initial experiments in nontranscriptional oscillations was the lack of Acetabularia, studies in the unicellular eukary- a straightforward procedure for the complete ote Paramecium multimicronucleatum showed exclusion of the transcriptional contributions that information about the circadian phase is from the nucleus. Nevertheless, there are several transferred during cell division over many gen- approaches and model systems that have been erations (Barnett 1966). These results were later successfully used over the years and these are validated in two additional model systems: in summarized in Table 1. the cyanobacteria, Synechococcus elongatus, In the 1960s, experiments performed in the which have a division time much shorter than unicellular alga Acetabularia for the first time 24 h (Kondo et al. 1997), and in mouse fibro-

2 Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 Downloaded from ieti ril as article this Cite http://cshperspectives.cshlp.org/ odSrn abPrpc Biol Perspect Harb Spring Cold

Timeline I: Nontranscriptional oscillators

Circadian phase Cold Spring Harbor information is retained

Symposium on during multiple cycle Two independent Autonomous redox onOctober6,2021-PublishedbyColdSpringHarborLaboratoryPress Quantitative Biology of cell division in oscillators indentified oscillations in anucleate “Biological Clocks” Paramecium in Gonyaulax human red blood cells

2018;10:a027698 1960 1961 1967 1985 1994 2005 2011

Persistence of GSH oscillations Reconstitution of photosynthetic rhythm in in anucleate platelets circadian oscillation of enucleated Acetabularia incubated in vitro cyanobacterial KaiC phosphorylation in vitro

Figure 1. of the key experiments performed in the field of nontranscriptional oscillators. The timeline begins with the first chronobiology meeting at the Cold Spring Harbor Symposia for Quantitative Biology, where the foundations of the field were laid. The next five yielded exciting discoveries in the area of nontranscriptional oscillators, culmi- Timekeeping Cellular nating with the finding of the cyanobacterial KaiC oscillator and the autonomous redox oscillations in peroxiredoxins in red blood cells (RBCs). GSH, glutathione. 3 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

N.B. Milev et al.

Table 1. Model systems and methodological approaches for studying nontranscriptional oscillations Method/approach Model system Overt rhythm References Anucleate cells (rhythms Red blood cells Rhythms in enzymatic Cornelius and Rensing measured in vitro) activity; rhythms in PRDX 1976; O’Neill and Reddy overoxidation; rhythms in 2011; Cho et al. 2014; NAD(P)H and ATP Homma et al. 2015 Platelets Rhythms in GSH in vitro Radha et al. 1985 Removal of the nucleus Acetabularia Rhythms in photosynthesis Sweeney and Haxo 1961; and chloroplast shape Mergenhagen and Schweiger 1975a “Clockless” cells/ Mouse embryonic stem Rhythms in 2-DG uptake Paulose et al. 2012 organisms cells precede TTFL expression Caenorhabditis elegans Rhythms in PRDX Olmedo et al. 2012 overoxidation; olfactory cued behavior; mRNA and protein levels Cell-free systems Kai oscillator Rhythms in KaiC Nakajima et al. 2005 phosphorylation Isolated membrane Rhythms in enzymatic activity Cornelius and Rensing fractions (RBCs) 1976 Pharmacological Acetabularia Rhythm in photosynthesis Mergenhagen and inhibition of persists Schweiger 1975b transcription and Bulla gouldina Transcription and translation Khalsa et al. 1992, 1996 translation required only in a phase- dependent manner Pharmacological REV-ERB agonist; CK1 Could be used to uncouple the Meng et al. 2010; Solt et al. manipulation of clock inhibitor transcriptional and 2012 activity metabolic oscillators Environmental Transcriptional arrest PRDX overoxidation rhythms; O’Neill et al. 2011; Feeney perturbations in Ostreococcus tauri Mg2þ rhythms et al. 2016 Treatment with light of Showed the existence of two Morse et al. 1994 different independent oscillators wavelengths in Gonyaulax Genetic manipulation of Constitutive bmal1 Rhythmicity in bmal1 Liu et al. 2008 core clock elements expression transcription is dispensable for functional TTFL Constitutive Restores molecular and Yang and Sehgal 2001 expression of tim in behavioral rhythmicity in Drosophila tim2/2 flies PRDX, Peroxiredoxin; GSH, glutathione; 2-DG, 2-deoxyglucose; TTFL, transcriptional/translation feedback loop; mRNA, messenger RNA; RBCs, red blood cells.

blasts cells (Nagoshi et al. 2004). Indeed, this nant during transcriptional arrest. spurred one of the biggest controversies in the Evidence for this model was shown in the uni- field of chronobiology. If transcription is the cellular photosynthetic alga, Ostreococcus tauri, sole timekeeping mechanisms in the cell, then in which transcription is completely abolished how do actively dividing cells keep track of time? under conditions of constant darkness (O’Neill A possible explanation is that the cell harbors a et al. 2011). Even after complete termination of second oscillator, which runs in parallel to the transcription, O. tauri cultures could restore transcriptional clock and becomes the domi- transcriptional rhythms once transferred to

4 Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

Cellular Timekeeping

constant light conditions. Perhaps, more im- tion for the overt rhythms in the nicotine ade- portantly, the phase of the oscillations is not nine dinucleotides, NADþ and NADPþ, based dictated by the transfer to light, which suggests on a purely nontranscriptional feedback loop. that a nontranscriptional mechanism runs in In this model, an increase in the levels of NADþ parallel to preserve the phase of the clock. drives Ca2þ efflux from the mitochondria, In 1974, Njus, Sulzman, and Hastings pro- which activates calmodulin in the cytoplasm. posed that the timekeeping mechanism is a pro- The activated Ca2þ-calmodulin complex, in perty of the cell membrane (Njus et al. 1974). turn, activates NADþ kinase and inhibits This model postulated that the ion gradients NADPþ phosphatase, thus generating a de- across biological membranes and the ion trans- crease in NADþ via its conversion to the phos- porters engage in a self-sustained feedback loop. phorylated form (Goto et al. 1985). Indeed, it is The most convincing results in support of this well established that the daily oscillations in model came from studies in the fruit fly Dro- Ca2þ levels are not a mere readout of clock func- sophila melanogaster, where electrical silencing tions but constitute an integral part of the time- of the pacemaker neurones causes complete loss keeping fundamental mechanism (O’Neill and of behavioral and transcriptional rhythmicity in Reddy 2012). constant darkness (Nitabach et al. 2002). How- In the late 1980s and early 1990s, experi- ever, data from other organisms did not prove ments in the unicellular dinoflagellate, Go- sufficiently compelling to attribute a central role nyaulax polyedra (Lingulodinium polyedrum), for the membrane model in the clock mecha- showed that the bioluminescence rhythms nism (Nitabach et al. 2005). It is interesting to shown by this organism are transcriptionally note that a recent study showed the presence of independent and are controlled at the transla- circadian oscillations in magnesium (Mg2þ) tional level (Morse et al. 1989). Later, it was also and potassium (Kþ) levels in cultured human shown that, in Gonyaulax, the rhythms in bio- cells and O. tauri under constant conditions. luminescence and phototaxis are powered by The latter was shown to show the rhythms two independent mechanisms as the phasing even under conditions of transcriptional arrest of the two oscillations can be segregated using (Feeney et al. 2016). Although the authors did different wavelengths of light (Roenneberg and not relate their findings to the membrane mod- Morse 1993; Morse et al. 1994). These studies el, it is plausible that a similar mechanism op- provided the first example of the coexistence of erates to generate these rhythms, at least in multiple independent oscillators in the same O. tauri. cell, which was later found to be the case for A series of experiments performed in red other unicellular organisms such as Neurospora blood cells (RBCs) in the late 1970s yielded crassa and S. elongatus (Bell-Pedersen et al. rather controversial results, mostly because of 2005). the lack of an appropriate experimental read- In the 1990s, experiments with the marine out. In 1976, Cornelius and Rensing showed mollusc, Bulla gouldiana, were used to test the rhythms in Mg2þ-dependent ATPase activity contribution of transcription and translation to in the membranes of RBCs kept in vitro (Cor- circadian rhythmicity. The optic nerves in this nelius and Rensing 1976). These results were organism show circadian oscillations in action later challenged by several studies failing to rep- potential firing, which can be recorded both in licate the observed oscillations in enzymatic vivo and in explanted eye cultures. By using activities (Mabood et al. 1978; Schrader and pharmacological inhibition of transcription Holzapfel 1980). Nevertheless, in 1985, it was and translation, it was shown that rhythmicity shown that human platelets incubated in vitro in this model system is dependent on the two display oscillations in the levels of the low-mo- processes but only during a certain time of the lecular weight thiol, glutathione (GSH) (Radha (in a phase-dependent manner). et al. 1985). In the same year, work in the single- A significant drawback of all these studies is cell flagellate, Euglena, proposed an explana- that they lacked conclusive mechanistic proof. If

Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 5 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

N.B. Milev et al.

the nucleus is not essential for maintaining cel- (O’Neill and Reddy 2011; O’Neill et al. 2011). lular rhythmicity, then what constitutes the These rhythms were found to occur in the anu- oscillator? The first compelling evidence for cleate human RBCs and in the unicellular alga an autonomous nontranscriptional oscillator O. tauri, even during transcriptional arrest. came in the early 2000s from studies in the cy- Moreover, these oscillations showed all the key anobacteria S. elongatus, the simplest organism characteristics of circadian clocks. But perhaps known to show daily oscillations. In cyanobac- the most exciting finding came a year later when teria, three genes (kaiA, kaiB, kaiC) and their it was shown that, in contrast to the divergent protein products were shown to constitute the evolution of the TTFL in different organisms, core oscillator that drives the rhythmic tran- PRDX redox oscillations are conserved across all scription of the whole genome (Nishiwaki domains of life (Edgar et al. 2012). The high et al. 2000; Nakajima et al. 2005; Tomita et al. phylogenetic conservation of the PRDX oscilla- 2005). In this three-component system, KaiC tions, together with the fact that they likely displays autokinase and autophosphatase activ- evolved to counteract oxidative stress, suggested ities; KaiA stimulates the kinase and inhibits the that this nontranscriptional timekeeping mech- phosphatase activities of KaiC, whereas KaiB anism might be as ancient as the first aerobic opposes the action of KaiA. The complex dy- organisms (O’Neill and Reddy 2011; O’Neill namic interactions between these three proteins et al. 2011; Edgar et al. 2012). These findings result in rhythms in the phosphorylation state of revived the interest in metabolic oscillators as KaiC, which, in turn, drives rhythmic genome- cellular timekeeping mechanisms and has wide transcriptional changes, including that of forced the reappraisal of the gene-centric model the kaiBC operon. Later, it was shown that the of the clockwork. rhythmic transcription of the kaiBC operon was Shortly after the discovery of the redox cycle not required for the oscillations in the phos- in PRDXs, it was shown that, in undifferentiat- phorylation of KaiC, suggesting that the cyano- ed mouse embryonic stem cells (mESCs), bacterial timekeeping mechanism does not use rhythms in the uptake of the glucose analog 2- a transcription–translation feedback loop, but deoxyglucose (2-DG) precede the development rather a purely nontranscriptional mechanism of a functional TTFL (Paulose et al. 2012). Even (Tomita et al. 2005). In a series of elegant bio- more interesting is the fact that these apparently chemical experiments, Nakajima and colleagues “clockless” cells showed rhythms in the tran- showed that the 24-h cycles of alternating phos- script of the glucose transporter, mGlut8. This phorylation and dephosphorylation of the KaiC observation implied that a putative metabolic protein could be reconstituted in vitro using a oscillator might drive rhythmic expression of simple four-component system consisting of certain genes, analogous to the Kai oscillator. the three Kai proteins and ATP (Nakajima et al. 2005). Although the Kai proteins are not conserved across taxa, these results provided a PEROXIREDOXINS: GUARDIANS AGAINST novel paradigm in circadian timekeeping; a OXIDATIVE STRESS, REDOX SIGNALING purely nontranscriptional process acts as the HUBS, AND CELLULAR primary oscillator, and this biochemical cycle Peroxiredoxins Are the Major Target drives rhythms in gene expression. of H O in the Cell Despite the general skepticism in the field 2 2 that nontranscriptional oscillators are compat- PRDXs are a large and evolutionarily conserved ible only with “simple” organisms, such as cya- family of thiol peroxidases that serve to main- nobacteria, the search for an evolutionary con- tain the cellular redox tone by keeping the levels served nontranscriptional oscillator continued. of reactive oxygen and nitrogen species (ROS/ In 2011, this search came to an end with the RNS) in check (Rhee 2016). PRDXs are also discovery of the circadian redox cycles in the increasingly being recognized for their role in antioxidant peroxiredoxin (PRDX) proteins integrating redox signaling into cellular physi-

6 Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

Cellular Timekeeping

ology by acting as peroxide sensors (Netto and the CysP does not rely on a partnering cysteine, Antunes 2016). The hallmark feature of this and the resolving power originates from low- family of antioxidant proteins is a conserved molecular-weight reductants such as ascorbate cysteine residue in the active site (also termed or GSH (Kang et al. 1998; Manevich et al. 2004). peroxidatic cysteine, CysP), which displays Mammalian cells express four typical 2-Cys unprecedentedly high reactivity toward biolog- PRDXs (PRDX1-4), one atypical (PRDX5), ical peroxides. The PRDXs owe their incredible and a single 1-Cys PRDX (PRDX6) (Rhee specificity and reactivity toward H2O2 to the et al. 2001). The different isoforms display highly conserved active site. This not only broad cellular distribution with PRDX 1, 2, activates the catalytic CysP in the reactive thio- and 6 being exclusively nucleocytoplasmic, late form but also stabilizes the transition PRDX3 located in the mitochondrion, PRDX4 state during catalysis (Perkins et al. 2015). residing in the endoplasmic reticulum (ER), This specificity dictates remarkably high reac- and PRDX5 being in the mitochondri- 7 9 21 21 tion rates on the order of 10 to 10 M sec , on, peroxisomes, and cytosol (Park et al. 2014). which make PRDXs the preferential targets for When the reactive cysteines of the 2-Cys H2O2 in the cell (Randall et al. 2013). Interest- PRDXs are in the disulfide form, the proteins ingly, the specificity of PRDXs to H2O2 is also are inactive and are recycled by the thioredoxin reflected in their relatively poor affinity toward (TRX) system. The redox-active cysteine couple other electrophiles, such as iodoacetamide in TRX extracts the disulfide bond from the (Peskin et al. 2007). Although the main sub- PRDXs via a thiol/disulfide exchange mecha- strate of PRDXs is H2O2, some members of nism and transfers it to the flavoprotein TRX the family react preferentially with peroxynitrite reductase (TrxR) (Chae et al. 1994). The oxi- and lipid peroxide species (Fisher 2011; Knoops dized TrxR utilizes a charge relay system to et al. 2011). transfer from nicotinamide adenine During catalysis, the reactive cysteine per- dinucleotide phosphate (NADPH) to the re- forms a nucleophilic attack on one of the oxy- dox-active cofactor flavin adenine dinucleotide gens in the H2O2 molecule, which leads to the (FADH) and finally to the reactive cysteine cou- oxidation of the CysP sulfhydryl (-SH) to a sul- ple in the active site of the reductase. Thus, the fenic acid (-SOH) intermediate. Although there TRX system (including the PRDXs) acts to are different classification systems that subdi- channel electrons from NADPH to H2O2,re- vide this diverse family of peroxidases (Poole ducing the latter to water (Fig. 2). and Nelson 2016), the most well-known criteri- In 99.3% of the time under unstressed con- on is the mechanism by which the catalytic in- ditions, the catalytic cycle of the PRDXs pro- termediate is resolved, which groups the PRDXs ceeds via the above-described fast catalytic into typical 2-Cys, atypical 2-Cys, and 1-Cys loop (Perkins et al. 2015). However, a rather (Rhee et al. 2001). In the typical 2-Cys PRDXs, unusual characteristic of the PRDXs is that the the CysP reacts with another conserved cysteine CysP sulfenic acid intermediate can undergo residue (termed the resolving cysteine, CysR) further oxidation, termed over/hyperoxidation, residing on a second PRDX monomer to form to form a sulfinic acid (-SO2H) (Rhee 2016). an intermolecular disulfide bond. The PRDXs This form of the enzyme is catalytically inactive from this subgroup exist as noncovalent homo- and is slowly regenerated back to the active sulf- dimers in a head-to-tail configuration, and thus hydryl form via an ATP-consuming reaction the CysP of one PRDX molecule is ideally catalyzed by the only known cellular sulfinic aligned with the CysR of the other monomer. acid reductase, sulfiredoxin (SRX) (Biteau et In the atypical 2-Cys PRDX class, the resolving al. 2003; Woo et al. 2003b; Rhee et al. 2007). cysteine is positioned on the same PRDX mol- The slow reaction rates of the recycling reaction ecule and thus the reaction between CysP and allow the hyperoxidized species to persist in the CysR results in an intramolecular disulfide cell on a timescale of , thus suggesting a bond. In the third group, the 1-Cys PRDXs, functional role for this form of the enzyme. In

Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 7 Downloaded from ..Mlve al. et Milev N.B. 8 http://cshperspectives.cshlp.org/

– S H2O2 PRDX HS SH SH PRDX S– + TRX Peroxide NADP SH scavenging

TrxR TRX loop

H2O S PRDX catalytic loop NADPH TRX S H2O2 onOctober6,2021-PublishedbyColdSpringHarborLaboratoryPress SOH S PRDX H2O PRDX HS SH HS PRDX S S– ieti ril as article this Cite PRDX S– PRDX inactivation loop

SO2H PRDX HS SH PRDX SRX S–

odSrn abPrpc Biol Perspect Harb Spring Cold ADP

ATP

Figure 2. Peroxiredoxins (PRDXs) are the primary H2O2 scavengers. The catalytic loop of 2-Cys PRDXs begins with the reaction between the reactive peroxidatic cysteine residues (CysP) and an incoming peroxide molecule (H2O2). This reaction results in the oxidation of CysP to a sulfenic acid (-SOH) intermediate and the detoxification of H2O2 to water. The sulfenic acid intermediate is relatively unstable and can continue its reaction route either via the catalytic or the inactivation loop. In the first case, the sulfenic acid intermediate reacts with another conserved residue, termed the resolving cysteine (CysR), originating from a second PRDX monomer to form an intramolecular disulfide bond. The thioredoxin (TRX) system acts to recycle the disulfide form of the enzyme via the action of TRX and TrxR, with electrons 2018;10:a027698 originating from nicotinamide adenine dinucleotide phosphate (NADPH). Occasionally, the sulfenic acid intermediate reacts with a second peroxide molecule, which results in the hyperoxidation to sulfinic acid (SO2H). This form of the enzyme is slowly recycled back to the active thiolate form by the enzyme sulfiredoxin (SRX) in an ATP-consuming reaction. The inactivation loop is considerably slower than the catalytic loop, which allows the accumulation of overoxidized PRDXs. The proportion of PRDX molecules entering the two pathways is dictated by intrinsic factors (e.g., structural character- istics) and dynamic parameters (the local H2O2 concentration). Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

Cellular Timekeeping

response to increased ROS levels or exogenous two forms of the protein displayed antiphasic application of H2O2, the relative proportion of oscillations, which indicated that PRDX6 is the hyperoxidized species increases and is fre- rhythmically modified. Initially it was suggested quently used as an indirect measure of oxidative that the forms could represent different phos- stress (Poynton and Hampton 2014). Initially, phorylation states of PRDX6, mostly because of the concept that an enzyme is inactivated by an the prevalence of this form of posttranslational increase in the levels of its substrate seemed regulation in the clockwork. However, around counterintuitive. However, soon it became ap- the same time, a series of studies from the Rhee parent that this paradox is, in fact, an evolution- laboratory identified the mechanism of revers- ary adaptation that allows the PRDX proteins to ible inactivation of PRDXs via hyperoxidation fulfill a broad range of cellular functions in sig- (Yanget al. 2002). More importantly, taking ad- nal transduction networks (Park et al. 2014) as vantage of the fact that the S-sulfinylation form molecular chaperones (Jang et al. 2004; Bar- of the enzyme is relatively stable and persists for ranco-Medina et al. 2009) and as intermediates several hours in vivo, the authors developed an between redox metabolism and the circadian antibody directed specifically against this form clock. The reason for the hyperoxidation lies of the enzymes (Woo et al. 2003a). The subse- in the structural rearrangement that is required quent turn of events can only be described with for disulfide bond formation (Wood et al. the words “applying the right tool in the right 2003). This reconfiguration allows the solvent model systems, asking the right questions.” exposure of the otherwise shielded sulfenic acid In 2011, it was shown that two of the most intermediate and therefore the occasional fur- powerful models for investigating nontran- ther oxidation. Therefore, the resolution reac- scriptional oscillations—the anucleate RBCs tion of the sulfenic acid intermediate is in equi- and the photosynthetic alga, O. tauri—show librium between disulfide bond formation and robust circadian oscillations in the S-sulfiny- hyperoxidation, and the factors that dictate the lated form of PRDXs. Importantly, the two direction include intrinsic biochemical charac- model systems allow the addressing of two crit- teristics, such as the rate of the unfolding and ical issues: (1) do nontranscriptional rhythms dynamic parameters, such as the local H2O2 occur in the absence of a nucleus, and (2) do concentration. these rhythms persist when transcription is stalled in nucleated systems? The first question was answered by assaying Circadian Oscillations in the S-Sulfinylation PRDXs hyperoxidation in RBCs under constant of the PRDXs environmental conditions (O’Neill and Reddy The story of the identification of the circadian 2011). Indeed, the results from these experi- hyperoxidation cycles in PRDXs began in 2006 ments showed that the PRDX redox oscillations when a global proteomics study identified that persisted for three full cycles under constant up to 20% of the mouse liver proteins show conditions and were unaffected by incubation circadian oscillations (Reddy et al. 2006). This at a lower free-running temperature. Even more result was interesting from a global perspective strikingly, these oscillations could be synchro- but what was even more fascinating was that nized by temperature cycles, whereby two many of the oscillating proteins did not display groups of RBCs entrained in the opposite tem- rhythms in the corresponding transcripts. The perature cycles displayed antiphasic PRDX os- study utilized two-dimensional difference gel cillations when subsequently released into free electrophoresis (2D-DiGE), which allows the running conditions. Because RBCs lack mito- segregation of proteins carrying posttransla- chondria, the auto-oxidation of hemoglobin, tional modifications (PTMs) from their un- which occurs naturally in these cells, was pro- modified counterparts. On examination of the posed as the primary source of H2O2. Consis- gels, the authors identified two spots corre- tently, the hemoglobin tetramer-to-dimer ratio sponding to different forms of PRDX6. The was also seen to oscillate in RBCs incubated in

Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 9 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

N.B. Milev et al.

vitro. However, one of the most important find- further investigations are required to unravel the ings was that the oscillations of PRDX hyper- oscillation-generating mechanism under nor- oxidation and hemoglobin oligomerization mal conditions. were paralleled by rhythms in the ratio of the The second question, concerning the possi- reduced redox cofactors NADH/NADPH as ble role of transcription in the hyperoxidation well as ATP. These results indicated that the re- rhythms of PRDXs in nucleated systems was dox/energy metabolism in RBCs shows self- answered by incubating O. tauri under condi- sustained circadian oscillations without any tions of transcriptional arrest (O’Neill et al. guidance from the transcription. 2011). It was unequivocally shown that the Although these initial studies strongly sug- PRDX oscillations persist under such condi- gested the role of hemoglobin in the generation tions, which indicated that, in this model sys- of the rhythmic oxidation of the PRDXs, this tem, normally expressing a functional TTFL, the was unequivocally shown a few years later by nontranscriptional rhythms in PRDX hyperox- work from the Rhee laboratory (Cho et al. idation are independent of the transcriptional 2014). Also, contrary to the expectations that clock. As already mentioned, the phase of the SRX plays an important role in the generation transcriptional clock in O. tauri does not reset of the oscillations, srx-null mice displayed es- after alleviation of the transcriptional arrest. sential normal PRDX rhythms, and the inacti- This observation indicates that the PRDX hy- vated form was shown to be subjected to degra- peroxidation cycles, or a putative metabolic os- dation by the 20S proteasome. This might cillator that they report on, not only act as the appear counterintuitive considering that RBCs primary timekeeping mechanism during peri- have a limited amount of PRDXs owing to the ods of attenuated transcription but that can also lack of transcription and translation. However, relay timing information back to the transcrip- less than 1% of total PRDX2 (the major PRDX tional clock (O’Neill et al. 2011). isoform in RBCs) undergoes daily oscillation, The icing on the cake of these series of dis- resulting in 30% loss of PRDX2 during the coveries was the demonstration that the PRDX life span (30–40 days) of mouse RBCs (Cho hyperoxidation cycles occur in organisms from et al. 2014). Therefore, it is possible that the all three domains of life, including archaea, eu- proteasome-mediated oscillation mechanism bacteria, plants, worms, flies, mice, and man has evolved for energy conservation because (Edgar et al. 2012; Olmedo et al. 2012). Impor- the SRX catalyzed reaction requires ATP. tantly, these rhythms were found to be indepen- The exact mechanism that generates the dent of the TTFL because clock mutant organ- 24-h period of the oscillations in RBCs is still isms display essentially normal rhythms (Edgar an open question. For instance, are the protea- et al. 2012). This not only cemented the role of somal degradation and hemoglobin auto-oxi- the PRDX oscillations in the clockwork but also dation required in a phase-dependent manner? challenged the ideas about the evolutionary or- Also, because the reduction of the disulfide and igins of circadian clocks. hyperoxidized forms of PRDXs are dependent If the genetic clock has been invented several on NADPH and ATP, respectively, is it possible in the evolutionary history, then it follows that the rhythmic accumulation of the inacti- that the primary timekeeping mechanism vated enzyme reflects the availability of the two might be encoded by the PRDX oscillations or metabolites and therefore of primary metabo- the process they report on. In this scenario, the lism? A recent study in RBCs of superoxide dis- transcriptional clock might have evolved to mutase-1 (Sod1)-deficient mice indicated that, meet the specific requirements of a given organ- under conditions of elevated ROS, the cycle in ism or system and rather fulfills species-specific PRDXs hyperoxidation is diminished (Homma functions in the adaptation response to the dis- et al. 2015). This clearly indicates that ROS ho- tinct environmental niche, whereas PRDX/ meostasis is important for the generation of metabolic oscillations might represent a prima- circadian cycles in PRDX hyperoxidation, but ry timekeeping mechanism. This is supported

10 Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

Cellular Timekeeping

by studies showing tissue-specific effects of principal issues associated with peroxide signal- clock function. For instance, disruption of the ing that has complicated the identification of molecular clockwork in the liver results in fast- novel regulatory targets: specificity, reactivity, ing-induced hypoglycemia, whereas the same and accessibility. The specificity in H2O2 signal- intervention in the pancreas impairs glucose ing cannot be achieved via molecular recogni- tolerance (Marcheva et al. 2010; Peek et al. tion because the oxidant has a very simple 2013). In this respect, analogies such as “time- chemical structure. Also, the unstable nature keeping” versus “time-telling” mechanisms of H2O2 prevents its reversible interaction would be appropriate (Rey and Reddy 2015). with protein targets. It follows that the selectiv- In simple terms, in the absence of a metabol- ity must be based on another physical property, ic/redox oscillator, the cells/organisms would such as hierarchy in the reaction rates (Winter- not be able to keep track of time. On the other bourn and Hampton 2008). This logic unveils hand, in the absence of a transcriptional oscil- yet another problem—although the reaction lator, the cells/organisms would not be able to between H2O2 and cellular thiols is highly ther- make use of the timing information. In a “basic” modynamically favorable, the extremely effec- system such as RBCs, the metabolic oscillator tive cellular peroxidases display unprecedented is both the “time-keeping” and “time-telling” kinetic superiority over other proteins. Kinetic mechanism with the only task of organizing the measurements indicate that most cellular thiols antioxidant defense around the clock. In more show moderate reaction rates with H2O2 on the 21 21 “complex” systems that coordinate a variety of order of 1–10 M sec , which are several or- processes, another layer of complexity needs to ders of magnitude lower than the ones displayed be added, and this is enabled by the diverse cyclic by peroxidases such as PRDXs and glutathi- 7 9 21 21 transcriptional programs fulfilled by the TTFL. one peroxidase (GPx) (10 –10 M sec ) There are many questions that remain to be (Randall et al. 2013). Moreover, the cellular answered concerning both the mechanistic and peroxidases are among the most abundant functional aspects of the oscillations in the S- proteins in the cell, with PRDXs making up to sulfinylation of PRDXs. Although it seems un- 1%–2% of the total mass in most cells (Chae likely that the oscillations in PRDX hyperoxida- et al. 1999), which indicates that the reaction tion serve as the timekeeping mechanism per se, between H2O2 and other cellular thiols is a it is highly possible that they are an important highly unlikely . Even though PRDXs act cog in the clockwork, potentially acting as a as the primary target of H2O2 in the cell, nu- coupling agent between the metabolic and tran- merous examples in the literature have shown scriptional oscillators. How can PRDXs fulfill that other proteins undergo reversible oxidation this role? The last has delivered a pleth- in response to the oxidant. How does this occur ora of examples supporting the important in the cell? It appears that there is no clear-cut role of this family of peroxidases in integrating answer, but the direct oxidation of target pro- ROS into cellular signaling. The subsequent sec- teins by H2O2 seems the least possible explana- tion will outline the different signaling mecha- tion (Randall et al. 2013). However, there are nisms and experimental evidence that argue in several different, but not mutually exclusive, favor of the functional role of PRDXs in the theories that offer plausible and, more impor- clockwork. tantly, experimentally validated solutions to this apparent paradox. Unsurprisingly, all of these mechanistic models revolve around the PRDX PRDXs as Mediators of Peroxide Signaling system (Fig. 3). Although H2O2 is now regarded as a bona fide physiological second messenger (Rhee 2006), Signaling via Hyperoxidation the exact details of the signal-transduction mechanism by which the oxidant exerts its ef- The structural peculiarity of the carboxy-termi- fects are still largely unknown. There are three nal tail of 2-Cys PRDXs, which enables the S-

Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 11 Downloaded from ..Mlve al. et Milev N.B. 12 http://cshperspectives.cshlp.org/

PRDX-TRX loops PRX-TRX loops coupled uncoupled

SH SH PRDX H O PRDX H2O – SH 2 –S SH S S– S– PRDX TRX PRDX TRX SH HS SH HS Peroxide Peroxide PRDX loop PRDX loop TRX loop scavenging TRX loop accumulation SH SH PRDX PRDX S S – HSO2 SO2H TRX S S onOctober6,2021-PublishedbyColdSpringHarborLaboratoryPress PRDX PRDX S TRX HS S S H2O2 H2O2 ieti ril as article this Cite

Redox relay Floodgate TRX redirection SH S HS S Protein HS TRX – PRDX Protein Protein S HS SH S HS S odSrn abPrpc Biol Perspect Harb Spring Cold PRDX HS SH S –S PRDX Protein S SH S S– Protein Protein PRDX S TRX S SH S HS

Figure 3. Peroxiredoxins (PRDXs) are mediators of reactive oxygen species (ROS) signaling. Schematic representation of the three main models for PRDXs signaling. The redox relay model: In this model, the oxidized PRDXs transfer oxidative equivalents to protein targets. This transfer reaction proceeds via the formation of a mixed disulfide intermediate between PRDXs and the client protein. The floodgate model: This model postulates that the inactivation of PRDXs via over- 2018;10:a027698 oxidation allows the local accumulation of H2O2, which can then react with other cellular targets. The thioredoxin (TRX) redirection model: This model proposes that the inactivation of PRDXs allows the accumulation of active TRX, which can then reduce other cellular targets. Note, that the redox relay model could occur under normal operational conditions when the PRDX and TRX loops are coupled. The floodgate and TRX redirection models are dependent on the uncoupling of the PRDX-TRX loop via the overoxidation of the peroxidases. Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

Cellular Timekeeping

sulfinylation of the active site cysteine, together is also required for the import process. In this with the slow rates of regeneration of the active context, it is interesting to note that analysis of enzymatic form by SRX, led some authors to gene regulatory networks associated with mam- suggest that the controlled inactivation of the malian circadian rhythms revealed that HSP90- peroxidases allows H2O2 to act as a second mes- FKBP complexes are central to the control of senger. This theory was coined the floodgate circadian gene expression by diverse environ- hypothesis, and was later supported by the find- mental signals (Yan et al. 2008). The imported ing that the hyperoxidation of mitochondrial SRX binds tightly to sulfinylated PRDX3 and PRDX3 is essential for the circadian rhythms reduces it. As the amount of sulfinylated in adrenal steroidogenesis (Wood et al. 2003; PRDX3 then declines, SRX becomes vulnerable Kil et al. 2012). Adrenocorticotropic hormone to degradation by the protease Lon, resulting in (ACTH)-induced corticosterone (CS) synthesis its down-regulation to basal levels and the con- occurs with a concomitant increase in H2O2, sequent onset of the next round of sufinylated driven by cytochrome P450 in the mitochon- PRDX3 accumulation and H2O2 release. dria. The increase in the levels of the oxidant Although this is the only existing example in results in the inactivation of PRDX3 by S-sulfi- support of the floodgate hypothesis, hyperoxi- nylation and leads to the accumulation of H2O2 dation has been found to enable other modes of in mitochondria and its overflow into the cyto- signaling. Analogous to the floodgate hypothe- sol, where it triggers the activation of p38 mito- sis, some investigators have proposed that when gen-activated protein kinase and consequently H2O2 levels exceed the antioxidant capacity of the down-regulation of both CS synthesis and the cell, S-sulfinylation acts as a molecular H2O2 production. Also, ACTH stimulates the switch that activates the chaperone-like proper- expression of SRX, which translocates to the ties of PRDXs at the expense of their peroxidase mitochondrial matrix and mediates the regen- function (Jang et al. 2004; Barranco-Medina eration of PRDX3, thus generating a regulatory et al. 2009). The important physiological role feedback loop. of the chaperone functions of 2-Cys PRDXs A daily oscillation of hyperoxidized PRDX3 was recently shown in yeast. Tsa1, the major 2- was also detected in brown adipose and heart Cys PRDX in Saccharomyces cerevisiae,was tissues, where a high level of fat metabolism is shown to be essential for the recruitment of accompanied by H2O2 production (Kil et al. the heat-shock chaperones Hsp70/Hsp104 to 2015). In addition, although only a small pro- protein aggregates, thus preventing the accu- portion of SRX was detected in mitochondria, mulation of aberrantly folded protein species the amount of mitochondrial SRX also under- in this organism (Hanze´n et al. 2016). Intrigu- goes daily oscillation in adrenal gland, as well as ingly, the recruitment of Hsp70/Hsp104 to pro- in brown adipose and heart tissues, with this tein aggregates resulting from aging, but not oscillation being antiphasic relative to that of heat stress, was shown to depend on the hyper- hyperoxidized PRDX3. These observations sug- oxidation of Tsa1, whereas the disaggregation of gest that the mitochondrial abundance of SRX the misfolded proteins only occurred after the is likely the key determinant of PRDX3 hyper- reduction of the S-sulfinylated form by SRX. oxidation and, consequently, also that of the Moreover, the overexpression of Tsa1 led to an timing and extent of mitochondrial H2O2 re- increase in life span in an Hsp70-dependent lease. The H2O2 released from mitochondria manner. Overall, these findings show the com- promotes the formation of a disulfide-linked plex nature of the PRDX system and underline complex between SRX and heat-shock protein the importance of this class of enzyme for the 90 (HSP90), and the resulting SRX–S–S– maintenance of cellular integrity. HSP90 complex is imported into mitochondria The third mechanism by which hyperoxida- by the translocase of the outer membrane tion might lead to signaling was recently shown (TOM) complex (Kil et al. 2015). A cochaper- in yeast. The inactivation of PRDXs allows the one of HSP90, FK506-binding protein (FKBP), accumulation of reduced (active) TRX, which

Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 13 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

N.B. Milev et al.

could, in turn, target other cellular proteins that whole PRDX pool into the disulfide-linked have undergone reversible oxidation. This form (Low et al. 2007). Also, the reduction by mechanism was found to be essential for the TRX was seen to be considerably slow, occurring inactivation of the yeast AP-1-like transcription gradually over a period of 20 min. This mecha- factor, Pap1, and the activation of the enzyme nism probably evolved to protect PRDX2 from methionine sulfoxide reductase (Mxr1) (Bozo- hyperoxidation because it has been shown re- net et al. 2005; Vivancos et al. 2005; Day et al. cently that, in erythrocytes, this form of the 2012). Also, expression of a hyperoxidation-re- enzyme is targeted for proteasomal degradation sistant form of the yeast 2-Cys PRDX, Tpx1, led rather than recycled by SRX (Cho et al. 2014). to a significant decrease in cell survival after Finally, it should be noted that the regula- exposure to H2O2 (Day et al. 2012). In support tion of PRDXs could be achieved via other ox- of the role of S-sulfinylation as a switch to allow idative modifications, such as S-nitrosylation repair of reversibly oxidized proteins, the dele- (Fang et al. 2007; Engelman et al. 2013; Chung tion of PRDX2 in Caenorhabditis elegans con- et al. 2017) and S-glutathionylation (Peskin fers increased resistance to oxidative stress (Ola´- et al. 2016). It also seems that, in addition to hova´ et al. 2008). However, the nematodes also being regulators of kinases, the activity of the displayed shortened life span, which underlines PRDXs is modulated by phosphorylation. For the important role of the PRDXs for longevity. example, controlled inactivation of PRDX1 by Interestingly, in vitro studies have revealed CDK1 allows the pericentrosomal buildup of that PRDX2, which is located in the cytosol, is H2O2 with potential consequences for protein more sensitive to hyperoxidation than mito- phosphatases such as Cdc14B (Lim et al. 2015). chondrial-located PRDX3 owing to the 10- fold slower formation of a disulfide bond with Beyond Hyperoxidation—Signaling via Redox its resolving cysteine (Haynes et al. 2013; Poyn- Relay Mechanism ton et al. 2016). In addition to the intrinsic biochemical properties of PRDXs, the hyperox- An alternative to the floodgate hypothesis is that idation is, in fact, dependent on a functional the fast reaction kinetics of PRDXs with H2O2 reducing system, which allows the enzymes to allow them to act as sensors and transducers spend more time in the sulfenic acid form. Con- of peroxide signaling via a thiol/disulfide ex- sistently, although the ER-localized PRDX4 dis- change mechanism, also termed redox relay. plays similar sensitivity to hyperoxidation to the In this model, PRDXs are first oxidized by cytosolic forms in vitro, an insignificant frac- H2O2, then form mixed disulfide intermediates tion of the enzyme becomes S-sulfinylated in with a client protein, and finally transfer the vivo. This is attributed to the low abundance oxidative equivalents, thereby oxidizing their of disulfide recycling systems in this organelle binding partner. This could either occur (Cao et al. 2014). The combination of different straight from the sulfenic acid form or from sensitivities toward hyperoxidation and the the disulfide form of 2-Cys PRDXs (Sobotta modulation of the activity of the disulfide re- et al. 2014). This model solves the problems ductase systems allows the PRDXs to be dynam- related to H2O2 signaling outlined at the begin- ically regulated to meet the specific demands of ning of this section. Namely, specificity is the different cellular compartments (Cox et al. provided by the direct protein–protein interac- 2009; Tavender et al. 2010) as well as different tions, the relatively nonreactive H2O2 molecule cell types (Low et al. 2007) and tissues (Ola´hova´ is activated by its conversion to the much more et al. 2008; Kil et al. 2012). In erythrocytes, for reactive sulfenic acid, and, finally, the competi- instance, where PRDX2 is one of the most abun- tion is eliminated because now PRDXs them- dant proteins, it was shown that the peroxidase selves act as second messengers. The pioneering acts mostly to detoxify low-grade H2O2 levels; work supporting this model was performed in 9 treatment with 5 mM H2O2 of 5 10 erythro- the yeast S. cerevisiae, where the thiol peroxidase cytes dm23 for 1 min is sufficient to convert the Orp1 was shown to form mixed disulfides with

14 Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

Cellular Timekeeping

the AP-1-like transcription factor Yap1, which the PDI and PRDX family was also observed ultimately leads to the activation of Yap1 via in the green alga Chlamydomonas reinhardtii, formation of an intramolecular disulfide (De- and, more interestingly, the complex between launay et al. 2002). Another example from yeast the two proteins was detected only during the is the H2O2-induced activation of the p38/JNK night phase (Filonova et al. 2013). homolog, Sty1, by the Schizosaccharomyces The PRDXs have often been described as pombe 2-Cys PRDX, Tpx1 (Veal et al. 2004). “the major sinks for H2O2.” Considering the Since these initial discoveries, participation in data presented above, this analogy is perhaps such signaling events has been described for all not the most accurate one. Our notion of these four mammalian typical 2-Cys PRDXs. For in- conserved thiol-dependent antioxidant pro- stance, PRDX1 was shown to transfer oxidative teins has evolved over the decade, and equivalents to the kinase ASK1, which results in now they can be rather classified as major cell- the formation of active homodimers and the signaling players that integrate redox signals subsequent phosphorylation of p38 (Jarvis into cellular physiology. Whether PRDXs are et al. 2012). Interestingly, this result adds an actively participating in the cellular timekeep- important detail to the mechanism of steroido- ing mechanism remains to be proven but, con- genesis in the mouse adrenal cortex described sidering the important role of the peroxidases in by Kil et al. (2012). It seems likely that the in- cell signaling, this idea seems highly plausible. activation of PRDX3 in the mitochondria al- lows the oxidation of PRDX1 in the cytosol, which then acts to activate ASK1 and in turn REDOX AROUND THE CLOCK p38, which then feeds back to induce the SRX- mediated recycling of PRDX3 (Jarvis et al. 2012; Overt Rhythms in Cellular Redox Couples Kil et al. 2012). This also strongly suggests that Oscillations in the Cellular Nicotinamide the movement of H O within the cell is medi- 2 2 Adenine Dinucleotide Redox Couples ated by the cellular peroxidases rather than by simple diffusion. More recently, a redox relay Nicotinamide adenine dinucleotide (NADþ) between PRDX2 and the oxidative stress-related and its phosphorylated form NAD phosphate transcription factor STAT3 was described (So- (NADPþ) are the two major hydride ion cofac- botta et al. 2014). In this work, it was shown that tors, which enable approximately 2000 cellular PRDX2 transfers oxidative equivalents to reactions (Opitz and Heiland 2015). NADþ STAT3, whereby STAT3 forms transcriptionally could either be synthesized de novo from tryp- inactive oligomers. In a similar fashion, PRDX2 tophan or salvaged using preformed com- also determines the oxidation status of the pounds such as niacin. NADPþ is synthesized human protein deglycase DJ-1 (Fernandez- directly via the phosphorylation of NADþ,cat- Caggiano et al. 2016). Of note, PRDX2 nuclear alyzed by NADþ kinase, and can be converted levels oscillate in HaCaT keratinocytes and back to NADþ via the action of NADPþ phos- overexpression of the protein in the nucleus phatase. Considering that the NADPþ/NADPH leads to damping of the oscillations in BMAL1 couple is maintained in the reduced state and (Avitabile et al. 2014; Ranieri et al. 2015). This serves mainly to shuffle hydride ions, this redox clearly indicates that the redox environment in couple could be regarded as the major reducing the nucleus affects the transcriptional oscillator power of the cell. On the other hand, the me- and suggests that PRDX2 might be involved in tabolism of NADþ is much more complex. This this process. Finally, the ER-resident PRDX4 is because, in addition to redox reactions, was shown to be involved in the oxidative fold- NADþ also participates as an enzyme cofactor ing of proteins by transferring oxidative equiv- in mono- and poly(ADP-ribosylation) as well as alents to several members of the protein disul- NADþ-dependent protein deacetylation. These fide isomerase (PDI) family (Tavender et al. so-called NADþ-consuming reactions convert 2010). The interaction between members of NADþ into nicotinamide mononucleotide,

Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 15 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

N.B. Milev et al.

which can be recycled back to NADþ via the stressed conditions, the GSH/GSSG ratio is salvage pathway. kept largely in the reduced state by the flavopro- In the fungus N. crassa, the NADþ/NADH tein, glutathione reductase (GR), which con- ratio was shown to resonate with the rhythms sumes reducing power in the form of NADPH. in conidiation (Brody and Harris 1973). Simi- Owing to its important role in antioxidant de- lar oscillations were observed for NADPþ/ fense, GSH has attracted much attention, and NADPH in plant seedlings kept under constant consistently its diurnal profile has been assayed conditions (Wagner and Frosch 1974). Diurnal in most major organ systems. Oscillations in variation in both redox couples was also ob- GSH have been observed in rodent liver (Isaacs served in rodent liver (Robinson et al. 1981; and Binkley 1977a,b; Robinson et al. 1981; Fa- Kaminsky et al. 1984), although these were as- rooqui and Ahmed 1984; Be´langer et al. 1991; sayed under normal feeding regimens and thus Neuschwander-Tetri and Rozin 1996; Li et al. might be partially driven by food intake. It is 1997; Baydas et al. 2002; Ponce et al. 2012), interesting to note that rats adapted to a high- brain (Farooqui and Ahmed 1984; Baydas protein diet displayed nearly unchanged oscil- et al. 2002; Wang et al. 2012; Kinoshita et al. lations in the cytoplasmic redox couples com- 2014), lung (Farooqui and Ahmed 1984; Pe- pared with controls, suggesting that food intake kovic-Vaughan et al. 2014), spleen (Farooqui might not be the sole driver of these rhythms. and Ahmed 1984), kidney (Farooqui and Ah- Indeed, oscillations in the levels of NAD(P)H med 1984), gut (Li et al. 1997), pancreas (Neu- were also observed in RBCs under constant con- schwander-Tetri and Rozin 1996), stomach ditions (O’Neill and Reddy 2011) and more (Farooqui and Ahmed 1984), blood plasma recently in human osteosarcoma U2OS cells (Blanco et al. 2007), and platelets (Radha et al. (Rey et al. 2016). The redox state of the central 1985). Despite the numerous reports on overt brain pacemaker, the suprachiasmatic nuclei rhythms in the levels of GSH, the exact molec- (SCN) as represented by the ratio between ular mechanism has not yet been elucidated, NAD(P)H/FADH, was seen to oscillate (Wang and it seems that the regulation might occur et al. 2012). More recently, with the use of non- in a tissue-specific manner. For example, fasting invasive methods, it was shown that the NAD/ diminishes GSH oscillations in the liver but not NADH ratio oscillates in vivo in stem cells and in the pancreas (Neuschwander-Tetriand Rozin HEK293T cells (Stringari et al. 2015; Huang 1996) and in the mesencephalon region of the et al. 2016). brain (Kinoshita et al. 2014). Interestingly, in Although the rhythms in pyridine dinucleo- the case of the latter, the oscillations in GSH tide redox couples might in part be driven by were attributed to a posttranslational mecha- the transcriptional clock in some of these model nism via a microRNA (miRNA), which controls systems, the fact that they have also been de- the levels of a neuron-specific membrane trans- scribed to persist in the absence of a nucleus porter for cysteine, the rate-limiting substrate strongly suggests that they might be driven by for GSH synthesis. The situation becomes nontranscriptional mechanisms. even more complex considering that platelets, which are naturally anucleate, show robust GSH oscillations even when kept in vitro. In stark Oscillations in Glutathione contrast, Pekovic-Vaughan et al. (2014) found GSH is a tripeptide consisting of L-g-glutamyl- that, in the lung, the transcriptional clock exerts L-cysteinyl-glycine and constitutes the most control over GSH metabolism via the transcrip- abundant low-molecular-weight thiol in eu- tion factor NRF2. One of the transcriptional karyotic cells. As such, GSH is the main thiol targets of NRF2 is glutamate cysteine ligase buffering system and has often been described (GCL), which catalyses the rate-limiting step as a rheostat of the cellular redox state. In cells, in GSH synthesis, and thus rhythmic transcrip- GSH exists in a redox couple with its oxidized tion of NRF2 was shown to result in rhythmic form, glutathione disulfide (GSSG). Under un- GCL transcription. Although nuclear import of

16 Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

Cellular Timekeeping

NRF2 has been shown to be a redox-regulated clock and, more importantly, that oscillations in process, in the lung, NRF2 displays rhythmic NAMPT generate rhythms in the cellular NADþ nuclear accumulation under unstressed condi- levels not only solidified the connection be- tions. A similar mechanism seems to be operat- tween the redox system and circadian rhythms, ing in fruit flies where the circadian rhythmicity but also indicated that cellular NADþ metabo- in GCL expression and activity, as well as GSH lism is much more dynamic than previously levels, were shown to be directly controlled by anticipated (Opitz and Heiland 2015). the transcriptional clock (Beaver et al. 2012). The first proof of a mechanistic link between It is important to note that the rhythms in NADþ metabolism and the circadian clock was GSH translate to daily oscillations in the sus- the demonstration that the NAD-dependent ceptibility to oxidant insult. Considering that deacetylase SIRT1 binds to the CLOCK:BMAL1 time-dependent protection against oxidative complex in a circadian fashion and regulates stress has been observed in both liver and the circadian transcriptional programs via the de- brain (Li et al. 1997; Kinoshita et al. 2014), and acetylation of the core clock transcription fac- that supplementation with N-acetyl-L-cysteine tors BMAL1 and PER2 and chromatin-associ- ameliorates the effects of oxidative stress (Kon- ated proteins (Asher et al. 2008; Nakahata et al. dratov et al. 2009; Berman et al. 2011), the in- 2008). Shortly after, two independent studies teraction between GSH metabolism and the established the important role of the cellular clock might be more important than previously clock in controlling the NAMPT-NAD-SIRT thought. axis by showing that the transcriptional oscilla- tor drives the cycling of the NAMPT transcript and protein, which, in turn, generates rhythms þ Redox Process as Outputs of the in the cellular NAD levels (Nakahata et al. Transcriptional Clock 2009; Ramsey et al. 2009). Of note, although fasted animals display NADþ oscillations in Clock-Mediated Control over the NAMPT- the liver, feeding induces 12-h harmonics in NAD-SIRT Axis this organ (Ramsey et al. 2009; Peek et al. As already described, in addition to its impor- 2013). The relationship between NADþ and tant function in redox metabolism, NADþ the circadian clock appears to be reciprocal as serves as a cofactor for several enzymes, such deletion of CD38, one of the major NADþ-con- as the protein deacetylases sirtuins (SIRTs) suming enzymes in the cell, results in altered and poly(ADP-ribose) polymerase 1 (PAPR1). behavioral and metabolic rhythms (Sahar The participation of NADþ in these reactions et al. 2011). Another NADþ-dependent en- results in the hydrolysis of the cofactor to the zyme, which has been shown to feed back to precursor nicotinamide (NAM). NADþ can affect clock function is PARP1. PARP1 was then be regenerated from NAM via a two-step shown to bind to CLOCK and transfer poly enzymatic pathway termed the NADþ salvage (ADP-ribose) moieties, thereby affecting the pathway. The first and rate-limiting step is cat- DNA-binding affinity of the transferrin (TF) alyzed by nicotinamide phosphoribosyltrans- and ultimately the phase of the clock (Asher ferase (NAMPT), which converts NAM into et al. 2010). Of note, the circadian activity of NAM mononucleotide (MNM) using ATP and PARP-1 was shown to be independent of a func- 5-phosphoribosyl 1-pyrophosphate (PRPP) as tional clock and rather regulated by feeding. þ cofactors. Although NAD can also be synthe- This is in line with the fact that the Km value sized de novo from the amino acid tryptophan of PARP1 for NADþ is relatively low (50– or directly from diet-derived nicotinic acid 59 mM), and thus NADþ levels are rarely limit- (NA), the salvage pathway appears to be the ing for PARP1 activity (Mendoza-Alvarez and major source of NADþ in cells (Canto´ et al. Alvarez-Gonzalez 1993; Canto´ et al. 2015). An- 2015). Therefore, the discovery that NAMPT other way by which PARP-1 can indirectly affect is under the direct control of the transcriptional the clock is by limiting the availability of NADþ

Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 17 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

N.B. Milev et al.

for the SIRTenzymes (Canto´ et al. 2015). This is are parallel to oscillations in the levels and ac- supported by the fact that deletion or pharma- tivity of a high proportion of antioxidant en- cological inhibition of PARP-1 results in signif- zymes, including SOD1, catalase, GPx, and the icantly elevated NADþ levels and increased PRDXs. Finally, studies in clock mutant organ- SIRT1 activity, indicating that the enzyme is a isms have shown that the circadian regulation of major NADþ consumer and in direct competi- the antioxidant defense systems is under the tion with the SIRTs for NADþ availability (Bai control of the transcriptional clock. For exam- et al. 2011). In fact, a recent study used compu- ple, ROS levels and protein carbonylation dis- tational simulations to show that increased play daily oscillations in wide-type flies but PARP-1 activity in response to DNA damage not in per01 mutants (Krishnan et al. 2008). might be able to induce SIRT1-dependent clock Bmal12/2 mice display reduced life span and phase shifts because of competition for limited symptoms of premature aging as well as in- NADþ supplies (Luna et al. 2015). creased ROS levels in several tissues (Kondratov In mammals, there are seven SIRTenzymes et al. 2006). Consistently, supplementation with (SIRT1-7) that display diverse cellular distribu- the antioxidant N-acetyl-L-cysteine ameliorates tion with SIRT1, SIRT6, and SIRT7 primarily in some of the overt effects in these mutant ani- the nucleus, SIRT3-5 in the mitochondria, and mals (Kondratov et al. 2009). The tau mutant in SIRT2 found in the cytoplasm (Canto´ et al. the Syrian hamster also displays increased levels 2015). Perhaps, more importantly, the SIRTs of oxidative damage, at least in the Harderian þ display widely different Km values to NAD , gland (Coto-Montes et al. 2001). Mutations in ranging from 100 mM (SIRT1/2/4/6) to the plant core clock regulator CCA1 affects the 1000 mM (SIRT3/5) (Canto´ et al. 2015). In- transcriptional regulation of many ROS-scav- deed, this sensitive range enables the family of enging enzymes, ROS homeostasis, and the tol- deacetylases to perform specific functions in erance to oxidative stress (Lai et al. 2012). different tissues and cellular compartments, de- It should be noted that in most cases the pending on the local daily oscillations in NADþ deletion of the clock gene leads to a general levels. Consistently, the intracellular rhythms in increase in intracellular ROS and oxidative NADþ were shown to modulate the activity of damage. This indicates that the effects might SIRT3, thereby inducing rhythmic cellular res- be attributable to noncircadian roles of these piration through modulation of mitochondrial transcription factors. For example, Bmal12/2 protein acetylation (Peek et al. 2013). Similarly, mice display significantly lower NADþ and SIRT6 was shown to aid the rhythmic recruit- NADH levels compared with wild-type (WT) ment of CLOCK:BMAL1 complexes and anoth- animals (Peek et al. 2013), and rhythmic expres- er transcription factor SREBP-1 to circadian sion of Bmal1 was shown to be dispensable for gene promoters, thereby regulating fatty acid molecular rhythmicity (Liu et al. 2008). Also, in metabolism in the liver (Masri et al. 2014). addition to the loss of rhythmicity in response to oxidant treatment, both the arrhythmic per01 and short-period perS mutant flies show higher Rhythms in Antioxidant Defense Systems average levels of lipid peroxidation and car- The important roles of the clock for maintain- bonylated proteins with aging (Krishnan et al. ing redox homeostasis have come from three 2009; Coto-Montes and Hardeland 2010). major lines of evidence. First, a high number of studies have shown that various oxidative Antioxidant Functions of Melatonin stress markers, such as DNA damage, lipid per- oxidation, and protein carbonylation, display Secreted primarily from the pineal gland, mel- diurnal oscillations (Hardeland et al. 2003; atonin is involved in the temporal coordination Wilking et al. 2013). This leads to the second of a plethora of processes, most notably in the line of evidence that comes from studies show- control of seasonal physiology (Challet 2015). ing that the rhythms in oxidative stress markers The antioxidant properties of melatonin were

18 Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

Cellular Timekeeping

first shown by studies reporting on the in vitro with the results in Neurospora, treatment with ROS-scavenging properties of the hormone exogenous oxidants has been shown to reset the (Iana˘s¸et al. 1991). Since then, the role of mel- phase of the transcriptional oscillator in mam- atonin in the antioxidant defense system has malian cells (Tamaru et al. 2013). In zebrafish, been unequivocally shown (Hardeland et al. light induces the production of H2O2, which, in 2003; Hardeland and Pandi-Perumal 2005; turn, acts as a second messenger to transduce Hardeland 2013). However, the circulating lev- the photic signals to the transcriptional oscilla- els of melatonin are too low to account for the tor (Hirayama et al. 2007). In cyanobacteria, the observed antioxidant effects of the hormone light-sensitive protein Ldp-1 functions as a cel- (Hardeland et al. 2003). In addition, one would lular redox sensor, which is capable of transduc- expect that if melatonin was to engage in direct ing redox signals to the core oscillator (Ivleva ROS scavenging, then the secretion of the hor- et al. 2005). In addition, treatment with oxi- mone would peak during the light phase in di- dized quinones resets the cyanobacterial clock, urnal animals to coincide with the maximal lo- and this effect was also showed using electro- comotor activity. In contrast, melatonin levels chemically controlled extracellular peak at the start of the dark phase in both noc- transfer, which enables the external modulation turnal and diurnal animals, which suggests that of the redox state (Lu et al. 2014). All these the hormone offers antioxidant support in an- studies indicate that the redox state can influ- other form. Several different modes of action ence circadian physiology. What are the molec- have been described for the antioxidant role ular mechanisms that enable these interactions? of melatonin, including up-regulation of the Below, we summarize the most well-known ex- levels of many ROS-scavenging enzymes (e.g., amples of the putative mechanistic links be- GPx, SOD1), down-regulation of pro-oxidant tween cellular redox and the transcriptional os- enzymes (e.g., lipoxygenase, NO synthase), cillator. and the activation of antioxidant transcription- al programs via stimulation of nuclear factor Redox Control over the Positive Elements of (NF)-kB, AP-1, and NRF2 (Luchetti et al. the Transcriptional Clock 2010). Indeed, it is possible that the antioxidant One of the first indications that the transcrip- properties of melatonin are a pleiotropic effect tional oscillator might be able to sense changes and unrelated to its role in circadian and sea- in the cellular redox state came from biochem- sonal physiology. However, the fact that mela- ical studies showing that the DNA-binding af- tonin supplementation has been shown to exert finity of CLOCK:BMAL1 and NPAS2:BMAL1 positive effects beyond its role in the clock sug- transcriptional complexes can be modulated gests that further investigation is required (Har- by varying the NAD(P)þ/NAD(P)H ratio in deland and Pandi-Perumal 2005). vitro (Rutter et al. 2001). The reduced cofactors were shown to increase binding, whereas their oxidized counterparts exerted an inhibitory ef- Redox Control over the Transcriptional Clock fect. Importantly, even subtle changes in the As already alluded to earlier, the connection NAD(P)þ/NAD(P)H ratio are capable of in- between the circadian clock and the redox poise ducing changes in the binding and dissociation, is reciprocal and several lines of evidence have indicating that these clock transcription factors shown that the redox state of the cell can influ- have the properties of bona fide redox sensors. ence the function of the molecular clockwork. These findings were later extended by showing For example, Sod-1-deficient Neurospora show that the 1-61 amino-terminal amino acids of more robust rhythms in the molecular clock- the bHLH domain of NPAS2 were sufficient to work (Yoshida et al. 2011). In addition, ROS sense NAD(P)H and that the NPAS2/BMAL1 can modulate the circadian phase and period transcriptional complexes can also sense in the fungi (Gyo¨ngyo¨si et al. 2013). Consistent changes in the cellular pH (Yoshii et al. 2013,

Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 19 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

N.B. Milev et al.

2015). The concentrations of NADH and mutation of a second cysteine residue in CRY1 NADPH used in these studies were considerably (Cys414) was shown to perturb both behavioral higher than the one measured in vivo and and molecular rhythms (Okano et al. 2009). whether NPAS2/CLOCK:BMAL1 are capable This residue is essential for Zn2þ coordination, of sensing redox changes in cells remains an and, although it does not participate in disul- open question (Zhang et al. 2002). However, fide bond formation, it is plausible that the ox- the direct sensing of changes in the NADþ/ idation of this cysteine in the apo-CRY1 might NADH ratio has been shown for carboxyl-ter- interfere with its interaction with PER2. To minus-binding protein (CtBP), indicating that complicate matters even further, it appears such a mechanism operates in vivo (Zhang et al. that Cys412 in CRY1 might be required for the 2002; Fjeld et al. 2003). Indeed, this would also interaction of the protein with FBXL3 (Xing allow the oscillations in NADþ levels to feed et al. 2013; Schmalen et al. 2014). The latter back to the transcriptional clock by directly promotes the degradation of the CRY proteins modulating the transcriptional activity of via its association with the ubiquitin E3 ligase NPAS2/CLOCK:BMAL1. Indirect evidence for complex SCF (Skp1-Cul1-F-box protein), and the importance of the nuclear redox state for the thus this interaction might have a functional functioning of the molecular clockwork came consequence. Of note, mutation of another cys- from a recent study showing that the nuclear teine residue in FBXL3, termed after hours overexpression of PRDX2 in HaCaT keratino- (Afh), results in long free-running rhythms cyte induces perturbations to the transcription- with the loss of circadian transcriptional oscil- al oscillations of Bmal1 (Ranieri et al. 2015). lations (Busino et al. 2007; Godinho et al. 2007), As described earlier, another potential way suggesting that the cellular redox state can im- in which oxidative stress might impinge on the pinge on the molecular clockwork via reversible molecular clockwork is through the activation oxidation of cysteine residues. Following the of PARP-1 by oxidation-induced DNA damage same line of thought, several proteins involved (Luna et al. 2015). Under such conditions, in the control of CRY1 stability have been shown PARP-1 is expected to consume relatively higher to harbor reactive cysteines. For instance, the amounts of NADþ and thus limit the availabil- metabolic sensor AMP-activated protein kinase ity of the cofactor for the SIRT deacetylase, (AMPK), which promotes the phosphoryla- which would, in turn, affect clock function. tion-dependent degradation of CRY1, has re- cently been shown to be regulated via a redox- active disulfide switch (Lamia et al. 2009). Sim- Redox Control over the Negative Arm of ilarly, CRY1 is stabilized by the deubiquitinase the TTFL herpes-virus-associated ubiquitin-specific pro- Recently, two independent studies reported on tease (HAUSP), which harbors an active site the crystal structure of the mouse PER2:CRY1 cysteine residue controlled via reversible oxida- complex (Nangle et al. 2014; Schmalen et al. tion (Lee et al. 2013). 2014). The structure of the complex between The involvement of redox regulation in the the circadian repressor proteins shows that a clockwork has also been implied in the control zinc ion (Zn2þ) binds in the interface between of the accessory loops of the transcriptional os- the two proteins and leads to the stabilization of cillator. It was shown that REV-ERBb harbors a the dimer. Interestingly, Schmalen et al. re- regulatory disulfide bond in the heme-binding ported that Zn2þ coordination is governed by pocket, which dictates the binding of the cofac- the redox state of a regulatory disulfide bond in tor and thus the activation of the TF (Gupta and CRY1. However, these results were questioned Ragsdale 2011). Although these cysteine resi- by Nangle et al., who showed that the mutation dues are not conserved in the closely related of one of the cysteines from the regulatory cou- REV-ERBa, a similar mechanism might still ple did not display a significantly different phe- be in place to control this TF because redox notype than the WT counterpart. Of note, the regulation appears to be a common theme in

20 Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

Cellular Timekeeping

the biology of the nuclear receptor superfamily complexes, whereas the depletion of CO was (Carter and Ragsdale 2014). shown to lead to global up-regulation of CLOCK:BMAL1-dependent gene expression. Interestingly, heme degradation via the HOs The Heme and Carbon Monoxide (CO) Loop consumes NADPH, thereby indicating a poten- Another reciprocal link between the circadian tial interaction between the cellular redox poise and redox systems centers on the metabolism of and the heme-CO-clock loop. the iron-containing porphyrin, heme. Heme forms the nonprotein part of hemoglobin but The Oxygen—HIF1a Axis also acts as a cofactor for many cellular enzymes and transcription factors. The rate-limiting en- One of the first indications that hypoxia and the zyme in heme biosynthesis, amino-levulinate circadian system are linked derived from the synthase 1 (Alas1), is under transcriptional reg- discovery that both clock and the hypoxia-in- ulation by NASP2:BMAL1 and the levels of the ducible factors (HIFs) are members of the basic porphyrin oscillate in a circadian manner. helix–loop–helix PER-ARNT-SIM (bHLH- Heme was shown to feed back to the core clock PAS) transcription factor superfamily. In addi- by affecting both the positive and negative limbs tion, it was shown that BMAL1 dimerizes with of the transcriptional oscillator. For instance, both HIF1a and HIF2a in vitro (Hogenesch binding of heme to the NPAS2:BMAL1 inhibits et al. 1998). The significance of this interaction the interaction of the complex with DNA in in physiological settings remained unexplored response to CO (Dioum et al. 2002; Ishida up until recently when three independent stud- et al. 2008). In contrast, an association of the ies showed the reciprocal interaction between cofactor with the nuclear hormone receptors the transcriptional oscillator and oxygen sens- REV-ERBa and REV-ERBb induces the recruit- ing via HIF1a, each one adding novel insights ment of the corepressor NcoR, which causes into the underlying mechanism of the coupling repression of target genes, including Bmal1 (Ra- and the physiological relevance of this interac- ghuram et al. 2007; Yin et al. 2007). Thus, heme tion in mammals (Adamovich et al. 2016; Peek exerts direct control over its levels by inhibiting et al. 2016; Wu et al. 2016). both NPAS2:BMAL1 activity and Bmal1 tran- Peek et al. (2016) investigated the interac- scription, thereby generating a regulatory loop. tion between hypoxia and circadian clocks in Finally, heme has been shown to bind to a novel muscle tissue owing to the significant implica- regulatory motif in PER2 (different from the tions of oxygen metabolism in this system. PAS domain), which leads to increased stability Exercise causes oxygen depletion in skeletal of the protein and altered association with muscle, which forces a metabolic switch to CRY1. Of note, the motif binds exclusively to primarily anaerobic glycolysis. This effect is ferric heme, and thus this interaction is gov- mediated by the HIF1a/b heterodimer, which erned by the redox state of the iron core of the senses the decrease in oxygen levels and in- cofactor (Yang et al. 2008). duces the transcription of genes important in A recent study added yet more detail to the glycolytic metabolism. The activity of HIF1a is mechanism of this feedback loop by showing directly proportional to the stability of the that heme is rhythmically degraded to CO by protein, which is under the control of the ox- the action of the enzyme heme oxygenase 1 and ygen-dependent prolyl-hydroxylases and the 2 (HO-1 and HO-2) (Klemz et al. 2016). More von Hippel–Lindau (VHL) E3 ubiquitin li- importantly, rhythmic CO generation was gase. Under normoxic conditions, HIF1a is shown to be essential for normal circadian continuously marked for degradation, whereas rhythmicity on both the molecular and the be- the decrease in oxygen levels during hypoxia havioral level. The primary action of CO was leads to the accumulation of the protein (Wei- shown to occur through inhibition of the tran- demann and Johnson 2008). The authors scriptional activity of CLOCK/NPAS2:BMAL1 showed that deletion of Bmal1 in myotubes

Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 21 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

N.B. Milev et al.

leads to the abrogation of the HIF1a-mediated Adamovich et al. (2016) focused primarily responses to hypoxia, whereas the absence of on the physiological aspects of the interaction the clock repressors CRY1/2 led to the stabili- between oxygen metabolism and the circadian zation of the TF. Conversely, pharmacological clock. The authors showed that oxygen con- or genetic stabilization of HIF1a altered the sumption is rhythmic in mice and, perhaps period of PER2 oscillations and led to an in- more importantly, that this translated to rhyth- crease in the transcript levels of several core mic oxygenation of the blood and the kidney clock genes. In addition, it was shown that (Adamovich et al. 2016). One particularly HIF1a binds to the promoter regions of Per2 interesting finding is that alternating 12-h and Cry1 and that the TF acts synergistically low- and high-oxygen cycles can entrain the with BMAL1 to induce expression of Per2. Fi- molecular clockwork in cultured cells in an nally, the authors showed the clear physiolog- HIF1a-dependent manner. To examine the rel- ical implication of this interaction by showing evance of these findings in the context of the that the circadian clock gates the hypoxic re- whole animal, the authors showed that the ap- sponse to exercise in mice. Considering the plication of a low oxygen pulse accelerates the major role of HIF1a in the metabolic switch adaptation of mice subjected to a jet lag proto- during hypoxia, these findings strongly suggest col. This observation unequivocally shows that that the TF might act as a coupling factor be- hypoxia is an important resetting cue and, to- tween the transcriptional clock and the redox gether with the findings of Peek et al. (2016), system of the cell. might help to explain the positive effect of By performing a series of ChIP-Seq experi- exercising in accelerating the adaption to jet ments, Wu et al. (2016) elaborated on the mech- lag (Mrosovsky and Salmon 1987). anism of the interaction between BMAL1 and HIF1a. Interestingly, they found BMAL1/ The Cellular Redox Poise and Circadian HIF1a co-occupancy at 20%–30% at BMAL1 Timekeeping: A Systems Affair binding sites and that this interaction is syner- gistically leading to an increase in the transcrip- Although some of the mechanistic details of the tional activation of these loci. Interestingly, the PRDX hyperoxidation cycles have now been de- circadian pathway was found to be significantly scribed, two important issues remain unre- overrepresented among the genes co-occupied solved. First, what is the significance of these by both TFs but not in genes occupied only by oscillations? Is the hyperoxidized form of the BMAL1. This implies that HIF1a is a dominant enzyme involved in signaling events or does regulator of the circadian clock during hypoxic the controlled inactivation of PRDXs allow sig- conditions. Reciprocally, the authors showed naling via rhythmic accumulation of H2O2? that the transcription of HIF1a displays circa- Second, what is the driving force behind these dian oscillations in mouse liver, supporting the oxidation cycles? Are they intrinsic to the bio- role of the clock in priming the response of chemical characteristics of PRDXs or are they a HIF1a to hypoxia. To solidify the physiological by-product of oscillations in a global cellular relevance of the findings, the authors investigat- parameter such as the cellular redox poise? ed the possible involvement of the circadian In RBCs, the PRDXs hyperoxidation cycles clock in the severity of the damage induced by are paralleled by robust oscillations in NADH, myocardial infarction (MI), one of the most NADPH, and ATP levels (O’Neill and Reddy well-known hypoxia-related pathologies. Con- 2011). Because the reduction of the disulfide sistent with previous reports showing higher and hyperoxidized forms of PRDXs are depen- mortality rates associated with morning heart dent on NADPH and ATP,respectively, it seems attacks (Portaluppi and Lemmer 2007), the au- plausible that the rhythmic accumulation of the thors showed that Per12/2/Per22/2 mutant inactivated enzyme reflects the availability of mice show more severe MI damage compared the two metabolites. In support of this hypoth- with WT controls. esis, recent work from our group identified the

22 Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

Cellular Timekeeping

highly conserved pentose phosphate pathway lator might be represented by the rhythmic ac- (PPP), the primary source of NADPH in the tivity of the highly conserved PPP pathway or cell, as a novel regulator of the cellular redox the interplay between the PPP and other prima- and transcriptional oscillations (Rey et al. ry metabolic pathways such as glycolysis. 2016). Using pharmacological and genetic ap- In fact, the existence of a redox oscillator has proaches, we found that inhibition of the PPP been previously described in the literature. In prolongs the period circadian rhythms in hu- 2005, Tu and colleagues showed that a continu- man cells, mouse tissues, and fruit flies. ous-culture system in S. cerevisiae shows robust Inhibition of the PPP leads to significant 4–5-h cycles of oxygen consumption under changes in the period length of the oscillations nutrient-limited growth conditions (Tu et al. in NADPH and the hyperoxidized form of 2005; Tu and McKnight 2007). These cycles PRDX in nucleated cells. This suggests that the were found to consist of three distinct phases: fluctuations in the reducing power of the cell are oxidative, reductive/building (R/B), and re- linked to the oscillations in PRDX. Because the ductive/charging (R/C). The authors found catalytic cycle of the TRX system relies on that over half of the S. cerevisiae genes are ex- NADPH as the ultimate electron donor, it is pressed periodically as a function of these met- possible that the activity of the system is reflect- abolic oscillations, and that specific clusters of ing the levels of the cofactor. This also implies genes are expressed in each phase. For example, that the hyperoxidation of PRDXs might de- cell division occurs only during the R/B phase pend on the activity of the TRX systems, where- (minimal oxygen consumption), and never dur- by the maximal activity of TRX coincides with ing the oxidative respiratory phase. The R/C the peak in PRDXs hyperoxidation. Although phase is defined by up-regulation of genes in- this might appear counterintuitive, active recy- volved in nonrespiratory modes of metabolism, cling of PRDXs via the TRX system has been which ensures the buildup of reducing power shown to be essential for the hyperoxidation before the respiratory burst in the subsequent of the peroxidases because the disulfide-linked oxidative phase. In an excellent review on the form is protected from further oxidation (Cao topic, the authors evolved the hypothesis that et al. 2014). such metabolic oscillations might be at the In addition to the alterations in the cellular heart of most biological cycles (Tu and redox oscillations, blocking of the PPP leads McKnight 2006). Based on the data summa- to profound period changes in the circadian rized above, it is tempting to speculate that gene-expression programs in cells. By perform- such a model could be applied in higher eukary- ing a series of ChIP experiments, we found otes. Indeed, the rhythms in NADPH shown in that this global effect is partially mediated by RBCs and U2OS cells under constant condi- increased activity of the circadian transcription tions indicate rhythmicity in the flux through factors BMAL1 and CLOCK, the redox-sensi- the oxidative branch of the PPP pathway. If we tive transcription factor NRF2, and the chro- accept such a model, then this would require matin remodeling histone acetyltransferase, that the “redox homoeostasis,” which in- P300. In addition, we found that inhibition fers that the system is static, be substituted by a of the PPP in primary mouse tissues recapit- term that factors in the diurnal aspects of the ulated the results seen with cells, thus clearly system, such as redox “homeodynamism.” showing the physiological significance of this Importantly, the question of what drives perturbation. these oxidation-reduction phases remains Finally, inhibition of the PPP in fruit flies open. Diurnal flux through the oxidative and leads to behavioral misalignment, which indi- reductive branches of metabolic pathways cates that the perturbation of this highly con- might in part explain this if a regulatory mech- served pathway affects rhythms at the molecu- anism that dictates the diurnal rate of activity of lar, tissue, and behavioral level. Collectively, these pathways is in place. These possibilities these data suggest that the putative redox oscil- remain to be tested.

Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 23 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

N.B. Milev et al.

Intersections with Tissue-Specific Clocks: protective mechanisms against genotoxicity The Redox Factor (Stringari et al. 2015).

Countless evidence has supported the impor- tant role of peripheral clocks for circadian CONCLUDING REMARKS rhythmicity. One of the emerging concepts in Retrospective the past couple of decades is that the circadian system shows pronounced tissue-specific roles. In the early days of the research into the molec- Indeed, the importance of the cellular redox ular mechanism of the circadian clock, the view state in the generation of tissue-specific oscilla- of most chronobiologists appeared more ba- tions has been shown for most major organ lanced. The following sentence taken from systems (Milev and Reddy 2015). Goto et al. (1985) embodies the view of the In the liver, the rhythmic activation of larger part of the field at the time: þ SIRT6 by NAD was shown to play an impor- Finally, we have proposed a model—not for tant role in lipid homeostasis (Masri et al. “the” clock—but for one oscillator in what is 2014). Deletion of Bmal1 in the liver leads to most probably a cellular “clockshop” .... increased anaerobic metabolism and lactic acid Unfortunately, as summarized in Lakin- production, whereas deletion of the TF in the Thomas (2006), not long after such state- muscle leads to a decrease in anaerobic metab- ments, this metamorphosed into: olism, suggesting that BMAL1 controls distinct transcriptional programs in these tissues (Peek All circadian oscillators are based on transcrip- tional feedback loops involving the canonical et al. 2016). Oscillations in the hyperoxidation clock genes. of the mitochondrial PRDX3 were shown to be essential for rhythmic steroidogenesis in the Unfortunately, even nowadays, after the discov- adrenal gland (Kil et al. 2012). Several redox ery of numerous examples of TTFL anomalies parameters such as the NADPH/FAD ratio and transcriptionally independent oscillations, and protein glutathionylation display rhythms the prevalent view remains largely unchanged in the SCN and these oscillations were shown (Fig. 4). It is of paramount importance that the to be essential for neuronal firing (Wang et al. chronobiologists of the adopt a more ob- 2012). Numerous studies have shown the prev- jective view and assume that the circadian sys- alence of overt oscillations of the redox state tem is most probably composed of several cou- in the hematologic system. RBCs display oscil- pled oscillators that work in unison in the cell to lations in PRDX hyperoxidation, in the pyri- generate a robust and highly adaptive cellular dine nucleotide levels, and hemoglobin auto- timekeeping mechanism. oxidation (O’Neill and Reddy 2011). Platelets display rhythms in GSH, which might be in- Perspective volved in the rhythmic activation of these cells (Radha et al. 1985; Scheer et al. 2011). The We believe that the future efforts in the field levels of GSH and sulfhydryl to disulfide ratio should focus on unraveling the coupling mech- in blood plasma have also been shown to os- anisms that exist between the different cogs in cillate (Blanco et al. 2007). These findings im- the clockwork, most notably between the pu- ply that the redox capacity of the blood reso- tative metabolic oscillator and the TTFL. Con- nates with the day–night cycles and implies sidering the ample evidence for the involve- functional consequences for a variety of pro- ment of the cellular redox system in circadian cesses, such as cell–cell interactions and timekeeping, we expect that this relationship wound healing (Sen and Roy 2008; Eble and will take an even more central position in the de Rezende 2014). Recently, oscillations in the future. However, we do not exclude the possi- NAD/NADH ratio were also shown in epider- bility that other of primary metabolism mal stem cells, implying circadian gating of the play an important role in the cellular time-

24 Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 Downloaded from ieti ril as article this Cite http://cshperspectives.cshlp.org/ odSrn abPrpc Biol Perspect Harb Spring Cold NADP+/ Antioxidant NADPH; α + PRDX-TRX O2-HlF1 Nutrients Heme-CO defense; NAD -SIRTs p300; NRF2 GSH; ROS onOctober6,2021-PublishedbyColdSpringHarborLaboratoryPress Coupling loops 2018;10:a027698

Metaboilc Transcriptional oscillator oscillator

Figure 4. The unifying model of the transcriptional and metabolic oscillators—coupled oscillators. In the cell, the metabolic and transcriptional oscillators are coupled to generate a robust timekeeper, which is shielded from both metabolic and transcriptional perturbations. The coupling of the two oscillators occurs via the various known and

potentially many unknown loops, which are individually dispensable but confer a high degree of flexibility to the clock. Timekeeping Cellular The high degree of interplay between the two oscillators in nucleated systems makes it extremely challenging to tease apart the two systems. However, manipulations of the coupling loops have provided ample evidence that the two oscillators are tightly connected. NADPH, Nicotinamide adenine dinucleotide phosphate; PRDX, peroxiredoxin; HIF,hypoxia-inducible factor; CO, carbon monoxide; GSH, glutathione; ROS, reactive oxygen species; SIRT, sirtuin. 25 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

N.B. Milev et al.

keeping mechanism and hope that no stones Bozonet SM, Findlay VJ, Day AM, Cameron J, Veal EA, will be left unturned. Morgan BA. 2005. Oxidation of a eukaryotic 2-Cys per- oxiredoxin is a molecular switch controlling the tran- scriptional response to increasing levels of hydrogen per- oxide. J Biol Chem 280: 23319–23327. REFERENCES Brody S, Harris S. 1973. Circadian rhythms in Neurospora: Spatial differences in pyridine nucleotide levels. Adamovich Y, Ladeuix B, Golik M, Koeners MP, Asher G. 180: 498–500. 2016. Rhythmic oxygen levels reset circadian clocks Busino L, Bassermann F, Maiolica A, Lee C, Nolan PM, through HIF1a. Cell Metab 25: 93–101. Godinho SIH, Draetta GF, Pagano M. 2007. SCFFbxl3 Asher G, Gatfield D, Stratmann M, Reinke H, Dibner C, controls the oscillation of the circadian clock by directing Kreppel F, Mostoslavsky R, Alt FW, Schibler U. 2008. the degradation of cryptochrome proteins. Science 316: SIRT1 regulates circadian clock gene expression through 900–904. PER2 deacetylation. Cell 134: 317–328. Canto´ C, Menzies KJ, Auwerx J. 2015. NADþ metabolism Asher G, Reinke H, Altmeyer M, Gutierrez-Arcelus M, Hot- and the control of energy homeostasis: A balancing act tiger MO, Schibler U. 2010. Poly(ADP-ribose) polymer- between mitochondria and the nucleus. Cell Metab 22: ase 1 participates in the phase entrainment of circadian 31–53. clocks to feeding. Cell 142: 943–953. Cao Z, Subramaniam S, Bulleid NJ. 2014. Lackof an efficient Avitabile D, Nicolussi A, Capriotti AL, Cucina A, Samperi R, endoplasmic reticulum-localized recycling system pro- Bizzarri M, Torrisi MR. 2014. Peroxiredoxin 2 nuclear tects peroxiredoxin IV from hyperoxidation. J Biol levels are regulated by circadian clock synchronization Chem 289: 5490–5498. in human keratinocytes. Int J Biochem Cell Biol 53:24– Carter EL, Ragsdale SW.2014. Modulation of nuclear recep- 34. tor function by cellular redox poise. J Inorg Biochem 133: Bai P, Canto´ C, Oudart H, Brunya´nszki A, Cen Y, Thomas 92–103. C, Yamamoto H, Huber A, Kiss B, Houtkooper RH, et al. 2011. PARP-1 inhibition increases mitochondrial Chae HZ, Chung SJ, Rhee SG. 1994. Thioredoxin-depen- metabolism through SIRT1 activation. Cell Metab 13: dent peroxide reductase from yeast. J Biol Chem 269: 461–468. 27670–27678. Barnett A. 1966. A of mating type reversals Chae HZ, Kim HJ, Kang SW, Rhee SG. 1999. Characteriza- in Paramecium multimicronucleatum, syngen 2, and its tion of three isoforms of mammalian peroxiredoxin that genetic control. J Cell Physiol 67: 239–270. reduce peroxides in the presence of thioredoxin. Diabetes Res Clin Pract 45: 101–112. Barranco-Medina S, La´zaro JJ, Dietz KJ. 2009. The oligo- meric conformation of peroxiredoxins links redox state Challet E. 2015. Keeping circadian time with hormones. to function. FEBS Lett 583: 1809–1816. Diabetes, Obesity Metab 17: 76–83. Baydas G, Gursu MF, Yilmaz S, Canpolat S, Yasar A, Cikim Cho CS, Yoon HJ, Kim JY, Woo HA, Rhee SG. 2014. Circa- G, Canatan H. 2002. Daily rhythm of glutathione perox- dian rhythm of hyperoxidized peroxiredoxin II is deter- idase activity, lipid peroxidation and glutathione levels in mined by hemoglobin autoxidation and the 20S protea- tissues of pinealectomized rats. Neurosci Lett 323: 195– some in red blood cells. Proc Natl Acad Sci 111: 12043– 198. 12048. Beaver LM, Klichko VI, Chow ES, Kotwica-Rolinska J, Wil- Chung MC, Alem F, Hamer SG, Narayanan A, Shatalin K, liamson M, Orr WC, Radyuk SN, Giebultowicz JM. 2012. Bailey C, Nudler E, Hakami RM. 2017. S-nitrosylation of Circadian regulation of glutathione levels and biosynthe- peroxiredoxin 1 contributes to viability of lung epithelial sis in Drosophila melanogaster. PLoS ONE 7: e50454. cells during Bacillus anthracis infection. Biochim Biophys Be´langer PM, Desgagne´ M, Bruguerolle B. 1991. Temporal Acta 1861: 3019–3029. variations in microsomal lipid peroxidation and in glu- Cornelius G, Rensing L. 1976. Daily rhythmic changes in tathione concentration of rat liver. Drug Metab Dispos 19: Mg2þ-dependent ATPase activity in human red blood cell 241–244. membranes in vitro. Biochem Biophys Res Commun 71: Bell-Pedersen D, Cassone VM, Earnest DJ, Golden SS, Har- 1269–1272. din PE, Thomas TL, Zoran MJ. 2005. Circadian rhythms Coto-Montes A, Hardeland R. 2010. Diurnal rhythm of from multiple oscillators: Lessons from diverse organ- protein carbonyl as an indicator of oxidative damage in isms. Nat Rev Genet 6: 544–556. Drosophila melanogaster: Influence of clock gene alleles Berman AE, Chan WY, Brennan AM, Reyes RC, Adler BL, and deficiencies in the formation of free-radical scaven- Suh SW,Kauppinen TM, Edling Y,Swanson RA. 2011. N- gers. Biol Rhythm Res 30: 383–391. acetylcysteine prevents loss of dopaminergic neurons in Coto-Montes A, Toma´s-Zapico C, Rodrı´guez-Colunga MJ, the EAAC12/2 mouse. Ann Neurol 69: 509–520. Tolivia-Cadrecha D, Martı´nez-Fraga J, Hardeland R, Toli- Biteau B, Labarre J, Toledano MB. 2003. ATP-dependent via D. 2001. Effects of the circadian mutation “tau” on the reduction of cysteine–sulphinic acid by S. cerevisiae sul- Harderian glands of Syrian hamsters. J Cell Biochem 83: phiredoxin. Nature 425: 980–984. 426–434. Blanco RA, Ziegler TR, Carlson BA, Cheng PY, Park Y, Cot- Cox AG, Pearson AG, Pullar JM, Jo¨nsson TJ, Lowther WT, sonis GA, Accardi CJ, Jones DP. 2007. Diurnal variation Winterbourn CC, Hampton MB. 2009. Mitochondrial in glutathione and cysteine redox states in human plas- peroxiredoxin 3 is more resilient to hyperoxidation ma. Am J Clin Nutr 86: 1016–1023. than cytoplasmic peroxiredoxins. Biochem J 421: 51–58.

26 Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

Cellular Timekeeping

Day AM, Brown JD, Taylor SR, Rand JD, Morgan BA, Veal Goto K, Laval-Martin DL, Edmunds LN. 1985. Biochemical EA. 2012. Inactivation of a peroxiredoxin by hydrogen modeling of an autonomously oscillatory circadian clock peroxide is critical for thioredoxin-mediated repair of in Euglena. Science 228: 1284–1288. oxidized proteins and cell survival. Mol Cell 45: 398–408. Gupta N, Ragsdale SW. 2011. Thiol-disulfide redox depen- Delaunay A, Pflieger D, Barrault MB, Vinh J, Toledano MB. dence of heme binding and heme ligand switching in b 2002. A thiol peroxidase is an H2O2 receptor and redox- nuclear hormone receptor Rev-erb . J Biol Chem 286: transducer in gene activation. Cell 111: 471–481. 4392–4403. Dioum EM, Rutter J, Tuckerman JR, Gonzalez G, Gilles- Gyo¨ngyo¨si N, Nagy D, Makara K, Ella K, Ka´ldi K. 2013. Gonzalez MA, McKnight SL. 2002. NPAS2: A gas-respon- Reactive oxygen species can modulate circadian phase sive transcription factor. Science 298: 2385–2387. and period in Neurospora crassa. Free Radic Biol Med 58: 134–143. Driessche TV. 1966. Circadian rhythms in Acetabularia: ´ Photosynthetic capacity and chloroplast shape. Exp Cell Hanzen S, Vielfort K, Yang J, Roger F, Andersson V, Zamar- bide-Fore´s S, Andersson R, Malm L, Palais G, Biteau B, et Res 42: 18–30. al. 2016. Lifespan control by redox-dependent recruit- Eble JA, de Rezende FF. 2014. Redox-relevant aspects of the ment of chaperones to misfolded proteins. Cell 166:1– extracellular matrix and its cellular contacts via integrins. 13. Antioxid Redox Signal 20: 1977–1993. Hardeland R. 2013. Chronobiology of melatonin beyond the Edgar RS, Green EW,Zhao Y,van Ooijen G, Olmedo M, Qin feedback to the suprachiasmatic nucleus—Consequences X, Xu Y, Pan M, Valekunja UK, Feeney KA, et al. 2012. to melatonin dysfunction. Int J Mol Sci 14: 5817–5841. Peroxiredoxins are conserved markers of circadian Hardeland R, Pandi-Perumal SR. 2005. Melatonin, a potent rhythms. Nature 485: 459–464. agent in antioxidative defense: Actions as a natural food Ehret CF, Trucco E. 1967. Molecular models for the circadi- constituent, gastrointestinal factor, drug and prodrug. an clock I. The chronon concept. J Theoret Biol 15: 240– Nutr Metab 2: 22. 262. Hardeland R, Coto-Montes A, Poeggeler B. 2003. Circadian Engelman R, Weisman-Shomer P, Ziv T, Xu J, Arne´r ESJ, rhythms, oxidative stress, and antioxidative defense Benhar M. 2013. Multilevel regulation of 2-Cys peroxi- mechanisms. Chronobiol Int 20: 921–962. redoxin reaction cycle by S-nitrosylation. J Biol Chem Hardin PE, Hall JC, Rosbash M. 1990. Feedback of the Dro- 288: 11312–11324. sophila period gene product on circadian cycling of its Fang J, Nakamura T, Cho DH, Gu Z, Lipton SA. 2007. S- messenger RNA levels. Nature 343: 536–540. nitrosylation of peroxiredoxin 2 promotes oxidative Haynes AC, Qian J, Reisz JA, Furdui CM, Lowther WT.2013. stress-induced neuronal cell death in Parkinson’s disease. Molecular basis for the resistance of human mitochon- Proc Natl Acad Sci 104: 18742–18747. drial 2-Cys peroxiredoxin 3 to hyperoxidation. J Biol Farooqui MY, Ahmed AE. 1984. Circadian periodicity of Chem 288: 29714–29723. tissue glutathione and its relationship with lipid peroxi- Hirayama J, Cho S, Sassone-Corsi P.2007. Circadian control dation in rats. Life Sci 34: 2413–2418. by the reduction/oxidation pathway: Catalase represses light-dependent clock gene expression in the zebrafish. Feeney KA, Hansen LL, Putker M, Olivares-Yan˜ez C, Day J, Proc Natl Acad Sci 104: 15747–15752. Eades LJ, Larrondo LF, Hoyle NP, O’Neill JS, van Ooijen G. 2016. Daily magnesium fluxes regulate cellular time- Hogenesch JB, Gu YZ, Jain S, Bradfield CA. 1998. The basic- keeping and energy balance. Nature 532: 375–379. helix–loop–helix-PAS orphan MOP3 forms transcrip- tionally active complexes with circadian and hypoxia fac- Fernandez-Caggiano M, Schro¨der E, Cho HJ, Burgoyne J, tors. Proc Natl Acad Sci 95: 5474–5479. Barallobre-Barreiro J, Mayr M, Eaton P. 2016. Oxidant- Homma T, Okano S, Lee J, Ito J, Otsuki N, Kurahashi T, induced interprotein disulfide formation in cardiac DJ-1 Kang ES, Nakajima O, Fujii J. 2015. SOD1 deficiency occurs via an interaction with peroxiredoxin 2. J Biol induces the systemic hyperoxidation of peroxiredoxin Chem 291: 10399–10410. in the mouse. Biochem Biophys Res Commun 463: Filonova A, Haemsch P, Gebauer C, Weisheit W, Wagner V. 1040–1046. 2013. Protein disulfide isomerase 2 of Chlamydomonas Huang G, Zhang Y,Shan Y,Cheliah Y,WangH, TakahashiJS. reinhardtii is involved in circadian rhythm regulation. 2016. Circadian oscillations of NADH redox state using Mol Plant 6: 1503–1517. heterologous metabolic sensor in mammalian cells. J Biol Fisher AB. 2011. Peroxiredoxin 6: A bifunctional enzyme Chem 291: 23906–23914. with glutathione peroxidase and phospholipase A2 activ- Iana˘s¸ O, Olinescu R, Ba˘descu I. 1991. Melatonin involve- ities. Antioxid Redox Signal 15: 831–844. ment in oxidative processes. Endocrinologie 29:147–153. Fjeld CC, Birdsong WT, Goodman RH. 2003. Differential Isaacs JT, Binkley F. 1977a. Cyclic AMP-dependent control binding of NADþ and NADH allows the transcriptional of the rat hepatic glutathione disulfide–sulfhydryl ratio. corepressor carboxyl-terminal binding protein to serve as Biochim Biophys Acta 498: 29–38. a metabolic sensor. Proc Natl Acad Sci 100: 9202–9207. Isaacs J, Binkley F. 1977b. Glutathione dependent control of Godinho SIH, Maywood ES, Shaw L, Tucci V, Barnard AR, protein disulfide-sulfhydryl content by subcellular frac- Busino L, Pagano M, Kendall R, Quwaillid MM, Romero tions of hepatic tissue. Biochim Biophys Acta 497: 192– MR, et al. 2007. The after-hours mutant reveals a role for 204. Fbxl3 in determining mammalian circadian period. Sci- Ishida M, Ueha T,Sagami I. 2008. Effects of mutations in the ence 316: 897–900. heme domain on the transcriptional activity and DNA-

Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 27 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

N.B. Milev et al.

binding activity of NPAS2. Biochem Biophys Res Commun Krishnan N, Kretzschmar D, Rakshit K, Chow E, Giebulto- 368: 292–297. wicz JM. 2009. The circadian clock gene period extends Ivleva NB, Bramlett MR, Lindahl PA, Golden SS. 2005. healthspan in aging Drosophila melanogaster. Aging 1: LdpA: A component of the circadian clock senses redox 937–948. state of the cell. EMBO J 24: 1202–1210. Lai AG, Doherty CJ, Mueller-Roeber B, Kay SA, Schippers Jang HH, Lee KO, Chi YH, Jung BG, Park SK, Park JH, Lee JHM, Dijkwel PP.2012. Circadian clock-associated 1 reg- JR, Lee SS, Moon JC, Yun JW,et al. 2004. Twoenzymes in ulates ROS homeostasis and oxidative stress responses. one; two yeast peroxiredoxins display oxidative stress-de- Proc Natl Acad Sci 109: 17129–17134. pendent switching from a peroxidase to a molecular Lakin-Thomas PL. 2006. Transcriptional feedback oscilla- chaperone function. Cell 117: 625–635. tors: Maybe, maybe not.... J Biol Rhythms 21: 83–92. Jarvis RM, Hughes SM, Ledgerwood EC. 2012. Peroxire- Lamia KA, Sachdeva UM, DiTacchio L, Williams EC, Alva- doxin 1 functions as a signal peroxidase to receive, trans- rez JG, Egan DF, Vasquez DS, Juguilon H, Panda S, Shaw duce, and transmit peroxide signals in mammalian cells. RJ, et al. 2009. AMPK regulates the circadian clock by Free Radic Biol Med 53: 1522–1530. cryptochrome phosphorylation and degradation. Science Kaminsky YG, Kosenko EA, Kondrashova MN. 1984. Anal- 326: 437–440. ysis of the circadian rhythm in energy metabolism of rat Lee JG, Baek K, Soetandyo N, Ye Y. 2013. Reversible inacti- liver. Int J Biochem 16: 629–639. vation of deubiquitinases by reactive oxygen species in Kang SW, Baines IC, Rhee SG. 1998. Characterization of a vitro and in cells. Nat Commun 4: 1568. mammalian peroxiredoxin that contains one conserved Li XM, Metzger G, Filipski E, Boughattas N, Lemaigre G, cysteine. J Biol Chem 273: 6303–6311. Hecquet B, Filipski J, Levi F. 1997. Pharmacologic mod- Khalsa SB, Block GD. 1992. Phase-shifting of a neuronal ulation of reduced glutathione circadian rhythms with circadian pacemaker in Bulla gouldiana by pentylenetet- buthionine sulfoximine: Relationship with cisplatin tox- razol. Comp Biochem Physiol C 101: 557–560. icity in mice. Toxicol Appl Pharmacol 143: 281–290. Khalsa SB, Whitmore D, Bogart B, Block GD. 1996. Evi- Lim JM, Lee KS, Woo HA, Kang D, Rhee SG. 2015. Control dence for a central role of transcription in the timing of the pericentrosomal H2O2 level by peroxiredoxin I is mechanism of a circadian clock. Am J Physiol 271: critical for mitotic progression. J Cell Biol 210: 23–33. C1646–C1651. Liu AC, Tran HG, Zhang EE, Priest AA, Welsh DK, Kay SA. Kil IS, Lee SK, Ryu KW,WooHA, Hu MC, Bae SH, Rhee SG. 2008. Redundant function of REV-ERBa and b and non- 2012. Feedback control of adrenal steroidogenesis via essential role for Bmal1 cycling in transcriptional regula- H O -dependent, reversible inactivation of peroxire- 2 2 tion of intracellular circadian rhythms. PLoS Genet 4: doxin III in mitochondria. Mol Cell 46: 584–594. e1000023. Kil IS, Ryu KW, Lee SY, Kim YY, Chu SY, Kim JH, Park S, Low FM, Hampton MB, Peskin AV, Winterbourn CC. 2007. Rhee SG. 2015. Circadian oscillation of sulfiredoxin in Peroxiredoxin 2 functions as a noncatalytic scavenger of the mitochondria. Mol Cell 59: 651–613. low-level hydrogen peroxide in the erythrocyte. Blood Kinoshita C, Aoyama K, Matsumura N, Kikuchi-Utsumi K, 109: 2611–2617. Watabe M, Nakaki T. 2014. Rhythmic oscillations of the microRNA miR-96-5p play a neuroprotective role by indi- Lu Y, Nishio K, Matsuda S, Toshima Y, Ito H, Konno T, rectly regulating glutathione levels. Nat Commun 5:3823. Ishihara K, Hashimoto K, Nakanishi S. 2014. Regulation of the cyanobacterial circadian clock by electrochemically Klemz R, Reischl S, Wallach T,Witte N, Ju¨rchott K, Klemz S, controlled extracellular electron transfer. Angew Chem Int Lang V,Lorenzen S, Knauer M, Heidenreich S, et al. 2016. Ed Engl 53: 2208–2211. Reciprocal regulation of carbon monoxide metabolism and the circadian clock. Nat Struct Mol Biol 24: 15–22. Luchetti F, Canonico B, Betti M, Arcangeletti M, Pilolli F, Piroddi M, Canesi L, Papa S, Galli F. 2010. Melatonin Knoops B, Goemaere J, Van der Eecken V,Declercq JP.2011. Peroxiredoxin 5: Structure, mechanism, and function of signaling and cell protection function. FASEB J 24: the mammalian atypical 2-Cys peroxiredoxin. Antiox Re- 3603–3624. dox Signal 15: 817–829. Luna A, McFadden GB, Aladjem MI, Kohn KW. 2015. Pre- þ Kondo T, Mori T, Lebedeva NV, Aoki S, Ishiura M, Golden dicted role of NAD utilization in the control of circadi- SS. 1997. Circadian rhythms in rapidly dividing cyano- an rhythms during DNA damage response. PLoS Comput bacteria. Science 275: 224–227. Biol 11: e1004144. Kondratov RV, Kondratova AA, Gorbacheva VY, Vykhova- Mabood SF, Newman PF, Nimmo IA. 1978. Circadian nets OV, Antoch MP. 2006. Early aging and -related rhythms in the activity of acetylcholinesterase of human pathologies in mice deficient in BMAL1, the core com- erythrocytes incubated in vitro. Biochem Soc Trans 6: ponent of the circadian clock. Genes Dev 20: 1868–1873. 305–308. Kondratov RV,Vykhovanets O, Kondratova AA, Antoch MP. Mandoli DF. 1998. Elaboration of body plan and phase 2009. Antioxidant N-acetyl-L-cysteine ameliorates symp- change during development of Acetabularia: How is the toms of premature aging associated with the deficiency of complex architecture of a giant unicell built? Annu Rev the circadian protein BMAL1. Aging 1: 979–987. Plant Physiol Plant Mol Biol 49: 173–198. Krishnan N, Davis AJ, Giebultowicz JM. 2008. Circadian Manevich Y, Feinstein SI, Fisher AB. 2004. Activation of the regulation of response to oxidative stress in Drosophila antioxidant enzyme 1-CYS peroxiredoxin requires gluta- melanogaster. Biochem Biophys Res Commun 374: 299– thionylation mediated by heterodimerization with pGST. 303. Proc Natl Acad Sci 101: 3780–3785.

28 Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

Cellular Timekeeping

Marcheva B, Ramsey KM, Buhr ED, Kobayashi Y, Su H, Ko Neuschwander-Tetri BA, Rozin T. 1996. Diurnal variability CH, Ivanova G, Omura C, Mo S, Vitaterna MH, et al. of cysteine and glutathione content in the pancreas and 2010. Disruption of the clock components CLOCK and liver of the mouse. Comp Biochem Physiol B Biochem Mol BMAL1 leads to hypoinsulinaemia and diabetes. Nature Biol 114: 91–95. 466: 627–631. Nishiwaki T, Iwasaki H, Ishiura M, Kondo T. 2000. Nucleo- Masri S, Rigor P, Cervantes M, Ceglia N, Sebastian C, Xiao tide binding and autophosphorylation of the clock pro- C, Roqueta-Rivera M, Deng C, Osborne TF,Mostoslavsky tein KaiC as a circadian timing process of cyanobacteria. R, et al. 2014. Partitioning circadian transcription by Proc Natl Acad Sci 97: 495–499. SIRT6 leads to segregated control of cellular metabolism. Nitabach MN, Blau J, Holmes TC. 2002. Electrical silencing Cell 158: 659–672. of Drosophila pacemaker neurons stops the free-running Mendoza-Alvarez H, Alvarez-Gonzalez R. 1993. Poly(ADP- circadian clock. Cell 109: 485–495. ribose) polymerase is a catalytic dimer and the auto- Nitabach MN, Holmes TC, Blau J. 2005. Membranes, ions, modification reaction is intermolecular. J Biol Chem and clocks: Testing the Njus–Sulzman–Hastings model of 268: 22575–22580. the circadian oscillator. Methods Enzymol 393: 682–693. Meng QJ, Maywood ES, Bechtold DA, Lu WQ, Li J, Gibbs JE, Njus D, Sulzman FM, Hastings JW. 1974. Membrane model Dupre SM, Chesham JE, Rajamohan F, Knafels J, et al. for the circadian clock. Nature 248: 116–120. 2010. Entrainment of disrupted circadian behavior Okano S, Akashi M, Hayasaka K, Nakajima O. 2009. Unusual through inhibition of casein kinase 1 (CK1) enzymes. circadian locomotor activity and pathophysiology in mu- Proc Natl Acad Sci 107: 15240–15245. tant CRY1 transgenic mice. Neurosci Lett 451: 246–251. Mergenhagen D, Schweiger HG. 1975a. Circadian rhythm of Ola´hova´ M, Taylor SR, Khazaipoul S, Wang J, Morgan BA, oxygen evolution in cell fragments of Acetabularia med- Matsumoto K, Blackwell TK, Veal EA. 2008. A redox- iterranea. Exp Cell Res 92: 127–130. sensitive peroxiredoxin that is important for longevity Mergenhagen D, Schweiger HG. 1975b. The effect of differ- has tissue- and stress-specific roles in stress resistance. ent inhibitors of transcription and translation on the Proc Natl Acad Sci 105: 19839–19844. expression and control of circadian rhythm in individual Olmedo M, O’Neill JS, Edgar RS, Valekunja UK, Reddy AB, cells of Acetabularia. Exp Cell Res 94: 321–326. Merrow M. 2012. Circadian regulation of olfaction and Milev NB, Reddy AB. 2015. Circadian redox oscillations and an evolutionarily conserved, nontranscriptional marker metabolism. Trends Endocrinol Metab 26: 430–437. in Caenorhabditis elegans. Proc Natl Acad Sci 109: 20479– 20484. Morse D, Milos PM, Roux E, Hastings JW. 1989. Circadian O’Neill JS, Reddy AB. 2011. Circadian clocks in human red regulation of bioluminescence in Gonyaulax involves blood cells. Nature 469: 498–503. translational control. Proc Natl Acad Sci 86: 172–176. O’Neill JS, Reddy AB. 2012. The essential role of cAMP/ Morse D, Hastings JW, Roenneberg T. 1994. Different phase 2 Ca þ signaling in mammalian circadian timekeeping. responses of the two circadian oscillators in Gonyaulax. J Biochem Soc Trans 40: 44–50. Biol Rhythms 9: 263–274. O’Neill JS, van Ooijen G, Dixon LE, Troein C, Corellou F, Mrosovsky N, Salmon PA. 1987. A behavioural method for Bouget FY,Reddy AB, Millar AJ. 2011. Circadian rhythms accelerating re-entrainment of rhythms to new light– persist without transcription in a eukaryote. Nature 469: dark cycles. Nature 330: 372–373. 554–558. Nagoshi E, Saini C, Bauer C, Laroche T, Naef F, Schibler U. Opitz CA, Heiland I. 2015. Dynamics of NADþ-metabo- 2004. Circadian gene expression in individual fibroblasts: lism: Everything but constant. Biochem Soc Trans 43: Cell-autonomous and self-sustained oscillators pass time 1127–1132. to daughter cells. Cell 119: 693–705. Park J, Lee S, Lee S, Kang SW. 2014. 2-Cys peroxiredoxins: Nakahata Y, Kaluzova M, Grimaldi B, Sahar S, Hirayama J, Emerging hubs determining redox dependency of mam- þ Chen D, Guarente LP,Sassone-Corsi P.2008. The NAD - malian signaling networks. Int J Cell Biol 2014: 1–10. dependent deacetylase SIRT1 modulates CLOCK-medi- Paulose JK, Rucker EB, Cassone VM. 2012. Toward the be- ated chromatin remodeling and circadian control. Cell ginning of time: Circadian rhythms in metabolism pre- 134: 329–340. cede rhythms in clock gene expression in mouse embry- Nakahata Y,Sahar S, Astarita G, Kaluzova M, Sassone-Corsi onic stem cells. PLoS ONE 7: e49555. þ P. 2009. Circadian control of the NAD salvage pathway Pavlidis T. 1969. Populations of interacting oscillators and by CLOCK-SIRT1. Science 324: 654–657. circadian rhythms. J Theor Biol 22: 418–436. Nakajima M, Imai K, Ito H, Nishiwaki T, Murayama Y,Iwa- Peek CB, Affinati AH, Ramsey KM, Kuo HY,Yu W,Sena LA, saki H, Oyama T, Kondo T. 2005. Reconstitution of cir- Ilkayeva O, Marcheva B, Kobayashi Y, Omura C, et al. cadian oscillation of cyanobacterial KaiC phosphoryla- 2013. Circadian clock NADþ cycle drives mitochondrial tion in vitro. Science 308: 414–415. oxidative metabolism in mice. Science 342: 1243417. Nangle SN, Rosensweig C, Koike N, Tei H, Takahashi JS, Peek CB, Levine DC, Cedernaes J, Taguchi A, Kobayashi Y, Green CB, Zheng N. 2014. Molecular assembly of the Tsai SJ, Bonar NA, McNulty MR, Ramsey KM, Bass J. period-cryptochrome circadian transcriptional repressor 2016. Circadian clock interaction with HIF1a mediates complex. eLife 3: e03674. oxygenic metabolism and anaerobic glycolysis in skeletal Netto LES, Antunes F. 2016. The roles of peroxiredoxin and muscle. Cell Metab 25: 86–92. thioredoxin in hydrogen peroxide sensing and in signal Pekovic-Vaughan V, Gibbs J, Yoshitane H, Yang N, Pathir- transduction. Mol Cells 39: 65–71. anage D, Guo B, Sagami A, Taguchi K, Bechtold D,

Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 29 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

N.B. Milev et al.

Loudon A, et al. 2014. The circadian clock regulates al. 2006. Circadian orchestration of the hepatic prote- rhythmic activation of the NRF2/glutathione-mediated ome. Curr Biol 16: 1107–1115. antioxidant defense pathway to modulate pulmonary fi- Rey G, Reddy AB. 2015. Interplay between cellular redox brosis. Genes Dev 28: 548–560. oscillations and circadian clocks. Diabetes Obes Metab Perkins A, Nelson KJ, Parsonage D, Poole LB, Karplus PA. 17: 55–64. 2015. Peroxiredoxins: Guardians against oxidative stress Rey G, Valekunja UK, Feeney KA, Wulund L, Milev NB, and modulators of peroxide signaling. Trends Biochem Sci Stangherlin A, Ansel-Bollepalli L, Velagapudi V, O’Neill 40: 435–445. JS, Reddy AB. 2016. The pentose phosphate pathway reg- Peskin AV, Low FM, Paton LN, Maghzal GJ, Hampton MB, ulates the circadian clock. Cell Metab 24: 462–473.

Winterbourn CC. 2007. The high reactivity of peroxire- Rhee SG. 2006. Cell signaling. H2O2, a necessary evil for cell doxin 2 with H2O2 is not reflected in its reaction with signaling. Science 312: 1882–1883. other oxidants and thiol reagents. J Biol Chem 282: Rhee SG. 2016. Overview on peroxiredoxin. Mol Cells 39:1– 11885–11892. 5. Peskin AV, Pace PE, Behring JB, Paton LN, Soethoudt M, Rhee SG, Kang SW, Chang TS, Jeong W, Kim K. 2001. Per- Bachschmid MM, Winterbourn CC. 2016. Glutathiony- oxiredoxin, a novel family of peroxidases. IUBMB Life 52: lation of the active site cysteines of peroxiredoxin 2 and 35–41. recycling by glutaredoxin. J Biol Chem 291: 3053–3062. Rhee SG, Jeong W, Chang TS, Woo HA. 2007. Sulfiredoxin, Ponce IT, Rezza IG, Delgado SM, Navigatore LS, Bonomi the cysteine sulfinic acid reductase specific to 2-Cys per- MR, Golini RL, Gimenez MS, Anzulovich AC. 2012. Dai- oxiredoxin: Its discovery, mechanism of action, and bio- ly oscillation of glutathione redox cycle is dampened in logical significance. Kidney Int Suppl 72: S3–S8. the nutritional vitamin A deficiency. Biol Rhythm Res 43: Robinson JL, Foustock S, Chanez M, Bois-Joyeux B, Peret J. 351–372. 1981. Circadian variation of liver metabolites and amino Poole LB, Nelson KJ. 2016. Distribution and features of the acids in rats adapted to a high protein, carbohydrate-free six classes of peroxiredoxins. Mol Cells 39: 53–59. diet. J Nutr 111: 1711–1720. Portaluppi F, Lemmer B. 2007. Chronobiology and chrono- Roenneberg T, Morse D. 1993. Two circadian oscillators in therapy of ischemic heart disease. Adv Drug Deliv Rev 59: one cell. Nature 362: 362–364. 952–965. Rutter J, Reick M, Wu LC, McKnight SL. 2001. Regulation of Poynton RA, Hampton MB. 2014. Peroxiredoxins as bio- clock and NPAS2 DNA binding by the redox state of NAD markers of oxidative stress. Biochim Biophys Acta 1840: cofactors. Science 293: 510–514. 906–912. Sahar S, Nin V,Barbosa MT,Chini EN, Sassone-Corsi P.2011. Poynton RA, Peskin AV,Haynes AC, Lowther WT, Hampton Altered behavioral and metabolic circadian rhythms in MB, Winterbourn CC. 2016. Kinetic analysis of struc- mice with disrupted NADþ oscillation. Aging 3: 794–802. tural influences on the susceptibility of peroxiredoxins ScheerFAJL,MichelsonAD,FrelingerAL,EvoniukH,Kelly 2 and 3 to hyperoxidation. Biochem J 473: 411–421. EE, McCarthy M, Doamekpor LA, Barnard MR, Steven AS. Radha E, Hill TD, Rao GH, White JG. 1985. Glutathione 2011. The human endogenous circadian system causes levels in human platelets display a circadian rhythm in greatest platelet activation during the biological morning vitro. Thrombosis Res 40: 823–831. independent of behaviors. PLoS ONE 6: e24549. Raghuram S, Stayrook KR, Huang P,Rogers PM, Nosie AK, Schmalen I, Reischl S, Wallach T, Klemz R, Grudziecki A, McClure DB, Burris LL, Khorasanizadeh S, Burris TP, Prabu JR, Benda C, Kramer A, Wolf E. 2014. Interaction Rastinejad F. 2007. Identification of heme as the ligand of circadian clock proteins CRY1 and PER2 is modulated for the orphan nuclear receptors REV-ERBa and REV- by zinc binding and disulfide bond formation. Cell 157: ERBb. Nat Struct Mol Biol 14: 1207–1213. 1203–1215. Ramsey KM, Yoshino J, Brace CS, Abrassart D, Kobayashi Y, Schrader LO, Holzapfel G. 1980. Circadian rhythms in hu- Marcheva B, Hong HK, Chong JL, Buhr ED, Lee C, et al. man erythrocytes in vitro not confirmed. J Interdisciplin- 2009. Circadian clock feedback cycle through NAMPT- ary Cycle Res 11: 199–207. þ mediated NAD biosynthesis. Science 324: 651- 654. Schweiger E, Wallraff HG, Schweiger HG. 1964. Endogenous Randall LM, Ferrer-Sueta G, Denicola A. 2013. Peroxiredox- circadian rhythm in cytoplasm of Acetabularia: Influence ins as preferential targets in H2O2-induced signaling. of the nucleus. Science 146: 658–659. Methods Enzymol 527: 41–63. Sen CK, Roy S. 2008. Redox signals in wound healing. Bio- Ranieri D, Avitabile D, Shiota M, Yokomizo A, Naito S, chim Biophys Acta 1780: 1348–1361. Bizzarri M, Torrisi MR. 2015. Nuclear redox imbalance Sobotta MC, Liou W, Sto¨cker S, Talwar D, Oehler M, Rup- affects circadian oscillation in HaCaT keratinocytes. Int J pert T, Scharf AN, Dick TP. 2014. Peroxiredoxin-2 and Biochem Cell Biol 65: 113–124. STAT3 form a redox relay for H2O2 signaling. Nat Chem Reddy AB, O’Neill JS. 2011. Metaclocks. Nature 12: 612– Biol 11: 64–70. 612. Solt LA, WangY,Banerjee S, Hughes T,Kojetin DJ, Lundasen Reddy AB, Rey G. 2014. Metabolic and nontranscriptional T, Shin Y, Liu J, Cameron MD, Noel R, et al. 2012. Reg- circadian clocks: Eukaryotes. Annu Rev Biochem 83: 165– ulation of circadian behaviour and metabolism by syn- 189. thetic REV-ERB agonists. Nature 485: 62–68. Reddy AB, Karp NA, Maywood ES, Sage EA, Deery M, Stringari C, Wang H, Geyfman M, Crosignani V, Kumar V, O’Neill JS, Wong GK, Chesham J, Odell M, Lilley KS, et TakahashiJS, Andersen B, Gratton E. 2015. In vivo single-

30 Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

Cellular Timekeeping

cell detection of metabolic oscillations in stem cells. Cell Woo HA, Kang SW, Kim HK, Yang KS, Chae HZ, Rhee SG. Rep 10:1–7. 2003a. Reversible oxidation of the active site cysteine of Sweeney BM, Haxo FT.1961. Persistence of a photosynthetic peroxiredoxins to cysteine sulfinic acid. Immunoblot de- rhythm in enucleated Acetabularia. Science 134: 1361– tection with antibodies specific for the hyperoxidized 1363. cysteine-containing sequence. J Biol Chem 278: 47361– Tamaru T, Hattori M, Ninomiya Y, Kawamura G, Vare`sG, 47364. Honda K, Mishra DP,WangB, Benjamin I, Sassone-Corsi Woo HA, Chae HZ, Hwang SC, Yang KS, Kang SW, Kim K, P, et al. 2013. ROS stress resets circadian clocks to coor- Rhee SG. 2003b. Reversing the inactivation of peroxire- dinate pro-survival signals. PLoS ONE 8: e82006. doxins caused by cysteine sulfinic acid formation. Science Tavender TJ, Springate JJ, Bulleid NJ. 2010. Recycling of 300: 653–656. peroxiredoxin IV provides a novel pathway for disulfide Wood ZA, Poole LB, Karplus PA. 2003. Peroxiredoxin evo- formation in the endoplasmic reticulum. EMBO J 29: lution and the regulation of hydrogen peroxide signaling. 4185–4197. Science 300: 650–653. Terborgh J, McLeod G. 1967. The photosynthetic rhythm of Woolum JC. 1991. A re-examination of the role of the nu- Acetabularia crenulata. I. Continuous measurements of cleus in generating the circadian rhythm in Acetabularia. oxygen exchange in alternating light–dark regimes and J Biol Rhythms 6: 129–136. in constant light of different intensities. Biol Bull 133: Wu Y, Tang D, Liu N, Xiong W, Huang H, Li Y, Ma Z, Zhao 659–669. H, Chen P, Qi X, et al. 2016. Reciprocal regulation be- Tomita J, Nakajima M, Kondo T, Iwasaki H. 2005. No tran- tween the circadian clock and hypoxia signaling at the scription–translation feedback in circadian rhythm of genome level in mammals. Cell Metab 25: 73–85. KaiC phosphorylation. Science 307: 251–254. Xing W,Busino L, Hinds TR, Marionni ST, Saifee NH, Bush Tu BP, McKnight SL. 2006. Metabolic cycles as an underly- MF,Pagano M, Zheng N. 2013. SCFFBXL3 ubiquitin ligase ing basis of biological oscillations. Nat Rev Mol Cell Biol targets cryptochromes at their cofactor pocket. Nature 7: 696–701. 496: 64–68. Tu BP, McKnight SL. 2007. The yeast metabolic cycle: In- Yan J, Wang H, Liu Y, Shao C. 2008. Analysis of gene regu- sights into the life of a eukaryotic cell. Cold Spring Harb latory networks in the mammalian circadian rhythm. Symp Quant Biol 72: 339–343. PLoS Comput Biol 4: e1000193. Tu BP,Kudlicki A, Rowicka M, McKnight SL. 2005. Logic of Yang Z, Sehgal A. 2001. Role of molecular oscillations in the yeast metabolic cycle: Temporal compartmentaliza- generating behavioral rhythms in Drosophila. Neuron tion of cellular processes. Science 310: 1152–1158. 29: 453–467. Vanden Driessche T. 1966. The role of the nucleus in the Yang KS, Kang SW, Woo HA, Hwang SC, Chae HZ, Kim K, circadian rhythms of Acetabularia mediterranea. Biochim Rhee SG. 2002. Inactivation of human peroxiredoxin I Biophys Acta 126: 456–470. during catalysis as the result of the oxidation of the cat- Van Ooijen G, Millar AJ. 2012. Non-transcriptional oscilla- alytic site cysteine to cysteine-sulfinic acid. J Biol Chem tors in circadian timekeeping. Trends Biochem Sci 37: 277: 38029–38036. 484–492. Yang J, Kim KD, Lucas A, Drahos KE, Santos CS, Mury SP, VealEA, Findlay VJ, Day AM, Bozonet SM, Evans JM, Quinn Capelluto GS, Finkielstein CV. 2008. A novel heme-reg- J, Morgan BA. 2004. A 2-Cys peroxiredoxin regulates ulatory motif mediates heme-dependent degradation of peroxide-induced oxidation and activation of a stress- the circadian factor period 2. Mol Cell Biol 28: 4697– activated MAP kinase. Mol Cell 15: 129–139. 4711. Vivancos AP,Castillo EA, Biteau B, Nicot C, Ayte´ J, Toledano Yin L, Wu N, Curtin JC, Qatanani M, Szwergold NR, Reid MB, Hidalgo E. 2005. A cysteine-sulfinic acid in peroxi- redoxin regulates H O -sensing by the antioxidant Pap1 RA, Waitt GM, Parks DJ, Pearce KH, Wisely GB, et al. 2 2 a pathway. Proc Natl Acad Sci 102: 8875–8880. 2007. Rev-erb , a heme sensor that coordinates metabol- ic and circadian pathways. Science 318: 1786–1789. Wagner E, Frosch S. 1974. Cycles in plants: Endogenous rhythmicity and energy transduction VI. Rhythmicity YoshidaY,Iigusa H, WangN, Hasunuma K. 2011. Cross-talk in reduced and oxidized pyridine nucleotide levels in between the cellular redox state and the circadian system seedlings of Chenopodium rubrum. J Interdisciplinary Cy- in Neurospora. PLoS ONE 6: e28227. cle Res 5: 231–239. Yoshii K, Ishijima S, Sagami I. 2013. Effects of NAD(P)H Wang TA, Yu YV, Govindaiah G, Ye X, Artinian L, Coleman and its derivatives on the DNA-binding activity of TP, Sweedler JV, Cox CL, Gillette MU. 2012. Circadian NPAS2, a mammalian circadian transcription factor. Bi- rhythm of redox state regulates excitability in suprachi- ochem Biophys Res Commun 437: 386–391. asmatic nucleus neurons. Science 337: 839–842. YoshiiK, TajimaF,Ishijima S, Sagami I. 2015. Changes in pH Weidemann A, Johnson RS. 2008. Biology of HIF-1a. Cell and NADPH regulate the DNA binding activity of neu- Death Differ 15: 621–627. ronal PAS domain protein 2, a mammalian circadian Wilking M, Ndiaye M, Mukhtar H, Ahmad N. 2013. Circa- transcription factor. Biochemistry 54: 250–259. dian rhythm connections to oxidative stress: Implications Young MW, Kay SA. 2001. Time zones: A comparative ge- for human health. Antioxid Redox Signal 19: 192–208. netics of circadian clocks. Nat Rev Genet 2: 702–715. Winterbourn CC, Hampton MB. 2008. Thiol chemistry and Zhang Q, Piston DW, Goodman RH. 2002. Regulation of specificity in redox signaling. Free Radic Biol Med 45: corepressor function by nuclear NADH. Science 295: 549–561. 1895–1897.

Cite this article as Cold Spring Harb Perspect Biol 2018;10:a027698 31 Downloaded from http://cshperspectives.cshlp.org/ on October 6, 2021 - Published by Cold Spring Harbor Laboratory Press

Cellular Timekeeping: It's Redox o'Clock

Nikolay B. Milev, Sue-Goo Rhee and Akhilesh B. Reddy

Cold Spring Harb Perspect Biol 2018; doi: 10.1101/cshperspect.a027698 originally published online August 4, 2017

Subject Collection Circadian Rhythms

Circadian Posttranscriptional Regulatory Coordination between Differentially Regulated Mechanisms in Mammals Circadian Clocks Generates Rhythmic Behavior Carla B. Green Deniz Top and Michael W. Young Design Principles of Phosphorylation-Dependent Introduction to Chronobiology Timekeeping in Eukaryotic Circadian Clocks Sandra J. Kuhlman, L. Michon Craig and Jeanne F. Koji L. Ode and Hiroki R. Ueda Duffy Interplay between Microbes and the Circadian Cellular Timekeeping: It's Redox o'Clock Clock Nikolay B. Milev, Sue-Goo Rhee and Akhilesh B. Paola Tognini, Mari Murakami and Paolo Reddy Sassone-Corsi A 50-Year Personal Journey: Location, Gene Molecular Mechanisms of Sleep Homeostasis in Expression, and Circadian Rhythms Flies and Mammals Michael Rosbash Ravi Allada, Chiara Cirelli and Amita Sehgal Regulating the Suprachiasmatic Nucleus (SCN) Membrane Currents, Gene Expression, and Circadian Clockwork: Interplay between Circadian Clocks Cell-Autonomous and Circuit-Level Mechanisms Charles N. Allen, Michael N. Nitabach and Erik D. Herzog, Tracey Hermanstyne, Nicola J. Christopher S. Colwell Smyllie, et al. Systems Chronobiology: Global Analysis of Gene The Plant Circadian Clock: From a Simple Regulation in a 24- Periodic World Timekeeper to a Complex Developmental Manager Jérôme Mermet, Jake Yeung and Felix Naef Sabrina E. Sanchez and Steve A. Kay

For additional articles in this collection, see http://cshperspectives.cshlp.org/cgi/collection/

Copyright © 2018 Cold Spring Harbor Laboratory Press; all rights reserved