Skip to main content

The circadian clock has roles in mesenchymal stem cell fate decision

Abstract

The circadian clock refers to the intrinsic biological rhythms of physiological functions and behaviours. It synergises with the solar cycle and has profound effects on normal metabolism and organismal fitness. Recent studies have suggested that the circadian clock exerts great influence on the differentiation of stem cells. Here, we focus on the close relationship between the circadian clock and mesenchymal stem cell fate decisions in the skeletal system. The underlying mechanisms include hormone signals and the activation and repression of different transcription factors under circadian regulation. Additionally, the clock interacts with epigenetic modifiers and non-coding RNAs and is even involved in chromatin remodelling. Although the specificity and safety of circadian therapy need to be further studied, the circadian regulation of stem cells can be regarded as a promising candidate for health improvement and disease prevention.

Background

Circadian rhythm, also known as the circadian clock, directs and maintains nearly 24-h periodic changes in behaviour and physiology [1]. The central pacemaker of endogenous circadian clocks lies in the SCN (suprachiasmatic nuclei) of the anterior hypothalamus, which is sensitive to environmental cues (called zeitgebers) and orchestrates peripheral synchronicity through hormonal and neural mechanisms [2, 3]. Light input is one of the essential entraining daily oscillations and triggers the sympathetic release of norepinephrine that synchronises the peripheral clocks (Fig. 1a) [4]. The molecular clock is essentially a multilevel transcription–translation feedback loop that integrates a number of regulatory cues, including Bmal1 (brain and muscle ARNT-like protein 1/aryl hydrocarbon receptor nuclear translocator-like protein 1), Clock, Per (period), Cry (cryptochrome), RORα/β/γ (retinoid-related orphan receptor α/β/γ), and Rev–erbα/β (Fig. 1b) [5].

Fig. 1
figure 1

a The circadian clock is distributed throughout the body. The retina perceives light information, which will be relayed to the SCN. In turn, the SCN entrains peripheral system clocks via neural, humoral, and metabolic cues [6]. b The multilevel transcription–translation feedback loop of circadian rhythm. The central loop is based on reciprocal regulation between the BMAL1-CLOCK complex and the PER complex. BMAL1 and CLOCK bind to E-boxes in the promoters of a subset of clock-controlled genes (CCGs) in a heterodimer form, promoting the expression of Per, Cry, Rev-erb, etc. Once PERs and CRY in the cytoplasm accumulate to a certain extent, they will enter the nucleus under GAPVD1 modulation and suppress the transcriptional activity of the BMAL1-CLOCK complex [7]. Casein kinases 1δ/ε phosphorylate PERs, which affects their nuclear translocation and stability [8]. In the second transcription loop, nuclear receptors RORα/β/γ and REV-ERBα/β compete for the ROR element of BMAL1 [9]

The circadian clock has a great influence on the differentiation process in stem cells derived from different sources [10]. Bmal1 knockdown changes the stem cells’ differentiation potential with an aberrant induction of differential gene expression in vitro, and it is essential for transcriptional programmes in cell differentiation [11]. However, how mesenchymal stem cell (MSC) differentiation responds to rhythmic changes timely is a fundamental question to be answered in bone metabolism. Vast amounts of research continue to emerge and are yet to be summarised. This review links new developments to earlier findings, aiming to clarify how circadian clock genes regulate stem cell differentiation fate in the skeletal system and hoping to exploit the MSC differentiation potential by targeting the circadian clock. This review opens up the possibility of a new strategy for MSC regulation and will potentially improve MSC differentiation efficiency for future clinical applications.

  1. 1.

    Stem cells and the circadian clock

Prior to the late 1970s, researchers sought to interrogate how stem cells controlled by the circadian clock perform tissue homeostasis and regenerative functions. Today, rhythmicity has been shown to be instrumental in stem cells’ self-renewal and activation in the event of injury and mammalian development [12]. Multiple studies have confirmed that circadian components exist in undifferentiated ESCs (embryonic stem cells), even before zygotic genome activation [13, 14]. However, ESCs show non-oscillating clock gene expression and are thought to lack clock function. A self-sustained oscillator occurs during postnatal development with the gradual appearance of CLOCK, which is controlled by Dicer/Dgcr8-mediated post-transcriptional mechanisms [15, 16]. Interestingly, Paulose et al. [17] identified the rhythmicity of glucose utilisation and glucose transporter mRNA in ESCs in the absence of clock gene oscillation. This meant that circadian variations in some physiological activities are independent of the core clock, such as peroxiredoxin hyperoxidation [18]. Additionally, there may be other clock regulatory mechanisms in ESCs, and core clock genes may have a non-clock function in ESCs. The deletion of Clock reduces proliferation and increases apoptosis in ESCs, while Bmal1 has been demonstrated to be essential for a proper lineage differentiation of ESCs. Depletion of Bmal1 adversely affects the subsequent differentiation, leading to an imbalance of lineage-specific or selective activation of genes with increased pluripotency transcription factor Nanog, which may be closely related to the change of PI3K, MAPK, and Wnt pathways [19].

Although induced pluripotent stem cells (iPSCs) do not display oscillating clock genes, expression rhythms arise spontaneously in their directed differentiation process [16], which can also be reversed if iPSCs are reprogrammed to pluripotency. This indicates that clock function may be a common feature of differentiated and mature cells [20].

Multipotent stem cells from different tissues show differential clock gene expression profiles with different amplitude ranges. For example, compared with bone marrow stem cells (BMSCs) and adipose tissue-derived mesenchymal stem cells (ADSCs), dental pulp stem cells (DPSCs) have higher Bmal1 expression. After circadian synchronisation by dexamethasone, PER2 peaked at 32 h in BMSCs, with a nadir at 16 h in ADSCs [10]. The circadian clock and daily changes in light and darkness operating at the RNA and protein levels control the proliferation and differentiation of various multipotent cells such as cardiac progenitor cells, stem cells from the epidermis, hair follicle, nervous system, and haematopoietic tissue [4, 12]. For example, the rhythmic Per and Cycle underlie fate decisions and regeneration of intestinal epithelial stem cells [21].

  1. 2.

    MSC differentiation exhibits a circadian pattern in the skeletal system

The circadian regulation of the skeletal system and the differentiation of MSCs have been widely recognised [22]. Luciferase reporter assays detect the 24-h rhythm of OCN mRNA, an osteogenesis marker [23]. BMP2 and RUNX2 change with the light cycle and the level of melatonin [24]. Serum bone turnover markers also show diurnal fluctuations [25], and the irregular light–dark cycle experienced by pregnant mice will affect the offspring’s circadian rhythm and lead to decreased bone perimeter, endocortical volume, and bone strength [26]. The dynamic balance of bone formation and resorption under strict time control ensures bone mass integrity and bone structure and function [27]. The fate change in MSCs resulting from a disrupted circadian clock is implicated in osteoporosis, elevated fracture risk, and osteoarthritis. Continuous exposure to artificial light significantly increases osteoporosis risk in humans and animals [28], and shift workers have a higher fracture risk [29]. Bioinformatic analysis of osteoarthritis patients’ transcriptome shows that clock imbalance may be the leading cause for these complications [30].

The intrinsic molecular clock underlies MSC differentiation at the genetic level. Clock gene (Bmal1, Pers, etc.) knockout and knockdown lead to the phase change of MSC differentiation. Disruption of Clock and Per2 results in a significant reduction in adipogenic differentiation without apparent changes in osteogenic differentiation [31]. BMAL1 oscillates in skeletal tissue and seems to promote osteogenesis and inhibit adipogenesis in a model-dependent manner. Bmal1−/− mice show low bone mass phenotypes with decreased bone trabeculae, bone density, and a thinner cortical bone, resulting from osteogenic MSC differentiation defects [32, 33], and the BMP2 signal may be a key target for BMAL1 [34]. Per−/− and Cry−/− mice show a significant increase in bone mass in both vertebrae and long bones [35]. This high bone phenotype worsens over time [36]. REV-ERBα, a negative regulator of BMAL1, gradually decreases with MSC osteogenesis, and its overexpression reverses the process [37]. However, the results are contradictory. RORα, a positive regulator of BMAL1, increases with MSC osteogenesis and late adipogenesis [38]. Bmal1 interference promotes adipogenesis with Wnt pathway down-regulation in preadipocytes. A high expression pattern of Bmal1 has been testified in mature adipocytes [39]. The knockdown of Bmal1 by RNA interference in mature 3T3-L1 cells only leads to a small amount of lipid droplet formation [40]. These findings suggest the distinct roles of the circadian clock at different differentiation stages that require further research.

The circadian clock regulates MSC differentiation by targeting hormones

Melatonin, synthesised in the pineal gland, is mainly stimulated by darkness and inhibited by light [41]. It keeps the bodily rhythms synchronised with the outside environment and has a positive effect on bone health [42, 43]. Decreased melatonin levels caused by ageing will increase the probability of adverse skeletal events [44], and this can be rescued by its timed administration. Melatonin’s capacity to reduce MSC oxidative stress-induced toxicity and stress-induced cellular senescence by increasing SIRT1 (sirtuin1) [45, 46] provides further evidence on its positive role in promoting bone health [47, 48]. It drives MSCs into an osteoblast lineage by inhibiting miRNA sponge circ-0003865 [49] and by taking part in osteogenesis signalling cascades, such as p38, BMP2, JNK, and Wnt-βcatenin pathways [50, 51]. Melatonin also increases chondrocyte cell volume and enhances chondrogenic differentiation through the Wnt pathway [52].

Parathyroid hormone (PTH) resets the skeletal clock by up-regulating Per mRNA expression [53], and its concentration in bodily fluids exhibits a circadian rhythm. PTH has been widely used in cartilage tissue engineering in recent years, as it inhibits chondrogenic differentiation but mitigates cartilage degeneration. Intermittent PTH protects cartilage by attenuating chondrocyte hypertrophy in 3D-printed tissue engineering scaffolds [54], although it reduces MSC cartilage terminal differentiation markers [55]. The deletion of PTH receptors in MSCs leads to decreased bone formation and increased bone marrow adipose tissue (BMAT), and intermittent PTH administration significantly reduced BMAT in control mice, also found in male patients with osteoporosis [56]. Research identifies that the number of bone marrow adipocytes in men with idiopathic osteoporosis decreased by 27% after 18 months of PTH treatment [56].

Glucocorticoids (GCs) mediate the transcription of Per1 and reset of a peripheral circadian rhythm [57]. SCN controls GC levels in plasma via the hypothalamic–pituitary–adrenal axis, and Per2 may be critical for this process [58]. Oscillation amplitudes of GCs are a crucial determinant of their biological function. The periodic GC-activated GC receptor directly influences the differentiation and functional integration of stem cells in hippocampi. GCs are directly or indirectly tied to osteoporosis by regulating PER2 [59]. Besides, GCs change the methylation profile of cell cycle regulation and Wnt signalling pathway genes [60].

The molecular clock influences MSC multilineage differentiation by regulating transcription factors

Various transcription factors spatiotemporally control the MSC differentiation in response to different stimuli. They directly interpret the genome, and their binding to it is the first step in DNA decoding. In bone metabolism, osteogenic factors, like RUNX2, guide MSCs to the osteogenic lineage; PPARγ promotes adipogenic differentiation; SOX9 participates in chondrogenesis. These transcription factors form a transcription network that plays a crucial role in MSC fate decisions. Network imbalance may lead to skeletal malformations and bone mass disorders. On the other hand, the mechanisms underlying these factors may provide solutions for the treatment of bone diseases.

Regulation and role of transcription factors in osteogenesis

RUNX2 is the ‘master switch’ of osteogenic differentiation, with a highly conserved DNA-binding runt domain. Runx −/− mice show a lack of osteogenic markers and complete blockage of intramembranous and endochondral ossification [61]. Cleidocranial dysplasia resulting from the mutation of RUNX2 is characterised by skeletal dysplasia, such as clavicle hypoplasia and abnormal clavicle [62]. Surprisingly, the overexpression of RUNX2 also shows bone defects in mice, including a dramatic reduction in osteoblast numbers with immature phenotypes. Therefore, RUNX2 plays a major role in MSCs’ early lineage commitment to osteoblasts and maintains the latter at an earlier stage to provide a large number of immature osteoblasts.

RUNX2 shows expression rhythmicity in various tissues, including the SCN and bone [63], and is involved in connecting the circadian clock and MSC fate decision. RUNX2 influences enamel formation in MSCs under the control of circadian pacemaker [64]. Some researchers suppose that RUNX2 relies on the activity of BMAL1, as the rhythm of RUNX2 mRNA disappears in mouse SCN without Bmal1 [63]. Still, some believe REV-ERBα is the upstream regulator of RUNX2 and disturbs the recruitment of RUNX2 to the targeted promoter. Knockdown of REV-ERBα enhances the differentiation and function of pre-osteoblasts [65].

miRNA is also one of the primary mechanisms underlying the rhythm of RUNX2. miR-7-5p disentangles CRY2’s inhibition on the BMAL1-CLOCK complex. The released BMAL1-CLOCK stimulates P300 transcription, promoting histone H3 acetylation and forming a transcription complex with RUNX2 to enhance bone formation (Fig. 2) [66]. Another miRNA, miR-433, shows robust circadian rhythmicity with a peak after dark and targets RUNX2 3’-UTR causing its degradation [67]. In miR-433-inhibited mice skulls, the mRNA levels of RUNX2 and its downstream osteocalcin increased significantly [68].

Fig. 2
figure 2

a Both PPARγ and its cofactor PGC-1α serve as master transcription factors in adipogenic differentiation and are tightly regulated by the circadian clock. b miR-7-5p liberates CRY from the BMAL1-CLOCK complex to promote the transcription of P300. P300 and RUNX2 act simultaneously on the targeted promoter. c BMAL1 promotes the expression of SOX9 and PER1 inhibits SOX6 expression

Besides acting as a transcription factor, RUNX2 also indirectly regulates the transcription of clock genes and/or clock-controlled genes (CCGs) by working as a scaffold to recruit growth factors and cytokines [63, 69], including P300, CCAAT/enhancer-binding protein, and histone deacetylase [70], with its protein–protein interacting domains.

To our knowledge, Osterix, cAMP-dependent transcription factor 4 (ATF4), Forkhead box O (FOXO), and other transcription factors in osteogenic differentiation have rarely been described in circadian rhythm studies. The role of these transcription factors in circadian rhythm remains to be identified.

Adipogenic master regulator PPARγ

Peroxisome proliferator-activated receptor γ (PPARγ) belongs to the nuclear hormone receptor superfamily with the highest expression in adipose tissue. It plays a vital role in the balance between osteogenic and adipogenic differentiation of MSCs [71,72,73]. PPARγ up-regulation promotes adipogenic differentiation and inhibits the generation of osteoblasts and chondrocytes [74], with its down-regulation having opposite effects [75]. Besides, the PPARγ agonist thiazolidinedione (TZD) is associated with bone loss and fracture [76]. Clinical experience and animal studies suggest that the fate change of MSCs associated with PPARγ may be the main reason for this.

PPARγ and its cofactor PGC-1α display a clear circadian expression pattern in many tissues, and PPARγ even acts as a crucial peripheral clock activator of the cardiovascular system and metabolism [77, 78]. Multiple randomised clinical trials have shown that the clock hormone melatonin up-regulates PPARγ [79, 80]. And the inhibition of Clock or Per2 changes the expression of osteocalcin in MSCs, in association with extraordinary increases in PPARγ [31]. BMAL1 and REV-ERBα are induced by PPARγ and PGC-1α [77, 81, 82]. PPARγ is deeply involved in the regulatory networks between circadian rhythms and MSC fate decisions (Fig. 2), with the mechanisms described hereafter.

Nocturnin−/− mice show deregulated circadian patterns in PPARγ gene expression and increased bone mass [83]. Nocturnin adapts to daily environmental changes by changing the organism’s metabolic state [84, 85]. It works as a pivotal mediator of mitochondrial function by dephosphorylating nicotinamide adenine dinucleotide phosphate (NADP+/NADPH) [86]. Nocturnin is highly expressed in MSCs, and its overexpression enhances the lipogenesis of preadipocytes 3T3-L1. It enhances PPARγ transcription activities and guides PPARγ nuclear translocation.

Per1 and Per2 have antagonistic influences on PPARγ activity. PER1 enhances PPARγ activity operating as its natural modulator, while PER2 directly and specifically inhibits the recruitment of PPARγ to the target promoter [87]. However, the PER2 regulation of adipogenesis has significant cell/tissue specificity [87], as it is regulated by a complicated network and works with other core clock products, such as PER1, BMAL1, and REV-ERBα. For example, a substantial reduction in PPARγ is found in per1/2−/− mice [88]. However, Per3 is believed to be a nonessential clock component with modest effects on circadian behaviours [89, 90], and it does inhibit adipogenesis in vivo in a clock manner [91]. The deletion of Per3 promotes adipogenesis by clock outputs [92]. PER3 binds to PPARγ and inhibits its activity [91, 93]. Increasing PPAR is observed in Per3−/− mice, which means that the differentiation arrest resulting from Per3 happens during the early stages.

D-box binding protein (DBP) and E4 promoter-binding protein 4 (E4BP4), contributors to circadian regulatory loops, interact with the active site of the PPARγ promoter to regulate its rhythmic expression and are enhanced by histone deacetylase inhibitor (HDACi) [94]. Decreased H3K9 acetylation of the DBP promoter leads to a lower level of DBP and a diminished binding between DBP and the PPARγ promoter, which gives rise to a lower level of PPARγ [95].

Chondrogenic factors

The circadian clock is critical to early chondrogenesis and endochondral ossification [96, 97]. Indian Hedgehog (Ihh) is key in governing chondrogenic differentiation, and its activity is regulated by PER1 and BMAL1 in the growth plate [97]. Bmal1 deficiency leads to the deregulated rhythmic expression of Ihh and Per [98]. Our previous studies show that Prg4 and related chondrogenic genes are downstream targets of Bmal1 in cartilage formation and chondrogenic differentiation [99]. Cartilage tissue engineering, which combines MSCs with other factors, such as collagen, is a promising therapy for degenerated temporomandibular joint and osteoarthritis. Sry-related HMG box 9 (SOX9), a key transcription regulator of chondrocyte differentiation [100], controls the synthesis of collagen type II, the hallmark of chondrogenic commitment [101]. Impaired circadian rhythm induces osteoarthritis resulting from SOX9 inhibition in articular chondrocytes, and the deletion of Bmal1 also leads to reduced SOX9 [102] (Table 1). Two additional SOX family members, SOX6 and SOX5, are also involved in chondrogenesis regulation. SOX6 is negatively regulated by its transactivator PER1 (Fig. 2). It shows a long-term decrease after exposure to 10 nM PTH for 1 h in chondrocyte-like ATDC5 cells with no distinct changes in level SOX5 and RUNX2 levels [103].

Table 1 Epigenetic modifying enzymes shared both in the circadian clock and in the MSC fate decisions

The circadian clock determines MSC fate through epigenetic events

Approximately 43% of mammalian genes oscillate in a circadian fashion, but no more than 30% of circadian mRNAs are driven by de novo transcription, indicating that post-transcriptional mechanisms mainly regulate the rhythmic expression of genes in a cell-/tissue-specific fashion [109, 110]. Epigenetics is an essential component of post-transcriptional regulation and has rapidly developed into various fields and disciplines in the twenty-first century, enhancing our understanding of gene regulation in health and disease.

The term ‘epigenetics’ refers to heritable alterations in gene expression without directly altering the nucleotide sequence [111]. The epigenome mediates a crosstalk between cells and the environment through covalent modifications in chromatin (also called epigenetic code) and epigenetic regulators, which include DNA methylation, histone modification, non-coding RNAs, etc. [112]. The epigenome opens or closes chromatin fibre around genes to control subsequent transcription factor binding, exerting great influence on MSC differentiation [113]. The circadian clock has also been implicated in stem cell differentiation, which far exceeds its daily timekeeping roles [114]. Interestingly, accumulating research shows the close relationship between the epigenetic mechanism and circadian rhythm. The rhythmic transcriptome is assisted by a specific epigenetic mechanism which cooperates with the clock to orchestrate daily oscillations [115]. Therefore, we postulate that epigenetic regulators function in the crosstalk between the circadian clock and stem cell fate decisions to determine how and when transcription oscillations occur. The coordinated mechanisms include the clock proteins and epigenetic modifying enzymes constituting complexes to regulate transcription and the collaborative interaction between the clock and non-coding RNAs or chromatin remodellers.

Circadian proteins form complexes with epigenetic modifying enzymes

  1. (1)

    Histone-modifying enzymes

Genome-wide expression analysis reveals the rhythm of global histone labels that, depending on an intact clock, are under tight regulation by enzymatic feedback loops [109]. The clock recruits epigenetic enzymes to the promoter of important transcription factors and key signalling molecules involved in differentiation. Besides, CLOCK, functioning as an epigenetic modifier, was subsequently found to have a histone acetyltransferase (HAT) activity greatly amplified by its combination with BMAL1 [116]. Additionally, clock-controlled SIRTs and EZH2 provide possibility in elucidating how the clock is linked to epigenetic mechanism in MSCs differentiation.

At the stage of transcriptional activation, the BMAL1-CLOCK complex acetylates the RUNX2 promoter by Clock itself and by recruiting acetyltransferases P300/CBP, thus promoting osteogenesis in MSCs [66]. At the repression stage, CRY (a constitutive component of the PER complex) occupies the C-terminal transactivation domain of BMAL1, leaving P300/CBP dissociated from its binding site [117]. Once CRY is loaded to the BMAL1-CLOCK complex, it directly recruits the SIN3 complex that comprises HDAC1 and HDAC2 to remove the acetylation tag of promoters (Fig. 3). Thus, the PER-SIN-HDACs pathway targeting histones H3 and H4 controls gene expression precisely and dynamically [118]. In addition, REV-ERBα and RORα exert downstream effects on the expression of NF-kB (which inhibits osteoblast differentiation) by recruiting HDACs [104]. REV-ERBα calls up co-repressors NCOR1/SMRT to ROR element. HDAC3, serving as the catalytic component of NCOR1/SMRT, rhythmically modulates histone deacetylation of the ROR element [119, 120].

Fig. 3
figure 3

The PER complex, competing with P300/CBP, binds the BMAL1-CLOCK complex. SIRT1 and REV-ERBα/β precisely regulate this process: SIRT1 prompts PER complex dissociation from the BMAL1-CLOCK complex and promotes PER2 degradation. It also plays a negative role in CLOCK HAT activity. REV-ERBα/β recruits the co-repressors NCOR1/SMRT to ROR element to promote P300/CBP binding with the BMAL1-CLOCK complex

Sirtuins (SIRTs) are class III NAD + dependent deacetylases that participate in many biological processes, including the circadian clock and cell differentiation [121,122,123]. This family has seven members (SIRT1–7) distributed across the cytoplasm, nucleus, and mitochondria. SIRT1 binds BMAL1-CLOCK in a circadian manner and counteracts the HAT activity of CLOCK [124]. Its deacetylation ability extends to BMAL1 and PER2 and relieves BMAL1-CLOCK from CRY and promotes PER2 degradation [125, 126]. SIRT1 requires the cofactors NAD + and NAMPT to function, which are both under tight circadian control [86, 127]. Thus, the deacetylase activity of SIRT1 exhibits a circadian pattern.

The MSC-specific overexpression of SIRT1 increased alveolar bone mass in vivo and promoted MSCs’ proliferation and osteogenic differentiation [128]. Aged mice with a MSC-specific knockout of SIRT1 show reduced subcutaneous fat and a progressive loss of trabecular and cortical bone volume [129]. SIRT1 exerts its effects mainly by deacetylating important transcription factors and signalling pathways, influencing their activity and localisation. For example, SIRT1 directly or indirectly deacetylates RUNX2and PPARγ [130, 131] and deacetylates β-catenin, facilitating its accumulation in the nucleus [132]. Along these lines, SIRT1, serving as a circadian hallmarker, causes dynamic changes in the epigenome to generate transcriptional responses controlling MSC differentiation.

Methylation is also a widespread histone modification in mammalian cells and is linked to transcriptional repression. EZH2, mediating gene silencing via trimethylation of H3K27, is directly regulated by the circadian clock via the E-BOX and RORE motifs in fish species [133]. EZH2 is a repressor of differentiation-associated genes in vitro. It inhibits osteogenesis by inhibiting BMP2 and inhibits chondrogenic differentiation by inhibiting ATOH8 [106, 107] (Table 1). However, in in-vivo animal studies, the additive effects of VDR‐mediated EZH2 transcriptional increase show increased osteogenesis, as EZH2 also inhibits MSC senescence by repressing p16 transcription [134]. MSCs are cells of mesodermal origin which derived from ESCs. The level of EZH2 in MSCs is significantly lower than that in ESCs, and a specific EZH2 inhibitor, GSK126, dramatically drives ESCs differentiation into MSCs by reducing H3K27me3 [135], suggesting that EZH2 levels in specific cell types are crucial for the balance of pluripotency maintenance, cell differentiation, and cell senescence.

  1. (2)

    DNA-modifying enzymes

Tissue-specific DNA methylation patterns play central roles in establishing cell identity during differentiation. Generally, DNA methylation is considered one of the hallmark mechanisms of gene silencing. For example, DNMT inhibitors, such as 5-aza-2'-deoxycytidine, induce gene activation and cell differentiation [136]. However, the genome-wide DNA methylation profiling of MSCs shows an increasing methylation level of the global genome and CpG sites in the process of osteogenic differentiation [137]. These results appear to be incompatible with the gene-silencing role of DNA methylation. Recent studies suggest that the function of DNA methylation varies depending on the methylation location. The methylation of the gene body, and not the promoter, promotes genome stability and even transcription [138], suggesting that promoter segments may not be the DNA methyltransferase target sites during MSCs differentiation.

In recent years, growing evidence in chronobiology hints at a close connection between the circadian rhythm and DNA methylation. DNMTs oscillate in a clock pattern in the mouse liver and brain [139], and excessive exposure to artificial light leads to tissue-specific changes in DNMTs expression [140]. Taken together, this evidence suggests that DNMT may exert great influence during clock-controlled differentiation.

Reciprocal regulation between the circadian clock and non-coding RNAs

miRNAs are a class of short non-coding RNAs with a length of about 22 nucleotides. The circadian clock regulates miRNAs expression and activity, and the clock itself is regulated by miRNAs at post-transcriptional levels [141]. And miRNAs taking part in clock regulation have been extensively reviewed [142, 143]. In healthy young males, some circulating miRNAs display 24-h rhythms with two expression patterns: one peaks at night and the other peaks during the daytime [144, 145]. There are 104 miRNAs differently expressed in CLOCK-mutant mice compared with the controlled group, and they are linked to transcription regulation and PI3K-AKT and MAPK signalling pathways [146], which are closely related to MSC differentiation.

However, some researchers argue that mature miRNAs do not oscillate in a circadian pattern. Studies show at least 57 pri-miRNAs oscillated in the mouse liver, with only four mature miRNAs showing the same rhythm as their primary transcript [147]. Interestingly, the rhythmic expression of Dicer (the essential enzyme for miRNA maturation) is found in various tissues [148]. Studies are contradictory and may be model-dependent. The inconsistent findings among these studies can be partly explained by variations in the study samples, methods of assessment, and, most importantly, tissue specificity, which may hint at the role of circadian miRNA in establishing cell identity.

Mounting research has unveiled the importance of miRNAs in the regulation of MSC differentiation. For example, the expression profile in patients with fractures or osteoporosis differs from that in healthy individuals. Hackl et al. summarized the different expression of miRNAs in patients with bone disease and the control group. They rasied the possibilility that circulating miRNAs are used as biomarkers for bone disease [149]. Furthermore, many miRNAs are shared between the circadian clock and MSCs fate decisions. A candidate miRNA set potentially involved in circadian control has been established by various means, including microarrays [150], CHIP [142], and in silico prediction analysis of databases [151]. We found that some of them also function in the differentiation process of MSCs, as summarised in the Additional file 1.

Long non-coding RNAs (lncRNAs) are non-coding RNAs with 2000–100,000 nucleotides in length. A mouse transcriptome analysis indicated that more than 1000 lncRNAs show a diurnal rhythm [152]. The expression of a substantial portion of rhythmic lncRNAs is pronounced in regions of enhancer clusters and regulates gene expression, showing high similarity in rhythm phase and expression patterns with genes in their proximity. For instance, lnc-CROT is highly expressed at super-enhancer regions and modulates long-range circadian gene regulation [153]. Accumulating research has shown that lncRNAs, such as lnc-MEG3 and lnc-XIST, take part in osteogenic differentiation [154, 155]. Unfortunately, there are few reports on the role of lncRNAs in the clock-controlled differentiation process, and follow-up bioinformatics studies should be carried out.

Circadian proteins influencing chromatin architecture

Different kinds of epigenetic mechanisms show a pile-up effect on gene expression levels by influencing chromatin architecture and accessibility, which influence the binding of transcription factors to the genome. The molecular clock, working with various epigenetic regulators, contributes to spatiotemporal balance and the oscillation of chromatin accessibility [109, 156]. For example, RORα/β, BMAL1, and other circadian proteins promote global chromatin decondensation during the activation phase via the SWI/SNF complex (one of the epigenetic remodelling complexes). In turn, this promotes the load of REV-ERB, which rapidly responds to structure changes and inhibits circadian protein expression [157]. Furthermore, the chromatin landscape shaped by circadian regulation is critical for MSCs self-renewal and multilineage differentiation [158].

Future perspective

Circadian dysfunction is associated with various diseases in mammals [159], such as myocardial infarctions (MIs) and bone diseases [160]. Interrupted light at night causes depressive-like behaviours in mice by the circadian ipRGCs-dpHb-NAc (retinal ganglion cells–the dorsal perihabenular nucleus–the nucleus accumbens) pathway [161]. Besides, the symptom complexity and the onset of various diseases show marked circadian rhythm. For example, the incidence of MIs at 9 a.m. is three times greater than that at 11 p.m., periods associated with the blood pressure rhythm [162]. Thus, manipulating and maintaining the circadian rhythm is important for disease prevention. The circadian rhythm of disease onset and chrono-drug discovery should be taken into account when deciding treatment times [163].

Luckily, the circadian rhythm can be remodelled by environmental cues. Light, temperature, eating frequency, and regular exercises are proved to be daily circadian temporal cues [164, 165]. Regular exercise rescued the disrupted circadian rhythm by increasing Per2 expression amplitude [166]. Time-restricted feeding increasing circadian amplitude prevents metabolic diseases and obesity in mice fed a high-fat diet [167]. Short-term caloric restriction is associated with cardio-protection [168]. This suggests that we may alter MSC function and restore bone function and structure by rectifying eating frequency and the sleep–wake pattern and reducing night-time exposure to artificial light to reset the circadian clock.

In addition to behavioural improvements, mechanical stress, biotic factors, and small molecule modifiers are promising interventions for a robust rhythm. Mechanical stress regulates the clock of skeletal muscle cells (C2C12 cells) by decreasing Per/Cry expression and increasing Bmal1 expression [169]. The Per1 inhibition of BMSC by artificial materials can be rescued by vitamin D supplements and osteogenic media [170]. Melatonin supplements, overcoming rhythm disruption, exhibit excellent treatment efficacy and safety as ideal candidates for preventing bone loss. GCs like dexamethasone are widely used to generate synchronised clocks in vitro. Some small molecule modifiers of circadian clocks alter the circadian period length, phase, and amplitude with favourable benefits. For example, Rev-erb activation by SR9009 is beneficial for the reparative process after ischaemia–reperfusion injury by attenuating NLRP3 inflammasome activation and immune infiltration, being also essential for protection against the development of cardiac hypertrophy and heart failure in mice [171, 172]. Furthermore, the proliferation and differentiation of neural stem cells are promoted by SR9009 in a dose-dependent manner through targeting Rev-erb [173]. However, clock is found in almost all tissues in mammals, and systematic influences on clock activity may interrupt normal physiological processes in healthy tissues. Thus, the specificity and safety of those modifiers should be fully reviewed before it is widely applied in clinical practice.

Additionally, different stem cells are sensitive to specific entraining cues with different circadian patterns. In DPSCs, mechanosensitive synchronisation is more accessible than dexamethasone synchronisation. BMSCs are more responsive to chemical stimulation than dexamethasone, and stem cells show different responses to entraining cues at certain times of the day [10, 162]. This hints at specific entraining choices for the circadian rhythm of different cells at specific times.

Our study has several limitations, since the interplay between the circadian rhythm and cell differentiation is a complicated network rather than a unidirectional hierarchical structure. How chromatin modifiers [174], systemic conditions, signalling pathways, kinases and phosphatases [175], RNA-binding factors [176], and other factors are involved in the whole regulatory network remains to be identified. However, the development of high-throughput techniques and novel sequencing technologies makes it possible to depict this underlying network, which is beneficial for a deeper understanding of MSC [1]. We may potentially improve MSC differentiation efficiency for future clinical applications.

Conclusions

The connections between the circadian rhythm and cell differentiation have been intensively researched in recent years. The underlying associated mechanisms remain to be summarised. This review elaborates on key coupling mechanisms, which include the following: (1) The circadian clock affects hormone secretion to regulate cell differentiation; (2) molecular clock core products directly modulate the activity and expression of key transcription factors; and (3) histone modifications, DNA modification, non-coding RNAs, and large multi-subunit chromatin remodelling complexes and other epigenetic modifications are critical nodes in the circadian regulation of cell fate decision. We offer an in-depth insight into the interplay between the circadian clock and cell differentiation in the skeletal system. Additionally, we suggest that incorporating circadian regulation (ranging from chemical, biological, and physical cues) may have profound benefits in optimising tissue engineering approaches and stem cell therapies.

Availability of data and materials

All data generated or analysed during this study are included in this published article.

Abbreviations

SCN:

Suprachiasmatic nuclei

Bmal1:

Brain and muscle ARNT-like protein 1

Per:

Period

Cry:

Cryptochrome

RORα/β/γ:

Retinoid-related orphan receptor α/β/γ

MSCs:

Mesenchymal stem cells

ESCs:

Embryonic stem cells

iPSCs:

Induced pluripotent stem cells

BMSCs:

Bone marrow stem cells

ADSCs:

Adipose tissue-derived mesenchymal stem cells

DPSCs:

Dental pulp stem cells

PTH:

Parathyroid hormone

BMAT:

Bone marrow adipose tissue

GCs:

Glucocorticoids

SOX9:

Sry-related HMG box 9

PPAR γ:

Peroxisome proliferator-activated receptor γ

TZD:

Thiazolidinedione

CCGs:

Clock-controlled genes

ATF4:

CAMP-dependent transcription factor 4

FOXO:

Forkhead box O

NADP + /NADPH:

Nicotinamide adenine dinucleotide phosphate

DBP:

D-box binding protein

E4BP4:

E4 promoter-binding protein 4

CBP:

CREB-binding protein

HDACs:

Histone deacetylases

SIRTs:

Sirtuins

EZH2:

Histone-lysine N-methyltransferase

DNMTs:

DNA methyltransferases

CpG:

Cytosine-phosphate-guanine

lncRNAs:

Long non-coding RNAs

MIs:

Myocardial infarctions

ipRGCs-dpHb-NAc:

Retinal ganglion cells–the dorsal perihabenular nucleus–the nucleus accumbens

References

  1. Chirico N, Van Laake LW, Sluijter JPG, et al. Cardiac circadian rhythms in time and space: the future is in 4D. Curr Opin Pharmacol. 2021;57:49–59.

    Article  CAS  PubMed  Google Scholar 

  2. Silver R, LeSauter J, Tresco PA, et al. A diffusible coupling signal from the transplanted suprachiasmatic nucleus controlling circadian locomotor rhythms. Nature. 1996;382(6594):810–3.

    Article  CAS  PubMed  Google Scholar 

  3. Yamazaki S, Numano R, Abe M, et al. Resetting central and peripheral circadian oscillators in transgenic rats. Science (New York, NY). 2000;288(5466):682–5.

    Article  CAS  Google Scholar 

  4. Fan SMY, Chang YT, Chen CL, et al. External light activates hair follicle stem cells through eyes via an ipRGC-SCN-sympathetic neural pathway. Proc Natl Acad Sci USA. 2018;115(29):E6880–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Dunlap JC. Molecular bases for circadian clocks. Cell. 1999;96(2):271–90.

    Article  CAS  PubMed  Google Scholar 

  6. Dibner C, Schibler U, Albrecht U. The mammalian circadian timing system: organization and coordination of central and peripheral clocks. Annu Rev Physiol. 2010;72:517–49.

    Article  CAS  PubMed  Google Scholar 

  7. Aryal RP, Kwak PB, Tamayo AG, et al. Macromolecular assemblies of the mammalian circadian clock. Mol Cell. 2017;67(5):770–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Narasimamurthy R, Hunt SR, Lu Y, et al. CK1δ/ε protein kinase primes the PER2 circadian phosphoswitch. Proc Natl Acad Sci USA. 2018;115(23):5986–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Zhang Y, Fang B, Emmett MJ, et al. Gene regulation. Discrete functions of nuclear receptor Rev-erbα couple metabolism to the clock. Science (New York, NY). 2015;348(6242):1488–92.

    Article  CAS  Google Scholar 

  10. Rogers EH, Fawcett SA, Pekovic-Vaughan V, et al. Comparing circadian dynamics in primary derived stem cells from different sources of human adult tissue. Stem Cells Int. 2017;2017:2057168.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  11. Ameneiro C, Moreira T, Fuentes-Iglesias A, et al. BMAL1 coordinates energy metabolism and differentiation of pluripotent stem cells. Life Sci Alliance. 2020;3(5).

  12. Janich P, Pascual G, Merlos-Suárez A, et al. The circadian molecular clock creates epidermal stem cell heterogeneity. Nature. 2011;480(7376):209–14.

    Article  CAS  PubMed  Google Scholar 

  13. Hamatani T, Carter MG, Sharov AA, et al. Dynamics of global gene expression changes during mouse preimplantation development. Dev Cell. 2004;6(1):117–31.

    Article  CAS  PubMed  Google Scholar 

  14. Delaunay F, Thisse C, Thisse B, et al. Differential regulation of Period 2 and Period 3 expression during development of the zebrafish circadian clock. Gene Expr Patterns. 2003;3(3):319–24.

    Article  CAS  PubMed  Google Scholar 

  15. Umemura Y, Koike N, Ohashi M, et al. Involvement of posttranscriptional regulation of in the emergence of circadian clock oscillation during mouse development. Proc Natl Acad Sci USA. 2017;114(36):E7479–88.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Dierickx P, Vermunt MW, Muraro MJ, et al. Circadian networks in human embryonic stem cell-derived cardiomyocytes. EMBO Rep. 2017;18(7):1199–212.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Paulose JK, Rucker EB, Cassone VM. Toward the beginning of time: circadian rhythms in metabolism precede rhythms in clock gene expression in mouse embryonic stem cells. PLoS ONE. 2012;7(11): e49555.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Putker M, O’Neill JS. Reciprocal control of the circadian clock and cellular redox state—a critical appraisal. Mol Cells. 2016;39(1):6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Gallardo A, Molina A, Asenjo HG, et al. The molecular clock protein Bmal1 regulates cell differentiation in mouse embryonic stem cells. Life Sci Alliance. 2020;3(5):e201900535.

    Article  PubMed  PubMed Central  Google Scholar 

  20. Umemura Y, Yoshida J, Wada M, et al. An in vitro ES cell-based clock recapitulation assay model identifies CK2α as an endogenous clock regulator. PLoS ONE. 2013;8(6): e67241.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Karpowicz P, Zhang Y, Hogenesch JB, et al. The circadian clock gates the intestinal stem cell regenerative state. Cell Rep. 2013;3(4):996–1004.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Dudek M, Meng Q-J. Running on time: the role of circadian clocks in the musculoskeletal system. Biochem J. 2014;463(1):1–8.

    Article  CAS  PubMed  Google Scholar 

  23. Gafni Y, Ptitsyn AA, Zilberman Y, et al. Circadian rhythm of osteocalcin in the maxillomandibular complex. J Dent Res. 2009;88(1):45–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Witt-Enderby PA, Slater JP, Johnson NA, et al. Effects on bone by the light/dark cycle and chronic treatment with melatonin and/or hormone replacement therapy in intact female mice. J Pineal Res. 2012;53(4):374–84.

    Article  CAS  PubMed  Google Scholar 

  25. Tsang K, Liu H, Yang Y, et al. Defective circadian control in mesenchymal cells reduces adult bone mass in mice by promoting osteoclast function. Bone. 2019;121:172–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Chaves I, van der Eerden B, Boers R, et al. Gestational jet lag predisposes to later-life skeletal and cardiac disease. Chronobiol Int. 2019;36(5):657–71.

    Article  CAS  PubMed  Google Scholar 

  27. Feng X, McDonald JM. Disorders of bone remodeling. Annu Rev Pathol. 2011;6:121–45.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Lucassen EA, Coomans CP, van Putten M, et al. Environmental 24-hr cycles are essential for health. Curr Biol CB. 2016;26(14):1843–53.

    Article  CAS  PubMed  Google Scholar 

  29. Feskanich D, Hankinson SE, Schernhammer ES. Nightshift work and fracture risk: the Nurses’ Health Study. Osteoporos Int J Establ Result Coop Between Eur Found Osteoporos Natl Osteoporos Found USA. 2009;20(4):537–42.

    Article  CAS  Google Scholar 

  30. Soul J, Dunn SL, Anand S, et al. Stratification of knee osteoarthritis: two major patient subgroups identified by genome-wide expression analysis of articular cartilage. Ann Rheum Dis. 2018;77(3):423.

    Article  CAS  PubMed  Google Scholar 

  31. Boucher H, Vanneaux V, Domet T, et al. Circadian clock genes modulate human bone marrow mesenchymal stem cell differentiation, migration and cell cycle. PLoS ONE. 2016;11(1): e0146674.

    Article  PubMed  PubMed Central  Google Scholar 

  32. Lieben L. Bone: The circadian clock controls bone remodelling. Nat Rev Rheumatol. 2016;12(3):132.

    Article  CAS  PubMed  Google Scholar 

  33. Min H-Y, Kim K-M, Wee G, et al. Bmal1 induces osteoblast differentiation via regulation of BMP2 expression in MC3T3-E1 cells. Life Sci. 2016;162:41–6.

    Article  CAS  PubMed  Google Scholar 

  34. Huang Z, Wei H, Wang X, et al. Icariin promotes osteogenic differentiation of BMSCs by upregulating BMAL1 expression via BMP signaling. Mol Med Rep. 2020;21(3):1590–6.

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Maronde E, Schilling AF, Seitz S, et al. The clock genes Period 2 and Cryptochrome 2 differentially balance bone formation. PLoS ONE. 2010;5(7): e11527.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  36. Fu L, Patel MS, Bradley A, et al. The molecular clock mediates leptin-regulated bone formation. Cell. 2005;122(5):803–15.

    Article  CAS  PubMed  Google Scholar 

  37. He Y, Lin F, Chen Y, et al. Overexpression of the circadian clock gene Rev-erbα affects murine bone mesenchymal stem cell proliferation and osteogenesis. Stem Cells Dev. 2015;24(10):1194–204.

    Article  CAS  PubMed  Google Scholar 

  38. Meyer T, Kneissel M, Mariani J, et al. In vitro and in vivo evidence for orphan nuclear receptor RORalpha function in bone metabolism. Proc Natl Acad Sci USA. 2000;97(16):9197–202.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Tal Y, Chapnik N, Froy O. Non-obesogenic doses of palmitate disrupt circadian metabolism in adipocytes. Adipocyte. 2019;8(1):392–400. 

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Shimba S, Ishii N, Ohta Y, et al. Brain and muscle Arnt-like protein-1 (BMAL1), a component of the molecular clock, regulates adipogenesis. Proc Natl Acad Sci USA. 2005;102(34):12071–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Stehle JH, Saade A, Rawashdeh O, et al. A survey of molecular details in the human pineal gland in the light of phylogeny, structure, function and chronobiological diseases. J Pineal Res. 2011;51(1):17–43.

    Article  CAS  PubMed  Google Scholar 

  42. Feng Z-Y, Yang S-D, Wang T, et al. Effect of melatonin for regulating mesenchymal stromal cells and derived extracellular vesicles. Front Cell Dev Biol. 2021;9: 717913.

    Article  PubMed  PubMed Central  Google Scholar 

  43. Lu X, Yu S, Chen G, et al. Insight into the roles of melatonin in bone tissue and bone-related diseases (Review). Int J Mol Med. 2021;47(5):1–19.

    Article  CAS  Google Scholar 

  44. Amstrup AK, Sikjaer T, Heickendorff L, et al. Melatonin improves bone mineral density at the femoral neck in postmenopausal women with osteopenia: a randomized controlled trial. J Pineal Res. 2015;59(2):221–9.

    Article  CAS  PubMed  Google Scholar 

  45. Munmun F, Witt-Enderby PA. Melatonin effects on bone: implications for use as a therapy for managing bone loss. J Pineal Res. 2021;71(1): e12749.

    Article  CAS  PubMed  Google Scholar 

  46. Chen W, Chen X, Chen AC, et al. Melatonin restores the osteoporosis-impaired osteogenic potential of bone marrow mesenchymal stem cells by preserving SIRT1-mediated intracellular antioxidant properties. Free Radic Biol Med. 2020;146:92–106.

    Article  CAS  PubMed  Google Scholar 

  47. Wu KK. Control of mesenchymal stromal cell senescence by tryptophan metabolites. Int J Mol Sci. 2021;22(2):697.

    Article  CAS  PubMed Central  Google Scholar 

  48. Cao Z, Geng X, Jiang X, et al. Melatonin attenuates AlCl-induced apoptosis and osteoblastic differentiation suppression by inhibiting oxidative stress in MC3T3-E1 cells. Biol Trace Elem Res. 2020;196(1):214–22.

    Article  CAS  PubMed  Google Scholar 

  49. Wang X, Chen T, Deng Z, et al. Melatonin promotes bone marrow mesenchymal stem cell osteogenic differentiation and prevents osteoporosis development through modulating circ_0003865 that sponges miR-3653-3p. Stem Cell Res Ther. 2021;12(1):150.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Zhang J, Jia G, Xue P, et al. Melatonin restores osteoporosis-impaired osteogenic potential of bone marrow mesenchymal stem cells and alleviates bone loss through the HGF/PTEN/Wnt/β-catenin axis. Ther Adv Chronic Dis. 2021;12:2040622321995685.

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Zhou Y, Wang C, Si J, et al. Melatonin up-regulates bone marrow mesenchymal stem cells osteogenic action but suppresses their mediated osteoclastogenesis via MT2-inactivated NF-κB pathway. Br J Pharmacol. 2020;177(9):2106–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Wang X, He T, He L, et al. Melatonin contributes to the hypertrophic differentiation of mesenchymal stem cell-derived chondrocytes via activation of the Wnt/β-catenin signaling pathway: melatonin promotes MSC-derived chondrocytes hypertrophy. Stem Cell Res Ther. 2021;12(1):467.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Hinoi E, Ueshima T, Hojo H, et al. Up-regulation of per mRNA expression by parathyroid hormone through a protein kinase A-CREB-dependent mechanism in chondrocytes. J Biol Chem. 2006;281(33):23632–42.

    Article  CAS  PubMed  Google Scholar 

  54. Music E, Futrega K, Palmer JS, et al. Intermittent parathyroid hormone (1–34) supplementation of bone marrow stromal cell cultures may inhibit hypertrophy, but at the expense of chondrogenesis. Stem Cell Res Ther. 2020;11(1):321.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Zhao W, Zou T, Cui H, et al. Parathyroid hormone (1–34) promotes the effects of 3D printed scaffold-seeded bone marrow mesenchymal stem cells on meniscus regeneration. Stem Cell Res Ther. 2020;11(1):328.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Fan Y, Hanai J-I, Le PT, et al. Parathyroid hormone directs bone marrow mesenchymal cell fate. Cell Metab. 2017;25(3):661–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Balsalobre A, Brown SA, Marcacci L, et al. Resetting of circadian time in peripheral tissues by glucocorticoid signaling. Science (New York, NY). 2000;289(5488):2344–7.

    Article  CAS  Google Scholar 

  58. Nakamura Y, Harama D, Shimokawa N, et al. Circadian clock gene Period2 regulates a time-of-day-dependent variation in cutaneous anaphylactic reaction. J Allergy Clin Immunol. 2011;127(4):1038–45.

    Article  CAS  PubMed  Google Scholar 

  59. Abe T, Sato T, Yoda T, et al. The period circadian clock 2 gene responds to glucocorticoids and regulates osteogenic capacity. Regen Ther. 2019;11:199–206.

    Article  PubMed  PubMed Central  Google Scholar 

  60. Schouten M, Bielefeld P, Garcia-Corzo L, et al. Circadian glucocorticoid oscillations preserve a population of adult hippocampal neural stem cells in the aging brain. Mol Psychiatry. 2020;25(7):1382–405.

    Article  CAS  PubMed  Google Scholar 

  61. Komori T. Runx2, an inducer of osteoblast and chondrocyte differentiation. Histochem Cell Biol. 2018;149(4):313–23.

    Article  CAS  PubMed  Google Scholar 

  62. Mundlos S, Otto F, Mundlos C, et al. Mutations involving the transcription factor CBFA1 cause cleidocranial dysplasia. Cell. 1997;89(5):773–9.

    Article  CAS  PubMed  Google Scholar 

  63. Reale ME, Webb IC, Wang X, et al. The transcription factor Runx2 is under circadian control in the suprachiasmatic nucleus and functions in the control of rhythmic behavior. PLoS ONE. 2013;8(1): e54317.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Athanassiou-Papaefthymiou M, Kim D, Harbron L, et al. Molecular and circadian controls of ameloblasts. Eur J Oral Sci. 2011;119(Suppl 1):35–40.

    Article  PubMed  PubMed Central  Google Scholar 

  65. Kim K, Kim JH, Kim I, et al. Rev-erbα negatively regulates osteoclast and osteoblast differentiation through p38 MAPK Signaling Pathway. Mol Cells. 2020;43(1):34–47.

    CAS  PubMed  PubMed Central  Google Scholar 

  66. Tang Z, Xu T, Li Y, et al. Inhibition of CRY2 by STAT3/miRNA-7-5p promotes osteoblast differentiation through upregulation of CLOCK/BMAL1/P300 expression. Mol Ther Nucleic Acids. 2020;19:865–76.

    Article  CAS  PubMed  Google Scholar 

  67. Kim E-J, Kang I-H, Lee JW, et al. MiR-433 mediates ERRγ-suppressed osteoblast differentiation via direct targeting to Runx2 mRNA in C3H10T1/2 cells. Life Sci. 2013;92(10):562–8.

    Article  CAS  PubMed  Google Scholar 

  68. Smith SS, Dole NS, Franceschetti T, et al. MicroRNA-433 dampens glucocorticoid receptor signaling, impacting circadian rhythm and osteoblastic gene expression. J Biol Chem. 2016;291(41):21717–28.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Lian JB, Stein JL, Stein GS, et al. Runx2/Cbfa1 functions: diverse regulation of gene transcription by chromatin remodeling and co-regulatory protein interactions. Connect Tissue Res. 2003;44(Suppl 1):141–8.

    Article  CAS  PubMed  Google Scholar 

  70. Jensen ED, Schroeder TM, Bailey J, et al. Histone deacetylase 7 associates with Runx2 and represses its activity during osteoblast maturation in a deacetylation-independent manner. J Bone Miner Res Off J Am Soc Bone Miner Res. 2008;23(3):361–72.

    Article  CAS  Google Scholar 

  71. Tontonoz P, Hu E, Spiegelman BM. Stimulation of adipogenesis in fibroblasts by PPAR gamma 2, a lipid-activated transcription factor. Cell. 1994;79(7):1147–56.

    Article  CAS  PubMed  Google Scholar 

  72. Ren D, Collingwood TN, Rebar EJ, et al. PPARgamma knockdown by engineered transcription factors: exogenous PPARgamma2 but not PPARgamma1 reactivates adipogenesis. Genes Dev. 2002;16(1):27–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Lehrke M, Lazar MA. The many faces of PPARgamma. Cell. 2005;123(6):993–9.

    Article  CAS  PubMed  Google Scholar 

  74. Long H, Zhu Y, Lin Z, et al. miR-381 modulates human bone mesenchymal stromal cells (BMSCs) osteogenesis via suppressing Wnt signaling pathway during atrophic nonunion development. Cell Death Dis. 2019;10(7):470.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  75. Fan J, Li J, Fan Q. Naringin promotes differentiation of bone marrow stem cells into osteoblasts by upregulating the expression levels of microRNA-20a and downregulating the expression levels of PPARγ. Mol Med Rep. 2015;12(3):4759–65.

    Article  CAS  PubMed  Google Scholar 

  76. Riche DM, King ST. Bone loss and fracture risk associated with thiazolidinedione therapy. Pharmacotherapy. 2010;30(7):716–27.

    Article  CAS  PubMed  Google Scholar 

  77. Wang N, Yang G, Jia Z, et al. Vascular PPARgamma controls circadian variation in blood pressure and heart rate through Bmal1. Cell Metab. 2008;8(6):482–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Yang X, Downes M, Yu RT, et al. Nuclear receptor expression links the circadian clock to metabolism. Cell. 2006;126(4):801–10.

    Article  CAS  PubMed  Google Scholar 

  79. González-González A, García Nieto E, González A, et al. Melatonin modulation of radiation and chemotherapeutics-induced changes on differentiation of breast fibroblasts. Int J Mol Sci. 2019;20(16):3935.

    Article  PubMed Central  CAS  Google Scholar 

  80. Shabani A, Foroozanfard F, Kavossian E, et al. Effects of melatonin administration on mental health parameters, metabolic and genetic profiles in women with polycystic ovary syndrome: a randomized, double-blind, placebo-controlled trial. J Affect Disord. 2019;250:51–6.

    Article  CAS  PubMed  Google Scholar 

  81. Fontaine C, Dubois G, Duguay Y, et al. The orphan nuclear receptor Rev-Erbalpha is a peroxisome proliferator-activated receptor (PPAR) gamma target gene and promotes PPARgamma-induced adipocyte differentiation. J Biol Chem. 2003;278(39):37672–80.

    Article  CAS  PubMed  Google Scholar 

  82. Liu C, Li S, Liu T, et al. Transcriptional coactivator PGC-1alpha integrates the mammalian clock and energy metabolism. Nature. 2007;447(7143):477–81.

    Article  CAS  PubMed  Google Scholar 

  83. Le PT, Bornstein SA, Motyl KJ, et al. A novel mouse model overexpressing Nocturnin results in decreased fat mass in male mice. J Cell Physiol. 2019;234(11):20228–39.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Chaturvedi P, Tyagi SC. Epigenetic mechanisms underlying cardiac degeneration and regeneration. Int J Cardiol. 2014;173(1):1–11.

    Article  PubMed  PubMed Central  Google Scholar 

  85. Kojima S, Gatfield D, Esau CC, et al. MicroRNA-122 modulates the rhythmic expression profile of the circadian deadenylase Nocturnin in mouse liver. PLoS ONE. 2010;5(6): e11264.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  86. Estrella MA, Du J, Chen L, et al. The metabolites NADP and NADPH are the targets of the circadian protein Nocturnin (Curled). Nat Commun. 2019;10(1):2367.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  87. Grimaldi B, Bellet MM, Katada S, et al. PER2 controls lipid metabolism by direct regulation of PPARγ. Cell Metab. 2010;12(5):509–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Adamovich Y, Rousso-Noori L, Zwighaft Z, et al. Circadian clocks and feeding time regulate the oscillations and levels of hepatic triglycerides. Cell Metab. 2014;19(2):319–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Bae K, Jin X, Maywood ES, et al. Differential functions of mPer1, mPer2, and mPer3 in the SCN circadian clock. Neuron. 2001;30(2):525–36.

    Article  CAS  PubMed  Google Scholar 

  90. Shearman LP, Jin X, Lee C, et al. Targeted disruption of the mPer3 gene: subtle effects on circadian clock function. Mol Cell Biol. 2000;20(17):6269–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Costa MJ, So AYL, Kaasik K, et al. Circadian rhythm gene period 3 is an inhibitor of the adipocyte cell fate. J Biol Chem. 2011;286(11):9063–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Aggarwal A, Costa MJ, Rivero-Gutiérrez B, et al. The circadian clock regulates adipogenesis by a Per3 crosstalk pathway to Klf15. Cell Rep. 2017;21(9):2367–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Wang T, Wang Z, Yang P, et al. PER1 prevents excessive innate immune response during endotoxin-induced liver injury through regulation of macrophage recruitment in mice. Cell Death Dis. 2016;7: e2176.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Suzuki C, Ushijima K, Ando H, et al. Induction of by a histone deacetylase inhibitor is involved in amelioration of insulin sensitivity via adipocyte differentiation in ob/ob mice. Chronobiol Int. 2019;36(7):955–68.

    Article  CAS  PubMed  Google Scholar 

  95. Ushijima K, Suzuki C, Kitamura H, et al. Expression of clock gene in omental and mesenteric adipose tissue in patients with type 2 diabetes. BMJ Open Diabetes Res Care. 2020;8(1):e001465.

    Article  PubMed  PubMed Central  Google Scholar 

  96. Alagha MA, Vágó J, Katona É, et al. A synchronized circadian clock enhances early chondrogenesis. Cartilage. 2021;13(2_suppl):53S-67S.

    Article  CAS  PubMed  Google Scholar 

  97. Takarada T, Kodama A, Hotta S, et al. Clock genes influence gene expression in growth plate and endochondral ossification in mice. J Biol Chem. 2012;287(43):36081–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Yu S, Tang Q, Xie M, et al. Circadian BMAL1 regulates mandibular condyle development by hedgehog pathway. Cell Prolif. 2020;53(1): e12727.

    Article  PubMed  Google Scholar 

  99. Cha S, Lee S-M, Wang J, et al. Enhanced circadian clock in MSCs-based cytotherapy ameliorates age-related temporomandibular joint condyle degeneration. Int J Mol Sci. 2021;22(19):10632.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  100. Bi W, Deng JM, Zhang Z, et al. Sox9 is required for cartilage formation. Nat Genet. 1999;22(1):85–9.

    Article  CAS  PubMed  Google Scholar 

  101. Hardingham TE, Oldershaw RA, Tew SR. Cartilage, SOX9 and Notch signals in chondrogenesis. J Anat. 2006;209(4):469–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Dudek M, Gossan N, Yang N, et al. The chondrocyte clock gene Bmal1 controls cartilage homeostasis and integrity. J Clin Investig. 2016;126(1):365–76.

    Article  PubMed  Google Scholar 

  103. Le NQ, Binh NT, Takarada T, et al. Negative correlation between Per1 and Sox6 expression during chondrogenic differentiation in pre-chondrocytic ATDC5 cells. J Pharmacol Sci. 2013;122(4):318–25.

    Article  PubMed  CAS  Google Scholar 

  104. Lee IK, Song H, Kim H, et al. RORα regulates cholesterol metabolism of CD8 T cells for anticancer immunity. Cancers. 2020;12(7):1733.

    Article  CAS  PubMed Central  Google Scholar 

  105. Qu P, Wang L, Min Y, et al. Vav1 regulates mesenchymal stem cell differentiation decision between adipocyte and chondrocyte via sirt1. Stem Cells (Dayton, Ohio). 2016;34(7):1934–46.

    Article  CAS  Google Scholar 

  106. Dudakovic A, Samsonraj RM, Paradise CR, et al. Inhibition of the epigenetic suppressor EZH2 primes osteogenic differentiation mediated by BMP2. J Biol Chem. 2020;295(23):7877–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Liu F, Song D-Y, Huang J, et al. Long non-coding RNA CIR inhibits chondrogenic differentiation of mesenchymal stem cells by epigenetically suppressing ATOH8 via methyltransferase EZH2. Mol Med. 2021;27(1):12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Nomura Y, Hara ES, Yoshioka Y, et al. DNA methylation-based regulation of human bone marrow-derived mesenchymal stem/progenitor cell chondrogenic differentiation. Cells Tissues Organs. 2019;207(3–4):115–26.

    Article  CAS  PubMed  Google Scholar 

  109. Koike N, Yoo S-H, Huang H-C, et al. Transcriptional architecture and chromatin landscape of the core circadian clock in mammals. Science (New York, NY). 2012;338(6105):349–54.

    Article  CAS  Google Scholar 

  110. Hurley JM, Loros JJ, Dunlap JC. Circadian oscillators: around the transcription-translation feedback loop and on to output. Trends Biochem Sci. 2016;41(10):834–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Browne CJ, Godino A, Salery M, et al. Epigenetic mechanisms of opioid addiction. Biol Psychiatry. 2020;87(1):22–33.

    Article  CAS  PubMed  Google Scholar 

  112. Richards EJ, Elgin SCR. Epigenetic codes for heterochromatin formation and silencing: rounding up the usual suspects. Cell. 2002;108(4):489–500.

    Article  CAS  PubMed  Google Scholar 

  113. Fu G, Ren A, Qiu Y, et al. Epigenetic regulation of osteogenic differentiation of mesenchymal stem cells. Curr Stem Cell Res Ther. 2016;11(3):235–46.

    Article  CAS  PubMed  Google Scholar 

  114. Brown SA. Circadian clock-mediated control of stem cell division and differentiation: beyond night and day. Development (Cambridge, England). 2014;141(16):3105–11.

    Article  CAS  Google Scholar 

  115. Orozco-Solis R, Aguilar-Arnal L. Circadian regulation of immunity through epigenetic mechanisms. Front Cell Infect Microbiol. 2020;10:96.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Doi M, Hirayama J, Sassone-Corsi P. Circadian regulator CLOCK is a histone acetyltransferase. Cell. 2006;125(3):497–508.

    Article  CAS  PubMed  Google Scholar 

  117. Xu H, Gustafson CL, Sammons PJ, et al. Cryptochrome 1 regulates the circadian clock through dynamic interactions with the BMAL1 C terminus. Nat Struct Mol Biol. 2015;22(6):476–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Duong HA, Robles MS, Knutti D, et al. A molecular mechanism for circadian clock negative feedback. Science (New York, NY). 2011;332(6036):1436–9.

    Article  CAS  Google Scholar 

  119. Yin L, Lazar MA. The orphan nuclear receptor Rev-erbalpha recruits the N-CoR/histone deacetylase 3 corepressor to regulate the circadian Bmal1 gene. Mol Endocrinol. 2005;19(6):1452–9.

    Article  CAS  PubMed  Google Scholar 

  120. Vaissière A, Berger S, Harrus D, et al. Molecular mechanisms of transcriptional control by Rev-erbα: an energetic foundation for reconciling structure and binding with biological function. Protein Sci. 2015;24(7):1129–46.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  121. Dernie F, Adeyoju D. A matter of time: circadian clocks in osteoarthritis and the potential of chronotherapy. Exp Gerontol. 2021;143: 111163.

    Article  CAS  PubMed  Google Scholar 

  122. Louvet L, Leterme D, Delplace S, et al. Sirtuin 1 deficiency decreases bone mass and increases bone marrow adiposity in a mouse model of chronic energy deficiency. Bone. 2020;136: 115361.

    Article  CAS  PubMed  Google Scholar 

  123. Shakibaei M, Shayan P, Busch F, et al. Resveratrol mediated modulation of Sirt-1/Runx2 promotes osteogenic differentiation of mesenchymal stem cells: potential role of Runx2 deacetylation. PLoS ONE. 2012;7(4): e35712.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Soni SK, Basu P, Singaravel M, et al. Sirtuins and the circadian clock interplay in cardioprotection: focus on sirtuin 1. Cell Mol Life Sci CMLS. 2021;78(6):2503–15.

    Article  CAS  PubMed  Google Scholar 

  125. Nakahata Y, Kaluzova M, Grimaldi B, et al. The NAD+dependent deacetylase SIRT1 modulates CLOCK-mediated chromatin remodeling and circadian control. Cell. 2008;134(2):329–40.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Asher G, Gatfield D, Stratmann M, et al. SIRT1 regulates circadian clock gene expression through PER2 deacetylation. Cell. 2008;134(2):317–28.

    Article  CAS  PubMed  Google Scholar 

  127. Ramsey KM, Yoshino J, Brace CS, et al. Circadian clock feedback cycle through NAMPT-mediated NAD+ biosynthesis. Science (New York, NY). 2009;324(5927):651–4.

    Article  CAS  Google Scholar 

  128. Artsi H, Cohen-Kfir E, Gurt I, et al. The Sirtuin1 activator SRT3025 down-regulates sclerostin and rescues ovariectomy-induced bone loss and biomechanical deterioration in female mice. Endocrinology. 2014;155(9):3508–15.

    Article  PubMed  CAS  Google Scholar 

  129. Simic P, Zainabadi K, Bell E, et al. SIRT1 regulates differentiation of mesenchymal stem cells by deacetylating β-catenin. EMBO Mol Med. 2013;5(3):430–40.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Zainabadi K, Liu CJ, Guarente L. SIRT1 is a positive regulator of the master osteoblast transcription factor, RUNX2. PLoS ONE. 2017;12(5): e0178520.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  131. Bäckesjö C-M, Li Y, Lindgren U, et al. Activation of Sirt1 decreases adipocyte formation during osteoblast differentiation of mesenchymal stem cells. Cells Tissues Organs. 2009;189(1–4):93–7.

    Article  PubMed  CAS  Google Scholar 

  132. Zhou Y, Zhou Z, Zhang W, et al. SIRT1 inhibits adipogenesis and promotes myogenic differentiation in C3H10T1/2 pluripotent cells by regulating Wnt signaling. Cell Biosci. 2015;5:61.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  133. Zhong Y, Ye Q, Chen C, et al. Ezh2 promotes clock function and hematopoiesis independent of histone methyltransferase activity in zebrafish. Nucleic Acids Res. 2018;46(7):3382–99.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Yang R, Chen J, Zhang J, et al. 1,25-Dihydroxyvitamin D protects against age-related osteoporosis by a novel VDR-Ezh2-p16 signal axis. Aging Cell. 2020;19(2): e13095.

    Article  CAS  PubMed  Google Scholar 

  135. Yu Y, Deng P, Yu B, et al. Inhibition of EZH2 promotes human embryonic stem cell differentiation into mesoderm by reducing H3K27me3. Stem Cell Rep. 2017;9(3):752–61.

    Article  CAS  Google Scholar 

  136. Lim HW, Iwatani M, Hattori N, et al. Resistance to 5-aza-2’-deoxycytidine in genic regions compared to non-genic repetitive sequences. J Reprod Dev. 2010;56(1):86–93.

    Article  CAS  PubMed  Google Scholar 

  137. Cao Y, Yang H, Jin L, et al. Genome-wide DNA methylation analysis during osteogenic differentiation of human bone marrow mesenchymal stem cells. Stem Cells Int. 2018;2018:8238496.

    PubMed  PubMed Central  Google Scholar 

  138. Jones PA. Functions of DNA methylation: islands, start sites, gene bodies and beyond. Nat Rev Genet. 2012;13(7):484–92.

    Article  CAS  PubMed  Google Scholar 

  139. Maekawa F, Shimba S, Takumi S, et al. Diurnal expression of Dnmt3b mRNA in mouse liver is regulated by feeding and hepatic clockwork. Epigenetics. 2012;7(9):1046–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Agbaria S, Haim A, Fares F, et al. Epigenetic modification in 4T1 mouse breast cancer model by artificial light at night and melatonin—the role of DNA-methyltransferase. Chronobiol Int. 2019;36(5):629–43.

    Article  CAS  PubMed  Google Scholar 

  141. Feng Y-Z, Yu Y, Zhou Y-F, et al. A natural variant of miR397 mediates a feedback loop in circadian rhythm. Plant Physiol. 2020;182(1):204–14.

    Article  CAS  PubMed  Google Scholar 

  142. Mehta N, Cheng HYM. Micro-managing the circadian clock: the role of microRNAs in biological timekeeping. J Mol Biol. 2013;425(19):3609–24.

    Article  CAS  PubMed  Google Scholar 

  143. Torres M, Becquet D, Franc J-L, et al. Circadian processes in the RNA life cycle. Wiley Interdiscip Rev RNA. 2018;9(3): e1467.

    Article  PubMed  CAS  Google Scholar 

  144. Assmann TS, Cuevas-Sierra A, Salas-Pérez F, et al. Crosstalk between circulating microRNAs and chronotypical features in subjects with metabolic syndrome. Chronobiol Int. 2020;37(7):1048–58.

    Article  CAS  PubMed  Google Scholar 

  145. Heegaard NHH, Carlsen AL, Lilje B, et al. Diurnal variations of human circulating cell-free micro-RNA. PLoS ONE. 2016;11(8): e0160577.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  146. Wang Y, Lv K, Zhao M, et al. Analysis of miRNA expression profiles in the liver of mutant mice. PeerJ. 2019;7: e8119.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  147. Wang H, Fan Z, Zhao M, et al. Oscillating primary transcripts harbor miRNAs with circadian functions. Sci Rep. 2016;6:21598.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  148. Yan Y, Salazar TE, Dominguez JM, et al. Dicer expression exhibits a tissue-specific diurnal pattern that is lost during aging and in diabetes. PLoS ONE. 2013;8(11): e80029.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Hackl M, Heilmeier U, Weilner S, et al. Circulating microRNAs as novel biomarkers for bone diseases—complex signatures for multifactorial diseases? Mol Cell Endocrinol. 2016;432:83–95.

    Article  CAS  PubMed  Google Scholar 

  150. Na YJ, Sung JH, Lee SC, et al. Comprehensive analysis of microRNA-mRNA co-expression in circadian rhythm. Exp Mol Med. 2009;41(9):638–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Figueredo DDS, Barbosa MR, Gitaí DLG, et al. Predicted microRNAs for mammalian circadian rhythms. J Biol Rhythms. 2013;28(2):107–16.

    Article  CAS  Google Scholar 

  152. Zhang R, Lahens NF, Ballance HI, et al. A circadian gene expression atlas in mammals: implications for biology and medicine. Proc Natl Acad Sci USA. 2014;111(45):16219–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Fan Z, Zhao M, Joshi PD, et al. A class of circadian long non-coding RNAs mark enhancers modulating long-range circadian gene regulation. Nucleic Acids Res. 2017;45(10):5720–38.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Sun H, Peng G, Wu H, et al. Long non-coding RNA MEG3 is involved in osteogenic differentiation and bone diseases (Review). Biomed Rep. 2020;13(1):15–21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Yu J, Xiao M, Ren G. Long non-coding RNA XIST promotes osteoporosis by inhibiting the differentiation of bone marrow mesenchymal stem cell by sponging miR-29b-3p that suppresses nicotinamide N-methyltransferase. Bioengineered. 2021;12(1):6057–69.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Takahashi JS. Transcriptional architecture of the mammalian circadian clock. Nat Rev Genet. 2017;18(3):164–79.

    Article  CAS  PubMed  Google Scholar 

  157. Zhu B, Gates LA, Stashi E, et al. Coactivator-dependent oscillation of chromatin accessibility dictates circadian gene amplitude via REV-ERB loading. Mol Cell. 2015;60(5):769–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Du W, Guo D, Du W. ATP-dependent chromatin remodeling complex in the lineage specification of mesenchymal stem cells. Stem Cells Int. 2020;2020:8839703.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  159. Deshayes N, Genty G, Berthelot F, et al. Human long-term deregulated circadian rhythm alters regenerative properties of skin and hair precursor cells. Eur J Dermatol EJD. 2018;28(4):467–75.

    Article  CAS  PubMed  Google Scholar 

  160. Wu Q-J, Sun H, Wen Z-Y, et al. Shift work and health outcomes: an umbrella review of systematic reviews and meta-analyses of epidemiological studies. J Clin Sleep Med JCSM Off Publ Am Acad Sleep Med. 2022;18(2):653–62.

    Google Scholar 

  161. An K, Zhao H, Miao Y, et al. A circadian rhythm-gated subcortical pathway for nighttime-light-induced depressive-like behaviors in mice. Nat Neurosci. 2020;23(7):869–80.

    Article  CAS  PubMed  Google Scholar 

  162. Janich P, Toufighi K, Solanas G, et al. Human epidermal stem cell function is regulated by circadian oscillations. Cell Stem Cell. 2013;13(6):745–53. 

    Article  CAS  PubMed  Google Scholar 

  163. Ohdo S. Chrono-drug discovery and development based on circadian rhythm of molecular, cellular and organ level. Biol Pharm Bull. 2021;44(6):747–61.

    Article  CAS  PubMed  Google Scholar 

  164. Parasram K, Karpowicz P. Time after time: circadian clock regulation of intestinal stem cells. Cell Mol Life Sci CMLS. 2020;77(7):1267–88.

    Article  CAS  PubMed  Google Scholar 

  165. Mishra HK, Ying NM, Luis A, et al. Circadian rhythms in bipolar disorder patient-derived neurons predict lithium response: preliminary studies. Mol Psychiatry. 2021;26(7):3383–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  166. Murphy BA, Wagner AL, McGlynn OF, et al. Exercise influences circadian gene expression in equine skeletal muscle. Vet J. 2014;201(1):39–45.

    Article  CAS  PubMed  Google Scholar 

  167. Hatori M, Vollmers C, Zarrinpar A, et al. Time-restricted feeding without reducing caloric intake prevents metabolic diseases in mice fed a high-fat diet. Cell Metab. 2012;15(6):848–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  168. Noyan H, El-Mounayri O, Isserlin R, et al. Cardioprotective signature of short-term caloric restriction. PLoS ONE. 2015;10(6): e0130658.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  169. Wang M, Yu D, Zheng L, et al. Mechanical stress affects circadian rhythm in skeletal muscle (C2C12 Myoblasts) by reducing Per/Cry gene expression and increasing Bmal1 gene expression. Med Sci Monit Int Med J Exp Clin Res. 2021;27: e928359.

    CAS  Google Scholar 

  170. Hassan N, McCarville K, Morinaga K, et al. Titanium biomaterials with complex surfaces induced aberrant peripheral circadian rhythms in bone marrow mesenchymal stromal cells. PLoS ONE. 2017;12(8): e0183359.

    Article  PubMed  PubMed Central  Google Scholar 

  171. Reitz CJ, Alibhai FJ, Khatua TN, et al. SR9009 administered for one day after myocardial ischemia-reperfusion prevents heart failure in mice by targeting the cardiac inflammasome. Commun Biol. 2019;2:353.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  172. Zhang L, Zhang R, Tien C-L, et al. REV-ERBα ameliorates heart failure through transcription repression. JCI Insight. 2017;2(17):e95177.

    Article  PubMed Central  Google Scholar 

  173. Shimozaki K. REV-ERB agonist SR9009 regulates the proliferation and neurite outgrowth/suppression of cultured rat adult hippocampal neural stem/progenitor cells in a concentration-dependent manner. Cell Mol Neurobiol. 2021.

  174. Sahar S, Sassone-Corsi P. The epigenetic language of circadian clocks. Handb Exp Pharmacol. 2013;217:29–44.

    Article  CAS  Google Scholar 

  175. Reischl S, Kramer A. Kinases and phosphatases in the mammalian circadian clock. FEBS Lett. 2011;585(10):1393–9.

    Article  CAS  PubMed  Google Scholar 

  176. Kowalska E, Ripperger JA, Muheim C, et al. Distinct roles of DBHS family members in the circadian transcriptional feedback loop. Mol Cell Biol. 2012;32(22):4585–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

Figures were created with BioRender.com.

Funding

This work was supported by grants from the National Natural Science Foundation of China (Nos. 81870745, 82171003).

Author information

Authors and Affiliations

Authors

Contributions

WZG wrote the initial manuscript draft; RL and MLY revised the manuscript draft; LXZ and JWZ revised the manuscript to make it clearer; XYW and YQY reviewed and corrected grammar mistakes; QZ edited and approved the final manuscript for publication. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Qing Zhao.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing financial or professional interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1

. Predicted clock-targeted miRNAs in mice and human which also working in the MSC differentiation

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Gao, W., Li, R., Ye, M. et al. The circadian clock has roles in mesenchymal stem cell fate decision. Stem Cell Res Ther 13, 200 (2022). https://doi.org/10.1186/s13287-022-02878-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13287-022-02878-0

Keywords