Skip to main content
Sage Choice logoLink to Sage Choice
. 2023 Jun 26;38(5):419–446. doi: 10.1177/07487304231178950

Circadian Regulation of the Neuroimmune Environment Across the Lifespan: From Brain Development to Aging

Ruizhuo Chen *,1, Brandy N Routh *,†,1, Andrew D Gaudet †,‡,§, Laura K Fonken *,†,2
PMCID: PMC10475217  NIHMSID: NIHMS1901348  PMID: 37357738

Abstract

Circadian clocks confer 24-h periodicity to biological systems, to ultimately maximize energy efficiency and promote survival in a world with regular environmental light cycles. In mammals, circadian rhythms regulate myriad physiological functions, including the immune, endocrine, and central nervous systems. Within the central nervous system, specialized glial cells such as astrocytes and microglia survey and maintain the neuroimmune environment. The contributions of these neuroimmune cells to both homeostatic and pathogenic demands vary greatly across the day. Moreover, the function of these cells changes across the lifespan. In this review, we discuss circadian regulation of the neuroimmune environment across the lifespan, with a focus on microglia and astrocytes. Circadian rhythms emerge in early life concurrent with neuroimmune sculpting of brain circuits and wane late in life alongside increasing immunosenescence and neurodegeneration. Importantly, circadian dysregulation can alter immune function, which may contribute to susceptibility to neurodevelopmental and neurodegenerative diseases. In this review, we highlight circadian neuroimmune interactions across the lifespan and share evidence that circadian dysregulation within the neuroimmune system may be a critical component in human neurodevelopmental and neurodegenerative diseases.

Keywords: circadian, neuroimmune, microglia, astrocyte, development, aging


The circadian system is an evolutionarily adaptive system that produces rhythmic anticipatory signals in response to the 24-h periodicity on Earth (Patke et al., 2020). These rhythmic anticipatory signals decide the “whens” in life: When do we stay alert? When do we need peak motor function? When do we have time and energy for digestion? When can we rest and clear debris from our systems? Importantly, activities that are regulated by the circadian system result in time-of-day differences in encountering pathogens and injury (Costantini et al., 2020). Thus, it follows that the circadian system also regulates the immune system. Immune processes are energetically expensive, involving the recruitment of specialized cells and production of inflammatory molecules to combat pathogens and clear up dead or damaged cells (Ganeshan and Chawla, 2014). Thus, circadian regulation of immune function can facilitate resource allocation and lead to optimized functioning (Scheiermann et al., 2013).

This review will focus on the circadian regulation of neuroimmune function. The immune system and the central nervous system (CNS) have tightly controlled bi-directional communication (Dantzer et al., 2008). Specialized neuroimmune cells, such as microglia and astrocytes, monitor and maintain the CNS environment throughout the lifespan (Ortinski et al., 2022). Neuroimmune cells not only protect the CNS from pathology but also participate in homeostatic processes which facilitate neuronal functioning and regulate behavior (Reemst et al., 2016). Neuroimmune cells are regulated by the circadian system (Fonken et al., 2015; Takayama et al., 2016), and disruption of circadian rhythmicity can lead to dysfunction in these cells (Erblich et al., 2011; Fonken et al., 2016; Inokawa et al., 2020; Takayama et al., 2017). In this review, we will discuss circadian regulation of the neuroimmune system across the lifespan, from the emergence of circadian rhythms in childhood to the dampening of circadian rhythms late in life. Neuroimmune and circadian dysfunction are implicated in multiple neurodevelopmental disorders (NDDs) and neurodegenerative diseases. Thus, exploring the intersection of circadian rhythms and the neuroimmune system could lead to new mechanistic and therapeutic insights.

The Circadian System Regulates Rhythmic Daily Activities to Maintain Homeostasis

In mammals, daily rhythms are controlled by the primary circadian oscillator located in the suprachiasmatic nucleus (SCN) of the hypothalamus. The SCN responds to external cues: light is a potent zeitgeber that entrains or “fine-tunes” the SCN to the diurnal environment. Light information is detected by a specialized population of melanopsin-containing intrinsically photosensitive retinal ganglion cells (ipRGCs). ipRGCs project to the SCN via a monosynaptic connection through the retinohypothalamic tract (Berson et al., 2002; Hattar et al., 2002). The SCN receives this direct photic input from the retina as well as indirect input from the intergeniculate leaflet (IGL) and the brainstem (Reghunandanan and Reghunandanan, 2006). In the absence of external light cues, SCN neurons maintain a synchronized, endogenous, free-running rhythm of approximately 24 h (Armstrong et al., 1986; Vriend and Reiter, 2015). Rhythmic SCN output is communicated to extra-SCN (or peripheral) molecular clocks that exist in nearly every cell via neural and humoral signals. Extra-SCN clocks have self-sustained circadian oscillations; however, their circadian phase is synchronized to the SCN by efferent information from the SCN including endocrine and autonomic signals, body temperature, exercise, and food intake (Astiz et al., 2019; Brown et al., 2002; Tahara et al., 2017). For example, the SCN drives the rhythmic secretion of hormones from the hypothalamus and pituitary. In turn, endocrine molecules such as melatonin, serotonin (5-hydroxytryptamine, 5-HT), and glucocorticoids act via cellular receptors to feedback on expression of genes regulating circadian clocks (reviewed by Vriend and Reiter, 2015) (Mistlberger et al., 2000; Vriend and Reiter, 2015; Woodruff et al., 2016). This endocrine feedback is critical as disruptions in these hormone rhythms can desynchronize circadian rhythms (Carroll et al., 2007; Daut and Fonken, 2019).

At the molecular level, 10%-40% of all protein-coding genes in rodents and more than 50% in non-human primates and human tissues show 24-h rhythmic oscillations in a highly tissue-specific manner (Baldi et al., 2021; Li et al., 2022; Mure et al., 2018; Ruben et al., 2018; Storch et al., 2002; Zhang et al., 2014). Oscillating genes are regulated by a feedback inhibition loop of “core-clock” genes and their proteins (summarized in Figure 1). The core-clock transcription factors BMAL1 and CLOCK dimerize and form the positive arm of this loop. The CLOCK-BMAL1 dimer binds to a DNA cis-element E-box to promote the transcription of various target genes, including Per and Cry. Once PER and CRY proteins accumulate in the cytosol to sufficient levels, they are transported into the nucleus to inhibit the binding of the BMAL1-CLOCK dimer to the E-box. This feedback loop is necessary for circadian oscillation (Bae et al., 2001; Ueda et al., 2005; Yoshitane et al., 2019). In addition, BMAL1-CLOCK activity is regulated by REV-ERB/ROR-binding element (RRE). ROR and REV-ERBα/β (a.k.a. Nr1d1/2) compete for the response element to activate or repress Bmal1 transcription (Guillaumond et al., 2005; Yin and Lazar, 2005) and thus form an additional stabilizing loop within the clock. REV-ERBs also adjust rhythms to exogenous factors (Jolley et al., 2014). The core-clock loop does not solely define cellular timing—this circadian transcriptional loop also controls the transcription of clock-controlled genes that regulate rhythmic biological and physiological activities.

Figure 1.

Figure 1.

Circadian-related feedback inhibition loops control gene oscillation in mammals. The transcription factors BMAL1 and CLOCK dimerize and form the positive arm of this loop. The dimer binds to a DNA cis-element E-box to promote transcription of various target genes, including per and cry, whose protein products inhibit the binding of the BMAL1-CLOCK dimer. In addition, BMAL1-CLOCK activity is regulated by REV-ERB/ROR-binding element (RRE). REV-ERBα/β represses the transcription of Bmal1 and in turn stabilizes rhythms (Yin and Lazar, 2005), and adjusts rhythms to exogenous factors (Jolley et al., 2014). The “core-clock” loop controls the production of clock-controlled genes that are involved in rhythmic biological and physiological activities. Abbreviation: RRE = REV-ERB/ROR-binding element.

Circadian rhythmicity helps maintain physiological homeostasis. The circadian system regulates the sleep-wake cycle, metabolic and cardiovascular processes, hormone rhythms, and immune function (Beersma and Gordijn, 2007; Edgar et al., 1993; Scheiermann et al., 2013; Vriend and Reiter, 2015; Zhang et al., 2020). Importantly, with the advent of modern electricity, humans are no longer constrained to natural light cycles. Disruption of the circadian system due to artificial light at night, shift work, and long-distance air travel increases the risk of metabolic dysfunction (Cheng et al., 2021), cardiovascular disease (Knutsson et al., 1999), poorer mental health (Torquati et al., 2019), and cancer (Straif et al., 2007). Importantly, many of these issues are likely associated with impaired immune responses in shift workers (Ruiz et al., 2020).

In summary, the SCN is the primary circadian oscillator of the circadian system. The SCN is entrained by external zeitgeber cues and synchronizes the extra-SCN molecular clocks. Together these body clocks maintain homeostasis of various daily rhythms including physiological activities such as immune responses.

Circadian Regulation of Immune Function

The circadian system potently regulates immune function. Halberg and colleagues (1960) published one of the first studies to suggest this association, revealing that an Escherichia coli lipopolysaccharide (LPS) challenge during a mouse’s inactive phase induces substantially higher (80%) mortality than injection during the active phase (20%) (Halberg et al., 1960). This notable time-of-day difference in sepsis-induced mortality has been confirmed in subsequent work (Curtis et al., 2015; Deng et al., 2018; Feigin et al., 1969; Hrushesky et al., 1994; Lang et al., 2021); this association is circadian in nature (Lang et al., 2021).

The circadian system modulates sickness behaviors and susceptibility to infection (Scheiermann et al., 2018; Wang et al., 2022). For example, there is diurnal regulation of the sickness response in both humans and rodents. In humans, peak presentation of symptoms in response to viral infection and allergic asthma tends to occur during the late evening to early morning (Smolensky et al., 1995). Rodents exhibit time-of-day differences in susceptibility to infection. Rats and hamsters administered LPS during the inactive compared with the active phase exhibit prolonged and exacerbated sickness behaviors (Fonken et al., 2015; Franklin et al., 2003; Prendergast et al., 2015). A similar time-of-day regulation exists in rodents following viral infection/mimics: infection during the inactive phase leads to higher cytokine production and mortality (Edgar et al., 2016; Ehlers et al., 2018; Gagnidze et al., 2016).

These time-of-day-dependent changes in behavior and mortality are likely due to circadian regulation of immune cell activation and mobilization. The circadian system regulates both innate and adaptive immune responses (Scheiermann et al., 2018). For example, innate immune cells such as macrophages display heightened pro-inflammatory response during the inactive phase when presentation of sickness behaviors is highest. Direct transcriptional regulation of TLR4 signaling by the circadian clock underlies this circadian inflammatory response in macrophages, regulating the release of the pro-inflammatory cytokine tumor necrosis factor alpha (TNF-α; Keller et al., 2009). Moreover, the circadian clock gene, Bmal1, regulates nuclear factor kappa B (NFκB) transcription in macrophages (Curtis et al., 2015). Time-of-day variation in immune responses also occurs with adaptive immune cells such as T cells (Bollinger et al., 2011) and B cells (Silver et al., 2012). The circadian clock regulates inflammatory responses within these cells as well as their migration to different tissues throughout the day (Druzd et al., 2017; Ince et al., 2019). Diurnal regulation of adaptive immune cells has crucial implications for human health and disease, as revealed in recent studies of vaccination timing: In subjects receiving a vaccine against influenza or severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), vaccines administered in the morning are more effective than those given in the afternoon (Hazan et al., 2023; Ince et al., 2023; Long et al., 2016; Zhang et al., 2021). Although a full discussion of circadian regulation of the innate and adaptive immune system is beyond the scope of this review, several excellent recent reviews have addressed this topic (Curtis et al., 2014; Scheiermann et al., 2018; Westwood et al., 2019).

The Neuroimmune System: Development and Function of Specialized Glial Cells Across the Lifespan

The CNS is immunologically unique. The neuroimmune system is tailored to maintain the health and homeostasis of the CNS, which is sensitive to immune stressors. Immune signals are both initiated in the brain (e.g., following a brain injury) and communicated to the brain during peripheral immune activation (e.g., following infection or injury). To protect relatively vulnerable CNS tissue from pathogens and potentially overzealous peripheral immune responses, the brain has a physical blood-brain barrier (BBB) and is considered immune privileged. However, immune privilege is relative and dynamic. Multiple innate and adaptive immune cells circulate in spaces surrounding the parenchyma (e.g., the choroid plexus, meningeal lymphatic vasculature, perivascular spaces, and the skull and bone marrow) and relay immune information to the brain (Rustenhoven and Kipnis, 2022). For example, the choroid plexus—located within the cerebral ventricles—houses a large population of antigen-presenting resident macrophages and dendritic cells and is a gateway for CD4+ T-cell trafficking into the cerebrospinal fluid (CSF) and brain parenchyma (Dani et al., 2021; Meeker et al., 2012). Moreover, recent work highlights that the skull and vertebral bone marrow are immune cell reservoirs for the CNS (Cugurra et al., 2021). Importantly, the CNS also contains specialized resident immune cells, which support neuronal functions and regulate inflammatory reactions in the brain. This review will focus primarily on 2 glial cells: microglia and astrocytes. Neuroinflammatory responses are also regulated by additional cell types (e.g., endothelial cells, pericytes, macrophages) that are beyond the scope of this review (see review by Villabona-Rueda et al., 2019).

Microglia Functions

Microglia are the resident innate immune cells of the CNS. With their unique highly motile processes, microglia actively survey their territory and recognize signals from “healthy” neurons to maintain a homeostatic state. Under homeostatic conditions, adult microglia actively support and sculpt their environment by engaging in phagocytosis and secreting factors including cytokines and neurotrophic factors. Homeostatic microglia are responsible for synaptic pruning, a process that is crucial for proper brain morphogenesis and learning. Microglia dysfunctions impair activity-dependent synaptic pruning (Hoshiko et al., 2012; Schafer et al., 2012). Other functions of homeostatic microglia include supporting myelination, neurogenesis, and regulating BBB permeability.

Microglia derive from erythro-myeloid progenitors (EMPs) that are generated in the yolk sac blood islands at E8.5 in mice (Ginhoux et al., 2010). Microglia proliferate and gradually colonize the brain perinatally, with distinct morphologies depending on spatiotemporal localization. In early brain development, microglia are crucial for vasculature development and neuronal circuit assembly (Casano and Peri, 2015; Mosser et al., 2017; Perdiguero and Geissmann, 2016; Reemst et al., 2016; Thion and Garel, 2017), participating in cell death, axonal tract formation, and morphogenesis in the brain (Butovsky et al., 2014; Hattori, 2022; Thion and Garel, 2017). Embryonic microglia depletion leads to higher neural progenitor cell (NPC) proliferation and neuronal oligodendrocyte differentiation in the hypothalamus (Marsters et al., 2020). Microglia also regulate the balance of neuronal connections within specific circuits. For example, microglia accumulate at dopaminergic axon terminals at E14.5 and limit the outgrowth of dopaminergic neurons to the striatum (Squarzoni et al., 2014). Embryonic depletion of microglia leads to an imbalance in dopaminergic neuron innervation and altered neocortical inhibitory interneuron positioning, which may influence postnatal synaptogenesis (Paolicelli et al., 2011; Squarzoni et al., 2014). Taken together, before eventually reaching adult-like patterns by the end of the second postnatal week in mice (Bennett et al., 2016), a critical function of microglia in the developing brain is to support proper wiring of brain circuits.

In the adult CNS, a major function of microglia is to maintain the integrity of the nervous system. Following immune system activation, microglia adapt altered molecular, morphological, and functional properties to initiate an immune response and repair the damaged tissue or eliminate the pathogen. Importantly, the responses of microglia to stimuli vary between developing and aged brains, with microglia taking on a more pro-inflammatory phenotype later in life (Bisht et al., 2016; Fonken and Gaudet, 2022).

Astrocyte Functions

Astrocytes are the most numerous cell type in the brain and critically regulate the brain environment. Astrocytic end-feet tightly ensheathe the vasculature and help maintain the BBB. Specifically, Aquaporin-4 (AQP4) channels on astrocyte end-feet regulate the glymphatic system to remove toxic metabolic byproducts from the brain parenchyma (Mestre et al., 2018). Astrocytes support neurotransmission by gliotransmitter release, storing and releasing glucose, regulating extracellular ion concentration (Harada et al., 2015; Walz, 2000), and recycling glutamate at the synaptic cleft (Schousboe and Waagepetersen, 2005). In response to high neural activity, astrocytes can increase oxygen supply by regulating neurovascular coupling (Iadecola, 2017; Otsu et al., 2015) and secrete glutathione to promote neuronal redox homeostasis and prevent excitotoxicity (Bolanos, 2016). Like microglia, astrocytes also mediate synapse maturation (Allen et al., 2012) and elimination (Chung et al., 2013; Vainchtein et al., 2018). Moreover, astrocytes coordinate communication between neurons and oligodendrocytes; for instance, in response to adenosine triphosphate (ATP) from neuronal activity, astrocytes release cytokines that trigger myelination by oligodendrocytes (Ishibashi et al., 2006).

During CNS injuries astrocytes respond quickly with a process called “astrogliosis,” with altered morphology, gene expression, and function to limit cascading neural damage (Sofroniew, 2014; Zamanian et al., 2012). For example, reactive astrocytes form glial scars that improve tissue repair (Anderson et al., 2016). In addition, although microglia are the predominant source of inflammatory cytokines in the brain, reactive astrocytes also produce inflammatory cytokines in response to immune challenges (Leone et al., 2006).

Historically, a neuro-centric view of the CNS has focused on neurons as the key components in regulating animal behavior. However, microglia and astrocytes sculpt and regulate the lifelong plasticity of the CNS, demonstrating glial cells are not merely passive supporters of neurological function but rather active influencers of neurophysiology and behavior. Improper function of these neuroimmune cells can affect behavior: for example, embryonic microglia depletion using CSF1R inhibitor PLX5622 induces hyperactivity and alters anxiety-like behaviors (Rosin et al., 2018). Microglia and astrocyte dysfunction can influence cognitive function (Bissonette et al., 2010; Politis et al., 2011; Wadhwa et al., 2017; Zhang et al., 2019), circadian rhythms (Brancaccio et al., 2017; Sominsky et al., 2021), and reward-related behaviors (Adeluyi et al., 2019; Bull et al., 2014; Northcutt et al., 2015) in adult animals (reviewed by Ortinski et al., 2022). Furthermore, neuroimmune dysregulation during vulnerable periods may have long-term effects. For example, early-life immune activation and immune activation in late aging are associated with long-lasting behavioral consequences.

Circadian Regulation of Microglia and Astrocytes

Like the broader immune system, the neuroimmune system is regulated by the circadian system (summarized in Figure 2). Microglia display diurnal differences in their morphology and function. For example, during the inactive phase, microglia are less “ramified” with thicker processes: a morphology associated with an elevated pro-inflammatory response (Barahona et al., 2022; Takayama et al., 2016). Microglia cytokine production also peaks during the inactive phase both at baseline (Taishi et al., 1997) and in response to exogenous stimuli such as LPS (Fonken et al., 2015). Moreover, microglia exhibit time-of-day variations in phagocytosis (Griffin et al., 2020). Purinergic receptor P2Y12 in the CNS is expressed exclusively by microglia and mediates microglial recognition of cellular adenosine diphosphate (ADP)/ATP, which is released during the normal activity of neurons, astrocytes, and oligodendroglia, or in response to tissue damage (Lin et al., 2020). The microglial molecular clock regulates P2Y12 receptor via lysosomal cysteine protease cathepsin S and modulates spine density and synaptic strength in a time-of-day manner, which is important for memory consolidation (Hayashi et al., 2013; Nakanishi et al., 2021). Disruption of the microglial clock impairs cathepsin S activity and, in turn, disturbs sleep and sociability (Hayashi et al., 2013; Takayama et al., 2017). Furthermore, depleting microglia using a diphtheria toxin-receptor-knock-in rat model disrupts daily oscillations in temperature, activity, and energy metabolism (Sominsky et al., 2021). These time-of-day variations in microglial activity persist under 24-h constant darkness (Hayashi et al., 2013), suggesting a circadian pattern independent of light cues.

Figure 2.

Figure 2.

Simplified representation of some of the circadian features of neuroimmune cells. Two major neuroimmune cells of the CNS, microglia and astrocytes, display diurnal rhythmicity in morphology and functions that are critical for maintaining diurnal homeostasis of the CNS. The peak timings of microglial and astrocytic key functions are highlighted on the 24-h clock. Abbreviations: CNS = central nervous system; GFAP = glial fibrillary acidic protein; BBB = blood-brain barrier.

At the molecular level, adult microglia exhibit rhythmic expression of clock genes, including Per1, Per2, Rev-erbα, and Bmal1 (Fonken et al., 2015; Hayashi et al., 2013). The nuclear receptor REV-ERBα is one factor that acts as a hub between the circadian and immune systems (Amir et al., 2018; Chang et al., 2019; Gibbs et al., 2012; Pariollaud et al., 2018; Yu et al., 2013). REV-ERBα regulates the immune system through binding to response elements in promoter regions of pro-inflammatory genes such as nucleotide-binding oligomerization domain (NOD), leucine rich repeat (LRR), and pyrin domain-containing protein 3 (Nlrp3) and interleukin 1β (Il1β) (Pourcet et al., 2018). The binding of REV-ERBα then inactivates NF-κB signaling (interleukin [IL]-1β, IL-18, and TNF-α) and represses the expression and activation of the NLRP3 inflammasome (Duez and Pourcet, 2021; Griffin et al., 2019; Pourcet et al., 2018). Therefore, daily REV-ERBα abundance regulates the timing of NLRP3 expression and inflammatory cytokines production (Pourcet et al., 2018). Importantly, the NLRP3 inflammasome links the complement system and IL-1β production (Laudisi et al., 2013). The complement system, as a central part of innate immunity, consists of protein complexes that opsonize pathogens and promote inflammation and phagocytosis. REV-ERBα expression is positively controlled by BMAL1 and, in turn, negatively regulates Bmal1 transcription (Figure 1). This BMAL1-REV-ERBα axis serves as a regulator of complement expression and synaptic phagocytosis in the brain (Griffin et al., 2020). A synthetic REV-ERBα agonist dampens microglial circadian responses and reduces phagocytosis (Wolff et al., 2020). Conversely, Rev-erbα deletion increases microglial reactivity and induces enhanced but arrhythmic phagocytosis (Griffin et al., 2020). Modulating BMAL1 similarly targets neuroimmune cells: microglia-specific Bmal1 knockdown reduces inflammation and increases phagocytosis, both of which may elicit synapse loss (Curtis et al., 2015; Griffin et al., 2020; Wang et al., 2020).

Interestingly, microglia cell state of reactivity also influences circadian rhythms in these cells: polarizing microglia toward a pro-inflammatory phenotype (by co-treatment with LPS and interferon [IFN]-γ) results in a blunted and shortened microglial PER2::LUC rhythm (Honzlova et al., 2023). Conversely, anti-inflammatory polarization via IL-4 treatment enhances microglial rhythmicity and, more importantly, increases the amplitude of PER2::LUC rhythm in the SCN. A similar phenotype has been previously reported in bone marrow–derived macrophages: PER2::LUC oscillation in macrophages can be suppressed by pro-inflammatory polarization and enhanced by anti-inflammatory polarization (Chen et al., 2020). This novel finding supports that microglia polarization may regulate rhythmicity at both cell and tissue levels.

Although a growing body of literature suggests that circadian machinery in microglia may contribute to their physiological functions and some related behaviors, whether microglia prune all synapses in a similar Bmal1- and Rev-erbα-dependent manner, and whether microglia synaptic pruning is dependent on sleep, remains unclear.

Similar to microglia, astrocytes display circadian patterns. Astrocytes first appear in the SCN at E20 in rats (Munekawa et al., 2000). Astrocytes isolated from neonatal rat brains at P1-P2 show rhythmic expression of circadian clock genes Bmal1 and Rev-erbα (Carver et al., 2014), which continues in adulthood (Wen et al., 2020). Astrocytes also display diurnal differences in functions and activities. Glial fibrillary acidic protein (GFAP) is a classic marker of astrocyte reactivity and is involved in many important astrocytic functions in the CNS (Tykhomyrov et al., 2016); during the inactive phase, astrocytes have intensified GFAP expression indicating higher inactive phase astrocytic reactivity (Leone et al., 2006; Santos et al., 2005). During the active phase in the SCN, astrocytes exhibit a higher level of intracellular calcium, which is antiphase to that in SCN neurons (Hastings et al., 2019; Noguchi et al., 2017).

Astrocyte rhythms may contribute to neuronal and behavioral rhythms. The circadian oscillation in astrocytic intracellular calcium is phase-locked to extracellular glutamate levels, through which astrocytes suppress the activity of SCN glutamatergic neurons (Brancaccio et al., 2017). Similarly, ex vivo cortical astrocytes entrain their daily rhythms to co-cultured cortical neurons in a Bmal1-dependent manner (Barca-Mayo et al., 2017; also see review by Hastings et al., 2023). SCN astrocytes are sufficient to establish circadian rhythms in locomotion, although their rhythms are not required. Loss of SCN astrocytic clock by conditional Bmal1 knockout lengthens, but does not abolish, circadian rhythms of locomotor activity. On the other hand, the shortened circadian period of locomotor activity in CK1ε-tau mutant mice is restored by increasing the period of SCN astrocytes via astrocyte-specific removal of the tau mutation (Tso et al., 2017). Furthermore, rescuing Cry1 in SCN astrocytes restores circadian rhythms in locomotion in Cry1/2 knockout mice (Brancaccio et al., 2019), suggesting that circadian rhythms in SCN astrocytes contribute to SCN function during the active phase. In addition to its effect on circadian rhythms of locomotor activity, studies using conditional Bmal1 deletion suggest that astrocytic rhythms regulate many physiological functions, such as glucose metabolism (Barca-Mayo et al., 2020) and cognition (Barca-Mayo et al., 2017).

Given these striking recent results highlighting the role of astrocytes in the circadian system, several knowledge gaps remain to be filled. For example, future studies should explore the extent that astrocytes “control” SCN neurons and whether depleting astrocytes would diminish (or reduce) the ability of SCN neurons to generate daily rhythms. In addition, the activity of glutamatergic neurons is known to be regulated by astrocytes, but a more thorough investigation on the type(s) of SCN neurons whose activities are regulated by astrocytes, as well as the pathways of the astrocytic regulation of SCN neuronal activities, will further our understanding of how neuroimmune cells like astrocytes contribute to daily rhythms in physiology and behavior.

The circadian regulation of microglia and astrocytes during early development remains understudied; however, the evidence presented below (see section “Circadian rhythms influence neuroimmune wiring of neuronal circuitry”) suggests that circadian dysregulation may interrupt critical functions of glia in the developing brain. Future studies should explore the emergence of diurnal changes in microglia and astrocytes during early development. This information could help illuminate mechanisms underlying how glial clocks influence the developing brain, and whether circadian-neuroimmune dysfunction contributes to NDDs.

Circadian Rhythms Emerge Early in Life, Influence Neuroimmune Modeling of Brain Circuits, and Affect Neurodevelopmental Outcomes

During prenatal and early postnatal periods, a massive number of cells are generated and form the brain and the spinal cord. Morphogenesis and accurate brain circuit wiring during these critical periods are essential for proper brain function throughout the lifespan. Glial cells support brain development and support developing brain circuits and structures. However, most of the evidence demonstrating circadian regulation in neuroimmune components comes from studies using adult animals, and few studies have explored how circadian mechanisms influence microglial and astrocytic roles in early development. Indeed, the strengthening of rhythmicity during the first postnatal weeks in rodents coincides with the maturation of brain circuits. This section of the review will discuss the emergence of circadian rhythms early in life, and how these rhythms may govern the neuroimmune sculpting of the CNS to influence animal behaviors into adulthood. We further discuss evidence implicating circadian neuroimmune dysregulation in the pathophysiology of NDDs, including autism spectrum disorder (ASD), bipolar disorder (BD), and schizophrenia (Figure 3).

Figure 3.

Figure 3.

Summary of circadian regulation of the neuroimmune environment in brain development and aging. Within the central nervous system, specialized glial cells such as microglia and astrocytes surveil and maintain the neuroimmune environment, and their activities vary in a time-of-day manner. Circadian dysregulation can alter immune function, thereby contributing to susceptibility to neurodevelopmental and neurodegenerative diseases. Abbreviations: SCN = suprachiasmatic nucleus; BBB = blood-brain barrier.

The function of microglia and astrocytes changes across the lifespan. Microglia and astrocytes support brain development. Circadian neuroimmune disruptions that disturb microglial and astrocytic clocks in utero or postnatally alter neuroimmune responses with long-lasting behavioral repercussions. This may play a role in the pathophysiology of neurodevelopmental disorders, including autism spectrum disorders, bipolar disorder, and schizophrenia. With brain aging, circadian function gradually declines, and neuroimmune function gradually shifts toward higher levels of basal inflammation. These processes are associated with progressively weakening circadian rhythms in microglia and astrocytes. These age-related changes in the circadian neuroimmune environment may predispose certain individuals to develop neurodegenerative disorders, including Alzheimer’s disease and Parkinson’s disease.

The Developmental Emergence of Circadian Rhythms

In rodents, neurogenesis in the SCN region of the embryonic hypothalamus happens between E10 and E18 (reviewed by Landgraf et al., 2014; Carmona-Alcocer et al., 2020). Development of SCN neurons displays a spatiotemporal order: a mid-SCN core forms earlier than the surrounding SCN-shell with a ventrolateral to dorsomedial gradient (Altman and Bayer, 1978, 1986; Davis et al., 1990; Kabrita and Davis, 2008). Following SCN neuronal development, by E17, mRNA expression of clock components are detectable in the mouse SCN region (Shearman et al., 1997; Shimomura et al., 2001); however, daily oscillation of these clock genes emerges only after birth (Ansari et al., 2009; Greiner et al., 2022; Li and Davis, 2005; Sladek et al., 2004). Likewise, rhythmic clock gene expression in developing extra-SCN tissue such as the heart (Sakamoto et al., 2002) and the liver (Sladek et al., 2007) is detected gradually after birth (except for transcriptional repressor gene Rev-erbα that shows rhythmic activity in embryonic liver at E20; Sladek et al., 2007). In contrast to transcriptomic analyses, rhythmic expression is observed in ex vivo embryonic SCN cells from clock gene-driven bioluminescent promoter PER::LUC as early as E12 (Saxena et al., 2007). A recent study reports self-sustained, synchronized PER2::LUC rhythm in SCN cells from E15.5 mice (Carmona-Alcocer et al., 2018). Together these and other studies suggest that the fetal SCN expresses rhythmic oscillations of metabolic and electrical activities (reviewed in Carmona-Alcocer et al., 2020). In ex vivo embryonic liver, heart, and kidney cells, rhythms have been observed by E18 (Dolatshad et al., 2010; Umemura et al., 2017). This may suggest the existence of rhythms at the cellular level in embryonic tissue and a lack of synchronized rhythms at the whole tissue level.

In adult animals, the SCN is entrained to the surrounding light-dark environment by direct photic input via melanopsin-containing ipRGCs. ipRGC neurogenesis occurs between E11 and E14 in mice (McNeill et al., 2011), and melanopsin expression is detectable in the retinal ganglion cell layer by E15 (Fahrenkrug et al., 2004; McNeill et al., 2011). By E17, ipRGC axons reach the optic chiasma near the ventral SCN and gradually innervate the SCN during the first postnatal week. Although ipRGCs are present prenatally, they are not functional until postnatal day 7 (P7) (McNeill et al., 2011). Thus, during early development, circadian rhythms in rodent pups likely depend primarily on maternal cues instead of direct photic input (Bates and Herzog, 2020; Takahashi and Deguchi, 1983). For example, in the 1980s, Reppert and Schwartz first showed that the embryonic SCN has a time-of-day difference in glucose utilization. This daily variation is influenced by the circadian phase in maternal SCN (Davis and Gorski, 1985; Reppert et al., 1984; Reppert and Schwartz, 1983, 1986). Maternal signals such as dopamine and melatonin directly trigger embryonic SCN metabolic rhythms (Davis and Mannion, 1988; Weaver and Reppert, 1995). Interestingly, recent work indicates that these maternal signals are sufficient, but may not be necessary for entrainment (see review by Bates and Herzog, 2020); this suggests that the maternal clock may provide a parallel pathway for the fetus to perceive or reinforce phase of daily timing. Likewise, recent studies indicate that maternal metabolic signals directly elicit growth factor signaling, whereas maternal activity/feeding rhythms entrain neuronal activities in the embryonic SCN, suggested by enrichment analysis of rhythmically expressed genes (Greiner et al., 2022). The influence of maternal activity/feeding rhythms on pup SCN activity persists during postnatal development until weaning age (Olejnikova et al., 2018). Furthermore, maternal-fetal synchronization is abolished when the maternal SCN is ablated (Davis and Gorski, 1988; Greiner et al., 2022; Reppert and Schwartz, 1986).

Circadian Rhythms Influence Neuroimmune Wiring of Neuronal Circuitry

Disruption of the circadian system can have detrimental effects on various body systems. Modern-day causes of circadian disruption—including jetlag and shift working schedules—are modeled in animals by manipulating the light-dark environment. Mimicking chronic jet lag in adult mice by repeatedly advancing the light phase for 4 h over a year desynchronizes SCN neurons and activates immune pathways (Inokawa et al., 2020).

Early in life, microglia and astrocytes play essential roles in many neurodevelopmental processes (Erblich et al., 2011), so circadian disruptions during critical periods of development may alter neurodevelopmental trajectories (Estes and McAllister, 2016; Knuesel et al., 2014). Animal models of maternal immune activation (MIA) (Bauman et al., 2014; Missig et al., 2020; Rose et al., 2017), which recapitulate symptoms of ASDs (Fernandezde et al., 2017) or schizophrenia (Meyer and Feldon, 2012), have revealed that prenatal immune activation can induce long-term upregulation in inflammatory gene expression (Rose et al., 2017) and decrease complement-dependent homeostatic synaptic pruning, potentially interfering with neurodevelopmental synaptic refinement (Fernandez de Cossio et al., 2017). The developmental timing, type of immune challenge, and specific immune receptors involved collectively determine the effect on development, with unique MIA protocols recapitulating symptoms from distinct NDDs. For example, Poly I:C (viral mimetic) given to pregnant rats activates TLR3 and TLR4 receptors, resulting in lasting autism-related social and anxiety-like behaviors, whereas a similarly timed imiquimod (IMQ) immune challenge activates TLR7 and results in opposite behaviors (Missig et al., 2020). In some cases, MIA effects are sex-dependent, underscoring the complex interactions between immune and neurodevelopmental programs early in life (Haida et al., 2019). As several NDDs display a sex bias (Aleman et al., 2003; Christensen et al., 2016; Ferri et al., 2018; Loomes et al., 2017), and symptoms may differ between males and females (Diflorio and Jones, 2010; Erol et al., 2015; Sommer et al., 2020; Werling and Geschwind, 2013), a better understanding of sex-specific effects of MIA is likely to enhance prevention and treatment options for both sexes.

Accordingly, recent evidence has revealed that disrupting environmental light/dark cycles in utero or postnatally alters neuroimmune responses with long-lasting behavioral repercussions, similar to the effects of MIA during gestation. Early-life exposure to constant light (LL), dim light at night (dLAN), and chronic jet lag (CJL) affects developmental trajectories.

Constant Light

From conception, the maternal immune environment influences embryonic development, so even in utero, developing offspring can be affected by their mother’s response to altered environmental light. Pregnant mice exposed to constant light (LL) exhibit decreased placental expression of Serpinf1 and upregulation of macrophage markers, Iba1 and CD11b (Clarkson-Townsend et al., 2021). If changes in placental macrophages (Edlow et al., 2019) are mirrored in embryonic microglia, these immune shifts may alter the neuroimmune environment even during the earliest periods of brain development. Early postnatal LL also shifts the neuroimmune environment by increasing astrocyte number and decreasing astrocytic arborization (Canal et al., 2009). Behaviorally, mice raised from birth to weaning under LL have weakened circadian rhythms in wheel running (Ohta et al., 2006), which persist into adulthood, long after the animals are switched to LD. Therefore, developmental LL disrupts the immune system and induces lasting changes in animal behaviors.

Light at Night

It is increasingly recognized in humans that artificial lighting at night, especially blue light from electronic devices, negatively affects sleep/wake cycles and mood (Bedrosian and Nelson, 2013). Using a mouse model of dLAN, we showed that temporary exposure to dLAN during juvenile or adolescent stages alters microglial reactivity and mood behaviors long into adulthood (Chen et al., 2021). Female mice were most susceptible to these early-life circadian disruptions, which primed the immune system and increased microglial reactivity to an LPS challenge in adulthood. These dLAN-induced neuroimmune shifts were accompanied by increased anxiety-like behaviors in female adults and increased depressive-like behaviors in both male and female adults (Chen et al., 2021). Another study found similar anxiety-like behaviors in adult mice after in utero or postnatal dLAN exposure (Borniger et al., 2014). These data suggest that early life is a sensitive period during which circadian disruption via aberrant timing of dim light can induce negative long-term developmental outcomes.

Chronic Jet Lag

In an elegant cross-fostering study, Smarr and colleagues (2017) recently explored the behavioral effects of pre- and postnatal chronodisruption using a murine CJL model. In this CJL study, pregnant dams or newborn offspring were exposed to 6 h phase advances every 4 days. In utero or postnatal CJL decreased social behaviors and increased stereotyped behaviors, similar to the phenotype observed in other animal models for ASD. Interestingly, CJL had the greatest impact during gestation. CJL effects were also magnified in the offspring who experienced both in utero and postnatal CJL, implying cumulative impacts over developmental time. Collectively, the body of experimental data in rodent models shows that environmental light cycles affect developmental neuroimmune and behavioral outcomes.

Circadian Neuroimmune Dysregulation in the Etiology of Neurodevelopmental Disorders

Epidemiological studies have repeatedly shown a connection between early-life immune challenges and NDDs, such as ASDs, schizophrenia, and BD (Atladottir et al., 2010; Brown and Derkits, 2010; Jiang et al., 2016). Viral or bacterial infection during pregnancy is associated with neurodevelopmental disruptions to the offspring. Gestational immune activation, measured by maternal plasma cytokine levels, correlates with autism risk (Goines et al., 2011; Jones et al., 2017). Inflammation resulting from birth complications or acute perinatal hypoxia similarly increases the risks of long-term neurological complications (Hagberg et al., 2015). Early-life immune challenges can result in a sensitized immune system throughout life and prolonged disruptions to neurodevelopmental processes with lasting cognitive and psychiatric consequences (Knuesel et al., 2014).

The circadian system may be critical for balancing inflammatory and homeostatic neuroimmune functions throughout early brain development, as dysregulation of circadian neuroimmune function early in life is implicated in the etiology of several NDDs. Individuals with NDDs commonly exhibit increased neuroinflammation and sleep disruptions, but it is unclear to what extent these are correlating symptoms or if they interact to exacerbate disorder pathology. Many human disorders, including NDDs, do not stem from a single cause but involve interlocking molecular pathways, called the interactome, in which cellular processes influence one another to jointly contribute toward disorder progression (Barabasi et al., 2011; Paci et al., 2021). The emerging field of network medicine treats disease states as a network perturbation and attempts to characterize diseases by their distinct alterations to the interactome. By exploring the impact of gene and protein pathways on one another, this approach can help identify novel targets for treatment and intervention.

The centrality of the circadian system to a wide range of physiological functions suggests that whether or not molecular clock dysfunction is a primary cause of NDD, it may play a role in mediating the progression of a disorder, through its interactions with other cellular processes (Mullegama et al., 2015). Here, we discuss evidence that circadian neuroimmune disruptions may play a role in the pathophysiology of NDDs, including ASDs, BD, and schizophrenia. We highlight cellular pathways which have been implicated in these disorders and discuss how the molecular clock is impacted by and may synergize with these pathways to exacerbate neuroinflammation and influence disorder progression.

Autism Spectrum Disorders

ASDs are a heterogenous group of early-onset NDDs characterized by social and communication impairments and restricted, repetitive behaviors (American Psychiatric Association [APA], 2013). Neuroinflammation may play a role in the heterogenous developmental disruptions that occur in ASD. Positron emission tomography (PET) scans in young adults with ASD reveal increased microglial reactivity across many brain regions, most markedly in the cerebellum (Suzuki et al., 2013). Postmortem samples from individuals with autism reveal increased levels of brain cytokines (Li et al., 2009; Wei et al., 2011) and altered microglial morphology compared with typically developing individuals (Lee et al., 2017).

Circadian disruption is common in individuals with ASD and may contribute to neuroinflammation. Most patients with ASD experience chronic sleep disruption (Veatch et al., 2015) and exhibit altered rhythms in body temperature, activity, and melatonin and serotonin secretion (Glickman, 2010; Melke et al., 2008; Rossignol and Frye, 2011). Genome-wide association studies have uncovered gene polymorphisms and allele variants that confer significant risk for ASDs (Satterstrom et al., 2020; Stessman et al., 2017). Many of these disrupt the circadian system either directly through the molecular clock (Yang et al., 2016) or by altering the central pace-making function of the SCN (Ingiosi et al., 2019; Ma et al., 2022). Mutations in Npas, Per2, and Rev-erb (Goto et al., 2017; Hu et al., 2009) and hypomethylation of the Rora promoter (Nguyen et al., 2010) are implicated as ASD risk factors.

Outside of core clock components, other ASD-related molecular targets, such as FMRP and mTOR, interrupt both neuroimmune and circadian systems. Transcriptional silencing of Fmr1, which encodes the fragile X mental retardation protein (FMRP), causes Fragile X Syndrome, which is the most common single-gene cause of ASD (Richter and Zhao, 2021). FMRP interacts with several circadian transcripts, and Fmr1 knockout mice show reduced hippocampal clock gene expression and time-of-day-specific memory impairments (Sawicka et al., 2019). FMRP-deficient microglia show exaggerated inflammatory responses to LPS (Parrott et al., 2021), and loss of FMRP results in an increased number of dendritic spines (Bagni and Greenough, 2005), possibly from defects in microglial-mediated synaptic pruning during development (Jawaid et al., 2018). The mTOR-eIF4E pathway provides another intriguing link between central clock dysfunction and inflammation/homeostasis imbalance in ASDs. The mTOR pathway is responsible for photic entrainment in the SCN (Cao et al., 2010), as well as regulating the molecular clock (Ramanathan et al., 2018; Zheng and Sehgal, 2010). Overactive mTOR-dependent protein synthesis increases autism-like behaviors in mice, and microglia-specific eIF4E overexpression results in increased microglial reactivity, increased phagocytosis, decreased spine pruning, and autism-like behaviors (Xu et al., 2020). Disruptions to either FMRP or mTOR signaling pathways impair circadian rhythmicity and alter neuroimmune function early in life, contributing to aberrant developmental processes.

Bipolar Disorder

BD is an affective disorder with typical onset in adolescence or early adulthood. BD is characterized by cyclic mood disturbances, alternating between mania and depression, with euthymic periods of normal mood function interspersed (Vieta et al., 2018). Maternal influenza infection increases the risk of offspring developing BD (Canetta et al., 2014; Parboosing et al., 2013). Depressive and manic episodes are associated with increased plasma concentrations of inflammatory cytokines (O’Brien et al., 2006), and postmortem analysis in prefrontal cortex from humans with BD revealed higher cytokine levels and increased astrogliosis and microgliosis than in healthy controls (Rao et al., 2010). RNA sequencing from human tissue has also implicated neuroimmune pathways in BD (Pacifico and Davis, 2017).

In the general population, disrupted circadian rhythms are associated with increased rates of mood disorders and poorer subjective well-being (Lyall et al., 2018). In BD, sleep disruptions occur in the early prodromal phase (i.e., mild symptoms experienced before the major signs of a disease state), sometimes occurring up to 10 years before the onset of the first manic episode (Conus et al., 2008). Furthermore, actigraphy studies demonstrate that circadian activity in bipolar individuals is distinct from that of control individuals, even in euthymic cycles (the stable mental state) (Salvatore et al., 2008). Also, genetic polymorphisms in core clock components occur at higher rates in individuals with affective disorders (Etain et al., 2011; Kripke et al., 2009; Severino et al., 2009), and a Per3 variable-number tandem repeat influences the onset age of BD (Benedetti et al., 2008). Collectively, these data suggest that circadian disruptions represent a persistent pathological feature of the disorder and not simply a symptom of mood state.

Disturbance to the molecular clock may cause the development of BD, as suggested by studies using gene disruption in mouse mutants. Mice with Clock gene knockdown develop mania-like behaviors, including hyperactivity and reduced anxiety (Roybal et al., 2007). When restricted to the ventral tegmental area (VTA) alone, Clock knockdown recapitulates both manic and depressive-like behaviors, analogous to those seen in humans with BD (Mukherjee et al., 2010). Early in life, the molecular clock influences the wiring of the dopaminergic system, which plays a role in motivation, attention, learning, and mood regulation. Through adulthood there are circadian rhythms in the transcription of monoamine oxidase A, which degrades dopamine in the synaptic cleft (Hampp et al., 2008). Clock knockdown interferes with this regulation, increasing dopaminergic signaling, which may underlie bipolar mood fluctuations. Behavioral phenotypes in Clock mutant mice are ameliorated by lithium treatment, which is a potent inhibitor of glycogen synthase kinase 3 (GSK-3), suggesting a central role for GSK-3 in modulating circadian neuroimmunity. In Drosophila, GSK-3 shortens the circadian period (Martinek et al., 2001). In rodents, GSK-3 is expressed in a circadian fashion and modulates the expression of multiple clock genes in SCN, promoting the phosphorylation and degradation of CRY2 (Kurabayashi et al., 2010), and dampening the amplitudes of BMAL1 and PER rhythms (Besing et al., 2015). GSK-3 is also known to increase microglial pro-inflammatory reactivity (Martinez et al., 2002). Lithium’s inhibition of GSK-3 activity may therefore act as a circadian neuroimmune therapy, working by increasing amplitude of Per2 rhythmic expression (Li et al., 2012) and decreasing inflammation.

Schizophrenia

Schizophrenia is a disabling psychiatric disorder involving disordered thinking, hallucinations, or delusions, and often first presents in adolescence or young adulthood (Tandon et al., 2008). Schizophrenia is associated with a decrease in gray matter, due to excessive developmental synaptic pruning (Prasad et al., 2016). The complement system regulates synaptic pruning by tagging synapses for removal with molecular signals that stimulate microglial engulfment (Schafer et al., 2012; Stevens et al., 2007). The complement component C4 is implicated in the altered pruning in schizophrenia, as greater schizophrenia risk is associated with the C4A allele compared with the C4B allele (Sekar et al., 2016). Serum complement analyte concentrations can be used as a biomarker of first-episode psychosis (Kopczynska et al., 2019) and induced microglia-like cells from patients with schizophrenia exhibit greater synaptic engulfment, which correlates with C4 risk variant (Sellgren et al., 2019).

Like ASD and BD, several lines of evidence suggest that circadian dysfunction may be a central contributor to pathology in schizophrenia. Individuals with schizophrenia often experience lifelong sleep difficulties and disrupted cycles of melatonin production (Johansson et al., 2016; Wulff et al., 2012). For adolescents with a familial risk of psychosis, circadian disturbances correlated with worse outcomes 1 year later (Lunsford-Avery et al., 2017). Patients with schizophrenia exhibit loss of rhythmicity in several core clock genes, including Clock, Per2, and Cry (Johansson et al., 2016), and postmortem cortical tissue from individuals with schizophrenia reveals reduced rhythmicity in typical circadian transcripts (Seney et al., 2019). Genetic risk factors may underlie this circadian arrhythmicity, as the Clock gene T3111C polymorphism is more highly prevalent in individuals with schizophrenia than in controls (Takao et al., 2007). One additional gene implicated in the etiology of schizophrenia (Hennah et al., 2003), Disrupted-in-Schizophrenia1 (Disc1), is also linked to the molecular clock. Disc1 was discovered through a gene-linkage study of a Scottish family with high penetrance of schizophrenia and adolescent affective and conduct disorders (St Clair et al., 1990). The DISC1 protein is a direct target of the BMAL1/CLOCK heterodimer, and it stabilizes BMAL1 activity through interactions with GSK-3β (Lee et al., 2021). Faulty DISC1 activity may alter circadian rhythms, disrupt neuroimmune homeostasis, and could shift microglia into a state of excessive pruning, contributing to schizophrenia pathology.

Here, we discussed links between circadian and neuroimmune function in human NDDs and possible mechanisms. Each of these disorders involves complex disruptions to the typically developing nervous system, with heterogenous symptomatology and multifaceted and interacting mechanisms. These disorders express comorbidly in the general population, and studies suggest that ASDs, BD, and schizophrenia share overlapping genetic and environmental risk mechanisms (Carroll and Owen, 2009; Gandal et al., 2018; Rees et al., 2021; Stefansson et al., 2009). Many of the molecular players already discussed have also been implicated in neurodegenerative disorders (Bie et al., 2019; Bleuze et al., 2021), which will be discussed in the next section. Circadian disruptions and altered neuroimmune pathways represent a common theme in a wide range of neurological disorders throughout the lifespan. Future studies should explore whether strengthening the circadian system can reduce neuroinflammation and alleviate symptoms in these disorders.

Age-Associated Circadian Changes and Neurodegenerative Diseases

The neuroimmune system plays multiple roles in promoting healthy brain development and optimal nervous system functioning throughout the lifespan. During natural aging, neuroimmune function gradually shifts toward higher levels of basal inflammation (Erraji-Benchekroun et al., 2005; Soreq et al., 2017). Both neuronal and glial transcription change over time, with a general decrease in neuronal transcription and an increase in immune-related transcripts with age (Boisvert et al., 2018; Grabert et al., 2016). In healthy aging, microglia increase the basal release of pro- inflammatory cytokines, such as IL-1α, TNF, and C1qA (Clarke et al., 2018), and astrocytes exhibit increasingly reactive transcriptional profiles with upregulation of complement and major histocompatibility complex (MHC) (Boisvert et al., 2018). These changes result in an increase in synaptic pruning and a gradual loss of dendritic spines and gray matter with age.

The aging neuroimmune system may also be compromised by progressively weakening circadian rhythms. Circadian amplitudes of activity and body temperature decrease with aging (Czeisler et al., 1992; Musiek et al., 2018), and aged adults wake earlier and have melatonin peaking earlier in the circadian cycle (Duffy et al., 2002). Elderly individuals experience more fragmented sleep (Carskadon et al., 1982), with over half of individuals aged 65 or older experiencing some form of chronic sleep disruption that is linked to negative health outcomes (Foley et al., 1995).

There are several mechanisms implicated in this circadian decline. First, clock gene oscillations throughout the brain dampen with aging (Chen et al., 2016). Second, there is a decline in SCN synchronization with aging, although there are mixed findings regarding whether this is due to the death of vasoactive intestinal polypeptide (VIP) neurons (Roberts et al., 2012; Wang et al., 2015) or the desynchronization of physiological properties of SCN neurons (Curran et al., 2019). Transplantation of fetal or young SCN tissue into aged hamsters can restore disrupted circadian rhythms and increase longevity, but the mechanisms for this effect are unclear. Electrophysiological recordings in rodent SCN demonstrate that neuronal membrane properties lose rhythmicity with age, due to decreasing GABAergic input and decreased A-type potassium currents (Farajnia et al., 2012).

Another mechanism for dampened rhythms with aging may involve a reduced expression of the NAD+-dependent deacetylase, sirtuin 1 (SIRT1) (Meng et al., 2020). Sirtuins were initially noted for their ability to increase longevity in yeast and have since been implicated in many aging-related processes across animal species, with a role in glucose metabolism, mitochondrial oxidation, telomere regulation, transcriptional repression, DNA repair, and maintaining genome integrity (Toiber et al., 2011). SIRT1 directly modulates the cellular clock by both binding the BMAL1-CLOCK heterodimer and deacetylating PER2 to enhance the amplitude of the feedback loop. SIRT1 protein levels naturally decline with age, contributing to weakening circadian rhythms, but transgenic mice overexpressing SIRT1 retain robust rhythms in gene expression, even in old age (Chang and Guarente, 2013).

Circadian Dysregulation of Neuroimmune Cells With Aging

Animal models show that the neuroimmune system loses circadian rhythmicity with aging, possibly contributing to the increased inflammation seen in immunosenescence. Hippocampal microglia from young rats (3 months) show robust circadian rhythms in clock genes and inflammatory markers, but aged rats (24 months) have reduced rhythms in Per genes and tonically increased inflammation, without circadian variation (Fonken et al., 2016). In addition, SIRT1 is also involved in regulating the immune system (Gamez-Garcia and Vazquez, 2021). Nuclear sirtuins inhibit microglial inflammation, but in microglia in the aged brain, an increase in nuclear cathepsin B degrades sirtuins, thereby increasing neuroinflammation (Meng et al., 2020). SIRT1 hereby contributes to both circadian weakening and neuroinflammation, significantly affecting longevity (Satoh et al., 2013).

Circadian regulation in the glymphatic system also decreases in the aged brain. The glymphatic system is a perivascular network that regulates CSF influx into the brain parenchyma and interstitial solute clearance by draining into the lymphatic system. In young animals, glymphatic clearance is highest during sleep, as slow-wave brain activity drives an increase in interstitial volume, allowing for the clearance of accumulated metabolites (Xie et al., 2013). However, the knockout of AQP4 channels on astrocytic end-feet abolishes glymphatic rhythmicity (Hablitz et al., 2020), highlighting the importance of AQP4 channels to proper timing of metabolite clearance. In human postmortem tissue, a decrease in perivascular AQP4 is associated with increasing age (Zeppenfeld et al., 2017), suggesting that human glymphatic function may similarly lose circadian rhythmicity in aging.

Circadian Components in Age-Related Neurodegenerative Disorders

Age-related changes in circadian neuroimmunity may predispose certain individuals to develop neurodegenerative disorders, which involve a toxic accumulation of protein aggregates, exaggerated and detrimental neuroinflammation, excessive neuronal death, and precipitous loss of cognitive or motor functions. Although protein misfolding and aggregation are pathological hallmarks of neurodegenerative diseases, there is no single driver of pathology; rather, degeneration involves interactions across multiple systems (Gan et al., 2018). Because the neuroimmune system plays a central role in phagocytosis and clearance of these protein aggregates, neuroimmune dysfunction, which may be exacerbated by underlying circadian dysrhythmia, may contribute to the progression of these diseases. Neurodegeneration is distinct from the “healthy” aging process. Here, we review the links between circadian and neuroimmune disruption in neurodegeneration, most of which are being studied in the context of two of the most common forms of neurodegeneration: Alzheimer’s disease (AD) and Parkinson’s disease (PD). Again, it is important to delineate correlation from causation. Neurodegenerative disorders often display concurrent shifts in both immune and circadian systems; accumulating evidence suggests that a pathological interaction of these dysfunctioning systems may play an additional role in disease progression.

Alzheimer’s Disease

AD is characterized by progressive memory impairment, along with psychiatric conditions like depression, aggression, and sleep disturbances. Pathological characteristics of AD include a buildup of protein aggregates, including intracellular neurofibrillary tangles and amyloid-beta (Aβ) plaques, diffusely affecting the entire brain, including the hippocampus, cortex, as well as hypothalamus (Ogomori et al., 1989). Inflammation is a key component of AD susceptibility. Inflammatory genes are highly implicated in AD, including complement receptor 1 and Trem2, which regulate phagocytosis and are important for the clearance of aβ plaques (Karch and Goate, 2015).

In humans, preclinical AD pathology is associated with fragmentation of sleep patterns, with 70% of AD patients exhibiting sleep disruptions, when measured by in-home actigraphy (Musiek et al., 2018), suggesting that sleep could be used as a biomarker for early AD diagnosis and intervention. For individuals with AD, sleep disturbances exacerbate behavioral symptoms, including aggression (Moran et al., 2005). Human postmortem samples suggest that aberrant rhythms in DNA methylation and Bmal1 expression could underlie circadian disruption in AD patients (Cronin et al., 2017).

While the SCN is less susceptible to plaque-induced cell death than the hippocampus or cortex (Stopa et al., 1999), this brain region displays a course of pathology in AD that is more severe than what occurs with healthy aging and is likely to play a role in the progression of AD (Van Erum et al., 2018). Through mild and moderate AD, some patients maintain typical circadian rhythmicity in hormone patterns including plasma cortisol (Hatfield et al., 2004), but AD patients have fewer SCN VIP neurons than age-matched controls (Stopa et al., 1999; Zhou et al., 1995). Severely affected individuals show significant death of VIP neurons and astrogliosis in the SCN (Stopa et al., 1999). Loss of vasopressin or neurotensin neurons in the SCN is associated with specific fragmentation in activity rhythms (Harper et al., 2008), suggesting that pathology underlies the progressive sleep disturbances associated with AD.

Circadian disruptions exaggerate neuroinflammation and increase neurotoxicity in AD rodent models with genetic mutations that alter amyloid or tau processing. Overexpressing amyloid precursor protein (APP) promotes an early AD phenotype in young animals. Two-month-old APP-knock-in mice have altered clock gene expression and increased time-of-day-dependent expression of inflammatory markers, including IL-6 and TNF-α (Ni et al., 2019). In another AD mouse model with the human tau variant, tauP301S, inhibiting microglial Sirt1 increases the expression of IL-1β in young mice and exacerbates memory deficits in aged mice (Cho et al., 2015). As previously discussed, Sirt1 deletion also disrupts circadian rhythms. As Sirt1 naturally declines with aging, dampened circadian rhythms and increased IL-1β expression may combine with other genetic risk factors to promote inflammatory pathology in individuals with AD.

Animal studies provide the strongest evidence that the circadian system directly modulates AD progression: in mice, a robust molecular clock facilitates the clearance of Aβ plaques in AD models, whereas circadian disruptions inhibit plaque clearance (Lee et al., 2020; Zhu et al., 2018). Bmal1 expression drives microglial phagocytosis of fAβ1-42 in immortalized BV-2 mouse microglial cells. In 5xFAD mice, pharmacological or siRNA inhibition of REV-ERBα/β increases Bmal1 expression and microglial fAβ1-42 clearance (Lee et al., 2020). Crossing the Rev-erba knockout mouse with the 5XFAD mouse decreases amyloid plaque burden and disease-associated microglial gene signatures (Lee et al., 2020). In the P301S mouse model of tauopathy, early-life sleep disruption increases glia cell reactivity, increases the accumulation of soluble tau, and hastens neurodegeneration in the hippocampus and amygdala, resulting in motor and memory impairments (Zhu et al., 2018).

Beyond local protein degradation mechanisms, circadian dysregulation inhibits the clearance of plaques through the glymphatic system in AD. Acute sleep deprivation increases the amount of Aβ accumulation in humans (Shokri-Kojori et al., 2018). AQP4 is implicated in the pathophysiology of AD due to its role in the clearance of Aβ plaques (Iliff et al., 2012). Furthermore, Aβ pathology and AD severity correlate with decreasing astrocytic AQP4, even when controlling for age (Zeppenfeld et al., 2017).

Parkinson’s Disease

PD is a progressive neurodegenerative disorder involving the death of dopaminergic neurons projecting from the substantia nigra to the basal ganglia resulting in disrupted motor planning, initiation, and execution. Motor symptoms include bradykinesia, tremor, rigidity, and postural instability, while non-motor features include cognitive impairment, emotional dysregulation, and sleep disruptions (Jankovic, 2008). Sleep irregularities including nighttime insomnia or daytime sleepiness are reported in a majority of individuals with PD (Stefani and Hogl, 2020), and cognitive, motor, and mood symptoms tend to increase at dusk. Termed “sundowning,” this time-of-day-dependent exacerbation of symptoms implies the involvement of circadian mechanisms in PD behaviors.

Several studies suggest that maintaining the integrity of the molecular clock in aging may reduce or delay neuronal death in PD. Rodent models of PD include intraperitoneal injection of 6-hydroxydopamine (6-OHDA) or 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) which causes acute death of dopaminergic neurons in the substantia nigra, motor impairments, and time-of-day dependent depression and anxiety-like behaviors. Bmal1 knockout mice exhibit increased microglial number and inflammatory cytokines like IL-1β and TNF-α after MPTP injection, exacerbating dopaminergic cell death, compared to wild-type mice (Liu et al., 2020). The opposing limb of the molecular clock, REV-ERBα, is also protective in PD models. Rev-erbα knockout mice exhibited worsened motor impairments and neuronal death in the substantia nigra (SNpc) and VTA after 6-OHDA injection compared to wild-type controls (Kim et al., 2018). Rev-erbα knockouts also display a prolonged increase in the number of microglia, but not astrocytes, after 6-OHDA injection. Local injection of the REV-ERBα antagonist, SR8278, into the VTA ameliorated these mood-related behaviors (Kim et al., 2022). Both Bmal1 and Rev-erbα may be neuroprotective for dopaminergic neurons in the striatum, although the mechanisms through which these opposing limbs may contribute to neuroprotection are unknown.

Future Perspectives

Additional research using adult animal models will unveil the mechanisms by which clock genes direct neuroimmune activities. Because the circadian system involves both cellular and system-level controls, it is difficult to disentangle the mechanisms linking circadian disruption and neuroinflammation. For example, global Rev-erbα knockout increases pro-inflammatory cytokines in the brain, but this inflammation may indirectly result from increased neuronal oxidation levels which then secondarily drive pro-inflammation (Griffin et al., 2019). Rather than global gene knockouts, cell-type-specific clock gene knockouts allow researchers to induce circadian disruptions in a single cell type and analyze its effects on network functioning and its broader impacts on animal behavior. For example, microglia-specific Bmal1 knockdown has revealed a cell-autonomous role for Bmal1 in mediating neuroinflammation and synaptic phagocytosis (Griffin et al., 2020). Continued work is needed using cell-type-specific clock gene knockouts to dissect cell-autonomous and nonautonomous mechanisms.

Microglia and astrocytes exhibit unique morphological and functional characteristics depending on the brain region (Grabert et al., 2016). Targeting cell-specific manipulations of clock genes to specific brain regions is also a research priority because regions might differ in sensitivity to circadian disruptions. For example, is the dopaminergic system more sensitive to circadian perturbations than other neuromodulatory systems? Dopaminergic dysfunction may be central to many psychiatric and neurodegenerative disorders because it requires tighter circadian regulation, hence may be more sensitive to circadian disruption than other systems.

Circadian disruptions alter neuroimmune homeostasis and interfere with healthy brain development or aging. Their involvement in a broad range of NDDs and neurodegenerative diseases highlights the need for a deeper exploration of circadian neuroimmune functions at all life stages. Despite the importance of early-life brain development, few studies have explored the role of circadian rhythms in early-life neuroimmune activities. How do maternal circadian rhythms influence fetal neuroimmune functions, and how might perinatal circadian disruption influence the immune system to contribute toward NDDs? At the other end of life, in aged individuals, a loss of circadian rhythmicity and increased pro-inflammatory responses characterize immunosenescence. Additional research on the extreme ages of life will help discern the importance of circadian rhythms in developmental processes and in neuroprotection.

The circadian system is central to human health and disease, and many pharmaceuticals target clock-controlled molecular mechanisms (Zhang et al., 2014). There is increasing awareness that timed dosing is an important consideration for treatments, such as cancer treatments (Dallmann et al., 2016). In addition, several pharmaceuticals have been developed that directly target the circadian system, including drugs that modulate the molecular clock (via RORs or REV-ERBs) as well as drugs that influence SCN phase, via melatonin or orexin receptors (Ruan et al., 2021). As our understanding of circadian neuroimmune mechanisms increases, circadian dosing or circadian modulators may be useful for the treatment of various neurological diseases.

Behavioral interventions that strengthen circadian rhythms are some of the most effective ways to protect cognitive and psychiatric health across all ages. Early morning light exposure (Tao et al., 2020), daytime exercise (Hughes et al., 2021), time-restricted diets (Balasubramanian et al., 2020; Hatori et al., 2012; Hu et al., 2019; Upadhyay et al., 2019), and proper sleep hygiene (Prince and Abel, 2013) are beneficial for combatting symptoms of depression, protecting against memory loss, and increasing longevity. The precise mechanisms connecting these behavioral interventions with the molecular clock and cognitive health are not fully known, and further investigation is needed into the role of the neuroimmune system in mediating these therapeutic strategies. In addition, it is unclear whether similar chronotherapies may support brain development early in life (Novakova et al., 2010; Prates et al., 2022) and ameliorate behavioral deficits in NDDs. Bright light therapy helps alleviate the symptoms of attention-deficit hyperactivity disorder (ADHD), and exercise is similarly beneficial for sleep and motor behaviors in children with autism (Brand et al., 2015; Turner and Johnson, 2013). Bolstering circadian rhythms early in life may help with immediate symptoms and could also ameliorate disorder trajectory. Early-life circadian reinforcement might lead to lasting effects on health, aging, and neurodegeneration.

Conclusions

The CNS directs a diverse repertoire of complex behaviors, including cognition, memory, personality, motor control, emotional regulation, and social interactions. As such, the rich social-emotional depth of the human lived experience relies on preserving the integrity of the CNS. Critically, the neuroimmune system maintains CNS homeostasis to optimize neural development and function throughout the lifespan. Neuroimmune cells such as microglia and astrocytes actively sculpt brain circuitry early in life, participating in axonal guidance, synaptogenesis, and spine pruning; in adulthood, the neuroimmune system facilitates learning and memory; and in aging, neuroimmune cells protect against the accumulation of misfolded proteins and neuronal death.

As this review highlights, neuroimmune responses to injury or illness are dynamic. Circadian regulation of neuroimmune activities is essential for protecting neural networks throughout the lifespan. On the other hand, circadian dysregulation may alter immune function at any age, disrupting CNS neuroimmune homeostasis, and thereby contributing to susceptibility to cognitive or psychiatric disorders. A growing body of evidence increasingly implicates circadian neuroimmune dysfunction in many neurodevelopmental and degenerative disorders—beginning before birth and extending throughout the lifespan. A deeper exploration of circadian neuroimmune functions across all ages will likely uncover novel avenues for treating a broad spectrum of human neurological conditions.

Acknowledgments

This work was supported by NIA grants R01AG078758 and R01AG062716 to L.K.F. and R01NS131806 to A.D.G. R.C. was supported by T32AA007471 and B.N.R. was supported by T32DA018926.

Footnotes

The author(s) have no potential conflicts of interest with respect to the research, authorship, and/or publication of this article.

References

  1. Adeluyi A, Guerin L, Fisher ML, Galloway A, Cole RD, Chan SSL, Wyatt MD, Davis SW, Freeman LR, Ortinski PI, et al. (2019) Microglia morphology and proinflammatory signaling in the nucleus accumbens during nicotine withdrawal. Sci Adv 5:eaax7031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Aleman A, Kahn RS, Selten JP. (2003) Sex differences in the risk of schizophrenia: evidence from meta-analysis. Arch Gen Psychiatry 60:565-571. [DOI] [PubMed] [Google Scholar]
  3. Allen NJ, Bennett ML, Foo LC, Wang GX, Chakraborty C, Smith SJ, Barres BA. (2012) Astrocyte glypicans 4 and 6 promote formation of excitatory synapses via GluA1 AMPA receptors. Nature 486:410-414. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Altman J, Bayer SA. (1978) Development of the diencephalon in the rat. I. Autoradiographic study of the time of origin and settling patterns of neurons of the hypothalamus. J Comp Neurol 182:945-971. [DOI] [PubMed] [Google Scholar]
  5. Altman J, Bayer SA. (1986) The development of the rat hypothalamus. Adv Anat Embryol Cell Biol 100:1-178. [PubMed] [Google Scholar]
  6. American Psychiatric Association (APA) (2013) The Diagnostic and Statistical Manual of Mental Disorders (5th ed.; DSM-5). Arlington (VA): American Psychiatric Publishing. [Google Scholar]
  7. Amir M, Chaudhari S, Wang R, Campbell S, Mosure SA, Chopp LB, Lu Q, Shang J, Pelletier OB, He Y, et al. (2018) REV-ERBalpha regulates TH17 cell development and autoimmunity. Cell Rep 25:3733-3749. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Anderson MA, Burda JE, Ren Y, Ao Y, O’Shea TM, Kawaguchi R, Coppola G, Khakh BS, Deming TJ, Sofroniew MV. (2016) Astrocyte scar formation aids central nervous system axon regeneration. Nature 532:195-200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Ansari N, Agathagelidis M, Lee C, Korf HW, von Gall C. (2009) Differential maturation of circadian rhythms in clock gene proteins in the suprachiasmatic nucleus and the pars tuberalis during mouse ontogeny. Eur J Neurosci 29:477-489. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Armstrong SM, Cassone VM, Chesworth MJ, Redman JR, Short RV. (1986) Synchronization of mammalian circadian rhythms by melatonin. J Neural Transm Suppl 21:375-394. [PubMed] [Google Scholar]
  11. Astiz M, Heyde I, Oster H. (2019) Mechanisms of communication in the mammalian circadian timing system. Int J Mol Sci 20:343. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Atladottir HO, Thorsen P, Ostergaard L, Schendel DE, Lemcke S, Abdallah M, Parner ET. (2010) Maternal infection requiring hospitalization during pregnancy and autism spectrum disorders. J Autism Dev Disord 40:1423-1430. [DOI] [PubMed] [Google Scholar]
  13. Bae K, Jin X, Maywood ES, Hastings MH, Reppert SM, Weaver DR. (2001) Differential functions of mPer1, mPer2, and mPer3 in the SCN circadian clock. Neuron 30:525-536. [DOI] [PubMed] [Google Scholar]
  14. Bagni C, Greenough WT. (2005) From mRNP trafficking to spine dysmorphogenesis: the roots of fragile X syndrome. Nat Rev Neurosci 6:376-387. [DOI] [PubMed] [Google Scholar]
  15. Balasubramanian P, DelFavero J, Ungvari A, Papp M, Tarantini A, Price N, de Cabo R, Tarantini S. (2020) Time-restricted feeding (TRF) for prevention of age-related vascular cognitive impairment and dementia. Ageing Res Rev 64:101189. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Baldi P, Alhassen W, Chen S, Nguyen H, Khoudari M, Alachkar A. (2021) Large-scale analysis reveals spatiotemporal circadian patterns of cilia transcriptomes in the primate brain. J Neurosci Res 99:2610-2624. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Barabási A-L, Gulbahce N, Loscalzo J. (2011) Network medicine: a network-based approach to human disease. Nat Rev Genet 12:56-68. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Barahona RA, Morabito S, Swarup V, Green KN. (2022) Cortical diurnal rhythms remain intact with microglial depletion. Sci Rep 12:114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Barca-Mayo O, Boender AJ, Armirotti A, De Pietri Tonelli D. (2020) Deletion of astrocytic BMAL1 results in metabolic imbalance and shorter lifespan in mice. Glia 68:1131-1147. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Barca-Mayo O, Pons-Espinal M, Follert P, Armirotti A, Berdondini L, De Pietri Tonelli D. (2017) Astrocyte deletion of Bmal1 alters daily locomotor activity and cognitive functions via GABA signalling. Nat Commun 8:14336. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Bates K, Herzog ED. (2020) Maternal-fetal circadian communication during pregnancy. Front Endocrinol 11:198. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Bauman MD, Iosif AM, Smith SE, Bregere C, Amaral DG, Patterson PH. (2014) Activation of the maternal immune system during pregnancy alters behavioral development of rhesus monkey offspring. Biol Psychiatry 75:332-341. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Bedrosian TA, Nelson RJ. (2013) Influence of the modern light environment on mood. Mol Psychiatry 18:751-757. [DOI] [PubMed] [Google Scholar]
  24. Beersma DG, Gordijn MC. (2007) Circadian control of the sleep-wake cycle. Physiol Behav 90:190-195. [DOI] [PubMed] [Google Scholar]
  25. Benedetti F, Dallaspezia S, Colombo C, Pirovano A, Marino E, Smeraldi E. (2008) A length polymorphism in the circadian clock gene Per3 influences age at onset of bipolar disorder. Neurosci Lett 445:184-187. [DOI] [PubMed] [Google Scholar]
  26. Bennett ML, Bennett FC, Liddelow SA, Ajami B, Zamanian JL, Fernhoff NB, Mulinyawe SB, Bohlen CJ, Adil A, Tucker A, et al. (2016) New tools for studying microglia in the mouse and human CNS. Proc Natl Acad Sci USA 113:E1738-E1746. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Berson DM, Dunn FA, Takao M. (2002) Phototransduction by retinal ganglion cells that set the circadian clock. Science 295:1070-1073. [DOI] [PubMed] [Google Scholar]
  28. Besing RC, Paul JR, Hablitz LM, Rogers CO, Johnson RL, Young ME, Gamble KL. (2015) Circadian rhythmicity of active GSK3 isoforms modulates molecular clock gene rhythms in the suprachiasmatic nucleus. J Biol Rhythms 30:155-160. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Bie B, Wu J, Foss JF, Naguib M. (2019) Activation of mGluR1 mediates C1q-dependent microglial phagocytosis of glutamatergic synapses in Alzheimer’s rodent models. Mol Neurobiol 56:5568-5585. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Bisht K, Sharma KP, Lecours C, Sanchez MG, El Hajj H, Milior G, Olmos-Alonso A, Gómez-Nicola D, Luheshi G, Vallières L, et al. (2016) Dark microglia: a new phenotype predominantly associated with pathological states. Glia 64:826-839. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Bissonette GB, Bae MH, Suresh T, Jaffe DE, Powell EM. (2010) Astrocyte-mediated hepatocyte growth factor/scatter factor supplementation restores GABAergic interneurons and corrects reversal learning deficits in mice. J Neurosci 30:2918-2923. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Bleuze L, Triaca V, Borreca A. (2021) FMRP-Driven neuropathology in autistic spectrum disorder and Alzheimer’s disease: a losing game. Front Mol Biosci 8:699613. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Boisvert MM, Erikson GA, Shokhirev MN, Allen NJ. (2018) The aging astrocyte transcriptome from multiple regions of the mouse brain. Cell Rep 22:269-285. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Bolanos JP. (2016) Bioenergetics and redox adaptations of astrocytes to neuronal activity. J Neurochem 139:115-125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Bollinger T, Leutz A, Leliavski A, Skrum L, Kovac J, Bonacina L, Benedict C, Lange T, Westermann J, Oster H, et al. (2011) Circadian clocks in mouse and human CD4+ T cells. PLoS ONE 6:e29801. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Borniger JC, McHenry ZD, Abi Salloum BA, Nelson RJ. (2014) Exposure to dim light at night during early development increases adult anxiety-like responses. Physiol Behav 133:99-106. [DOI] [PubMed] [Google Scholar]
  37. Brancaccio M, Edwards MD, Patton AP, Smyllie NJ, Chesham JE, Maywood ES, Hastings MH. (2019) Cell-autonomous clock of astrocytes drives circadian behavior in mammals. Science 363:187-192. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Brancaccio M, Patton AP, Chesham JE, Maywood ES, Hastings MH. (2017) Astrocytes control circadian timekeeping in the suprachiasmatic nucleus via glutamatergic signaling. Neuron 93:1420-1435. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Brand S, Jossen S, Holsboer-Trachsler E, Puhse U, Gerber M. (2015) Impact of aerobic exercise on sleep and motor skills in children with autism spectrum disorders—a pilot study. Neuropsychiatr Dis Treat 11:1911-1920. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Brown AS, Derkits EJ. (2010) Prenatal infection and schizophrenia: a review of epidemiologic and translational studies. Am J Psychiatry 167:261-280. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Brown SA, Zumbrunn G, Fleury-Olela F, Preitner N, Schibler U. (2002) Rhythms of mammalian body temperature can sustain peripheral circadian clocks. Curr Biol 12:1574-1583. [DOI] [PubMed] [Google Scholar]
  42. Bull C, Freitas KC, Zou S, Poland RS, Syed WA, Urban DJ, Minter SC, Shelton KL, Hauser KF, Negus SS, et al. (2014) Rat nucleus accumbens core astrocytes modulate reward and the motivation to self-administer ethanol after abstinence. Neuropsychopharmacology 39:2835-2845. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Butovsky O, Jedrychowski MP, Moore CS, Cialic R, Lanser AJ, Gabriely G, Koeglsperger T, Dake B, Wu PM, Doykan CE, et al. (2014) Identification of a unique TGF-beta-dependent molecular and functional signature in microglia. Nat Neurosci 17:131-143. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Canal MM, Mohammed NM, Rodriguez JJ. (2009) Early programming of astrocyte organization in the mouse suprachiasmatic nuclei by light. Chronobiol Int 26:1545-1558. [DOI] [PubMed] [Google Scholar]
  45. Canetta SE, Bao Y, Co MD, Ennis FA, Cruz J, Terajima M, Shen L, Kellendonk C, Schaefer CA, Brown AS. (2014) Serological documentation of maternal influenza exposure and bipolar disorder in adult offspring. Am J Psychiatry 171:557-563. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Cao R, Li A, Cho HY, Lee B, Obrietan K. (2010) Mammalian target of rapamycin signaling modulates photic entrainment of the suprachiasmatic circadian clock. J Neurosci 30:6302-6314. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Carmona-Alcocer V, Abel JH, Sun TC, Petzold LR, Doyle FJ, 3rd, Simms CL, Herzog ED. (2018) Ontogeny of circadian rhythms and synchrony in the suprachiasmatic nucleus. J Neurosci 38:1326-1334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Carmona-Alcocer V, Rohr KE, Joye DAM, Evans JA. (2020) Circuit development in the master clock network of mammals. Eur J Neurosci 51:82-108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Carroll BJ, Cassidy F, Naftolowitz D, Tatham NE, Wilson WH, Iranmanesh A, Liu PY, Veldhuis JD. (2007) Pathophysiology of hypercortisolism in depression. Acta Psychiatr Scand Suppl 433:90-103. [DOI] [PubMed] [Google Scholar]
  50. Carroll LS, Owen MJ. (2009) Genetic overlap between autism, schizophrenia and bipolar disorder. Genome Med 1:102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Carskadon MA, Brown ED, Dement WC. (1982) Sleep fragmentation in the elderly: relationship to daytime sleep tendency. Neurobiol Aging 3:321-327. [DOI] [PubMed] [Google Scholar]
  52. Carver KA, Lourim D, Tryba AK, Harder DR. (2014) Rhythmic expression of cytochrome P450 epoxygenases CYP4x1 and CYP2c11 in the rat brain and vasculature. Am J Physiol Cell Physiol 307:C989-998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Casano AM, Peri F. (2015) Microglia: multitasking specialists of the brain. Dev Cell 32:469-477. [DOI] [PubMed] [Google Scholar]
  54. Chang C, Loo CS, Zhao X, Solt LA, Liang Y, Bapat SP, Cho H, Kamenecka TM, Leblanc M, Atkins AR, et al. (2019) The nuclear receptor REV-ERBalpha modulates Th17 cell-mediated autoimmune disease. Proc Natl Acad Sci USA 116:18528-18536. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Chang HC, Guarente L. (2013) SIRT1 mediates central circadian control in the SCN by a mechanism that decays with aging. Cell 153:1448-1460. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Chen CY, Logan RW, Ma T, Lewis DA, Tseng GC, Sibille E, McClung CA. (2016) Effects of aging on circadian patterns of gene expression in the human prefrontal cortex. Proc Natl Acad Sci USA 113:206-211. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Chen R, Weitzner AS, McKennon LA, Fonken LK. (2021) Light at night during development in mice has modest effects on adulthood behavior and neuroimmune activation. Behav Brain Res 405:113171. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Chen S, Fuller KK, Dunlap JC, Loros JJ. (2020) A pro- and anti-inflammatory axis modulates the macrophage circadian clock. Front Immunol 11:867. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Cheng WJ, Liu CS, Hu KC, Cheng YF, Karhula K, Harma M. (2021) Night shift work and the risk of metabolic syndrome: findings from an 8-year hospital cohort. PLoS ONE 16:e0261349. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Cho SH, Chen JA, Sayed F, Ward ME, Gao F, Nguyen TA, Krabbe G, Sohn PD, Lo I, Minami S, et al. (2015) SIRT1 deficiency in microglia contributes to cognitive decline in aging and neurodegeneration via epigenetic regulation of IL-1beta. J Neurosci 35:807-818. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Christensen DL, Baio J, Van Naarden Braun K, Bilder D, Charles J, Constantino JN, Daniels J, Durkin MS, Fitzgerald RT, Kurzius-Spencer M, et al. (2016) Prevalence and characteristics of autism spectrum disorder among children aged 8 years—autism and developmental disabilities monitoring network, 11 sites, United States, 2012. MMWR Surveill Summ 65:1-23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Chung WS, Clarke LE, Wang GX, Stafford BK, Sher A, Chakraborty C, Joung J, Foo LC, Thompson A, Chen C, et al. (2013) Astrocytes mediate synapse elimination through MEGF10 and MERTK pathways. Nature 504:394-400. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Clarke LE, Liddelow SA, Chakraborty C, Munch AE, Heiman M, Barres BA. (2018) Normal aging induces A1-like astrocyte reactivity. Proc Natl Acad Sci USA 115:E1896-E1905. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Clarkson-Townsend DA, Bales KL, Hermetz KE, Burt AA, Pardue MT, Marsit CJ. (2021) Developmental chronodisruption alters placental signaling in mice. PLoS ONE 16:e0255296. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Conus P, Ward J, Hallam KT, Lucas N, Macneil C, McGorry PD, Berk M. (2008) The proximal prodrome to first episode mania—a new target for early intervention. Bipolar Disord 10:555-565. [DOI] [PubMed] [Google Scholar]
  66. Costantini C, Renga G, Sellitto F, Borghi M, Stincardini C, Pariano M, Zelante T, Chiarotti F, Bartoli A, Mosci P, et al. (2020) Microbes in the era of circadian medicine. Front Cell Infect Microbiol 10:30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Cronin P, McCarthy MJ, Lim ASP, Salmon DP, Galasko D, Masliah E, De Jager PL, Bennett DA, Desplats P. (2017) Circadian alterations during early stages of Alzheimer’s disease are associated with aberrant cycles of DNA methylation in BMAL1. Alzheimers Dement 13:689-700. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Cugurra A, Mamuladze T, Rustenhoven J, Dykstra T, Beroshvili G, Greenberg ZJ, Baker W, Papadopoulos Z, Drieu A, Blackburn S, et al. (2021) Skull and vertebral bone marrow are myeloid cell reservoirs for the meninges and CNS parenchyma. Science 373:eabf7844. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Curran JA, Buhl E, Tsaneva-Atanasova K, Hodge JJL. (2019) Age-dependent changes in clock neuron structural plasticity and excitability are associated with a decrease in circadian output behavior and sleep. Neurobiol Aging 77:158-168. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Curtis AM, Bellet MM, Sassone-Corsi P, O’Neill LA. (2014) Circadian clock proteins and immunity. Immunity 40:178-186. [DOI] [PubMed] [Google Scholar]
  71. Curtis AM, Fagundes CT, Yang G, Palsson-McDermott EM, Wochal P, McGettrick AF, Foley NH, Early JO, Chen L, Zhang H, et al. (2015) Circadian control of innate immunity in macrophages by miR-155 targeting Bmal1. Proc Natl Acad Sci USA 112:7231-7236. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Czeisler CA, Dumont M, Duffy JF, Steinberg JD, Richardson GS, Brown EN, Sánchez R, Ríos CD, Ronda JM. (1992) Association of sleep-wake habits in older people with changes in output of circadian pacemaker. Lancet 340:933-936. [DOI] [PubMed] [Google Scholar]
  73. Dallmann R, Okyar A, Levi F. (2016) Dosing-time makes the poison: circadian regulation and pharmacotherapy. Trends Mol Med 22:430-445. [DOI] [PubMed] [Google Scholar]
  74. Dani N, Herbst RH, McCabe C, Green GS, Kaiser K, Head JP, Cui J, Shipley FB, Jang A, Dionne D, et al. (2021) A cellular and spatial map of the choroid plexus across brain ventricles and ages. Cell 184:3056-3074. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Dantzer R, O’Connor JC, Freund GG, Johnson RW, Kelley KW. (2008) From inflammation to sickness and depression: when the immune system subjugates the brain. Nat Rev Neurosci 9:46-56. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Daut RA, Fonken LK. (2019) Circadian regulation of depression: a role for serotonin. Front Neuroendocrinol 54:100746. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Davis FC, Gorski RA. (1985) Development of hamster circadian rhythms. I. Within-litter synchrony of mother and pup activity rhythms at weaning. Biol Reprod 33:353-362. [DOI] [PubMed] [Google Scholar]
  78. Davis FC, Gorski RA. (1988) Development of hamster circadian rhythms: role of the maternal suprachiasmatic nucleus. J Comp Physiol A 162:601-610. [DOI] [PubMed] [Google Scholar]
  79. Davis FC, Mannion J. (1988) Entrainment of hamster pup circadian rhythms by prenatal melatonin injections to the mother. Am J Physiol 255:R439-R448. [DOI] [PubMed] [Google Scholar]
  80. Davis FC, Boada R, LeDeaux J. (1990) Neurogenesis of the hamster suprachiasmatic nucleus. Brain Res 519:192-199. [DOI] [PubMed] [Google Scholar]
  81. Deng W, Zhu S, Zeng L, Liu J, Kang R, Yang M, Cao L, Wang H, Billiar TR, Jiang J, et al. (2018) The circadian clock controls immune checkpoint pathway in sepsis. Cell Rep 24:366-378. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Diflorio A, Jones I. (2010) Is sex important? Gender differences in bipolar disorder. Int Rev Psychiatry 22:437-452. [DOI] [PubMed] [Google Scholar]
  83. Dolatshad H, Cary AJ, Davis FC. (2010) Differential expression of the circadian clock in maternal and embryonic tissues of mice. PLoS ONE 5:e9855. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Druzd D, Matveeva O, Ince L, Harrison U, He W, Schmal C, Herzel H, Tsang AH, Kawakami N, Leliavski A, et al. (2017) Lymphocyte circadian clocks control lymph node trafficking and adaptive immune responses. Immunity 46:120-132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Duez H, Pourcet B. (2021) Nuclear receptors in the control of the NLRP3 inflammasome pathway. Front Endocrinol (Lausanne) 12:630536. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Duffy JF, Zeitzer JM, Rimmer DW, Klerman EB, Dijk DJ, Czeisler CA. (2002) Peak of circadian melatonin rhythm occurs later within the sleep of older subjects. Am J Physiol Endocrinol Metab 282:E297-E303. [DOI] [PubMed] [Google Scholar]
  87. Edgar DM, Dement WC, Fuller CA. (1993) Effect of SCN lesions on sleep in squirrel monkeys: evidence for opponent processes in sleep-wake regulation. J Neurosci 13:1065-1079. [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Edgar RS, Stangherlin A, Nagy AD, Nicoll MP, Efstathiou S, O’Neill JS, Reddy AB. (2016) Cell autonomous regulation of herpes and influenza virus infection by the circadian clock. Proc Natl Acad Sci USA 113:10085-10090. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Edlow AG, Glass RM, Smith CJ, Tran PK, James K, Bilbo S. (2019) Placental macrophages: a window into fetal microglial function in maternal obesity. Int J Dev Neurosci 77:60-68. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Ehlers A, Xie W, Agapov E, Brown S, Steinberg D, Tidwell R, Sajol G, Schutz R, Weaver R, Yu H, et al. (2018) BMAL1 links the circadian clock to viral airway pathology and asthma phenotypes. Mucosal Immunol 11:97-111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Erblich B, Zhu L, Etgen AM, Dobrenis K, Pollard JW. (2011) Absence of colony stimulation factor-1 receptor results in loss of microglia, disrupted brain development and olfactory deficits. PLoS ONE 6:e26317. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Erol A, Winham SJ, McElroy SL, Frye MA, Prieto ML, Cuellar-Barboza AB, Fuentes M, Geske J, Mori N, Biernacka JM, et al. (2015) Sex differences in the risk of rapid cycling and other indicators of adverse illness course in patients with bipolar I and II disorder. Bipolar Disord 17:670-676. [DOI] [PubMed] [Google Scholar]
  93. Erraji-Benchekroun L, Underwood MD, Arango V, Galfalvy H, Pavlidis P, Smyrniotopoulos P, Mann JJ, Sibille E. (2005) Molecular aging in human prefrontal cortex is selective and continuous throughout adult life. Biol Psychiatry 57:549-558. [DOI] [PubMed] [Google Scholar]
  94. Estes ML, McAllister AK. (2016) Maternal immune activation: implications for neuropsychiatric disorders. Science 353:772-777. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Etain B, Milhiet V, Bellivier F, Leboyer M. (2011) Genetics of circadian rhythms and mood spectrum disorders. Eur Neuropsychopharmacol 21:S676-682. [DOI] [PubMed] [Google Scholar]
  96. Fahrenkrug J, Nielsen HS, Hannibal J. (2004) Expression of melanopsin during development of the rat retina. Neuroreport 15:781-784. [DOI] [PubMed] [Google Scholar]
  97. Farajnia S, Michel S, Deboer T, vanderLeest HT, Houben T, Rohling JH, Ramkisoensing A, Yasenkov R, Meijer JH. (2012) Evidence for neuronal desynchrony in the aged suprachiasmatic nucleus clock. J Neurosci 32:5891-5899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Feigin RD, San Joaquin VH, Haymond MW, Wyatt RG. (1969) Daily periodicity of susceptibility of mice to pneumococcal infection. Nature 224:379-380. [DOI] [PubMed] [Google Scholar]
  99. Fernandez de Cossio L, Guzman A, van der Veldt S, Luheshi GN. (2017) Prenatal infection leads to ASD-like behavior and altered synaptic pruning in the mouse offspring. Brain Behav Immun 63:88-98. [DOI] [PubMed] [Google Scholar]
  100. Ferri SL, Abel T, Brodkin ES. (2018) Sex differences in autism spectrum disorder: a review. Curr Psychiatry Rep 20:9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Foley DJ, Monjan AA, Brown SL, Simonsick EM, Wallace RB, Blazer DG. (1995) Sleep complaints among elderly persons: an epidemiologic study of three communities. Sleep 18:425-432. [DOI] [PubMed] [Google Scholar]
  102. Fonken LK, Gaudet AD. (2022) Neuroimmunology of healthy brain aging. Curr Opin Neurobiol 77:102649. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Fonken LK, Frank MG, Kitt MM, Barrientos RM, Watkins LR, Maier SF. (2015) Microglia inflammatory responses are controlled by an intrinsic circadian clock. Brain Behav Immun 45:171-179. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Fonken LK, Kitt MM, Gaudet AD, Barrientos RM, Watkins LR, Maier SF. (2016) Diminished circadian rhythms in hippocampal microglia may contribute to age-related neuroinflammatory sensitization. Neurobiol Aging 47:102-112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Franklin AE, Engeland CG, Kavaliers M, Ossenkopp KP. (2003) Lipopolysaccharide-induced hypoactivity and behavioral tolerance development are modulated by the light-dark cycle in male and female rats. Psychopharmacology 170:399-408. [DOI] [PubMed] [Google Scholar]
  106. Gagnidze K, Hajdarovic KH, Moskalenko M, Karatsoreos IN, McEwen BS, Bulloch K. (2016) Nuclear receptor REV-ERBalpha mediates circadian sensitivity to mortality in murine vesicular stomatitis virus-induced encephalitis. Proc Natl Acad Sci USA 113:5730-5735. [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Gamez-Garcia A, Vazquez BN. (2021) Nuclear Sirtuins and the Aging of the Immune System. Genes 12:1856. [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Gan L, Cookson MR, Petrucelli L, La Spada AR. (2018) Converging pathways in neurodegeneration, from genetics to mechanisms. Nat Neurosci 21:1300-1309. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Gandal MJ, Zhang P, Hadjimichael E, Walker RL, Chen C, Liu S, Won H, van Bakel H, Varghese M, Wang Y, et al. (2018) Transcriptome-wide isoform-level dysregulation in ASD, schizophrenia, and bipolar disorder. Science 362:eaat8127. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Ganeshan K, Chawla A. (2014) Metabolic regulation of immune responses. Annu Rev Immunol 32:609-634. [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. Gibbs JE, Blaikley J, Beesley S, Matthews L, Simpson KD, Boyce SH, Farrow SN, Else KJ, Singh D, Ray DW, et al. (2012) The nuclear receptor REV-ERBalpha mediates circadian regulation of innate immunity through selective regulation of inflammatory cytokines. Proc Natl Acad Sci USA 109:582-587. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Ginhoux F, Greter M, Leboeuf M, Nandi S, See P, Gokhan S, Mehler MF, Conway SJ, Guan Ng L, Stanley ER, et al. (2010) Fate mapping analysis reveals that adult microglia derive from primitive macrophages. Science 330:841-845. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Glickman G. (2010) Circadian rhythms and sleep in children with autism. Neurosci Biobehav Rev 34:755-768. [DOI] [PubMed] [Google Scholar]
  114. Goines PE, Croen LA, Braunschweig D, Yoshida CK, Grether J, Hansen R, Kharrazi M, Ashwood P, Van de Water J. (2011) Increased midgestational IFN-gamma, IL-4 and IL-5 in women bearing a child with autism: a case-control study. Mol Autism 2:13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Goto M, Mizuno M, Matsumoto A, Yang Z, Jimbo EF, Tabata H, Yamagata T, Nagata KI. (2017) Role of a circadian-relevant gene NR1D1 in brain development: possible involvement in the pathophysiology of autism spectrum disorders. Sci Rep 7:43945. [DOI] [PMC free article] [PubMed] [Google Scholar]
  116. Grabert K, Michoel T, Karavolos MH, Clohisey S, Baillie JK, Stevens MP, Freeman TC, Summers KM, McColl BW. (2016) Microglial brain region-dependent diversity and selective regional sensitivities to aging. Nat Neurosci 19:504-516. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Greiner P, Houdek P, Sladek M, Sumova A. (2022) Early rhythmicity in the fetal suprachiasmatic nuclei in response to maternal signals detected by omics approach. PLoS Biol 20:e3001637. [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Griffin P, Dimitry JM, Sheehan PW, Lananna BV, Guo C, Robinette ML, Hayes ME, Cedeño MR, Nadarajah CJ, Ezerskiy LA, et al. (2019) Circadian clock protein Rev-erbalpha regulates neuroinflammation. Proc Natl Acad Sci USA 116:5102-5107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. Griffin P, Sheehan PW, Dimitry JM, Guo C, Kanan MF, Lee J, Zhang J, Musiek ES. (2020) REV-ERBalpha mediates complement expression and diurnal regulation of microglial synaptic phagocytosis. eLife 9:e58765. [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Guillaumond F, Dardente H, Giguere V, Cermakian N. (2005) Differential control of Bmal1 circadian transcription by REV-ERB and ROR nuclear receptors. J Biol Rhythms 20:391-403. [DOI] [PubMed] [Google Scholar]
  121. Hablitz LM, Pla V, Giannetto M, Vinitsky HS, Staeger FF, Metcalfe T, Nguyen R, Benrais A, Nedergaard M. (2020) Circadian control of brain glymphatic and lymphatic fluid flow. Nat Commun 11:4411. [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Hagberg H, Mallard C, Ferriero DM, Vannucci SJ, Levison SW, Vexler ZS, Gressens P. (2015) The role of inflammation in perinatal brain injury. Nat Rev Neurol 11:192-208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Haida O, Al Sagheer T, Balbous A, Francheteau M, Matas E, Soria F, Fernagut PO, Jaber M. (2019) Sex-dependent behavioral deficits and neuropathology in a maternal immune activation model of autism. Transl Psychiatry 9:124. [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. Halberg F, Johnson EA, Brown BW, Bittner JJ. (1960) Susceptibility rhythm to E. coli endotoxin and bioassay. Proc Soc Exp Biol Med 103:142-144. [DOI] [PubMed] [Google Scholar]
  125. Hampp G, Ripperger JA, Houben T, Schmutz I, Blex C, Perreau-Lenz S, Brunk I, Spanagel R, Ahnert-Hilger G, Meijer JH, et al. (2008) Regulation of monoamine oxidase A by circadian-clock components implies clock influence on mood. Curr Biol 18:678-683. [DOI] [PubMed] [Google Scholar]
  126. Harada K, Kamiya T, Tsuboi T. (2015) Gliotransmitter release from astrocytes: functional, developmental, and pathological implications in the brain. Front Neurosci 9:499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Harper DG, Stopa EG, Kuo-Leblanc V, McKee AC, Asayama K, Volicer L, Kowall N, Satlin A. (2008) Dorsomedial SCN neuronal subpopulations subserve different functions in human dementia. Brain 131:1609-1617. [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. Hastings MH, Brancaccio M, Gonzalez-Aponte MF, Herzog ED. (2023) Circadian rhythms and astrocytes: the good, the bad, and the ugly. Annu Rev Neurosci. doi: 10.1146/annurev-neuro-100322-112249. [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Hastings MH, Maywood ES, Brancaccio M. (2019) The mammalian circadian timing system and the suprachiasmatic nucleus as its pacemaker. Biology 8:13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Hatfield CF, Herbert J, van Someren EJ, Hodges JR, Hastings MH. (2004) Disrupted daily activity/rest cycles in relation to daily cortisol rhythms of home-dwelling patients with early Alzheimer’s dementia. Brain 127:1061-1074. [DOI] [PubMed] [Google Scholar]
  131. Hatori M, Vollmers C, Zarrinpar A, DiTacchio L, Bushong EA, Gill S, Leblanc M, Chaix A, Joens M, Fitzpatrick JAJ, et al. (2012) Time-restricted feeding without reducing caloric intake prevents metabolic diseases in mice fed a high-fat diet. Cell Metab 15:848-860. [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Hattar S, Liao HW, Takao M, Berson DM, Yau KW. (2002) Melanopsin-containing retinal ganglion cells: architecture, projections, and intrinsic photosensitivity. Science 295:1065-1070. [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Hattori Y. (2022) The behavior and functions of embryonic microglia. Anat Sci Int 97:1-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. Hayashi Y, Koyanagi S, Kusunose N, Okada R, Wu Z, Tozaki-Saitoh H, Ukai K, Kohsaka S, Inoue K, Ohdo S, et al. (2013) The intrinsic microglial molecular clock controls synaptic strength via the circadian expression of cathepsin S. Sci Rep 3:2744. [DOI] [PMC free article] [PubMed] [Google Scholar]
  135. Hazan G, Duek OA, Alapi H, Mok H, Ganninger AT, Ostendorf EM, Gierasch C, Chodick G, Greenberg D, Haspel JA. (2023) Biological rhythms in COVID-19 vaccine effectiveness in an observational cohort study of 1.5 million patients. J Clin Invest. 133(11):e167339. [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Hennah W, Varilo T, Kestila M, Paunio T, Arajarvi R, Haukka J, Parker A, Martin R, Levitzky S, Partonen T, et al. (2003) Haplotype transmission analysis provides evidence of association for DISC1 to schizophrenia and suggests sex-dependent effects. Hum Mol Genet 12:3151-3159. [DOI] [PubMed] [Google Scholar]
  137. Honzlova P, Semenovykh K, Sumova A. (2023) The circadian clock of polarized microglia and its interaction with mouse brain oscillators. Cell Mol Neurobiol 43:1319-1333. [DOI] [PMC free article] [PubMed] [Google Scholar]
  138. Hoshiko M, Arnoux I, Avignone E, Yamamoto N, Audinat E. (2012) Deficiency of the microglial receptor CX3CR1 impairs postnatal functional development of thalamocortical synapses in the barrel cortex. J Neurosci 32:15106-15111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  139. Hrushesky WJ, Langevin T, Kim YJ, Wood PA. (1994) Circadian dynamics of tumor necrosis factor alpha (cachectin) lethality. J Exp Med 180:1059-1065. [DOI] [PMC free article] [PubMed] [Google Scholar]
  140. Hu D, Mao Y, Xu G, Liao W, Ren J, Yang H, Yang J, Sun L, Chen H, Wang W, et al. (2019) Time-restricted feeding causes irreversible metabolic disorders and gut microbiota shift in pediatric mice. Pediatr Res 85:518-526. [DOI] [PMC free article] [PubMed] [Google Scholar]
  141. Hu VW, Sarachana T, Kim KS, Nguyen A, Kulkarni S, Steinberg ME, Lee NH. (2009) Gene expression profiling differentiates autism case-controls and phenotypic variants of autism spectrum disorders: evidence for circadian rhythm dysfunction in severe autism. Autism Res 2:78-97. [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Hughes ATL, Samuels RE, Bano-Otalora B, Belle MDC, Wegner S, Guilding C, Northeast RC, Irvine Loudon AS, Gigg J, Piggins HD. (2021) Timed daily exercise remodels circadian rhythms in mice. Commun Biol 4:761. [DOI] [PMC free article] [PubMed] [Google Scholar]
  143. Iadecola C. (2017) The neurovascular unit coming of age: a journey through neurovascular coupling in health and disease. Neuron 96:17-42. [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. Iliff JJ, Wang M, Liao Y, Plogg BA, Peng W, Gundersen GA, Benveniste H, Vates GE, Deane R, Goldman SA, et al. (2012) A paravascular pathway facilitates CSF flow through the brain parenchyma and the clearance of interstitial solutes, including amyloid beta. Sci Transl Med 4:147ra111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. Ince LM, Barnoud C, Lutes LK, Pick R, Wang C, Sinturel F, Chen C-S, de Juan A, Weber J, Holtkamp SJ, et al. (2023) Influence of circadian clocks on adaptive immunity and vaccination responses. Nat Commun 14:476. [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. Ince LM, Zhang Z, Beesley S, Vonslow RM, Saer BR, Matthews LC, Begley N, Gibbs JE, Ray DW, Loudon ASI. (2019) Circadian variation in pulmonary inflammatory responses is independent of rhythmic glucocorticoid signaling in airway epithelial cells. FASEB J 33:126-139. [DOI] [PMC free article] [PubMed] [Google Scholar]
  147. Ingiosi AM, Schoch H, Wintler T, Singletary KG, Righelli D, Roser LG, Peixoto L. (2019) Shank3 modulates sleep and expression of circadian transcription factors. eLife 8:e42819. [DOI] [PMC free article] [PubMed] [Google Scholar]
  148. Inokawa H, Umemura Y, Shimba A, Kawakami E, Koike N, Tsuchiya Y, Ohashi M, Minami Y, Cui G, Asahi T, et al. (2020) Chronic circadian misalignment accelerates immune senescence and abbreviates lifespan in mice. Sci Rep 10:2569. [DOI] [PMC free article] [PubMed] [Google Scholar]
  149. Ishibashi T, Dakin KA, Stevens B, Lee PR, Kozlov SV, Stewart CL, Fields RD. (2006) Astrocytes promote myelination in response to electrical impulses. Neuron 49:823-832. [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Jankovic J. (2008) Parkinson’s disease: clinical features and diagnosis. J Neurol Neurosurg Psychiatry 79:368-376. [DOI] [PubMed] [Google Scholar]
  151. Jawaid S, Kidd GJ, Wang J, Swetlik C, Dutta R, Trapp BD. (2018) Alterations in CA1 hippocampal synapses in a mouse model of fragile X syndrome. Glia 66:789-800. [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Jiang HY, Xu LL, Shao L, Xia RM, Yu ZH, Ling ZX, Yang F, Deng M, Ruan B. (2016) Maternal infection during pregnancy and risk of autism spectrum disorders: a systematic review and meta-analysis. Brain Behav Immun 58:165-172. [DOI] [PubMed] [Google Scholar]
  153. Johansson AS, Owe-Larsson B, Hetta J, Lundkvist GB. (2016) Altered circadian clock gene expression in patients with schizophrenia. Schizophr Res 174:17-23. [DOI] [PubMed] [Google Scholar]
  154. Jolley CC, Ukai-Tadenuma M, Perrin D, Ueda HR. (2014) A mammalian circadian clock model incorporating daytime expression elements. Biophys J 107:1462-1473. [DOI] [PMC free article] [PubMed] [Google Scholar]
  155. Jones KL, Croen LA, Yoshida CK, Heuer L, Hansen R, Zerbo O, DeLorenze GN, Kharrazi M, Yolken R, Ashwood P, Van de Water J. (2017) Autism with intellectual disability is associated with increased levels of maternal cytokines and chemokines during gestation. Mol Psychiatry 22:273-279. [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Kabrita CS, Davis FC. (2008) Development of the mouse suprachiasmatic nucleus: determination of time of cell origin and spatial arrangements within the nucleus. Brain Res 1195:20-27. [DOI] [PMC free article] [PubMed] [Google Scholar]
  157. Karch CM, Goate AM. (2015) Alzheimer’s disease risk genes and mechanisms of disease pathogenesis. Biol Psychiatry 77:43-51. [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Keller M, Mazuch J, Abraham U, Eom GD, Herzog ED, Volk HD, Kramer A, Maier B. (2009) A circadian clock in macrophages controls inflammatory immune responses. Proc Natl Acad Sci USA 106:21407-21412. [DOI] [PMC free article] [PubMed] [Google Scholar]
  159. Kim J, Jang S, Choi M, Chung S, Choe Y, Choe HK, Son GH, Rhee K, Kim K. (2018) Abrogation of the circadian nuclear receptor REV-ERBalpha exacerbates 6-hydroxydopamine-induced dopaminergic neurodegeneration. Mol Cells 41:742-752. [DOI] [PMC free article] [PubMed] [Google Scholar]
  160. Kim J, Park I, Jang S, Choi M, Kim D, Sun W, Choe Y, Choi J-W, Moon C, Park SH, et al. (2022) Pharmacological rescue with SR8278, a circadian nuclear receptor REV-ERBalpha antagonist as a therapy for mood disorders in parkinson’s disease. Neurotherapeutics 19:592-607. [DOI] [PMC free article] [PubMed] [Google Scholar]
  161. Knuesel I, Chicha L, Britschgi M, Schobel SA, Bodmer M, Hellings JA, Toovey S, Prinssen EP. (2014) Maternal immune activation and abnormal brain development across CNS disorders. Nat Rev Neurol 10:643-660. [DOI] [PubMed] [Google Scholar]
  162. Knutsson A, Hallquist J, Reuterwall C, Theorell T, Akerstedt T. (1999) Shiftwork and myocardial infarction: a case-control study. Occup Environ Med 56:46-50. [DOI] [PMC free article] [PubMed] [Google Scholar]
  163. Kopczynska M, Zelek W, Touchard S, Gaughran F, Di Forti M, Mondelli V, Murray R, O’Donovan MC, Morgan BP. (2019) Complement system biomarkers in first episode psychosis. Schizophr Res 204:16-22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Kripke DF, Nievergelt CM, Joo E, Shekhtman T, Kelsoe JR. (2009) Circadian polymorphisms associated with affective disorders. J Circadian Rhythms 7:2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  165. Kurabayashi N, Hirota T, Sakai M, Sanada K, Fukada Y. (2010) DYRK1A and glycogen synthase kinase 3beta, a dual-kinase mechanism directing proteasomal degradation of CRY2 for circadian timekeeping. Mol Cell Biol 30:1757-1768. [DOI] [PMC free article] [PubMed] [Google Scholar]
  166. Landgraf D, Koch CE, Oster H. (2014) Embryonic development of circadian clocks in the mammalian suprachiasmatic nuclei. Front. Neuroanat 8:143. [DOI] [PMC free article] [PubMed] [Google Scholar]
  167. Lang V, Ferencik S, Ananthasubramaniam B, Kramer A, Maier B. (2021) Susceptibility rhythm to bacterial endotoxin in myeloid clock-knockout mice. eLife 10:e62469. [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. Laudisi F, Spreafico R, Evrard M, Hughes TR, Mandriani B, Kandasamy M, Morgan BP, Sivasankar B, Mortellaro A. (2013) Cutting edge: the NLRP3 inflammasome links complement-mediated inflammation and IL-1beta release. J Immunol 191:1006-1010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  169. Lee AS, Azmitia EC, Whitaker-Azmitia PM. (2017) Developmental microglial priming in postmortem autism spectrum disorder temporal cortex. Brain Behav Immun 62:193-202. [DOI] [PubMed] [Google Scholar]
  170. Lee J, Kim DE, Griffin P, Sheehan PW, Kim DH, Musiek ES, Yoon SY. (2020) Inhibition of REV-ERBs stimulates microglial amyloid-beta clearance and reduces amyloid plaque deposition in the 5XFAD mouse model of Alzheimer’s disease. Aging Cell 19:e13078. [DOI] [PMC free article] [PubMed] [Google Scholar]
  171. Lee SB, Park J, Kwak Y, Park YU, Nhung TTM, Suh BK, Woo Y, Suh Y, Cho E, Cho S, et al. (2021) Disrupted-in-schizophrenia 1 enhances the quality of circadian rhythm by stabilizing BMAL1. Transl Psychiatry 11:110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  172. Leone MJ, Marpegan L, Bekinschtein TA, Costas MA, Golombek DA. (2006) Suprachiasmatic astrocytes as an interface for immune-circadian signalling. J Neurosci Res 84:1521-1527. [DOI] [PubMed] [Google Scholar]
  173. Li J, Lu WQ, Beesley S, Loudon AS, Meng QJ. (2012) Lithium impacts on the amplitude and period of the molecular circadian clockwork. PLoS ONE 7:e33292. [DOI] [PMC free article] [PubMed] [Google Scholar]
  174. Li J, Nie P, Turck CW, Wang GZ. (2022) Gene networks under circadian control exhibit diurnal organization in primate organs. Commun Biol 5:764. [DOI] [PMC free article] [PubMed] [Google Scholar]
  175. Li X, Davis FC. (2005) Developmental expression of clock genes in the Syrian hamster. Brain Res Dev Brain Res 158:31-40. [DOI] [PubMed] [Google Scholar]
  176. Li X, Chauhan A, Sheikh AM, Patil S, Chauhan V, Li XM, Ji L, Brown T, Malik M. (2009) Elevated immune response in the brain of autistic patients. J Neuroimmunol 207:111-116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  177. Lin SS, Tang Y, Illes P, Verkhratsky A. (2020) The safeguarding microglia: central role for P2Y12 receptors. Front Pharmacol 11:627760. [DOI] [PMC free article] [PubMed] [Google Scholar]
  178. Liu WW, Wei SZ, Huang GD, Liu LB, Gu C, Shen Y, Wang X-H, Xia S-T, Xie A-M, Hu L-F, Liu CF. (2020) BMAL1 regulation of microglia-mediated neuroinflammation in MPTP-induced Parkinson’s disease mouse model. FASEB J 34:6570-6581. [DOI] [PubMed] [Google Scholar]
  179. Long JE, Drayson MT, Taylor AE, Toellner KM, Lord JM, Phillips AC. (2016) Morning vaccination enhances antibody response over afternoon vaccination: a cluster-randomised trial. Vaccine 34:2679-2685. [DOI] [PMC free article] [PubMed] [Google Scholar]
  180. Loomes R, Hull L, Mandy WPL. (2017) What is the male-to-female ratio in autism spectrum disorder? A systematic review and meta-analysis. J Am Acad Child Adolesc Psychiatry 56:466-474. [DOI] [PubMed] [Google Scholar]
  181. Lunsford-Avery JR, Goncalves B, Brietzke E, Bressan RA, Gadelha A, Auerbach RP, Mittal VA. (2017) Adolescents at clinical-high risk for psychosis: circadian rhythm disturbances predict worsened prognosis at 1-year follow-up. Schizophr Res 189:37-42. [DOI] [PMC free article] [PubMed] [Google Scholar]
  182. Lyall LM, Wyse CA, Graham N, Ferguson A, Lyall DM, Cullen B, Celis Morales CA, Biello SM, Mackay D, Ward J, et al. (2018) Association of disrupted circadian rhythmicity with mood disorders, subjective wellbeing, and cognitive function: a cross-sectional study of 91 105 participants from the UK Biobank. Lancet Psychiatry 5:507-514. [DOI] [PubMed] [Google Scholar]
  183. Ma Z, Eaton M, Liu Y, Zhang J, Chen X, Tu X, Shi Y, Que Z, Wettschurack K, Zhang Z, et al. (2022) Deficiency of autism-related Scn2a gene in mice disrupts sleep patterns and circadian rhythms. Neurobiol Dis 168:105690. [DOI] [PMC free article] [PubMed] [Google Scholar]
  184. McNeill DS, Sheely CJ, Ecker JL, Badea TC, Morhardt D, Guido W, Hattar S. (2011) Development of melanopsin-based irradiance detecting circuitry. Neural Dev 6:8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  185. Marsters CM, Nesan D, Far R, Klenin N, Pittman QJ, Kurrasch DM. (2020) Embryonic microglia influence developing hypothalamic glial populations. J Neuroinflammation 17:146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  186. Martinek S, Inonog S, Manoukian AS, Young MW. (2001) A role for the segment polarity gene shaggy/GSK-3 in the Drosophila circadian clock. Cell 105:769-779. [DOI] [PubMed] [Google Scholar]
  187. Martinez A, Castro A, Dorronsoro I, Alonso M. (2002) Glycogen synthase kinase 3 (GSK-3) inhibitors as new promising drugs for diabetes, neurodegeneration, cancer, and inflammation. Med Res Rev 22:373-384. [DOI] [PubMed] [Google Scholar]
  188. Meeker RB, Williams K, Killebrew DA, Hudson LC. (2012) Cell trafficking through the choroid plexus. Cell Adh Migr 6:390-396. [DOI] [PMC free article] [PubMed] [Google Scholar]
  189. Melke J, Goubran Botros H, Chaste P, Betancur C, Nygren G, Anckarsater H, Rastam M, Stahlberg O, Gillberg IC, Delorme R, et al. (2008) Abnormal melatonin synthesis in autism spectrum disorders. Mol Psychiatry 13:90-98. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Meng J, Liu Y, Xie Z, Qing H, Lei P, Ni J. (2020) Nucleus distribution of cathepsin B in senescent microglia promotes brain aging through degradation of sirtuins. Neurobiol Aging 96:255-266. [DOI] [PubMed] [Google Scholar]
  191. Mestre H, Hablitz LM, Xavier AL, Feng W, Zou W, Pu T, Monai H, Murlidharan G, Castellanos Rivera RM, Simon MJ, et al. (2018) Aquaporin-4-dependent glymphatic solute transport in the rodent brain. eLife 7:e40070. [DOI] [PMC free article] [PubMed] [Google Scholar]
  192. Meyer U, Feldon J. (2012) To poly(I:C) or not to poly(I:C): advancing preclinical schizophrenia research through the use of prenatal immune activation models. Neuropharmacology 62:1308-1321. [DOI] [PubMed] [Google Scholar]
  193. Missig G, Robbins JO, Mokler EL, McCullough KM, Bilbo SD, McDougle CJ, Carlezon WA., Jr (2020) Sex-dependent neurobiological features of prenatal immune activation via TLR7. Mol Psychiatry 25:2330-2341. [DOI] [PMC free article] [PubMed] [Google Scholar]
  194. Mistlberger RE, Antle M, Glass J, Miller J. (2000) Behavioral and serotonergic regulation of circadian rhythms. Biological Rhythm Research 31:240-283. [Google Scholar]
  195. Moran M, Lynch CA, Walsh C, Coen R, Coakley D, Lawlor BA. (2005) Sleep disturbance in mild to moderate Alzheimer’s disease. Sleep Med 6:347-352. [DOI] [PubMed] [Google Scholar]
  196. Mosser CA, Baptista S, Arnoux I, Audinat E. (2017) Microglia in CNS development: shaping the brain for the future. Prog Neurobiol 149-150:1-20. [DOI] [PubMed] [Google Scholar]
  197. Mukherjee S, Coque L, Cao JL, Kumar J, Chakravarty S, Asaithamby A, Graham A, Gordon E, Enwright JF, DiLeone RJ, et al. (2010) Knockdown of Clock in the ventral tegmental area through RNA interference results in a mixed state of mania and depression-like behavior. Biol Psychiatry 68:503-511. [DOI] [PMC free article] [PubMed] [Google Scholar]
  198. Mullegama SV, Pugliesi L, Burns B, Shah Z, Tahir R, Gu Y, Nelson DL, Elsea SH. (2015) MBD5 haploinsufficiency is associated with sleep disturbance and disrupts circadian pathways common to Smith-Magenis and fragile X syndromes. Eur J Hum Genet 23:781-789. [DOI] [PMC free article] [PubMed] [Google Scholar]
  199. Munekawa K, Tamada Y, Iijima N, Hayashi S, Ishihara A, Inoue K, Tanaka M, Ibata Y. (2000) Development of astroglial elements in the suprachiasmatic nucleus of the rat: with special reference to the involvement of the optic nerve. Exp Neurol 166:44-51. [DOI] [PubMed] [Google Scholar]
  200. Mure LS, Le HD, Benegiamo G, Chang MW, Rios L, Jillani N, Ngotho M, Kariuki T, Dkhissi-Benyahya O, Cooper HM, et al. (2018) Diurnal transcriptome atlas of a primate across major neural and peripheral tissues. Science 359:eaao0318. [DOI] [PMC free article] [PubMed] [Google Scholar]
  201. Musiek ES, Bhimasani M, Zangrilli MA, Morris JC, Holtzman DM, Ju YS. (2018) Circadian rest-activity pattern changes in aging and preclinical Alzheimer disease. JAMA Neurol 75:582-590. [DOI] [PMC free article] [PubMed] [Google Scholar]
  202. Nakanishi H, Ni J, Nonaka S, Hayashi Y. (2021) Microglial circadian clock regulation of microglial structural complexity, dendritic spine density and inflammatory response. Neurochem Int 142:104905. [DOI] [PubMed] [Google Scholar]
  203. Nguyen A, Rauch TA, Pfeifer GP, Hu VW. (2010) Global methylation profiling of lymphoblastoid cell lines reveals epigenetic contributions to autism spectrum disorders and a novel autism candidate gene, RORA, whose protein product is reduced in autistic brain. FASEB J 24:3036-3051. [DOI] [PMC free article] [PubMed] [Google Scholar]
  204. Ni J, Wu Z, Meng J, Saito T, Saido TC, Qing H, Nakanishi H. (2019) An impaired intrinsic microglial clock system induces neuroinflammatory alterations in the early stage of amyloid precursor protein knock-in mouse brain. J Neuroinflammation 16:173. [DOI] [PMC free article] [PubMed] [Google Scholar]
  205. Noguchi T, Leise TL, Kingsbury NJ, Diemer T, Wang LL, Henson MA, Welsh DK. (2017) Calcium circadian rhythmicity in the suprachiasmatic nucleus: cell autonomy and network modulation. eNeuro 4:ENEURO.0160-17.2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  206. Northcutt AL, Hutchinson MR, Wang X, Baratta MV, Hiranita T, Cochran TA, Pomrenze MB, Galer EL, Kopajtic TA, Li CM, et al. (2015) DAT isn’t all that: cocaine reward and reinforcement require Toll-like receptor 4 signaling. Mol Psychiatry 20:1525-1537. [DOI] [PMC free article] [PubMed] [Google Scholar]
  207. Novakova M, Sladek M, Sumova A. (2010) Exposure of pregnant rats to restricted feeding schedule synchronizes the SCN clocks of their fetuses under constant light but not under a light-dark regime. J Biol Rhythms 25:350-360. [DOI] [PubMed] [Google Scholar]
  208. O’Brien SM, Scully P, Scott LV, Dinan TG. (2006) Cytokine profiles in bipolar affective disorder: focus on acutely ill patients. J Affect Disord 90:263-267. [DOI] [PubMed] [Google Scholar]
  209. Ogomori K, Kitamoto T, Tateishi J, Sato Y, Suetsugu M, Abe M. (1989) Beta-protein amyloid is widely distributed in the central nervous system of patients with Alzheimer’s disease. Am J Pathol 134:243-251. [PMC free article] [PubMed] [Google Scholar]
  210. Ohta H, Mitchell AC, McMahon DG. (2006) Constant light disrupts the developing mouse biological clock. Pediatr Res 60:304-308. [DOI] [PubMed] [Google Scholar]
  211. Olejnikova L, Polidarova L, Behuliak M, Sladek M, Sumova A. (2018) Circadian alignment in a foster mother improves the offspring’s pathological phenotype. J Physiol 596:5757-5775. [DOI] [PMC free article] [PubMed] [Google Scholar]
  212. Ortinski PI, Reissner KJ, Turner J, Anderson TA, Scimemi A. (2022) Control of complex behavior by astrocytes and microglia. Neurosci Biobehav Rev 137:104651. [DOI] [PMC free article] [PubMed] [Google Scholar]
  213. Otsu Y, Couchman K, Lyons DG, Collot M, Agarwal A, Mallet JM, Pfrieger FW, Bergles DE, Charpak S. (2015) Calcium dynamics in astrocyte processes during neurovascular coupling. Nat Neurosci 18:210-218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  214. Paci P, Fiscon G, Conte F, Wang R-S, Farina L, Loscalzo J. (2021) Gene co-expression in the interactome: moving from correlation toward causation via an integrated approach to disease module discovery. NPJ Syst Biology Appl. 7(1):3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  215. Pacifico R, Davis RL. (2017) Transcriptome sequencing implicates dorsal striatum-specific gene network, immune response and energy metabolism pathways in bipolar disorder. Mol Psychiatry 22:441-449. [DOI] [PubMed] [Google Scholar]
  216. Paolicelli RC, Bolasco G, Pagani F, Maggi L, Scianni M, Panzanelli P, Giustetto M, Ferreira TA, Guiducci E, Dumas L, et al. (2011) Synaptic pruning by microglia is necessary for normal brain development. Science 333:1456-1458. [DOI] [PubMed] [Google Scholar]
  217. Parboosing R, Bao Y, Shen L, Schaefer CA, Brown AS. (2013) Gestational influenza and bipolar disorder in adult offspring. JAMA Psychiatry 70:677-685. [DOI] [PubMed] [Google Scholar]
  218. Pariollaud M, Gibbs JE, Hopwood TW, Brown S, Begley N, Vonslow R, Poolman T, Guo B, Saer B, Jones DH, et al. (2018) Circadian clock component REV-ERBalpha controls homeostatic regulation of pulmonary inflammation. J Clin Invest 128:2281-2296. [DOI] [PMC free article] [PubMed] [Google Scholar]
  219. Parrott JM, Oster T, Lee HY. (2021) Altered inflammatory response in FMRP-deficient microglia. iScience 24:103293. [DOI] [PMC free article] [PubMed] [Google Scholar]
  220. Patke A, Young MW, Axelrod S. (2020) Molecular mechanisms and physiological importance of circadian rhythms. Nat. Rev Mol Cell Biol 21:67-84. [DOI] [PubMed] [Google Scholar]
  221. Perdiguero EG, Geissmann F. (2016) The development and maintenance of resident macrophages. Nat Immunol 17:2-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  222. Politis M, Pavese N, Tai YF, Kiferle L, Mason SL, Brooks DJ, Tabrizi SJ, Barker RA, Piccini P. (2011) Microglial activation in regions related to cognitive function predicts disease onset in Huntington’s disease: a multimodal imaging study. Hum Brain Mapp 32:258-270. [DOI] [PMC free article] [PubMed] [Google Scholar]
  223. Pourcet B, Zecchin M, Ferri L, Beauchamp J, Sitaula S, Billon C, Delhaye S, Vanhoutte J, Mayeuf-Louchart A, Thorel Q, et al. (2018) Nuclear receptor subfamily 1 group D member 1 regulates circadian activity of NLRP3 inflammasome to reduce the severity of fulminant hepatitis in mice. Gastroenterology 154:1449-1464. [DOI] [PMC free article] [PubMed] [Google Scholar]
  224. Prasad KM, Burgess AM, Keshavan MS, Nimgaonkar VL, Stanley JA. (2016) Neuropil pruning in early-course schizophrenia: immunological, clinical, and neurocognitive correlates. Biol Psychiatry Cogn Neurosci Neuroimaging 1:528-538. [DOI] [PMC free article] [PubMed] [Google Scholar]
  225. Prates KV, Pavanello A, Gongora AB, Moreira VM, de Moraes AMP, Rigo KP, Vieira E, Mathias PCF. (2022) Time-restricted feeding during embryonic development leads to metabolic dysfunction in adult rat offspring. Nutrition 103-104:111776. [DOI] [PubMed] [Google Scholar]
  226. Prendergast BJ, Cable EJ, Stevenson TJ, Onishi KG, Zucker I, Kay LM. (2015) Circadian disruption alters the effects of lipopolysaccharide treatment on circadian and ultradian locomotor activity and body temperature rhythms of female siberian hamsters. J Biol Rhythms 30:543-556. [DOI] [PMC free article] [PubMed] [Google Scholar]
  227. Prince TM, Abel T. (2013) The impact of sleep loss on hippocampal function. Learn Mem 20:558-569. [DOI] [PMC free article] [PubMed] [Google Scholar]
  228. Ramanathan C, Kathale ND, Liu D, Lee C, Freeman DA, Hogenesch JB, Cao R, Liu AC. (2018) mTOR signaling regulates central and peripheral circadian clock function. PLoS Genet 14:e1007369. [DOI] [PMC free article] [PubMed] [Google Scholar]
  229. Rao JS, Harry GJ, Rapoport SI, Kim HW. (2010) Increased excitotoxicity and neuroinflammatory markers in postmortem frontal cortex from bipolar disorder patients. Mol Psychiatry 15:384-392. [DOI] [PMC free article] [PubMed] [Google Scholar]
  230. Reemst K, Noctor SC, Lucassen PJ, Hol EM. (2016) The indispensable roles of microglia and astrocytes during brain development. Front Hum Neurosci 10:566. [DOI] [PMC free article] [PubMed] [Google Scholar]
  231. Rees E, Creeth HDJ, Hwu HG, Chen WJ, Tsuang M, Glatt SJ, Rey R, Kirov G, Walters JTR, Holmans P, et al. (2021) Schizophrenia, autism spectrum disorders and developmental disorders share specific disruptive coding mutations. Nat Commun 12:5353. [DOI] [PMC free article] [PubMed] [Google Scholar]
  232. Reghunandanan V, Reghunandanan R. (2006) Neurotransmitters of the suprachiasmatic nuclei. J Circadian Rhythms 4:2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  233. Reppert SM, Schwartz WJ. (1983) Maternal coordination of the fetal biological clock in utero. Science 220:969-971. [DOI] [PubMed] [Google Scholar]
  234. Reppert SM, Schwartz WJ. (1986) Maternal suprachiasmatic nuclei are necessary for maternal coordination of the developing circadian system. J Neurosci 6:2724-2729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  235. Reppert SM, Coleman RJ, Heath HW, Swedlow JR. (1984) Pineal N-acetyltransferase activity in 10-day-old rats: a paradigm for studying the developing circadian system. Endocrinology 115:918-925. [DOI] [PubMed] [Google Scholar]
  236. Richter JD, Zhao X. (2021) The molecular biology of FMRP: new insights into fragile X syndrome. Nat Rev Neurosci 22:209-222. [DOI] [PMC free article] [PubMed] [Google Scholar]
  237. Roberts DE, Killiany RJ, Rosene DL. (2012) Neuron numbers in the hypothalamus of the normal aging rhesus monkey: stability across the adult lifespan and between the sexes. J Comp Neurol 520:1181-1197. [DOI] [PMC free article] [PubMed] [Google Scholar]
  238. Rose DR, Careaga M, Van de Water J, McAllister K, Bauman MD, Ashwood P. (2017) Long-term altered immune responses following fetal priming in a non-human primate model of maternal immune activation. Brain Behav Immun 63:60-70. [DOI] [PMC free article] [PubMed] [Google Scholar]
  239. Rosin JM, Vora SR, Kurrasch DM. (2018) Depletion of embryonic microglia using the CSF1R inhibitor PLX5622 has adverse sex-specific effects on mice, including accelerated weight gain, hyperactivity and anxiolytic-like behaviour. Brain Behav Immun 73:682-697. [DOI] [PubMed] [Google Scholar]
  240. Rossignol DA, Frye RE. (2011) Melatonin in autism spectrum disorders: a systematic review and meta-analysis. Dev Med Child Neurol 53:783-792. [DOI] [PubMed] [Google Scholar]
  241. Roybal K, Theobold D, Graham A, DiNieri JA, Russo SJ, Krishnan V, Chakravarty S, Peevey J, Oehrlein N, Birnbaum S, et al. (2007) Mania-like behavior induced by disruption of CLOCK. Proc Natl Acad Sci USA 104:6406-6411. [DOI] [PMC free article] [PubMed] [Google Scholar]
  242. Ruan W, Yuan X, Eltzschig HK. (2021) Circadian rhythm as a therapeutic target. Nat Rev Drug Discov 20:287-307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  243. Ruben MD, Wu G, Smith DF, Schmidt RE, Francey LJ, Lee YY, Anafi RC, Hogenesch JB. (2018) A database of tissue-specific rhythmically expressed human genes has potential applications in circadian medicine. Sci Transl Med 10:eaat8806. [DOI] [PMC free article] [PubMed] [Google Scholar]
  244. Ruiz FS, Rosa DS, Zimberg IZ, Dos Santos Quaresma MV, Nunes JO, Apostolico JS, Weckx LY, Souza AR, Narciso FV, Fernandes-Junior SA, et al. (2020) Night shift work and immune response to the meningococcal conjugate vaccine in healthy workers: a proof of concept study. Sleep Med 75:263-275. [DOI] [PubMed] [Google Scholar]
  245. Rustenhoven J, Kipnis J. (2022) Brain borders at the central stage of neuroimmunology. Nature 612:417-429. [DOI] [PMC free article] [PubMed] [Google Scholar]
  246. Sakamoto K, Oishi K, Nagase T, Miyazaki K, Ishida N. (2002) Circadian expression of clock genes during ontogeny in the rat heart. Neuroreport 13:1239-1242. [DOI] [PubMed] [Google Scholar]
  247. Salvatore P, Ghidini S, Zita G, De Panfilis C, Lambertino S, Maggini C, Baldessarini RJ. (2008) Circadian activity rhythm abnormalities in ill and recovered bipolar I disorder patients. Bipolar Disord 10:256-265. [DOI] [PubMed] [Google Scholar]
  248. Santos J, Araújo J, Cunha M, Costa S, Barbosa A, Mesquita J, Costa M. (2005) Circadian variation in GFAP immunoreactivity in the mouse suprachiasmatic nucleus. Biological Rhythm Research 36:141-150. [Google Scholar]
  249. Satoh A, Brace CS, Rensing N, Cliften P, Wozniak DF, Herzog ED, Yamada KA, Imai S. (2013) Sirt1 extends life span and delays aging in mice through the regulation of Nk2 homeobox 1 in the DMH and LH. Cell Metab 18:416-430. [DOI] [PMC free article] [PubMed] [Google Scholar]
  250. Satterstrom FK, Kosmicki JA, Wang J, Breen MS, De Rubeis S, An JY, Peng M, Collins R, Grove J, Klei L, et al. (2020) Large-scale exome sequencing study implicates both developmental and functional changes in the neurobiology of autism. Cell 180:568-584. [DOI] [PMC free article] [PubMed] [Google Scholar]
  251. Sawicka K, Hale CR, Park CY, Fak JJ, Gresack JE, Van Driesche SJ, Kang JJ, Darnell JC, Darnell RB. (2019) FMRP has a cell-type-specific role in CA1 pyramidal neurons to regulate autism-related transcripts and circadian memory. eLife 8:e46919. [DOI] [PMC free article] [PubMed] [Google Scholar]
  252. Saxena MT, Aton SJ, Hildebolt C, Prior JL, Abraham U, Piwnica-Worms D, Herzog ED. (2007) Bioluminescence imaging of period1 gene expression in utero. Mol Imaging 6:68-72. [PubMed] [Google Scholar]
  253. Schafer DP, Lehrman EK, Kautzman AG, Koyama R, Mardinly AR, Yamasaki R, Ransohoff RM, Greenberg ME, Barres BA, Stevens B. (2012) Microglia sculpt postnatal neural circuits in an activity and complement-dependent manner. Neuron 74:691-705. [DOI] [PMC free article] [PubMed] [Google Scholar]
  254. Scheiermann C, Gibbs J, Ince L, Loudon A. (2018) Clocking in to immunity. Nat Rev Immunol 18:423-437. [DOI] [PubMed] [Google Scholar]
  255. Scheiermann C, Kunisaki Y, Frenette PS. (2013) Circadian control of the immune system. Nat Rev Immunol 13:190-198. [DOI] [PMC free article] [PubMed] [Google Scholar]
  256. Schousboe A, Waagepetersen HS. (2005) Role of astrocytes in glutamate homeostasis: implications for excitotoxicity. Neurotox Res 8:221-225. [DOI] [PubMed] [Google Scholar]
  257. Sekar A, Bialas AR, de Rivera H, Davis A, Hammond TR, Kamitaki N, Tooley K, Presumey J, Baum M, Van Doren V, et al. (2016) Schizophrenia risk from complex variation of complement component 4. Nature 530:177-183. [DOI] [PMC free article] [PubMed] [Google Scholar]
  258. Sellgren CM, Gracias J, Watmuff B, Biag JD, Thanos JM, Whittredge PB, Fu T, Worringer K, Brown HE, Wang J, et al. (2019) Increased synapse elimination by microglia in schizophrenia patient-derived models of synaptic pruning. Nat Neurosci 22:374-385. [DOI] [PMC free article] [PubMed] [Google Scholar]
  259. Seney ML, Cahill K, Enwright JF, 3rd, Logan RW, Huo Z, Zong W, Tseng G, McClung CA. (2019) Diurnal rhythms in gene expression in the prefrontal cortex in schizophrenia. Nat Commun 10:3355. [DOI] [PMC free article] [PubMed] [Google Scholar]
  260. Severino G, Manchia M, Contu P, Squassina A, Lampus S, Ardau R, Chillotti C, Del Zompo M. (2009) Association study in a Sardinian sample between bipolar disorder and the nuclear receptor REV-ERBalpha gene, a critical component of the circadian clock system. Bipolar Disord 11:215-220. [DOI] [PubMed] [Google Scholar]
  261. Shearman LP, Zylka MJ, Weaver DR, Kolakowski LF, Jr, Reppert SM. (1997) Two period homologs: circadian expression and photic regulation in the suprachiasmatic nuclei. Neuron 19:1261-1269. [DOI] [PubMed] [Google Scholar]
  262. Shimomura H, Moriya T, Sudo M, Wakamatsu H, Akiyama M, Miyake Y, Shibata S. (2001) Differential daily expression of Per1 and Per2 mRNA in the suprachiasmatic nucleus of fetal and early postnatal mice. Eur J Neurosci 13:687-693. [DOI] [PubMed] [Google Scholar]
  263. Shokri-Kojori E, Wang GJ, Wiers CE, Demiral SB, Guo M, Kim SW, Lindgren E, Ramirez V, Zehra A, Freeman C, et al. (2018) Beta-Amyloid accumulation in the human brain after one night of sleep deprivation. Proc Natl Acad Sci USA 115:4483-4488. [DOI] [PMC free article] [PubMed] [Google Scholar]
  264. Silver AC, Arjona A, Hughes ME, Nitabach MN, Fikrig E. (2012) Circadian expression of clock genes in mouse macrophages, dendritic cells, and B cells. Brain Behav Immun 26:407-413. [DOI] [PMC free article] [PubMed] [Google Scholar]
  265. Sladek M, Jindrakova Z, Bendova Z, Sumova A. (2007) Postnatal ontogenesis of the circadian clock within the rat liver. Am J Physiol Regul Integr Comp Physiol 292:R1224-R1229. [DOI] [PubMed] [Google Scholar]
  266. Sladek M, Sumova A, Kovacikova Z, Bendova Z, Laurinova K, Illnerova H. (2004) Insight into molecular core clock mechanism of embryonic and early postnatal rat suprachiasmatic nucleus. Proc Natl Acad Sci USA 101:6231-6236. [DOI] [PMC free article] [PubMed] [Google Scholar]
  267. Smarr BL, Grant AD, Perez L, Zucker I, Kriegsfeld LJ. (2017) Maternal and early-life circadian disruption have long-lasting negative consequences on offspring development and adult behavior in mice. Sci Rep 7:3326. [DOI] [PMC free article] [PubMed] [Google Scholar]
  268. Smolensky MH, Reinberg A, Labrecque G. (1995) Twenty-four hour pattern in symptom intensity of viral and allergic rhinitis: treatment implications. J Allergy Clin Immunol 95:1084-1096. [DOI] [PMC free article] [PubMed] [Google Scholar]
  269. Sofroniew MV. (2014) Astrogliosis. Cold Spring Harb Perspect Biol 7:a020420. [DOI] [PMC free article] [PubMed] [Google Scholar]
  270. Sominsky L, Dangel T, Malik S, De Luca SN, Singewald N, Spencer SJ. (2021) Microglial ablation in rats disrupts the circadian system. FASEB J 35:e21195. [DOI] [PubMed] [Google Scholar]
  271. Sommer IE, Tiihonen J, van Mourik A, Tanskanen A, Taipale H. (2020) The clinical course of schizophrenia in women and men-a nation-wide cohort study. NPJ Schizophr 6:12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  272. Soreq L, UK Brain Expression Consortium, North American Brain Expression Consortium, Rose J, Soreq E, Hardy J, Trabzuni D, Cookson MR, Smith C, Ryten M, Patani R, et al. (2017) Major shifts in glial regional identity are a transcriptional hallmark of human brain aging. Cell Rep 18:557-570. [DOI] [PMC free article] [PubMed] [Google Scholar]
  273. Squarzoni P, Oller G, Hoeffel G, Pont-Lezica L, Rostaing P, Low D, Bessis A, Ginhoux F, Garel S. (2014) Microglia modulate wiring of the embryonic forebrain. Cell Rep 8:1271-1279. [DOI] [PubMed] [Google Scholar]
  274. St Clair D, Blackwood D, Muir W, Carothers A, Walker M, Spowart G, Gosden C, Evans HJ. (1990) Association within a family of a balanced autosomal translocation with major mental illness. Lancet 336:13-16. [DOI] [PubMed] [Google Scholar]
  275. Stefani A, Hogl B. (2020) Sleep in Parkinson’s disease. Neuropsychopharmacology 45:121-128. [DOI] [PMC free article] [PubMed] [Google Scholar]
  276. Stefansson H, Ophoff RA, Steinberg S, Andreassen OA, Cichon S, Rujescu D, Werge T, Pietiläinen OPH, Mors O, Mortensen PB, et al. (2009) Common variants conferring risk of schizophrenia. Nature 460:744-747. [DOI] [PMC free article] [PubMed] [Google Scholar]
  277. Stessman HA, Xiong B, Coe BP, Wang T, Hoekzema K, Fenckova M, Kvarnung M, Gerdts J, Trinh S, Cosemans N, et al. (2017) Targeted sequencing identifies 91 neurodevelopmental-disorder risk genes with autism and developmental-disability biases. Nat Genet 49:515-526. [DOI] [PMC free article] [PubMed] [Google Scholar]
  278. Stevens B, Allen NJ, Vazquez LE, Howell GR, Christopherson KS, Nouri N, Micheva KD, Mehalow AK, Huberman AD, Stafford B, et al. (2007) The classical complement cascade mediates CNS synapse elimination. Cell 131:1164-1178. [DOI] [PubMed] [Google Scholar]
  279. Stopa EG, Volicer L, Kuo-Leblanc V, Harper D, Lathi D, Tate B, Satlin A. (1999) Pathologic evaluation of the human suprachiasmatic nucleus in severe dementia. J Neuropathol Exp Neurol 58:29-39. [DOI] [PubMed] [Google Scholar]
  280. Storch KF, Lipan O, Leykin I, Viswanathan N, Davis FC, Wong WH, Weitz CJ. (2002) Extensive and divergent circadian gene expression in liver and heart. Nature 417:78-83. [DOI] [PubMed] [Google Scholar]
  281. Straif K, Baan R, Grosse Y, Secretan B, El Ghissassi F, Bouvard V, Altieri A, Benbrahim-Tallaa L, Cogliano V, WHO International Agency For Research on Cancer Monograph Working Group (2007) Carcinogenicity of shift-work, painting, and fire-fighting. Lancet Oncol 8:1065-1066. [DOI] [PubMed] [Google Scholar]
  282. Suzuki K, Sugihara G, Ouchi Y, Nakamura K, Futatsubashi M, Takebayashi K, Yoshihara Y, Omata K, Matsumoto K, Tsuchiya KJ, et al. (2013) Microglial activation in young adults with autism spectrum disorder. JAMA Psychiatry 70:49-58. [DOI] [PubMed] [Google Scholar]
  283. Tahara Y, Aoyama S, Shibata S. (2017) The mammalian circadian clock and its entrainment by stress and exercise. J Physiol Sci 67:1-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  284. Taishi P, Bredow S, Guha-Thakurta N, Obal F, Jr, Krueger JM. (1997) Diurnal variations of interleukin-1 beta mRNA and beta-actin mRNA in rat brain. J Neuroimmunol 75:69-74. [DOI] [PubMed] [Google Scholar]
  285. Takahashi K, Deguchi T. (1983) Entrainment of the circadian rhythms of blinded infant rats by nursing mothers. Physiol Behav 31:373-378. [DOI] [PubMed] [Google Scholar]
  286. Takao T, Tachikawa H, Kawanishi Y, Mizukami K, Asada T. (2007) CLOCK gene T3111C polymorphism is associated with Japanese schizophrenics: a preliminary study. Eur Neuropsychopharmacol 17:273-276. [DOI] [PubMed] [Google Scholar]
  287. Takayama F, Hayashi Y, Wu Z, Liu Y, Nakanishi H. (2016) Diurnal dynamic behavior of microglia in response to infected bacteria through the UDP-P2Y6 receptor system. Sci Rep 6:30006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  288. Takayama F, Zhang X, Hayashi Y, Wu Z, Nakanishi H. (2017) Dysfunction in diurnal synaptic responses and social behavior abnormalities in cathepsin S-deficient mice. Biochem Biophys Res Commun 490:447-452. [DOI] [PubMed] [Google Scholar]
  289. Tandon R, Keshavan MS, Nasrallah HA. (2008) Schizophrenia, “just the facts” what we know in 2008. 2. Epidemiology and etiology. Schizophr Res 102:1-18. [DOI] [PubMed] [Google Scholar]
  290. Tao L, Jiang R, Zhang K, Qian Z, Chen P, Lv Y, Yao Y. (2020) Light therapy in non-seasonal depression: an update meta-analysis. Psychiatry Res 291:113247. [DOI] [PubMed] [Google Scholar]
  291. Thion MS, Garel S. (2017) On place and time: microglia in embryonic and perinatal brain development. Curr Opin Neurobiol 47:121-130. [DOI] [PubMed] [Google Scholar]
  292. Toiber D, Sebastian C, Mostoslavsky R. (2011) Characterization of nuclear sirtuins: molecular mechanisms and physiological relevance. Handb Exp Pharmacol 206:189-224. [DOI] [PubMed] [Google Scholar]
  293. Torquati L, Mielke GI, Brown WJ, Burton NW, Kolbe-Alexander TL. (2019) Shift work and poor mental health: a meta-analysis of longitudinal studies. Am J Public Health 109:e13-e20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  294. Tso CF, Simon T, Greenlaw AC, Puri T, Mieda M, Herzog ED. (2017) Astrocytes regulate daily rhythms in the suprachiasmatic nucleus and behavior. Curr Biol 27:1055-1061. [DOI] [PMC free article] [PubMed] [Google Scholar]
  295. Turner KS, Johnson CR. (2013) Behavioral interventions to address sleep disturbances in children with autism spectrum disorders: a review. Topics in Early Childhood Special Education 33:144-152. [Google Scholar]
  296. Tykhomyrov AA, Pavlova AS, Nedzvetsky VS. (2016) Glial fibrillary acidic protein (GFAP): on the 45th anniversary of its discovery. Neurophysiology 48:54-71. [Google Scholar]
  297. Ueda HR, Hayashi S, Chen W, Sano M, Machida M, Shigeyoshi Y, Iino M, Hashimoto S. (2005) System-level identification of transcriptional circuits underlying mammalian circadian clocks. Nat Genet 37:187-192. [DOI] [PubMed] [Google Scholar]
  298. Umemura Y, Koike N, Ohashi M, Tsuchiya Y, Meng QJ, Minami Y, Hara M, Hisatomi M, Yagita K. (2017) Involvement of posttranscriptional regulation of Clock in the emergence of circadian clock oscillation during mouse development. Proc Natl Acad Sci USA 114:E7479-E7488. [DOI] [PMC free article] [PubMed] [Google Scholar]
  299. Upadhyay A, Anjum B, Godbole NM, Rajak S, Shukla P, Tiwari S, Sinha RA, Godbole MM. (2019) Time-restricted feeding reduces high-fat diet associated placental inflammation and limits adverse effects on fetal organ development. Biochem Biophys Res Commun 514:415-421. [DOI] [PubMed] [Google Scholar]
  300. Vainchtein ID, Chin G, Cho FS, Kelley KW, Miller JG, Chien EC, Liddelow SA, Nguyen PT, Nakao-Inoue H, Dorman LC, et al. (2018) Astrocyte-derived interleukin-33 promotes microglial synapse engulfment and neural circuit development. Science 359:1269-1273. [DOI] [PMC free article] [PubMed] [Google Scholar]
  301. Van Erum J, Van Dam D, De Deyn PP. (2018) Sleep and Alzheimer’s disease: a pivotal role for the suprachiasmatic nucleus. Sleep Med Rev 40:17-27. [DOI] [PubMed] [Google Scholar]
  302. Veatch OJ, Maxwell-Horn AC, Malow BA. (2015) Sleep in autism spectrum disorders. Curr Sleep Med Rep 1:131-140. [DOI] [PMC free article] [PubMed] [Google Scholar]
  303. Vieta E, Berk M, Schulze TG, Carvalho AF, Suppes T, Calabrese JR, Gao K, Miskowiak KW, Grande I. (2018) Bipolar disorders. Nat Rev Dis Primers 4:18008. [DOI] [PubMed] [Google Scholar]
  304. Villabona-Rueda A, Erice C, Pardo CA, Stins MF. (2019) The evolving concept of the Blood Brain Barrier (BBB): from a single static barrier to a heterogeneous and dynamic relay center. Front Cell Neurosci 13:405. [DOI] [PMC free article] [PubMed] [Google Scholar]
  305. Vriend J, Reiter RJ. (2015) Melatonin feedback on clock genes: a theory involving the proteasome. J Pineal Res 58:1-11. [DOI] [PubMed] [Google Scholar]
  306. Wadhwa M, Prabhakar A, Ray K, Roy K, Kumari P, Jha PK, Kishore K, Kumar S, Panjwani U. (2017) Inhibiting the microglia activation improves the spatial memory and adult neurogenesis in rat hippocampus during 48 h of sleep deprivation. J Neuroinflammation 14:222. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  307. Walz W. (2000) Role of astrocytes in the clearance of excess extracellular potassium. Neurochem Int 36:291-300. [DOI] [PubMed] [Google Scholar]
  308. Wang C, Lutes LK, Barnoud C, Scheiermann C. (2022) The circadian immune system. Sci Immunol 7:eabm2465. [DOI] [PubMed] [Google Scholar]
  309. Wang JL, Lim AS, Chiang WY, Hsieh WH, Lo MT, Schneider JA, Buchman AS, Bennett DA, Hu K, Saper CB. (2015) Suprachiasmatic neuron numbers and rest-activity circadian rhythms in older humans. Ann Neurol 78:317-322. [DOI] [PMC free article] [PubMed] [Google Scholar]
  310. Wang XL, Wolff SEC, Korpel N, Milanova I, Sandu C, Rensen PCN, Kooijman S, Cassel J-C, Kalsbeek A, Boutillier A-L, et al. (2020) Deficiency of the circadian clock gene Bmal1 reduces microglial immunometabolism. Front Immunol 11:586399. [DOI] [PMC free article] [PubMed] [Google Scholar]
  311. Weaver DR, Reppert SM. (1995) Definition of the developmental transition from dopaminergic to photic regulation of c-fos gene expression in the rat suprachiasmatic nucleus. Brain Res Mol Brain Res 33:136-148. [DOI] [PubMed] [Google Scholar]
  312. Wei H, Zou H, Sheikh AM, Malik M, Dobkin C, Brown WT, Li X. (2011) IL-6 is increased in the cerebellum of autistic brain and alters neural cell adhesion, migration and synaptic formation. J Neuroinflammation 8:52. [DOI] [PMC free article] [PubMed] [Google Scholar]
  313. Wen S, Ma D, Zhao M, Xie L, Wu Q, Gou L, Zhu C, Fan Y, Wang H, Yan J. (2020) Spatiotemporal single-cell analysis of gene expression in the mouse suprachiasmatic nucleus. Nat Neurosci 23:456-467. [DOI] [PubMed] [Google Scholar]
  314. Werling DM, Geschwind DH. (2013) Sex differences in autism spectrum disorders. Curr Opin Neurol 26:146-153. [DOI] [PMC free article] [PubMed] [Google Scholar]
  315. Westwood ML, O’Donnell AJ, de Bekker C, Lively CM, Zuk M, Reece SE. (2019) The evolutionary ecology of circadian rhythms in infection. Nat Ecol Evol 3:552-560. [DOI] [PMC free article] [PubMed] [Google Scholar]
  316. Wolff SEC, Wang XL, Jiao H, Sun J, Kalsbeek A, Yi CX, Gao Y. (2020) The effect of Rev-erbalpha agonist SR9011 on the Immune response and cell metabolism of microglia. Front Immunol 11:550145. [DOI] [PMC free article] [PubMed] [Google Scholar]
  317. Woodruff ER, Chun LE, Hinds LR, Spencer RL. (2016) Diurnal corticosterone presence and phase modulate clock gene expression in the male rat prefrontal cortex. Endocrinology 157:1522-1534. [DOI] [PMC free article] [PubMed] [Google Scholar]
  318. Wulff K, Dijk DJ, Middleton B, Foster RG, Joyce EM. (2012) Sleep and circadian rhythm disruption in schizophrenia. Br J Psychiatry 200:308-316. [DOI] [PMC free article] [PubMed] [Google Scholar]
  319. Xie L, Kang H, Xu Q, Chen MJ, Liao Y, Thiyagarajan M, O’Donnell J, Christensen DJ, Nicholson C, Iliff JJ, et al. (2013) Sleep drives metabolite clearance from the adult brain. Science 342:373-377. [DOI] [PMC free article] [PubMed] [Google Scholar]
  320. Xu ZX, Kim GH, Tan JW, Riso AE, Sun Y, Xu EY, Liao G-Y, Xu H, Lee S-H, Do N-Y, et al. (2020) Elevated protein synthesis in microglia causes autism-like synaptic and behavioral aberrations. Nat Commun 11:1797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  321. Yang Z, Matsumoto A, Nakayama K, Jimbo EF, Kojima K, Nagata K, Iwamoto S, Yamagata T. (2016) Circadian-relevant genes are highly polymorphic in autism spectrum disorder patients. Brain Dev 38:91-99. [DOI] [PubMed] [Google Scholar]
  322. Yin L, Lazar MA. (2005) The orphan nuclear receptor Rev-erbalpha recruits the N-CoR/histone deacetylase 3 corepressor to regulate the circadian Bmal1 gene. Mol Endocrinol 19:1452-1459. [DOI] [PubMed] [Google Scholar]
  323. Yoshitane H, Asano Y, Sagami A, Sakai S, Suzuki Y, Okamura H, Iwasaki W, Ozaki H, Fukada Y. (2019) Functional D-box sequences reset the circadian clock and drive mRNA rhythms. Commun Biol 2:300. [DOI] [PMC free article] [PubMed] [Google Scholar]
  324. Yu X, Rollins D, Ruhn KA, Stubblefield JJ, Green CB, Kashiwada M, Rothman PB, Takahashi JS, Hooper LV. (2013) TH17 cell differentiation is regulated by the circadian clock. Science 342:727-730. [DOI] [PMC free article] [PubMed] [Google Scholar]
  325. Zamanian JL, Xu L, Foo LC, Nouri N, Zhou L, Giffard RG, Barres BA. (2012) Genomic analysis of reactive astrogliosis. J Neurosci 32:6391-6410. [DOI] [PMC free article] [PubMed] [Google Scholar]
  326. Zeppenfeld DM, Simon M, Haswell JD, D’Abreo D, Murchison C, Quinn JF, Grafe MR, Woltjer RL, Kaye J, Iliff JJ. (2017) Association of perivascular localization of aquaporin-4 with cognition and Alzheimer disease in aging brains. JAMA Neurol 74:91-99. [DOI] [PubMed] [Google Scholar]
  327. Zhang H, Liu Y, Liu D, Zeng Q, Li L, Zhou Q, Li M, Mei J, Yang N, Mo S, et al. (2021) Time of day influences immune response to an inactivated vaccine against SARS-CoV-2. Cell Res 31:1215-1217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  328. Zhang J, Chatham JC, Young ME. (2020) Circadian regulation of cardiac physiology: rhythms that keep the heart beating. Annu Rev Physiol 82:79-101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  329. Zhang R, Lahens NF, Ballance HI, Hughes ME, Hogenesch JB. (2014) A circadian gene expression atlas in mammals: implications for biology and medicine. Proc Natl Acad Sci USA 111:16219-16224. [DOI] [PMC free article] [PubMed] [Google Scholar]
  330. Zhang Y, Xu H, Zhang F, Shao F, Ellenbroek B, Wang J, Wang W. (2019) Deficiencies of microglia and TNFalpha in the mPFC-mediated cognitive inflexibility induced by social stress during adolescence. Brain Behav Immun 79:256-266. [DOI] [PubMed] [Google Scholar]
  331. Zheng X, Sehgal A. (2010) AKT and TOR signaling set the pace of the circadian pacemaker. Curr Biol 20:1203-1208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  332. Zhou JN, Hofman MA, Swaab DF. (1995) VIP neurons in the human SCN in relation to sex, age, and Alzheimer’s disease. Neurobiol Aging 16:571-576. [DOI] [PubMed] [Google Scholar]
  333. Zhu Y, Zhan G, Fenik P, Brandes M, Bell P, Francois N, Shulman K, Veasey S. (2018) Chronic sleep disruption advances the temporal progression of tauopathy in P301S mutant mice. J Neurosci 38:10255-10270. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Journal of Biological Rhythms are provided here courtesy of SAGE Publications

RESOURCES