Skip to main content
  • Mini-review
  • Published:

Role of GABA in the regulation of the central circadian clock of the suprachiasmatic nucleus

Abstract

In mammals, circadian rhythms, such as sleep/wake cycles, are regulated by the central circadian clock located in the suprachiasmatic nucleus (SCN) of the hypothalamus. The SCN consists of thousands of individual neurons, which exhibit circadian rhythms. They synchronize with each other and produce robust and stable oscillations. Although several neurotransmitters are expressed in the SCN, almost all SCN neurons are γ-amino butyric acid (GABA)-ergic. Several studies have attempted to understand the roles of GABA in the SCN; however, precise mechanisms of the action of GABA in the SCN are still unclear. GABA exhibits excitatory and/or inhibitory characteristics depending on the circadian phase or region in the SCN. It can both synchronize and destabilize cellular circadian rhythms in individual SCN cells. Differing environmental light conditions, such as a long photoperiod, result in the decoupling of circadian oscillators of the dorsal and ventral SCN. This is due to high intracellular chloride concentrations in the dorsal SCN. Because mice with functional GABA deficiency, such as vesicular GABA transporter- and glutamate decarboxylase-deficient mice, are neonatal lethal, research has been limited to pharmacological approaches. Furthermore, different recording methods have been used to understand the roles of GABA in the SCN. The excitability of GABAergic neurons also changes during the postnatal period. Although there are technical difficulties in understanding the functions of GABA in the SCN, technical developments may help uncover new roles of GABA in circadian physiology and behavior.

The central circadian clock: the suprachiasmatic nucleus (SCN)

Several physiological functions in our body exhibit oscillations on various time scales, including electrical activity in the brain, heart rate, breathing, and the sleep/wake cycle. Among them, circadian rhythms are defined as approximately 24-h oscillations in physiology and behavior. In mammals, the suprachiasmatic nucleus (SCN) of the hypothalamus is known as the central circadian pacemaker. Circadian behavioral rhythms were abolished by SCN lesions [71, 105], and restored by implantation of the SCN [56]. Importantly, the restored circadian period was identical to that of the donor, rather than the host [91, 107]. In addition, implantation of the SCN contained in a semipermeable polymeric capsule also restored circadian behavioral rhythms in SCN-lesioned hamsters, indicating that diffusible signals from the SCN control circadian behavior [102].

The SCN contains approximately 20,000 neurons [113] that have heterogeneous circadian properties. In dispersed SCN cell culture, individual SCN neurons exhibit autonomous oscillations. However, the circadian period, phase, and amplitude, differ from cell to cell [34, 37, 120]. In cultured SCN slices, circadian rhythms of individual cells synchronize with each other and they express stable circadian oscillations [33, 36, 79, 84] (Fig. 1). Cellular networks in the SCN are important for synchronization [121] and the stability of circadian oscillation in individual cells [59]. Synchronized circadian rhythms within the SCN are responsible for coordinating peripheral circadian oscillators [123], and exhibiting circadian behavior. In the SCN, there are multiple oscillators. Environmental light/dark conditions change the coupling between these regional oscillators in the SCN, which is critical for the output for the circadian behavior and peripheral clock [35, 41, 78, 82, 114].

Fig. 1
figure 1

PER2::LUC rhythms in dispersed SCN cells, SCN slices, and the SCN of freely moving mice. Per2 protein fusion luciferase activity was measured using bioluminescence imaging from dispersed (left) and slice (middle) SCN culture using an EM-CCD camera. Circadian rhythms of individual SCN cells in dispersed (lower left) and slice (lower middle) cultures are expressed as pseudo colors, in which red and blue indicate peak and trough circadian rhythm phases. PER2::LUC rhythms in the SCN in freely moving mice were measured using an optical fiber under constant darkness for 4 days. Neuronal networks are important for SCN cellular circadian rhythm synchronization. Scale bars represent 200 μm

Circadian rhythms of individual SCN cells are generated by transcription-translation feedback loops involving several clock genes and protein products [93]. In this feedback loop, the positive elements, BMAL1 and CLOCK form a heterodimer that initiates the transcription of genes that contain E-box enhancer sequences, including Period (Per) and Cryptochrome (Cry) [12, 23, 53]. The protein products of Per and Cry then suppress transactivation by the BMAL1/CLOCK heterodimer [96, 100]. This clock machinery is widely observed on the single cell level.

Anatomical properties of the SCN

The SCN is divided into dorsal (shell) and ventral (core) subdivisions. Retinal projections to the ventral SCN [11, 108] and dorsal SCN [29] are observed. The SCN contains several neurotransmitters located within specific regions (Fig. 2). This general organization has been studied in hamsters, mice, rats, and humans [1, 14, 62, 74]. Arginine vasopressin (AVP) neurons are located in the SCN shell. Conversely, the SCN core expresses several neuropeptides, such as vasoactive intestinal peptides (VIP), gastrin-releasing peptide (GRP), and calbindin [1, 103].

Fig. 2
figure 2

Schematic diagram of circadian organization in the SCN. Light input is transmitted to the ventral SCN through the retinohypothalamic tract (RHT). Several neuropeptides are expressed in the SCN and almost all SCN cells are GABAergic neurons. Circadian rhythms in individual SCN cells synchronize via neurotransmitters, and synchronized circadian rhythms in the SCN regulate behaviors such as sleep and wake cycles

Remarkably, almost all SCN neurons express γ-amino butyric acid (GABA) [1, 73]. GABAergic neurons in the SCN contain GABA vesicular transporters (VGAT), [8] and GABA synthesizing enzymes, such as glutamate decarboxylase (GAD) [73, 83]. Two distinct isoforms of GAD (GAD65 and GAD67) are found within the SCN, and they have different molecular weights and subcellular distributions. A study using immunohistochemistry determined that GAD65 and GAD67 are highly expressed in the ventral SCN of rats, similar to the distribution of VGAT [8]. GABA receptors, GABAA and GABAB, are also found in the SCN [8, 21, 106]. GABAB receptors are mainly expressed in the dorsal area in the SCN, which generally corresponds to the region of AVP positive neurons [8].

The timing of GABA synthesis, trafficking, and release are important for the modulation of circadian rhythms in the SCN. It was reported that mRNA levels of the GABA synthesizing enzyme GAD65 were higher in the light than in the dark period, but mRNA levels of GAD67 were not rhythmic in the SCN [38]. Another group also reported that both GAD65 and GAD67 mRNA showed circadian oscillations under constant darkness [13]. VGAT contents in the SCN did not exhibit circadian rhythms, but expression levels were attenuated under constant darkness compared with light dark conditions [18]. Furthermore, intercellular signals, such as VIP, could modulate GABA release in the SCN [24, 44].

Excitatory and/or inhibitory effects of GABA in the SCN

Although the effects of GABA on cellular activity in the SCN have been studied for more than 30 years, results remain controversial. For example, one study demonstrated that GABA had inhibitory effects on almost all SCN neurons [65], while several studies have reported that GABA has both excitatory and inhibitory effects in the SCN [15, 22, 57, 101, 116]. Wanger et al. reported that in SCN tissue slices, the application of GABA decreased firing frequency at night, and increased firing frequency during the day. This result indicates that the effects of GABA in SCN neurons are dependent on the circadian phase [116]. However, some investigators have reported no difference in the excitatory and inhibitory effects of GABA during the subjective day compared to night [15, 26, 57, 58, 65]. Similar to the application of GABA, GABA agonists, such as muscimol and baclofen, induce excitatory or inhibitory effects in the SCN [26, 39, 70], while GABA has also been shown to produce either excitatory or inhibitory effects on firing across the circadian day [22]. These effects were inhibited by the GABAA receptor antagonist bicuculline [2, 15, 22, 26, 65]. Regional differences in the effects of GABA in the SCN have also been reported. Short-term application of bicuculline increased neuronal activity in the ventral SCN, but decreased activity in dorsal SCN slices [2]. These pharmacological approaches have shown inconsistent observations regarding the acute effects of GABA in the SCN. Furthermore, these agents may represent non-specific pharmacological actions. For example, the GABAA receptor antagonist, bicuculline is generally known to block small-conductance calcium-activated potassium channels [45, 51] and does so in SCN neurons [110]. It is necessary to consider this non-specific effect of these drugs.

Coupling circadian oscillators in the SCN

Individual SCN neurons exhibit autonomous circadian rhythms even when isolated in culture [119, 120]. The circadian period, phase, and amplitude differ from each other in dispersed cell culture, although their rhythms are synchronized in SCN slices [33, 36, 79, 84] and in vivo [42, 122]. Due to the involvement of heterogeneous oscillators in the SCN, individual SCN cells must couple to each other. Synchronized circadian rhythms in the SCN entrain light–dark cycles to adapt to environmental light–dark conditions. It is thought that GABA may be involved in mediating circadian rhythm coupling in individual SCN neurons.

Synchronizer or destabilizer of cellular circadian rhythms

Liu and Reppert demonstrated that the application of GABA on dispersed SCN cultured cells completely inhibited spontaneous firing at all circadian phases. The application of GABA after the peak phase of neuronal activity rhythms induced a large phase delay. The GABAA agonist muscimol induced phase shifts of neuronal activity rhythms in SCN neurons, whereas the application of the GABAB agonist bacrofen had no effect, suggesting that this phase-dependent shift is mediated by GABAA receptors. Additionally, GABA was applied daily to dispersed SCN cultured cells to assess the effects of circadian rhythms in individual SCN neurons. After administration of GABA for 3 h every 24 h for 5 days, circadian rhythms of spontaneous firings in individual SCN neurons synchronized. This result has been interpreted as evidence that the daily application of GABA synchronized circadian firing rhythms in the dispersed SCN cell culture [60] (Fig. 3a). However, these data do not exclude the effects of other transmitters, such as neuropeptides.

Fig. 3
figure 3

Possible roles of GABA in SCN cellular coupling. a GABA acts as both a synchronizer and destabilizer. b A phase delay shift results in bimodal neuronal activity in the intact SCN slice; however, after the application of GABAA receptor antagonists, circadian rhythms in the dorsal and ventral SCN are dissociated. c Long-day photoperiods change the intracellular chloride concentration in the dorsal SCN and decoupled circadian rhythms between the dorsal and ventral SCN. d Hypothetical model of the astrocytic-neuronal intercellular axis in the SCN

Monitoring gene expression or protein products of clock genes in the SCN using bioluminescence reporters has provided insights into the understanding of circadian rhythms [121, 123, 125]. Based on the evidence for the role of GABA in the synchronization of circadian rhythms in the SCN [60], antagonizing GABA receptors may induce desynchronization of cellular circadian rhythms. However, long-term application of GABAA (bicuculline) and GABAB (saclofen) receptor antagonists into the Per1 promoter-driven luciferase reporter (Per1-luc) SCN slices gradually increased the amplitude of circadian rhythms compared with controls at the tissue level [4]. This increased amplitude of circadian rhythms was due to increased amplitude of cellular circadian rhythms and decreased cycle-to-cycle period variation (increased precision of cellular rhythms). GABA receptor antagonists (bicuculline and saclofen) also increased the amplitude of circadian firing rhythms in dispersed cultured cells [4]. Similarly, blocking GABAA receptor signaling with the application of gabazine onto the Per2 protein fusion luciferase reporter (PER2::LUC) SCN slices decreased circadian period variability in individual cells compared to vehicle-treated controls [22]. These results suggest that GABA destabilizes circadian oscillations in the SCN (Fig. 3a).

Coupling of dorsal and ventral circadian oscillators

Constant light causes splitting of circadian behavioral rhythms [89]. The circadian rhythms of the shell and core, or left and right SCN, exhibited a 180° antiphase during splitting under constant light conditions [82, 124]. Another study showed that an abrupt change in the light–dark cycle disrupted synchronous oscillation of circadian components in the rat SCN [76]. After a phase delay shift of light–dark cycles, clock gene expression rhythms shifted rapidly in the ventrolateral SCN, whereas this shift occurred more slowly in the dorsomedial SCN [76]. Several researchers have demonstrated a role of GABA in the re-synchronization of the dorsal and ventral regions of the SCN after manipulation of environmental light–dark cycles.

Albus et al. revealed the role of GABA in coupling dorsal and ventral SCN circadian rhythms in acute SCN slices. Measurements of neuronal activity in the SCN following a 6-h phase delay in the light/dark schedule revealed bimodal patterns of activity rhythms. Furthermore, when the SCN was cut horizontally, separating the slice into dorsal and ventral areas, the peak phase of neuronal activity rhythms in the ventral SCN was significantly advanced compared with that in the dorsal SCN. Continuous application of a GABAA receptor antagonist (bicuculline) yielded similar results [2] (Fig. 3b). These results indicate that GABA is important for coupling circadian rhythms in the dorsal and ventral SCN.

Photoperiodic changes in SCN cellular networks

Environmental light–dark conditions change depending on the season. This photoperiodic change is important for seasonal reproduction in some animals [81, 126]. The long-day photoperiod also changes cellular networks in the SCN and decouples circadian oscillatory cell groups [30, 41, 77]. Recently, the role of GABAA receptors in coupling SCN dorsal and ventral circadian oscillators was reported to occur during the long-day photoperiod [19] (Fig. 3c). PER2::LUC reporter mice were exposed to 20 h of light and 4 h of dark (LD20:4), and the PER2::LUC bioluminescence was measured from the SCN slice. In these conditions, the circadian phase of the SCN core was advanced compared with that of the shell. This phase difference between the dorsal and ventral region in the SCN gradually returned to an organizational state similar to that observed under LD12:12 conditions. To investigate the role of GABAA signaling in the re-synchronization of dorsal and ventral oscillators after LD20:4 conditions, the GABAA receptor antagonist, bicuculline, was applied to the SCN slice. Bicuculline attenuated the re-synchronization when the circadian phase between dorsal and ventral regions was out of phase, but not when it was in-phase. These results were interpreted as evidence that GABAA signaling contributes to the synchronization of circadian rhythms between the dorsal and ventral SCN in a state-dependent manner. An alternative interpretation would suggest that GABA may acutely modulate firing without having much, if any, effect on circadian phase. These results also suggest that GABA is sufficient for the synchronization of dispersed SCN cells [60] and that the absence of GABA does not desynchronize cellular rhythms under steady-state networks in the SCN [4].

Intracellular chloride concentrations and cellular coupling

Long-day photoperiods also change the excitability [20], and levels of chloride transporter expression in SCN neurons. This determines the difference in circadian phase and period between the dorsal and ventral SCN [75] (Fig. 3c). It was observed that under long-day conditions, the phase difference in Bmal1 promoter-driven luciferase reporter (Bmal1-Eluc) circadian rhythms between the dorsal and ventral SCN were increased, and the circadian period of the dorsal SCN was decreased compared with the ventral SCN. Myung et al. also measured intracellular chloride concentrations in the SCN, using N-(ethoxycarbonylmethyl)-6-methoxyquinoliniumbromide fluorescence, and found that intracellular chloride concentrations were increased under long-day photoperiods. These results are due to a higher expression ratio of sodium/potassium/chloride cotransporter (NKCC1)/potassium/chloride cotransporters (KCC2), in the dorsal than the ventral SCN. Because NKCC1 is a chloride importer, a high ratio of NKCC1/KCC2 results in more GABA-induced excitation. KCC2 is expressed exclusively in VIP and GRP neurons, whereas NKCC1 is expressed in VIP, GRP, and AVP neurons within the SCN [7]. Recently, Klett and Allen reported that intracellular chloride concentrations were higher during the day than at night in both AVP- and VIP-positive neurons [52]. The prevalence of GABA excitation and inhibition is dependent on the level of chloride transporter expression and may affect the coupling of dorsal and ventral SCN circadian oscillations.

Astrocytes and GABA

In addition to neurons, astrocytes in the SCN are involved in circadian rhythms. Individual astrocytes display circadian rhythms entrained to daily temperature cycles [90]. Moreover, astrocytes co-cultured with adult SCN explants sustained the rhythms of astrocytes, suggesting that diffusible factors from the SCN are sufficient to entrain circadian oscillations in astrocytes [90]. Furthermore, VIP expressed in SCN neurons entrains astrocyte circadian rhythms [63]. It is known that astrocytes release ATP into the extracellular space and it has been shown that cultured astrocytes display daily oscillations of extracellular ATP concentrations that are under circadian control [64]. Astrocytes also regulate neuronal networks through the reuptake and release of various transmitters including glutamate [27, 88]; however, the functional mechanisms of these transmitters in astrocytes have yet to be identified. Recently, three groups reported that astrocytes regulate SCN and behavioral circadian rhythms [5, 10, 112]. Astrocyte-specific deletion of the Bmal1 gene led to reduced expression of astrocytic GABA transporter 1 and 3 (GAT1 and GAT3), suggesting a potential impairment in the clearance of extracellular GABA released by neurons [5]. Brancaccio et al. proposed a new model of circadian timekeeping in the SCN. They hypothesized that in the SCN, glutamate released from astrocytes maintains higher intracellular calcium levels, specifically in pre-synaptic terminals, through the activation of NMDA receptors (NR2C), which subsequently facilitates neuronal GABA release [10] (Fig. 3d). Because the peak phase of circadian calcium rhythms in astrocytes was observed at night, GABA release from neurons via glutamate release from astrocytes would increase at night, resulting in a decrease in neuronal activity at night.

Circadian outputs and GABA functions

Circadian rhythms in SCN neurons synchronize via several neurotransmitters, such as AVP, VIP, and GRP [3, 66, 67]. To regulate sleep/wake cycles, SCN circadian rhythms need to send outputs to peripheral circadian oscillators. Neuronal activity in the SCN is one of the most important input and/or output signals from molecular oscillations. Application of the sodium channel blocker, tetrodotoxin, into the SCN in vivo resulted in arrhythmicity in behavior without affecting the circadian oscillation [97]. In addition, optogenetic stimulation of the SCN in vivo modulated circadian behavioral rhythms [46]. GABA, therefore, may regulate the transmission of circadian output from the SCN by changing neuronal excitability. In contrast, it is also reported that GABA is involved in the synaptic plasticity changes in the retinohypothalamic tract-SCN synapses [69]. These results indicate that GABA modulates both input and output of circadian oscillation in the SCN.

Identification of SCN circadian outputs is important for understanding the mechanisms underlying the regulation of behavioral rhythms. Previously, SCN efferent projections have been investigated [50, 54, 104, 117, 118]. Neurons in the dorsal SCN project densely to the preoptic area, paraventricular nucleus (PVN), dorsomedial hypothalamus, and subparaventricular zone (SPVZ). In contrast, dense projections from the core are limited to the peri-suprachiasmatic area (PSCN), lateral SPVZ, and ventral tuberal area (VTU) [54]. Interestingly, the circadian rhythms of neuronal activity outside the SCN in vivo were in antiphase compared with the SCN in nocturnal animals [42, 80]. Importantly, in both diurnal and nocturnal animals, the peak phase of neuronal activity rhythms in the SCN is observed during the day. However, in diurnal animals, the peak phase of neuronal activity outside of the SCN is in-phase compared with that of the SCN [95], suggesting that circadian information from the SCN is transferred to output brain areas, and that this mechanism is different in nocturnal and diurnal animals. The mechanisms for switching day-night information from the SCN, however, have not been identified.

The PVN receives an efferent projection from the SCN. The PVN contains several neuropeptides, such as corticotrophin-releasing factor, oxytocin, and AVP, and regulates endocrine and autonomic functions. Electrical stimulation of the SCN evoked monosynaptic inhibitory postsynaptic potentials, as well as excitatory postsynaptic potentials in the PVN, indicating that GABA and glutamate are important mediators of fast monosynaptic transmission from the SCN to the PVN [31, 32]. Tousson and Meissl used a multi-electrode array dish to demonstrate that humoral factors are responsible for circadian rhythms in the PVN [111]. They observed circadian rhythms in PVN neuronal activity measured in brain slices containing both PVN and SCN. When the SCN was removed from the brain slice, PVN circadian rhythms disappeared and were subsequently restored by co-cultured SCN grafts.

The SCN regulates circadian rhythms of plasma melatonin concentrations via a multi-synaptic pathway, including the PVN, sympathetic preganglionic neurons of the spinal cord, and noradrenergic sympathetic neurons of the superior cervical ganglion [72]. The melatonin concentration is increased at night and decreased during the day. Blocking GABAergic transmission to the PVN results in inhibition of melatonin synthesis in the pineal gland [48, 49], suggesting that GABAergic signals from the SCN have an inhibitory effect on melatonin synthesis during the day. Importantly, melatonin administration suppressed spontaneous firing in the SCN [61, 99]. Neuronal and humoral regulation of PVN neuronal activity by the SCN is important for temporal integration of physiological events.

Variety of circadian recording methods

Several studies have attempted to understand the function of GABA in the SCN, using both ex vivo and in vivo methods (Fig. 4). Primary culture techniques (dispersed and slice) are useful for measuring SCN circadian rhythms. Typically, the SCN from neonatal animals is used for this experiment [34, 60, 79, 120]; however, acute SCN slices from adult animals have also been used for recording circadian rhythms [2, 25, 116]. Neonatal SCN tissue allows researchers to measure circadian rhythms over long periods; however, the cellular properties and neuronal networks in the SCN may not be the same in the adult. In general, NKCC1 expression predominates in immature neurons, in which the intracellular concentration of chloride ions is relatively high, whereas KCC2 expression predominates in mature neurons. These developmental changes of chloride transporters modulate excitability to GABA depending on age [9]. Developmental effects of GABA on SCN circadian oscillations have not been fully identified. Recently, however, we reported developmental changes of neuronal networks in the SCN, as Ono et al. demonstrated that Cryptochrome (Cry), a clock gene, is involved in cellular coupling in the adult SCN [84, 85]. GABA may have an important role in cellular coupling and circadian outputs in the SCN, depending on the developmental stage.

Fig. 4
figure 4

Variety of circadian rhythm recording techniques. a Different parameters can be measured using different methods. Clock gene expression rhythms can be measured with bioluminescence reporters. Cytosolic calcium concentrations can be measured with fluorescent probes. Neuronal activity can be recorded using patch clamp or multi-electrode array dish techniques. b SCN cell preparations for recording circadian rhythms. Individual SCN properties may depend on the coupling strength and the age of mice

GABA functions have been studied using various recording methods, such as neuronal activity recording, calcium imaging, and bioluminescence imaging [2, 4, 40, 43, 60, 101, 116]. These recording methods are useful for measuring circadian oscillations in the SCN. However, circadian oscillation in measurements such as gene expression, intracellular calcium levels, and neuronal activity do not show the same properties. For instance, Vansteensel et al. measured Per1-luc and neuronal activity rhythms in both acute SCN slices and in vivo, during a 6-h light/dark phase advance schedule. Per1 and neuronal activity rhythms in SCN slices demonstrated a phase shift immediately after the 6-h phase advance, whereas in vivo neuronal activity did not immediately shift, indicating that neuronal activity was dissociated between brain slices and in vivo [115]. Ono et al. also demonstrated that circadian rhythms of the clock genes, Per1 and Bmal1, were dissociated in cultured SCN slices [86]. We have successfully simultaneously measured Per1, Bmal1, cytosolic calcium ions, and neuronal activity in SCN slices and found that the circadian period of Bmal1-ELuc rhythms is shorter than that of Per1-luc rhythms. Furthermore, the circadian period of calcium and neuronal activity rhythms were intermediate between Per1 and Bmal1 rhythms. Simultaneous and multifunctional recording of circadian oscillation is a useful tool for understanding the hierarchal structure of SCN networks in addition to the roles of GABA within these networks.

Perspectives

All studies that have addressed the function of GABA in the SCN have only used pharmacological approaches. Because mice with genetic GABA deficiencies, such as VGAT- and GAD65/67-deficient mice, cannot survive after birth [47, 94], it is difficult to assess the roles of GABA in the SCN using a genetic approach. It is possible to measure SCN circadian rhythms obtained from embryos of VGAT- or GAD65/67-deficient mice. However, this slice culture approach is not enough to understand mechanisms of circadian rhythms as they relate to behavior. A conditional knockout method could potentially provide a solution to this problem. Recently, several Cre driver mice have been used to conditionally disrupt genes of interest in the SCN [6, 28, 55, 68]. Creating conditional KO mice is useful; however, there are currently no SCN-specific Cre driver mice. While Neuromedin-S is specifically expressed in the SCN, not all SCN neurons express this neuropeptide [55]. Recently, a Crispr/Cas9 genome editing method, a powerful tool for manipulating genes of interest in various species [17], was developed and has been used to manipulate signaling in specific cell types of the SCN [112]. Adeno-associated viral driven Crispr/Cas9 techniques may be applicable to the study of GABA function specifically in the SCN [92]. Since transfection and expression efficiency are not perfect, however, it is difficult to control viral diffusion outside of the SCN. Therefore, alternative methods are required for SCN-specific functional GABA deficiency.

To understand the roles of GABA in circadian physiology and behavior, it is important to identify neuronal networks and output pathways from the SCN. Several new methods have been developed to accomplish this. For example, the glycoprotein-deleted (DG) rabies virus is a useful tool for studies of neural circuits [87]. In combination with Cre mice, we can identify direct input and output pathways of specific neurons [98]. However, because neurons infected with the rabies virus are killed by approximately 14 days after infection [87], manipulations or recordings of neuronal activity are limited to this short time window. Recently, self-inactivating rabies virus has been developed, which may allow us to further understand neuronal brain networks [16]. AAV-mediated retrograde tracing methods have also been developed, which could allow us to label, manipulate, and measure retrograde-labeled neurons [109]. Identification of a local GABA circuit in the SCN and long-range brain networks may be important for the understanding of circadian physiology and behavior in the future.

References

  1. Abrahamson EE, Moore RY (2001) Suprachiasmatic nucleus in the mouse: retinal innervation, intrinsic organization and efferent projections. Brain Res 916:172–191

    CAS  PubMed  Google Scholar 

  2. Albus H, Vansteensel MJ, Michel S, Block GD, Meijer JH (2005) A GABAergic mechanism is necessary for coupling dissociable ventral and dorsal regional oscillators within the circadian clock. Curr Biol 15:886–893

    CAS  PubMed  Google Scholar 

  3. Aton SJ, Colwell CS, Harmar AJ, Waschek J, Herzog ED (2005) Vasoactive intestinal polypeptide mediates circadian rhythmicity and synchrony in mammalian clock neurons. Nat Neurosci 8:476–483

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Aton SJ, Huettner JE, Straume M, Herzog ED (2006) GABA and Gi/o differentially control circadian rhythms and synchrony in clock neurons. Proc Natl Acad Sci USA 103:19188–19193

    CAS  PubMed  PubMed Central  Google Scholar 

  5. 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. Nature Commun 8:14336

    CAS  Google Scholar 

  6. Bedont JL, LeGates TA, Slat EA, Byerly MS, Wang H, Hu J, Rupp AC, Qian J, Wong GW, Herzog ED, Hattar S, Blackshaw S (2014) Lhx1 controls terminal differentiation and circadian function of the suprachiasmatic nucleus. Cell Rep 7:609–622

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Belenky MA, Sollars PJ, Mount DB, Alper SL, Yarom Y, Pickard GE (2010) Cell-type specific distribution of chloride transporters in the rat suprachiasmatic nucleus. Neuroscience 165:1519–1537

    CAS  PubMed  Google Scholar 

  8. Belenky MA, Yarom Y, Pickard GE (2008) Heterogeneous expression of gamma-aminobutyric acid and gamma-aminobutyric acid-associated receptors and transporters in the rat suprachiasmatic nucleus. J Comp Neurol 506:708–732

    CAS  PubMed  Google Scholar 

  9. Ben-Ari Y (2002) Excitatory actions of GABA during development: the nature of the nurture. Nat Rev Neurosci 3:728–739

    CAS  PubMed  Google Scholar 

  10. 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):e1425

    Google Scholar 

  11. Bryant DN, LeSauter J, Silver R, Romero MT (2000) Retinal innervation of calbindin-D28 K cells in the hamster suprachiasmatic nucleus: ultrastructural characterization. J Biol Rhythms 15:103–111

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Bunger MK, Wilsbacher LD, Moran SM, Clendenin C, Radcliffe LA, Hogenesch JB, Simon MC, Takahashi JS, Bradfield CA (2000) Mop3 is an essential component of the master circadian pacemaker in mammals. Cell 103:1009–1017

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Cagampang FR, Rattray M, Powell JF, Campbell IC, Coen CW (1996) Circadian changes of glutamate decarboxylase 65 and 67 mRNA in the rat suprachiasmatic nuclei. NeuroReport 7:1925–1928

    CAS  PubMed  Google Scholar 

  14. Card JP, Moore RY (1984) The suprachiasmatic nucleus of the golden hamster: immunohistochemical analysis of cell and fiber distribution. Neuroscience 13:415–431

    CAS  PubMed  Google Scholar 

  15. Choi HJ, Lee CJ, Schroeder A, Kim YS, Jung SH, Kim JS, Kim DY, Son EJ, Han HC, Hong SK, Colwell CS, Kim YI (2008) Excitatory actions of GABA in the suprachiasmatic nucleus. J Neurosc 28:5450–5459

    CAS  Google Scholar 

  16. Ciabatti E, Gonzalez-Rueda A, Mariotti L, Morgese F, Tripodi M (2017) Life-long genetic and functional access to neural circuits using self-inactivating rabies virus. Cell 170(382–392):e314

    Google Scholar 

  17. Cong L, Ran FA, Cox D, Lin SL, Barretto R, Habib N, Hsu PD, Wu XB, Jiang WY, Marraffini LA, Zhang F (2013) Multiplex genome engineering using CRISPR/Cas systems. Science (New York, N.Y.) 339:819–823

    CAS  Google Scholar 

  18. Darna M, Schmutz I, Richter K, Yelamanchili SV, Pendyala G, Holtje M, Albrecht U, Ahnert-Hilger G (2009) Time of day-dependent sorting of the vesicular glutamate transporter to the plasma membrane. J Biol Chem 284:4300–4307

    CAS  PubMed  Google Scholar 

  19. Evans JA, Leise TL, Castanon-Cervantes O, Davidson AJ (2013) Dynamic interactions mediated by nonredundant signaling mechanisms couple circadian clock neurons. Neuron 80:973–983

    CAS  PubMed  Google Scholar 

  20. Farajnia S, van Westering TL, Meijer JH, Michel S (2014) Seasonal induction of GABAergic excitation in the central mammalian clock. Proc Natl Acad Sci USA 111:9627–9632

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Francois-Bellan AM, Hery M, Faudon M, Hery F (1989) Analysis of the inhibitory effect of oestradiol on functional GABA/5-HT relationship in the rat suprachiasmatic area. J Neuroendocrinol 1:415–422

    CAS  PubMed  Google Scholar 

  22. Freeman GM Jr, Krock RM, Aton SJ, Thaben P, Herzog ED (2013) GABA networks destabilize genetic oscillations in the circadian pacemaker. Neuron 78:799–806

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Gekakis N, Staknis D, Nguyen HB, Davis FC, Wilsbacher LD, King DP, Takahashi JS, Weitz CJ (1998) Role of the CLOCK protein in the mammalian circadian mechanism. Science (New York, N.Y.) 280:1564–1569

    CAS  Google Scholar 

  24. Gompf HS, Irwin RP, Allen CN (2006) Retrograde suppression of GABAergic currents in a subset of SCN neurons. Eur J Neurosci 23:3209–3216

    PubMed  Google Scholar 

  25. Green DJ, Gillette R (1982) Circadian rhythm of firing rate recorded from single cells in the rat suprachiasmatic brain slice. Brain Res 245:198–200

    CAS  PubMed  Google Scholar 

  26. Gribkoff VK, Pieschl RL, Wisialowski TA, Park WK, Strecker GJ, de Jeu MT, Pennartz CM, Dudek FE (1999) A reexamination of the role of GABA in the mammalian suprachiasmatic nucleus. J Biol Rhythms 14:126–130

    CAS  PubMed  Google Scholar 

  27. Halassa MM, Haydon PG (2010) Integrated brain circuits: astrocytic networks modulate neuronal activity and behavior. Annu Rev Physiol 72:335–355

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Hatori M, Gill S, Mure LS, Goulding M, O’Leary DD, Panda S (2014) Lhx1 maintains synchrony among circadian oscillator neurons of the SCN. Elife 3:e03357

    PubMed  PubMed Central  Google Scholar 

  29. Hattar S, Kumar M, Park A, Tong P, Tung J, Yau KW, Berson DM (2006) Central projections of melanopsin-expressing retinal ganglion cells in the mouse. J Comp Neurol 497:326–349

    PubMed  PubMed Central  Google Scholar 

  30. Hazlerigg DG, Ebling FJ, Johnston JD (2005) Photoperiod differentially regulates gene expression rhythms in the rostral and caudal SCN. Curr Biol 15:R449–R450

    CAS  PubMed  Google Scholar 

  31. Hermes ML, Buijs RM, Renaud LP (1996) Electrophysiology of suprachiasmatic nucleus projections to hypothalamic paraventricular nucleus neurons. Prog Brain Res 111:241–252

    CAS  PubMed  Google Scholar 

  32. Hermes ML, Coderre EM, Buijs RM, Renaud LP (1996) GABA and glutamate mediate rapid neurotransmission from suprachiasmatic nucleus to hypothalamic paraventricular nucleus in rat. J Physiol 496(Pt 3):749–757

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Herzog ED, Aton SJ, Numano R, Sakaki Y, Tei H (2004) Temporal precision in the mammalian circadian system: a reliable clock from less reliable neurons. J Biol Rhythms 19:35–46

    PubMed  Google Scholar 

  34. Herzog ED, Takahashi JS, Block GD (1998) Clock controls circadian period in isolated suprachiasmatic nucleus neurons. Nat Neurosci 1:708–713

    CAS  PubMed  Google Scholar 

  35. Honma S (2018) The mammalian circadian system: a hierarchical multi-oscillator structure for generating circadian rhythm. J Physiol Sci. https://doi.org/10.1007/s12576-018-0597-5

    Article  PubMed  Google Scholar 

  36. Honma S, Nakamura W, Shirakawa T, Honma K (2004) Diversity in the circadian periods of single neurons of the rat suprachiasmatic nucleus depends on nuclear structure and intrinsic period. Neurosci Lett 358:173–176

    CAS  PubMed  Google Scholar 

  37. Honma S, Shirakawa T, Katsuno Y, Namihira M, Honma K (1998) Circadian periods of single suprachiasmatic neurons in rats. Neurosci Lett 250:157–160

    CAS  PubMed  Google Scholar 

  38. Huhman KL, Hennessey AC, Albers HE (1996) Rhythms of glutamic acid decarboxylase mRNA in the suprachiasmatic nucleus. J Biol Rhythms 11:311–316

    CAS  PubMed  Google Scholar 

  39. Ikeda M, Toyoda H, Yamada J, Okabe A, Sato K, Hotta Y, Fukuda A (2003) Differential development of cation-chloride cotransporters and Cl- homeostasis contributes to differential GABAergic actions between developing rat visual cortex and dorsal lateral geniculate nucleus. Brain Res 984:149–159

    CAS  PubMed  Google Scholar 

  40. Ikeda M, Yoshioka T, Allen CN (2003) Developmental and circadian changes in Ca2+ mobilization mediated by GABAA and NMDA receptors in the suprachiasmatic nucleus. Eur J Neurosci 17:58–70

    PubMed  Google Scholar 

  41. Inagaki N, Honma S, Ono D, Tanahashi Y, Honma K (2007) Separate oscillating cell groups in mouse suprachiasmatic nucleus couple photoperiodically to the onset and end of daily activity. Proc Natl Acad Sci USA 104:7664–7669

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Inouye ST, Kawamura H (1979) Persistence of circadian rhythmicity in a mammalian hypothalamic “island” containing the suprachiasmatic nucleus. Proc Natl Acad Sci USA 76:5962–5966

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Irwin RP, Allen CN (2009) GABAergic signaling induces divergent neuronal Ca2+ responses in the suprachiasmatic nucleus network. Eur J Neurosci 30:1462–1475

    PubMed  PubMed Central  Google Scholar 

  44. Itri J, Colwell CS (2003) Regulation of inhibitory synaptic transmission by vasoactive intestinal peptide (VIP) in the mouse suprachiasmatic nucleus. J Neurophysiol 90:1589–1597

    CAS  PubMed  Google Scholar 

  45. Johnson SW, Seutin V (1997) Bicuculline methiodide potentiates NMDA-dependent burst firing in rat dopamine neurons by blocking apamin-sensitive Ca2+-activated K+ currents. Neurosci Lett 231:13–16

    CAS  PubMed  Google Scholar 

  46. Jones JR, Tackenberg MC, McMahon DG (2015) Manipulating circadian clock neuron firing rate resets molecular circadian rhythms and behavior. Nat Neurosci 18:373–375

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Kakizaki T, Oriuchi N, Yanagawa Y (2015) GAD65/GAD67 double knockout mice exhibit intermediate severity in both cleft palate and omphalocele compared with GAD67 knockout and VGAT knockout mice. Neuroscience 288:86–93

    CAS  PubMed  Google Scholar 

  48. Kalsbeek A, Drijfhout WJ, Westerink BH, van Heerikhuize JJ, van der Woude TP, van der Vliet J, Buijs RM (1996) GABA receptors in the region of the dorsomedial hypothalamus of rats are implicated in the control of melatonin and corticosterone release. Neuroendocrinology 63:69–78

    CAS  PubMed  Google Scholar 

  49. Kalsbeek A, Garidou ML, Palm IF, Van Der Vliet J, Simonneaux V, Pevet P, Buijs RM (2000) Melatonin sees the light: blocking GABA-ergic transmission in the paraventricular nucleus induces daytime secretion of melatonin. Eur J Neurosci 12:3146–3154

    CAS  PubMed  Google Scholar 

  50. Kalsbeek A, Palm IF, La Fleur SE, Scheer FA, Perreau-Lenz S, Ruiter M, Kreier F, Cailotto C, Buijs RM (2006) SCN outputs and the hypothalamic balance of life. J Biol Rhythms 21:458–469

    CAS  PubMed  Google Scholar 

  51. Khawaled R, Bruening-Wright A, Adelman JP, Maylie J (1999) Bicuculline block of small-conductance calcium-activated potassium channels. Pflugers Arch 438:314–321

    CAS  PubMed  Google Scholar 

  52. Klett NJ, Allen CN (2017) Intracellular chloride regulation in AVP+ and VIP+ neurons of the suprachiasmatic nucleus. Sci Rep 7:10226

    PubMed  PubMed Central  Google Scholar 

  53. Kume K, Zylka MJ, Sriram S, Shearman LP, Weaver DR, Jin X, Maywood ES, Hastings MH, Reppert SM (1999) mCRY1 and mCRY2 are essential components of the negative limb of the circadian clock feedback loop. Cell 98:193–205

    CAS  PubMed  Google Scholar 

  54. Leak RK, Moore RY (2001) Topographic organization of suprachiasmatic nucleus projection neurons. J Comp Neurol 433:312–334

    CAS  PubMed  Google Scholar 

  55. Lee IT, Chang AS, Manandhar M, Shan Y, Fan J, Izumo M, Ikeda Y, Motoike T, Dixon S, Seinfeld JE, Takahashi JS, Yanagisawa M (2015) Neuromedin S-producing neurons act as essential pacemakers in the suprachiasmatic nucleus to couple clock neurons and dictate circadian rhythms. Neuron 85:1086–1102

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Lehman MN, Silver R, Gladstone WR, Kahn RM, Gibson M, Bittman EL (1987) Circadian rhythmicity restored by neural transplant. Immunocytochemical characterization of the graft and its integration with the host brain. J Neurosci 7:1626–1638

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Liou SY, Albers HE (1990) Single unit response of neurons within the hamster suprachiasmatic nucleus to GABA and low chloride perfusate during the day and night. Brain Res Bull 25:93–98

    CAS  PubMed  Google Scholar 

  58. Liou SY, Shibata S, Albers HE, Ueki S (1990) Effects of GABA and anxiolytics on the single unit discharge of suprachiasmatic neurons in rat hypothalamic slices. Brain Res Bull 25:103–107

    CAS  PubMed  Google Scholar 

  59. Liu AC, Welsh DK, Ko CH, Tran HG, Zhang EE, Priest AA, Buhr ED, Singer O, Meeker K, Verma IM, Doyle FJ 3rd, Takahashi JS, Kay SA (2007) Intercellular coupling confers robustness against mutations in the SCN circadian clock network. Cell 129:605–616

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Liu C, Reppert SM (2000) GABA synchronizes clock cells within the suprachiasmatic circadian clock. Neuron 25:123–128

    CAS  PubMed  Google Scholar 

  61. Liu C, Weaver DR, Jin X, Shearman LP, Pieschl RL, Gribkoff VK, Reppert SM (1997) Molecular dissection of two distinct actions of melatonin on the suprachiasmatic circadian clock. Neuron 19:91–102

    CAS  PubMed  Google Scholar 

  62. Mai JK, Kedziora O, Teckhaus L, Sofroniew MV (1991) Evidence for subdivisions in the human suprachiasmatic nucleus. J Comp Neurol 305:508–525

    CAS  PubMed  Google Scholar 

  63. Marpegan L, Krall TJ, Herzog ED (2009) Vasoactive intestinal polypeptide entrains circadian rhythms in astrocytes. J Biol Rhythms 24:135–143

    CAS  PubMed  PubMed Central  Google Scholar 

  64. Marpegan L, Swanstrom AE, Chung K, Simon T, Haydon PG, Khan SK, Liu AC, Herzog ED, Beaule C (2011) Circadian regulation of ATP release in astrocytes. J Neurosci 31:8342–8350

    CAS  PubMed  PubMed Central  Google Scholar 

  65. Mason R, Biello SM, Harrington ME (1991) The effects of GABA and benzodiazepines on neurones in the suprachiasmatic nucleus (SCN) of Syrian hamsters. Brain Res 552:53–57

    CAS  PubMed  Google Scholar 

  66. Maywood ES, Chesham JE, O’Brien JA, Hastings MH (2011) A diversity of paracrine signals sustains molecular circadian cycling in suprachiasmatic nucleus circuits. Proc Natl Acad Sci USA 108:14306–14311

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Maywood ES, Reddy AB, Wong GK, O’Neill JS, O’Brien JA, McMahon DG, Harmar AJ, Okamura H, Hastings MH (2006) Synchronization and maintenance of timekeeping in suprachiasmatic circadian clock cells by neuropeptidergic signaling. Curr Biol 16:599–605

    CAS  PubMed  Google Scholar 

  68. Mieda M, Ono D, Hasegawa E, Okamoto H, Honma K, Honma S, Sakurai T (2015) Cellular clocks in AVP neurons of the SCN are critical for interneuronal coupling regulating circadian behavior rhythm. Neuron 85:1103–1116

    CAS  PubMed  Google Scholar 

  69. Moldavan MG, Allen CN (2013) GABAB receptor-mediated frequency-dependent and circadian changes in synaptic plasticity modulate retinal input to the suprachiasmatic nucleus. J Physiol 591:2475–2490

    CAS  PubMed  PubMed Central  Google Scholar 

  70. Moldavan MG, Irwin RP, Allen CN (2006) Presynaptic GABA(B) receptors regulate retinohypothalamic tract synaptic transmission by inhibiting voltage-gated Ca2+ channels. J Neurophysiol 95:3727–3741

    CAS  PubMed  Google Scholar 

  71. Moore RY, Eichler VB (1972) Loss of a circadian adrenal corticosterone rhythm following suprachiasmatic lesions in the rat. Brain Res 42:201–206

    CAS  PubMed  Google Scholar 

  72. Moore RY, Klein DC (1974) Visual pathways and the central neural control of a circadian rhythm in pineal serotonin N-acetyltransferase activity. Brain Res 71:17–33

    CAS  PubMed  Google Scholar 

  73. Moore RY, Speh JC (1993) GABA is the principal neurotransmitter of the circadian system. Neurosci Lett 150:112–116

    CAS  PubMed  Google Scholar 

  74. Moore RY, Speh JC, Leak RK (2002) Suprachiasmatic nucleus organization. Cell Tissue Res 309:89–98

    CAS  PubMed  Google Scholar 

  75. Myung J, Hong S, DeWoskin D, De Schutter E, Forger DB, Takumi T (2015) GABA-mediated repulsive coupling between circadian clock neurons in the SCN encodes seasonal time. Proc Natl Acad Sci USA 112:E3920–E3929

    CAS  PubMed  PubMed Central  Google Scholar 

  76. Nagano M, Adachi A, Nakahama K, Nakamura T, Tamada M, Meyer-Bernstein E, Sehgal A, Shigeyoshi Y (2003) An abrupt shift in the day/night cycle causes desynchrony in the mammalian circadian center. J Neurosci 23:6141–6151

    CAS  PubMed  PubMed Central  Google Scholar 

  77. Naito E, Watanabe T, Tei H, Yoshimura T, Ebihara S (2008) Reorganization of the suprachiasmatic nucleus coding for day length. J Biol Rhythms 23:140–149

    PubMed  Google Scholar 

  78. Nakamura TJ, Takasu NN, Nakamura W (2016) The suprachiasmatic nucleus: age-related decline in biological rhythms. J Physiol Sci 66:367–374

    CAS  PubMed  Google Scholar 

  79. Nakamura W, Honma S, Shirakawa T, Honma K (2002) Clock mutation lengthens the circadian period without damping rhythms in individual SCN neurons. Nat Neurosci 5:399–400

    CAS  PubMed  Google Scholar 

  80. Nakamura W, Yamazaki S, Nakamura TJ, Shirakawa T, Block GD, Takumi T (2008) In vivo monitoring of circadian timing in freely moving mice. Curr Biol 18:381–385

    CAS  PubMed  Google Scholar 

  81. Nakao N, Ono H, Yamamura T, Anraku T, Takagi T, Higashi K, Yasuo S, Katou Y, Kageyama S, Uno Y, Kasukawa T, Iigo M, Sharp PJ, Iwasawa A, Suzuki Y, Sugano S, Niimi T, Mizutani M, Namikawa T, Ebihara S, Ueda HR, Yoshimura T (2008) Thyrotrophin in the pars tuberalis triggers photoperiodic response. Nature 452:317–322

    CAS  PubMed  Google Scholar 

  82. Ohta H, Yamazaki S, McMahon DG (2005) Constant light desynchronizes mammalian clock neurons. Nat Neurosci 8:267–269

    CAS  PubMed  Google Scholar 

  83. Okamura H, Berod A, Julien JF, Geffard M, Kitahama K, Mallet J, Bobillier P (1989) Demonstration of GABAergic cell bodies in the suprachiasmatic nucleus: in situ hybridization of glutamic acid decarboxylase (GAD) mRNA and immunocytochemistry of GAD and GABA. Neurosci Lett 102:131–136

    CAS  PubMed  Google Scholar 

  84. Ono D, Honma S, Honma K (2013) Cryptochromes are critical for the development of coherent circadian rhythms in the mouse suprachiasmatic nucleus. Nature Commun 4:1666

    Google Scholar 

  85. Ono D, Honma S, Honma K (2016) Differential roles of AVP and VIP signaling in the postnatal changes of neural networks for coherent circadian rhythms in the SCN. Sci Adv 2:e1600960

    PubMed  PubMed Central  Google Scholar 

  86. Ono D, Honma S, Nakajima Y, Kuroda S, Enoki R, Honma KI (2017) Dissociation of Per1 and Bmal1 circadian rhythms in the suprachiasmatic nucleus in parallel with behavioral outputs. Proc Natl Acad Sci USA 114:E3699–E3708

    CAS  PubMed  PubMed Central  Google Scholar 

  87. Osakada F, Mori T, Cetin AH, Marshel JH, Virgen B, Callaway EM (2011) New rabies virus variants for monitoring and manipulating activity and gene expression in defined neural circuits. Neuron 71:617–631

    CAS  PubMed  PubMed Central  Google Scholar 

  88. Perea G, Navarrete M, Araque A (2009) Tripartite synapses: astrocytes process and control synaptic information. Trends Neurosci 32:421–431

    CAS  PubMed  Google Scholar 

  89. Pittendrigh CS, Daan S (1976) Functional-analysis of circadian pacemakers in nocturnal rodents. 5. Pacemaker structure—clock for all seasons. J Comp Physiol 106:333–355

    Google Scholar 

  90. Prolo LM, Takahashi JS, Herzog ED (2005) Circadian rhythm generation and entrainment in astrocytes. J Neurosci 25:404–408

    CAS  PubMed  PubMed Central  Google Scholar 

  91. Ralph MR, Foster RG, Davis FC, Menaker M (1990) Transplanted suprachiasmatic nucleus determines circadian period. Science (New York, N.Y.) 247:975–978

    CAS  Google Scholar 

  92. Ran FA, Cong L, Yan WX, Scott DA, Gootenberg JS, Kriz AJ, Zetsche B, Shalem O, Wu XB, Makarova KS, Koonin EV, Sharp PA, Zhang F (2015) In vivo genome editing using Staphylococcus aureus Cas9. Nature 520:U186–U198

    Google Scholar 

  93. Reppert SM, Weaver DR (2002) Coordination of circadian timing in mammals. Nature 418:935–941

    CAS  PubMed  Google Scholar 

  94. Saito K, Kakizaki T, Hayashi R, Nishimaru H, Furukawa T, Nakazato Y, Takamori S, Ebihara S, Uematsu M, Mishina M, Miyazaki J, Yokoyama M, Konishi S, Inoue K, Fukuda A, Fukumoto M, Nakamura K, Obata K, Yanagawa Y (2010) The physiological roles of vesicular GABA transporter during embryonic development: a study using knockout mice. Mol Brain 3:40

    CAS  PubMed  PubMed Central  Google Scholar 

  95. Sato T, Kawamura H (1984) Circadian rhythms in multiple unit activity inside and outside the suprachiasmatic nucleus in the diurnal chipmunk (Eutamias sibiricus). Neurosci Res 1:45–52

    CAS  PubMed  Google Scholar 

  96. Sato TK, Yamada RG, Ukai H, Baggs JE, Miraglia LJ, Kobayashi TJ, Welsh DK, Kay SA, Ueda HR, Hogenesch JB (2006) Feedback repression is required for mammalian circadian clock function. Nat Genet 38:312–319

    CAS  PubMed  PubMed Central  Google Scholar 

  97. Schwartz WJ, Gross RA, Morton MT (1987) The suprachiasmatic nuclei contain a tetrodotoxin-resistant circadian pacemaker. Proc Natl Acad Sci USA 84:1694–1698

    CAS  PubMed  PubMed Central  Google Scholar 

  98. Schwarz LA, Miyamichi K, Gao XJ, Beier KT, Weissbourd B, DeLoach KE, Ren J, Ibanes S, Malenka RC, Kremer EJ, Luo L (2015) Viral-genetic tracing of the input-output organization of a central noradrenaline circuit. Nature 524:88–92

    CAS  PubMed  PubMed Central  Google Scholar 

  99. Scott FF, Belle MD, Delagrange P, Piggins HD (2010) Electrophysiological effects of melatonin on mouse Per1 and non-Per1 suprachiasmatic nuclei neurones in vitro. J Neuroendocrinol 22:1148–1156

    CAS  PubMed  Google Scholar 

  100. Shearman LP, Sriram S, Weaver DR, Maywood ES, Chaves I, Zheng B, Kume K, Lee CC, van der Horst GT, Hastings MH, Reppert SM (2000) Interacting molecular loops in the mammalian circadian clock. Science (New York, N.Y.) 288:1013–1019

    CAS  Google Scholar 

  101. Shirakawa T, Honma S, Katsuno Y, Oguchi H, Honma KI (2000) Synchronization of circadian firing rhythms in cultured rat suprachiasmatic neurons. Eur J Neurosci 12:2833–2838

    CAS  PubMed  Google Scholar 

  102. Silver R, LeSauter J, Tresco PA, Lehman MN (1996) A diffusible coupling signal from the transplanted suprachiasmatic nucleus controlling circadian locomotor rhythms. Nature 382:810–813

    CAS  PubMed  Google Scholar 

  103. Silver R, Romero MT, Besmer HR, Leak R, Nunez JM, LeSauter J (1996) Calbindin-D28 K cells in the hamster SCN express light-induced Fos. NeuroReport 7:1224–1228

    CAS  PubMed  Google Scholar 

  104. Stephan FK, Berkley KJ, Moss RL (1981) Efferent connections of the rat suprachiasmatic nucleus. Neuroscience 6:2625–2641

    CAS  PubMed  Google Scholar 

  105. Stephan FK, Zucker I (1972) Circadian rhythms in drinking behavior and locomotor activity of rats are eliminated by hypothalamic lesions. Proc Natl Acad Sci USA 69:1583–1586

    CAS  PubMed  PubMed Central  Google Scholar 

  106. Strecker GJ, Wuarin JP, Dudek FE (1997) GABAA-mediated local synaptic pathways connect neurons in the rat suprachiasmatic nucleus. J Neurophysiol 78:2217–2220

    CAS  PubMed  Google Scholar 

  107. Sujino M, Masumoto KH, Yamaguchi S, van der Horst GT, Okamura H, Inouye ST (2003) Suprachiasmatic nucleus grafts restore circadian behavioral rhythms of genetically arrhythmic mice. Curr Biol 13:664–668

    CAS  PubMed  Google Scholar 

  108. Tanaka M, Hayashi S, Tamada Y, Ikeda T, Hisa Y, Takamatsu T, Ibata Y (1997) Direct retinal projections to GRP neurons in the suprachiasmatic nucleus of the rat. NeuroReport 8:2187–2191

    CAS  PubMed  Google Scholar 

  109. Tervo DG, Hwang BY, Viswanathan S, Gaj T, Lavzin M, Ritola KD, Lindo S, Michael S, Kuleshova E, Ojala D, Huang CC, Gerfen CR, Schiller J, Dudman JT, Hantman AW, Looger LL, Schaffer DV, Karpova AY (2016) A designer AAV variant permits efficient retrograde access to projection neurons. Neuron 92:372–382

    CAS  PubMed  PubMed Central  Google Scholar 

  110. Teshima K, Kim SH, Allen CN (2003) Characterization of an apamin-sensitive potassium current in suprachiasmatic nucleus neurons. Neuroscience 120:65–73

    CAS  PubMed  Google Scholar 

  111. Tousson E, Meissl H (2004) Suprachiasmatic nuclei grafts restore the circadian rhythm in the paraventricular nucleus of the hypothalamus. J Neurosci 24:2983–2988

    CAS  PubMed  PubMed Central  Google Scholar 

  112. 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

    CAS  PubMed  PubMed Central  Google Scholar 

  113. Van den Pol AN (1980) The hypothalamic suprachiasmatic nucleus of rat: intrinsic anatomy. J Comp Neurol 191:661–702

    PubMed  Google Scholar 

  114. VanderLeest HT, Houben T, Michel S, Deboer T, Albus H, Vansteensel MJ, Block GD, Meijer JH (2007) Seasonal encoding by the circadian pacemaker of the SCN. Curr Biol 17:468–473

    CAS  PubMed  Google Scholar 

  115. Vansteensel MJ, Yamazaki S, Albus H, Deboer T, Block GD, Meijer JH (2003) Dissociation between circadian Per1 and neuronal and behavioral rhythms following a shifted environmental cycle. Curr Biol 13:1538–1542

    CAS  PubMed  Google Scholar 

  116. Wagner S, Castel M, Gainer H, Yarom Y (1997) GABA in the mammalian suprachiasmatic nucleus and its role in diurnal rhythmicity. Nature 387:598–603

    CAS  PubMed  Google Scholar 

  117. Watts AG, Swanson LW (1987) Efferent projections of the suprachiasmatic nucleus: II. Studies using retrograde transport of fluorescent dyes and simultaneous peptide immunohistochemistry in the rat. J Comp Neurol 258:230–252

    CAS  PubMed  Google Scholar 

  118. Watts AG, Swanson LW, Sanchez-Watts G (1987) Efferent projections of the suprachiasmatic nucleus: I. Studies using anterograde transport of Phaseolus vulgaris leucoagglutinin in the rat. J Comp Neurol 258:204–229

    CAS  PubMed  Google Scholar 

  119. Webb AB, Angelo N, Huettner JE, Herzog ED (2009) Intrinsic, nondeterministic circadian rhythm generation in identified mammalian neurons. Proc Natl Acad Sci USA 106:16493–16498

    CAS  PubMed  PubMed Central  Google Scholar 

  120. Welsh DK, Logothetis DE, Meister M, Reppert SM (1995) Individual neurons dissociated from rat suprachiasmatic nucleus express independently phased circadian firing rhythms. Neuron 14:697–706

    CAS  PubMed  Google Scholar 

  121. Yamaguchi S, Isejima H, Matsuo T, Okura R, Yagita K, Kobayashi M, Okamura H (2003) Synchronization of cellular clocks in the suprachiasmatic nucleus. Science (New York, N.Y.) 302:1408–1412

    CAS  Google Scholar 

  122. Yamazaki S, Kerbeshian MC, Hocker CG, Block GD, Menaker M (1998) Rhythmic properties of the hamster suprachiasmatic nucleus in vivo. J Neurosci 18:10709–10723

    CAS  PubMed  PubMed Central  Google Scholar 

  123. Yamazaki S, Numano R, Abe M, Hida A, Takahashi R, Ueda M, Block GD, Sakaki Y, Menaker M, Tei H (2000) Resetting central and peripheral circadian oscillators in transgenic rats. Science (New York, N.Y.) 288:682–685

    CAS  Google Scholar 

  124. Yan L, Foley NC, Bobula JM, Kriegsfeld LJ, Silver R (2005) Two antiphase oscillations occur in each suprachiasmatic nucleus of behaviorally split hamsters. J Neurosci 25:9017–9026

    CAS  PubMed  PubMed Central  Google Scholar 

  125. Yoo SH, Yamazaki S, Lowrey PL, Shimomura K, Ko CH, Buhr ED, Siepka SM, Hong HK, Oh WJ, Yoo OJ, Menaker M, Takahashi JS (2004) PERIOD2:LUCIFERASE real-time reporting of circadian dynamics reveals persistent circadian oscillations in mouse peripheral tissues. Proc Natl Acad Sci USA 101:5339–5346

    CAS  PubMed  PubMed Central  Google Scholar 

  126. Yoshimura T, Yasuo S, Watanabe M, Iigo M, Yamamura T, Hirunagi K, Ebihara S (2003) Light-induced hormone conversion of T4 to T3 regulates photoperiodic response of gonads in birds. Nature 426:178–181

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

This work was supported in part by the Uehara Memorial Foundation, the Nakajima Foundation, GSK Japan Research Grant 2015, the Project for Developing Innovation Systems of the MEXT, and Creation of Innovation Centers for Advanced Interdisciplinary Research Areas Program, Ministry of Education, Culture, Sports, Science and Technology, Japan, and JSPS KAKENHI (Nos. 15H04679, 26860156, 15K12763).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Daisuke Ono.

Ethics declarations

Ethical approval

All applicable international, national, and/or institutional guidelines for the care and use of animals were followed. This article does not contain any studies with human participants performed by any of the authors.

Conflict of interest

The authors declare no conflicts of interest.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ono, D., Honma, Ki., Yanagawa, Y. et al. Role of GABA in the regulation of the central circadian clock of the suprachiasmatic nucleus. J Physiol Sci 68, 333–343 (2018). https://doi.org/10.1007/s12576-018-0604-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s12576-018-0604-x

Keywords