Abstract

LKB1 is known as a master kinase for 14 kinases related to the adenosine monophosphate (AMP)-activated protein kinase (AMPK). Two of them (SIK3 and AMPKa) have previously been implicated in sleep regulation. We generated loss-of-function (LOF) mutants for Lkb1 in both Drosophila and mice. Sleep, but not circadian rhythms, was reduced in Lkb1-mutant flies and in flies with neuronal deletion of Lkb1. Genetic interactions between Lkb1 and AMPK T184A mutants in Drosophila sleep or those between Lkb1 and SIK3 T196A mutants in Drosophila viability have been observed. Sleep was reduced in mice after virally mediated reduction of Lkb1 in the brain. Electroencephalography (EEG) analysis showed that non-rapid eye movement (NREM) sleep and sleep need were both reduced in Lkb1-mutant mice. These results indicate that LKB1 plays a physiological role in sleep regulation conserved from flies to mice.

Introduction

Human Peutz-Jeghers syndrome (PJS) (Peutz 1921; Jeghers et al. 1949) is an autosomal dominant disorder with gastrointestinal (GI) polyps and increased cancer risk of multiple tissues (Tomlinson and Houlston 1997; Westerman et al. 1999). The gene mutated in, and responsible for, PJS encodes the liver kinase B1 (LKB1, also known as STK11) (Hemminki et al. 1997; 1998; Jenne et al. 1998). Lkb1 is thus a tumor suppressor gene, mutated in multiple cancers, especially the GI (Mehenni et al. 1998; Bardeesy et al. 2002; Jishage et al. 2002; Miyoshi et al. 2002; Hearle et al. 2006) and lung adenocarcinoma (Sanchez-Cespedes et al. 2002; Carretero et al. 2004; Ji et al. 2007; Matsumoto et al. 2007; Gill et al. 2011; Skoulidis et al. 2015), cervical cancer (Wingo et al. 2009), ovarian cancer (Tanwar et al. 2014), breast cancer (Shen et al. 2002; Hearle et al. 2006; Sengupta et al. 2017), pancreatic cancer (Morton et al. 2010), and melanoma (Guldberg et al. 1999; Rowan et al. 1999).

LKB1 phosphorylates threonine 172 (T172) of the α subunit of adenosine monophosphate (AMP)-activated protein kinase (AMPKα) (Hawley et al. 2003; Hong et al. 2003; Sutherland et al. 2003; Woods et al. 2003; Lizcano et al. 2004; Shaw et al. 2004, 2005; Sakamoto et al. 2005), and positively regulates the activity of AMPK.

AMPK is a well-known kinase (Beg et al. 1973; Carlson and Kim 1973; Ingebritsen et al. 1978; Yeh and Kim 1980; Ferrer et al. 1985; Carling et al. 1987, 1989; Munday et al. 1988) with important physiological and pathological roles (Hardie 2014; Lopez and Dieguez 2014; Hardie et al. 2016; Herzig and Shaw 2018). The α, β, and γ subunits of AMPK form a heterotrimeric complex (Davies et al. 1994; Mitchelhill et al. 1994; Michell et al. 1996). The catalytic α subunit is regulated by phosphorylation at T172 of AMPKα2 or T183 of AMPKα1 (Hawley et al. 1996).

There are 12 additional mammalian AMPK-related kinases (ARKs), similar to the α subunit of AMPK, all regulated at the site equivalent to AMPK-T172 (Lizcano et al. 2004). LKB1 and its associated proteins STE20-related adaptor (STRAD) and mouse protein 25 (MO25) have been reported to phosphorylate all 14 ARKs (Lizcano et al. 2004), making LKB1 a master kinase for ARKs (Lizcano et al. 2004; Alessi et al. 2006; Shackelford and Shaw 2009). We have recently found that more than 20 kinases in the STE20 family of mammalian serine-threonine kinases could phosphorylate ARKs in vitro, though the physiological roles of STE20 kinases in ARK phosphorylation remain unknown (Liu, Wang, Cui, Gao, et al. 2022; Liu, Wang, Cui, et al. 2022).

Some ARKs have been reported to regulate sleep. In mice, inhibitors of AMPK were found to decrease sleep, whereas activators of AMPK were found to increase sleep (Chikahisa et al. 2009). In flies, knockdown of AMPKβ in neurons decreased the total amount of sleep and resulted in fragmented sleep (Nagy et al. 2018). Knockdown of AMPKα in a specific pair of neurons suppressed sleep (Yurgel et al. 2019).

Studies in mice have shown that sleep was increased in gain-of-function (GOF) mutations in the salt inducible kinase (SIK) 3 (Funato et al. 2016), and sleep need was reduced in GOF mutants of SIK 1, 2, and 3 (Funato et al. 2016; Honda et al. 2018; Park et al. 2020). Sleep was also decreased when SIK3 was downregulated in flies (Funato et al. 2016). Here we investigated the functional role of LKB1 in regulating sleep in flies and mice.

Materials and methods

Fly lines and rearing conditions

Flies were reared on standard corn meal at 25°C, 60% humidity and kept in 12 hours (h) light/12 h dark (LD) conditions. 57C10-Gal4, nos-phiC31, hs-Cre on X were from the Bloomington Stock Center. vas-Cas9 was a gift from Dr J. Ni (Tsinghua University, Beijing). Upstream activation sequence (UAS)-Cas9, UAS-Ampk, UAS-Ampk-T184A, UAS-Ampk-T184E, Sik3-flag, Sik3–T196A-flag, and Sik3-T196E-flag flies were from our laboratory.

Flies used in behavioral assays were backcrossed into our isogenized Canton S background for 7 generations. All results of sleep analysis in this paper were obtained from female flies.

Generation of KO, KI, and transgenic lines

Total RNA was extracted from isoCS by TRIzol reagent (Invitrogen). Using the PrimeScript II 1st Strand cDNA Synthesis Kit (Takara), we reverse-transcribed the extracted mRNA into cDNA. The UAS-Lkb1 flies was constructed by inserting the coding sequence (CDS) of CG9374 amplified from cDNA into the pACU2 plasmid (a gift from the Jan Lab at UCSF) (Han et al. 2011) before being inserted into the attP40 site.

The UAS-Lkb1-sgRNA construct was designed by inserting the sgRNAs into pMt: sgRNA3XEF vectors based on pACU2, with rice tRNA separating the different sgRNAs. CRISPR-Gold website was used to design 3 sgRNAs of Lkb1 (Supplementary Fig. 3) (Chu et al. 2016; Poe et al. 2019). The construct was inserted into the attP40 site.

KO and KI lines were generated as described previously (Deng et al. 2019). Knockout flies were generated with the CRISPR/Cas9 system. Two different sgRNAs were constructed with U6b-sgRNA plasmids. The 5′ homologous arm and the 3′ homologous arm of ∼2 kb amplified from the wt fly genome were inserted into a pBSK plasmid for homologous recombination repair. The cassette of attP-3P3-RFP was introduced in the middle. sgRNA plasmids and the donor plasmids were injected into vas-Cas9 embryos to introduce attP-3P3-RFP into the genome at the region of interest and replaced it by homologous recombination. 3P3-RFP served as a marker to screen for the correct flies. Primers across the homologous arms were designed to verify the sequences by PCR and DNA sequencing. attP site was introduced into the genome with 3P3-RFP-LoxP. For KI files, the nos-phiC31 virgin females were first crossed with knock-out males and the pBSK plasmid inserted with attB-T2A-Gal4-miniwhite-LoxP cassette was injected into the female embryos. Miniwhite serves as a marker to screen for the correct flies, which could be excised by the Cre/LoxP system. Primers were designed to verify the sequence by PCR and DNA sequencing.

Quantitative PCR

Total RNA was extracted from 30 flies of 5–7 days old by TRIzol reagent (Invitrogen). The genomic template was removed using DNase (Takara). cDNA was reverse-transcribed using Takara’s PrimeScript II 1st Strand cDNA synthesis kit (Takara). Quantitative PCR was carried out with TransStart Top Green qPCR SuperMix kit (TransGen) in the Bio-Rad PCR system (CFX96 Touch Deep Well). The sequences of primers used to detect Lkb1 and RP49(endogenous control) mRNA are

Lkb1-F: 5′-GCCGTCAAGATCCTGACTA-3′

Lkb1-R: 5′-CTCCGCTGGACCAGATG-3′

Rp49-F: 5′-CGACGCTTCAAGGGACAGTATC-3′

Rp49-R: 5′-TCCGACCAGGTTACAAGAACTCTC-3′

Drosophila sleep analysis

Drosophila sleep analysis was performed as described previously (Qian et al. 2017; Dai et al. 2019). 5–7 days old flies were placed in a 65 mm × 5 mm clear glass tube with one end containing food and the other end plugging with cotton. All flies were recorded by video-cameras. Before sleep measurement, flies were entrained to an LD cycle at 25°C, 60% humidity for at least 2 days, and infrared LED light was used to ensure constant illumination when lights off. Immobility longer than 5 min was defined as one sleep event (Hendricks et al. 2000; Shaw et al. 2000). Information of fly location was tracked and sleep parameters were analyzed using Matlab (Mathworks), from which dead flies were removed. Sleep duration, sleep bout duration, sleep bout number, and sleep latency for each LD were analyzed. Each experiment was repeated at least 3 times.

Drosophila circadian analysis

Flies were reared and recorded in the same condition as sleep assay as described in papers from our laboratory (Qian et al. 2017; Dai et al. 2021), except that the condition was constant darkness. Six to 8 days activity was measured and calculated in ActogramJ (Klarsfeld et al. 2003). Rhythmic strength, power and period were calculated by Chi-square method.

Immunoblot analysis

Mouse brains were quickly dissected and washed with phosphate buffer saline on ice. Lysis buffer (20 mM HEPES, 10 mM KCl, 1.5 mM MgCl2, 1 mM EDTA, 1 mM EGTA, 1 mM DTT, freshly supplemented with a protease and phosphatase inhibitors cocktail) were used to homogenize brains by homogenizer (Wiggens, D-500 pro) at 4°C. Brain homogenates were centrifuged at 14,000 revolution per minutes for 15 min at 4°C. The supernatant was transferred to a new microtube and quantified with bicinchoninic acid assay (Thermo Fisher, 23225). The supernatant was analyzed by SDS-PAGE and proteins were transferred to a nitrocellulose membrane (GE Healthcare, #BA85). Membranes were incubated for 1 h in a blocking solution (tris-buffered saline containing 0.1% Tween-20, 5% milk). Primary antibodies were anti-LKB1 (cell signaling, #3047) and anti-ACTIN (Santa Cruz, sc-8342).

Retro-orbital injection in mice

Mice were reared at controlled temperature and humidity conditions with 12 h light/12 h dark cycle. Food and water were provided ad libitum. Lkb1fl/fl mice were from the Jackson Laboratory (JAX #014143). They contained loxP sites flanking exons 3–6 of Lkb1 gene (Nakada et al. 2010). adeno-associated viral (AAV)-PHP.eB-hSyn-Cre-green fluorescent protein (GFP) and AAV-PHP.eB-hSyn-GFP virus were from Chinese Institute for Brain Research, Beijing. All results of sleep analysis in this paper were obtained from female mice.

Total 0.2 ml/10 g avertin was injected intraperitoneally into the mice for anesthetization. Rodent eyes were protruded by gentle downward pressure to the skin on the dorsal and ventral sides of the eye. The operator inserted the needle beveled downward into the retro-orbital sinus at the medial corner of the eye (Yardeni et al. 2011). The AAV-PHP.eB virus was injected for whole brain infection (Chan et al. 2017).

Mouse sleep analysis

Mouse sleep analysis was described in a previous article from our laboratory (Zhang et al. 2018). Eight-week-old mice were selected for retro-orbital injection. One week after viral injection, EEG and EMG electrode implantation procedures were performed. Mice were allowed to recover for more than 5 days individually and placed in a recording cage and tethered to an omni-directional arm (RWD Life Science Inc.) with connection cable for 2 days of habituation before recording. EEG and EMG data were recorded with custom software at a sampling frequency of 200 Hz for 2 consecutive days to analyze sleep/wake behavior under baseline conditions. The recording chamber was maintained at 12 h LD cycle and controlled temperature (24–25°C). EEG/EMG data were initially processed by Accusleep (Barger et al. 2019) before manual correction in SleepSign to improve accuracy. WAKE was scored as high amplitude and variable EMG and fast and low amplitude EEG. Nonrapid eye movement (NREM) was scored as high amplitude δ (1–4 Hz) frequency EEG and low EMG tonus. REM was scored as a complete silent of EMG signals and low amplitude high frequency θ (6–9 Hz)-dominated EEG signal.

For power spectrum analysis, EEG was subjected to fast Fourier transform analysis with a Hamming window method by SleepSign, yielding power spectra between 0 and 25 Hz with a 0.39 Hz bin resolution. Epochs containing movement artifacts were marked, included in sleep duration analysis but excluded from the power spectra analysis. Power spectra for each vigilance state represents the mean power distribution of this state during a 24-h baseline recording. The δ-power density of NREMs per hour represents the average of δ-power density as a percentage of δ-band power (1–4 Hz) to total power (0–25 Hz) for each NREM epoch contained in an hour.

Statistics

All statistical analyses were performed with Prism 7 (GraphPad Software). Differences in means between samples larger than 2 groups were analyzed using ordinary 1-way ANOVA. Unpaired t test was used for 2 groups comparison. Power spectrum between different lines was compared by 2-way ANOVA followed by Turkey’s multiple comparisons test. n.s. denotes P > 0.05; *P < 0.05; **P < 0.01; and ***P < 0.001 for all statistical results in this paper.

Results

Sleep phenotypes of Drosophila Lkb1 mutants

Null mutants for Lkb1 are lethal in Drosophila (Martin and St Johnston 2003). We had generated a Lkb1 knockout (“lkb1T1”) line (Supplementary Figs. 1a and 3b) and found that lkb1T1/T1 mutation was lethal in the pupa stage. The level of Lkb1 mRNA was reduced in the heterozygous lkb1T1/+ flies (Supplementary Fig. 1b). We then tested whether the heterozygous lkb1T1 had any phenotype in sleep using flies kept in 12 hours (h) light/12 h dark (LD) cycles (Supplementary Fig. 1c). While lkb1T1/+ flies were not significantly different from the wild-type (wt) flies in sleep bout numbers (Supplementary Fig. 1e), or daytime sleep duration (Supplementary Fig. 1d), daytime sleep bout duration (Supplementary Fig. 1f), lkb1T1/+ flies showed significantly lower nighttime sleep duration (Supplementary Fig. 1d), nighttime sleep bout duration (Supplementary Fig. 1f), and longer latency to sleep (Supplementary Fig. 1g). Thus, there was dosage-sensitive physiological requirement of Lkb1 in nighttime sleep.

We tried to, and succeeded in, generating lkb1T2, a hypomorphic mutation for Lkb1 in flies (Fig. 1a;Supplementary Fig. 3, a and b). Lkb1 mRNA was significantly reduced in lkb1T2/+ and lkb1T2/T2 flies (Fig. 1b). During the day, lkb1T2/T2 flies were not significantly different from the lkb1T2/+ and wt flies in sleep duration (Fig. 1, c and d), sleep bout number (Fig. 1e), sleep bout duration (Fig. 1f), or latency to sleep (Fig. 1g). During the night, not only lkb1T2/T2 flies showed significantly reduced sleep duration (Fig. 1, c and d), highly reduced sleep bout duration (Fig. 1f) and highly increased latency (Fig. 1g) than the wt flies, but also the heterozygous lkb1T2/+ flies were significantly different from the wt flies in all these parameters (Fig. 1, c–g), indicating a dosage sensitive requirement for Lkb1.

Fig. 1.

Sleep phenotypes of Lkb1 knock-down mutant flies. a) A diagram illustrating the Lkb1 insertion mutant lkb1T2. b) Relative Lkb1 mRNA levels in lkb1T2/T2 (red), lkb1T2/+ (blue), and wt (black) flies. c) Sleep profiles of lkb1T2/T2 (red, n = 42), lkb1T2/+ (blue, n = 44), and wt (black, n = 44) flies in a 12 h light/12 h dark (LD) cycle. Statistical analysis of sleep duration, sleep bout number, sleep bout duration, and latency to sleep in lkb1T2/T2 (red, n = 42), lkb1T2/+ (blue, n = 44), and wt (black, n = 44) flies. Open bars denote daytime, filled bars denote nighttime. d) Sleep duration. Nighttime sleep durations of lkb1T2/T2 mutants were significantly less than those in lkb1T2/+ and wt flies. e) Sleep bout number. Daytime sleep bout number of lkb1T2/T2 mutants was less than that of wt flies. f) Sleep bout duration. Nighttime sleep bout duration of lkb1T2/T2 mutants was significantly less than those of lkb1T2/+ and wt flies. g) Latency to sleep. Latency to sleep after light-off of lkb1T2/T2 mutants was significantly prolonged than lkb1T2/+ and wt flies. One-way ANOVA was used. n.s. denotes P > 0.05, *P < 0.05, **P < 0.01, ***P < 0.001. Error bars represent standard error of the mean (SEM).

We examined the phenotypes of lkb1T1/T2. Consistent with the lkb1T2/T2, the mRNA levels of Lkb1 were significantly reduced in lkb1T1/T2 compared with wt, lkb1T1/+ and lkb1T2/+, and even lower than that in lkb1T2/T2 (Supplementary Fig. 2a).

The sleep phenotype in lkb1T1/T2 flies was also consistent with lkb1T2/T2, with highly reduced nighttime sleep duration (Supplementary Fig. 2, b and c), highly reduced sleep bout duration (Supplementary Fig. 2e) and highly increased latency to sleep (Supplementary Fig. 2f), when compared with wt, lkb1T1/+, and lkb1T2/+ flies.

Results of sleep analysis of lkb1T1/+, lkb1T2/+, lkb1T2/T2, and lkb1T1/T2 mutant flies all consistently support that Lkb1 plays a physiological role in promoting sleep.

Rescue of sleep phenotypes by Lkb1 in flies

We inserted the sequence of the yeast transcription factor Gal4 into the lkb1T2 mutant flies, flanking the lkb1 promoter, and obtained lkb1T2-Gal4 flies (Fig. 2a). We also generated UAS-Lkb1 flies in which the Lkb1 CDS was expressed under the control of the UAS (Brand and Perrimon 1993). Because Gal4 protein binds to the UAS, the expression of Lkb1 in flies resulting from the crosses between lkb1T2-Gal4 flies and UAS-Lkb1 flies would be under the control of the endogenous Lkb1 promoter. Indeed, expression of Lkb1 mRNA was restored when lkb1T2-Gal4 and UAS-Lkb1 were present in the same flies (Fig. 2b), whereas Lkb1 mRNA was less in UAS-Lkb1; lkb1T2/T2 mutant flies, and lkb1T2-Gal4/lkb1T2-Gal4 flies than that in the wt. UAS-Lkb1 alone could not restore Lkb1 mRNA expression level to that in wt flies (Fig. 2b).

Fig. 2.

Rescue of sleep loss in lkb1T2/T2 by Lkb1. a) A diagram of lkb1T2-Gal4: a cDNA for the yeast Gal4 gene inserted in the lkb1T2 knockdown mutants. b) Relative Lkb1 mRNA levels in lkb1T2-Gal4 (blue), UAS-Lkb1; lkb1T2-Gal4 (red), UAS-Lkb1; lkb1T2 (yellow), and wt (black) flies. In lkb1T2-Gal4 homozygous flies, UAS-Lkb1 cDNA driven by Gal4 to rescue sleep phenotypes of lkb1 knockdown mutants. c) Sleep profiles of UAS-Lkb1; lkb1T2-Gal4 (red, n = 45), UAS-Lkb1; lkb1T2 (yellow, n = 47), lkb1T2-Gal4 (blue, n = 46), and wt (black, n = 36) flies. Statistical analysis of sleep duration, sleep bout number, sleep bout duration and latency to sleep in UAS-Lkb1; lkb1T2-Gal4 (red, n = 45), UAS-Lkb1; lkb1T2 (yellow, n = 47), lkb1T2-Gal4 (blue, n = 46), and wt (black, n = 36) flies. Open bars denote daytime, filled bars nighttime. d) Sleep duration. Nighttime sleep duration of UAS-Lkb1; lkb1T2-Gal4 was similar to that of wt mutants, both significantly higher than UAS-Lkb1; lkb1T2 and lkb1T2/T2-Gal4 flies. e) Sleep bout number. Nighttime sleep bout number of UAS-Lkb1; lkb1T2-Gal4 was similar to the wt but significantly higher than UAS-Lkb1; lkb1T2 and lkb1T2-Gal4 flies. f) Sleep bout duration. Nighttime sleep bout duration of UAS-Lkb1; lkb1T2-Gal4 was similar to the wt but significantly higher than UAS-Lkb1; lkb1T2 and lkb1T2-Gal4 flies. g) Latency to sleep. Latency after light-off of UAS-Lkb1; lkb1T2-Gal4 was similar to the wt but significantly shorter than UAS-Lkb1; lkb1T2 and lkb1T2-Gal4 flies. One-way ANOVA was used. n.s. denotes P > 0.05, * P < 0.05, ** P < 0.01, ***P < 0.001. Error bars represent SEM.

Both daytime and nighttime sleep durations were less in lkb1T2-Gal4/lkb1T2-Gal4 flies than those in wt flies (Fig. 2c). Introduction of UAS-Lkb1 in lkb1T2/T2 flies or lkb1T2-Gal4 alone could not restore sleep. When both lkb1T2-Gal4 and UAS-Lkb1 were present, nighttime sleep durations were restored (Fig. 2d). Nighttime sleep bout number, nighttime sleep bout duration, and nighttime latency were restored when both lkb1T2-Gal4 and UAS-Lkb1 were present, but not when lkb1T2-Gal4 or UAS-Lkb1 alone was present (Fig. 2, e–g).

These results support that the sleep phenotypes of lkb1T2/T2 were attributable to the reduction of Lkb1 mRNA expression in these flies.

Sleep phenotypes of flies carrying neuronal deletion of the Lkb1 gene

To determine whether Lkb1 functions in neurons, we used the CRISPR-Cas9 system to delete Lkb1 from neurons specifically (Supplementary Fig. 4). A pan-neuronal Gal4 driver (57C10-Gal4) was used to control the expression of small guide RNA (sgRNA) targetting Lkb1 in neurons. Compared with 57C10-Gal4>UAS-Cas9 alone or 57C10-Gal4>UAS-Lkb1-sgRNA alone, when both UAS-Cas9 and UAS-Lkb1-sgRNA were present in flies, nighttime sleep duration (Fig. 3b) and nighttime sleep bout duration (Fig. 3d) were significantly reduced and nighttime sleep latency significantly lengthened (Fig. 3e). Daytime sleep duration, bout number, bout duration and latency were not significantly affected by neuronal gene targeting of Lkb1 (Fig. 3, b–e).

Fig. 3.

Sleep phenotypes of mutants from whose neurons Lkb1 was targeted. a) Sleep profiles of UAS-Lkb1-sgRNA/57C10-Gal4;+/UAS-Cas9 (red, n = 44), UAS-Lkb1-sgRNA/57C10-Gal4 (blue, n = 41), and 57C10-Gal4/+;+/UAS-Cas9 (black, n = 45) flies. Statistical analysis of sleep duration, sleep bout number, sleep bout duration, and latency to sleep in UAS-Lkb1-sgRNA/57C10-Gal4;+/UAS-Cas9 (red, n = 44), UAS-Lkb1-sgRNA/57C10-Gal4 (blue, n = 41), and 57C10-Gal4/+;+/UAS-Cas9 (black, n = 45) flies. Open bars denote daytime, filled bars nighttime. b) Sleep duration. Daytime and nighttime sleep duration of UAS-Lkb1-sgRNA/57C10-Gal4;+/UAS-Cas9 was significantly less than those of UAS-Lkb1-sgRNA/57C10-Gal4 and 57C10-Gal4/+;+/UAS-Cas9 flies. c) Sleep bout number. Daytime and nighttime sleep bout numbers of UAS-Lkb1-sgRNA/57C10-Gal4;+/UAS-Cas9 was not significantly from those of UAS-Lkb1-sgRNA/57C10-Gal4 and 57C10-Gal4/+;+/UAS-Cas9 flies. d) Sleep bout duration. Nighttime sleep bout duration of UAS-Lkb1-sgRNA/57C10-Gal4;+/UAS-Cas9 was significantly less than that of UAS-Lkb1-sgRNA/57C10-Gal4 and 57C10-Gal4/+;+/UAS-Cas9 flies. e) Latency to sleep. Latency to sleep after light-off of UAS-Lkb1-sgRNA/57C10-Gal4;+/UAS-Cas9 was longer than that of 57C10-Gal4/+;+/UAS-Cas9 which was not significantly different from that of UAS-Lkb1-sgRNA/57C10-Gal4 flies. One-way ANOVA was used. n.s. denotes P > 0.05, *P < 0.05, **P < 0.01, ***P < 0.001. Error bars represent SEM.

We also investigated any potential effect that overexpression of Lkb1 in neurons might cause (Supplementary Fig. 5a). We detected no phenotype resulting from neuronal overexpression of Lkb1 on daytime and nighttime sleep duration, sleep bout number, sleep bout duration, or latency (Supplementary Fig. 5, b–f).

In all 3 series of experiments (Figs. 1–3), nighttime sleep phenotypes were more obvious than daytime sleep phenotypes. These results strongly indicate that Lkb1 expression in neurons are required physiologically for sleep, especially nighttime sleep.

Genetic interactions between Lkb1 and Ampk or Sik3 in flies

To examine potential genetic interactions of Lkb1 with either Ampk or Sik3, we combined the loss-of-function (LOF) Lkb1 mutation lkb1T2 with specific point mutations in either Ampk or Sik3.

The regulatory site T184 in Drosophila AMPK and T196 in Drosophila SIK3 were equivalent to T172 of mammalian AMPK2 and T221 of mammalian SIK3, important for their activities. When the endogenous T184 in fly AMPK was mutated to alanine (A) or glutamic acid (E), flies were lethal. We therefore introduced T184A and T184E mutations into an Ampk transgene whose expression was controlled by UAS. We introduced UAS-Ampk, UAS-Ampk-T184A, and UAS-Ampk-T184E into lkb1T2/T2 flies and used a pan-neuronal driver to express them in neurons (Fig. 4). Neuronal overexpression of Ampk-T184E (Fig. 4a) and UAS-Ampk (Fig. 4c) in lkb1T2 flies did not significantly change the sleep phenotypes of lkb1T2/T2 flies, but neuronal overexpression of UAS-Ampk-T184A (Fig. 4b) in lkb1T2/T2 flies further decreased nighttime sleep duration.

Fig. 4.

Genetic interactions between lkb1 and ampk. a) Sleep profiles of UAS-Ampk-T184E/57C10; lkb1T2 (red, n = 19), 57C10; lkb1T2 (yellow, n = 24), UAS-Ampk-T184E; lkb1T2 (blue, n = 22) and wt (black, n = 24) flies. b) Sleep profiles of UAS-Ampk-T184A/57C10; lkb1T2 (red, n = 24), 57C10; lkb1T2 (yellow, n = 24), UAS-Ampk-T184A; lkb1T2 (blue, n = 23) and wt (black, n = 24) flies. c) Sleep profiles of UAS-Ampk/57C10; lkb1T2 (red, n = 24), 57C10; lkb1T2 (yellow, n = 24), UAS-Ampk; lkb1T2 (blue, n = 11) and wt (black, n = 24) flies. (d–g) Statistical analysis of sleep duration, sleep bout number, sleep bout duration, and latency to sleep in wt (black, n = 24), 57C10; lkb1T2 (yellow, n = 24), UAS-Ampk-T184E; lkb1T2 (blue, n = 22), UAS-Ampk-T184E/57C10; lkb1T2 (red, n = 19), UAS-Ampk-T184A; lkb1T2 (blue, n = 23), UAS-Ampk-T184A/57C10; lkb1T2 (red, n = 24), UAS-Ampk; lkb1T2 (blue, n = 11), and UAS-Ampk/57C10; lkb1T2 (red, n = 24) flies. Open bars denote daytime, filled bars nighttime; n.s. not shown. One-way ANOVA was used. n.s. denotes P > 0.05, *P < 0.05, **P < 0.01, ***P < 0.001. Error bars represent SEM.

Point mutations of Sik3-flag, Sik3-T196A-flag, and Sik3-T196E-flag were constructed in Drosophila. When the endogenous T196 in Sik3 was mutated to A or E, we could get homozygous flies. Upon crossing to lkb1T2/T2, Sik3-T196A-flag; lkb1T2/T2 were homozygous lethal. The cross of Sik3-T196E-flag into the lkb1T2/T2 background generated viable flies, with no detectable change in sleep (Fig. 5).

Fig. 5.

Genetic interactions between lkb1 and sik3. a) Sleep profiles of lkb1T2 (red, n = 45), Sik3-T196E-flag; lkb1T2 (dark red, n = 46), Sik3-flag; lkb1T2 (orange, n = 48), Sik3-T196E-flag (blue, n = 45), Sik3-flag (dark blue, n = 44) and wt (black, n = 46) flies. (b–e) Statistical analysis of sleep duration, sleep bout number, sleep bout duration, and latency to sleep in wt (black, n = 46), Sik3-flag (dark blue, n = 44), Sik3-T196E-flag (blue, n = 45), Sik3-flag; lkb1T2 (orange, n = 48), Sik3-T196E-flag; lkb1T2 (dark red, n = 46), and lkb1T2 (red, n = 45) flies. Open bars denote daytime, filled bars nighttime. Statistics for groups Sik3-flag; lkb1T2 and Sik3-T196E-flag; lkb1T2 were presented. One-way ANOVA was used. n.s. denotes P > 0.05, *P < 0.05, **P < 0.01, ***P < 0.001. Error bars represent SEM.

Allele-specific genetic interactions between Lkb1 and Ampk in sleep, or between Lkb1 and Sik3 in viability, suggest, but do not prove, regulatory relationships between Lkb1 and Ampk in sleep or Sik3 in viability.

Circadian rhythm in Lkb1 mutant flies

The transcription factor differentiated embryo-chondrocyte 1 regulates circadian rhythm and can negatively regulate the transcription of Lkb1 and subsequently reduce AMPK activity (Sato et al. 2015).

We tested whether the circadian rhythm was affected in Lkb1 mutant flies. Lkb1 mutant flies were not different from wt flies in period length (Supplementary Fig. 7b). Relative rhythmic power was increased in lkb1T2/+ and lkb1T2/T2 mutants than wt flies (Supplementary Fig. 7).

Sleep phenotypes in Lkb1 conditional knockout mice

To investigate potential involvement of Lkb1 in regulating sleep of mammalian animals, we obtained Lkb1fl/fl mice in which the loxP sites flanked exons 3–6 of the Lkb1 gene (Nakada et al. 2010). To delete the Lkb1 gene from these mice, we injected AAV constructs expressing either the Cre recombinase together with the GFP or GFP alone to infect the mouse brain. Cre-GFP or GFP was under the control of a neuronal specific promoter human Synapsin I (hSyn) (in AAV-PHP.eB-hSyn-Cre-GFP or AAV-PHP.eB-hSyn-GFP).

We analyzed the expression of LKB1 protein in mice (Fig. 6, a and b). Injection of Cre-GFP expressing virus into wt or Lkb1fl/+ mice did not reduce LKB1 protein expression in the brain. Neither did injection of only GFP expressing virus into Lkb1fl/fl mice. This conclusion was reached by examination of either several mouse brains combined (Fig. 6a), or individual mouse brains (Fig. 6b).

Fig. 6.

Sleep phenotypes of lkb1 conditional knockout mice. a) Levels of LKB1 protein from Lkb1fl/fl mice injected with AAV-hSyn-Cre-GFP virus (Cre+ Lkb1fl/fl, n = 4), Lkb1fl/fl mice injected with AAV-hSyn-GFP virus (GFP+ Lkb1fl/fl, n = 3), Lkb1fl/+ mice injected with AAV-hSyn-Cre-GFP virus (Cre+ Lkb1fl/+, n = 2) and Lkb1+/+ mice injected with AAV-hSyn-Cre-GFP virus (Cre+ Lkb1+/+, n = 2). These mice were among those used for EEG recording and analysis. b) Levels of LKB1 protein in individual mice (genotypes labeled: Cre+ Lkb1fl/fl, GFP+ Lkb1fl/fl, Cre+ Lkb1fl/+, and Cre+ Lkb1+/+). These mice were the same mice as those in (a) but presented individually. Statistical analysis of wake duration, NREM duration and REM duration in Cre+ Lkb1fl/fl (red, n = 10), GFP+ Lkb1fl/fl (yellow, n = 5), Cre+ Lkb1fl/+ (blue, n = 7), and Cre+ Lkb1+/+ (black, n = 4) mice in a 12:12 LD cycle. White background denotes daytime, gray background nighttime. c) Wake duration. Daytime wake duration of Cre+ Lkb1fl/fl mice was higher than those of GFP+ Lkb1fl/fl, Cre+ Lkb1fl/+, or Cre+ Lkb1+/+ mice. Night-time wake duration of Cre+ Lkb1fl/fl mice was higher than that of GFP+ Lkb1fl/fl mice. e) NREM duration. Daytime NREM duration of Cre+ Lkb1fl/fl mice was lower than those of GFP+ Lkb1fl/fl, Cre+ Lkb1fl/+, and Cre+ Lkb1+/+ mice. Night-time NREM duration of Cre+ Lkb1fl/fl mice was lower than that of GFP+ Lkb1fl/fl mice. g) REM duration. Daytime and nighttime REM durations of Cre+ Lkb1fl/fl mice was not significantly different from those of GFP+ Lkb1fl/fl, Cre+ Lkb1fl/+, and Cre+ Lkb1+/+ mice. EEG power spectrum of (d) WAKE, (f) NREM, and (h) REM states in Cre+ Lkb1fl/fl (red, n = 8), GFP+ Lkb1fl/fl (yellow, n = 5), Cre+ Lkb1fl/+ (blue, n = 6), and Cre+ Lkb1+/+ (black, n = 4) mice. (i) NREM δ-power density of Cre+ Lkb1fl/fl (red, n = 8), GFP+ Lkb1fl/fl (yellow, n = 5), Cre+ Lkb1fl/+ (blue, n = 6), and Cre+ Lkb1+/+ (black, n = 4) mice. One-way ANOVA was used in (c, e, g) for comparison of Cre+ Lkb1fl/fl, Cre+ Lkb1fl/+, and Cre+ Lkb1+/+ mice. Unpaired t test was used in (c, e, g) for comparison of Cre+ Lkb1fl/fl, and GFP+ Lkb1fl/fl mice. Two-way ANOVA followed by Turkey’s multiple comparisons test was used in (d, f, h, i). n.s. denotes P > 0.05, *P < 0.05, **P < 0.01, ***P < 0.001. Error bars represent SEM.

Functionally, only when the Cre-GFP expressing virus was injected into Lkb1fl/fl mice, wake duration was significantly increased during daytime (Fig. 6c;Supplementary Fig. 6a), NREM sleep duration was significantly decreased during daytime (Fig. 6e;Supplementary Fig. 6b). Controls (Cre-GFP injection into wt or Lkb1fl/+ mice, GFP injection into Lkb1fl/fl mice) did not significantly changed any sleep phenotypes.

REM sleep duration was not significantly affected by Cre-GFP injection into Lkb1fl/fl mice (Fig. 6g;Supplementary Fig. 6c).

Power density in the 1–4 Hz range (δ power density) of NREM is a commonly accepted indicator of sleep need (Borbely et al. 1981; Borbely 1982; Daan et al. 1984; Tobler and Borbely 1986; Dijk et al. 1987; Werth et al. 1996; Franken et al. 2001). We found that NREM δ power density was significantly reduced when the Cre-GFP expressing virus was injected into Lkb1fl/fl mice (Fig. 6f). Analysis over 24 h indicated that significant reduction was observed over most of the daily cycle (Fig. 6i).

Discussion

Our results indicate that LKB1 is required for sleep regulation: it plays an important and conserved role by promoting sleep in both flies and mice. LKB1 plays this role in neurons in both species because gene targeting of Lkb1 in neurons led to reduction of sleep. In mice, with the additional advantage of EEG analysis, we find that LKB1 regulates sleep need as indicated by reduced NREM δ power density in Lkb1 knockdown mutant mice.

Sleep is important for animals. Sleep regulation is accomplished through 2 processes: circadian and sleep homeostatic (Borbely 1982; Borbely et al. 2016). The circadian clock regulates the timing of sleep, and homeostatic process regulates the sleep drive. Molecular mechanisms of the circadian clock have been revealed through research in Drosophila and other organisms (Hendricks et al. 2000; Shaw et al. 2000; Nitabach and Taghert 2008; Allada and Chung 2010; Mohawk et al. 2012). Although many sleep-related genes have been identified in sleep regulation (Cirelli 2009; Allada et al. 2017; Jan et al. 2020), our understanding of the mechanisms underlying sleep homeostatic regulation remains limited (Allada et al. 2017; Donlea et al. 2017).

Multiple regions in Drosophila and mouse brains have been implicated in sleep regulation. In flies, sleep is regulated by several regions including: the ILNv and DN1 clock neurons which are important for circadian control of sleep (Parisky et al. 2008; Shang et al. 2008; Sheeba et al. 2008; Chung et al. 2009; Shang et al. 2013; Kunst et al. 2014). And the mushroom bodies, the dorsal of fan-shaped body, the ellipsoid body, the pars intercerebralis, and glia (Joiner et al. 2006; Foltenyi et al. 2007; Crocker et al. 2010; Donlea et al. 2011; Guo et al. 2011; Seugnet et al. 2011; Liu et al. 2012; Ueno et al. 2012; Yi et al. 2013; Donlea et al. 2014; Park et al. 2014; Chen et al. 2015; Liu et al. 2016; Pimentel et al. 2016). In mammals, sleep is regulated by monoaminergic, cholinergic, glutamatergic, and GABAergic neurons that are distributed in multiple regions including the brain stem, the preoptic hypothalamus, the lateral hypothalamus, and the basal forebrain (Weber and Dan 2016; Saper and Fuller 2017; Scammell et al. 2017). It will be interesting to investigate whether Lkb1 functions in all or a limited subset of neurons to regulate sleep.

Lkb1 as a master kinase can regulate the activities of ARKs by phosphorylating the site in the active T loop equivalent to AMPK-T172 (Lizcano et al. 2004). Because both AMPK and SIK3 are involved in sleep regulation, it will be interesting to investigate downstream kinases mediating the function of Lkb1 in sleep regulation. Is it SIK3, AMPK, or other members of the ARKs? Our findings of allele-specific genetic interactions between Lkb1 and Ampk suggest that they could be upstream and downstream of each other in regulating sleep. Because of the lethality of double mutation combination of Sik3 and lkb1, we cannot rule out that Sik3 may also be downstream of Lkb1 in regulating Drosophila sleep. The Ca2+/calmodulin-dependent protein kinase kinase-2 (CaMKK2, also known as CaMKKβ) could phosphorylate AMPKα-T172 (Hawley et al. 2005; Hurley et al. 2005; Woods et al. 2005; Anderson et al. 2008), but CaMKK2 could not phosphorylate the equivalent sites in the other ARKs, including SIK3 (Fogarty et al. 2010). It will be interesting to investigate whether and how CaMKK2 regulates sleep.

In Drosophila, LKB1 functions through SIK3 which phosphorylates histone deacetylase 4 (HDAC4) to regulate lipid storage (Choi et al. 2015). It will be interesting to investigate whether HDAC4 is downstream of LKB1 in sleep regulation.

More importantly, an important question for further studies is whether Lkb1 regulation of sleep is related to its regulation of metabolism. Changes in energy homeostasis directly and reversibly influence the sleep/wake cycle (Collet et al. 2016). Some molecules involved in metabolism regulate sleep (Taheri et al. 2004; Bjorness and Greene 2009; Thimgan et al. 2010; Gerstner et al. 2011; Nixon et al. 2015; Grubbs et al. 2020). In Drosophila, starvation suppresses sleep without building up sleep drive (Thimgan et al. 2010). Lkb1 and its downstream components are involved in regulating metabolism, with examples such as LKB1-AMPK signaling in the liver regulating glucose homeostasis (Shaw et al. 2005), SIK3-HDAC4 regulating energy balance in Drosophila (Wang et al. 2011). Either LKB1 has 2 independent roles in sleep and metabolism or that its 2 roles are related.

Our recent in vitro biochemical discoveries of STE20 phosphorylation of AMPK and SIK3 (and other ARKs) raise more questions about physiological significance of any STE20 or any other ARK in sleep (Liu, Wang, Cui, Gao, et al. 2022; Liu, Wang, Cui, et al. 2022).

Data availability

Strains and plasmids are available upon request. All data necessary for confirming the conclusions of the article are present within the article and its supplementary data.

Supplemental material is available at GENETICS online.

Acknowledgments

We are grateful to Dr Bowen Deng for Ampk flies, to Ping-ping Yan, Lan Wang, and Yong-hui Zhang for technical assistance, to Drs Wei Yang and En-xin Zhou for Drosophila video tracing programs, to Dan Wang for mouse rearing, to members of the Rao lab for discussion, to the Bloomington Drosophila Stock Center for flies, to the Jackson Laboratory for mice, to CIBR, Peking-Tsinghua Center for Life Sciences, IDG/McGovern Institute for Brain Research at Peking University and Changping Laboratory for support.

Conflicts of interest

The authors declare no conflict of interests.

Literature cited

Alessi
DR
,
Sakamoto
K
,
Bayascas
JR.
LKB1-dependent signaling pathways
.
Annu Rev Biochem
.
2006
;
75
:
137
163
.

Allada
R
,
Chung
BY.
Circadian organization of behavior and physiology in Drosophila
.
Annu Rev Physiol
.
2010
;
72
:
605
624
.

Allada
R
,
Cirelli
C
,
Sehgal
A.
Molecular mechanisms of sleep homeostasis in flies and mammals
.
Cold Spring Harb Perspect Biol
.
2017
;
9
(
8
):
a027730
.

Anderson
KA
,
Ribar
TJ
,
Lin
F
,
Noeldner
PK
,
Green
MF
,
Muehlbauer
MJ
,
Witters
LA
,
Kemp
BE
,
Means
AR.
Hypothalamic CaMKK2 contributes to the regulation of energy balance
.
Cell Metab
.
2008
;
7
(
5
):
377
388
.

Bardeesy
N
,
Sinha
M
,
Hezel
AF
,
Signoretti
S
,
Hathaway
NA
,
Sharpless
NE
,
Loda
M
,
Carrasco
DR
,
DePinho
RA.
Loss of the Lkb1 tumour suppressor provokes intestinal polyposis but resistance to transformation
.
Nature
.
2002
;
419
(
6903
):
162
167
.

Barger
Z
,
Frye
CG
,
Liu
D
,
Dan
Y
,
Bouchard
KE.
Robust, automated sleep scoring by a compact neural network with distributional shift correction
.
PLoS One
.
2019
;
14
(
12
):
e0224642
.

Beg
ZH
,
Allmann
DW
,
Gibson
DM.
Modulation of 3-hydroxy-3-methylglutaryl coenzyme A reductase activity with cAMP and with protein fractions of rat liver cytosol
.
Biochem Biophys Res Commun
.
1973
;
54
(
4
):
1362
1369
.

Bjorness
TE
,
Greene
RW.
Adenosine and sleep
.
Curr Neuropharmacol
.
2009
;
7
(
3
):
238
245
.

Borbely
AA.
A two process model of sleep regulation
.
Hum Neurobiol
.
1982
;
1
(
3
):
195
204
.

Borbely
AA
,
Baumann
F
,
Brandeis
D
,
Strauch
I
,
Lehmann
D.
Sleep deprivation: effect on sleep stages and EEG power density in man
.
Electroencephalogr Clin Neurophysiol
.
1981
;
51
(
5
):
483
495
.

Borbely
AA
,
Daan
S
,
Wirz-Justice
A
,
Deboer
T.
The two-process model of sleep regulation: a reappraisal
.
J Sleep Res
.
2016
;
25
(
2
):
131
143
.

Brand
AH
,
Perrimon
N.
Targeted gene expression as a means of altering cell fates and generating dominant phenotypes
.
Development
.
1993
;
118
(
2
):
401
415
.

Carling
D
,
Clarke
PR
,
Zammit
VA
,
Hardie
DG.
Purification and characterization of the AMP-activated protein-kinase. Copurification of acetyl-CoA carboxylase kinase and 3-hydroxy-3-methylglutaryl-CoA reductase kinase-activities
.
Eur J Biochem
.
1989
;
186
(
1–2
):
129
136
.

Carling
D
,
Zammit
VA
,
Hardie
DG.
A common bicyclic protein-kinase cascade inactivates the regulatory enzymes of fatty-acid and cholesterol-biosynthesis
.
FEBS Lett
.
1987
;
223
(
2
):
217
222
.

Carlson
CA
,
Kim
KH.
Regulation of hepatic acetyl coenzyme A carboxylase by phosphorylation and dephosphorylation
.
J Biol Chem
.
1973
;
248
(
1
):
378
380
.

Carretero
J
,
Medina
PP
,
Pio
R
,
Montuenga
LM
,
Sanchez-Cespedes
M.
Novel and natural knockout lung cancer cell lines for the LKB1/STK11 tumor suppressor gene
.
Oncogene
.
2004
;
23
(
22
):
4037
4040
.

Chan
KY
,
Jang
MJ
,
Yoo
BB
,
Greenbaum
A
,
Ravi
N
,
Wu
WL
,
Sanchez-Guardado
L
,
Lois
C
,
Mazmanian
SK
,
Deverman
BE
, et al. 
Engineered AAVs for efficient noninvasive gene delivery to the central and peripheral nervous systems
.
Nat Neurosci
.
2017
;
20
(
8
):
1172
1179
.

Chen
WF
,
Maguire
S
,
Sowcik
M
,
Luo
W
,
Koh
K
,
Sehgal
A.
A neuron-glia interaction involving GABA transaminase contributes to sleep loss in sleepless mutants
.
Mol Psychiatry
.
2015
;
20
(
2
):
240
251
.

Chikahisa
S
,
Fujiki
N
,
Kitaoka
K
,
Shimizu
N
,
Sei
H.
Central AMPK contributes to sleep homeostasis in mice
.
Neuropharmacology
.
2009
;
57
(
4
):
369
374
.

Choi
S
,
Lim
DS
,
Chung
J.
Feeding and fasting signals converge on the LKB1-SIK3 pathway to regulate lipid metabolism in Drosophila
.
PLoS Genet
.
2015
;
11
(
5
):
e1005263
.

Chu
VT
,
Graf
R
,
Wirtz
T
,
Weber
T
,
Favret
J
,
Li
X
,
Petsch
K
,
Tran
NT
,
Sieweke
MH
,
Berek
C
, et al. 
Efficient CRISPR-mediated mutagenesis in primary immune cells using CrispRGold and a C57BL/6 Cas9 transgenic mouse line
.
Proc Natl Acad Sci U S A
.
2016
;
113
(
44
):
12514
12519
.

Chung
BY
,
Kilman
VL
,
Keath
JR
,
Pitman
JL
,
Allada
R.
The GABA(a) receptor EDL acts in peptidergic PDF neurons to promote sleep in Drosophila
.
Curr Biol
.
2009
;
19
(
5
):
386
390
.

Cirelli
C.
The genetic and molecular regulation of sleep: from fruit flies to humans
.
Nat Rev Neurosci
.
2009
;
10
(
8
):
549
560
.

Collet
TH
,
van der Klaauw
AA
,
Henning
E
,
Keogh
JM
,
Suddaby
D
,
Dachi
SV
,
Dunbar
S
,
Kelway
S
,
Dickson
SL
,
Farooqi
IS
, et al. 
The sleep/wake cycle is directly modulated by changes in energy balance
.
Sleep
.
2016
;
39
(
9
):
1691
1700
.

Crocker
A
,
Shahidullah
M
,
Levitan
IB
,
Sehgal
A.
Identification of a neural circuit that underlies the effects of octopamine on sleep: wake behavior
.
Neuron
.
2010
;
65
(
5
):
670
681
.

Daan
S
,
Beersma
DG
,
Borbely
AA.
Timing of human sleep: recovery process gated by a circadian pacemaker
.
Am J Physiol
.
1984
;
246
(
2 Pt 2
):
R161
R183
.

Dai
X
,
Zhou
E
,
Yang
W
,
Mao
R
,
Zhang
W
,
Rao
Y.
Molecular resolution of a behavioral paradox: sleep and arousal are regulated by distinct acetylcholine receptors in different neuronal types in Drosophila
.
Sleep
.
2021
;
44
(
7
):
0161
8105
.

Dai
X
,
Zhou
E
,
Yang
W
,
Zhang
X
,
Zhang
W
,
Rao
Y.
D-serine made by serine racemase in Drosophila intestine plays a physiological role in sleep
.
Nat Commun
.
2019
;
10
(
1
):
2041
1723
.

Davies
SP
,
Hawley
SA
,
Woods
A
,
Carling
D
,
Haystead
TAJ
,
Hardie
DG.
Purification of the AMP-activated protein-kinase on ATP-gamma-sepharose and analysis of its subunit structure
.
Eur J Biochem
.
1994
;
223
(
2
):
351
357
.

Deng
B
,
Li
Q
,
Liu
X
,
Cao
Y
,
Li
B
,
Qian
Y
,
Xu
R
,
Mao
R
,
Zhou
E
,
Zhang
W
, et al. 
Chemoconnectomics: mapping chemical transmission in Drosophila
.
Neuron
.
2019
;
101
(
5
):
876
893.e4
.

Dijk
DJ
,
Beersma
DG
,
Daan
S.
EEG power density during nap sleep: reflection of an hourglass measuring the duration of prior wakefulness
.
J Biol Rhythms
.
1987
;
2
(
3
):
207
219
.

Donlea
JM
,
Alam
MN
,
Szymusiak
R.
Neuronal substrates of sleep homeostasis; lessons from flies, rats and mice
.
Curr Opin Neurobiol
.
2017
;
44
:
228
235
.

Donlea
JM
,
Pimentel
D
,
Miesenbock
G.
Neuronal machinery of sleep homeostasis in Drosophila
.
Neuron
.
2014
;
81
(
4
):
860
872
.

Donlea
JM
,
Thimgan
MS
,
Suzuki
Y
,
Gottschalk
L
,
Shaw
PJ.
Inducing sleep by remote control facilitates memory consolidation in Drosophila
.
Science
.
2011
;
332
(
6037
):
1571
1576
.

Ferrer
A
,
Caelles
C
,
Massot
N
,
Hegardt
FG.
Activation of rat-liver cytosolic 3-hydroxy-3-methylglutaryl coenzyme A reductase kinase by adenosine 5'-monophosphate
.
Biochem Bioph Res Co
.
1985
;
132
(
2
):
497
504
.

Fogarty
S
,
Hawley
SA
,
Green
KA
,
Saner
N
,
Mustard
KJ
,
Hardie
DG.
Calmodulin-dependent protein kinase kinase-beta activates AMPK without forming a stable complex: synergistic effects of Ca2+ and AMP
.
Biochem J
.
2010
;
426
(
1
):
109
118
.

Foltenyi
K
,
Greenspan
RJ
,
Newport
JW.
Activation of EGFR and ERK by rhomboid signaling regulates the consolidation and maintenance of sleep in Drosophila
.
Nat Neurosci
.
2007
;
10
(
9
):
1160
1167
.

Franken
P
,
Chollet
D
,
Tafti
M.
The homeostatic regulation of sleep need is under genetic control
.
J Neurosci
.
2001
;
21
(
8
):
2610
2621
.

Funato
H
,
Miyoshi
C
,
Fujiyama
T
,
Kanda
T
,
Sato
M
,
Wang
Z
,
Ma
J
,
Nakane
S
,
Tomita
J
,
Ikkyu
A
, et al. 
Forward-genetics analysis of sleep in randomly mutagenized mice
.
Nature
.
2016
;
539
(
7629
):
378
383
.

Gerstner
JR
,
Vanderheyden
WM
,
Shaw
PJ
,
Landry
CF
,
Yin
JCP.
Fatty-acid binding proteins modulate sleep and enhance long-term memory consolidation in Drosophila
.
PLoS One
.
2011
;
6
(
1
):
e15890
.

Gill
RK
,
Yang
SH
,
Meerzaman
D
,
Mechanic
LE
,
Bowman
ED
,
Jeon
HS
,
Chowdhuri
SR
,
Shakoori
A
,
Dracheva
T
,
Hong
KM
, et al. 
Frequent homozygous deletion of the LKB1/STK11 gene in non-small cell lung cancer
.
Oncogene
.
2011
;
30
(
35
):
3784
3791
.

Grubbs
JJ
,
Lopes
LE
,
van der Linden
AM
,
Raizen
DM.
A salt-induced kinase is required for the metabolic regulation of sleep
.
PLoS Biol
.
2020
;
18
(
4
):
e3000220
.

Guldberg
P
,
thor Straten
P
,
Ahrenkiel
V
,
Seremet
T
,
Kirkin
AF
,
Zeuthen
J.
Somatic mutation of the Peutz-Jeghers syndrome gene, LKB1/STK11, in malignant melanoma
.
Oncogene
.
1999
;
18
(
9
):
1777
1780
.

Guo
F
,
Yi
W
,
Zhou
MM
,
Guo
AK.
Go signaling in mushroom bodies regulates sleep in Drosophila
.
Sleep
.
2011
;
34
(
3
):
273
281
.

Han
C
,
Jan
LY
,
Jan
YN.
Enhancer-driven membrane markers for analysis of nonautonomous mechanisms reveal neuron-glia interactions in Drosophila
.
Proc Natl Acad Sci U S A
.
2011
;
108
(
23
):
9673
9678
.

Hardie
DG.
AMP-activated protein kinase: maintaining energy homeostasis at the cellular and whole-body levels
.
Annu Rev Nutr
.
2014
;
34
:
31
55
.

Hardie
DG
,
Schaffer
BE
,
Brunet
A.
AMPK: an energy-sensing pathway with multiple inputs and outputs
.
Trends Cell Biol
.
2016
;
26
(
3
):
190
201
.

Hawley
SA
,
Boudeau
J
,
Reid
JL
,
Mustard
KJ
,
Udd
L
,
Makela
TP
,
Alessi
DR
,
Hardie
DG.
Complexes between the LKB1 tumor suppressor, STRAD alpha/beta and MO25 alpha/beta are upstream kinases in the AMP-activated protein kinase cascade
.
J Biol
.
2003
;
2
(
4
):
28
.

Hawley
SA
,
Davison
M
,
Woods
A
,
Davies
SP
,
Beri
RK
,
Carling
D
,
Hardie
DG.
Characterization of the AMP-activated protein kinase kinase from rat liver and identification of threonine 172 as the major site at which it phosphorylates AMP-activated protein kinase
.
J Biol Chem
.
1996
;
271
(
44
):
27879
27887
.

Hawley
SA
,
Pan
DA
,
Mustard
KJ
,
Ross
L
,
Bain
J
,
Edelman
AM
,
Frenguelli
BG
,
Hardie
DG.
Calmodulin-dependent protein kinase kinase-beta is an alternative upstream kinase for AMP-activated protein kinase
.
Cell Metab
.
2005
;
2
(
1
):
9
19
.

Hearle
N
,
Schumacher
V
,
Menko
FH
,
Olschwang
S
,
Boardman
LA
,
Gille
JJ
,
Keller
JJ
,
Westerman
AM
,
Scott
RJ
,
Lim
W
, et al. 
Frequency and spectrum of cancers in the Peutz-Jeghers syndrome
.
Clin Cancer Res
.
2006
;
12
(
10
):
3209
3215
.

Hemminki
A
,
Markie
D
,
Tomlinson
I
,
Avizienyte
E
,
Roth
S
,
Loukola
A
,
Bignell
G
,
Warren
W
,
Aminoff
M
,
Hoglund
P
, et al. 
A serine/threonine kinase gene defective in Peutz-Jeghers syndrome
.
Nature
.
1998
;
391
(
6663
):
184
187
.

Hemminki
A
,
Tomlinson
I
,
Markie
D
,
Jarvinen
H
,
Sistonen
P
,
Bjorkqvist
AM
,
Knuutila
S
,
Salovaara
R
,
Bodmer
W
,
Shibata
D
, et al
Localization of a susceptibility locus for Peutz-Jeghers syndrome to 19p using comparative genomic hybridization and targeted linkage analysis
.
Nat Genet
.
1997
;
15
(
1
):
87
90
.

Hendricks
JC
,
Finn
SM
,
Panckeri
KA
,
Chavkin
J
,
Williams
JA
,
Sehgal
A
,
Pack
AI.
Rest in Drosophila is a sleep-like state
.
Neuron
.
2000
;
25
(
1
):
129
138
.

Herzig
S
,
Shaw
RJ.
AMPK: guardian of metabolism and mitochondrial homeostasis
.
Nat Rev Mol Cell Biol
.
2018
;
19
(
2
):
121
135
.

Honda
T
,
Fujiyama
T
,
Miyoshi
C
,
Ikkyu
A
,
Hotta-Hirashima
N
,
Kanno
S
,
Mizuno
S
,
Sugiyama
F
,
Takahashi
S
,
Funato
H
, et al. 
A single phosphorylation site of SIK3 regulates daily sleep amounts and sleep need in mice
.
Proc Natl Acad Sci U S A
.
2018
;
115
(
41
):
10458
10463
.

Hong
SP
,
Leiper
FC
,
Woods
A
,
Carling
D
,
Carlson
M.
Activation of yeast Snf1 and mammalian AMP-activated protein kinase by upstream kinases
.
Proc Natl Acad Sci U S A
.
2003
;
100
(
15
):
8839
8843
.

Hurley
RL
,
Anderson
KA
,
Franzone
JM
,
Kemp
BE
,
Means
AR
,
Witters
LA.
The Ca2+/calmodulin-dependent protein kinase kinases are AMP-activated protein kinase kinases
.
J Biol Chem
.
2005
;
280
(
32
):
29060
29066
.

Ingebritsen
TS
,
Lee
HS
,
Parker
RA
,
Gibson
DM.
Reversible modulation of the activities of both liver microsomal hydroxymethylglutaryl coenzyme A reductase and its inactivating enzyme. Evidence for regulation by phosphorylation-dephosphorylation
.
Biochem Biophys Res Commun
.
1978
;
81
(
4
):
1268
1277
.

Jan
M
,
O'Hara
BF
,
Franken
P.
Recent advances in understanding the genetics of sleep
.
F1000Research
.
2020
;
9
:
214
.

Jeghers
H
,
Mc
KV
,
Katz
KH.
Generalized intestinal polyposis and melanin spots of the oral mucosa, lips and digits; a syndrome of diagnostic significance
.
N Engl J Med
.
1949
;
241
(
26
):
1031
1036
.

Jenne
DE
,
Reimann
H
,
Nezu
J
,
Friedel
W
,
Loff
S
,
Jeschke
R
,
Muller
D
,
Back
W
,
Zimmer
M.
Peutz-Jeghers syndrome is caused by mutations in a novel serine threonine kinase
.
Nat Genet
.
1998
;
18
(
1
):
38
44
.

Ji
HB
,
Ramsey
MR
,
Hayes
DN
,
Fan
C
,
McNamara
K
,
Kozlowski
P
,
Torrice
C
,
Wu
MC
,
Shimamura
T
,
Perera
SA
, et al. 
LKB1 modulates lung cancer differentiation and metastasis
.
Nature
.
2007
;
448
(
7155
):
807
U807
.

Jishage
K
,
Nezu
J
,
Kawase
Y
,
Iwata
T
,
Watanabe
M
,
Miyoshi
A
,
Ose
A
,
Habu
K
,
Kake
T
,
Kamada
N
, et al. 
Role of Lkb1, the causative gene of Peutz-Jegher's syndrome, in embryogenesis and polyposis
.
Proc Natl Acad Sci U S A
.
2002
;
99
(
13
):
8903
8908
.

Joiner
WJ
,
Crocker
A
,
White
BH
,
Sehgal
A.
Sleep in Drosophila is regulated by adult mushroom bodies
.
Nature
.
2006
;
441
(
7094
):
757
760
.

Klarsfeld
A
,
Leloup
JC
,
Rouyer
F.
Circadian rhythms of locomotor activity in Drosophila
.
Behav Process
.
2003
;
64
(
2
):
161
175
.

Kunst
M
,
Hughes
ME
,
Raccuglia
D
,
Felix
M
,
Li
M
,
Barnett
G
,
Duah
J
,
Nitabach
MN.
Calcitonin gene-related peptide neurons mediate sleep-specific circadian output in Drosophila
.
Curr Biol
.
2014
;
24
(
22
):
2652
2664
.

Liu
QL
,
Liu
S
,
Kodama
L
,
Driscoll
MR
,
Wu
MN.
Two dopaminergic neurons signal to the dorsal fan-shaped body to promote wakefulness in Drosophila
.
Curr Biol
.
2012
;
22
(
22
):
2114
2123
.

Liu
S
,
Liu
QL
,
Tabuchi
M
,
Wu
MN.
Sleep drive is encoded by neural plastic changes in a dedicated circuit
.
Cell
.
2016
;
165
(
6
):
1347
1360
.

Liu
YX
,
Wang
TV
,
Cui
Y
,
Gao
S
,
Rao
Y.
Biochemical purification uncovers mammalian sterile 3 (MST3) as a new protein kinase for multifunctional protein kinases AMPK and SIK3
.
J Biol Chem
.
2022
;
298
(
5
):
101929
.

Liu
YX
,
Wang
TV
,
Cui
Y
,
Li
C
,
Jiang
L
,
Rao
Y.
STE20 phosphorylation of AMPK related kinases revealed by biochemical purifications combined with genetics
.
J Biol Chem
.
2022
;
298
(
5
):
101928
.

Lizcano
JM
,
Goransson
O
,
Toth
R
,
Deak
M
,
Morrice
NA
,
Boudeau
J
,
Hawley
SA
,
Udd
L
,
Makela
TP
,
Hardie
DG
, et al. 
LKB1 is a master kinase that activates 13 kinases of the AMPK subfamily, including MARK/PAR-1
.
EMBO J
.
2004
;
23
(
4
):
833
843
.

Lopez
M
,
Dieguez
C.
Cellular energy sensors: AMPK and beyond
.
Mol Cell Endocrinol
.
2014
;
397
(
1–2
):
1
3
.

Martin
SG
,
St Johnston
D.
A role for Drosophila LKB1 in anterior-posterior axis formation and epithelial polarity
.
Nature
.
2003
;
421
(
6921
):
379
384
.

Matsumoto
S
,
Iwakawa
R
,
Takahashi
K
,
Kohno
T
,
Nakanishi
Y
,
Matsuno
Y
,
Suzuki
K
,
Nakamoto
M
,
Shimizu
E
,
Minna
JD
, et al. 
Prevalence and specificity of LKB1 genetic alterations in lung cancers
.
Oncogene
.
2007
;
26
(
40
):
5911
5918
.

Mehenni
H
,
Gehrig
C
,
Nezu
J
,
Oku
A
,
Shimane
M
,
Rossier
C
,
Guex
N
,
Blouin
JL
,
Scott
HS
,
Antonarakis
SE.
Loss of LKB1 kinase activity in Peutz-Jeghers syndrome, and evidence for allelic and locus heterogeneity
.
Am J Hum Genet
.
1998
;
63
(
6
):
1641
1650
.

Michell
BJ
,
Stapleton
D
,
Mitchelhill
KI
,
House
CM
,
Katsis
F
,
Witters
LA
,
Kemp
BE.
Isoform-specific purification and substrate specificity of the 5'-AMP-activated protein kinase
.
J Biol Chem
.
1996
;
271
(
45
):
28445
28450
.

Mitchelhill
KI
,
Stapleton
D
,
Gao
G
,
House
C
,
Michell
B
,
Katsis
F
,
Witters
LA
,
Kemp
BE.
Mammalian AMP-activated protein-kinase shares structural and functional homology with the catalytic domain of yeast Snf1 protein-kinase
.
J Biol Chem
.
1994
;
269
(
4
):
2361
2364
.

Miyoshi
H
,
Nakau
M
,
Ishikawa
T
,
Seldin
MF
,
Oshima
M
,
Taketo
MM.
Gastrointestinal hamartomatous polyposis in Lkb1 heterozygous knockout mice
.
Cancer Res
.
2002
;
62
(
8
):
2261
2266
.

Mohawk
JA
,
Green
CB
,
Takahashi
JS.
Central and peripheral circadian clocks in mammals
.
Annu Rev Neurosci
.
2012
;
35
:
445
462
.

Morton
JP
,
Jamieson
NB
,
Karim
SA
,
Athineos
D
,
Ridgway
RA
,
Nixon
C
,
Mckay
CJ
,
Carter
R
,
Brunton
VG
,
Frame
MC
, et al. 
Lkb1 haploinsufficiency cooperates with KRAS to promote pancreatic cancer through suppression of p21-dependent growth arrest
.
Gastroenterology
.
2010
;
139
(
2
):
586
597
.

Munday
MR
,
Carling
D
,
Hardie
DG.
Negative interactions between phosphorylation of acetyl-coA carboxylase by the cyclic AMP-dependent and AMP-activated protein-kinases
.
FEBS Lett
.
1988
;
235
(
1–2
):
144
148
.

Nagy
S
,
Maurer
GW
,
Hentze
JL
,
Rose
M
,
Werge
TM
,
Rewitz
K.
AMPK signaling linked to the schizophrenia-associated 1q21.1 deletion is required for neuronal and sleep maintenance
.
PLoS Genet
.
2018
;
14
(
12
):
e1007623
.

Nakada
D
,
Saunders
TL
,
Morrison
SJ.
Lkb1 regulates cell cycle and energy metabolism in haematopoietic stem cells
.
Nature
.
2010
;
468
(
7324
):
653
658
.

Nitabach
MN
,
Taghert
PH.
Organization of the Drosophila circadian control circuit
.
Curr Biol
.
2008
;
18
(
2
):
R84
93
.

Nixon
JP
,
Mavanji
V
,
Butterick
TA
,
Billington
CJ
,
Kotz
CM
,
Teske
JA.
Sleep disorders, obesity, and aging: the role of orexin
.
Ageing Res Rev
.
2015
;
20
:
63
73
.

Parisky
KM
,
Agosto
J
,
Pulver
SR
,
Shang
YH
,
Kuklin
E
,
Hodge
JJL
,
Kang
K
,
Liu
X
,
Garrity
PA
,
Rosbash
M
, et al. 
PDF cells are a GABA-responsive wake-promoting component of the Drosophila sleep circuit
.
Neuron
.
2008
;
60
(
4
):
672
682
.

Park
M
,
Miyoshi
C
,
Fujiyama
T
,
Kakizaki
M
,
Ikkyu
A
,
Honda
T
,
Choi
J
,
Asano
F
,
Mizuno
S
,
Takahashi
S
, et al. 
Loss of the conserved PKA sites of SIK1 and SIK2 increases sleep need
.
Sci Rep
.
2020
;
10
(
1
):
8676
.

Park
S
,
Sonn
JY
,
Oh
Y
,
Lim
C
,
Choe
J.
SIFamide and SIFamide receptor define a novel neuropeptide signaling to promote sleep in Drosophila
.
Mol Cells
.
2014
;
37
(
4
):
295
301
.

Peutz
JLA.
Very remarkable case of familial polyposis of mucous membrane of intestinal tract and nasopharynx accompanied by peculiar pigmentations of skin and mucous membrane
.
Nederl Maandschr Geneesk
.
1921
;
10
:
134
146
.

Pimentel
D
,
Donlea
JM
,
Albot
CBT
,
Ong
SMS
,
Hurston
AJFT
,
Miesenbock
G.
Operation of a homeostatic sleep switch
.
Nature
.
2016
;
536
(
7616
):
333
337
.

Poe
AR
,
Wang
B
,
Sapar
ML
,
Ji
H
,
Li
K
,
Onabajo
T
,
Fazliyeva
R
,
Gibbs
M
,
Qiu
Y
,
Hu
Y
, et al. 
Robust CRISPR/Cas9-mediated tissue-specific mutagenesis reveals gene redundancy and perdurance in Drosophila
.
Genetics
.
2019
;
211
(
2
):
459
472
.

Qian
Y
,
Cao
Y
,
Deng
B
,
Yang
G
,
Li
J
,
Xu
R
,
Zhang
D
,
Huang
J
,
Rao
Y.
Sleep homeostasis regulated by 5HT2b receptor in a small subset of neurons in the dorsal fan-shaped body of Drosophila
.
Elife
.
2017
;
6
:
e26519
.

Rowan
A
,
Bataille
V
,
MacKie
R
,
Healy
E
,
Bicknell
D
,
Bodmer
W
,
Tomlinson
I.
Somatic mutations in the Peutz-Jeghers (LKB1/STKII) gene in sporadic malignant melanomas
.
J Invest Dermatol
.
1999
;
112
(
4
):
509
511
.

Sakamoto
K
,
McCarthy
A
,
Smith
D
,
Green
KA
,
Hardie
DG
,
Ashworth
A
,
Alessi
DR.
Deficiency of LKB1 in skeletal muscle prevents AMPK activation and glucose uptake during contraction
.
EMBO J
.
2005
;
24
(
10
):
1810
1820
.

Sanchez-Cespedes
M
,
Parrella
P
,
Esteller
M
,
Nomoto
S
,
Trink
B
,
Engles
JM
,
Westra
WH
,
Herman
JG
,
Sidransky
D.
Inactivation of LKB1/STK11 is a common event in adenocarcinomas of the lung
.
Cancer Res
.
2002
;
62
(
13
):
3659
3662
.

Saper
CB
,
Fuller
PM.
Wake-sleep circuitry: an overview
.
Curr Opin Neurobiol
.
2017
;
44
:
186
192
.

Sato
F
,
Muragaki
Y
,
Zhang
YP.
DEC1 negatively regulates AMPK activity via LKB1
.
Biochem Bioph Res Co
.
2015
;
467
(
4
):
711
716
.

Scammell
TE
,
Arrigoni
E
,
Lipton
JO.
Neural circuitry of wakefulness and sleep
.
Neuron
.
2017
;
93
(
4
):
747
765
.

Sengupta
S
,
Nagalingam
A
,
Muniraj
N
,
Bonner
MY
,
Mistriotis
P
,
Afthinos
A
,
Kuppusamy
P
,
Lanoue
D
,
Cho
S
,
Korangath
P
, et al. 
Activation of tumor suppressor LKB1 by honokiol abrogates cancer stem-like phenotype in breast cancer via inhibition of oncogenic Stat3
.
Oncogene
.
2017
;
36
(
41
):
5709
5721
.

Seugnet
L
,
Suzuki
Y
,
Merlin
G
,
Gottschalk
L
,
Duntley
SP
,
Shaw
PJ.
Notch signaling modulates sleep homeostasis and learning after sleep deprivation in Drosophila
.
Curr Biol
.
2011
;
21
(
10
):
835
840
.

Shackelford
DB
,
Shaw
RJ.
The lkb1-ampk pathway: metabolism and growth control in tumour suppression
.
Nat Rev Cancer
.
2009
;
9
(
8
):
563
575
.

Shang
YH
,
Donelson
NC
,
Vecsey
CG
,
Guo
F
,
Rosbash
M
,
Griffith
LC.
Short neuropeptide f is a sleep-promoting inhibitory modulator
.
Neuron
.
2013
;
80
(
1
):
171
183
.

Shang
YH
,
Griffith
LC
,
Rosbash
M.
Light-arousal and circadian photoreception circuits intersect at the large PDF cells of the Drosophila brain
.
Proc Natl Acad Sci U S A
.
2008
;
105
(
50
):
19587
19594
.

Shaw
PJ
,
Cirelli
C
,
Greenspan
RJ
,
Tononi
G.
Correlates of sleep and waking in Drosophila melanogaster
.
Science
.
2000
;
287
(
5459
):
1834
1837
.

Shaw
RJ
,
Kosmatka
M
,
Bardeesy
N
,
Hurley
RL
,
Witters
LA
,
DePinho
RA
,
Cantley
LC.
The tumor suppressor LKB1 kinase directly activates amp-activated kinase and regulates apoptosis in response to energy stress
.
Proc Natl Acad Sci U S A
.
2004
;
101
(
10
):
3329
3335
.

Shaw
RJ
,
Lamia
KA
,
Vasquez
D
,
Koo
SH
,
Bardeesy
N
,
DePinho
RA
,
Montminy
M
,
Cantley
LC.
The kinase LKB1 mediates glucose homeostasis in liver and therapeutic effects of metformin
.
Science
.
2005
;
310
(
5754
):
1642
1646
.

Sheeba
V
,
Fogle
KJ
,
Kaneko
M
,
Rashid
S
,
Chou
YT
,
Sharma
VK
,
Holmes
TC.
Large ventral lateral neurons modulate arousal and sleep in Drosophila
.
Curr Biol
.
2008
;
18
(
20
):
1537
1545
.

Shen
Z
,
Wen
XF
,
Lan
F
,
Shen
ZZ
,
Shao
ZM.
The tumor suppressor gene LKB1 is associated with prognosis in human breast carcinoma
.
Clin Cancer Res
.
2002
;
8
(
7
):
2085
2090
.

Skoulidis
F
,
Byers
LA
,
Diao
LX
,
Papadimitrakopoulou
VA
,
Tong
P
,
Izzo
J
,
Behrens
C
,
Kadara
H
,
Parra
ER
,
Canales
JR
, et al. 
Co-occurring genomic alterations define major subsets of KRAS-mutant lung adenocarcinoma with distinct biology, immune profiles, and therapeutic vulnerabilities
.
Cancer Discov
.
2015
;
5
(
8
):
860
877
.

Sutherland
CM
,
Hawley
SA
,
McCartney
RR
,
Leech
A
,
Stark
MJR
,
Schmidt
MC
,
Hardie
DG.
Elm1p is one of three upstream kinases for the saccharomyces cerevisiae SNF1 complex
.
Curr Biol
.
2003
;
13
(
15
):
1299
1305
.

Taheri
S
,
Lin
L
,
Austin
D
,
Young
T
,
Mignot
E.
Short sleep duration is associated with reduced leptin, elevated ghrelin, and increased body mass index (BMI)
.
Sleep
.
2004
;
27
:
146
147
.

Tanwar
PS
,
Mohapatra
G
,
Chiang
S
,
Engler
DA
,
Zhang
LH
,
Kaneko-Tarui
T
,
Ohguchi
Y
,
Birrer
MJ
,
Teixeira
JM.
Loss of LKB1 and PTEN tumor suppressor genes in the ovarian surface epithelium induces papillary serous ovarian cancer
.
Carcinogenesis
.
2014
;
35
(
3
):
546
553
.

Thimgan
MS
,
Suzuki
Y
,
Seugnet
L
,
Gottschalk
L
,
Shaw
PJ.
The perilipin homologue, lipid storage droplet 2, regulates sleep homeostasis and prevents learning impairments following sleep loss
.
PLoS Biol
.
2010
;
8
(
8
):
e1000466
.

Tobler
I
,
Borbely
AA.
Sleep EEG in the rat as a function of prior waking
.
Electroencephalogr Clin Neuro
.
1986
;
64
(
1
):
74
76
.

Tomlinson
IP
,
Houlston
RS.
Peutz-Jeghers syndrome
.
J Med Genet
.
1997
;
34
(
12
):
1007
1011
.

Ueno
T
,
Tomita
J
,
Tanimoto
H
,
Endo
K
,
Ito
K
,
Kume
S
,
Kume
K.
Identification of a dopamine pathway that regulates sleep and arousal in Drosophila
.
Nat Neurosci
.
2012
;
15
(
11
):
1516
1523
.

Wang
B
,
Moya
N
,
Niessen
S
,
Hoover
H
,
Mihaylova
MM
,
Shaw
RJ
,
Yates
JR
III
,
Fischer
WH
,
Thomas
JB
,
Montminy
M.
A hormone-dependent module regulating energy balance
.
Cell
.
2011
;
145
(
4
):
596
606
.

Weber
F
,
Dan
Y.
Circuit-based interrogation of sleep control
.
Nature
.
2016
;
538
(
7623
):
51
59
.

Werth
E
,
Dijk
DJ
,
Achermann
P
,
Borbely
AA.
Dynamics of the sleep EEG after an early evening nap: experimental data and simulations
.
Am J Physiol
.
1996
;
271
(
3 Pt 2
):
R501
510
.

Westerman
AM
,
Entius
MM
,
de Baar
E
,
Boor
PP
,
Koole
R
,
van Velthuysen
ML
,
Offerhaus
GJ
,
Lindhout
D
,
de Rooij
FW
,
Wilson
JH.
Peutz-Jeghers syndrome: 78-year follow-up of the original family
.
Lancet
.
1999
;
353
(
9160
):
1211
1215
.

Wingo
SN
,
Gallardo
TD
,
Akbay
EA
,
Liang
MC
,
Contreras
CM
,
Boren
T
,
Shimamura
T
,
Miller
DS
,
Sharpless
NE
,
Bardeesy
N
, et al. 
Somatic LKB1 mutations promote cervical cancer progression
.
PLoS One
.
2009
;
4
(
4
):
e5137
.

Woods
A
,
Dickerson
K
,
Heath
R
,
Hong
SP
,
Momcilovic
M
,
Johnstone
SR
,
Carlson
M
,
Carling
D.
Ca2+/calmodulin-dependent protein kinase kinase-beta acts upstream of AMP-activated protein kinase in mammalian cells
.
Cell Metab
.
2005
;
2
(
1
):
21
33
.

Woods
A
,
Johnstone
SR
,
Dickerson
K
,
Leiper
FC
,
Fryer
LG
,
Neumann
D
,
Schlattner
U
,
Wallimann
T
,
Carlson
M
,
Carling
D.
LKB1 is the upstream kinase in the AMP-activated protein kinase cascade
.
Curr Biol
.
2003
;
13
(
22
):
2004
2008
.

Yardeni
T
,
Eckhaus
M
,
Morris
HD
,
Huizing
M
,
Hoogstraten-Miller
S.
Retro-orbital injections in mice
.
Lab Anim (NY)
.
2011
;
40
(
5
):
155
160
.

Yeh
LA
,
Kim
KH.
Regulation of acetyl-coA carboxylase - properties of coA activation of acetyl-coA carboxylase
.
Proc Natl Acad Sci U S A
.
1980
;
77
(
6
):
3351
3355
.

Yi
W
,
Zhang
YP
,
Tian
YJ
,
Guo
J
,
Li
Y
,
Guo
AK.
A subset of cholinergic mushroom body neurons requires go signaling to regulate sleep in Drosophila
.
Sleep
.
2013
;
36
(
12
):
1809
1821
.

Yurgel
ME
,
Kakad
P
,
Zandawala
M
,
Nassel
DR
,
Godenschwege
TA
,
Keene
AC.
A single pair of leucokinin neurons are modulated by feeding state and regulate sleep-metabolism interactions
.
PLoS Biol
.
2019
;
17
(
2
):
e2006409
.

Zhang
X
,
Yan
H
,
Luo
Y
,
Huang
Z
,
Rao
Y.
Thermoregulation-independent regulation of sleep by serotonin revealed in mice defective in serotonin synthesis
.
Mol Pharmacol
.
2018
;
93
(
6
):
657
664
.

This article is published and distributed under the terms of the Oxford University Press, Standard Journals Publication Model (https://academic.oup.com/journals/pages/open_access/funder_policies/chorus/standard_publication_model)
Editor: H Bellen
H Bellen
Editor
Search for other works by this author on: