Transcriptomic analyses reveal rhythmic and CLOCK-driven pathways in human skeletal muscle

Laurent Perrin, Ursula Loizides-Mangold, Stéphanie Chanon, Cédric Gobet, Nicolas Hulo, Laura Isenegger, Benjamin D Weger, Eugenia Migliavacca, Aline Charpagne, James A Betts, Jean-Philippe Walhin, Iain Templeman, Keith Stokes, Dylan Thompson, Kostas Tsintzas, Maud Robert, Cedric Howald, Howard Riezman, Jerome N Feige, Leonidas G Karagounis, Jonathan D Johnston, Emmanouil T Dermitzakis, Frédéric Gachon, Etienne Lefai, Charna Dibner, Laurent Perrin, Ursula Loizides-Mangold, Stéphanie Chanon, Cédric Gobet, Nicolas Hulo, Laura Isenegger, Benjamin D Weger, Eugenia Migliavacca, Aline Charpagne, James A Betts, Jean-Philippe Walhin, Iain Templeman, Keith Stokes, Dylan Thompson, Kostas Tsintzas, Maud Robert, Cedric Howald, Howard Riezman, Jerome N Feige, Leonidas G Karagounis, Jonathan D Johnston, Emmanouil T Dermitzakis, Frédéric Gachon, Etienne Lefai, Charna Dibner

Abstract

Circadian regulation of transcriptional processes has a broad impact on cell metabolism. Here, we compared the diurnal transcriptome of human skeletal muscle conducted on serial muscle biopsies in vivo with profiles of human skeletal myotubes synchronized in vitro. More extensive rhythmic transcription was observed in human skeletal muscle compared to in vitro cell culture as a large part of the in vivo mRNA rhythmicity was lost in vitro. siRNA-mediated clock disruption in primary myotubes significantly affected the expression of ~8% of all genes, with impact on glucose homeostasis and lipid metabolism. Genes involved in GLUT4 expression, translocation and recycling were negatively affected, whereas lipid metabolic genes were altered to promote activation of lipid utilization. Moreover, basal and insulin-stimulated glucose uptake were significantly reduced upon CLOCK depletion. Our findings suggest an essential role for the circadian coordination of skeletal muscle glucose homeostasis and lipid metabolism in humans.

Keywords: RNA sequencing; circadian oscillators; human; human biology; human skeletal muscle; medicine.

Conflict of interest statement

LP, UL, SC, NH, LI, JW, IT, KS, DT, KT, MR, CH, HR, JJ, ED, FG, EL, CD No competing interests declared, CG Is a full-time employee of the Nestlé Institute of Health Sciences SA. BW Benjamin D Weger: Is a full-time employee of the Nestlé Institute of Health Sciences SA. EM Eugenia Migliavacca: Is a full-time employee of the Nestlé Institute of Health Sciences SA. AC Aline Charpagne: Is a full-time employee of the Nestlé Institute of Health Sciences SA. JB Has been a consultant for PepsiCo (Quaker) and Kellogg's. JF Jerome N Feige: Is a full-time employee of the Nestlé Institute of Health Sciences SA. LK Is an employee of Nestec Ltd.

© 2018, Perrin et al.

Figures

Figure 1.. Rhythmic gene expression in human…
Figure 1.. Rhythmic gene expression in human skeletal muscle.
(A) Heat map showing genes rhythmic at the pre-mRNA and mRNA level (R-I.R-E: upper panel), at the pre-mRNA level only (R-I: middle panel), and at the mRNA level only (R-E: lower panel). Standardized relative gene expression is indicated in green (low) and red (high) and ordered by their respective phase. (B) mRNA half-life proxy by exon/intron ratio showing lower stability for genes with rhythmic mRNA (R–E) profiles. (C) Amplitude distribution of genes that are rhythmic only at the mRNA level (R-E, blue), the pre-mRNA level (R-I, red), or rhythmic for both (R-I.R-E). Genes with higher amplitude of transcription at the pre-mRNA level have a higher probability to be rhythmic at the mRNA level (R-I.R-E). (D) Number of genes in each group in relation to the -log10 BH corrected p-value; dashed line indicates threshold of 0.05. (E) Phase distribution at the pre-mRNA and mRNA level for the three groups described in (A). (F) Phase distribution for genes activated by acute muscle exercise (red), inflammation (blue), or both (green). (G) Temporal expression of core clock components, and (H) key muscle transcription factors. N = 10 human muscle biopsy donors. (I) Phase distribution of predicted rhythmic DNA motif activity.
Figure 1—figure supplement 1.. Temporal gene expression…
Figure 1—figure supplement 1.. Temporal gene expression in human skeletal muscle.
(A) Temporal expression of genes involved in vesicular trafficking; (B) Temporal expression of genes involved in glucose uptake; (C) Temporal expression of genes involved in lipid metabolism analyzed by RNA-seq. N = 10 human muscle biopsy donors.
Figure 2.. Disruption of the circadian oscillator…
Figure 2.. Disruption of the circadian oscillator impacts on functional gene expression hSKM.
Differential gene expression analysis between hSKM bearing a disrupted (siCLOCK) or intact (siControl) circadian clock. Comparison of the median gene expression at all analyzed circadian time points between the two groups. A total of 16,776 genes were detected by RNA-seq as expressed above the threshold of counts per million (CPM) >3. (A) Core clock genes; (B) 15,446 genes remained unchanged (dark blue), and 1330 genes exhibited a significantly different expression level upon siCLOCK-mediated clock disruption (light blue), with 588 being up-regulated (orange) and 742 down-regulated (grey) (FDR <0.05 and p-value <0.05). Altered genes comprised those related to glycerophospholipid and triglyceride metabolism, storage and transport (C) inositol phosphate metabolism (D) and sphingolipid metabolism and storage (E). (F) Relative levels of PC, PE, PI, PS, Cer GlcCer, SM, CL and TAG, analyzed by mass spectrometry based lipidomics in hSKM cells transfected with either siControl (orange bar) or siCLOCK (blue bar). Total phosphatidylcholine (PC) and glycosylceramide (GlcCer) levels are significantly decreased or increased, respectively, upon siCLOCK transfection. Data are mean ± SEM, N = 4 (# p-value <0.05). (*) for FDR <0.05, (**) for FDR <0.01, (***) for FDR <0.001.
Figure 2—figure supplement 1.. Study design.
Figure 2—figure supplement 1.. Study design.
Primary human myotubes were transfected with 20 nM of non-targeting siRNA (siControl) or with same amount of siCLOCK at day 6 of differentiation. 24 hr following siRNA transfection, in vitro synchronization by 10 µM forskolin pulse during 1 hr was conducted for all the dishes, except for those collected at time point 0 hr. Cells were harvested in duplicate dishes for each time point and for each donor, every 2 hr during 48 hr. Extracted RNA samples were subjected to RNA sequencing analysis or to real-time quantitative PCR (RT-qPCR).
Figure 2—figure supplement 2.. siRNA-mediated CLOCK knockdown…
Figure 2—figure supplement 2.. siRNA-mediated CLOCK knockdown leads to a flattening of the Bmal1-luc circadian oscillation amplitude in hSKM.
Human primary myoblasts were transduced with lentiviral particles expressing Bmal1-luc and differentiated into myotubes. siControl or siCLOCK transfection was performed 24 hr prior synchronization with forskolin. (A) CLOCK mRNA was measured by RT-qPCR in samples used for RNA-seq and collected at time points 0 hr (before synchronization), 24 hr and 48 hr. Data were normalized to the mean of 9S and HPRT. CLOCK expression was reduced by 82 ± 4% (mean of the three time ± SEM, N = 2, done in duplicates; (***) p-value <0.001) in siCLOCK-transfected cells. (B) Bmal1-luc bioluminescence profile shows an important reduction of amplitude in siCLOCK-transfected myotubes compared to controls. Representative Bmal1-luc oscillation profiles are shown for non-transfected (light blue line), siControl (black line), and siCLOCK (red line) transfected myotubes. Bmal1-luc oscillation profiles were recorded in duplicates (two experiments, one donor per experiment).
Figure 3.. Basal and insulin-induced glucose uptake…
Figure 3.. Basal and insulin-induced glucose uptake by hSKM is downregulated in the absence of a functional circadian clock.
(A) CLOCK mRNA was measured in hSKM cells transfected with siControl or siCLOCK by RT-qPCR and normalized to the mean of 9S and HPRT. CLOCK expression was reduced by 91 ± 2% (mean ± SEM, N = 4; (***) p-value <0.001) in siCLOCK-transfected cells. Protein levels of CLOCK (B) and 14-3-3θ (C) were assessed by western blot. Left panel: representative western blot; right panel protein quantification over all monoplicates (mean ± SEM, N = 5). CLOCK and 14-3-3θ protein levels were reduced by 75 ± 2%, and 28 ± 8%, respectively. (D) Glucose uptake rates (in pmol/mg.min) measured in the absence (basal) or presence (insulin) of insulin (1 hr, 100 nM) in siControl or siCLOCK-transfected cells. Note significant reduction of basal (31 ± 3%) and insulin-stimulated glucose uptake (28 ± 3%). Data are mean ± SEM of four independent experiments using myotubes from four different donors (same as for A-C). (*) p-value <0.05, (**) p-value <0.01, (***) p-value <0.001.
Figure 3—figure supplement 1.. TBC1D4/AS160 protein levels…
Figure 3—figure supplement 1.. TBC1D4/AS160 protein levels are not affected by siCLOCK.
Representative western blot for TBC1D4/AS160 in hSKM cells transfected with either siControl or siCLOCK for 72 hr. Quantification of TBC1D4/AS160 protein levels shows no statistically significant difference of upon CLOCK disruption. Data are mean ± SEM (N = 5, monoplicates).
Figure 4.. Temporal gene expression analysis in…
Figure 4.. Temporal gene expression analysis in human skeletal myotubes bearing a disrupted or functional circadian clock.
(A) Total of 12,985 genes were identified by RNA-seq as expressed above log2 RPKM >0. Genes were subjected to the circadian analysis, adapted for high-resolution datasets over 48 hr. Genes were categorized as rhythmic or non-rhythmic (NR) (left diagram) and rhythmic genes (994) were grouped into 15 models (right panel). Genes that were non-rhythmic in either one of the 15 models (11,991 genes) are represented in model 16. (B) Heat maps for genes assigned to models 1 to 4. Relative expression is indicated in green (low) and red (high). (C) Individual temporal expression profiles of core clock genes ARNTL, NR1D1, NR1D2, CRY1, CRY2, PER1, PER2 and PER3 in siControl or siCLOCK-transfected cells. (D) Upper panel: Representative examples for genes assigned to models 1–4. Lower panel: Circadian amplitude quantification of siControl and siCLOCK in models 1–4.
Figure 4—figure supplement 1.. Comparison of rhythmic…
Figure 4—figure supplement 1.. Comparison of rhythmic transcript profiles between the two donors.
(A) More rhythmic transcripts are found in donor two compared to donor one in both siControl and siCLOCK condition either due to the genetic differences between both donors or due to a better response to the forskolin pulse used for synchronization. (B) Circadian amplitude quantification for the summary of models 1–4 in siControl and siCLOCK. (C) Examples of genes qualified as rhythmic in model 1–4.
Figure 4—figure supplement 2.. Temporal profiles of…
Figure 4—figure supplement 2.. Temporal profiles of core clock transcript expression.
(A) Expression profile of NPAS2, TEF and BHLHE41 assessed by RNA-seq. (B) ARNTL (BMAL1), NR1D1 (REVERBα) and PER3 measured by RT-qPCR. mRNA levels were analyzed every 4 hr over 48 hr for each donor to confirm the RNA-seq results. Data were normalized to the mean of 9S and HPRT. For donor one data represent monoplicates, whereas for donor two data were done in duplicates. Donor 1: circle; donor 2: triangle; siControl: black line; siCLOCK: red line.
Figure 4—figure supplement 3.. Genes involved in…
Figure 4—figure supplement 3.. Genes involved in cell cycle regulation exhibit circadian expression profile in hSKM.
CCNB1, CDC6, CDC7, CDC20, CDK1, E2F1, E2F7, MCM2, MCM5, MCM7, PLK1, PRC1, RBL1/p107 and TPX2 expression profiles obtained by RNA-seq. All genes shown were classified in model 1 except for E2F1 and MCM5 (model 3).
Figure 5.. Comparison between the circadian transcriptome…
Figure 5.. Comparison between the circadian transcriptome of human skeletal muscle and human primary myotubes.
(A) Scatter plot, representing the amplitude of expression in relation to the corrected p-value for genes that were rhythmic in vivo (human muscle biopsies). Genes that were also rhythmic in vitro (hSKM, models 1–4) are colored in red. Blue dots represent genes with a p-value <0.01 and log2 amp >0.5. (B) Phase distribution plot of genes rhythmic in muscle biopsies and primary myotubes shows enrichment at the 04:00 time point. (C) Examples of genes, involved in glucose homeostasis and muscle regeneration, that are rhythmic in vivo and in vitro (RNA-seq data, N = 10).
Figure 6.. Schema summarizing impact of clock…
Figure 6.. Schema summarizing impact of clock disruption on muscle metabolic function.
Clock disruption leads to impaired insulin sensitivity and decrease in glucose uptake (1), causes a dysregulation of genes involved in vesicle trafficking (2) and impacts lipid metabolism and lipid metabolite oscillations (3) as reported in Loizides-Mangold et al. (2017).
Author response image 1.. Comparison of our…
Author response image 1.. Comparison of our in vivo vastus lateralis rhythmic dataset with mouse skeletal muscle rhythmic data (Dyar et al., 2014; Zhang et al., 2014).
Blue = our in vivo RNAseq data; red = (Dyar et al., 2014) soleus; green = (Zhang et al., 2014). The numbers presented in the Venn diagram correspond to the total number of rhythmically expressed genes for each segment (exons only). For the human in vivo dataset (our work), this corresponds to R–E + R–I.R–E, with the cutoff described in the manuscript (Figure 1). For (Dyar et al., 2014) (microarray) and (Zhang et al., 2014) (microarray) rhythmic genes were identified using a Benjamini–Hochberg q–value

Author response image 2.

Author response image 2.

Author response image 2.

Author response image 3.. Examples of the…

Author response image 3.. Examples of the gene expression profiles from our database qualified as…

Author response image 3.. Examples of the gene expression profiles from our database qualified as circadian by JTK_CYCLE.
All figures (16)
Author response image 2.
Author response image 2.
Author response image 3.. Examples of the…
Author response image 3.. Examples of the gene expression profiles from our database qualified as circadian by JTK_CYCLE.

References

    1. Aas V, Bakke SS, Feng YZ, Kase ET, Jensen J, Bajpeyi S, Thoresen GH, Rustan AC. Are cultured human myotubes far from home? Cell and Tissue Research. 2013;354:671–682. doi: 10.1007/s00441-013-1655-1.
    1. Adamovich Y, Rousso-Noori L, Zwighaft Z, Neufeld-Cohen A, Golik M, Kraut-Cohen J, Wang M, Han X, Asher G. Circadian clocks and feeding time regulate the oscillations and levels of hepatic triglycerides. Cell Metabolism. 2014;19:319–330. doi: 10.1016/j.cmet.2013.12.016.
    1. Agley CC, Rowlerson AM, Velloso CP, Lazarus NL, Harridge SDR. Isolation and quantitative immunocytochemical characterization of primary myogenic cells and fibroblasts from human skeletal muscle. Journal of Visualized Experiments. 2015;95:52049. doi: 10.3791/52049.
    1. Albrecht U. Timing to perfection: the biology of central and peripheral circadian clocks. Neuron. 2012;74:246–260. doi: 10.1016/j.neuron.2012.04.006.
    1. An D, Toyoda T, Taylor EB, Yu H, Fujii N, Hirshman MF, Goodyear LJ. TBC1D1 regulates insulin- and contraction-induced glucose transport in mouse skeletal muscle. Diabetes. 2010;59:1358–1365. doi: 10.2337/db09-1266.
    1. Andrews JL, Zhang X, McCarthy JJ, McDearmon EL, Hornberger TA, Russell B, Campbell KS, Arbogast S, Reid MB, Walker JR, Hogenesch JB, Takahashi JS, Esser KA. CLOCK and BMAL1 regulate MyoD and are necessary for maintenance of skeletal muscle phenotype and function. PNAS. 2010;107:19090–19095. doi: 10.1073/pnas.1014523107.
    1. Aronson D, Violan MA, Dufresne SD, Zangen D, Fielding RA, Goodyear LJ. Exercise stimulates the mitogen-activated protein kinase pathway in human skeletal muscle. Journal of Clinical Investigation. 1997;99:1251–1257. doi: 10.1172/JCI119282.
    1. Aronson D, Wojtaszewski JF, Thorell A, Nygren J, Zangen D, Richter EA, Ljungqvist O, Fielding RA, Goodyear LJ. Extracellular-regulated protein kinase cascades are activated in response to injury in human skeletal muscle. American Journal of Physiology-Cell Physiology. 1998;275:C555–C561. doi: 10.1152/ajpcell.1998.275.2.C555.
    1. Asher G, Sassone-Corsi P. Time for food: the intimate interplay between nutrition, metabolism, and the circadian clock. Cell. 2015;161:84–92. doi: 10.1016/j.cell.2015.03.015.
    1. Atger F, Gobet C, Marquis J, Martin E, Wang J, Weger B, Lefebvre G, Descombes P, Naef F, Gachon F. Circadian and feeding rhythms differentially affect rhythmic mRNA transcription and translation in mouse liver. PNAS. 2015;112:E6579–E6588. doi: 10.1073/pnas.1515308112.
    1. Baati N, Feillet-Coudray C, Fouret G, Vernus B, Goustard B, Coudray C, Lecomte J, Blanquet V, Magnol L, Bonnieu A, Koechlin-Ramonatxo C. Myostatin deficiency is associated with lipidomic abnormalities in skeletal muscles. Biochimica Et Biophysica Acta (BBA) - Molecular and Cell Biology of Lipids. 2017;1862:1044–1055. doi: 10.1016/j.bbalip.2017.06.017.
    1. Bates D, Mächler M, Bolker B, Walker S. Fitting Linear Mixed-Effects Models Usinglme4. Journal of Statistical Software. 2015;67 doi: 10.18637/jss.v067.i01.
    1. Benedict C, Shostak A, Lange T, Brooks SJ, Schiöth HB, Schultes B, Born J, Oster H, Hallschmid M. Diurnal rhythm of circulating nicotinamide phosphoribosyltransferase (Nampt/visfatin/PBEF): impact of sleep loss and relation to glucose metabolism. The Journal of Clinical Endocrinology & Metabolism. 2012;97:E218–E222. doi: 10.1210/jc.2011-2241.
    1. Bergstrom J. Muscle Electrolytes in Man - Determined by Neutron Activation Analysis on Needle Biopsy Specimens - Study on Normal Subjects, Kidney Patients, and Patients with Chronic Diarrhoea. Scandinavian Journal of Clinical and Laboratory Investigation. 1962;14
    1. Catoire M, Mensink M, Boekschoten MV, Hangelbroek R, Müller M, Schrauwen P, Kersten S. Pronounced effects of acute endurance exercise on gene expression in resting and exercising human skeletal muscle. PLoS ONE. 2012;7:e51066. doi: 10.1371/journal.pone.0051066.
    1. Chanon S, Durand C, Vieille-Marchiset A, Robert M, Dibner C, Simon C, Lefai E. Glucose uptake measurement and response to insulin stimulation in in vitro cultured human primary myotubes. Journal of Visualized Experiments. 2017 doi: 10.3791/55743.
    1. Chatterjee S, Ma K. Circadian clock regulation of skeletal muscle growth and repair. F1000Research. 2016;5:1549. doi: 10.12688/f1000research.9076.1.
    1. Coleman SK, Rebalka IA, D'Souza DM, Deodhare N, Desjardins EM, Hawke TJ. Myostatin inhibition therapy for insulin-deficient type 1 diabetes. Scientific Reports. 2016;6:32495. doi: 10.1038/srep32495.
    1. Coomer M, Essop MF. Differential hexosamine biosynthetic pathway gene expression with type 2 diabetes. Molecular Genetics and Metabolism Reports. 2014;1:158–169. doi: 10.1016/j.ymgmr.2014.03.003.
    1. DeBruyne JP, Weaver DR, Reppert SM. CLOCK and NPAS2 have overlapping roles in the suprachiasmatic circadian clock. Nature Neuroscience. 2007;10:543–545. doi: 10.1038/nn1884.
    1. DeFronzo RA, Jacot E, Jequier E, Maeder E, Wahren J, Felber JP. The effect of insulin on the disposal of intravenous glucose. Results from indirect calorimetry and hepatic and femoral venous catheterization. Diabetes. 1981;30:1000–1007. doi: 10.2337/diab.30.12.1000.
    1. DeFronzo RA, Tripathy D. Skeletal muscle insulin resistance is the primary defect in type 2 diabetes. Diabetes Care. 2009;32 Suppl 2:S157–S163. doi: 10.2337/dc09-S302.
    1. Delevoye C, Miserey-Lenkei S, Montagnac G, Gilles-Marsens F, Paul-Gilloteaux P, Giordano F, Waharte F, Marks MS, Goud B, Raposo G. Recycling endosome tubule morphogenesis from sorting endosomes requires the kinesin motor KIF13A. Cell Reports. 2014;6:445–454. doi: 10.1016/j.celrep.2014.01.002.
    1. Dibner C, Schibler U. Circadian timing of metabolism in animal models and humans. Journal of Internal Medicine. 2015;277:513–527. doi: 10.1111/joim.12347.
    1. Dobin A, Davis CA, Schlesinger F, Drenkow J, Zaleski C, Jha S, Batut P, Chaisson M, Gingeras TR. STAR: ultrafast universal RNA-seq aligner. Bioinformatics. 2013;29:15–21. doi: 10.1093/bioinformatics/bts635.
    1. Dugani CB, Randhawa VK, Cheng AW, Patel N, Klip A. Selective regulation of the perinuclear distribution of glucose transporter 4 (GLUT4) by insulin signals in muscle cells. European Journal of Cell Biology. 2008;87:337–351. doi: 10.1016/j.ejcb.2008.02.009.
    1. Dyar KA, Ciciliot S, Tagliazucchi GM, Pallafacchina G, Tothova J, Argentini C, Agatea L, Abraham R, Ahdesmäki M, Forcato M, Bicciato S, Schiaffino S, Blaauw B. The calcineurin-NFAT pathway controls activity-dependent circadian gene expression in slow skeletal muscle. Molecular Metabolism. 2015;4:823–833. doi: 10.1016/j.molmet.2015.09.004.
    1. Dyar KA, Ciciliot S, Wright LE, Biensø RS, Tagliazucchi GM, Patel VR, Forcato M, Paz MI, Gudiksen A, Solagna F, Albiero M, Moretti I, Eckel-Mahan KL, Baldi P, Sassone-Corsi P, Rizzuto R, Bicciato S, Pilegaard H, Blaauw B, Schiaffino S. Muscle insulin sensitivity and glucose metabolism are controlled by the intrinsic muscle clock. Molecular Metabolism. 2014;3:29–41. doi: 10.1016/j.molmet.2013.10.005.
    1. Fecchi K, Volonte D, Hezel MP, Schmeck K, Galbiati F. Spatial and temporal regulation of GLUT4 translocation by flotillin-1 and caveolin-3 in skeletal muscle cells. The FASEB Journal. 2006;20:705–707. doi: 10.1096/fj.05-4661fje.
    1. Fischer KM, Cottage CT, Wu W, Din S, Gude NA, Avitabile D, Quijada P, Collins BL, Fransioli J, Sussman MA. Enhancement of myocardial regeneration through genetic engineering of cardiac progenitor cells expressing Pim-1 kinase. Circulation. 2009;120:2077–2087. doi: 10.1161/CIRCULATIONAHA.109.884403.
    1. Frederick DW, Loro E, Liu L, Davila A, Chellappa K, Silverman IM, Quinn WJ, Gosai SJ, Tichy ED, Davis JG, Mourkioti F, Gregory BD, Dellinger RW, Redpath P, Migaud ME, Nakamaru-Ogiso E, Rabinowitz JD, Khurana TS, Baur JA. Loss of NAD homeostasis leads to progressive and reversible degeneration of skeletal muscle. Cell Metabolism. 2016;24:269–282. doi: 10.1016/j.cmet.2016.07.005.
    1. Frias MA, Thoreen CC, Jaffe JD, Schroder W, Sculley T, Carr SA, Sabatini DM. mSin1 is necessary for Akt/PKB phosphorylation, and its isoforms define three distinct mTORC2s. Current Biology. 2006;16:1865–1870. doi: 10.1016/j.cub.2006.08.001.
    1. Friedmann-Bette B, Schwartz FR, Eckhardt H, Billeter R, Bonaterra G, Kinscherf R. Similar changes of gene expression in human skeletal muscle after resistance exercise and multiple fine needle biopsies. Journal of Applied Physiology. 2012;112:289–295. doi: 10.1152/japplphysiol.00959.2011.
    1. Gachon F, Loizides-Mangold U, Petrenko V, Dibner C. Glucose homeostasis: regulation by peripheral circadian clocks in rodents and humans. Endocrinology. 2017;158:1074–1084. doi: 10.1210/en.2017-00218.
    1. Garten A, Schuster S, Penke M, Gorski T, de Giorgis T, Kiess W. Physiological and pathophysiological roles of NAMPT and NAD metabolism. Nature Reviews Endocrinology. 2015;11:535–546. doi: 10.1038/nrendo.2015.117.
    1. Goueli BS, Powell MB, Finger EC, Pfeffer SR. TBC1D16 is a Rab4A GTPase activating protein that regulates receptor recycling and EGF receptor signaling. PNAS. 2012;109:15787–15792. doi: 10.1073/pnas.1204540109.
    1. Guillaumond F, Gréchez-Cassiau A, Subramaniam M, Brangolo S, Peteri-Brünback B, Staels B, Fiévet C, Spelsberg TC, Delaunay F, Teboul M. Kruppel-like factor KLF10 is a link between the circadian clock and metabolism in liver. Molecular and Cellular Biology. 2010;30:3059–3070. doi: 10.1128/MCB.01141-09.
    1. Haldar SM, Jeyaraj D, Anand P, Zhu H, Lu Y, Prosdocimo DA, Eapen B, Kawanami D, Okutsu M, Brotto L, Fujioka H, Kerner J, Rosca MG, McGuinness OP, Snow RJ, Russell AP, Gerber AN, Bai X, Yan Z, Nosek TM, Brotto M, Hoppel CL, Jain MK. Kruppel-like factor 15 regulates skeletal muscle lipid flux and exercise adaptation. PNAS. 2012;109:6739–6744. doi: 10.1073/pnas.1121060109.
    1. Hansen J, Timmers S, Moonen-Kornips E, Duez H, Staels B, Hesselink MK, Schrauwen P. Synchronized human skeletal myotubes of lean, obese and type 2 diabetic patients maintain circadian oscillation of clock genes. Scientific Reports. 2016;6:35047. doi: 10.1038/srep35047.
    1. Harfmann BD, Schroder EA, Esser KA. Circadian rhythms, the molecular clock, and skeletal muscle. Journal of Biological Rhythms. 2015;30:84–94. doi: 10.1177/0748730414561638.
    1. Harfmann BD, Schroder EA, Kachman MT, Hodge BA, Zhang X, Esser KA. Muscle-specific loss of Bmal1 leads to disrupted tissue glucose metabolism and systemic glucose homeostasis. Skeletal Muscle. 2016;6:12. doi: 10.1186/s13395-016-0082-x.
    1. Hartler J, Trötzmüller M, Chitraju C, Spener F, Köfeler HC, Thallinger GG. Lipid data analyzer: unattended identification and quantitation of lipids in LC-MS data. Bioinformatics. 2011;27:572–577. doi: 10.1093/bioinformatics/btq699.
    1. Hendershott MC, Vale RD. Regulation of microtubule minus-end dynamics by CAMSAPs and Patronin. PNAS. 2014;111:5860–5865. doi: 10.1073/pnas.1404133111.
    1. Hodge BA, Wen Y, Riley LA, Zhang X, England JH, Harfmann BD, Schroder EA, Esser KA. The endogenous molecular clock orchestrates the temporal separation of substrate metabolism in skeletal muscle. Skeletal Muscle. 2015;5:17. doi: 10.1186/s13395-015-0039-5.
    1. Huang P, Altshuller YM, Hou JC, Pessin JE, Frohman MA. Insulin-stimulated plasma membrane fusion of Glut4 glucose transporter-containing vesicles is regulated by phospholipase D1. Molecular Biology of the Cell. 2005;16:2614–2623. doi: 10.1091/mbc.E04-12-1124.
    1. Hughes ME, DiTacchio L, Hayes KR, Vollmers C, Pulivarthy S, Baggs JE, Panda S, Hogenesch JB. Harmonics of circadian gene transcription in mammals. PLoS Genetics. 2009;5:e1000442. doi: 10.1371/journal.pgen.1000442.
    1. Hughes ME, Hogenesch JB, Kornacker K. JTK_CYCLE: an efficient nonparametric algorithm for detecting rhythmic components in genome-scale data sets. Journal of Biological Rhythms. 2010;25:372–380. doi: 10.1177/0748730410379711.
    1. Jaldin-Fincati JR, Pavarotti M, Frendo-Cumbo S, Bilan PJ, Klip A. Update on GLUT4 Vesicle Traffic: A Cornerstone of Insulin Action. Trends in Endocrinology & Metabolism. 2017;28:597–611. doi: 10.1016/j.tem.2017.05.002.
    1. Jordan SD, Kriebs A, Vaughan M, Duglan D, Fan W, Henriksson E, Huber AL, Papp SJ, Nguyen M, Afetian M, Downes M, Yu RT, Kralli A, Evans RM, Lamia KA. CRY1/2 selectively repress PPARδ and limit exercise capacity. Cell Metabolism. 2017;26:243–255. doi: 10.1016/j.cmet.2017.06.002.
    1. Jović M, Kean MJ, Dubankova A, Boura E, Gingras AC, Brill JA, Balla T. Endosomal sorting of VAMP3 is regulated by PI4K2A. Journal of Cell Science. 2014;127:3745–3756. doi: 10.1242/jcs.148809.
    1. Kalsbeek A, la Fleur S, Fliers E. Circadian control of glucose metabolism. Molecular Metabolism. 2014;3:372–383. doi: 10.1016/j.molmet.2014.03.002.
    1. Kennaway DJ, Owens JA, Voultsios A, Boden MJ, Varcoe TJ. Metabolic homeostasis in mice with disrupted Clock gene expression in peripheral tissues. American Journal of Physiology-Regulatory, Integrative and Comparative Physiology. 2007;293:R1528–R1537. doi: 10.1152/ajpregu.00018.2007.
    1. Kleppe R, Martinez A, Døskeland SO, Haavik J. The 14-3-3 proteins in regulation of cellular metabolism. Seminars in Cell & Developmental Biology. 2011;22:713–719. doi: 10.1016/j.semcdb.2011.08.008.
    1. Krishnaiah SY, Wu G, Altman BJ, Growe J, Rhoades SD, Coldren F, Venkataraman A, Olarerin-George AO, Francey LJ, Mukherjee S, Girish S, Selby CP, Cal S, Er U, Sianati B, Sengupta A, Anafi RC, Kavakli IH, Sancar A, Baur JA, Dang CV, Hogenesch JB, Weljie AM. Clock regulation of metabolites reveals coupling between transcription and metabolism. Cell Metabolism. 2017;25:961–974. doi: 10.1016/j.cmet.2017.03.019.
    1. Laing EE, Johnston JD, Möller-Levet CS, Bucca G, Smith CP, Dijk DJ, Archer SN. Exploiting human and mouse transcriptomic data: Identification of circadian genes and pathways influencing health. BioEssays. 2015;37:544–556. doi: 10.1002/bies.201400193.
    1. Lavallée G, Andelfinger G, Nadeau M, Lefebvre C, Nemer G, Horb ME, Nemer M. The Kruppel-like transcription factor KLF13 is a novel regulator of heart development. The EMBO Journal. 2006;25:5201–5213. doi: 10.1038/sj.emboj.7601379.
    1. Leto D, Saltiel AR. Regulation of glucose transport by insulin: traffic control of GLUT4. Nature Reviews Molecular Cell Biology. 2012;13:383–396. doi: 10.1038/nrm3351.
    1. Liu AC, Tran HG, Zhang EE, Priest AA, Welsh DK, Kay SA. Redundant function of REV-ERBalpha and beta and non-essential role for Bmal1 cycling in transcriptional regulation of intracellular circadian rhythms. PLoS Genetics. 2008;4:e1000023. doi: 10.1371/journal.pgen.1000023.
    1. Lluís F, Roma J, Suelves M, Parra M, Aniorte G, Gallardo E, Illa I, Rodríguez L, Hughes SM, Carmeliet P, Roig M, Muñoz-Cánoves P. Urokinase-dependent plasminogen activation is required for efficient skeletal muscle regeneration in vivo. Blood. 2001;97:1703–1711. doi: 10.1182/blood.V97.6.1703.
    1. Loizides-Mangold U, Koren-Gluzer M, Skarupelova S, Makhlouf AM, Hayek T, Aviram M, Dibner C. Paraoxonase 1 (PON1) and pomegranate influence circadian gene expression and period length. Chronobiology International. 2016;33:453–461. doi: 10.3109/07420528.2016.1154067.
    1. Loizides-Mangold U, Perrin L, Vandereycken B, Betts JA, Walhin JP, Templeman I, Chanon S, Weger BD, Durand C, Robert M, Paz Montoya J, Moniatte M, Karagounis LG, Johnston JD, Gachon F, Lefai E, Riezman H, Dibner C. Lipidomics reveals diurnal lipid oscillations in human skeletal muscle persisting in cellular myotubes cultured in vitro. PNAS. 2017;114:E8565–E8574. doi: 10.1073/pnas.1705821114.
    1. Lück S, Thurley K, Thaben PF, Westermark PO. Rhythmic degradation explains and unifies circadian transcriptome and proteome data. Cell Reports. 2014;9:741–751. doi: 10.1016/j.celrep.2014.09.021.
    1. Manderson AP, Kay JG, Hammond LA, Brown DL, Stow JL. Subcompartments of the macrophage recycling endosome direct the differential secretion of IL-6 and TNFalpha. The Journal of Cell Biology. 2007;178:57–69. doi: 10.1083/jcb.200612131.
    1. Mannic T, Meyer P, Triponez F, Pusztaszeri M, Le Martelot G, Mariani O, Schmitter D, Sage D, Philippe J, Dibner C. Circadian clock characteristics are altered in human thyroid malignant nodules. The Journal of Clinical Endocrinology & Metabolism. 2013;98:4446–4456. doi: 10.1210/jc.2013-2568.
    1. Mansueto G, Armani A, Viscomi C, D'Orsi L, De Cegli R, Polishchuk EV, Lamperti C, Di Meo I, Romanello V, Marchet S, Saha PK, Zong H, Blaauw B, Solagna F, Tezze C, Grumati P, Bonaldo P, Pessin JE, Zeviani M, Sandri M, Ballabio A. Transcription factor EB controls metabolic flexibility during exercise. Cell Metabolism. 2017;25:182–196. doi: 10.1016/j.cmet.2016.11.003.
    1. Marcheva B, Ramsey KM, Buhr ED, Kobayashi Y, Su H, Ko CH, Ivanova G, Omura C, Mo S, Vitaterna MH, Lopez JP, Philipson LH, Bradfield CA, Crosby SD, JeBailey L, Wang X, Takahashi JS, Bass J. Disruption of the clock components CLOCK and BMAL1 leads to hypoinsulinaemia and diabetes. Nature. 2010;466:627–631. doi: 10.1038/nature09253.
    1. Martin TF. PI(4,5)P₂-binding effector proteins for vesicle exocytosis. Biochimica Et Biophysica Acta (BBA) - Molecular and Cell Biology of Lipids. 2015;1851:785–793. doi: 10.1016/j.bbalip.2014.09.017.
    1. Matyash V, Liebisch G, Kurzchalia TV, Shevchenko A, Schwudke D. Lipid extraction by methyl-tert-butyl ether for high-throughput lipidomics. Journal of Lipid Research. 2008;49:1137–1146. doi: 10.1194/jlr.D700041-JLR200.
    1. McCarthy JJ, Andrews JL, McDearmon EL, Campbell KS, Barber BK, Miller BH, Walker JR, Hogenesch JB, Takahashi JS, Esser KA. Identification of the circadian transcriptome in adult mouse skeletal muscle. Physiological Genomics. 2007;31:86–95. doi: 10.1152/physiolgenomics.00066.2007.
    1. McGinnis GR, Tang Y, Brewer RA, Brahma MK, Stanley HL, Shanmugam G, Rajasekaran NS, Rowe GC, Frank SJ, Wende AR, Abel ED, Taegtmeyer H, Litovsky S, Darley-Usmar V, Zhang J, Chatham JC, Young ME. Genetic disruption of the cardiomyocyte circadian clock differentially influences insulin-mediated processes in the heart. Journal of Molecular and Cellular Cardiology. 2017;110:80–95. doi: 10.1016/j.yjmcc.2017.07.005.
    1. Mi H, Huang X, Muruganujan A, Tang H, Mills C, Kang D, Thomas PD. PANTHER version 11: expanded annotation data from gene ontology and reactome pathways, and data analysis tool enhancements. Nucleic Acids Research. 2017;45:D183–D189. doi: 10.1093/nar/gkw1138.
    1. Miller BH, McDearmon EL, Panda S, Hayes KR, Zhang J, Andrews JL, Antoch MP, Walker JR, Esser KA, Hogenesch JB, Takahashi JS. Circadian and CLOCK-controlled regulation of the mouse transcriptome and cell proliferation. PNAS. 2007;104:3342–3347. doi: 10.1073/pnas.0611724104.
    1. Muoio DM, Newgard CB. Mechanisms of disease:Molecular and metabolic mechanisms of insulin resistance and beta-cell failure in type 2 diabetes. Nature Reviews Molecular Cell Biology. 2008;9:193–205. doi: 10.1038/nrm2327.
    1. Mure LS, Le HD, Benegiamo G, Chang MW, Rios L, Jillani N, Ngotho M, Kariuki T, Dkhissi-Benyahya O, Cooper HM, Panda S. Diurnal transcriptome atlas of a primate across major neural and peripheral tissues. Science. 2018;359:eaao0318. doi: 10.1126/science.aao0318.
    1. Neve B, Fernandez-Zapico ME, Ashkenazi-Katalan V, Dina C, Hamid YH, Joly E, Vaillant E, Benmezroua Y, Durand E, Bakaher N, Delannoy V, Vaxillaire M, Cook T, Dallinga-Thie GM, Jansen H, Charles MA, Clément K, Galan P, Hercberg S, Helbecque N, Charpentier G, Prentki M, Hansen T, Pedersen O, Urrutia R, Melloul D, Froguel P. Role of transcription factor KLF11 and its diabetes-associated gene variants in pancreatic beta cell function. PNAS. 2005;102:4807–4812. doi: 10.1073/pnas.0409177102.
    1. Pachkov M, Balwierz PJ, Arnold P, Ozonov E, van Nimwegen E. SwissRegulon, a database of genome-wide annotations of regulatory sites: recent updates. Nucleic Acids Research. 2013;41:D214–D220. doi: 10.1093/nar/gks1145.
    1. Peek CB, Levine DC, Cedernaes J, Taguchi A, Kobayashi Y, Tsai SJ, Bonar NA, McNulty MR, Ramsey KM, Bass J. Circadian clock interaction with HIF1α mediates oxygenic metabolism and anaerobic glycolysis in skeletal muscle. Cell Metabolism. 2017;25:86–92. doi: 10.1016/j.cmet.2016.09.010.
    1. Perrin L, Loizides-Mangold U, Skarupelova S, Pulimeno P, Chanon S, Robert M, Bouzakri K, Modoux C, Roux-Lombard P, Vidal H, Lefai E, Dibner C. Human skeletal myotubes display a cell-autonomous circadian clock implicated in basal myokine secretion. Molecular Metabolism. 2015;4:834–845. doi: 10.1016/j.molmet.2015.07.009.
    1. Petrenko V, Saini C, Perrin L, Dibner C. Parallel measurement of circadian clock gene expression and hormone secretion in human primary cell cultures. Journal of Visualized Experiments. 2016;117 doi: 10.3791/54673.
    1. Pulimeno P, Mannic T, Sage D, Giovannoni L, Salmon P, Lemeille S, Giry-Laterriere M, Unser M, Bosco D, Bauer C, Morf J, Halban P, Philippe J, Dibner C. Autonomous and self-sustained circadian oscillators displayed in human islet cells. Diabetologia. 2013;56:497–507. doi: 10.1007/s00125-012-2779-7.
    1. Pullen N, Dennis PB, Andjelkovic M, Dufner A, Kozma SC, Hemmings BA, Thomas G. Phosphorylation and activation of p70s6k by PDK1. Science. 1998;279:707–710.
    1. Quinlan AR, Hall IM. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics. 2010;26:841–842. doi: 10.1093/bioinformatics/btq033.
    1. Ramm G, Larance M, Guilhaus M, James DE. A role for 14-3-3 in insulin-stimulated GLUT4 translocation through its interaction with the RabGAP AS160. Journal of Biological Chemistry. 2006;281:29174–29180. doi: 10.1074/jbc.M603274200.
    1. Ramsey KM, Yoshino J, Brace CS, Abrassart D, Kobayashi Y, Marcheva B, Hong HK, Chong JL, Buhr ED, Lee C, Takahashi JS, Imai S, Bass J. Circadian clock feedback cycle through NAMPT-mediated NAD+ biosynthesis. Science. 2009;324:651–654. doi: 10.1126/science.1171641.
    1. Roach WG, Chavez JA, Mîinea CP, Lienhard GE. Substrate specificity and effect on GLUT4 translocation of the Rab GTPase-activating protein Tbc1d1. Biochemical Journal. 2007;403:353–358. doi: 10.1042/BJ20061798.
    1. Robinson MD, McCarthy DJ, Smyth GK. edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics. 2010;26:139–140. doi: 10.1093/bioinformatics/btp616.
    1. Robinson MD, Oshlack A. A scaling normalization method for differential expression analysis of RNA-seq data. Genome Biology. 2010;11:R25. doi: 10.1186/gb-2010-11-3-r25.
    1. Robinson MD, Smyth GK. Small-sample estimation of negative binomial dispersion, with applications to SAGE data. Biostatistics. 2008;9:321–332. doi: 10.1093/biostatistics/kxm030.
    1. Rose AJ, Jeppesen J, Kiens B, Richter EA. Effects of contraction on localization of GLUT4 and v-SNARE isoforms in rat skeletal muscle. American Journal of Physiology-Regulatory, Integrative and Comparative Physiology. 2009;297:R1228–R1237. doi: 10.1152/ajpregu.00258.2009.
    1. Saini C, Brown SA, Dibner C. Human peripheral clocks: applications for studying circadian phenotypes in physiology and pathophysiology. Frontiers in Neurology. 2015;6:95. doi: 10.3389/fneur.2015.00095.
    1. Saini C, Petrenko V, Pulimeno P, Giovannoni L, Berney T, Hebrok M, Howald C, Dermitzakis ET, Dibner C. A functional circadian clock is required for proper insulin secretion by human pancreatic islet cells. Diabetes, Obesity and Metabolism. 2016;18:355–365. doi: 10.1111/dom.12616.
    1. Sakamoto K, Holman GD. Emerging role for AS160/TBC1D4 and TBC1D1 in the regulation of GLUT4 traffic. American Journal of Physiology-Endocrinology and Metabolism. 2008;295:E29–E37. doi: 10.1152/ajpendo.90331.2008.
    1. Schiaffino S, Blaauw B, Dyar KA. The functional significance of the skeletal muscle clock: lessons from Bmal1 knockout models. Skeletal Muscle. 2016;6:33. doi: 10.1186/s13395-016-0107-5.
    1. Spoelstra K, Wikelski M, Daan S, Loudon AS, Hau M. Natural selection against a circadian clock gene mutation in mice. PNAS. 2016;113:686–691. doi: 10.1073/pnas.1516442113.
    1. Stenmark H, Parton RG, Steele-Mortimer O, Lütcke A, Gruenberg J, Zerial M. Inhibition of rab5 GTPase activity stimulates membrane fusion in endocytosis. The EMBO Journal. 1994;13:1287–1296.
    1. Szekeres F, Chadt A, Tom RZ, Deshmukh AS, Chibalin AV, Björnholm M, Al-Hasani H, Zierath JR. The Rab-GTPase-activating protein TBC1D1 regulates skeletal muscle glucose metabolism. American Journal of Physiology-Endocrinology and Metabolism. 2012;303:E524–E533. doi: 10.1152/ajpendo.00605.2011.
    1. Talbot J, Maves L. Skeletal muscle fiber type: using insights from muscle developmental biology to dissect targets for susceptibility and resistance to muscle disease. Wiley Interdisciplinary Reviews: Developmental Biology. 2016;5:518–534. doi: 10.1002/wdev.230.
    1. Tan Z, Zhou LJ, Mu PW, Liu SP, Chen SJ, Fu XD, Wang TH. Caveolin-3 is involved in the protection of resveratrol against high-fat-diet-induced insulin resistance by promoting GLUT4 translocation to the plasma membrane in skeletal muscle of ovariectomized rats. The Journal of Nutritional Biochemistry. 2012;23:1716–1724. doi: 10.1016/j.jnutbio.2011.12.003.
    1. Taylor MV, Hughes SM. Mef2 and the skeletal muscle differentiation program. Seminars in Cell & Developmental Biology. 2017;72:33–44. doi: 10.1016/j.semcdb.2017.11.020.
    1. Teng S, Stegner D, Chen Q, Hongu T, Hasegawa H, Chen L, Kanaho Y, Nieswandt B, Frohman MA, Huang P. Phospholipase D1 facilitates second-phase myoblast fusion and skeletal muscle regeneration. Molecular Biology of the Cell. 2015;26:506–517. doi: 10.1091/mbc.E14-03-0802.
    1. Thomson DM, Winder WW. AMP-activated protein kinase control of fat metabolism in skeletal muscle. Acta Physiologica. 2009;196:147–154. doi: 10.1111/j.1748-1716.2009.01973.x.
    1. Van Thienen R, D'Hulst G, Deldicque L, Hespel P. Biochemical artifacts in experiments involving repeated biopsies in the same muscle. Physiological Reports. 2014;2:e00286. doi: 10.14814/phy2.286.
    1. Volchuk A, Sargeant R, Sumitani S, Liu Z, He L, Klip A. Cellubrevin is a resident protein of insulin-sensitive GLUT4 glucose transporter vesicles in 3T3-L1 adipocytes. Journal of Biological Chemistry. 1995;270:8233–8240. doi: 10.1074/jbc.270.14.8233.
    1. Wang J, Symul L, Yeung J, Gobet C, Sobel J, Lück S, Westermark PO, Molina N, Naef F. Circadian clock-dependent and -independent posttranscriptional regulation underlies temporal mRNA accumulation in mouse liver. PNAS. 2018;115:E1916–E1925. doi: 10.1073/pnas.1715225115.
    1. Wehrens SMT, Christou S, Isherwood C, Middleton B, Gibbs MA, Archer SN, Skene DJ, Johnston JD. Meal timing regulates the human circadian system. Current Biology. 2017;27:1768–1775. doi: 10.1016/j.cub.2017.04.059.
    1. Yoshitane H, Ozaki H, Terajima H, Du NH, Suzuki Y, Fujimori T, Kosaka N, Shimba S, Sugano S, Takagi T, Iwasaki W, Fukada Y. CLOCK-controlled polyphonic regulation of circadian rhythms through canonical and noncanonical E-boxes. Molecular and Cellular Biology. 2014;34:1776–1787. doi: 10.1128/MCB.01465-13.
    1. Zeileis A, Leisch F, Hornik K, Kleiber C. strucchange: An R Package for Testing for Structural Change in Linear Regression Models. Journal of Statistical Software. 2002;7 doi: 10.18637/jss.v007.i02.
    1. Zhang R, Lahens NF, Ballance HI, Hughes ME, Hogenesch JB. A circadian gene expression atlas in mammals: implications for biology and medicine. PNAS. 2014;111:16219–16224. doi: 10.1073/pnas.1408886111.

Source: PubMed

3
購読する