Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

NAD+ cellular redox and SIRT1 regulate the diurnal rhythms of tyrosine hydroxylase and conditioned cocaine reward

Abstract

The diurnal regulation of dopamine is important for normal physiology and diseases such as addiction. Here we find a novel role for the CLOCK protein to antagonize CREB-mediated transcriptional activity at the tyrosine hydroxylase (TH) promoter, which is mediated by the interaction with the metabolic sensing protein, Sirtuin 1 (SIRT1). Additionally, we demonstrate that the transcriptional activity of TH is modulated by the cellular redox state, and daily rhythms of redox balance in the ventral tegmental area (VTA), along with TH transcription, are highly disrupted following chronic cocaine administration. Furthermore, CLOCK and SIRT1 are important for regulating cocaine reward and dopaminergic (DAergic) activity, with interesting differences depending on whether DAergic activity is in a heightened state and if there is a functional CLOCK protein. Taken together, we find that rhythms in cellular metabolism and circadian proteins work together to regulate dopamine synthesis and the reward value for drugs of abuse.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

References

  1. Cribbet MR, Logan RW, Edwards MD, Hanlon E, Bien Peek C, Stubblefield JJ, et al. Circadian rhythms and metabolism: from the brain to the gut and back again. Ann N Y Acad Sci. 2016;1385:21–40.

    Article  PubMed  PubMed Central  Google Scholar 

  2. Eckel-Mahan K, Sassone-Corsi P. Metabolism and the circadian clock converge. Physiol Rev. 2013;93:107–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Vengeliene V, Cannella N, Takahashi T, Spanagel R. Metabolic shift of the kynurenine pathway impairs alcohol and cocaine seeking and relapse. Psychopharmacology. 2016;233:3449–59.

    Article  CAS  PubMed  Google Scholar 

  4. Barandas R, Landgraf D, McCarthy MJ, Welsh DK. Circadian clocks as modulators of metabolic comorbidity in psychiatric disorders. Curr Psychiatry Rep. 2015;17:98.

    Article  PubMed  Google Scholar 

  5. Kaidanovich-Beilin O, Cha DS, McIntyre RS. Crosstalk between metabolic and neuropsychiatric disorders. F1000 Biol Rep. 2012;4:14.

    Article  PubMed  PubMed Central  Google Scholar 

  6. Karatsoreos IN. Links between circadian rhythms and psychiatric disease. Front Behav Neurosci. 2014;8:162.

    Article  PubMed  PubMed Central  Google Scholar 

  7. Takahashi JS, Hong HK, Ko CH, McDearmon EL. The genetics of mammalian circadian order and disorder: implications for physiology and disease. Nat Rev Genet. 2008;9:764–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Abarca C, Albrecht U, Spanagel R. Cocaine sensitization and reward are under the influence of circadian genes and rhythm. Proc Natl Acad Sci USA. 2002;99:9026–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Andretic R, Chaney S, Hirsh J. Requirement of circadian genes for cocaine sensitization in Drosophila. Science. 1999;285:1066–8.

    Article  CAS  PubMed  Google Scholar 

  10. Jansen HT, Sergeeva A, Stark G, Sorg BA. Circadian discrimination of reward: evidence for simultaneous yet separable food- and drug-entrained rhythms in the rat. Chronobiol Int. 2012;29:454–68.

    Article  CAS  PubMed  Google Scholar 

  11. Sleipness EP, Sorg BA, Jansen HT. Diurnal differences in dopamine transporter and tyrosine hydroxylase levels in rat brain: dependence on the suprachiasmatic nucleus. Brain Res. 2007;1129:34–42.

    Article  CAS  PubMed  Google Scholar 

  12. Halbout B, Perreau-Lenz S, Dixon CI, Stephens DN, Spanagel R. Per1(Brdm1) mice self-administer cocaine and reinstate cocaine-seeking behaviour following extinction. Behav Pharmacol. 2011;22:76–80.

    Article  CAS  PubMed  Google Scholar 

  13. Ozburn AR, Falcon E, Twaddle A, Nugent AL, Gillman AG, Spencer SM, et al. Direct regulation of diurnal Drd3 expression and cocaine reward by NPAS2. Biol Psychiatry. 2015;77:425–33.

    Article  CAS  PubMed  Google Scholar 

  14. Ozburn AR, Larson EB, Self DW, McClung CA. Cocaine self-administration behaviors in ClockDelta19 mice. Psychopharmacology. 2012;223:169–77.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Chung S, Lee EJ, Yun S, Choe HK, Park SB, Son HJ, et al. Impact of circadian nuclear receptor REV-ERBalpha on midbrain dopamine production and mood regulation. Cell. 2014;157:858–68.

    Article  CAS  PubMed  Google Scholar 

  16. Ferris MJ, Espana RA, Locke JL, Konstantopoulos JK, Rose JH, Chen R, et al. Dopamine transporters govern diurnal variation in extracellular dopamine tone. Proc Natl Acad Sci USA. 2014;111:E2751–2759.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Shumay E, Fowler JS, Wang GJ, Logan J, Alia-Klein N, Goldstein RZ, et al. Repeat variation in the human PER2 gene as a new genetic marker associated with cocaine addiction and brain dopamine D2 receptor availability. Transl Psychiatry. 2012;2:e86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Sidor MM, Spencer SM, Dzirasa K, Parekh PK, Tye KM, Warden MR, et al. Daytime spikes in dopaminergic activity drive rapid mood-cycling in mice. Mol Psychiatry. 2015;20:1406–19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Sorg BA, Chen SY, Kalivas PW. Time course of tyrosine hydroxylase expression after behavioral sensitization to cocaine. J Pharmacol Exp Ther. 1993;266:424–30.

    CAS  PubMed  Google Scholar 

  20. Spanagel R, Pendyala G, Abarca C, Zghoul T, Sanchis-Segura C, Magnone MC, et al. The clock gene Per2 influences the glutamatergic system and modulates alcohol consumption. Nat Med. 2005;11:35–42.

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  22. Zhang R, Lahens NF, Ballance HI, Hughes ME, Hogenesch JB. A circadian gene expression atlas in mammals: implications for biology and medicine. Proc Natl Acad Sci USA. 2014;111:16219–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Asher G, Gatfield D, Stratmann M, Reinke H, Dibner C, Kreppel F, et al. SIRT1 regulates circadian clock gene expression through PER2 deacetylation. Cell. 2008;134:317–28.

    Article  CAS  PubMed  Google Scholar 

  24. Nakahata Y, Kaluzova M, Grimaldi B, Sahar S, Hirayama J, Chen D, et al. The NAD+-dependent deacetylase SIRT1 modulates CLOCK-mediated chromatin remodeling and circadian control. Cell. 2008;134:329–40.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Nakahata Y, Sahar S, Astarita G, Kaluzova M, Sassone-Corsi P. Circadian control of the NAD+salvage pathway by CLOCK-SIRT1. Science. 2009;324:654–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Ramsey KM, Yoshino J, Brace CS, Abrassart D, Kobayashi Y, Marcheva B, et al. Circadian clock feedback cycle through NAMPT-mediated NAD+biosynthesis. Science. 2009;324:651–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Houtkooper RH, Canto C, Wanders RJ, Auwerx J. The secret life of NAD+: an old metabolite controlling new metabolic signaling pathways. Endocr Rev. 2010;31:194–223.

    Article  CAS  PubMed  Google Scholar 

  28. Hirrlinger J, Dringen R. The cytosolic redox state of astrocytes: maintenance, regulation and functional implications for metabolite trafficking. Brain Res Rev. 2010;63:177–88.

    Article  CAS  PubMed  Google Scholar 

  29. Ying W. NAD+ and NADH in brain functions, brain diseases and brain aging. Front Biosci. 2007;12:1863–88.

    Article  CAS  PubMed  Google Scholar 

  30. Henley J, Poo MM. Guiding neuronal growth cones using Ca2+ signals. Trends Cell Biol. 2004;14:320–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Konur S, Ghosh A. Calcium signaling and the control of dendritic development. Neuron. 2005;46:401–5.

    Article  CAS  PubMed  Google Scholar 

  32. Liang M, Yin XL, Wang LY, Yin WH, Song NY, Shi HB, et al. NAD+ attenuates bilirubin-induced hyperexcitation in the ventral cochlear nucleus by inhibiting excitatory neurotransmission and neuronal excitability. Front Cell Neurosci. 2017;11:21.

    PubMed  PubMed Central  Google Scholar 

  33. Tamsett TJ, Picchione KE, Bhattacharjee A. NAD+ activates KNa channels in dorsal root ganglion neurons. J Neurosci. 2009;29:5127–34.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Ferguson D, Koo JW, Feng J, Heller E, Rabkin J, Heshmati M, et al. Essential role of SIRT1 signaling in the nucleus accumbens in cocaine and morphine action. J Neurosci. 2013;33:16088–98.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Ferguson D, Shao N, Heller E, Feng J, Neve R, Kim HD, et al. SIRT1-FOXO3a regulate cocaine actions in the nucleus accumbens. J Neurosci. 2015;35:3100–11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Kim HD, Hesterman J, Call T, Magazu S, Keeley E, Armenta K, et al. SIRT1 mediates depression-like behaviors in the nucleus accumbens. J Neurosci. 2016;36:8441–52.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Libert S, Pointer K, Bell EL, Das A, Cohen DE, Asara JM, et al. SIRT1 activates MAO-A in the brain to mediate anxiety and exploratory drive. Cell. 2011;147:1459–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Renthal W, Kumar A, Xiao G, Wilkinson M, Covington HE 3rd, Maze I, et al. Genome-wide analysis of chromatin regulation by cocaine reveals a role for sirtuins. Neuron. 2009;62:335–48.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Aksenov MY, Aksenova MV, Nath A, Ray PD, Mactutus CF, Booze RM. Cocaine-mediated enhancement of Tat toxicity in rat hippocampal cell cultures: the role of oxidative stress and D1 dopamine receptor. Neurotoxicology. 2006;27:217–28.

    Article  CAS  PubMed  Google Scholar 

  40. Lopez-Pedrajas R, Ramirez-Lamelas DT, Muriach B, Sanchez-Villarejo MV, Almansa I, Vidal-Gil L, et al. Cocaine promotes oxidative stress and microglial-macrophage activation in rat cerebellum. Front Cell Neurosci. 2015;9:279.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Muriach M, Lopez-Pedrajas R, Barcia JM, Sanchez-Villarejo MV, Almansa I, Romero FJ. Cocaine causes memory and learning impairments in rats: involvement of nuclear factor kappa B and oxidative stress, and prevention by topiramate. J Neurochem. 2010;114:675–84.

    Article  CAS  PubMed  Google Scholar 

  42. Pomierny-Chamiolo L, Moniczewski A, Wydra K, Suder A, Filip M. Oxidative stress biomarkers in some rat brain structures and peripheral organs underwent cocaine. Neurotox Res. 2013;23:92–102.

    Article  CAS  PubMed  Google Scholar 

  43. Requardt RP, Wilhelm F, Rillich J, Winkler U, Hirrlinger J. The biphasic NAD(P)H fluorescence response of astrocytes to dopamine reflects the metabolic actions of oxidative phosphorylation and glycolysis. J Neurochem. 2010;115:483–92.

    Article  CAS  PubMed  Google Scholar 

  44. Requardt RP, Hirrlinger PG, Wilhelm F, Winkler U, Besser S, Hirrlinger J. Ca(2)(+) signals of astrocytes are modulated by the NAD(+)/NADH redox state. J Neurochem. 2012;120:1014–25.

    CAS  PubMed  Google Scholar 

  45. Sordi AO, Pechansky F, Kessler FH, Kapczinski F, Pfaffenseller B, Gubert C, et al. Oxidative stress and BDNF as possible markers for the severity of crack cocaine use in early withdrawal. Psychopharmacology. 2014;231:4031–9.

    Article  CAS  PubMed  Google Scholar 

  46. Uys JD, Knackstedt L, Hurt P, Tew KD, Manevich Y, Hutchens S, et al. Cocaine-induced adaptations in cellular redox balance contributes to enduring behavioral plasticity. Neuropsychopharmacology. 2011;36:2551–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Berke JD, Hyman SE. Addiction, dopamine, and the molecular mechanisms of memory. Neuron. 2000;25:515–32.

    Article  CAS  PubMed  Google Scholar 

  48. Nutt DJ, Lingford-Hughes A, Erritzoe D, Stokes PR. The dopamine theory of addiction: 40 years of highs and lows. Nat Rev Neurosci. 2015;16:305–12.

    Article  CAS  PubMed  Google Scholar 

  49. Volkow ND. Imaging the addicted brain: from molecules to behavior. J Nucl Med. 2004;45:13N–16N. 19N-20N, 22N passim

    PubMed  Google Scholar 

  50. Volkow ND, Morales M. The brain on drugs: from reward to addiction. Cell. 2015;162:712–25.

    Article  CAS  PubMed  Google Scholar 

  51. Margolis EB, Lock H, Hjelmstad GO, Fields HL. The ventral tegmental area revisited: is there an electrophysiological marker for dopaminergic neurons? J Physiol. 2006;577(Pt 3):907–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Nair-Roberts RG, Chatelain-Badie SD, Benson E, White-Cooper H, Bolam JP, Ungless MA. Stereological estimates of dopaminergic, GABAergic and glutamatergic neurons in the ventral tegmental area, substantia nigra and retrorubral field in the rat. Neuroscience. 2008;152:1024–31.

    Article  CAS  PubMed  Google Scholar 

  53. Swanson LW. The projections of the ventral tegmental area and adjacent regions: a combined fluorescent retrograde tracer and immunofluorescence study in the rat. Brain Res Bull. 1982;9:321–53.

    Article  CAS  PubMed  Google Scholar 

  54. Kabanova A, Pabst M, Lorkowski M, Braganza O, Boehlen A, Nikbakht N, et al. Function and developmental origin of a mesocortical inhibitory circuit. Nat Neurosci. 2015;18:872–82.

    Article  CAS  PubMed  Google Scholar 

  55. Morales M, Margolis EB. Ventral tegmental area: cellular heterogeneity, connectivity and behaviour. Nat Rev Neurosci. 2017;18:73–85.

    Article  CAS  PubMed  Google Scholar 

  56. Zhang JT, Ma SS, Yip SW, Wang LJ, Chen C, Yan CG, et al. Decreased functional connectivity between ventral tegmental area and nucleus accumbens in Internet gaming disorder: evidence from resting state functional magnetic resonance imaging. Behav Brain Funct. 2015;11:37.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Sleipness EP, Jansen HT, Schenk JO, Sorg BA. Time-of-day differences in dopamine clearance in the rat medial prefrontal cortex and nucleus accumbens. Synapse. 2008;62:877–85.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Sleipness EP, Sorg BA, Jansen HT. Time of day alters long-term sensitization to cocaine in rats. Brain Res. 2005;1065:132–7.

    Article  CAS  PubMed  Google Scholar 

  59. Luscher C, Malenka RC. Drug-evoked synaptic plasticity in addiction: from molecular changes to circuit remodeling. Neuron. 2011;69:650–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Nestler EJ, Carlezon WA Jr.. The mesolimbic dopamine reward circuit in depression. Biol Psychiatry. 2006;59:1151–9.

    Article  CAS  PubMed  Google Scholar 

  61. Logan RW, McClung CA. Animal models of bipolar mania: the past, present and future. Neuroscience. 2016;321:163–88.

    Article  CAS  PubMed  Google Scholar 

  62. Biguet NF, Buda M, Lamouroux A, Samolyk D, Mallet J. Time course of the changes of TH mRNA in rat brain and adrenal medulla after a single injection of reserpine. EMBO J. 1986;5:287–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Fossom LH, Carlson CD, Tank AW. Stimulation of tyrosine hydroxylase gene transcription rate by nicotine in rat adrenal medulla. Mol Pharmacol. 1991;40:193–202.

    CAS  PubMed  Google Scholar 

  64. Tank AW, Lewis EJ, Chikaraishi DM, Weiner N. Elevation of RNA coding for tyrosine hydroxylase in rat adrenal gland by reserpine treatment and exposure to cold. J Neurochem. 1985;45:1030–3.

    Article  CAS  PubMed  Google Scholar 

  65. Cambi F, Fung B, Chikaraishi D. 5’ flanking DNA sequences direct cell-specific expression of rat tyrosine hydroxylase. J Neurochem. 1989;53:1656–9.

    Article  CAS  PubMed  Google Scholar 

  66. Iwata N, Kobayashi K, Sasaoka T, Hidaka H, Nagatsu T. Structure of the mouse tyrosine hydroxylase gene. Biochem Biophys Res Commun. 1992;182:348–54.

    Article  CAS  PubMed  Google Scholar 

  67. Kobayashi K, Kaneda N, Ichinose H, Kishi F, Nakazawa A, Kurosawa Y, et al. Structure of the human tyrosine hydroxylase gene: alternative splicing from a single gene accounts for generation of four mRNA types. J Biochem. 1988;103:907–12.

    Article  CAS  PubMed  Google Scholar 

  68. Gekakis N, Staknis D, Nguyen HB, Davis FC, Wilsbacher LD, King DP, et al. Role of the CLOCK protein in the mammalian circadian mechanism. Science. 1998;280:1564–9.

    Article  CAS  PubMed  Google Scholar 

  69. Kumaki Y, Ukai-Tadenuma M, Uno KD, Nishio J, Masumoto KH, Nagano M, et al. Analysis and synthesis of high-amplitude Cis-elements in the mammalian circadian clock. Proc Natl Acad Sci USA. 2008;105:14946–51.

    Article  PubMed  PubMed Central  Google Scholar 

  70. Yoo SH, Ko CH, Lowrey PL, Buhr ED, Song EJ, Chang S, et al. A noncanonical E-box enhancer drives mouse Period2 circadian oscillations in vivo. Proc Natl Acad Sci USA. 2005;102:2608–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Yoshitane H, Ozaki H, Terajima H, Du NH, Suzuki Y, Fujimori T, et al. CLOCK-controlled polyphonic regulation of circadian rhythms through canonical and noncanonical E-boxes. Mol Cell Biol. 2014;34:1776–87.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Lazaroff M, Patankar S, Yoon SO, Chikaraishi DM. The cyclic AMP response element directs tyrosine hydroxylase expression in catecholaminergic central and peripheral nervous system cell lines from transgenic mice. J Biol Chem. 1995;270:21579–89.

    Article  CAS  PubMed  Google Scholar 

  73. Olson VG, Zabetian CP, Bolanos CA, Edwards S, Barrot M, Eisch AJ, et al. Regulation of drug reward by cAMP response element-binding protein: evidence for two functionally distinct subregions of the ventral tegmental area. J Neurosci. 2005;25:5553–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Roybal K, Theobold D, Graham A, DiNieri JA, Russo SJ, Krishnan V, et al. Mania-like behavior induced by disruption of CLOCK. Proc Natl Acad Sci USA. 2007;104:6406–11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Ozburn AR, Falcon E, Mukherjee S, Gillman A, Arey R, Spencer S, et al. The role of clock in ethanol-related behaviors. Neuropsychopharmacology. 2013;38:2393–2400.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. McClung CA, Sidiropoulou K, Vitaterna M, Takahashi JS, White FJ, Cooper DC, et al. Regulation of dopaminergic transmission and cocaine reward by the Clock gene. Proc Natl Acad Sci USA. 2005;102:9377–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Parekh PK, Becker-Krail D, Sundaravelu P, Ishigaki S, Okado H, Sobue G, et al. Altered GluA1 (Gria1) function and accumbal synaptic plasticity in the ClockDelta19 model of bipolar mania. Biol Psychiatry. 2017. [Epub ahead of print].

  78. Baur JA, Sinclair DA. Therapeutic potential of resveratrol: the in vivo evidence. Nat Rev Drug Discov. 2006;5:493–506.

    Article  CAS  PubMed  Google Scholar 

  79. Spencer S, Falcon E, Kumar J, Krishnan V, Mukherjee S, Birnbaum SG, et al. Circadian genes period 1 and period 2 in the nucleus accumbens regulate anxiety-related behavior. Eur J Neurosci. 2013;37:242–50.

    Article  PubMed  Google Scholar 

  80. Stromsdorfer KL, Yamaguchi S, Yoon MJ, Moseley AC, Franczyk MP, Kelly SC, et al. NAMPT-mediated NAD(+) biosynthesis in adipocytes regulates adipose tissue function and multi-organ insulin sensitivity in mice. Cell Rep. 2016;16:1851–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Yoshino J, Imai S. Accurate measurement of nicotinamide adenine dinucleotide (NAD(+)) with high-performance liquid chromatography. Methods Mol Biol. 2013;1077:203–15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Coque L, Mukherjee S, Cao JL, Spencer S, Marvin M, Falcon E, et al. Specific role of VTA dopamine neuronal firing rates and morphology in the reversal of anxiety-related, but not depression-related behavior in the ClockDelta19 mouse model of mania. Neuropsychopharmacology. 2011;36:1478–88.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. King DP, Vitaterna MH, Chang AM, Dove WF, Pinto LH, Turek FW, et al. The mouse clock mutation behaves as an antimorph and maps within the W19H deletion, distal of Kit. Genetics. 1997;146:1049–60.

    CAS  PubMed  PubMed Central  Google Scholar 

  84. King DP, Zhao Y, Sangoram AM, Wilsbacher LD, Tanaka M, Antoch MP, et al. Positional cloning of the mouse circadian clock gene. Cell. 1997;89:641–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Lewis-Tuffin LJ, Quinn PG, Chikaraishi DM. Tyrosine hydroxylase transcription depends primarily on cAMP response element activity, regardless of the type of inducing stimulus. Mol Cell Neurosci. 2004;25:536–47.

    Article  CAS  PubMed  Google Scholar 

  86. Piech-Dumas KM, Tank AW. CREB mediates the cAMP-responsiveness of the tyrosine hydroxylase gene: use of an antisense RNA strategy to produce CREB-deficient PC12 cell lines. Brain Res Mol Brain Res. 1999;70:219–30.

    Article  CAS  PubMed  Google Scholar 

  87. Aguilar-Arnal L, Katada S, Orozco-Solis R, Sassone-Corsi P. NAD(+)-SIRT1 control of H3K4 trimethylation through circadian deacetylation of MLL1. Nat Struct Mol Biol. 2015;22:312–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Canto C, Gerhart-Hines Z, Feige JN, Lagouge M, Noriega L, Milne JC, et al. AMPK regulates energy expenditure by modulating NAD+ metabolism and SIRT1 activity. Nature. 2009;458:1056–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Bitterman KJ, Anderson RM, Cohen HY, Latorre-Esteves M, Sinclair DA. Inhibition of silencing and accelerated aging by nicotinamide, a putative negative regulator of yeast sir2 and human SIRT1. J Biol Chem. 2002;277:45099–107.

    Article  CAS  PubMed  Google Scholar 

  90. Berger F, Lau C, Dahlmann M, Ziegler M. Subcellular compartmentation and differential catalytic properties of the three human nicotinamide mononucleotide adenylyltransferase isoforms. J Biol Chem. 2005;280:36334–41.

    Article  CAS  PubMed  Google Scholar 

  91. Koch-Nolte F, Fischer S, Haag F, Ziegler M. Compartmentation of NAD+-dependent signalling. FEBS Lett. 2011;585:1651–6.

    Article  CAS  PubMed  Google Scholar 

  92. Hasmann M, Schemainda I. FK866, a highly specific noncompetitive inhibitor of nicotinamide phosphoribosyltransferase, represents a novel mechanism for induction of tumor cell apoptosis. Cancer Res. 2003;63:7436–42.

    CAS  PubMed  Google Scholar 

  93. Hu Y, Liu J, Wang J, Liu Q. The controversial links among calorie restriction, SIRT1, and resveratrol. Free Radic Biol Med. 2011;51:250–6.

    Article  CAS  PubMed  Google Scholar 

  94. Dietrich JB, Mangeol A, Revel MO, Burgun C, Aunis D, Zwiller J. Acute or repeated cocaine administration generates reactive oxygen species and induces antioxidant enzyme activity in dopaminergic rat brain structures. Neuropharmacology. 2005;48:965–74.

    Article  CAS  PubMed  Google Scholar 

  95. Gimenez-Xavier P, Gomez-Santos C, Castano E, Francisco R, Boada J, Unzeta M, et al. The decrease of NAD(P)H has a prominent role in dopamine toxicity. Biochim Biophys Acta. 2006;1762:564–74.

    Article  CAS  PubMed  Google Scholar 

  96. Macedo DS, de Vasconcelos SM, dos Santos RS, Aguiar LM, Lima VT, Viana GS, et al. Cocaine alters catalase activity in prefrontal cortex and striatum of mice. Neurosci Lett. 2005;387:53–56.

    Article  CAS  PubMed  Google Scholar 

  97. Numa R, Kohen R, Poltyrev T, Yaka R. Tempol diminishes cocaine-induced oxidative damage and attenuates the development and expression of behavioral sensitization. Neuroscience. 2008;155:649–58.

    Article  CAS  PubMed  Google Scholar 

  98. Bellet MM, Orozco-Solis R, Sahar S, Eckel-Mahan K, Sassone-Corsi P. The time of metabolism: NAD+, SIRT1, and the circadian clock. Cold Spring Harb Symp Quant Biol. 2011;76:31–38.

    Article  CAS  PubMed  Google Scholar 

  99. Aguilar-Arnal L, Ranjit S, Stringari C, Orozco-Solis R, Gratton E, Sassone-Corsi P. Spatial dynamics of SIRT1 and the subnuclear distribution of NADH species. Proc Natl Acad Sci USA. 2016;45:12715–20.

  100. Fulco M, Schiltz RL, Iezzi S, King MT, Zhao P, Kashiwaya Y, et al. Sir2 regulates skeletal muscle differentiation as a potential sensor of the redox state. Mol Cell. 2003;12:51–62.

    Article  CAS  PubMed  Google Scholar 

  101. Prozorovski T, Schulze-Topphoff U, Glumm R, Baumgart J, Schroter F, Ninnemann O, et al. Sirt1 contributes critically to the redox-dependent fate of neural progenitors. Nat Cell Biol. 2008;10:385–94.

    Article  CAS  PubMed  Google Scholar 

  102. Revollo JR, Grimm AA, Imai S. The NAD biosynthesis pathway mediated by nicotinamide phosphoribosyltransferase regulates Sir2 activity in mammalian cells. J Biol Chem. 2004;279:50754–63.

    Article  CAS  PubMed  Google Scholar 

  103. Obrietan K, Impey S, Smith D, Athos J, Storm DR. Circadian regulation of cAMP response element-mediated gene expression in the suprachiasmatic nuclei. J Biol Chem. 1999;274:17748–56.

    Article  CAS  PubMed  Google Scholar 

  104. Travnickova-Bendova Z, Cermakian N, Reppert SM, Sassone-Corsi P. Bimodal regulation of mPeriod promoters by CREB-dependent signaling and CLOCK/BMAL1 activity. Proc Natl Acad Sci USA. 2002;99:7728–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Chang HC, Guarente L. SIRT1 mediates central circadian control in the SCN by a mechanism that decays with aging. Cell. 2013;153:1448–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Hughes ME, Hong HK, Chong JL, Indacochea AA, Lee SS, Han M, et al. Brain-specific rescue of Clock reveals system-driven transcriptional rhythms in peripheral tissue. PLoS Genet. 2012;8:e1002835.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. DeBruyne JP, Weaver DR, Reppert SM. CLOCK and NPAS2 have overlapping roles in the suprachiasmatic circadian clock. Nat Neurosci. 2007;10:543–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Gu YZ, Hogenesch JB, Bradfield CA. The PAS superfamily: sensors of environmental and developmental signals. Annu Rev Pharmacol Toxicol. 2000;40:519–61.

    Article  CAS  PubMed  Google Scholar 

  109. Hogenesch JB, Chan WK, Jackiw VH, Brown RC, Gu YZ, Pray-Grant M, et al. Characterization of a subset of the basic-helix-loop-helix-PAS superfamily that interacts with components of the dioxin signaling pathway. J Biol Chem. 1997;272:8581–93.

    Article  CAS  PubMed  Google Scholar 

  110. Reick M, Garcia JA, Dudley C, McKnight SL. NPAS2: an analog of clock operative in the mammalian forebrain. Science. 2001;293:506–9.

    Article  CAS  PubMed  Google Scholar 

  111. Landgraf D, Wang LL, Diemer T, Welsh DK. NPAS2 compensates for loss of clock in peripheral circadian oscillators. PLoS Genet. 2016;12:e1005882.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Kennaway DJ, Voultsios A, Varcoe TJ, Moyer RW. Melatonin and activity rhythm responses to light pulses in mice with the Clock mutation. Am J Physiol Regul Integr Comp Physiol. 2003;284:R1231–1240.

    Article  CAS  PubMed  Google Scholar 

  113. Ochi M, Sono S, Sei H, Oishi K, Kobayashi H, Morita Y, et al. Sex difference in circadian period of body temperature in Clock mutant mice with Jcl/ICR background. Neurosci Lett. 2003;347:163–6.

    Article  CAS  PubMed  Google Scholar 

  114. Vitaterna MH, King DP, Chang AM, Kornhauser JM, Lowrey PL, McDonald JD, et al. Mutagenesis and mapping of a mouse gene, Clock, essential for circadian behavior. Science. 1994;264:719–25.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Ripperger JA, Schibler U. Rhythmic CLOCK-BMAL1 binding to multiple E-box motifs drives circadian Dbp transcription and chromatin transitions. Nat Genet. 2006;38:369–74.

    Article  CAS  PubMed  Google Scholar 

  116. Vitaterna MH, Ko CH, Chang AM, Buhr ED, Fruechte EM, Schook A, et al. The mouse Clock mutation reduces circadian pacemaker amplitude and enhances efficacy of resetting stimuli and phase-response curve amplitude. Proc Natl Acad Sci USA. 2006;103:9327–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Redlin U, Hattar S, Mrosovsky N. The circadian Clock mutant mouse: impaired masking response to light. J Comp Physiol A Neuroethol Sens Neural Behav Physiol. 2005;191:51–59.

    Article  PubMed  Google Scholar 

  118. Garcia JA, Zhang D, Estill SJ, Michnoff C, Rutter J, Reick M, et al. Impaired cued and contextual memory in NPAS2-deficient mice. Science. 2000;288:2226–30.

    Article  CAS  PubMed  Google Scholar 

  119. Zhang Y, Xue Y, Meng S, Luo Y, Liang J, Li J, et al. Inhibition of lactate transport erases drug memory and prevents drug relapse. Biol Psychiatry. 2016;79:928–39.

    Article  CAS  PubMed  Google Scholar 

  120. Abe N, Uchida S, Otsuki K, Hobara T, Yamagata H, Higuchi F, et al. Altered sirtuin deacetylase gene expression in patients with a mood disorder. J Psychiatr Res. 2011;45:1106–12.

    Article  PubMed  Google Scholar 

  121. Kishi T, Fukuo Y, Kitajima T, Okochi T, Yamanouchi Y, Kinoshita Y, et al. SIRT1 gene, schizophrenia and bipolar disorder in the Japanese population: an association study. Genes Brain Behav. 2011;10:257–63.

    Article  CAS  PubMed  Google Scholar 

  122. Kishi T, Yoshimura R, Fukuo Y, Kitajima T, Okochi T, Matsunaga S, et al. The CLOCK gene and mood disorders: a case-control study and meta-analysis. Chronobiol Int. 2011;28:825–33.

    Article  CAS  PubMed  Google Scholar 

  123. Nivoli A, Porcelli S, Albani D, Forloni G, Fusco F, Colom F, et al. Association between sirtuin 1 gene rs10997870 polymorphism and suicide behaviors in bipolar disorder. Neuropsychobiology. 2016;74:1–7.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank Mark Brown and Mariah Hildebrand for animal care and genotyping. We also thank Eric Nestler, Deveroux Ferguson, and Rachel Neve for the mCREB and SIRT1 viruses, and Gary Thomas for the SIRT1 partial constructs. We thank Joe Takahashi for the ClockΔ19 mice and CLOCK expression constructs. This work was funded by DA039865, DA037636, MH106460, MH082876, DA023988, DA042886, NS058339, NARSAD, and IMHRO to CAM, and DA038564, DA041872, and NARSAD to RWL.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Colleen A. McClung.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Electronic supplementary material

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Logan, R.W., Parekh, P.K., Kaplan, G.N. et al. NAD+ cellular redox and SIRT1 regulate the diurnal rhythms of tyrosine hydroxylase and conditioned cocaine reward. Mol Psychiatry 24, 1668–1684 (2019). https://doi.org/10.1038/s41380-018-0061-1

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41380-018-0061-1

This article is cited by

Search

Quick links