• Aucun résultat trouvé

Neuronal circuit involved in inhibitory control

3) How does the potentiation of OFC-striatum impact on the neuronal activity in the striatum?

4.5 Neuronal circuit involved in inhibitory control

Clinical studies show that addicted patients suffer from a lack of inhibitory control (Feil et al., 2010). In the Go/NoGo task, anterior cingulate cortex (ACC) is active during both go and no-go trials, while ventrolateral and dorsolateral prefrontal cortex (VLPFC and DLPFC, respectively) are more active during the no-go trials (Liddle et al., 2001). Similar results were obtained with the color-word Stroop task (Kerns et al., 2004; MacDonald et al., 2000).

These results suggest VLPFC and DLPFC are particularly involved in response inhibition.

Cocaine addicted patients display more errors than healthy subjects in the both Go/NoGo (Kaufman et al., 2003) and Stroop task (Bolla et al., 2004). This worse performance has been attributed to the hypoactivity in the ACC or DFPFC (Bolla et al., 2004; Kaufman et al., 2003). The causal relation between the loss of inhibitory control and hypoactivity in the ACC or DLPFC has not been tested in human studies. However, in a preclinical study, silencing of the ACC, PL or IL diminished the inhibition of reward pursuit under threat of punishment (Verharen et al., 2019), suggesting that these brain regions are crucial for the inhibitory control. Prolonged cocaine self-administration induced massive reduction of firing rate in the PL, which is more profound in persevering animals (Chen et al., 2013).

The optogenetic inhibition of the PL facilitats compulsive cocaine seeking, indicating that the hypoactivity of the PL is casually related to the compulsive behavior (Chen et al., 2013).

It has yet to be determined how cocaine self-administration induces the reduction of the firing rate in the PL. The inhibition of the PL pyramidal neurons projecting to the NAc promotes reward seeking associated with punishments (Domingo-rodriguez et al., 2020;

Kim et al., 2017) but not the VTA projection (Kim et al., 2017), suggesting that the PL to NAc pathway is particularly important for the inhibitory control. However, it has been unclear how neuronal plasticity control the inhibitory control. Clinical studies show that addicted patients have lower D2 receptor availability in the striatum (Volkow et al., 2009).

Although we have not measured the expression of D2 receptor, this needs to be measured in the laboratory animals. The role of D2 receptors in the addiction has been investigated for decades but it’s still unclear. Pharmacological experiments shows that D2 receptors have higher affinity than D1 (Marcellino et al., 2012). Because of its high affinity, D2 receptors are tonically active at the basal state. Aversive events, such as foot shock

76

triggers the dip of dopamine (de Jong et al., 2018), and the dip of dopamine triggers the potentiation in D2-MSNs (Iino et al., 2020). In other words, D2 receptors are necessary to detect the dip of dopamine. The deletion of D2 receptors in NAc abolishes the active avoidance behavior (Danjo et al., 2014). We would like to hypothesize that when there is less D2 receptor availability in NAc, patients or laboratory animals are less sensitive to the aversive events. As a result, they don’t stop drug seeking despite negative consequences because of the insensitivity to negative events.

4.6 Conclusion

Combining optogenetic dopamine neurons self-stimulation with seek-take chained schedule, we confirmed that the potentiation at OFC-striatum is crucial for compulsive reward seeking. We also measured neuronal plasticity at mPFC-striatum and M1-striatum, detecting no differences between naïve, renouncing and persevering mice. This negative result suggests the septicity of the potentiation at OFC-striatum pathway. To investigate the impact of the synaptic potentiation on neuronal activity, we performed fiber photometry experiment in the striatum. Correlating with the plasticity, persevering mice showed stronger neural activity in the dorsal striatum at the end of seeking behaviour than renouncers. The activity was sustained during punishment sessions in persevering mice while in renouncers, the peak amplitude was diminished, suggesting that punishment induced neuronal adaptations exclusively to renouncers. To test the causality between the potentiation at OFC-striatum and strong neuronal activity in the striatum, chemogenetic inhibition was combined with the photometry experiment. The inhibition of the OFC flattered calcium signal observed in the striatum and attenuated compulsive reward seeking. Finally, to investigate the contribution of the neuronal activity in the striatum to compulsive reward seeking, we used inhibitory opsin in the dorsal striatum, allowing time locked inhibition. The inhibition at the end of the seeking behaviour attenuated compulsive reward seeking. This result has raised the idea that OFC-striatum encodes the value of the state, which was degraded by the inhibition.

77

Altogether, this work expanded the knowledge of how synaptic plasticity at OFC-striatum control the neuronal activity and eventually compulsive reward seeking. This knowledge might be inspiring the development of more efficient therapy of addiction.

5.References

Ago, Y., Takuma, K., and Matsuda, T. (2016). Methamphetamine-Induced Hyperlocomotion: A Focus on the Role of the Prefrontal Serotonergic System.

Neuropathology of Drug Addictions and Substance Misuse, Chapter 31.

Ann, M., and Wand, G. (2012). Stress and the HPA Axis: Role of Glucocorticoids in Alcohol Dependence. Alcohol Res. 34(4), 468–483.

Anthony, J.C., Warner, L.A., and Kessler, R.C. (1994). Comparative Epidemiology of Dependence on Tobacco, Alcohol, Controlled Substances, and Inhalants: Basic Findings From the National Comorbidity Survey. Exp. Clin. Psychopharmacol. 2, 244–

268.

Badiani, A., Oates, M.M., and Robinson, T.E. (2000). Modulation of morphine

sensitization in the rat by contextual stimuli. Psychopharmacology (Berl). 151, 273–282.

Balleine, B.W. (2019). The Meaning of Behavior: Discriminating Reflex and Volition in the Brain. Neuron 104, 47–62.

78

Balleine, B.W., and O’Doherty, J.P. (2010). Human and rodent homologies in action control: Corticostriatal determinants of goal-directed and habitual action.

Neuropsychopharmacology 35, 48–69.

Bariselli, S., Miyazaki, N.L., Creed, M.C., and Kravitz, A. V. (2020). Orbitofrontal-striatal potentiation underlies cocaine-induced hyperactivity. Nat. Commun. 11, 3996.

Belin-Rauscent, A., Fouyssac, M., Bonci, A., and Belin, D. (2016). How preclinical

models evolved to resemble the diagnostic criteria of drug addiction. Biol. Psychiatry 79, 39–46.

Belin, D., and Everitt, B.J. (2008). Cocaine Seeking Habits Depend upon Dopamine-Dependent Serial Connectivity Linking the Ventral with the Dorsal Striatum. Neuron 57, 432–441.

Belin, D., Jonkman, S., Dickinson, A., Robbins, T.W., and Everitt, B.J. (2009). Parallel and interactive learning processes within the basal ganglia: Relevance for the

understanding of addiction. Behav. Brain Res. 199, 89–102.

Belin, D., Berson, N., Balado, E., Piazza, P.V., and Deroche-Gamonet, V. (2011). High-novelty-preference rats are predisposed to compulsive cocaine self-administration.

Neuropsychopharmacology 36, 569–579.

Bell, K., and Kalivas, P.W. (1996). Context-specific cross-sensitization between systemic cocaine and intra-accumbens AMPA infusion in the rat. Psychopharmacology (Berl).

127, 377–383.

Bertran-Gonzalez, J., Bosch, C., Maroteaux, M., Matamales, M., Hervé, D., Valjent, E., and Girault, J.A. (2008). Opposing patterns of signaling activation in dopamine D1 and D2 receptor-expressing striatal neurons in response to cocaine and haloperidol. J.

Neurosci. 28, 5671–5685.

Bocklisch, C., Pascoli, V., Wong, J.C.Y., House, D.R.C., Yvon, C., de Roo, M., Tan, K.R., and Lüscher, C. (2013). Cocaine disinhibits dopamine neurons by potentiation of GABA transmission in the ventral tegmental area. Science 341, 1521–1525.

Boekhoudt, L., Wijbrans, E.C., Man, J.H.K., Luijendijk, M.C.M., de Jong, J.W., van der Plasse, G., Vanderschuren, L.J.M.J., and Adan, R.A.H. (2018). Enhancing excitability of

79

dopamine neurons promotes motivational behaviour through increased action initiation.

Eur. Neuropsychopharmacol. 28, 171–184.

Bolla, K., Ernst, M., Kiehl, K., Mouratidis, M., Eldreth, D., Contoreggi, C., Matochik, J., Kurian, V., M.H.S., J.C., Kimes, A., et al. (2004). Prefrontal Cortical Dysfunction in Abstinent Cocaine Abusers. J. Neuropsychiatr. 16, 456–464.

Bolloni, C., Badas, P., Corona, G., and Diana, M. (2018). Transcranial magnetic stimulation for the treatment of cocaine addiction : evidence to date. Subst. Abuse Rehabil. 2018:9, 11–21.

Borgland, S.L., Taha, S.A., Sarti, F., Fields, H.L., and Bonci, A. (2006). Orexin A in the VTA is critical for the induction of synaptic plasticity and behavioral sensitization to cocaine. Neuron 49, 589–601.

Boudreau, A.C., and Wolf, M.E. (2005). Behavioral Sensitization to Cocaine Is Associated with Increased AMPA Receptor Surface Expression in the Nucleus Accumbens. 25, 9144–9151.

Bradfield, L.A., Leung, B.K., Liang, S., Boldt, S., and Balleine, B.W. (2020). Goal-directed action transiently depends on dorsal hippocampus. Nat. Neurosci. 23, 1194–

1197.

Bromberg-martin, E.S., Matsumoto, M., and Hikosaka, O. (2010). Dopamine in Motivational Control : Rewarding, Aversive, and Alerting. Neuron 68, 815–834.

Brown, H.D., Mccutcheon, J.E., Cone, J.J., Ragozzino, M.E., and Roitman, M.F. (2011).

Primary food reward and reward-predictive stimuli evoke different patterns of phasic dopamine signaling throughout the striatum. Eur. J. Neurosci. 34, 1997–2006.

Brown, M.T.C., Bellone, C., Mameli, M., Labouèbe, G., Bocklisch, C., Balland, B., Dahan, L., Luján, R., Deisseroth, K., and Lüscher, C. (2010). Drug-driven AMPA

receptor redistribution mimicked by selective dopamine neuron stimulation. PLoS One 5, e15870.

Brown, M.T.C., Tan, K.R., O’Connor, E.C., Nikonenko, I., Muller, D., and Lüscher, C.

(2012). Ventral tegmental area GABA projections pause accumbal cholinergic interneurons to enhance associative learning. Nature 492, 452–456.

80

Brown, R.M., Kupchik, Y.M., Spencer, S., Garcia-Keller, C., Spanswick, D.C., Lawrence, A.J., Simonds, S.E., Schwartz, D.J., Jordan, K.A., Jhou, T.C., et al. (2017). Addiction-like Synaptic Impairments in Diet-Induced Obesity. Biol. Psychiatry 81, 797–806.

Calipari, E.S., Bagot, R.C., Purushothaman, I., Davidson, T.J., Yorgason, J.T., Peña, C.J., Walker, D.M., Pirpinias, S.T., Guise, K.G., Ramakrishnan, C., et al. (2016). In vivo imaging identifies temporal signature of D1 and D2 medium spiny neurons in cocaine reward. Proc. Natl. Acad. Sci. 113, 2726–2731.

Chakraborty, S., Brierley, D.I., and Ungless, M.A. (2009). Phasic excitation of dopamine neurons in ventral VTA by noxious stimuli. Proc. Natl. Acad. Sci. 106, 4894-4899

Chen, B.T., Bowers, M.S., Martin, M., Hopf, F.W., Guillory, A.M., Carelli, R.M., Chou, J.K., and Bonci, A. (2008). Cocaine but Not Natural Reward Self-Administration nor Passive Cocaine Infusion Produces Persistent LTP in the VTA. Neuron 59, 288–297.

Chen, B.T., Yau, H.-J., Hatch, C., Kusumoto-Yoshida, I., Cho, S.L., Hopf, F.W., and Bonci, A. (2013). Rescuing cocaine-induced prefrontal cortex hypoactivity prevents compulsive cocaine seeking. Nature 496, 359–362.

Di Chiara, G., and Imperato, A. (1988). Drugs abused by humans preferentially increase synaptic dopamine concentrations in the mesolimbic system of freely moving rats. Proc.

Natl. Acad. Sci. U. S. A. 85, 5274–5278.

Chornock, W.M., Stitzer, M.L., Gross, J., and Leischow, S. (1992). Experimental model of smoking re-exposure: effects on relapse. Psychopharmacology (Berl). 108, 495–500.

Chudasama, Y., and Robbins, T.W. (2003). Dissociable contributions of the orbitofrontal and infralimbic cortex to pavlovian autoshaping and discrimination reversal learning:

Further evidence for the functional heterogeneity of the rodent frontal cortex. J.

Neurosci. 23, 8771–8780.

Claus, E.D.I. neurobiological phenotypes associated with alcohol use disorder severity, Ewing, S.W.F., Filbey, F.M., Sabbineni, A., and Hutchison, K.E. (2011). Identifying Neurobiological Phenotypes Associated with Alcohol Use Disorder Severity.

Neuropsychopharmacology 36, 2086–2096.

Cohen, J.Y., Haesler, S., Vong, L., Lowell, B.B., and Uchida, N. (2012).

Neuron-type-81

specific signals for reward and punishment in the ventral tegmental area. Nature 482, 85–88.

Wallance, B.C. (1989). Psychological and Environmental Determinants of Relapse in Crack Cocaine Smokers. J. Subst. Abuse Treat. 6, 95–106.

Corbit, L.H., Nie, H., and Janak, P.H. (2014). Habitual responding for alcohol depends upon both AMPA and D2 receptor signaling in the dorsolateral striatum. Front. Behav.

Neurosci. 8, 301.

Corre, J., Zessen, R. van, Loureiro, M., Patriarchi, T., Tian, L., Pascoli, V., and Lüscher, C. (2018). Dopamine neurons projecting to medial shell of the nucleus accumbens drive heroin reinforcement. eLife 7,e39945.

Covey, D.P., and Cheer, J.F. (2019). Accumbal Dopamine Release Tracks the

Expectation of Dopamine Neuron-Mediated Reinforcement. Cell Rep. 27, 481-490.e3.

Creed, M., Pascoli, V.J., and Lüscher, C. (2015). Refining deep brain stimulation to emulate optogenetic treatment of synaptic pathology. Science 347, 659-664

Creed, M., Ntamati, N.R., Chandra, R., Lobo, M.K., and Lüscher, C. (2016).

Convergence of Reinforcing and Anhedonic Cocaine Effects in the Ventral Pallidum.

Neuron, 92, 214–226.

Dabney, W., Kurth-Nelson, Z., Uchida, N., Starkweather, C.K., Hassabis, D., Munos, R., and Botvinick, M. (2020). A distributional code for value in dopamine-based

reinforcement learning. Nature 577, 671–675.

Dalley, J.W., Fryer, T.D., Brichard, L., Robinson, E.S.J., Theobald, D.E.H., Lääne, K., Peña, Y., Murphy, E.R., Shah, Y., Probst, K., et al. (2007). Nucleus Accumbens D2/3 Receptors Predict Trait Impulsivity and Cocaine Reinforcement. Science. 315, 1267–

1270.

Danjo, T., Yoshimi, K., Funabiki, K., Yawata, S., and Nakanishi, S. (2014). Aversive behavior induced by optogenetic inactivation of ventral tegmental area dopamine neurons is mediated by dopamine D2 receptors in the nucleus accumbens. Proc. Natl.

Acad. Sci. 111, 6455–6460.

82

Deguchi, Y., Harada, M., Shinohara, R., Lazarus, M., Cherasse, Y., Urade, Y., Yamada, D., Sekiguchi, M., Watanabe, D., Furuyashiki, T., et al. (2016). mDia and ROCK Mediate Actin-Dependent Presynaptic Remodeling Regulating Synaptic Efficacy and Anxiety.

Cell Rep. 17, 2405–2417.

Deroche-Gamonet, V., Belin, D., and Piazza, P.V. (2004). Evidence for addiction-like behavior in the rat. Science. 305, 1014–1017.

Dolan, R.J., and Dayan, P. (2013). Goals and habits in the brain. Neuron 80, 312–325.

Domingo-rodriguez, L., Azua, I.R. De, Dominguez, E., Senabre, E., Serra, I., Kummer, S., Navandar, M., Baddenhausen, S., Hofmann, C., Andero, R., et al. (2020). A specific prelimbic-nucleus accumbens pathway controls resilience versus vulnerability to food addiction. Nat. Commun. 11, 782

Edwards, N.J., Tejeda, H.A., Pignatelli, M., Zhang, S., McDevitt, R.A., Wu, J., Bass, C.E., Bettler, B., Morales, M., and Bonci, A. (2017). Circuit specificity in the inhibitory architecture of the VTA regulates cocaine-induced behavior. Nat. Neurosci. 20, 438–

448.

Eichenbaum, H., Dudchenko, P., Wood, E., Shapiro, M., and Tanila, H. (1999). The Hippocampus , Memory , and Place Cells : Is It Spatial Memory or a Memory Space ? Neuron. 23, 209–226.

Everitt, B.J. (2014). Neural and psychological mechanisms underlying compulsive drug seeking habits and drug memories - indications for novel treatments of addiction. Eur. J.

Neurosci. 40, 2163–2182.

Everitt, B.J., and Robbins, T.W. (2013). From the ventral to the dorsal striatum:

Devolving views of their roles in drug addiction. Neurosci. Biobehav. Rev. 37, 1946–

1954.

Everitt, B.J., and Robbins, T.W. (2016). Drug Addiction: Updating Actions to Habits to Compulsions Ten Years On. Annu. Rev. Psychol. 67, 23–50.

Fanous, S., Goldart, E.M., Theberge, F.R.M., Bossert, J.M., Shaham, Y., and Hope, B.T.

(2012). Role of orbitofrontal cortex neuronal ensembles in the expression of incubation of heroin craving. J. Neurosci. 32, 11600–11609.

83

Faure, A., Haberland, U., Condé, F., and El Massioui, N. (2005). Lesion to the

nigrostriatal dopamine system disrupts stimulus-response habit formation. J. Neurosci.

25, 2771–2780.

Feil, J., Sheppard, D., Fitzgerald, P.B., Yücel, M., Lubman, D.I., and Bradshaw, J.L.

(2010). Addiction, compulsive drug seeking, and the role of frontostriatal mechanisms in regulating inhibitory control. Neurosci. Biobehav. Rev. 35, 248–275.

Fletcher, P.J., Azampanah, A., and Korth, K.M. (2002). Activation of 5-HT1B receptors in the nucleus accumbens reduces self-administration of amphetamine on a progressive ratio schedule. Pharmacol. Biochem. Behav. 71, 717–725.

Francis, T.C., Yano, H., Demarest, T.G., Shen, H., and Bonci, A. (2019).

High-Frequency Activation of Nucleus Accumbens D1-MSNs Drives Excitatory Potentiation on D2-MSNs. Neuron 103, 432–444.

Gerfen, C.R., and Surmeier, D.J. (2011). Modulation of Striatal Projection Systems by Dopamine. Annu. Rev. Neurosci. 34, 441–466.

Gibson, G.D., Prasad, A.A., Jean-Richard-dit-Bressel, P., Yau, J.O.Y., Millan, E.Z., Liu, Y., Campbell, E.J., Lim, J., Marchant, N.J., Power, J.M., et al. (2018).Distinct

Accumbens Shell Output Pathways Promote versus Prevent Relapse to Alcohol Seeking. Neuron 98, 512–520.

Girardeau, P., Navailles, S., Durand, A., Guillem, K., Ahmed, S.H., and Vouillac-mendoza, C. (2019). Neuropharmacology Relapse to cocaine use persists following extinction of drug-primed craving. Neuropharmacology 155, 185–193.

Goldstein, R.Z., and Volkow, N.D. (2002). Drug addiction and its underlying

neurobiological basis: Neuroimaging evidence for the involvement of the frontal cortex.

Am. J. Psychiatry 159, 1642–1652.

Goldstein, R.Z., and Volkow, N.D. (2011). Dysfunction of the prefrontal cortex in

addiction: Neuroimaging findings and clinical implications. Nat. Rev. Neurosci. 12, 652–

669.

Goldstein, R.Z., Tomasi, D., Alia-Klein, N., Cottone, L.A., Zhang, L., Telang, F., and Volkow, N.D. (2007). Subjective sensitivity to monetary gradients is associated with

84

frontolimbic activation to reward in cocaine abusers. Drug Alcohol Depend. 87, 233–240.

Goto, A., Nakahara, I., Yamaguchi, T., Kamioka, Y., Sumiyama, K., Matsuda, M., Nakanishi, S., and Funabiki, K. (2015). Circuit-dependent striatal PKA and ERK signaling underlies rapid behavioral shift in mating reaction of male mice. Proc. Natl.

Acad. Sci. U. S. A. 112, 6718–6723.

Graybiel, A.M. (2008). Habits, rituals, and the evaluative brain. Annu. Rev. Neurosci. 31, 359–387.

Gremel, C.M., and Costa, R.M. (2013). Orbitofrontal and striatal circuits dynamically encode the shift between goal-directed and habitual actions. Nat. Commun. 4, 2264.

Gremel, C.M., Chancey, J.H., Atwood, B.K., Luo, G., Neve, R., Ramakrishnan, C., Deisseroth, K., Lovinger, D.M., and Costa, R.M. (2016). Endocannabinoid Modulation of Orbitostriatal Circuits Gates Habit Formation. Neuron 90, 1312-1324.

Gruber, A.J., and McDonald, R.J. (2012). Context, emotion, and the strategic pursuit of goals: Interactions among multiple brain systems controlling motivated behavior. Front.

Behav. Neurosci. 6, 50.

de Guglielmo, G., Kallupi, M., Pomrenze, M.B., Crawford, E., Simpson, S., Schweitzer, P., Koob, G.F., Messing, R.O., and George, O. (2019). Inactivation of a CRF-dependent amygdalofugal pathway reverses addiction-like behaviors in alcohol-dependent rats.

Nat. Commun. 10, 1238.

Guillem, K., and Ahmed, S.H. (2018). Preference for Cocaine is Represented in the Orbitofrontal Cortex by an Increased Proportion of Cocaine Use-Coding Neurons.

Cereb. Cortex 28, 819–832.

Guillem, K., Brenot, V., Durand, A., and Ahmed, S.H. (2018). Neuronal representation of individual heroin choices in the orbitofrontal cortex. Addict. Biol. 23, 880–888.

Haber, S.N. (2017). Corticostriatal Circuitry. Neurosci. 21st Century 1721–1741.

Haber, S.N., Fudge, J.L., and McFarland, N.R. (2000). Striatonigrostriatal pathways in primates form an ascending spiral from the shell to the dorsolateral striatum. J.

Neurosci. 20, 2369–2382.

85

Han, D.D., and Gu, H.H. (2006). Comparison of the monoamine transporters from human and mouse in their sensitivities to psychostimulant drugs. BMC Pharmacol. 6, 6.

Harada, M., Hiver, A., Pascoli, V., and Lüscher, C. (2019). Cortico-striatal synaptic plasticity underlying compulsive reward seeking. BioRxiv.doi:

https://doi.org/10.1101/789495

Hart, G., and Balleine, B.W. (2016). Consolidation of Goal-Directed Action Depends on MAPK/ERK Signaling in Rodent Prelimbic Cortex. J. Neurosci. 36, 11974–11986.

Hart, G., Leung, B.K., and Balleine, B.W. (2014). Dorsal and ventral streams: The distinct role of striatal subregions in the acquisition and performance of goal-directed actions. Neurobiol. Learn. Mem. 108, 104–118.

Hart, G., Bradfield, L.A., and Balleine, B.W. (2018). Prefrontal cortico-striatal

disconnection blocks the acquisition of goal-directed action. J. Neurosci. 38, 1311-1322.

Hearing, M.C., Jedynak, J., Ebner, S.R., Ingebretson, A., Asp, A.J., Fischer, R.A., Schmidt, C., Larson, E.B., and Thomas, M.J. (2016). Reversal of morphine-induced cell-type–specific synaptic plasticity in the nucleus accumbens shell blocks reinstatement.

Proc. Natl. Acad. Sci. U. S. A. 113, 757–762.

Hirokawa, J., Vaughan, A., Masset, P., Ott, T., and Kepecs, A. (2019). Frontal cortex neuron types categorically encode single decision variables. Nature 576, 446–451.

Hnasko, T.S., Sotak, B.N., and Palmiter, R.D. (2005). Morphine reward in dopamine-deficient mice. Nature 438, 854–857.

Hu, D.A.N., and Amsel, A. (1995). A simple test of the vicarious trial-and-error hypothesis of hippocampal function. Proc. Natl. Acad. Sci. 92, 5506–5509.

Huang, Y.H., Lin, Y., Mu, P., Lee, B.R., Brown, T.E., Wayman, G., Marie, H., Liu, W., Yan, Z., Sorg, B. a, et al. (2009). In vivo cocaine experience generates silent synapses.

Neuron 63, 40–47.

Hunnicutt, B.J., Jongbloets, B.C., Birdsong, W.T., Gertz, K.J., Zhong, H., and Mao, T.

(2016). A comprehensive excitatory input map of the striatum reveals novel functional organization. eLife 5,e19103.

86

Huys, Q.J.M., Maia, T. V, and Frank, M.J. (2016). Computational psychiatry as a bridge between neuro- science and clinical applications. Nat. Neurosci. 19, 404-413.

Iino, Y., Sawada, T., Yamaguchi, K., Tajiri, M., Ishii, S., Kasai, H., and Yagishita, S.

(2020). Dopamine D2 receptors in discrimination learning and spine enlargement.

Nature 579, 555-560.

Iversen, S.D., and Mishkin, M. (1970). Perseverative interference in monkeys following selective lesions of the inferior prefrontal convexity. Exp. Brain Res. 11, 376–386.

de Jong, J.W., Afjei, S.A., Pollak Dorocic, I., Peck, J.R., Liu, C., Kim, C.K., Tian, L., Deisseroth, K., and Lammel, S. (2018). A Neural Circuit Mechanism for Encoding Aversive Stimuli in the Mesolimbic Dopamine System. Neuron 101, 133-151.

Jonkman, S., Pelloux, Y., and Everitt, B.J. (2012). Differential Roles of the Dorsolateral and Midlateral Striatum in Punished Cocaine Seeking. J. Neurosci. 32, 4645–4650.

Kasanetz, F., Deroche-Gamonet, V., Berson, N., Balado, E., Lafourcade, M., Manzoni, O., and Piazza, P.V. (2010). Transition to addiction is associated with a persistent impairment in synaptic plasticity. Science. 328, 1709–1712.

Kaufman, A.M., Geiller, T., and Losonczy, A. (2020). A Role for the Locus Coeruleus in Hippocampal CA1 Place Cell Reorganization during Spatial Reward Learning. Neuron.

105, 1019-1026.

Kaufman, J.N., Ross, T.J., Stein, E.A., and Garavan, H. (2003). Cingulate hypoactivity in cocaine users during a GO-NOGO task as revealed by event-related functional magnetic resonance imaging. J. Neurosci. 23, 7839–7843.

Keiflin, R., and Janak, P.H. (2015). Dopamine Prediction Errors in Reward Learning and Addiction: From Theory to Neural Circuitry. Neuron 88, 247–263.

Kerns, J.G., Cohen, J.D., MacDonald, A.W., Cho, R.Y., Stenger, V.A., and Carter, C.S.

(2004). Anterior Cingulate Conflict Monitoring and Adjustments in Control. Science. 303, 1023–1026.

Kim, C.K., Ye, L., Jennings, J.H., Yoo, A.W., Ramakrishnan, C., Deisseroth, K., Kim, C.K., Ye, L., Jennings, J.H., Pichamoorthy, N., et al. (2017). Molecular and

Circuit-87

Dynamical Identification of Top-Down Neural Mechanisms for Restraint of Article Molecular and Circuit-Dynamical Identification of Top-Down Neural Mechanisms for Restraint of Reward Seeking. Cell. 170, 1013-1027.

Koob, G.F. (2015). The dark side of emotion: The addiction perspective. Eur. J.

Pharmacol. 753, 73–87.

Koob, G.F. (2017). Antireward, compulsivity, and addiction: seminal contributions of Dr.

Athina Markou to motivational dysregulation in addiction. Psychopharmacology (Berl).

234, 1315–1332.

Kourrich, S., Rothwell, P.E., Klug, J.R., and Thomas, M.J. (2007). Cocaine Experience Controls Bidirectional Synaptic Plasticity in the Nucleus Accumbens. J. Neurosci. 27, 7921–7928.

Kuhn, B.N., Kalivas, P.W., and Bobadilla, A.C. (2019). Understanding Addiction Using Animal Models. Front. Behav. Neurosci. 13, 262.

Lahiri, A.K., and Bevan, M.D. (2020). Dopaminergic Transmission Rapidly and Persistently Enhances Excitability of D1 Receptor-Expressing Striatal Projection Neurons. Neuron 106, 277-290.

Lee, B.R., Ma, Y.-Y., Huang, Y.H., Wang, X., Otaka, M., Ishikawa, M., Neumann, P. a, Graziane, N.M., Brown, T.E., Suska, A., et al. (2013). Maturation of silent synapses in amygdala-accumbens projection contributes to incubation of cocaine craving. Nat.

Neurosci. 16, 1644–1651.

Lee, J., Wang, W., and Sabatini, B.L. (2020a). Anatomically segregated basal ganglia pathways allow parallel behavioral modulation. Nat. Neurosci. 23, 1388-1398

Lee, K., Claar, L.D., Hachisuka, A., Bakhurin, K.I., Nguyen, J., Trott, J.M., Gill, J.L., and Masmanidis, S.C. (2020b). Temporally restricted dopaminergic control of

reward-conditioned movements. Nat. Neurosci. 23. 209-216.

Lever, C., Wills, T., Cacucci, F., Burgess, N., and Keefe, J.O. (2002). Long-term plasticity in hippocampal place-cell representation of environmental geometry. Nature 416, 90-94.

88

Liddle, P.F., Kiehl, K.A., and Smith, A.M. (2001). Event-related fMRI study of response inhibition. Hum. Brain Mapp. 12, 100–109.

Lucantonio, F., Takahashi, Y.K., Hoffman, A.F., Chang, C.Y., Bali-Chaudhary, S.,

Shaham, Y., Lupica, C.R., and Schoenbaum, G. (2014). Orbitofrontal activation restores insight lost after cocaine use. Nat. Neurosci. 17, 1092–1099.

Lucantonio, F., Kambhampati, S., Haney, R.Z., Atalayer, D., Rowland, N.E., Shaham, Y., and Schoenbaum, G. (2015). Effects of Prior Cocaine Versus Morphine or Heroin Self-Administration on Extinction Learning Driven by Overexpectation Versus Omission of Reward. Biol. Psychiatry 77, 912–920.

Lüscher, C., and Ungless, M.A. (2006). The mechanistic classification of addictive drugs.

PLoS Med. 3, e437.

Lüscher, C., Robbins, T.W., and Everitt, B.J. (2020). The transition to compulsion in addiction. Nat. Rev. Neurosci. 21, 247–263.

Ma, T., Cheng, Y., Roltsch Hellard, E., Wang, X., Lu, J., Gao, X., Huang, C.C.Y., Wei, X.Y., Ji, J.Y., and Wang, J. (2018). Bidirectional and long-lasting control of alcohol-seeking behavior by corticostriatal LTP and LTD. Nat. Neurosci. 21, 373-383.

Ma, Y., Lee, B.R., Wang, X., Guo, C., Liu, L., Cui, R., Lan, Y., Balcita-pedicino, J.J., Wolf, M.E., Sesack, S.R., et al. (2014). Bidirectional Modulation of Incubation of Cocaine Craving by Silent Synapse-Based Remodeling of Prefrontal Cortex to Accumbens

Projections. Neuron 83, 1453–1467.

MacDonald, A.W., Cohen, J.D., Andrew Stenger, V., and Carter, C.S. (2000).

Dissociating the role of the dorsolateral prefrontal and anterior cingulate cortex in cognitive control. Science. 288, 1835–1838.

Marcellino, D., Kehr, J., Agnati, L.F., and Fuxe, K. (2012). Increased affinity of dopamine for D 2-like versus D 1-like receptors. Relevance for volume transmission in interpreting

Marcellino, D., Kehr, J., Agnati, L.F., and Fuxe, K. (2012). Increased affinity of dopamine for D 2-like versus D 1-like receptors. Relevance for volume transmission in interpreting

Documents relatifs