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:

Bidirectional and long-lasting control of alcohol-seeking behavior by corticostriatal LTP and LTD

Abstract

Addiction is proposed to arise from alterations in synaptic strength via mechanisms of long-term potentiation (LTP) and depression (LTD). However, the causality between these synaptic processes and addictive behaviors is difficult to demonstrate. Here we report that LTP and LTD induction altered operant alcohol self-administration, a motivated drug-seeking behavior. We first induced LTP by pairing presynaptic glutamatergic stimulation with optogenetic postsynaptic depolarization in the dorsomedial striatum, a brain region known to control goal-directed behavior. Blockade of this LTP by NMDA-receptor inhibition unmasked an endocannabinoid-dependent LTD. In vivo application of the LTP-inducing protocol caused a long-lasting increase in alcohol-seeking behavior, while the LTD protocol decreased this behavior. We further identified that optogenetic LTP and LTD induction at cortical inputs onto striatal dopamine D1 receptor-expressing neurons controlled these behavioral changes. Our results demonstrate a causal link between synaptic plasticity and alcohol-seeking behavior and suggest that modulation of this plasticity may inspire a therapeutic strategy for addiction.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Fig. 1: oPSD facilitated induction of NMDAR-dependent LTP and eCB LTD in DMS slices.
Fig. 2: oPSD facilitated corticostriatal LTP induction in the DMS.
Fig. 3: In vivo optogenetic LTP induction in the DMS produced a long-lasting increase in operant alcohol self-administration.
Fig. 4: In vivo LTD induction caused a long-lasting reduction of alcohol-seeking behavior in an eCB-dependent manner.
Fig. 5: Corticostriatal LTP was preferentially induced in DMS D1-MSNs.
Fig. 6: Corticostriatal LTD was preferentially induced in DMS D1-MSNs.
Fig. 7: Selective in vivo LTP or LTD induction in D1-MSNs produced long-lasting control of alcohol-seeking behavior.
Fig. 8: Model of bidirectional and long-lasting control of alcohol-seeking behavior by corticostriatal plasticity.

Similar content being viewed by others

References

  1. Volkow, N. D. & Morales, M. The brain on drugs: from reward to addiction. Cell 162, 712–725 (2015).

    Article  CAS  PubMed  Google Scholar 

  2. Everitt, B. J. & Robbins, T. W. Drug addiction: updating actions to habits to compulsions ten years on. Annu. Rev. Psychol. 67, 23–50 (2016).

    Article  PubMed  Google Scholar 

  3. Koob, G. F. & Volkow, N. D. Neurocircuitry of addiction. Neuropsychopharmacology 35, 217–238 (2010).

    Article  PubMed  Google Scholar 

  4. Lüscher, C. & Malenka, R. C. Drug-evoked synaptic plasticity in addiction: from molecular changes to circuit remodeling. Neuron 69, 650–663 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  5. Yin, H. H. & Knowlton, B. J. The role of the basal ganglia in habit formation. Nat. Rev. Neurosci. 7, 464–476 (2006).

    Article  CAS  PubMed  Google Scholar 

  6. Lovinger, D. M. Neurotransmitter roles in synaptic modulation, plasticity and learning in the dorsal striatum. Neuropharmacology 58, 951–961 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Balleine, B. W. & O’Doherty, J. P. Human and rodent homologies in action control: corticostriatal determinants of goal-directed and habitual action. Neuropsychopharmacology 35, 48–69 (2010).

    Article  PubMed  Google Scholar 

  8. Ma, T., Barbee, B., Wang, X. & Wang, J. Alcohol induces input-specific aberrant synaptic plasticity in the rat dorsomedial striatum. Neuropharmacology 123, 46–54 (2017).

    Article  CAS  PubMed  Google Scholar 

  9. Wang, J. et al. Ethanol-mediated facilitation of AMPA receptor function in the dorsomedial striatum: implications for alcohol drinking behavior. J. Neurosci. 32, 15124–15132 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Wang, J. et al. Long-lasting adaptations of the NR2B-containing NMDA receptors in the dorsomedial striatum play a crucial role in alcohol consumption and relapse. J. Neurosci. 30, 10187–10198 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Corbit, L. H., Chieng, B. C. & Balleine, B. W. Effects of repeated cocaine exposure on habit learning and reversal by N-acetylcysteine. Neuropsychopharmacology 39, 1893–1901 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Pascoli, V. et al. Contrasting forms of cocaine-evoked plasticity control components of relapse. Nature 509, 459–464 (2014).

    Article  CAS  PubMed  Google Scholar 

  13. Ma, Y. Y. et al. Bidirectional modulation of incubation of cocaine craving by silent synapse-based remodeling of prefrontal cortex to accumbens projections. Neuron 83, 1453–1467 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Gerfen, C. R. & Surmeier, D. J. Modulation of striatal projection systems by dopamine. Annu. Rev. Neurosci. 34, 441–466 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Shen, W., Flajolet, M., Greengard, P. & Surmeier, D. J. Dichotomous dopaminergic control of striatal synaptic plasticity. Science 321, 848–851 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Cheng, Y. et al. Distinct synaptic strengthening of the striatal direct and indirect pathways drives alcohol consumption. Biol. Psychiatry 81, 918–929 (2017).

    Article  CAS  PubMed  Google Scholar 

  17. MacAskill, A. F., Cassel, J. M. & Carter, A. G. Cocaine exposure reorganizes cell type- and input-specific connectivity in the nucleus accumbens. Nat. Neurosci. 17, 1198–1207 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Klapoetke, N. C. et al. Independent optical excitation of distinct neural populations. Nat. Methods 11, 338–346 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Fenno, L., Yizhar, O. & Deisseroth, K. The development and application of optogenetics. Annu. Rev. Neurosci. 34, 389–412 (2011).

    Article  CAS  PubMed  Google Scholar 

  20. Surmeier, D. J., Plotkin, J. & Shen, W. Dopamine and synaptic plasticity in dorsal striatal circuits controlling action selection. Curr. Opin. Neurobiol. 19, 621–628 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Kreitzer, A. C. & Malenka, R. C. Endocannabinoid-mediated rescue of striatal LTD and motor deficits in Parkinson’s disease models. Nature 445, 643–647 (2007).

    Article  CAS  PubMed  Google Scholar 

  22. Mattis, J. et al. Principles for applying optogenetic tools derived from direct comparative analysis of microbial opsins. Nat. Methods 9, 159–172 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  23. Yizhar, O. et al. Neocortical excitation/inhibition balance in information processing and social dysfunction. Nature 477, 171–178 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Mathur, B. N. & Lovinger, D. M. Endocannabinoid-dopamine interactions in striatal synaptic plasticity. Front. Pharmacol. 3, 66 (2012).

  25. Gerdeman, G. L., Ronesi, J. & Lovinger, D. M. Postsynaptic endocannabinoid release is critical to long-term depression in the striatum. Nat. Neurosci. 5, 446–451 (2002).

    Article  CAS  PubMed  Google Scholar 

  26. Wu, Y. W. Input- and cell-type-specific endocannabinoid-dependent LTD in the striatum. Cell Rep 10, 75–87 (2015).

    Article  CAS  PubMed  Google Scholar 

  27. Minatohara, K., Akiyoshi, M. & Okuno, H. Role of immediate-early genes in synaptic plasticity and neuronal ensembles underlying the memory trace. Front. Mol. Neurosci. 8, 78 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  28. Girault, J. A., Valjent, E., Caboche, J. & Hervé, D. ERK2: a logical AND gate critical for drug-induced plasticity? Curr. Opin. Pharmacol. 7, 77–85 (2007).

    Article  CAS  PubMed  Google Scholar 

  29. Ye, L. et al. Wiring and molecular features of prefrontal ensembles representing distinct experiences. Cell 165, 1776–1788 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Sun, X. & Lin, Y. Npas4: linking neuronal activity to memory. Trends Neurosci. 39, 264–275 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Wang, J. et al. Ethanol induces long-term facilitation of NR2B-NMDA receptor activity in the dorsal striatum: implications for alcohol drinking behavior. J. Neurosci. 27, 3593–3602 (2007).

    Article  CAS  PubMed  Google Scholar 

  32. Kash, T. L., Baucum, A. J. II, Conrad, K. L., Colbran, R. J. & Winder, D. G. Alcohol exposure alters NMDAR function in the bed nucleus of the stria terminalis. Neuropsychopharmacology 34, 2420–2429 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Rosenmund, C. & Stevens, C. F. Definition of the readily releasable pool of vesicles at hippocampal synapses. Neuron 16, 1197–1207 (1996).

    Article  CAS  PubMed  Google Scholar 

  34. Lovinger, D. M., & Kash, T. L. Mechanisms of neuroplasticity and ethanol’seffects on plasticity in the striatum and bed nucleus of the stria terminalis. Alcohol Res 37, 109–124 (2015).

    PubMed  PubMed Central  Google Scholar 

  35. McCool, B. A. Ethanol modulation of synaptic plasticity. Neuropharmacology 61, 1097–1108 (2011).

    Article  CAS  PubMed  Google Scholar 

  36. Lobo, M. K. & Nestler, E. J. The striatal balancing act in drug addiction: distinct roles of direct and indirect pathway medium spiny neurons. Front. Neuroanat. 5, 41 (2011).

  37. Wills, T. A. et al. GluN2B subunit deletion reveals key role in acute and chronic ethanol sensitivity of glutamate synapses in bed nucleus of the stria terminalis. Proc. Natl. Acad. Sci. USA 109, E278–E287 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Zalocusky, K. A. et al. Nucleus accumbens D2R cells signal prior outcomes and control risky decision-making. Nature 531, 642–646 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Wang, Z. et al. Dopaminergic control of corticostriatal long-term synaptic depression in medium spiny neurons is mediated by cholinergic interneurons. Neuron 50, 443–452 (2006).

    Article  CAS  PubMed  Google Scholar 

  40. Calabresi, P, Pisani, A, Mercuri, N. B. & Bernardi, G. Long-term potentiation in the striatum is unmasked by removing the voltage-dependent magnesium block of NMDA receptor channels.Eur. J. Neurosci 4, 929–935 (1992).

    Article  PubMed  Google Scholar 

  41. Mancini, M. et al. Memantine alters striatal plasticity inducing a shift of synaptic responses toward long-term depression. Neuropharmacology 101, 341–350 (2016).

    Article  CAS  PubMed  Google Scholar 

  42. Huang, Y. H. et al. In vivo cocaine experience generates silent synapses. Neuron 63, 40–47 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Dong, Y. & Nestler, E. J. The neural rejuvenation hypothesis of cocaine addiction. Trends Pharmacol. Sci. 35, 374–383 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Lee, B. R. et al. Maturation of silent synapses in amygdala-accumbens projection contributes to incubation of cocaine craving. Nat. Neurosci. 16, 1644–1651 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Ma, Y. Y. et al. Re-silencing of silent synapses unmasks anti-relapse effects of environmental enrichment. Proc. Natl. Acad. Sci. USA 113, 5089–5094 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Creed, M., Pascoli, V. J. & Lüscher, C. Addiction therapy. Refining deep brain stimulation to emulate optogenetic treatment of synaptic pathology. Science 347, 659–664 (2015).

    Article  CAS  PubMed  Google Scholar 

  47. Creed, M., Ntamati, N. R., Chandra, R., Lobo, M. K. & Lüscher, C. Convergence of reinforcing and anhedonic cocaine effects in the ventral pallidum. Neuron 92, 214–226 (2016).

    Article  CAS  PubMed  Google Scholar 

  48. Wang, J. et al. Alcohol elicits functional and structural plasticity selectively in dopamine D1 receptor-expressing neurons of the dorsomedial striatum. J. Neurosci. 35, 11634–11643 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Sidhpura, N. & Parsons, L. H. Endocannabinoid-mediated synaptic plasticity and addiction-related behavior. Neuropharmacology 61, 1070–1087 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Wei, X. et al. Dopamine D1 or D2 receptor-expressing neurons in the central nervous system. Addict. Biol. https://doi.org/10.1111/adb.12512 (2017).

  51. Huang, C. C. Y. et al. Stroke triggers nigrostriatal plasticity and increases alcohol consumption in rats. Sci. Rep. 7, 2501 (2017).

  52. Yin, H.H., Park, B. S., Adermark, L. & Lovinger, D.M. Ethanol reverses the direction of long-term synaptic plasticity in the dorsomedial striatum. Eur. J. Neurosci. 25, 3226–3232 (2007).

    Article  PubMed  Google Scholar 

  53. Cruikshank, S. J., Urabe, H., Nurmikko, A. V. & Connors, B. W. Pathway-specific feedforward circuits between thalamus and neocortex revealed by selective optical stimulation of axons. Neuron 65, 230–245 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Chen, T. W. et al. Ultrasensitive fluorescent proteins for imaging neuronal activity. Nature 499, 295–300 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Wang, J., Yeckel, M. F., Johnston, D. & Zucker, R. S. Photolysis of postsynaptic caged Ca2+ can potentiate and depress mossy fiber synaptic responses in rat hippocampal CA3 pyramidal neurons. J. Neurophysiol. 91, 1596–1607 (2004).

    Article  PubMed  Google Scholar 

  56. June, H. L. & Gilpin, N. W. Operant self-administration models for testing the neuropharmacological basis of ethanol consumption in rats. Curr. Protoc. Neurosci. 9, 11–26 (2010). Chapter 9, Unit12.

  57. Carnicella, S., Amamoto, R. & Ron, D. Excessive alcohol consumption is blocked by glial cell line-derived neurotrophic factor. Alcohol 43, 35–43 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Simms, J. A., Bito-Onon, J. J., Chatterjee, S. & Bartlett, S. E. Long-Evans rats acquire operant self-administration of 20% ethanol without sucrose fading. Neuropsychopharmacology 35, 1453–1463 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Zhao, X. et al. Regulation of lipogenesis by cyclin-dependent kinase 8-mediated control of SREBP-1. J. Clin. Invest. 122, 2417–2427 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank W. Griffith, D. DuBois, M. Han, R. Smith, Y. Zhu, J. Artz and A. Mahnke for discussions and technical assistance. This research was supported by NIAAA R01AA021505 (J.W.), NIAAA U01AA025932 (J.W.) and NIDDK R01DK095013 (J.Y.J.).

Author information

Authors and Affiliations

Authors

Contributions

J.W. conceived, designed and supervised all experiments in the study. J.W. and T.M. wrote a draft of the manuscript. J.W., T.M. and Y.C. revised the manuscript. T.M. also designed and performed electrophysiology experiments and analyzed the data. Y.C. and E.R.H. designed and performed the behavior experiments and analyzed the data. X.W. conducted behavior experiments. J.-Y.J. supervised the collection and analyses of biochemical experiments. J.L. and X.G. performed and analyzed the biochemical experiments. C.C.Y.H. conducted the electrophysiology experiment for virus validation and analyzed the data. Y.C. and X.-Y.W. conducted confocal imaging experiments and analyzed the data. All authors discussed the results and commented on the manuscript.

Corresponding author

Correspondence to Jun Wang.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Integrated supplementary information

Supplementary Figure 1 Pairing of eHFS with iPSD did not induce LTP in DMS slices of adult rats.

(a) Schematic illustration of LTP-inducing protocols. Presynaptic eHFS consists of 4 trains of stimuli at an interval of 20 sec, and each train contains 100 pulses at 100 Hz. eHFS was delivered alone or paired with somatic current injection-induced (iPSD) or optogenetically induced postsynaptic depolarization (oPSD). (b) Paired eHFS+iPSD did not induce LTP in DMS slices of adult rats (102.43 ± 6.47 % of baseline, t(6) = -0.38, P = 0.72; paired t test. n = 7 slices from 5 rats). Whole-cell current-clamp recordings were conducted to measure excitatory postsynaptic potentials (EPSPs). Step currents (250 pA, 1 sec) were injected through the patch pipette during four trains of eHFS. Sulpiride was bath-applied to block LTD and favor LTP induction. Data are presented as mean ± s.e.m.

Supplementary Figure 2 oPSD induced greater Ca2+ transients in distal dendrites than iPSD.

AAV-GCaMP6s and AAV-C1V1-eYFP were co-infused into the DMS, and whole-cell current-clamp recordings were conducted in C1V1-expressing neurons. A 2-sec step current was injected through the patch pipette, and 1-sec light stimulation (590 nm) was delivered through the objective lens. (a) Sample whole-cell recording showing action potentials elicited by iPSD or oPSD. Note that iPSD-induced a train of spikes, whereas oPSD induced a few spikes, which may be due to depolarization block1. (b) Representative traces of dendritic Ca2+ transients induced by iPSD (left) or by oPSD (right). (c) oPSD induced significantly greater Ca2+ transients in the distal dendrites than did iPSD. Left, comparison of dendritic Ca2+ transients induced by iPSD and oPSD. t(13) = -3.34, **P = 0.0054. Right, comparison of normalized dendritic Ca2+ transients to the somatic ones. t(13) = -3.15, **P = 0.0077; Two-sided unpaired t test. n = 8 neurons, 5 rats (iPSD) and 7 neurons, 3 rats (oPSD). Data are presented as mean ± s.e.m. 1Kleinlogel, S. et al. Ultra light-sensitive and fast neuronal activation with the Ca2+-permeable channelrhodopsin CatCh. Nat Neurosci 14, 513-518 (2011).

Supplementary Figure 3 oPSD enhanced NMDAR-mediated EPSPs (EPSPNMDA) and Ca2+ influx.

AAV-C1V1-eYFP (a-e), AAV-Chrimson-tdTomato and AAV-GCaMP6s (f, g) were infused into the DMS. (a) Sample confocal image of an Alexa Fluor 594-filled C1V1-eYFP-expressing neuron showing the placement of stimulating (150 µm from the soma) and recording electrodes. (b) Representative traces depicting optogenetic-mediated EPSP in the distal dendrites in response to single synaptic stimulation. After blockade of the AMPAR-mediated EPSP (EPSPAMPA) with NBQX (10 µM), simultaneous presynaptic stimulation and oPSD elicited a depolarization (C1V1-mediated response [Vc1v1] + EPSPNMDA) that was reduced by bath application of APV (Vc1v1). The optogenetic-mediated EPSPNMDA was calculated by digital subtraction of Vc1v1 from Vc1v1 + EPSPNMDA. (c) Comparison of the amplitudes of the electrically and optogenetically induced responses in the absence (E+O) and presence (E+O+APV) of APV. t(9) = 4.93, ***P = 0.00082; n = 10 neurons, 3 rats. (d) Top, sample EPSP traces in response to eHFS in the presence (color) and absence (black) of APV. Bottom, traces were generated by digital subtraction of the EPSP with APV from that without APV. (e) Paired eHFS+oPSD induced a greater area under EPSPNMDA than eHFS alone. t(7) = -4.50, **P = 0.0028; n = 8 neurons, 4 rats per group. (f) Representative traces of Ca2+ transients in distal dendrites in response to eHFS (left) and eHFS+oPSD (right) with and without APV application. (g) eHFS+oPSD induced greater NMDAR-mediated Ca2+ transients than that of eHFS. t(11) = 4.30, **P = 0.0012; n = 12 neurons, 3 rats per group. Two-sided paired t test for c, e, g. Data are presented as mean ± s.e.m.

Supplementary Figure 4 Dopamine signaling regulated optogenetic LTP in DMS slices.

(a) Optogenetic induction of LTP in the presence of the dopamine D1R antagonist, SCH 23390 (10 µM, left), and agonist, SKF 38393 (20 µM, right) in DMS slices of rats. AAV-C1V1-eYFP was infused into the DMS. The grey lines are the control LTP from Fig. 1d for reference. Left, suppression of LTP by SCH 23390 (108.65 ± 1.95% of baseline, t(5) = -4.44, P = 0.0068; n = 6 slices, 6 rats). Right, enhancement of LTP by SKF 38393 (128.84 ± 2.69% of baseline, t(7) = -10.70, P < 0.0001; n = 8 slices, 5 rats). Sulpiride was bath-applied to block LTD and favor LTP induction. (b) Comparison of the magnitudes of LTP in control and in the presence of SCH 23390 and of SKF 38393 in rats. The grey dots are from Fig. 1d for comparison (SCH 23390 vs Control: t(14) = 2.33, *P = 0.035; SKF 38393 vs Control: t(16) = -2.56, *P = 0.021). n = 10 slices from 6 rats (Control), 6 slices from 6 rats (SCH 23390), and 8 slices from 5 rats (SKF 38393). Two-sided paired t test for a; two-sided unpaired t test for b. Data are presented as mean ± s.e.m.

Supplementary Figure 5 Verification of the fidelity of spiking in mPFC neurons expressing Chronos and optical fEPSP/PS in DMS slices.

(a-e) AAV-Chronos-GFP was infused into the mPFC and coronal sections containing the mPFC or DMS were prepared eight weeks after infusion. (a) Schematic illustration of whole-cell recording in Chronos-expressing mPFC neurons and field recording in the DMS area containing the Chronos-expressing fibers in separated experiments. (b) Whole-cell recording showing that Chronos-expressing mPFC neurons can fire spikes in response to a train of light pulses (2 ms) at 50 Hz. Scale bars: 20 ms, 10 mV. (c) 405-nm light-driven spike probability over a range of frequencies. F(3,12) = 17.51, *P < 0.05, ***P < 0.001; n = 6 neurons, 2 rats. (d) Field recording demonstrating that Chronos-expressing mPFC fibers in the DMS can exert fEPSP/PS in response to 50-Hz train stimulation of light. Scale bars: 20 ms, 10 µV. (e) An example recording of a light-evoked fEPSP/PS, which was blocked by a mixture of AMPAR and NMDAR antagonists CNQX+CPP. Note that the remaining response in the presence of CNQX and CPP was partially blocked by TTX and completely by reducing light intensity to 2% of the original light intensity. Scale bars: 5 ms, 0.2 mV. One-way RM ANOVA for c. Data are presented as mean ± s.e.m.

Supplementary Figure 6 LTD was observed in whole-cell recording of D1-MSNs but not in field recording.

(a) Paired oHFS+oPSD in the presence of MK801 (50 µM) failed to induce LTD in DMS slices (99.76 ± 4.99% of baseline, t(7) = 0.05, P = 0.96; n = 8 slices, 6 rats). The gray line is the averaged data from Fig. 2c for comparison. (b) In presence of the D2R antagonist, sulpiride, and APV, paired oHFS+oPSD failed to induce LTD (99.49 ± 2.25% of baseline, t(5) = 0.22, P = 0.83; n = 6 slices, 3 rats). (c) Optogenetic induction of corticostriatal LTD in DMS D1-MSNs. AAV-Chronos-GFP was infused into the mPFC for selective stimulation of this input; the retrograde AAV5-Cre was infused into the SNr and AAV-Flex-Chrimson-tdTomato into the DMS for selective oPSD of D1-MSNs. Whole-cell recording was conducted in tdTomato-positive neurons to measure EPSPs before and after paired oHFS+oPSD. The pairing in the presence of MK801 and raclopride induced LTD in D1-MSNs (78.28 ± 2.60% of baseline, t(6) = 8.35, P = 0.00016; n = 7 neurons, 4 rats). Two-sided paired t test for a-c. Data are presented as mean ± s.e.m.

Supplementary Figure 7 In vivo delivery of the optogenetic LTP-inducing protocol facilitated alcohol- but not sucrose-seeking behavior.

(a) In vivo optogenetic LTP induction persistently increased alcohol deliveries. F(5,54) = 3.83, P = 0.0048, *P < 0.05 vs. baseline (BL). n = 14 rats. (b,c) In vivo LTP induction increased blood alcohol concentration (BAC) at 30 min (b) and 2 days (c) after induction. 30 min: t(5) = -4.25, P = 0.0081, n = 6 rats; 2 days: t(5) = -3.36, P = 0.02, n = 6 rats. *P < 0.05 and **P < 0.01 vs. baseline (BL). (d) In vivo oHFS stimulation alone did not persistently alter lever presses for alcohol (left; F(4,17) = 4.38, P = 0.013, n = 6 rats), alcohol deliveries (middle; F(4,17) = 4.67, P = 0.01, n = 6 rats), or alcohol intake (right; F(4,17) = 0.59, P = 0.67, n = 6 rats). *P < 0.05. (e) In vivo oPSD stimulation alone did not alter lever presses for alcohol (left; F(4,26) = 0.63, P = 0.65, n = 8 rats), alcohol deliveries (middle; F(4,26) = 0.41, P = 0.80, n = 8 rats), or alcohol intake (right; F(4,26) = 0.6, P = 0.66, n = 8 rats). (f) In vivo LTP induction in the presence of a D1R antagonist, SCH 23390 (SCH, 0.01 mg/kg), failed to alter lever presses (left; F(4,40) = 0.66, P = 0.63), alcohol deliveries (middle; F(4,39) = 1.26, P = 0.30), or alcohol intake (right; F(4,39) = 0.58, P = 0.68). n = 12 rats. (g) In vivo LTP induction in the presence of MK801 failed to alter lever presses (left; t(4) = -2.09, P = 0.11), alcohol deliveries (middle; t(4) = -2.07, P = 0.11), or alcohol intake (right; t(4) = -0.60, P = 0.58). n = 5 rats. (h) Systematic administration of SCH 23309 did not alter lever presses (left; t(13) = 0.16, P = 0.87), alcohol deliveries (middle; t(13) = 0.03, P = 0.98), or alcohol intake (right; t(13) = -0.85, P = 0.41), as compared to baseline (BL). n = 14 rats. (i) In vivo LTP induction did not alter sucrose deliveries (left) or intake (right). Deliveries: F(4,19) = 0.84, P = 0.51; intake: F(4,18) = 1.94, P = 0.15, n = 6 rats. One-way RM ANOVA for a,d-f,i; two-sided paired t test for b, c, g, and h. Data are presented as mean ± s.e.m.

Supplementary Figure 8 Operant alcohol self-administration facilitated NMDAR activity and corticostriatal LTP induction in the DMS.

(a) Operant alcohol or sucrose self-administration increased the AMPAR/NMDAR ratio, as compared with age-matched water controls. Note that the data for the EtOH and sucrose groups (grey dots) are the same as in Fig. 3e and 3g, respectively. Rats in the water control group did not receive operant training. F(2,34) = 11.13, P = 0.00019; **P < 0.01 and ***P < 0.001 versus Water; n = 13 neurons from 3 rats (Water), 13 neurons from 5 rats (EtOH), and 11 neurons from 5 rats (Sucrose). (b) Schematic illustration of the experimental procedure: After infusion of AAV-Chronos into the mPFC and AAV-Chrimson into the DMS as Fig. 3a, rats underwent operant training but without in vivo light stimulation. Twenty-four hours after the last alcohol or sucrose exposure, DMS slices were prepared. (c) Operant alcohol self-administration caused higher NMDAR activity than did operant sucrose self-administration. Comparison of input-output relation for NMDAR-EPSCs between alcohol and sucrose groups. F(1, 94) = 9.62, ##P = 0.0049; n = 13 neurons from 4 rats (Sucrose) and 13 neurons from 5 rats (EtOH). (d) Paired oHFS+oPSD induced a LTP in DMS slices from alcohol-exposed [109.66 ± 1.92% of baseline (BL), t(9) = -5.02, P = 0.00072], but not from sucrose-exposed rats [100.81 ± 1.69% of BL, t(6) = -0.48, P = 0.65; Two-sided paired t test]. ##P = 0.0052; two-sided unpaired t test. n = 10 slices from 8 rats (EtOH) and 7 slices from 4 rats (Sucrose). One-way ANOVA followed by SNK test for a; two-way RM ANOVA followed by SNK test for c. Data are presented as mean ± s.e.m.

Supplementary Figure 9 In vivo LTP-inducing protocol increased AMPAR EPSC rectification.

(a) Representative traces recorded at -70, 0, and +40 mV in the control (Ctrl) and LTP groups. Control rats received training of operant alcohol self-administration, but without in vivo light stimulation. LTP rats received the same operant training and in vivo LTP induction two days before the recording. (b) I/V curves and rectification index (RI) of light-evoked AMPAR-EPSCs at -70, 0, and +40 mV show that in vivo LTP induction increased the RI of AMPAR-EPSCs. The curves were plotted by normalizing the EPSC amplitudes at +40 and 0 mV to the amplitude at -70 mV. t(23) = 3.54, **P = 0.0018 for difference in AMPAR-EPSCs at +40 mV; t(23) = -3.53, **P = 0.0018 for RI. n = 12 neurons from 3 rats (Ctrl) and 13 neurons from 3 rats (LTP). Two-sided unpaired t test. Data are presented as mean ± s.e.m.

Supplementary Figure 10 In vivo delivery of the optogenetic LTD-inducing protocol reduced alcohol-seeking behavior in an endocannabinoid manner.

(a) In vivo delivery of LTD-inducing protocol (oHFS+oPSD+MK801+raclopride) in the DMS produced a long-lasting decrease in alcohol deliveries. F(5,38) = 4.27, P = 0.0035. *P < 0.05, **P < 0.01 vs. baseline (BL). n = 9 rats. (b-d) In vivo oHFS stimulation in presence of the cocktail (MK801+raclopride) did not alter lever presses (b; F(4,27) = 0.50, P = 0.74), alcohol deliveries (c; F(4,27) = 0.32, P = 0.86), or alcohol intake (d; F(4,27) = 0.53, P = 0.72). n = 9 rats. (e-g) Administration of the cocktail (MK801+raclopride) did not affect lever presses (e; F(4,21) = 2.12, P = 0.11), alcohol deliveries (f; F(4,21) = 1.43, P = 0.26), or alcohol intake (g; F(4,21) = 0.42, P = 0.79). n = 8 rats. (h-i) In vivo optogenetic delivery of the LTD-inducing protocol (oHFS+oPSD+MK801+raclopride) in the presence of AM251 failed to alter alcohol deliveries (h; F(4,31) = 0.46, P = 0.76), or alcohol intake (i; F(4,32) = 0.42, P = 0.79). n = 9 rats. (j-l) Systemic administration of AM251 (0.3 mg/kg) did not affect lever presses (j; F(2,17) = 0.33, P = 0.73), alcohol deliveries (k; F(2,17) = 1.05, P = 0.37), or alcohol intake (l; F(2,17) = 2.31, P = 0.13), as compared to before (baseline, BL) and after (Post) treatment. n = 10 rats. One-way RM ANOVA for all figures. Data are presented as mean ± s.e.m.

Supplementary Figure 11 In vivo delivery of the LTP- and LTD-inducing protocols to mPFC inputs onto D1-MSNs distinctly modulated alcohol seeking-behavior.

(a) In vivo delivery of the LTP-inducing protocol (oHFS+oPSD) to the mPFC input onto DMS D1-MSNs induced a persistent increase in alcohol-deliveries. *P < 0.05 vs. baseline (BL). F(4,23) = 4.77, P = 0.0060; n = 8 rats. (b, c) In vivo oPSD of D1-MSNs alone did not alter alcohol deliveries (b; F(4,18) = 0.32, P = 0.86) or intake (c; F(4,18) = 0.76, P = 0.56; n = 6 rats). (d) In vivo delivery of LTD-inducing protocol (oHFS+oPSD+MK801+raclopride) to the mPFC input onto DMS D1-MSNs produced a long-lasting attenuation in alcohol deliveries. *P < 0.05 vs. baseline (BL). F(4,23) = 4.99, P = 0.0048; n = 7 rats. (e,f) In vivo delivery of the LTD-inducing protocol in the presence of AM251 (oHFS+oPSD+MK801+raclopride+AM251) failed to alter alcohol deliveries (e; F(4,20) = 0.24, P = 0.91) or intake (f; F(4,19) = 1.67, P = 0.20; n = 7 rats). One-way RM ANOVA for all figures. Data are presented as mean ± s.e.m.

Supplementary Figure 12 In vivo delivery of the LTP- or LTD-inducing protocol did not alter inactive lever presses.

(a-b) In vivo delivery of the optogenetic LTP-inducing protocol did not change inactive lever presses for alcohol (a, F(5,54) = 0.41, P = 0.84; n = 14 rats; related to Fig. 3b) or for sucrose (b, F(4,18) = 0.31, P = 0.87; n = 6 rats; related to Fig. 3c). (c) In vivo delivery of the optogenetic LTD-inducing protocol did not change inactive lever presses for alcohol. F(5,30) = 1.05, P = 0.41; n = 9 rats; related to Fig. 4b. (d) In vivo delivery of the optogenetic LTD-inducing protocol in the presence of AM251 did not change inactive lever presses for alcohol. F(4,30) = 1.33, P = 0.28; n = 9 rats; related to Fig. 4c. (e-f) In vivo delivery of the LTP-inducing protocol (oHFS+oPSD) (e, F(4,25) = 0.53, P = 0.72; n = 8 rats; related to Fig. 7c) or oPSD (f, F(4,16) = 1.44, P = 0.27; n = 6 rats; related to Fig. 7d) to the mPFC input onto DMS D1-MSNs did not affect inactive lever presses for alcohol. (g) In vivo delivery of the LTD-inducing protocol to the mPFC input onto DMS D1-MSNs did not alter inactive lever presses for alcohol. F(4,18) = 1.59, P = 0.22; n = 6 rats; related to Fig. 7e. (h) In vivo delivery of the LTD-inducing protocol to the mPFC input onto DMS D1-MSNs in the presence of AM251 decreased inactive lever presses for alcohol only at 30 min, but not on day 2, 4, or 7. *P < 0.05 versus baseline (BL). F(4,23) = 5.04, P = 0.0046; n = 7 rats; related to Fig. 7f. (i-j) In vivo delivery of oHFS (i, F(4,17) = 1.86, P = 0.16; n = 6 rats; related to Supplementary Figure 7d) or oPSD (j, F(4,25) = 1.39, P = 0.26; n = 8 rats; related to Supplementary Figure 7e) did not cause any change on inactive lever presses for alcohol. (k-l) In vivo delivery of the LTP-inducing protocol in the presence of SCH 23390 (k, F(4,43) = 0.57, P = 0.69; n = 12 rats; related to Supplementary Figure 7f) or MK801 (l, t(4) = 0, P = 1; n = 5 rats; related to Supplementary Figure 7g) did not change the inactive lever presses for alcohol. (m) Systematic administration of SCH 23390 did not alter inactive lever presses at 30 min (t(13) = 0.16, P = 0.87; n = 13 rats; related to Supplementary Figure 7h). (n) In vivo delivery of oHFS with administration of MK801 and raclopride did not change inactive lever presses for alcohol. F(4,28) = 1.17, P = 0.34; n = 9 rats; related to Supplementary Figure 10b. (o) Systematic administration of a cocktail of MK801 and raclopride did not affect the inactive lever presses for alcohol. F(4,24) = 0.96, P = 0.45; n = 8 rats; related to Supplementary Figure 10e. (p) Systematic administration of AM251 (0.3 mg/kg) did not alter inactive lever presses for alcohol. F(2,18) = 2.04, P = 0.16; n = 10 rats; related to Supplementary Figure 10j. One-way RM ANOVA for Figures a-k and n-p; two-sided paired t test for Figures l and m. Data are presented as mean ± s.e.m.

Supplementary information

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ma, T., Cheng, Y., Roltsch Hellard, E. et al. Bidirectional and long-lasting control of alcohol-seeking behavior by corticostriatal LTP and LTD. Nat Neurosci 21, 373–383 (2018). https://doi.org/10.1038/s41593-018-0081-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-018-0081-9

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing