Sei sulla pagina 1di 133

1521-0081/67/4/8721004$25.

00
PHARMACOLOGICAL REVIEWS
Copyright 2015 by The American Society for Pharmacology and Experimental Therapeutics

http://dx.doi.org/10.1124/pr.115.010967
Pharmacol Rev 67:8721004, October 2015

ASSOCIATE EDITOR: MARKKU KOULU

Mechanisms of Action and Persistent Neuroplasticity


by Drugs of Abuse
Esa R. Korpi, Bjrnar den Hollander, Usman Farooq, Elena Vashchinkina, Ramamoorthy Rajkumar, David J. Nutt, Petri Hyyti,
and Gavin S. Dawe
Department of Pharmacology, Faculty of Medicine, University of Helsinki, Finland (E.R.K., B.d.H., E.V., P.H.); Department of
Pharmacology, Yong Loo Lin School of Medicine, National University Health System, Neurobiology and Ageing Programme, Life Sciences
Institute, National University of Singapore, Singapore, and SINAPSE, Singapore Institute for Neurotechnology, Singapore (E.R.K., R.R.,
G.S.D.); Interdepartmental Neuroscience Program, Yale University, New Haven, Connecticut (U.F.); and Centre for
Neuropsychopharmacology, Division of Brain Sciences, Burlington Danes Building, Imperial College London, London.
United Kingdom (D.J.N.)

This project was partially funded by the Academy of Finland, the Sigrid Juselius foundation, the Finnish Foundation for Alcohol Studies,
the Jane and Aatos Erkko Foundation, the Foundations Professor pool, and the Orion Research Foundation.
Address correspondence to: Dr. Esa R. Korpi, Department of Pharmacology, Faculty of Medicine, Biomedicum Helsinki, POB 63
(Haartmaninkatu 8), FI-00014 University of Helsinki, Finland. E-mail esa.korpi@helsinki.fi or Gavin S. Dawe, Department of Pharmacology,
Yong Loo Lin School of Medicine, Building MD3, #04-01Y, 16 Medical Drive, Singapore 117600. Email: gavin_dawe@nuhs.edu.sg.
dx.doi.org/10.1124/pr.115.010967
872

Downloaded from pharmrev.aspetjournals.org at Nyu Med Centr Lib on October 8, 2015

Abstract. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 875
I. Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 875
A. Different Forms of Neuroplasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 876
B. Short-term Neuroplasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 878
C. Main Forms of Long-term Neuroplasticity: Long-term Potentiation and Long-term
Depression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 879
1. Presynaptic Forms of Long-term Plasticity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 880
2. Postsynaptic Forms of Long-term Plasticity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 880
3. a-Amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid and N-methyl-D-aspartate
Receptor Phosphorylation.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 880
4. a-Amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid and N-methyl-D-aspartate
Receptor Trafficking. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 881
D. Forms of Long-term Plasticity at Inhibitory Synapses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 881
E. Developmental Maturation of the Brain in Rodents and Humans . . . . . . . . . . . . . . . . . . . . . . . . . 882
F. Developmental Neuroplasticity: Critical Periods, Reopening of Plasticity in Adults . . . . . . . . 883
G. Neurotrophins as Regulators of Neuroplasticity: Brain-Derived Neurotrophic
Factor as an Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 885
H. Animal Models of Drug Reinforcement and Addiction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 885
II. Actions and Persistent Effects of Specific Drugs of Abuse. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 886
A. Cocaine, a Stimulant and Local Anesthetic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 886
1. Persistent Ventral Tegmental Area Neuroplasticity after a Single Dose. . . . . . . . . . . . . . . 887
2. Changes in the Nucleus Accumbens Mediate Cocaine Addiction and Relapse.. . . . . . . . . 890
3. Altered Gene Expression in the Nucleus Accumbens in Cocaine Seeking and Relapse. 893
4. Progressive Involvement of the Dorsal Striatum.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 895
5. Changes in Lateral Habenula and Rostromedial Tegmental Nucleus Contribute
to Aversive Symptoms of Cocaine Withdrawal. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 895
6. Complex Alterations in the PFC. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 896
7. Hippocampal Functional Changes.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 897
8. Bed Nucleus of Stria Terminalis and Amygdala are Involved in
Stress-Induced Reinstatement of Cocaine Seeking. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 897
9. Effects of Adolescent Cocaine Exposure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 898
10. Human Imaging Studies. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 898
11. Limited Evidence of Cocaine Neurotoxicity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 899

Drug-Induced Neuroplasticity

873

12. Potential Mechanisms in Cocaine Neurotoxicity and Neuroprotection. . . . . . . . . . . . . . . . . 900


13. Methylphenidate and Novel Psychoactive Substances.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 900
14. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 900
B. Amphetamine-type Psychostimulants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 900
1. Many Stimulant Amphetamines for Abuse and Therapy. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 902
2. Molecular Targets and Mechanisms of Action of Amphetamines. . . . . . . . . . . . . . . . . . . . . . 902
3. Comparison of the Mechanisms of Action of Amphetamine versus Methamphetamine. 902
4. Effects of Amphetamine at Plasmalemmal Dopamine Transporter. . . . . . . . . . . . . . . . . . . . 903
5. Regulation of Dopamine Efflux Via the Dopamine Transporter by Protein Kinase
C and Ca2+/Calmodulin Kinase II Phosphorylation, and by Reactive Oxygen Species. . 904
6. Effects on Secretory Vesicles. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 904
7. Other Amphetamine Targets: Monoamine Oxidase, Tyrosine Hydroxylase, other Monoamine Transporters, Trace Amine-Associated Receptor 1, and Sigma Receptors. . . . . . . 905
8. Action of Substituted Amphetamines and Cathinones. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 905
9. Action of Stimulants Common in Clinical Use. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 906
10. Efficacy and Addiction Liability of Amphetamines in Clinical Use. . . . . . . . . . . . . . . . . . . . 906
11. Neuroplasticity Related to Sensitization and Addiction to Amphetamines. . . . . . . . . . . . . 907
12. Glutamate and Amphetamine/Methamphetamine Reinforcement/Extinction/Reinstatement.. 908
13. Long-term Neuroadaptation/Neurotoxicity after Exposure to High Doses of
Psychostimulants. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 909
a. Effects on brain structure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 909
b. Effects on neurochemistry. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 910
c. Effects on behavior. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 910
14. Limited Evidence for Long-term Effects of Substituted Cathinones in Rodent
Models. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 911
15. Mechanisms Involved in Long-term Adaptation and Neurotoxicity. . . . . . . . . . . . . . . . . . . . 911
a. Oxidative stress: DA oxidation and drug metabolites as sources of reactive species.911
b. Mitochondrial dysfunction: ATP deficit, superoxide leakage, and apoptosis.. . . . . . . . 912

ABBREVIATIONS: AC, adenylyl cyclase; ADHD, attention deficit hyperactivity disorder; 2-AG, 2-arachidonoylglycerol; ALDH,
aldehyde dehydrogenase; AMPA, a-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid; AMPH, amphetamine; BDNF, brain-derived
neurotrophic factor; BLA, basolateral complex of the amygdala; BNST, bed nucleus of stria terminalis; BOLD, blood oxygen level
dependent; BZ, benzodiazepine; CAM, cell adhesion molecule; CaMKII, Ca2+/calmodulin kinase II; CCK, cholecystokinin; CeA, central
nucleus of the amygdala; CNS, central nervous system; CPP, conditioned place preference; CREB, cAMP response element-binding
protein; CRF, corticotropin-releasing factor; D1R, dopamine-1 receptor; D2R, dopamine-2 receptor; DA, dopamine; DARPP-32, DA- and
cAMP-regulated phosphoprotein; DAT, dopamine transporter; DBI, diazepem binding inhibitor; DGL, diacylglycerol lipase; DOI,
2,5-dimethoxy-4-iodoamphetamine; DSE, depolarization-induced suppression of excitation; DSI, depolarization-induced suppression of
inhibition; eCB, endocannabinoid; ECM, extracellular matrix; EEG, electroencephalography; EPSC, excitatory postsynaptic currents;
ER, endoplasmic reticulum; ERK, extracellular signal-regulated kinase; ETC, electron transport chain; FC, frontal cortex; fMRI,
functional magnetic resonance imaging; G9a, lysine dimethyltransferase G9a; GABA, g-aminobutyric acid; GHB, gammahydroxybutyrate; GPCR, G protein-coupled receptor; HC, hippocampus; HDAC, histone deacetylase; HFS, high-frequency stimulation;
HPPD, hallucinogen persisting perception disorder; 5-HT, 5-hydroxytryptamine, serotonin; ICSS, intracranial self-stimulation; iGluR,
iontropic glutamate receptor; IL, infralimbic cortical region; IP3, inositol 1, 4,5-trisphosphate; IPN, interpeduncular nucleus; IPSC,
inhibitory postsynaptic currents; IPSP, inhibitory postsynaptic potential; JNK, c-Jun N-terminal kinases; Kal-7, postsynaptic densitylocalized kalirin-7; KCC, potassium-chloride cotransporter; KO, knockout; LA, lateral nucleus of amygdala; LC, locus ceruleus; LDTg,
laterodorsal tegmental nucleus; LDX, lisdexamfetamine; LFS, low-frequency stimulation; LHb, lateral habenula; LSD, lysergic acid
diethylamide; LTD, long-term depression; LTP, long-term potentiation; MAO, monoamine oxidase; MAPK, mitogen-activated protein
kinase; MDMA, 3,4-methylenedioxymethamphetamine (ecstasy); MDMC, methylone; METH, methamphetamine; mGlu, metabotropic
glutamate receptor; MHb, medial habenula; MK-801, dizocilpine; 4-MMC, mephedrone; MPH, methylphenidate; MRI, magnetic
resonance imaging; MSN, medium spiny neuron; mTOR, mammalian target of rapamycin; NAc, nucleus accumbens; nAChR, nicotinic
acetylcholine receptors; NE, norepinephrine; NET, norepinephrine transporter; NFkB, nuclear factor kappa-B; NMDA, N-methyl-Daspartate; NO, nitric oxide; NOP, nociceptin receptor; N/OFQ, nociceptin/orphanin FQ; NPY, neuropeptide Y; OFC, orbitofrontal cortex;
OR, odds ratio; OX, orexin; P, postnatal day; PAG, periaqueductal gray area; PBR, peripheral benzodiazepine receptor; PCP,
phencyclidine; PET, positron emission tomography; PFC, prefrontal cortex; PI3K, phosphatidylinositide 3-kinase; PKA, protein kinase
A; PKC, protein kinase C; PKMzeta, atypical PKC; PLC, phospholipase C; PNN, perineuronal nets; PrL, prelimbic cortical region; PSANCAM, polysialylated neuronal cell adhesion molecule; PSD, postsynaptic density; Pv, parvalbumin; R, receptor; RGS, regulator of G
protein signaling; rmTg, rostromedial tegmental nucleus; ROS, reactive oxygen species; SA, self-administration; SERT, serotonin
transporter; SN, substantia nigra; TA1, trace amine-associated receptor 1; TH, tyrosine hydroxylase; THC, tetrahydrocannabinol;
TLR4, toll-like receptor 4; TPH, tryptophan hydroxylase; TrkB, tropomyosin receptor kinase B; TSPO, mitochondrial translocator
protein; VGCC, voltage-gated calcium channel; vHC, ventral hippocampus; VMAT-2, vesicular monoamine transporter 2; VTA, ventral
tegmental area.

874

Korpi et al.

C.

D.

E.

F.

G.

H.

c. Excitotoxicity: glutamate receptor involvement and intracellular excitotoxic


processes.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 913
d. Hyperthermia: a catalyst of other toxic processes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 913
e. Other
factors:
inflammation,
blood-brain
barrier
disruption,
protective
preconditioning. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 914
16. Evidence of Long-term Plasticity and Cognitive Effects after Stimulant Use in
Humans.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 914
a. Cognitive behaviors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 914
b. Brain imaging of receptors, transporters and activity/metabolism.. . . . . . . . . . . . . . . . . 915
c. Magnetic resonance imaging studies. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 915
d. Magnetic resonance spectroscopy and analysis of postmortem tissue.. . . . . . . . . . . . . . 916
e. Prospective and experimental studies. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 916
17. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 917
Nicotine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 917
1. Nicotinic Acetylcholine Receptors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 917
2. Upregulation of Nicotinic Receptors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 918
3. Nicotine-priming Effects and Upregulation of Catecholamine Synthesis. . . . . . . . . . . . . . . 919
4. Nicotine and Neuroplasticity within the Ventral Tegmental Area. . . . . . . . . . . . . . . . . . . . . 920
5. Nicotine and Neuroplasticity in the Lateral Hypothalamic Orexin System. . . . . . . . . . . . 921
6. Nicotine and Glutamatergic Neuroplasticity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 921
7. Nicotine and the Habenula. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 922
8. Long-term Sequelae of Adolescent Exposure to Nicotine. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 923
9. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 924
Neural Adaptations Induced by Ethanol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 925
1. Cell Membrane Ion Channels as Primary Targets of Ethanol. . . . . . . . . . . . . . . . . . . . . . . . . 925
2. Ethanol-Induced Changes in Glutamatergic Transmission. . . . . . . . . . . . . . . . . . . . . . . . . . . . 928
3. Ethanol-Induced Changes in g-Aminobutyric Acidergic Transmission. . . . . . . . . . . . . . . . . 929
4. Ethanol-Induced Changes in Neuropeptide Mechanisms. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 931
5. Role of Acetaldehyde in Ethanols Reinforcing Actions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 933
6. Structural Plasticity in Alcohol Dependence. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 933
7. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 934
Benzodiazepines and Other GABAergic Drugs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 934
1. Molecular Targets for Benzodiazepines.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 935
2. Neuroplasticity Induced by Benzodiazepines and Related Compounds. . . . . . . . . . . . . . . . 936
3. Behavioral After-effects of g-Aminobutyric Acid A Drugs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 939
4. Effects of Flumazenil on Benzodiazepine and Alcohol Tolerance. . . . . . . . . . . . . . . . . . . . . . 939
5. Treatment of Addictions by g-Aminobutyric Acid B Receptor Agonists. . . . . . . . . . . . . . . . 940
6. g-Hydroxybutyrate as a Drug of Abuse and a Therapeutic Compound.. . . . . . . . . . . . . . . . 941
7. Anesthetics and Neuroplasticity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 942
8. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 943
N-methyl-D-aspartate Receptor Antagonists . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 943
1. Ketamine: a Dissociative Anesthetic with Rapid Antidepressant Effects in Patients. . . 944
2. Phencyclidine. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 946
3. Dizocilpine, a Prototypic Noncompetitive N-methyl-D-aspartate Receptor Antagonist. . 948
4. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 949
Opioid-Induced Neural Adaptations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 949
1. Desensitization and Internalization in Opioid Tolerance and Dependence. . . . . . . . . . . . . 950
2. Cellular Signaling in Opioid Tolerance and Dependence. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 950
3. Novel Mechanisms in Between-systems Adaptations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 951
4. Opioid-Induced Changes in Excitatory and Inhibitory Synaptic Plasticity. . . . . . . . . . . . . 952
5. Chronic Opioids and Synaptic Plasticity.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 953
6. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 955
Cannabinoids: Multiple Mechanisms and Possible Indications . . . . . . . . . . . . . . . . . . . . . . . . . . . . 955
1. Endocannabinoid System as an Endogenous Lipid Messenger System.. . . . . . . . . . . . . . . . 956
2. Short-term Plasticity Involving Retrograde Endocannabinoid Signaling. . . . . . . . . . . . . . . 957
3. Long-term Plasticity and CB1 Receptors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 958
4. Cannabinoid Effects on Brain Development. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 958

Drug-Induced Neuroplasticity

875

5. Cannabinoid-Induced Cognitive Impairment. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 959


6. Rewarding and Aversive Behaviors, Specific Actions on VTA DA Neurons. . . . . . . . . . . . 960
7. Cannabinoids during Adolescence: Increased Risk for Schizophrenia?. . . . . . . . . . . . . . . . . 961
8. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 962
I. Hallucinogens . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 962
1. Effects of Serotonergic Hallucinogens Are Mediated by the 5-HT2A Receptor. . . . . . . . . . 962
2. Hallucinogen ActionA Perturbation of Sensory Gating or Desynchronization
of Cortical Rhythms?... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 963
3. Behavioral Effects and Addiction Potential. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 964
4. Long-term Residual Effects of Serotonergic Hallucinogens: the Role of
Neuroplasticity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 964
a. Sensory processing. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 964
b. Mood and anxiety. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 965
5. Scopolamine: Another Hallucinogen Revisited for Depression. . . . . . . . . . . . . . . . . . . . . . . . . 966
6. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 966
III. General Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 966
A. Novel Methods for Future Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 967
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 969

AbstractAdaptation of the nervous system to


different chemical and physiologic conditions is important for the homeostasis of brain processes and for
learning and remembering appropriate responses to
challenges. Although processes such as tolerance and
dependence to various drugs of abuse have been known
for a long time, it was recently discovered that even
a single pharmacologically relevant dose of various
drugs of abuse induces neuroplasticity in selected
neuronal populations, such as the dopamine neurons
of the ventral tegmental area, which persist long after
the drug has been excreted. Prolonged (self-) administration of drugs induces gene expression, neurochemical, neurophysiological, and structural changes in
many brain cell populations. These region-specific
changes correlate with addiction, drug intake, and

conditioned drugs effects, such as cue- or stressinduced reinstatement of drug seeking. In rodents,
adolescent drug exposure often causes significantly
more behavioral changes later in adulthood than
a corresponding exposure in adults. Clinically the
most impairing and devastating effects on the brain
are produced by alcohol during fetal development. In
adult recreational drug users or in medicated patients,
it has been difficult to find persistent functional or
behavioral changes, suggesting that heavy exposure to
drugs of abuse is needed for neurotoxicity and for
persistent emotional and cognitive alterations. This
review describes recent advances in this important
area of research, which harbors the aim of translating
this knowledge to better treatments for addictions and
related neuropsychiatric illnesses.

I. Introduction

and ethnic groups studied in the United States. Individuals with alcohol or illicit drug abuse or dependence
in the past year constituted 5% of the adolescents aged
between 12 and 17 and more than 8% of all individuals
aged 12 or older (SAMSHA, 2014a). Of the racial/ethnic
groups studied among those aged 12 and older, Asians
had the lowest proportion of binge alcohol drinkers and
past-month illicit drug users, likely partly due to
societal/cultural traditions. These findings on rather
widespread drug and alcohol exposures at young ages
are very alarming, because brain development continues well past the age of 17 years and because the
exposure to several different drugs, such as illicit drugs,
cigarettes, and alcohol, appears to concentrate on the
same individuals (SAMSHA, 2014b). Therefore, it is
possible that harmful effects from early drug use will
prevail later in life, because drugs of abuse induce
different modulations in brain circuitries (adaptation,
plasticity, learning, and memory) due to their pharmacological actions and due to behavioral/social effects
associated with their use and settings. Thus, we will

Brain diseases are associated with an enormous cost


to affected individuals, their families, and the society. In
Europe, it has been estimated that the total yearly cost
of brain diseases in 2010 was close to 800 billion euros
(Olesen et al., 2012), of which addictions are responsible
for as much as anxiety disorders, with only dementia
and mood illnesses costing more (DiLuca and Olesen,
2014). Drug abuse produces both direct and indirect
costs to the society, although many of the drugs are also
clinically used to treat various patient groups. The
purpose of this review is to present up-to-date knowledge of the mechanisms of action of the main drugs of
abuse and to reveal the possible long-term alterations in
the nervous system associated with the use and abuse of
various drugs acting on the brain, also paying attention
to the trajectory of brain development.
Addiction is a complex phenomenon, which is not only
dependent on pharmacological mechanisms, but also
has a societal/cultural dimension. This is reflected in
the proportion of drug addiction among different age

876

Korpi et al.

explore the data on different drugs and how they


persistently affect brain functions if the drug exposure
occurs during adolescence.
The review first gives a general basic introduction to
neuroplasticity. Then sections on different drugs follow
and they have different focuses, because the various
drugs do not invoke exactly similar mechanisms or
adaptations in the nervous system. The review will end
with a short general summary.
Certain mechanisms seem to be common for many
drugs, for example, activation of the extracellular
signal-regulated kinase (ERK) pathway in specific brain
structures is necessary for effects of and tolerance to
cocaine, nicotine, MDMA, phencyclidine, alcohol, and
cannabinoids after both acute and chronic treatments in
rodents (Kyosseva et al., 2001; Salzmann et al., 2003;
Valjent et al., 2004; Rubino et al., 2005; Tonini et al.,
2006; Schroeder et al., 2008). One human postmortem
brain study suggested that cocaine, cannabis, and/or
phencyclidine abuse all decrease transcription of
calmodulin-related genes and increase transcription of
genes related to lipid/cholesterol and Golgi/endoplasmic
reticulum (ER) function in the anterior prefrontal cortex
(PFC), which may underlie changes in synaptic function
and plasticity (Lehrmann et al., 2006). However, postmortem brain regional gene expression profiling in
alcohol-dependent patients has indicated widely differing sets of affected genes between different brain
regions (Flatscher-Bader et al., 2005, 2006), with alcoholics nevertheless being easily separable from nonalcoholic controls and smokers (Flatscher-Bader et al.,
2010). The same situation was found for other drugs of
abuse (Albertson et al., 2004, 2006). Significant alterations in glutamate and GABA receptor mRNAs were
found in postmortem brains of alcohol-dependent subjects, but the changes differed from one brain region to
another (Jin et al., 2011, 2014a,b; Bhandage et al.,
2014). For these obvious reasons, we reviewed the
literature for each drug according to the specific brain
regions where alterations have been observed most
commonly in preclinical experiments. These brain areas
and pathways, which serve for positive and negative
reinforcing behavioral and emotional effects and for
goal-directed and habitual drug seeking, are illustrated
in Fig. 1. However, it will be necessary in the future to
study the plasticity and mechanisms more precisely
at the level of different neuronal populations and
subpopulations.
As is usual in addiction-related reviews and research,
the dopamine (DA) mechanisms are center stage. The
reader is referred to recent reviews on various aspects
of DA as a neurotransmitter and a regulator of motor,
cognitive, and motivated behavior (Bjorklund and Dunnett,
2007; Beaulieu and Gainetdinov, 2011; Salamone and
Correa, 2012). DA is important for a wide number of
brain functions, and it will be impossible to cover each
and every aspect of the DA mechanisms in drug actions

and adaptations. The focus in this review will be more


on the neuroplasticity of the glutamate synapses on DA
neurons.
Pharmacological actions of drugs are mediated by
specific receptors. On the other hand, the effects of the
drugs of abuse are often modulated by experimental
settings and expectancies in both preclinical studies
and human experiments. Drug intake in rodents is
dependent, for example, on cage conditions, with the
effects differing between opioids and stimulants (Badiani
et al., 2011). Voluntary self-administration (SA) (selfstimulation or cocaine) induces different brain regional
activations (Porrino et al., 1984), or more prolonged
synaptic molecular adaptations, than experimentergiven stimulation or drug injections (Chen et al., 2008).
Placebo/nocebo effects are real in human studies, and the
expectancy of strong or uncertain drug effects are known
to affect human brain imaging results in response to
acute drugs (Volkow et al., 2010). For example, in
smokers, positive and negative beliefs on nicotine content
of cigarettes influenced functional magnetic resonance
imaging (fMRI)-scanned striatal responses to value and
reward prediction errors during an investment task (Gu
et al., 2015), indicating that beliefs can affect cognitive
performance also under the drug-associated states. These
issues indicate that, in addition to direct pharmacological
actions, abused drugs may also change basic learning and
memory processes, for example, by conditioning and
environmental factors.
A. Different Forms of Neuroplasticity
The ability of the brain to remodel its connections
functionally and structurally in response to individual
experience has been described by the concept of neuroplasticity. Neuroplasticity occurs on a variety of levels
ranging from molecular changes in synapses to largescale changes involved in neurocircuitry remapping.
Synaptic plasticity refers to adaptive changes in the
strength of synaptic connections. On the basis of its time
frame, synaptic plasticity has been classified as shortterm (acts on a timescale of milliseconds to minutes)
and long-term (hours to days) plasticity. Short-term
plasticity is achieved through transient changes, such
as facilitation or depression of a synaptic connection,
which then quickly return to their initial state. However, repeated stimulation causes a persistent change
in the connection to achieve long-term plasticity.
Hebbian or activity-dependent plasticity is the most
studied form of long-term plasticity. It occurs when
presynaptic stimulation coincides with postsynaptic
depolarization (Hebb, 1949; Bi and Poo, 2001). The
best-known example of Hebbian plasticity is N-methylD-aspartate receptor (NMDAR)-mediated long-term potentiation (LTP). Importantly, this form of LTP occurs
only in synapses that actively contribute to the induction process, so it is input and synapse specific. The
term anti-Hebbian plasticity currently describes

Drug-Induced Neuroplasticity

877

Fig. 1. Brain regions of major importance for the acute and chronic addictive effects of drugs of abuse. The top drawing shows many pathways that
participate in rewarding/drug-seeking behavior. The middle drawing depicts several pathways/brain regions that are particularly related to addictionrelated aversive behaviors (wide connections of the locus ceruleus to other brain areas are not shown). The bottom drawing illustrates the role of midbrainstriatal-cortical loops in goal-directed and habitual drug taking. See Fig. 10 for additional pathways involved in nicotine neurotoxicity. Amy, amygdala;
BNST, bed nucleus of stria terminalis; CeA, central nucleus of amygdala; DLS, dorsolateral striatum; DMS, dorsomedial striatum; HC, hippocampus; LC,
locus ceruleus; LDTg, laterodorsal tegmental nucleus; LH, lateral hypothalamus; LHb, lateral habenula; mPFC, medial prefrontal cortex; NAc, nucleus
accumbens; OFC, orbitofrontal cortex; rmTg, rostromedullary tegmental nucleus; SN, substantia nigra; VTA, ventral tegmental area.

either long-term depression (LTD) (Nelson, 2004) or LTP


that occurs when presynaptic activation coincides with
postsynaptic inactivity (Kullmann and Lamsa, 2007).
In rodents, the neuronal organization remains immature at birth (see below). The process that describes
changes in neuronal organization during development as a result of environmental interactions and
experience/learning-induced neural changes is known
as developmental plasticity. To maintain the balance
between neuronal excitation and inhibition, homeostatic plasticity regulates the overall activity of complex
circuits by specifically regulating the destabilizing

effects of developmental and learning processes. Otherwise, activity-dependent forms of plasticity could drive
neural activity toward runaway excitation or quiescence (Miller, 1996; Turrigiano, 1999; Turrigiano and
Nelson, 2004).
Metaplasticity refers to plasticity of synaptic plasticity, which describes changes in the ability to induce
further synaptic plasticity (Abraham and Bear, 1996).
For example, prolonged exposure to cocaine induces
a population of silent glutamatergic synapses in the
nucleus accumbens (NAc) that form sites for future
plasticity (reviewed in Lee and Dong, 2011).

878

Korpi et al.

To help appreciate the effects of abused drugs on


synaptic plasticity, in the following paragraphs we will
briefly introduce mechanisms of the most common
presynaptic and postsynaptic forms of synaptic
plasticity.
B. Short-term Neuroplasticity
Short-term plasticity can appear as a transient facilitation, depression, or augmentation and posttetanic
potentiation in synaptic strength that lasts for up to
a few minutes (reviewed in Fioravante and Regehr,
2011). Despite the variety in synaptic neurotransmitters, all forms of short-term plasticity are primarily
governed by presynaptic mechanisms associated with
fluctuations of presynaptic residual Ca2+, which acts on
one or more molecular targets, resulting in the changes
in neurotransmitter release (Fig. 2). Enhancement of

Ca2+ channel activity and increases in the probability of


Ca2+ influx, altered vesicle pool properties, local depletion of Ca2+ buffers, and increases in quantal size of
neurotransmitter release contribute to short-term facilitation (Fioravante and Regehr, 2011). On the other
hand, vesicle depletion, inactivation of neurotransmitter release sites, and Ca2+ channels contribute to shortterm synaptic depression (Neher and Sakaba, 2008). In
addition, glial-neuronal interactions impact on shortterm synaptic plasticity by controlling the speed and
extent of neurotransmitter clearance from the synaptic
cleft (Bergles et al., 1999) as well as by astroglial release
of substances that can affect synaptic efficacy (Araque
et al., 2001). Most importantly, there is retrograde
communication between post- and presynaptic terminals: both endocannabinoids (reviewed in Wilson
and Nicoll, 2002; Kano et al., 2009) and nitric oxide

Fig. 2. Typical induction protocols and main factors regulating mechanisms of short-term plasticity. (A) Brief paired-pulse (PP) stimulation induces
short-term facilitation of neurotransmission by transiently increasing the Ca2+-dependent readily releasable pool (RRP) of synaptic vesicles. (B) Shortterm depression of neurotransmission can be induced by frequent tetanic stimulation, which transiently depletes synaptic vesicles. (C). Depolarizationinduced suppression of inhibition (DSI) or similarly that of excitation (DSE; not shown) is a locally induced transient depression of neurotransmission
that is dependent on retrograde eCB signaling. Depolarization-induced suppression of inhibition/excitation begins with stimulation of excitatory
neuronal connections inducing eCB synthesis, and 2-AG in particular then moves to activate presynaptic CB1 receptors on surrounding neurons. This
induces local, transient depression of inhibitory or excitatory (not shown) neurotransmission.

Drug-Induced Neuroplasticity

(NO)-guanylyl cyclase signaling (Sammut et al., 2010)


contribute in a general transient local process that
tunes neurotransmitter release and different aspects of
synaptic dynamics both in inhibitory and excitatory
synapses (discussed in more detail in sections II.G and
II.H). Short depolarization of a neuron may cause
a transient suppression of excitation or inhibition of
that and neighboring neurons, called depolarizationinduced suppression of excitation or inhibition (DSE

879

and DSI), which was found to be dependent on retrograde encocannabinoid (eCB) signaling (see section II.H
on cannabinoids).
C. Main Forms of Long-term Neuroplasticity:
Long-term Potentiation and Long-term Depression
Long-term forms of synaptic plasticity can appear as
potentiation (long-term potentiation, LTP) or depression (long-term depression, LTD) in synaptic strength

Fig. 3. Main factors regulating pre- and postsynaptic plasticity mechanisms, such as LTP and LTD, after most typical induction protocols. (A)
Presynaptic LTP is often measured using paired-pulse (PP) stimulation e.g., at 50-ms interstimulus intervals (ISI). This leads to presynaptic Ca2+
influx via VGCCs, activation of adenylate cyclase (AC) and phosphorylation of synaptic vesicular proteins, such as Rab3a and RIM1a, leading to
increased transmitter release. (B) During low-frequency stimulation (LFS; often used for LTD induction) weak presynaptic depolarization triggers only
a modest Ca2+ influx that activates phosphatases (calcineurin, protein phosphatase 1, and protein phosphatase 2) in the postsynaptic cell. These
phosphatases dephosphorylate AMPA and NMDA receptors, thus promoting receptor removal from the membrane. (C) During high-frequency
stimulation (HFS; often used for LTP induction) strong presynaptic depolarization triggers robust glutamate release promoting Ca2+ influx via AMPA
and NMDA receptors into the postsynaptic cell. This Ca2+ influx activates protein kinases (FYN, PKA, PKC, PKMz, and CaMKII) that phosphorylate
receptors and promote stabilization of receptors at the membrane. These kinases also activate local protein synthesis, which leads to insertion of new
receptors in the membrane. (D) By inducing robust glutamate release with HFS up to 300 Hz, activation of postsynaptic group I mGlu receptors leads
to AMPA receptor internalization. (E) The LTP that is dependent on retrograde endocannabinoid (eCB) signaling consists postsynaptic activation that
induces eCB synthesis, especially increasing 2-AG, which then acts presynaptically on CB1 receptors limiting transmitter release (for more details, see
Fig. 12). This form of plasticity can be experimentally evoked, e.g., in the striatal neurons by 100-Hz HFS.

880

Korpi et al.

that last for hours to weeks. In contrast to short-term


plasticity, the nature of long-term forms of plasticity
involves both pre- and postsynaptic alterations (Fig. 3).
1. Presynaptic Forms of Long-term Plasticity.
Presynaptic LTP has been best studied at hippocampal
CA3 mossy fiber synapses (Weisskopf et al., 1994; Nicoll
and Schmitz, 2005), but similar forms of LTP have been
found in multiple brain areas (Salin et al., 1996; CastroAlamancos and Calcagnotto, 1999; Lopez de Armentia
and Sah, 2007). This presynaptic LTP appears at both
excitatory and inhibitory synapses and does not require
postsynaptic NMDARs (but see Yeckel et al., 1999).
Instead, presynaptic LTP appears to be induced by an
activity-dependent rise of presynaptic residual calcium
that results in a rise in cAMP and subsequent activation
of protein kinase A (PKA). This, in turn, modifies the
functions of proteins that act to coordinate synaptic
vesicle interactions with the presynaptic active zone,
leading to a long-lasting increase in neurotransmitter
release (Castillo, 2012).
Endocannabinoid-mediated long-term depression
(eCB-LTD) is a widely expressed form of long-term
plasticity at both excitatory and inhibitory synapses.
Brief robust neuronal stimulation triggers the synthesis
of eCBs, lipophilic molecules that travel retrogradely
across the synapse to activate the presynaptic CB1
cannabinoid receptors (see section II.H), which suppresses neurotransmitter release via a wide range of
effector molecules, including voltage-dependent calcium channels (VGCC), potassium channels, PKA, p38
mitogen-activated protein kinase (MAPK), and c-Jun
N-terminal kinases (JNK) (reviewed in Howlett et al.,
2002). Importantly, CB1 receptors activation per se is
not sufficient for eCB-LTD induction. Rather, the presynaptic terminal integrates multiple signals to generate eCB-LTD (Heifets and Castillo, 2009). CB1 receptors
also mediate short-term plasticity of DSI and DSE (Fig.
2). Interestingly, some hippocampal inhibitory synapses can undergo both short- and long-term forms of eCBmediated plasticity, where the time frame of depression
(short-term versus long-term manner) depends on the
downstream signaling pathways (Heifets and Castillo,
2009). Particularly, cAMP/PKA-dependent signaling
has been shown to be necessary only for eCB-LTD but
not for DSI (Chevaleyre and Castillo, 2003).
2. Postsynaptic Forms of Long-term Plasticity.
NMDAR-dependent LTP and LTD are the best understood forms of long-lasting synaptic plasticity (Bliss
and Lomo, 1973; Hayashi et al., 2000). Early phases of
LTP and LTD are mediated by a redistribution of both
a-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid
receptors (AMPAR) and NMDARs at the postsynaptic
membrane and/or by changes in presynaptic transmitter release. With time, such changes are consolidated by
structural alterations, which require synthesis of new
proteins (Kasai et al., 2010).

In in vitro experiments on acute brain slices, LTP and


LTD can be induced by distinct patterns of activity
(reviewed in Holscher, 1999; Luscher and Malenka,
2012) (Fig. 3). High-frequency stimulation (HFS) of the
presynaptic cell (tetanic pulses at 50100 Hz) is commonly used to induce LTP. This stimulation protocol
causes a strong postsynaptic depolarization that removes
the Mg2+ block of NMDARs, allowing timed Ca2+ influx.
This triggers the downstream molecular cascades inducing LTP. In contrast, low-frequency stimulation (LFS,
at 0.15 Hz) often induces LTD. Typically, it causes only
a weak postsynaptic depolarization that results in a modest but prolonged Ca2+ influx triggering the downstream
molecular cascades driving to LTD.
Since the original discovery of LTP in the hippocampus (HC) (Bliss and Lomo, 1973), LTP and LTD have
been observed in a variety of other brain regions
including the ventral tegmental area (VTA), NAc,
PFC, and amygdala ex vivo (Luscher and Malenka,
2011) and in vivo (Canals et al., 2009; Zhang et al.,
2015b). Different brain regions appear to exhibit different forms of LTP and LTD, and therefore, synapses
recruit different signaling pathways to accomplish their
functions. The most studied forms of neuroplasticity are
NMDAR-dependent LTP and LTD, with the NMDARs
providing the major pathway for Ca2+ influx (Huang
and Kandel, 1996). There is also NMDAR-independent
LTP and LTD, in which VGCC (Kato et al., 2009) or
GluA2 subunit-lacking AMPARs (Lamsa et al., 2007)
provide the Ca2+ influx triggering induction.
In the early phases of LTP, elevated Ca2+ triggers
persistent activation of protein kinases including PKA,
Ca2+/calmodulin kinase II (CaMKII), and protein kinase C (PKC). A striking feature of CaMKIIa is its
capacity for autophosphorylation at threonine residue
Thr286, which keeps this kinase activated even in the
absence of Ca2+ (Giese et al., 1998). During this stage,
atypical protein kinase C (PKMz) may also become
autonomously active (Ling et al., 2006; Sacktor, 2008).
Autonomously active and other protein kinases use
phosphorylation to carry out the two major mechanisms
underlying the expression of LTP: first, they phosphorylate existing AMPARs and NMDARs to increase their
activity, and second, they mediate the insertion of new
receptors into the postsynaptic membrane (see below for
more detailed description).
3. a-Amino-3-hydroxy-5-methyl-4-isoxazolepropionic
acid and N-methyl-D-aspartate Receptor Phosphorylation.
Huganir and coworkers made detailed experiments
on the regulation of AMPAR trafficking and function
by phosphorylation during LTP/LTD (reviewed in
Shepherd and Huganir, 2007). During LTP, PKA and
CaMKII are recruited to phosphorylate serine residues
in the GluA1 subunit at Ser831 (Mammen et al., 1997)
and Ser845 (Roche et al., 1996) and the GluA2 subunit
at Ser880 (Chung et al., 2000), promoting receptor
insertion and synaptic potentiation (Malenka and Bear,

Drug-Induced Neuroplasticity

2004; Huganir and Nicoll, 2013). Inhibition of PKA and


CaMKII or removal of the above-mentioned phosphorylation sites from AMPAR subunits can impair LTP
(Lee et al., 2003; Malenka and Bear, 2004). Importantly,
knock-in mutant mice that lack both Ser831 and Ser845
phosphorylation sites on GluA1 subunits demonstrated
impaired memory in behavioral tests (Olivito et al.,
2014).
Conversely, during LTD, phosphatases calcineurin
and protein phosphatases 1 and 2 are recruited to
dephosphorylate AMPARs, promoting their removal
from the synapse (Malenka and Bear, 2004; Huganir
and Nicoll, 2013).
Phosphorylation of NMDARs is also essential during
LTP/LTD expression because it contributes to stabilization of the receptors at the synapse by forming more
stable binding to postsynaptic density (PSD) proteins.
In fact, Fyn and Src kinases phosphorylate GluN2B
subunit at Tyr1472, which prevents the clathrin adaptor protein AP-2 from binding to the GluN2B subunit,
thus blocking endocytosis (Zhang et al., 2008).
4. a-Amino-3-hydroxy-5-methyl-4-isoxazolepropionic
acid and N-methyl-D-aspartate Receptor Trafficking.
During LTP, new GluA2 subunit-lacking AMPARs are
rapidly transported to the cell surface under the influence of protein kinases (Malinow and Malenka, 2002;
Luscher and Malenka, 2012). Importantly, these new
AMPARs exhibit some unique properties: first, they
have greater single-channel conductance and, second,
greater permeability to Ca2+ ions, which facilitates
Ca2+-dependent signaling events. Noteworthy, this
early insertion of new Ca2+-permeable non-GluA2containing AMPARs to the synapse is independent of
protein synthesis. This is achieved by having a nonsynaptic pool of AMPARs adjacent to the postsynaptic
membrane (Kauer and Malenka, 2007). Moreover,
NMDARs are also actively replaced during LTP/LTD
(Barria and Malinow, 2002). However, the data on
NMDARs are still puzzling due to wide variability in
expression, kinetics, and regulation sites (reviewed in
Shipton and Paulsen, 2014). Particularly interesting for
LTP expression are the GluN2B subunit-containing
NMDARs because of their special association with CaMKII. This interaction maintains CaMKII close to its
substrates, such as AMPARs, to initiate the phosphorylation events that support synaptic strengthening.
Postsynaptic LTD is associated with receptor withdrawal from the synapses (Hayashi et al., 2000; Bellone
and Luscher, 2006). LTD can also be generated by
activation of VGCCs (Christie et al., 1997) or metabotropic glutamate receptors (mGlu) (Bellone and Luscher,
2006) and does not require NMDAR activation. Although
the above cellular model of LTP/LTD presented above is
explained entirely by its postsynaptic mechanisms, additional events may also occur presynaptically, for example,
via alteration in probability of glutamate vesicle release

881

by different retrograde signaling mechanisms (reviewed


in Luscher and Malenka, 2012).
With time, the insertion/removal of additional receptors is likely to result in the rearrangement of the PSD
with different scaffolding proteins, spine ultrastructure, or even spine density that requires gene transcription and protein synthesis. In fact, spines associated
with synapses that underwent LTP were found to
be enlarged (Matsuzaki et al., 2004; Holtmaat and
Svoboda, 2009; Kasai et al., 2010), whereas spines that
underwent LTD were shrunken or even disappeared
(Nagerl et al., 2004; Kasai et al., 2010). Although these
structural changes are not absolutely required for
neuroplasticity, the stabilization and maintenance of
existing synapses needs activity of several kinases such
as PKA, CaMKIV, PKMz, and ERK that induce protein
synthesis either locally in the dendrites from prefabricated mRNAs or by nuclear transcription (Sacktor,
2008). These mechanisms are also involved in synaptic
tagging, an activity-driven molecular labeling of synapses to be strengthened (Frey and Morris, 1997; Sajikumar
and Frey, 2004; Sajikumar and Korte, 2011).
D. Forms of Long-term Plasticity at
Inhibitory Synapses
Several forms of NMDA receptor-dependent neuroplasticity (both LTP and LTD) have been described at
inhibitory GABAergic synapses (reviewed in Castillo
et al., 2011). Depending on interneuron subtype and
brain region, inhibitory synaptic plasticity results in
changes either in presynaptic GABA release or in
postsynaptic g-aminobutyric acid A (GABAA) receptor
responsiveness. Interneurons also show glutamatergic
synapse plasticity, which has characteristics that are
partly different from those of principal neurons
(reviewed in Kullmann and Lamsa, 2011).
Presynaptic inhibitory plasticity is mediated via retrograde messengers produced in an activity-dependent
manner (by afferent stimulation of excitatory inputs) in
the postsynaptic neuron and transferred back across the
synapse to modulate presynaptic GABA release. The most
studied presynaptic form of the plasticity is the eCBmediated inhibitory LTD, in which released eCBs inhibit
presynaptic GABA release by acting on CB1 cannabinoid
receptors (Chevaleyre et al., 2006; Adermark et al., 2009)
(Fig. 3). Another form of inhibitory plasticity, NOmediated inhibitory LTP, requires NO as a retrograde
messenger to stimulate presynaptic guanylyl cyclase,
resulting in the potentiation of presynaptic GABA release
(Nugent and Kauer, 2008). Brain-derived neurotrophic
factor (BDNF)-tropomyosin receptor kinase B (TrkB)mediated inhibitory LTP recruits BDNF that activates
its presynaptic receptor TrkB and potentiates GABA
release (Xu et al., 2010). It is evident that GABAergic
synapses from interneuron populations are very plastic
and actively participate in brain circuit refinement, learning, and memory formation (Chen et al., 2015).

882

Korpi et al.

Postsynaptic inhibitory plasticity is also mediated by


diverse mechanisms in different synapses. Similar to
glutamatergic receptors, phosphorylation of postsynaptic GABAA receptors by protein kinases, including PKA,
PKC, CaMKII, and Src, leads to changes in the integral
anion channel function (reviewed in Nakamura et al.,
2015). The effects of phosphorylation are dynamic
(increases and decreases in receptor activity) and dependent on the subunit composition of the GABAA
receptor in a fashion that is not fully known at present.
Moreover, phosphorylation of cation-chloride cotransporters such as Na+-Cl2 cotransporters, Na+-K+-2Cl2
cotransporters, and K+-Cl2 cotransporters (KCCs)
affects their activity and consequently the amplitude
of GABAA receptor-mediated responses by altering the
transmembrane anion gradients (reviewed in Kaila
et al., 2014). The function of GABAA receptors is further
dynamically regulated by constitutive exo- and endocytosis of receptors at GABAergic synapses (reviewed in
Michels and Moss, 2007).
E. Developmental Maturation of the Brain in Rodents
and Humans
Brain development is a delicate and complicated process
that can be adversely affected by drugs during pregnancy,
leading to teratogenic effects (reviewed in Manent et al.,
2008; Jutras-Aswad et al., 2009; Holbrook and Rayburn,
2014; Ross et al., 2015). In the present review, there will be
only short notes on drug actions on fetal brain development, such as fetal alcohol effects during pregnancy or
early postnatal period in rodents (Valenzuela et al., 2012).
A thorough comparative analysis of early brain maturation in different species, including rodents and human,
recently appeared (Workman et al., 2013). It is possible
that vulnerability for drug abuse may be greater the
younger the exposure/use of drugs is started.
Developmental neuroplasticity is based on genetic
programs in addition to activity-dependent processes,
and after the maturation, the neuroplasticity is expected
to be more homeostatic, dependent on molecular interactions in brain cells, neurites, spines, and synapses
within a relatively rigid extracellular milieu. The relevant question to the review of drug-induced neuroplasticity is the vulnerability of the brain to drug effects
during periods of childhood, adolescence, and adulthood.
These periods are often defined in humans as adolescence, starting at 10 to 12 years of age and ending at 18 to
20 years. Childhood precedes it and adulthood comes
after it, and these periods can be further subdivided into
shorter periods by using behavioral and brain maturation
indexes (Bossong and Niesink, 2010; Pressler and Auvin,
2013; Semple et al., 2013). For the present review an
important aspect to consider is how rodent models at
specific postnatal ages compare with human childhood,
adolescence, and adulthood. Figure 4 gives a rough
estimation of how to relate differently timed rodent
experiments to human brain developmental stages. The

Fig. 4. Phases of brain development and comparable ages in humans and


rodents (modified from Semple et al., 2013 and Flurkey and Currer, 2004).
(A) Number of spines (per 50 mm of the apical dendrite) of the human
prefrontal cortical pyramidal neurons in layers IIIc and V are shown by
curves estimated from (Petanjek et al., 2011). The figure illustrates normal
behavioral features (in boxes) and the onsets of attention deficit
hyperactivity disorder (ADHD) and schizophrenia. It is important to note
that puberty in female rats takes place roughly from P28 (vaginal opening)
to the first estrus cycle on P40 and in male rats from P40 (balanopreputial
separation) to spermatozoa maturation on P60, respectively (Schneider,
2008). (B) Postnatal development of the number of neurons (yellow traces)
and nonneuronal cells (blue) for whole rat brain (WB), cerebral cortex (Ctx,
with the scale in parentheses), and cerebellum (Cb) (Bandeira et al., 2009;
Mortera and Herculano-Houzel, 2012) indicates rapid changes in rodents
during the first 3 postnatal weeks, the cell numbers remaining rather
stable after that. (C) Timelines for the brain developmental processes
important for the functions of neuronal and various glial cells.

figure depicts various brain cell developmental processes,


emergence of selective behavioral phenotypes, and the
development of the behavioral modulation of brain processes by external stimuli in humans and rodents.
Several precautions have to be kept in mind, though.
First, there is a large variation between the developmental program timescale for different brain regions,
with the cortical regions and the cerebellum developing
last (Huttenlocher, 1979; 1990; Huttenlocher and
Dabholkar, 1997). Second, there are genetic, sex, and
disease vulnerability differences between the phases of
brain development (Thompson et al., 2001; Vasileiadis
et al., 2009; Asato et al., 2010; Brenhouse and Andersen,
2011; Dennis and Thompson, 2013). Third, various
neurotransmitter systems develop at different timelines

Drug-Induced Neuroplasticity

in different brain regions, with the DAergic and


GABAergic systems of the cortex developing late (Tseng
et al., 2007; Hashimoto et al., 2009; Kilb, 2012; Manitt
et al., 2013; Ouellet and de Villers-Sidani, 2014). Fourth,
various neurotrophic factors and their receptors have
differential expression profiles in various brain regions
(Perovic et al., 2013). Thus, the prolonged effects of
drugs of abuse are also dependent on the developmental
time of the exposure to drugs (Stanwood and Levitt,
2004). Furthermore, activating and stress-relieving
effects including neonatal handling, maternal care,
juvenile play, enriched environment, social contacts or
lack thereof, and lesioning effects on the brain also have
long-term effects and their critical times (Meaney et al.,
1988; Liu et al., 1997; Caldji et al., 1998; Bouwmeester
et al., 2002; Sale et al., 2007; Tseng et al., 2009; Foscarin
et al., 2012). A good example is the differential activation of the caudate-putamen, NAc, and PFC of rats at
different ages by a stress-inducing dose of the GABAA
receptor benzodiazepine-site inverse agonist FG-7142
(Lyss et al., 1999) known to induce a stress-like activation
of the mesocortical DA system in adult rats. The DA
system has a delayed maturation in the cortex, becoming
mature by postnatal day 60 (Kalsbeek et al., 1988). At
postnatal day 10, only the NAc is strongly activated by
FG-7142, at day 18 all regions start to be activated, and
finally at 45 and 100 days of age, the PFC is the brain
region that is activated the most.
Potential for neuroplasticity is greater in young than
adult rodents (Hensch, 2005; Bavelier et al., 2010;
Hensch and Bilimoria, 2012), as is demonstrated, for
example, by easier induction of the NMDAR-dependent
LTP using tetanic stimulation in the NAc neurons of
3-week-old than adult mice (Schramm et al., 2002).
Furthermore, the brain growth-promoting and limiting
responses to injury are blunted and dysregulated in
elderly (2224 months) versus adult (24 months) rats
(Li and Carmichael, 2006; Li et al., 2010b).
Perhaps an extreme illustration of the prolonged
neurodevelopmental process and normal potential for
neuroplasticity in humans may be the profiles of layer
IIIc and V spine densities in normal PFC (Petanjek
et al., 2011), which is illustrated in Fig. 4A. The spine
densities were determined using Golgi staining of postmortem brain sections at various ages. The peak density
of dendritic spines is seen at the age of 512 years, after
which the density slowly reduces to stable levels after
30 years of age. This extended spine-pruning period is
consistent with the results of human functional magnetic resonance imaging and connectivity studies that
have indicated prolonged developmental changes in the
cortical areas (Thompson et al., 2005; Knickmeyer et al.,
2008; Dosenbach et al., 2010). A corresponding delayed
development of the structural connectivity between basolateral amygdala and mPFC has been observed in the rat
(Cunningham et al., 2002), with a wide range of different
experiences, including the effects of various drugs of abuse,

883

known to induce long-lasting plasticity in the PFC also in the


adulthood (Kolb and Gibb, 2015).
F. Developmental Neuroplasticity: Critical Periods,
Reopening of Plasticity in Adults
The development and critical period plasticity of the
sensory systems, particularly the visual cortex, have been
well described. In rats and mice, critical periods of
plasticity in the visual, somatosensory barrel, and auditory
cortices take place early in the development of these
sensory systems. Importantly, some of the associated
plasticity mechanisms are reused or reopened in adult
neuronal plasticity. Drugs inhibiting or stimulating the
GABAA receptors can promote or delay the critical periods,
respectively (Hensch, 2005). Synaptic pruning eliminates
extra nonfunctional connections formed during initial
overproduction of synaptic connections, while strengthening functionally important ones (Changeux and Danchin,
1976). In the normal adult brain, neural plasticity is
needed for maintaining neuronal network excitability,
large-scale regulation of cortical and subcortical circuits,
and fine-scale experience-dependent refinement and
maintenance of local circuits (Griffen et al., 2012).
The extracellular matrix (ECM) network is an active
participant in brain function and neuroplasticity
(reviewed in Pavlov et al., 2004). The ECM consists of
various molecules (glycoproteins and proteoglycans)
secreted by all cells that are assembled inside this
matrix (Bosman and Stamenkovic, 2003). ECM network
regulates synaptogenesis, consolidation, strengthening,
and maintenance of synapses (Fields and Itoh, 1996).
During development, ECM participates in neuronal
differentiation, neuronal movement, guidance for growing axons, and synaptogenesis (Pavlov et al., 2004). In
the mature brain, ECM is crucial not only for anchoring
of neurons and organization of brain regions (structural
function) but also for transducing a wide range of
signals to the neurons (Thalhammer and Cingolani,
2014). Interaction between the ECM and cell is facilitated by cell adhesion molecules (CAMs), particularly
by integrins. Integrins are heterodimeric transmembrane glycoprotein receptors that mediate ECM-cell
interactions via binding ECM proteins and cellular
transmembrane proteins (e.g., ion channels and growth
factor receptors). ECM plays an important role in
regulating actin polymerization, spine morphology,
and GluA2 subunit trafficking in culture (Wu and
Reddy, 2012). In regard to this review, suppression of
the synthesis of integrin-linked kinase increases the
level of Ser845 phosphorylated GluA1 subunitcontaining receptors and spine density in the NAc and
blocks cocaine-induced sensitization (Chen et al., 2010).
Moreover, disruption of integrin expression in the NAc
interfered with trafficking of GluA2 subunit-containing
receptors, which affected cocaine-induced conditioned
place preference (CPP) and reinstated drug seeking
(Wiggins et al., 2011).

884

Korpi et al.

Perineuronal nets (PPN), first described by Camillo


Golgi (reviewed in Celio et al., 1998), develop around
specific neuronal populations to protect highly active
neurons and to preserve their structure in an
experience-dependent manner during pruning of extra
connections in the visual cortex and many other parts of
the central nervous system (CNS) (Kwok et al., 2011;
Soleman et al., 2013; Ye and Miao, 2013). PNNs are
primarily composed of chondroitin sulfate proteoglycans (e.g., aggrecan, versican, neurocan, brevican, and
phosphacan), hyaluronan, link proteins such as cartilage
link protein 1 Crtl1 or Hapln1 (Carulli et al., 2010), and
tenascins. In the cortex, PNNs selectively surround the
parvalbumin (Pv)-containing fast-spiking GABAergic
interneurons (Ye and Miao, 2013), starting after postnatal day 10 (P10) and achieving adult levels by P42.
Thus, the PNNs develop during the critical period, and
after closing, the critical period for plasticity can be
reopened by degradation of the PNNs (Pizzorusso et al.,
2002). Although the negatively charged components of the
PNN can have multiple molecular interactions and functions, for the maturation of Pv-interneurons and regulation of critical periods, the PNN provides a high-affinity
binding site for the lysine-arginine containing domain of
the transcription factor Otx2 (orthodenticle homeobox 2)
(Beurdeley et al., 2012). Once Otx2 reaches the Pvinterneurons of the visual cortex, being synthesized in
the retina or the lateral geniculate of the visual tract
(Sugiyama et al., 2008), it induces expression of Pv, GAD65
(glutamate decarboxylase 65 kDa), GABAA a1 subunit,
and Kv3.1b K+ channels specifically in these target
neurons (Sugiyama et al., 2008). This process is experience
dependent, that is it cannot begin before eye opening, after
which the critical period for setting ocular dominance and
visual acuity is started (Fagiolini and Hensch, 2000;
Sugiyama et al., 2009). Pv-interneurons in the visual
cortex mature in connections (e.g., forming perisomatic
synapses on principal neurons) and intrinsic firing properties (e.g., fast spiking) by the end of the critical period,
thus allowing the firm functional cortical connections for
binocular visual acuity to be formed (Kuhlman et al.,
2010). PNNs interact with receptors such as NgR1 (NoGo
receptor) and leukocyte common antigene-related phosphatase (Akbik et al., 2013; Ye and Miao, 2013).
PNN formation in the basolateral amygdala (BLA)
stabilizes fear memories in adults, unlike in the developing
brain when they can be readily extinguished (Gogolla
et al., 2009). Degradation of PNN by the enzyme chondroitinase ABC in the BLA, left the fear conditioning
intact, but after extinction the animals failed to spontaneously recover the fear and to express freezing under the
fear-related context, indicating erasure of the memory by
degradation of the PNN. Two recent papers indicate that
PNNs affect drug- and experience-induced memory in
rodents. First, Karpova et al. (2011) showed that fluoxetine
treatment during extinction training in fear-conditioned
adult mice led to erasure of fear memory, associated with

a reduction in the proportion of Pv-containing interneurons with PNN expression in the BLA, HC, and infralimbic
cortex (IL). In the BLA, fluoxetine treatment also increased the expression of polysialynated neuronal cell
adhesion molecule (PSA-NCAM) and decreased KCC2
levels in the BLA and HC, with all these data suggesting
that fluoxetine treatment with extinction training induced
a juvenile-like state in the interneurons of the fear
circuitry. Second, infusions of chondroitinase ABC into
the BLA or the central nucleus of the amygdala (CeA) of
adult rats during extinction of morphine- and cocaineinduced CPP prevented priming-induced reinstatement
(Xue et al., 2014). This effect was associated with increased
levels of AMPAR GluA1 and GluA2, but not GluA3, and
increased BDNF levels in the BLA. Thus, the extracellular
matrix and its restructuring seem to be involved in the
modulation of specific memories that can also be associated with persistent drug effects. Perhaps, changes in the
matrix may be developed into future therapies for enhancing or reversing plasticity (Castren et al., 2012). However,
we should learn more about how the extracellular matrix
induces or restricts plasticity and when would be the right
time to increase or decrease the matrix components in
appropriate neuronal circuits.
One of the major anionic extracellular components
around mature interneurons is formed by PSA-NCAM
molecules with antiadhesive properties. Expression of
PSA-NCAM has been described in a subpopulation of
GABAergic calbindin-positive and somatostatinpositive neurons in the adult rat and mouse cerebral
cortex, without expression in the principal neurons or
other interneuron types (Gomez-Climent et al., 2011).
These PSA-NCAM-positive interneurons had reduced
synaptic connections as judged from confocal immunohistochemical counting of synaptophysin-positive perisomatic and peridendritic puncta in close apposition to
GAD67-expressing cells. This suggests that expression
of PSA restricts the synaptic connections of the interneurons in mature brain. Interestingly, DAergic agonists increase (Castillo-Gomez et al., 2008) and also
serotonergic drugs increase (or downregulate stresselevated) (Wedzony et al., 2013) production of PSA in
the mPFC. NCAM is needed for hippocampal LTP that
is deficient in NCAM-KO mice (Caroni et al., 2012;
Stoenica et al., 2006). Intraventral mPFC infusion of
endoneuraminidase, which cleaves PSA from NCAM to
mice that were trained to extinguish operant alcoholseeking behavior, strongly attenuated extinction (Barker
et al., 2012), suggesting that reduced levels of PSANCAM restricts the plasticity needed for extinction.
Similarly to PNNs, extracellular NCAM and PSANCAM mechanisms also participate in retaining the
mature structures, although they are also needed for
novel connections and pathways of plasticity. They do
offer a number of possible, but still deficiently characterized, targets for drug action, and they may be important
for understanding impairments of interneuron function

Drug-Induced Neuroplasticity

in depression and schizophrenia (Nacher et al., 2013;


Wedzony et al., 2013).
G. Neurotrophins as Regulators of Neuroplasticity:
Brain-Derived Neurotrophic Factor as an Example
BDNF is an important trophic factor for nerve cells in
the CNS (Park and Poo, 2013), and it serves here as an
example of the multiple roles that the neurotrophins are
playing in orchestrating development, maintenance,
and plasticity of neurogenesis and neural connections.
It is secreted from various cells activity dependently
and exerts effects on local synaptic plasticity.
In short, mature BDNF is a 14-kDa protein that is
cleaved from prepro-BDNF expressed both in neurons
and glial cells in the CNS. BDNF gene has IX promoters
in humans and rodents, and at least the promoter IV is
activity dependent with Ca2+ and cAMP response
elements (Park and Poo, 2013). BDNF preferentially
acts via activation of tropomyosin receptor kinase B
(TrkB) (Minichiello, 2009). Like many of the trophic
factor receptors, TrkB belongs to the family of receptor
tyrosine kinases. Signaling of BDNF via TrkB leads to
activation of Ras-ERK, phosphatidyl inositol 3-kinase
(PI3K) and phospholipase C, gamma (PLCg) pathways,
which serve for survival, differentiation, and synaptic
plasticity phenomena in neurons (Minichiello, 2009).
Pro-BDNF activates preferentially P75 neurotrophin
receptor and often has opposite or different effects than
mature BDNF (Lu et al., 2005).
BDNF has important interactions with the GABA
system. It is needed for the GABA system to become
inhibitory during development (Huang et al., 1999), but
its activation in the mature nervous system may lead to
reduction of KCC2 expression and subsequent change to
GABAergic excitation. This has been described during
development of neuropathic pain in the spinal cord, with
activated microglial cells releasing BDNF, which makes
the circuitry abnormally sensitive to stimuli due to
downregulation of KCC2 expression (Coull et al., 2005)
and due to LTP induction via GluN2B upregulation
(Ding et al., 2014). In similar preclinical models, certain
benzodiazepine-site agonists, the efficacy of which is
supposed to be dependent on KCC2-driven Cl2 gradients, surprisingly show strong analgesic effects by
ablating neuropathic pain symptoms (Knabl et al.,
2008; Paul et al., 2014a), stressing that multiple parallel
pathways are involved (Zeilhofer et al., 2012). The KCC2
downregulation and GABAergic excitation also appear to
switch rewarding responses in the CPP test to intra-VTA
GABAA agonist muscimol from DA-sensitive to DAinsensitive (or vice versa for the antagonist bicuculline)
in opioid-dependent rats (Laviolette et al., 2004). This
condition could be mimicked by intra-VTA infusion of
BDNF (Vargas-Perez et al., 2009). Furthermore, more
recently it was shown that in stressed animals, specific
hypothalamic neurons are excited by GABA due to stressinduced phosphorylation and endocytosis of KCC2

885

(Sarkar et al., 2011). These striking results appear to


link chronic drug exposure, stress, and aversive pain
stimuli to pathologic neuroadaptation in BDNF and
GABA systems and indicate that boosting the chloride
gradients might be an attractive drug development
target (Gagnon et al., 2013).
Although GABAA receptor-mediated excitation or
depolarization in principal neurons is known to be
important in early neuronal development (Ben-Ari
et al., 2007; Blaesse et al., 2009) and perhaps in specific
pathologic conditions in the mature nervous system
(Medina et al., 2014), to our knowledge there is still
little evidence that drugs acting on GABAA receptors
would induce tolerance via reversed Cl2 gradients
during in vivo treatments. In addition, acute behavioral
effects of drugs with preferential action on extrasynaptic GABAA receptors, including gaboxadol, were not
affected in mice hypomorphic to KCC2, although the
responses to the synaptic GABAA receptor agonist
diazepam were strongly blunted (Tornberg et al.,
2007). Interneurons express KCC2 already early in
development (Batista-Brito et al., 2008), consistent with
shunting inhibition in the interneurons already at
postnatal days ,10 (Banke and McBain, 2006). The
stabilized role of specific interneuron populations, such
as Pv-interneurons, may become even more important
at the end of the critical periods (Hensch, 2014).
7,8-Dihydroxyflavone, a specific small molecule agonist of TrkB receptors, has a wide activity in various
preclinical tests for cognitive and emotional behaviors
(Baker-Andresen et al., 2013), but to our knowledge, it
has not yet been tested for modulation of drug-induced
plasticity. Perhaps only together with a systemically
active antagonist of TrkB receptors, the roles of BDNFTrkB signaling in various phases of drug-induced
plasticity would become better established. However,
as mentioned above, in the developing CNS during the
critical periods of cortical circuitry formation BDNF
signaling is indispensable for maturation of GABAergic
inhibition (Huang et al., 1999) (reviewed in Kuczewski
et al., 2010).
As discussed above, inhibitory interneurons have an
important role in neuroplasticity, and their diversity,
development, and functions were recently characterized
in the cortical regions (Klausberger and Somogyi, 2008;
Ascoli et al., 2008; Lapray et al., 2012), but only initially
in the midbrain (Olson and Nestler, 2007; Nair-Roberts
et al., 2008). Because the midbrain GABAergic interneurons even have different embryonic origins to the
cortical interneurons, likely subpopulations and characteristics of midbrain interneurons will be an interesting target for research and drug effects.
H. Animal Models of Drug Reinforcement
and Addiction
Animal models are critical for understanding the
neurobiological basis of drug actions and drug addiction.

886

Korpi et al.

Because addiction is a human phenomenon, there are


no complete animal models for it. However, some
characteristics of this syndrome can be satisfactorily
modeled in laboratory animals. It is generally agreed
that when viewed as experimental preparations for
studying the human syndrome, animal models exhibit
good construct and predictive validity, even if their face
validity can be disputed (for a comprehensive review,
see Sanchis-Segura and Spanagel, 2006).
Conceptualizations of the development of addiction hold
that drugs are initially taken voluntarily because of their
positive reinforcing or rewarding effects, and therefore
models have been constructed for evaluating the ability of
drugs to act as reinforcers or conditioned reinforcers,
including a variety of animal models of drug SA. Behaviorally, they can be classified as operant models based on
operant conditioning and nonoperant models that are
restricted to various oral SA procedures used particularly
for measuring ethanol consumption. Traditionally, these
models were based on a two-bottle choice between water
and ethanol solution, but it was subsequently found that
manipulation of the temporal access to ethanol can
increase the consumed ethanol during drinking bouts,
leading to development of binge-drinking models (Wise,
1973; Rhodes et al., 2005; Simms et al., 2008).
Operant models are based on the delivery of a reinforcer
contingently to the completion of the required response
that is determined by the reinforcement schedule. Reinforcement schedules exert powerful control over behavior and can be used for evaluating the nonspecific and
motivational drug effects. For example, progressive ratio
schedules can measure the reinforcing efficacy of a drug,
whereas the second-order schedules are useful for studying the role of conditioned reinforcers in maintaining
behavior. Various routes of SA can be used, e.g., intravenous, intracranial, intragastric, or oral. In particular, the intravenous drug SA animal model is considered
to be predictive of drug abuse potential in humans.
In conditioned preference paradigms, drugs effects (unconditioned stimulus) are paired repeatedly with a previous neutral stimulus. Through Pavlovian conditioning,
the neutral stimulus acquires the ability to act as a conditioned stimulus. Most commonly, conditioned preference is
studied using CPP procedures. These procedures use an
apparatus with two or more separate distinctive compartments. Access to one compartment is paired with drug
injections, whereas the other compartment is accessible
after vehicle injections. After repeated pairings of the drug
and the vehicle with their compartments, the animal is
allowed to move freely in all compartments during the test
session. Increase in the time spent in the drug-associated
compartment is taken as a measure of CPP.
Animals will perform tasks to self-administer electrical trains of stimulation to many different brain
areas, particularly along the medial forebrain bundle,
leading to the idea that direct activation of brain reveals
the circuitry involved in reinforcement from natural

rewards. Various intracranial self-stimulation (ICSS)


procedures have been used for mapping the neural
circuitry mediating drug reinforcement and pharmacological suppression thereof. Many abused drugs acutely
decrease the ICSS reward threshold, whereas increased
thresholds have been seen during withdrawal after
induction of dependence.
Compulsive drug seeking and relapse to drug use are
considered to be the hallmarks of addiction, and
therefore attempts have been made to model these
aspects of addiction in animal models. Particularly for
modeling drug-seeking behavior, reinstatement procedures are now being commonly used. These procedures
consist of the initial phase of operant drug SA, the
extinction phase during which operant responses have
no programmable consequences, and the reinstatement
phase during which the extinguished responding is
reinstated either with a priming dose of the drug,
conditioned stimuli associated with the drug, stressful
footshocks, or pharmacological stressors. Drug seeking
induced by re-exposure to drug-associated cues has
been shown to progressively increase over several
weeks after withdrawal from drug SA, a phenomenon
that has been called incubation of drug craving (Lu
et al., 2004), which is reminiscent of the alcohol
deprivation effect found earlier in rats after a period of
abstinence (Sinclair, 1972). The compulsive features of
drug taking, such as resistance to adverse consequences
and escalated intake, are accentuated in models in which
the daily access to a drug is extended (Ahmed and Koob,
1998; Deroche-Gamonet et al., 2004; Vanderschuren and
Everitt, 2004).
Psychomotor or behavioral sensitization refers to the
progressive increase in locomotion when abused drugs
are administered repeatedly. This phenomenon is in
essence nonassociative, but the context of drug administration has been demonstrated to have a role in the
induction and expression of sensitization. It has been
theorized that psychomotor sensitization also entails
increased incentive salience of stimuli associated with
drug effects, and therefore sensitization is implicated in
transition to compulsive drug seeking and taking
(Robinson and Berridge, 1993).
II. Actions and Persistent Effects of Specific
Drugs of Abuse
A. Cocaine, a Stimulant and Local Anesthetic
Cocaine, derived from the leaves of the coca plant, is
a very potent psychostimulant. Cocaine and its derivatives, can be inhaled, injected into the bloodstream, or
smoked, with the duration and intensity of effects
depending on the route of administration. Its effects
include euphoria (which might eventually be replaced
by anxiety), hyperactiveness, suppression of appetite,
local anesthesia, and possible sudden death due
to cardiac arrest that makes it considerably more

Drug-Induced Neuroplasticity

dangerous than other psychostimulants (for comprehensive review, see Grabowski, 1984). In addition,
cocaine has been associated with a variety of cardiovascular disorders, including myocardial infarction, heart
failure, cardiomyopathies, arrhythmias, aortic dissection, and endocarditis (Schwartz et al., 2010).
Cocaine is a DA, norepinephrine (NE), and serotonin
reuptake inhibitor (Fig. 5; Table 1). It also blocks
voltage-dependent sodium channels, which produces
local anesthesia, and also has been shown to cause
locomotor depression rather than stimulation (Reith
et al., 1985). Cocaine increases plasma catecholamine
levels; therefore, it exhibits sympathomimetic properties. As a consequence, it is a potent vasoconstrictor of
blood vessels in the brain (Kaufman et al., 1998; Volkow
et al., 1996; Wallace et al., 1996). The addictive
potential of cocaine is widely considered to be mainly
because of its block of DA reuptake (Sulzer, 2011). The
vast majority of studies have focused on elucidating the
mechanisms via which DA reuptake suppression leads
to addictive behaviors.
In this regard, the holy grail of the field has been to
understand lasting changes induced in the brain due to
this reuptake inhibition, which outlasts the presence of
the drug in the system. The main rationale for this
approach is that addictive behaviors persist well beyond
the presence of the drug in the system. One idea that
has gained considerable experimental support in the
past decade or so is that the normal learning processes
in the brain (involving the mesocorticolimbic DA system) are "hijacked" in addiction to reinforce the acquisition of the addictive drug. As is discussed in this
review, experimental results have lent considerable
support to this notion; however, recently emerging
evidence indicates this is only part of the picture (Nutt
et al., 2015). Furthermore, emerging evidence indicates
that other structures play a role in addiction as well. In
this review, the effects of cocaine on lasting changes in
the brain are discussed in a brain area-dependent
fashion, because the evidence indicates different brain
regions might play different roles in addiction (Thomas
et al., 2001; Ungless et al., 2001; Conrad et al., 2008;
Kalivas, 2009; Koob and Volkow, 2010; Luscher and
Malenka, 2011; Pascoli et al., 2012, 2014).
1. Persistent Ventral Tegmental Area Neuroplasticity
after a Single Dose. The VTA, along with the adjacent
substantia nigra (SN), forms the principle source of
DAergic projection neurons in the brain. Located in the
midbrain, it projects to the many forebrain regions
including the ventral striatum (NAc), amygdala, lateral
habenula (LHb), mPFC, ventral hippocampus (vHC),
and bed nucleus of stria terminalis (BNST) (see Fig. 1).
Although the majority of its projections are DAergic in
nature, significant populations of GABAergic and glutamatergic projection neurons have been observed as
well (Nair-Roberts et al., 2008; Stuber et al., 2010;
Hnasko and Edwards, 2012). It is believed that the VTA

887

Fig. 5. The effect of stimulant drugs on dopaminergic neurotransmission


is characterized by elevation of synaptic dopamine (DA) levels and
subsequent increases in postsynaptic neuronal responses. Cocaine blocks
the dopamine transporters (DAT), resulting in increased synaptic DA
levels due to competitive reuptake inhibition. Amphetamine and related
substances (AMPH), such as methamphetamine, are taken up into the
presynaptic neuron via the DAT, also resulting in increased synaptic DA
levels due to competitive reuptake inhibition. Because AMPHs are
lipophilic weak bases they are also capable of entering the cytoplasm via
diffusion through the membrane. Furthermore, once inside the cytoplasm
they enter secretory vesicles and cause an efflux of DA from them.
AMPHs are capable of reversing the function of DAT, causing further
increases in synaptic DA levels. In contrast, cocaine and related
stimulants do not enter the presynaptic neuron and enhance synaptic
DA levels only by inhibiting its reuptake. Either way, the end result is
enhanced signaling through DA D1- and D2-like receptors, leading to
activation or inhibition, respectively, of adenylate cyclase (AC) and
further downstream cellular responses.

plays a role in assigning "salience" to rewarding or


punishing events, and the more salient (or surprising)
an event is, the greater the chance that it is learned.
DAergic neurons in the VTA drastically increase firing
in response to unexpected rewards or punishments and
the cues associated with them (Mirenowicz and Schultz,
1996; Ungless et al., 2004; Brischoux et al., 2009).
Therefore, a drug that evokes a persistent increase in
DA neuronal firing or leads to a long-lasting potentiation of excitatory inputs to DA neuronspossibly by
inducing LTPcould in theory drive motivational
behaviors such as are observed toward salient events
(Wolf, 2002).
Initial evidence indicated that a single, noncontingent (experimenter administered) dose of cocaine led to

888

Korpi et al.
TABLE 1
Inhibition of transporter ligand binding and reuptake by cocaine and other stimulants

(A) The concentration ranges of the amphetamines AMPH, METH, and MDMA; the cathinones methcathinone, 4-MMC, and MDMC as well as cocaine and MPH needed to
displace typical DAT, SERT, NET, and VMAT2 ligands (Ki in mM). (B) The concentration ranges required to inhibit monoamine reuptake via the same transporters (IC50 in
mM). Note that methodological differences exist between the studies reported in the table.
Drug

DAT

SERT

NET

A. Inhibition of ligand binding to monoamine transporters, Ki range (mM)


Cocaine
0.20.7
0.40.7
0.40.5
Methylphenidate (MPH)
0.10.3
.1.10
0.10.3
AMPH
0.55.7
2438
0.11.0

VMAT2

References

0.1
39

(Fone and Nutt, 2005; Han and Gu, 2006)


(Bymaster et al., 2002; Fone and Nutt, 2005)
(Teng et al., 1998; Han and Gu, 2006; John and Jones,
2007; Simmler et al., 2013)
(Han and Gu, 2006; Nickell et al., 2010; Eshleman et al.,
2013; Simmler et al., 2013)
(Han and Gu, 2006; Montgomery et al., 2007b;
Eshleman et al., 2013; Iversen et al., 2013)
(Eshleman et al., 2013; Iversen et al., 2013; Simmler
et al., 2013)
(Angoa-Perez et al., 2013a; Eshleman et al., 2013;
Simmler et al., 2013)
(Eshleman et al., 2013; Simmler et al., 2013)

METH

0.54.0

9.3240

0.14.3

80-920

MDMA

1.78.3

0.62.5

0.61.8

661

Methcathinone

0.63.3

1.8.100

1.54.5

.1000

4-MMC

1.54.8

18.30

12.35

.1000

16.25

.1000

0.133.3

MDMC
2.75.0
.3095
B. Inhibition of monoamine reuptake, IC50 range (mM)
Cocaine
0.090.46
0.481.05
Methylphenidate (MPH)

0.150.26

16

0.47

AMPH
METH

0.091.6
0.010.47

.13.4
1.312

0.060.07
0.010.65

12
4.711

MDMA

0.050.48

0.0535

0.056.6

5.813

Methcathinone

0.140.36

0.0235

0.030.51

112

4-MMC

0.050.97

0.120.51

0.050.49

3.4116

MDMC

0.131.2

0.235.75

0.030.51

112

a potentiation of excitatory synaptic transmission onto


DA neurons in the VTA, and the changes observed in
response to repeated cocaine administration were similar (Ungless et al., 2001; Saal et al., 2003; Borgland
et al., 2004). This potentiation was observed by measuring the ratio of AMPAR to NMDAR-mediated currents (AMPAR/NMDAR ratio) in a neuron clamped at
a fixed voltage ex vivo, 24 hour after cocaine had been
administered in vivo. One advantage of this protocol,
which has been widely used since in the field for cocaine
and other drugs of abuse, is that lasting changes induced
in vivo can be observed (as the drug has cleared the
system). This potentiation of excitatory synaptic transmission returns to normal levels when assessed a week
later (Ungless et al., 2001). This astounding finding
indicates that a single dose of cocaine might be sufficient
to induce lasting synaptic changes in the VTA and
possibly alter activity in downstream targets. In this
regard, a recent report further characterized the earlier
findings and demonstrated that only VTA DA neurons
projecting to the NAc shell and not to the mPFC are
potentiated on exposure to cocaine (Lammel et al., 2011).
Glutamate receptor neuroplasticity could not be elicited
in the VTA-mPFC neurons even by repeated exposure to
cocaine, whereas, in contrast to cocaine, an aversive
stimulus (a formalin injection) potentiated the synapses
of the mPFC projection in the VTA. Furthermore,
optogenetic activation of inputs from the laterodorsal

(Bennett et al., 1995; Carroll et al., 1995; Meltzer et al.,


2006)
(Gatley et al., 1996)
(Markowitz et al., 2006)
(Fone and Nutt, 2005; Baumann et al., 2013;)
(Cozzi et al., 1999; Baumann et al., 2012; Eshleman
et al., 2013)
(Cozzi et al., 1999; Baumann et al., 2012; Eshleman
et al., 2013)
(Cozzi et al., 1999; Fleckenstein et al., 1999; Baumann
et al., 2012; Eshleman et al., 2013)
(Lopez-Arnau et al., 2012; Baumann et al., 2012, 2013;
Eshleman et al., 2013)
(Cozzi et al., 1999; Fone and Nutt, 2005; Baumann
et al., 2012; Lopez-Arnau et al., 2012; Eshleman
et al., 2013)

tegmental nucleus (LDTg, which preferentially synapse


onto VTA DA neurons projecting to the NAc lateral shell)
elicits reward, whereas activation of inputs from the LHb
(preferentially synapsing on DAergic projections to the
mPFC and GABAergic neurons in the VTA) results in
aversion (Lammel et al., 2012). These results suggest that
only the VTA DA neurons involved in reward processing
are affected (further discussed within this section concerning cocaine effects in the NAc and mPFC).
The increase in AMPAR/NMDAR ratio is due to
increased AMPAR currents, which is mainly due to
insertion of Ca2+-permeable GluA2 subunit-lacking
AMPAR receptors (Fig. 6) (Mameli et al., 2011). However, NMDAR-dependent currents are reduced as well
(reviewed in Luscher, 2013). This reduction is due to
insertion of quasi-Ca2+-impermeable NMDARs containing GluN3A and GluN2B subunits. It has been hypothesized that GluN3A-containing NMDARs might play
a role in cocaine-evoked AMPAR plasticity as well
(Yuan et al., 2013). Therefore, removing the GluN3Acontaining NMDARs might reduce cocaine-evoked
AMPAR plasticity and mitigate its effects on downstream targets. Activation of mGlu1 receptors in the
VTA removed these NMDARs (Yuan et al., 2013).
Furthermore, functionally overriding mGlu1 receptors
in vivo made cocaine-evoked synaptic potentiation in the
VTA persistent (that is, it did not return to normal after
7 days) (Mameli et al., 2009). This suggests mGlu1

Drug-Induced Neuroplasticity

Fig. 6. Neuronal plasticity such as long-term potentiation (LTP) or longterm depression (LTD) after exposure to acute and subchronic stimulant
drugs may occur via several distinct mechanisms. Acute cocaine exposure
induces postsynaptic LTP that may occur when novel Ca2+-permeable nonGluA2 subunit-containing AMPA receptors (AMPAR) are inserted into the
postsynaptic membrane. This process is mediated by Ca2+/calmodulindependent protein kinase II (CaMKII) after its activation by Ca2+ and is also
influenced by signaling from G protein-coupled orexin and corticotropinreleasing factor (CRF) receptors and by recruitment of GluN2B and GluN3A
subunit-containing NMDA receptors (NMDAR). Synaptic LTD occurs due to
internalization of membrane-expressed AMPARs in response to elevated
Ca2+ levels or activation of metabotropic glutamate receptors (mGlu).
Another type of LTD is due to on-demand synthesis of endocannabinoids
(eCB) such as 2-arachidonoylglycerol and anandamide (shown in green) that
activate presynaptic cannabinoid CB1 receptors, resulting in inhibition of
presynaptic neurotransmitter release (not shown here; please, see Fig. 12).
Subchronic cocaine treatment may generate silent synapses mediated by
selective recruitment of GluN2B subunit-containing NMDARs into new
synaptic locations. These sites might then maturate into functional synapses
after synaptic targeting of new AMPARs.

signaling plays an important role in regulating plasticity


in the VTA, and manipulating its activity could potentially provide a means of reducing downstream changes
induced by VTA activation in the initial vulnerable
stages of usage of drugs. In addition to redistribution of
receptors, another attractive model for plasticity in the
CNS is dendritic reorganization and dendritic spine
remodeling (which influences dendritic input to the
neuron). In this regard, spine density in a subset of
VTA DA neurons (possibly the VTA-NAc DA neurons,
but this was not investigated) has been shown to increase
in response to a single noncontingent dose of cocaine and,
strikingly, only these cells exhibited an increase in
AMPAR/NMDAR ratio as well (Sarti et al., 2007).
One possible mechanism that is suggested to maintain
LTP in various neurons by enhancing AMPAR synaptic
targeting is via prolonged induction of PKMz activation
(Sacktor, 2011). Interestingly, intracellular levels of
PKMz, a member of the atypical PKC subfamily, are
increased in the VTA neurons by cocaine treatment.
Suppressing this increase in PKMz levels reversed
cocaine-enhanced excitatory synaptic transmission, the

889

enhanced AMPAR/NMDAR ratio in the VTA, and the


cocaine-induced locomotor sensitization (Ho et al., 2012;
Velez-Hernandez et al., 2013). The role of PKMz, however, is still uncertain. The above findings have been
based only on the inhibition of PKMz activity by a zetainhibitory peptide. This inhibitor may not be specific for
PKMz, and genetic inactivation models of PKMz activity
have failed to alter memory processes that were suggested to be dependent on PKMz (Lee et al., 2013a; Volk
et al., 2013). Other atypical PKC members might compensate the genetic inactivation of PKMz, suggesting
redundance of regulating mechanisms (Jalil et al., 2015).
Optogenetic activation of GABA neurons in the VTA
results in aversion and disrupts reward consumption
(Tan et al., 2012a; van Zessen et al., 2012). This
indicates that in addition to DA neurons in the VTA,
GABA neurons play a pivotal role in reward processing
as well. Bocklisch et al. (2013) found that inhibitory
dopamine-1 receptor (D1R)-expressing medium spiny
neurons (MSNs) from the NAc that strongly project to
GABAergic interneurons in the VTA exhibited cocaineevoked synaptic potentiation. This activation of MSNs
led to a disinhibition of VTA DA neurons. Therefore,
cocaine evokes increased DAergic output from the VTA,
via indirect, disinhibitory mechanisms as well.
It is evident that the drug-induced changes in the
brain are not simply dependent on the drug alone, but
the interaction of the drug with the environment the
animal is in (for discussion, see Badiani, 2013; Luscher
and Malenka, 2011). In this regard, the effects of two
neuropeptides, corticotropin-releasing factor (CRF) and
orexin (OX), which play a role in stress, arousal, and
hunger, on potentiation of cocaine-evoked synaptic plasticity, have been widely studied. Chronic noncontingent
cocaine exposure leads to a potentiation of CRFdependent AMPAR and NMDAR neurotransmission in
the VTA (Hahn et al., 2009), whereas an OX1 receptor
antagonist blocks the potentiation of excitatory neurotransmission onto DAergic neurons observed in the VTA
of cocaine-treated animals (Borgland et al., 2006). Intriguingly, OX1 antagonist simultaneously blocked locomotor sensitization as well. Activation of OX2 receptors,
on the other hand, increased release of glutamate on to
VTA DA neurons and potentiated NMDAR currents
(Borgland et al., 2008). Nonetheless, these neuropeptides
actively modulate cocaine-induced changes in VTA DA
neurotransmission (the role of CRF and OX in VTAmediated reinstatement of drug seeking is further
discussed below), emphasizing the importance of the
interaction of the environment with the drug of abuse.
The majority of the studies have focused on the role of
the VTA in initial drug taking, which might predispose
the individual to later compulsive drug seeking. However, there is a preponderance of evidence that indicates
the VTA plays a role in drug relapse as well. A variety of
factors contribute to reinstatement of cocaine seeking
after drug-seeking behaviors have been extinguished.

890

Korpi et al.

These include stress, cues related to the drug, the


context the drug was administered in, and a "priming"
injection of the drug. In this regard, footshock a
stressorreinstates cocaine seeking by releasing CRF,
which in turn leads to an activation of VTA DA neurons
via the release of glutamate (Wang et al., 2005). CRF
neurotransmission, via CRF1 and CRF2 receptors, in
the VTA promotes reinstatement, in long-access and
short-access SA protocols, respectively (Wang et al.,
2007; Blacktop et al., 2011; reviewed in Koob and
Zorrilla, 2010; Wise and Morales, 2010). Interestingly,
activation of OX1 receptors increases reinstatement of
cocaine seeking as well, in an independent manner from
the effects of CRF (Borgland et al., 2010; Baimel and
Borgland, 2012). Therefore, neuropeptides not only
modulate initial cocaine intake but play a major role
in relapse as well.
The studies discussed thus far involve administration
of cocaine by the experimenter to the animal in a noncontingent manner (nondependent on the behavior of
the animal). Studies have remarkably shown that the
effects of noncontingent and contingent administrations
on the brain are different. In the VTA, SA of cocaine
leads to a persistent synaptic potentiation, which lasts
for up to 3 months of abstinence (Chen et al., 2008).
Therefore, further studies are required to elucidate the
mechanisms underlying these disparate changes in the
VTA.
2. Changes in the Nucleus Accumbens Mediate
Cocaine Addiction and Relapse. The NAc, principally
owing to its connectivity, is believed to play a role in
initiation of movements based on emotional salience of
encountered events. The major excitatory inputs to the
NAc are from the mPFC, BLA, vHC, and thalamus.
Furthermore, the NAc receives dense projections from
the VTA, mostly DA-releasing neurons. Anatomic studies reveal a repetitive "loop-like" mesostriatal circuitry
(Haber et al., 2000). Medial aspects of the VTA project to
the shell of the NAc, projecting back to more lateral
aspects of the VTA that in turn sends projections to the
NAc core. The core then projects to the SN that projects
more dorsally to the striatum and so on. Indeed, this
loop-like circuitry has been hypothesized to be critically
involved in addictive behaviors, by virtue of its role in
reinforcement learning and eventual habit formation
(for discussion, see Everitt and Robbins, 2005). The
majority of ventral and dorsal striatal neurons are
projecting MSNs, whose activity is mostly regulated
by excitatory inputs from cortical and limbic areas
(Sesack and Grace, 2010).
Unlike in the VTA, multiple noncontingent doses of
cocaine administration are required to elicit synaptic
plasticity in excitatory synapses in the NAc (Thomas
et al., 2001), consistent with the hypothesis that the
NAc is affected due to the action of cocaine on the VTA
and its projections. Interestingly, cocaine is intracranially self-administered only into the NAc or mPFC, not

into the VTA (Wise, 1996). The direction of the synaptic


plasticity induced due to repeated cocaine administration was a huge controversy in the field that was
recently resolved. Earlier work (Thomas et al., 2001;
Beurrier and Malenka, 2002), showed a reduction in the
AMPAR/NMDAR current ratios 1014 days after repeated cocaine exposure on administration of a challenge dose of cocaine. In line with these findings,
because mPFC-NAc synapses (a major excitatory input)
on MSNs exhibited a reduced AMPAR/NMDAR ratio
(indicative of an LTD state), further LTD could not be
elicited via electrical means and the amplitude of miniEPSCs was reduced. Activation of D1Rs by DA led to an
enhanced reduction in the AMPAR-mediated synaptic
transmission, but not NMDAR-mediated synaptic
transmission in the NAc shell (Beurrier and Malenka,
2002). On the other hand, if a challenge dose of cocaine
was not administered, a potentiation of NAc excitatory
synapses is observed (Kourrich et al., 2007). Therefore,
these results together indicate that NAc excitatory
synaptic transmission is potentiated gradually on withdrawal from repeated cocaine exposure and becomes
depressed on administration of a cocaine challenge dose
(the transient depression returns to normal levels in 7
days). The mechanisms underlying the withdrawalinduced increase in AMPAR/NMDAR ratio are still
under active investigation. Using coimmunoprecipitation studies and dissociated NAc-mPFC cultures, it was
earlier shown that an increased number of AMPARs are
recruited to the cell membrane (reviewed in Wolf and
Ferrario, 2010), with no detectable change in the subunit composition of the AMPARs. However, longer
withdrawal periods were recently shown to lead to
expression of GluA2 subunit-lacking AMPARs (which
are Ca2+ permeable). Furthermore, the expression of
these receptors was correlated with incubation of cocaine craving (defined as the progressive increase of
cue-induced cocaine seeking with time of withdrawal
and abstinence). Pharmacological inactivation of GluA2
subunit-lacking AMPARs blocked the incubation of
cocaine craving as well (Conrad et al., 2008). This
indicates that Ca2+-permeable AMPARs in the NAc
are involved in increased vulnerability of relapse with
longer withdrawal times. Moreover, the increased expression of GluA2-lacking AMPARs is preceded by
a reduction in mGlu1 receptor expression. Indeed,
systemic mGlu1 activation reverses cue-induced cocaine
craving by resetting AMPAR-mediated neurotransmission to control levels (McCutcheon et al., 2011a; Loweth
et al., 2014).
Of note here are the parallelisms between the VTA
and NAc, including the increased expression of GluA2lacking AMPARs in response to a single cocaine dose
and extended withdrawal from SA, respectively, and
their reversal by mGlu1 activation (Fig. 6). These
results taken together suggest that mGlu receptor
modulators have therapeutic potential for curbing

Drug-Induced Neuroplasticity

addiction. However, it is important to note that the


expression of GluA2-lacking AMPARs in the NAc is
only observed after long-access cocaine SA protocols
and not after short-access or noncontingent protocols
(McCutcheon et al., 2011b; Purgianto et al., 2013). In
theory, an increase in AMPAR/NMDAR ratio could be
also mediated by a decrease in the number or function of
NMDAR receptors as well. In this regard, Ferrario et al.
(2012) found that the percentage of spines mediating
NMDAR currents is reduced, but not other characteristics of NMDAR function, during the abstinence from
cocaine as well.
The importance of dissecting the specific roles of
inputs to the NAc is increasingly being realized. However, intriguingly, Stuber et al. (2011) and Britt et al.
(2012) demonstrated that three major inputs to the NAc
from the vHC, mPFC, and BLA induce self-stimulation
behavior when animals are set for specific optogenetic
stimulation of these pathways. These results suggest
that intensity of excitatory input and not the origin are
important for motivated behavior, which, remarkably,
could be induced by direct optogenetic stimulation of the
NAc MSNs (Britt et al., 2012).
Another previously overlooked change is the emergence of silent synapses. Silent synapses do not express
AMPARs, and owing to the presence of NMDARs only
which are deactivated at resting membrane and slightly
depolarized potentialsare silent under normal conditions. Initially, it was believed that these synapses
were only a characteristic of the developing brain.
However, Huang et al. (2009) showed that a salient
experience such as a subchronic, noncontingent, exposure to cocaine generates de novo silent synapses in the
NAc after a short withdrawal (1 or 2 days after 5-day
cocaine treatment) when examined ex vivo (Fig. 6).
Furthermore, increasing the number of days cocaine
was administered also increased the number of silent
synapses generated, which returned to normal levels
after a prolonged withdrawal. The generation of silent
synapses recruits a CREB-dependent insertion of novel
GluN2B subunit-containing NMDARs into the postsynapses, prevention of which impairs the development of
cocaine-elicited locomotor sensitization (Brown et al.,
2011). It is tempting to speculate here that some of the
cocaine-generated silent synapses can undergo a maturation process to form fully functional synapses. Furthermore, the initial decrease in the AMPAR/NMDAR
ratio observed in the NAc on repeated cocaine administration or a challenge dose of cocaine might be due
to increased silent synapses (reviewed in Wolf and
Ferrario, 2010).
It is extremely interesting that this phenomenon of
generation of silent synapses is observed in the adult
brain; however, what is the role of silent synapses in
addiction? Two recent studies attempted to determine
the role of silent synapses in the NAc in incubation of
cocaine craving (Lee et al., 2013b; Ma et al., 2014).

891

Consistent with the earlier findings, they observed an


increase in silent synapses in the BLA-NAc shell
projections during early withdrawal from cocaine SA
regimens (Lee et al., 2013b) by comparing synaptic
failure rates in NAc MSNs in response to selective
optogenetic minimal stimulation. Furthermore, they
discovered a similar time course for reduction in silent
synapses and emergence of GluA2-lacking AMPARs
(critical for cue-induced cocaine craving as discussed
before). Inhibiting GluA2-lacking AMPARs increased
the number of silent synapses after extended withdrawal, suggesting that the silent synapses generated
in early withdrawal will express GluA2-lacking
AMPARs later. Interestingly, an optical LTD protocol
on BLA-NAc axons internalized the GluA2-lacking
AMPARs and reduced the increase in cocaine seeking
during withdrawal.
Ma et al. (2014) studied the role of silent synapses in
the other major excitatory projections to the NAc from
the mPFC. Silent synapses are generated in both
prelimbic subregion of the mPFC (PrL)-NAc core and
infralimbic subregion of the mPFC (IL)-NAc shell
projections during early withdrawal. Extended withdrawal led to a recruitment of GluA2-lacking AMPARs
in IL-NAc synapses (showing more rectification) but to
a recruitment of GluA2-containing AMPARs in PrLNAc synapses in NAc slices (showing less rectification).
Behaviorally, internalization of these receptors in ILNAc and PrL-NAc synapses, by selective optogenetic
LTD stimulation protocols, led rats to show an increase
and decrease in enhanced cocaine craving during
abstinence, respectively (further discussed in the PFC
section). Of note here are the parallels between the
similar effect of PrL and BLA manipulations on initial
cocaine seeking discussed earlier (Britt et al., 2012;
Stuber et al., 2011) and on incubation of cocaine craving
(Lee et al., 2013b; Ma et al., 2014); that is, motivated
behaviors require an excitatory drive to the NAc, which
can be from either the BLA or the PrL. Similarly,
cocaine withdrawal leads to an increase in BLA-NAc
and PrL-NAc silent synapses, reversal of which reduces
the enhancement of cocaine seeking during withdrawal/
abstinence.
These results taken together indicate that silent
synapses generated in the NAc due to chronic cocaine
exposure might eventually express AMPARs, which
owing to their higher conductivity, make the NAc
neurons more excitable, leading to increased cocaine
craving with time.
The role of silent synapses in the NAc might not be
limited to enhancement of cocaine craving during
withdrawal. A series of studies conducted by Hope and
colleagues (Koya et al., 2009, 2012) showed that induction of context-specific psychomotor sensitization
on noncontingent repeated administration of cocaine
is dependent on a very small subset of neurons (an
ensemble consisting of 23% of neurons in the NAc), which

892

Korpi et al.

can be visualized in c-fos-lacZ transgenic rats. Selective inactivation of these neurons by b-galactosidaseinduced formation of a cell toxin daunorubicin impaired
psychomotor sensitization. These highly active neurons
produced higher levels of silent synapses on cocaine
sensitization than the general neuron population,
which unlike the whole NAc persisted up to 611 days
after sensitization. However, the precise role of these
silent synapses is yet to be determined.
Although many earlier reports have studied the role
of NAc in cocaine addiction without considering the
heterogeneity in the NAc, it is increasingly being
realized that the NAc has distinct subcircuits. The main
population of neurons in the NAc is the MSNs. These
constitute 95% of NAc neurons and are the major class
of projection neurons from the NAc. MSNs have been
further subdivided depending on whether they express
D1Rs or dopamine-2 receptors (D2Rs). Akin to the muchstudied dorsal striatum, these two classes of MSNs form
distinct projection pathways. D1R-expressing MSNs
constitute the "direct pathway" that projects directly
to the midbrain, whereas D2R-expressing MSNs form
the "indirect pathway" that projects to the midbrain via
the ventral pallidum. Moreover, these classes of MSNs
have different electrophysiological and neurochemical
characteristics (Cepeda et al., 2008; Planert et al.,
2013). Hence, a host of recent studies have discovered
differing roles subserved by these MSNs in addictive
behaviors. Nonspecific optogenetic inhibition of NAc
core neurons reduces cocaine-primed cocaine reinstatement (Stefanik et al., 2013b); however, this effect is only
replicated by inhibiting the pallidal projections from the
NAc (indirect pathway) and not the nigral projections
(direct pathway) (Stefanik et al., 2013a). However,
another study found that optogenetic activation of
channel rhodopsin-2 expressing D2R-MSNs in the NAc
suppressed responding for cocaine, whereas their inhibition using a chemogenetic approach (designer
receptor exclusively activated by a designer drugmediated protocol) enhanced the motivation (higher
breakpoints of responding) to work for cocaine (Bock
et al., 2013).
The differences observed in these results might be
due to the different stimulation protocols employed and
to different behavioral tests. Stefanik et al. (2013a)
stimulated NAc-pallidal axons, whereas Bock et al.
(2013) manipulated D2R-MSNs directly. Alternatively,
D2R-MSNs might only be involved in compulsive cocaine SA [as measured by increased breakpoints for
responding to cocaine in Bock et al. (2013)] and not
cocaine-primed reinstatement behavior, which was
used in Stefanik et al. (2013a). In this regard, one study
(Lobo et al., 2010) selectively deleted the BDNF receptor TrkB in either D1R- or D2R-MSNs. BDNF has
been widely implicated in mechanisms underlying
synaptic plasticity. By using the CPP paradigm, genetic
deletion of TrkB expression in D1R- or D2R-MSNs

increased and decreased the rewarding properties of


cocaine, respectively. These effects were mimicked by
optogenetic activation of D1R- or D2R-MSNs, respectively. Another ingenious study (Pascoli et al., 2014)
found that incubation of cue-associated cocaine craving
is correlated with rectifying AMPAR currents and a reduced AMPAR/NMDAR ratio of mPFC-NAc shell synapses on to D1R-MSNs. On the other hand, AMPAR/
NMDAR ratio increased at vHC-NAc D1R-MSNs. Optogenetic reversal of these changes in both inputs, using
optogenetic LTD protocols, abolished cue-induced cocaine seeking. Intriguingly, LTD induction at the
mPFC-NAc synapses impaired response discrimination
in the cocaine-seeking task, whereas a reduction in
responses was observed when LTD was induced at
vHC-NAc synapses. These studies taken together indicate that D1R-MSNs are involved in promoting initial
cocaine-seeking behaviors and reinstatement of cocaineseeking behaviors as well. More specifically, inputs from
the mPFC to the NAc shell D1R-expressing neurons
might promote responding to the correct lever, whereas
vHC inputs might promote responding per se. D2R-MSNs
initially inhibit cocaine seeking, but might promote
compulsive cocaine seeking. Nonetheless, further studies
are required to clarify the precise roles of these subpopulations of NAc MSNs in addictive behaviors.
In concert with the changes in AMPAR expression in
the NAc, withdrawal from chronic noncontingent cocaine exposure results in structural changes in MSN
dendrites. A short withdrawal period (2 days) results in
increased dendritic spine formationmediated by a
variety of transcription factors and epigenetic mechanisms as discussed laterin both D1R- and D2R-MSNs,
whereas this increase is only sustained in D1R-MSNs
after prolonged withdrawal of 30 days (Lee et al., 2006).
In addition, both D1R- and D2R-MSNs exhibit dendritic
remodeling on chronic cocaine exposure, which can be
mediated by D1R activation (Li et al., 2012; Zhang et al.,
2012). Furthermore, spine structure varies considerably as well. Withdrawal from repeated noncontingent
cocaine exposure results in increased spine diameter.
Subsequent cocaine priming elicits a biphasic change in
spine diameter: transiently further increasing the size
of the spine (45 minute after injection) and then reducing it when observed 24 hours later. This change in
spine diameter is paralleled by changes in AMPAR
expression, which has been shown to transiently increase (30 minutes after administration) on cocaine reexposure before decreasing (24 hours later); and then
subsequently increasing again as the duration of withdrawal is prolonged (Boudreau et al., 2007; Kourrich
et al., 2007; Anderson et al., 2008).
Reorganization of the cytoskeletal architecture of the
neuron, or lack thereof, is critical for synaptic plasticity
and stability. Likewise, the structural changes in
dendrites of MSNs are associated with changes in the
cytoskeletal architecture as well. Withdrawal from

Drug-Induced Neuroplasticity

repeated cocaine administration produces lasting elevation in polymerized filamentous actin in the NAc, due
to suppression of actin cycling by reducing the active
form of Rac1, a Rho GTPase that controls actin remodeling (Dietz et al., 2012). Reducing actin polymerization
or increasing depolymerization results in reduced dendritic spine density and explains the transient increase
in spine head diameter on cocaine priming and behavioral sensitization (Toda et al., 2010; Dietz et al., 2012).
Indeed, knockout of Rac1 is sufficient to increase the
density of immature dendritic spines in the NAc. Hence,
these studies establish a mechanistic link between the
behavioral responses, changes in spine morphology, and
actin cycling.
In agreement with these findings, PSD-localized
kalirin-7 (Kal-7; a brain-specific guanine nucleotide
exchange factor) and its downstream effectors Rac1
and p21-activated kinase PAK are increased in the NAc,
alongside with AMPAR surface expression and spine
density on withdrawal from a repeated cocaine administration regimen (Wang et al., 2013d). Single-hairpin
RNA (shRNA)-mediated silencing of Kal-7 in the NAc
before cocaine exposure eliminates all these changes
(increased AMPAR expression and spine density). This
Kal-7 silencing, however, failed to prevent locomotor
sensitization indicating the involvement of other mechanisms in eliciting this behavior. Incentive sensitization, the ability to rapidly learn to self-administer
cocaine, was also impaired. Interestingly, constitutive
Kal-7 knockout enhanced cocaine SA (Kiraly et al.,
2013), which might be due to compensations by other
kalirin isoforms.
In addition to intracellular biochemical pathways
associated with the cytoskeletal architecture, extracellular CAMs are affected by cocaine as well. In this
regard, integrins are CAMs in the ECM that are
involved in the regulation of synaptic plasticity, including glutamate receptor trafficking. Mirroring the
other changes in the NAc, b3 integrin subunit is reduced
in the NAc core 24 hours after a SA cocaine protocol.
However, after 3 weeks of abstinence and extinction, the
levels are elevated (Wiggins et al., 2011). Moreover,
cocaine priming further regulates the levels of b3
integrin. Daily NAc microinfusions of a peptide ligand
that mimics a protein binding to integrins before SA or
reinstatement, inhibited cocaine-induced drug seeking
and normalizes the surface expression of AMPARs.
Repeated cocaine exposure and the subsequent withdrawal results in other changes in MSNs synaptic and
extrasynaptic neurotransmission, in addition to alterations in AMPAR expression, dendritic arborization,
silent synapses, spine head length, and spine density. In
general, D1R-MSNs exhibit decreased membrane excitability on withdrawal from repeated cocaine exposure,
whereas D2R-MSNs show no change in membrane
excitability (Kim et al., 2011). Furthermore, cocaine
disrupts the normal glutamate homeostasis necessary

893

for maintenance of appropriate levels of glutamate in


the synaptic and extrasynaptic (a much larger pool)
spaces. In this regard, the extracellular glutamate in
the NAc is regulated by cysteine/glutamate exchanger
xCT and glutamate transporter GLT-1 in glia (Baker
et al., 2003; Knackstedt et al., 2010). Withdrawal from
repeated cocaine injections or SA substantially reduces
extracellular glutamate, glutamate uptake, and the
expression of these transporters (Baker et al., 2003;
Knackstedt et al., 2010). Extrasynaptic glutamate
normally binds to presynaptic mGlu2/3 receptors and
reduces synaptic glutamate release. Therefore, chronic
cocaine exposure results in an increase in the synaptically releasable pool of glutamate (reviewed in Kalivas,
2009; Wolf and Ferrario, 2010). Importantly, N-acetylcysteine, a well-tolerated prodrug for increasing cysteine levels and used to replenish liver glutathione levels
in paracetamol (acetaminophen) intoxication, has been
shown to enhance extrasynaptic glutamate levels and to
reduce relapse to cocaine in rodents and in early studies
on cocaine addicts (Baker et al., 2003; Schmaal et al.,
2012; LaRowe et al., 2013). Hence, cocaine abuse results
in a host of changes in MSNs in the NAc; indeed, it has
been suggested that these changes in glutamate transmission and membrane excitability are linked to
AMPAR regulation (Wolf and Ferrario, 2010) to mediate the behavioral phenotype.
3. Altered Gene Expression in the Nucleus Accumbens
in Cocaine Seeking and Relapse. Not all individuals
have an equal propensity to develop addiction. In this
regard, approximately half of a persons risk for addiction is genetic (Goldstein, 2001; Wang et al., 2012a).
Therefore, long-lasting alterations in gene regulation
upon cocaine exposure, via transcriptional or epigenetic
mechanisms, are likely to play a pivotal role in addiction
(reviewed in Robison and Nestler, 2011). Transcription
factors are proteins that regulate gene expression in
response to intracellular signaling pathways by binding
to certain sites on the DNA and subsequently altering
the expression of the genes. Genetic material in the
nucleus is packed into chromatin complexes (DNA,
histones, and other proteins), generally split in to two
classes: euchromatin (loosely packed DNA around
histones) and heterochromatin (densely packed, hence
less active), depending on the activity states of the
chromatin. Epigenetic changes are considered alterations in this packing and thus alter activity of different
regions of DNA. Regulation of genetic activity via
transcriptional and epigenetic mechanisms involves
a myriad of processes; the field has focused mainly on
a few mechanisms that will be discussed here, including
notable work on transcription factors like DFosB,
CREB, NFkB, and MEF2, and epigenetic histone
acetylation and methylation. Because the majority of
these alterations in levels of transcription factors and
epigenetic regulation are observed on repeated exposure to cocaine, most of the studies have focused on the

894

Korpi et al.

NAc. Hence, a considerable amount of work is still


needed in this promising, budding field to elucidate the
role of numerous other factors and structures.
Expression of the Fos family of transcription factors
in the NAc is increased transiently on acute cocaine
exposure. However, only DFosB (which is only slightly
induced on acute exposure), due to its longer half-life,
accumulates on repeated drug exposure selectively in
NAc D1R-MSNs (Hope et al., 1994; Moratalla et al.,
1996; Kelz et al., 1999; reviewed in Nestler et al., 2001).
Induction of DFosB chronically in D1R-MSNs or nonselectively in the MSNs increases the rewarding properties of cocaine and other natural rewards as assessed
by CPP, enhances the locomotor response, and promotes
SA; surprisingly, a short-term change in DFosB reduces
the rewarding properties of cocaine as assessed by CPP
(McClung and Nestler, 2003; Grueter et al., 2013).
Intriguingly, overexpression of DFosB in D1R-MSNs
decreased the AMPAR/NMDAR ratio and increased
silent synapses on these neurons in the NAc shell and
core, whereas overexpression in D2R-MSNs resulted in
increased excitatory synaptic strength as assessed by
the AMPAR/NMDAR ratio and in decreased silent
synapses in the NAc shell only. Furthermore, density
of immature spines in D1R- but not D2R-MSNs increased as well on respective overexpression of DFosB in
these neuronal populations (Grueter et al., 2013).
Taken together, these findings indicate that DFosB
plays a critical role in mediating the rewarding and
locomotion-inducing properties of cocaine.
Alterations in induction of DFosB in the NAc might
play a role in relapse as well. Damez-Werno et al. (2012)
demonstrated that after prolonged withdrawal from
repeated cocaine administration, DFosB induction and
accumulation in the NAc was enhanced by cocaine reexposure. This increased induction might be due to
sustained elevation of DFosB, and increased dendritic
spine density as well, in D1R-MSNs after prolonged
withdrawal. D2R-MSNs, however, only exhibited increased DFosB and dendritic spine density after short
withdrawal periods (2 days) on cessation of a repeated
cocaine administration regimen (Lee et al., 2006).
CREB, another transcription factor, seems to increase in both major subtypes of NAc MSNs. CREB,
unlike DFosB, reduces the sensitivity to the rewarding
effects of cocaine (tolerance) and increases SA and
relapse via negative reinforcement. For instance, the
k-opioid peptide dynorphin, regulated by CREB, suppresses DA transmission to the NAc via the activation of
k-opioid receptors, thus impairing rewarding behaviors
(Carlezon et al., 1998; Larson et al., 2011; Muschamp
et al., 2011; Ehrich et al., 2014). Furthermore, the
induction of DFosB in the NAc is dependent on CREB
and serum response factor (Vialou et al., 2012). Deletion
of CREB and serum response factor, but not either one
alone, eliminates induction of DFosB on cocaine exposure in the NAc, reduces the rewarding properties of

moderate doses of cocaine as assessed by CPP, and blocks


locomotor sensitization at higher doses of cocaine. Interestingly, deletion of CREB alone has the opposite
effect, that is, an increase in behavioral sensitization and
CPP in response to cocaine (Carlezon et al., 1998; Walters
and Blendy, 2001; McClung and Nestler, 2003). The
sustained expression of DFosB in the NAc and the
accompanied increase in dendritic spines and behavioral
responses, requires CaMKIIa, which is increased on
chronic cocaine exposure as well (Robison et al., 2013).
Importantly, DFosB and CaMKIIa are increased in the
postmortem NAc of human cocaine addicts compared
with matched normal controls without histories of major
psychiatric illnesses (Robison et al., 2013).
Other transcription factors have been implicated in
cocaine addiction as well. For instance, reduction of
myocyte enhancer factor 2 activity in the NAc is required for cocaine-induced increased spine density and
possibly is linked to behavioral sensitization as well
(Pulipparacharuvil et al., 2008). Similarly, nuclear
factor kappa-B (NFkB), involved in maintaining cell
structure, is induced by chronic cocaine exposure.
Elevation and reduction of NFkB signaling in the NAc
increased and decreased (and also prevented cocaineinduced increase in) the number of dendritic spines on
NAc neurons, respectively. Furthermore, inhibition of
NFkB makes cocaine less rewarding (Russo et al., 2009).
The N-terminus of the histone (the histone tail) is
particularly susceptible to modifications. Relevant histone tail modifications involve the enzymes histone
acetyltransferases, whereas histone deacetylases (HDACs)
catalyze the removal of acetyl groups. Similarly, histone
methyltransferases and histone demethylases catalyze
the addition and removal of methyl groups, respectively. Generally, acetylation is associated with increased activity and methylation with reduced gene
activity.
Earlier studies found conflicting results on cocaineinduced behavioral responses with inhibition of HDAC.
Two studies found that systemic treatment with HDAC
inhibitors, trichostatin A and phenylbutyrate, reduced
the motivation for SA of cocaine and cocaine-plus-cued
reinstatement of drug seeking (Romieu et al., 2008,
2011), whereas a third one (Kumar et al., 2005) reported
increased rewarding effects of cocaine as assessed by
CPP and stimulation of cocaine-induced locomotion.
Furthermore, chronic cocaine exposure increased
DFosB, cdk5, and BDNF in the NAc (as discussed
earlier). This increase is mirrored by an increase in
H3 acetylation at promoters of cdk5 (regulated by
DFosB) and BDNF (Kumar et al., 2005). Although these
studies indicated a link between histone modifications
and transcription factors in cocaine addiction, their role
was not clear because of the conflicting results.
However, the use of specific experimental manipulations in more recent studies has resolved this controversy. Kennedy et al. (2013) found that only prolonged

Drug-Induced Neuroplasticity

knockdown of HDAC1, not HDAC2 or HDAC3, in the


NAc reduced cocaine-induced behavioral plasticity.
Overexpression of HDAC3, 4, or 5 in the NAc, on the
other hand, suppressed cocaine-induced acquisition of
CPP; that is, the cocaine-context association took longer
to form (Rogge et al., 2013). Furthermore, chronic
cocaine exposure or stress decreases HDAC5 function
in the NAc; conversely, loss of HDAC5 in the NAc
increases the responsiveness to chronic cocaine or stress
as assessed by the CPP paradigm (Renthal et al., 2007).
The role of methylation, and the associated enzymes,
in cocaine addiction has been studied as well. Repeated
cocaine exposure reduces overall levels of histone 3
lysine 9 dimethylation and lysine dimethyltransferase
G9a in the NAc as well (Maze et al., 2010). DFosB
downregulates G9a, which results in reduced histone 3
lysine 9 dimethylation. Experimental reduction of G9a
increases dendritic spine plasticity in the NAc and
rewarding properties of cocaine as assessed by CPP.
This reduction in G9a is observed in both D1R- and D2RMSNs. However, the effects of experimental manipulation of G9a in the MSN subpopulations are different.
Similar to a nonspecific downregulation of G9a in MSNs
(Maze et al., 2010), knockout of G9a in D2R-MSNs
increased cocaine reward and locomotor sensitization,
whereas knockout in D1R-MSNs had the opposite
effects (Maze et al., 2014). Intriguingly, knockout of
G9a in D2R-MSNs resulted in these neurons having
a phenotype similar to D1R-MSNs.
On another note, experimental reduction of G9a in
the NAc also results in a depression-like phenotype.
This phenotype emerges on chronic cocaine exposure as
well. Indeed, increasing G9a activity in the NAc after
chronic cocaine exposure rescues the depression-like
phenotype (Covington et al., 2011). It has been suggested that G9a reduction by chronic cocaine exposure
might result in development of stress-related illness in
addicts. Further studies are warranted to determine if
this is indeed the case.
4. Progressive Involvement of the Dorsal Striatum.
A popular hypothesis for addiction proposes that addiction initially depends on the ventral striatum (including
the NAc), but progressively engages and hijacks the
circuitry in the dorsal striatum, which is normally
required for habitual learning. It is suggested that this
shift might mirror the change from voluntary to involuntary (or habitual) drug intake in addiction (Robbins
and Everitt, 1999; Everitt and Robbins, 2005, 2013).
This hypothesis fits well with the human literature on
cocaine addiction, which implicates the dorsal striatum
in cocaine addiction. Because a substantial amount of
human imaging studies compare cocaine addicts to
nonaddicted individuals, assuming this hypothesis
were true, it is not surprising that altered activity in
the dorsal striatum is observed. Consistent with this
hypothesis (and the human literature), Ito et al. (2000,
2002) observed increased DA release in the dorsal

895

striatum, but not NAc, on presentation of cocaineassociated cues after prolonged exposure to cocaine.
Similarly, blocking (or transiently inactivating) DA
receptors of dorsal striatum, but not those of NAc,
reduced the cocaine-seeking behaviors after protracted
cocaine exposure (Vanderschuren et al., 2005; Zapata
et al., 2010). On the basis of these findings, the authors
postulated that initial dependence on the VTA-NAc
pathway might be replaced by an increasing involvement of the nigrostriatal pathway as drug seeking
becomes more habitual (that is, on prolonged exposure
to the drug). It is well known that midbrain DAergic
projections to the striatum and backprojections to the
midbrain form a "spiraling" cascade: medial VTA projects to the NAc, which backprojects to lateral VTA, and
so on (progressively involving the dorsal striatum and
SN as well, Fig. 1)(Haber et al., 2000). To test whether
this circuitry might be involved in cocaine addiction, an
elegant study disconnected various parts of this serial
circuitry using alpha-fluphentixol (an antipsychotic
with strong antagonism at D1, D2, D3, and histamine
H1 receptors) and found a selective reduction in habitual drug seeking (Belin and Everitt, 2008).
Although the ventral striatum is critical for voluntary
behavior, a host of studies also implicate the dorsal
striatum in goal-directed (action-outcome) and habitual
(stimulus-response) behaviors. More specifically, the
dorsomedial striatum is involved in goal-directed behavior, whereas the dorsolateral striatum is involved in
habitual behavior (Yin et al., 2004, 2005a,b; Corbit and
Janak, 2010; Corbit et al., 2012) (Fig. 1). In agreement
with these findings, repeated cocaine exposure has been
shown to perturb the electrophysiological properties of
neurons in the dorsomedial striatum and to reduce goaldirected behavior as well. Furthermore, administration
of N-acetylcysteine reversed the behavioral deficits and
the associated changes in properties of dorsomedial
striatal neurons (Corbit et al., 2014) (see also discussion
in section II.D on alcohol).
5. Changes in Lateral Habenula and Rostromedial
Tegmental Nucleus Contribute to Aversive Symptoms of
Cocaine Withdrawal. Although "withdrawal symptoms"
from cocaine are not as pronounced as from some other
drugs of abuse, a depression-like phenotype is observed,
which might result in increased drug intake (Barr et al.,
2002; Koob and Le Moal, 2008). Thus, given the role of
the lateral habenula (LHb) in depressive disorders, it
is not surprising that an interest in its role in drug
addiction has recently emerged (Li et al., 2011;
Maroteaux and Mameli, 2012; Zuo et al., 2013; Lecca
et al., 2014). In this regard, earlier studies demonstrated that the LHb inhibits the VTA via its projections to the GABAergic rostromedial tegmental
nucleus (rmTg), promoting avoidance behaviors (Jhou
et al., 2009b; Kaufling et al., 2009; Balcita-Pedicino
et al., 2011; Stamatakis and Stuber, 2012) (Fig. 1). By
using an ex vivo preparation similar to earlier studies in

896

Korpi et al.

the VTA and NAc (discussed earlier), it was recently


shown that two prior in vivo, noncontingent doses of
cocaine selectively enhance glutamatergic transmission
onto LHb neurons projecting to the rmTg (as well as
increasing excitability of these neurons by reducing K+
currents) (Meye et al., 2013, 2015). These changes
persisted for 1 week, whereas a chronic cocaine regimen
(5 noncontingent doses) results in similar persistent
changes for 2 weeks (at which point, depressive-like
"withdrawal" behaviors were observed). Furthermore,
the changes were dependent on GluA1 subunitcontaining AMPAR trafficking in these neurons. Its
inhibition by virally expressed GluA1 C-terminal peptide resulted in a reduction in depression-like behaviors
on withdrawal and in persistent neuronal changes in
the LHb (Meye et al., 2015). Therefore, these studies
implicate the LHb-rmTg-VTA axis in the negative
symptoms associated with cocaine withdrawal. Interestingly, initially acute cocaine results in a reduction of
activity of a subpopulation of LHb neurons, followed by
increased activity, perhaps in putative rmTg-projecting
neurons (Lecca et al., 2014), suggesting that these
neurons might disinhibit VTA DA neurons during the
early pleasant experience on cocaine intake, but might
later reverse roles, leading to withdrawal symptoms.
Intriguingly, stimulation of the midbrain DAergic
structures modulates activity of pain-associated LHb
neurons as well, indicating that there are functionally
relevant connections from the VTA to LHb as well (Shen
et al., 2012). Obviously, further studies are required to
determine the roles of various LHb pathways in cocaine
addiction. In addition, to the aforementioned role of
LHb in cocaine addiction, stimulants have been shown
to induce strong excitation of LHb efferents to the rmTg
(further discussed in section II.A.11).
6. Complex Alterations in the PFC. The role of the
PFC in cocaine addiction is more complex. A single dose
of cocaine fails to elicit increased glutamate receptor
neuroplasticity in those VTA DA neurons that project
the mPFC (Lammel et al., 2011), although the same
treatment induces a prolonged increase (at least 21
days) in AMPAR/NMDAR current ratios of those VTA
DA neurons that project to NAc medial shell. The
mesocortical DA neurons might be more responsive to
aversive stimuli (Lammel et al., 2011). These findings
indicate that the PFC might not be involved in initial
cocaine-seeking behaviors.
DA neurons in the VTA responsive to aversive stimuli
project to the mPFC (Lammel et al., 2012). Interestingly, there is considerable evidence from early studies
implicating excitatory neurotransmission from the
mPFC to the NAc in cue-induced reinstatement (for
discussion, see Kalivas, 2009). However, the role of
mPFC inputs might be subtler. Optogenetic inhibition
of PrL-NAc core projections reduces cocaine-primed
reinstatement of cocaine seeking, and activation of the
PrL pathways is required for the cue-associated

transient increase of dendritic size and synaptic strength


in the NAc (Gipson et al., 2013; Stefanik et al., 2013b;
Shen et al., 2014). Furthermore, as discussed earlier,
Pascoli et al. (2014) found that optogenetic reversal of
cocaine-evoked plasticity in mPFC-NAc shell D1R-MSN
synapses attenuates response discrimination in cueinduced cocaine reinstatement. In addition, Ma et al.
(2014) present evidence that PrL-NAc inputs promote,
whereas IL-NAc inputs reduce, the enhancement of
cocaine craving during withdrawal. PrL neurons show
enhanced responsiveness to cocaine-associated cues, unlike the IL neurons (West et al., 2014). BDNF in the IL
promotes extinction of CPP (Otis et al., 2014). Taken
together, these findings indicate that PrL-NAc activation
promotes cue-induced cocaine seeking, whereas IL-NAc
activation promotes extinction of cocaine seeking.
On the other hand, pyramidal neurons in the PrL of the
mPFC in "compulsive" cocaine-seeking rats (a subset of
rodents that will seek cocaine despite negative consequences; more truly representative of addictive individuals) are less excitable after an extended SA protocol.
This compulsive cocaine seeking is increased or decreased
by optogenetic inhibition and activation of PrL pyramidal
neurons, respectively (Chen et al., 2013a). It is intriguing
to speculate that in these compulsive cocaine-seeking
rats, VTA-PrL synaptic excitability, which should convey
aversive values of encountered stimuli (Lammel et al.,
2011), is downregulated as well. This would reduce
aversion and hence enable cocaine seeking even at the
cost of "aversive consequences." However, further studies
are needed to determine if this indeed is the case. Another
possible explanation of these findings might be that these
rodents, due to reduced activity in the PrL, exhibit
reduced flexibility in switching behaviors (a function
subserved by the mPFC)or perseveranceon introduction of aversive consequences. Indeed, Allen and Leri
(2014) found that lesions of the PrL (which might have
similar functional consequences as hypoactivity of PrL
pyramidal neurons) resulted in increased perseverance
after chronic cocaine exposure.
Similar to the hypoactivity observed in the PrL, cocaine
SA also leads to a hypoactivity of pyramidal neurons in
the orbitofrontal cortex (OFC; although the underlying
mechanisms are different), a locus of decision-making and
insight in the brain (Lucantonio et al., 2014). Furthermore, this correlated with a lack of insight in the animals
(lack of expectation of a larger reward summation, when
two different cues were presented at the same time, as
shown by unchanged responding), which was normalized
by optogenetic stimulation of the OFC neurons. Moreover,
the VTA-OFC-BLA forms a functional circuit for contextinduced drug reinstatement. Specifically, contextual reinstatement of cocaine seeking in extinguished rats is
dependent on OFC-BLA activity and D1R signaling in the
OFC (Lasseter et al., 2011, 2014), as demonstrated by
local inactivation with microinfusions of GABAA/GABAB
agonist cocktail and D1R antagonist.

Drug-Induced Neuroplasticity

7. Hippocampal Functional Changes. Although the


HC is one of the principal regions in the brain involved
in spatial and contextual memory processing, not much
attention has been paid to the role of the HC in cocaine
addiction. An earlier study (Vorel et al., 2001) observed
reinstatement of cocaine-seeking behaviors after theta
burst electrical stimulation of the ventral subiculum
(part of the vHC). This reinstatement was mediated by
HC glutamatergic projections to the VTA, as it could be
mimicked by intra-VTA infusion of NMDA and blocked
by infusion of an NMDAR antagonist. Similarly, a trisynaptic pathway originating from the CA3 of the dorsal
HC to the VTA (via the lateral septum), which excites
DA neurons, is required for reinstatement of cocaine
seeking on exposure to the context (Luo et al., 2011).
Optogenetically inducing LTD in vHC-NAc D1R-MSN
synapses reduces the number of responses on cueassociated reinstatement of cocaine seeking, in contrast
to induction of LTD on mPFC-NAc D1R-MSN synapses,
which diminishes response discrimination (Pascoli
et al., 2014). Taken together, these findings indicate
that the HC plays a role in reinstatement of cocaineseeking behaviors. However, further studies are required to discern the precise roles of specific subregions
in the HC in cocaine addiction.
Like in other neuropsychiatric disorders, dysfunctional
adult neurogenesis has been observed to play a role in
cocaine addiction as well. Reducing hippocampal neurogenesis increases cocaine SA. Furthermore, neurogenesis
is also negatively correlated to chronic cocaine intake and
seeking behaviors (Noonan et al., 2008, 2010). Cocaine SA
reduces dentate gyrus neurogenesis, which returns back
to baseline during extinction (Deschaux et al., 2014). LFS
of the dorsal or ventral HC after extinction session
prevented the increase in neurogenesis and actually
enhanced cocaine seeking after a priming dose. Further
studies are required to elucidate the precise mechanisms
underlying these effects.
Selective reactivation of the dorsal dentate gyrus
neurons originally involved in encoding a rewarding
memory in an aversive environment reverses the valence
of that rewarding memory (its ability to evoke appetitive
behaviors as assessed by CPP) (Redondo et al., 2014). A
similar reactivation of the neurons originally active in the
BLA did not reverse the valence of the memory (Redondo
et al., 2014). However, selective ablation (CREB-Cre in
iDT mice expressing cell-specific diphtheria toxin receptor after cre-recombinase induction) or silencing
(chemicogenetic silencing with CREB-hM4Di designer
receptor exclusively activated by a designer drug receptors) of "cocaine-memory-trace"-activated neurons in the
lateral nucleus of amygdala (LA; about 10% of the
neurons that increased expression of CREB in response
to cocaine administration) disrupted the expression of
a previously acquired cocaine-context-related memory,
because the animals failed to show CPP (Hsiang et al.,
2014). Interestingly, when these animals were taken

897

through extinction session and tested later for cocaineinduced reinstatement, the small number of previously
activated LA neurons was needed also for this kind of
drug memory. Taken together, these findings indicate
that in these HC and amygdala structures, a small subset
of neurons are required for reward processing and are
possibly hijacked during cocaine-seeking behaviors.
The hippocampal formation plays a role in spatial
navigation and memory consolidation. In this regard,
HC "place cells" fire in a specific location in an environment (OKeefe and Dostrovsky, 1971; Colgin et al.,
2008). When the animal traverses space, sequential
HC place cell firing is observed, encoding the animals
trajectory. These sequences are replayed during subsequent sleep and quiet wakefulness, which are proposed to consolidate this memory (Skaggs and
McNaughton, 1996; Davidson et al., 2009). However, it
was recently demonstrated that replay or reactivation
of HC activity does not simply depend on experience,
but might be modulated by the behavioral state of the
animal, including inputs from the VTA, and the environment (Gupta et al., 2010; McNamara et al., 2014).
Such off-line reactivation of neurons associated with
rewarding behaviors is observed in the NAc as well
(Lansink et al., 2008), possibly providing consolidating
memories with a motivational component. Interestingly, using multineuron recordings, reactivation of
HC place cells leads to activation of striatal reward
site-related neurons as ensembles in sleep, followed
shortly after HC sharp wave-ripple complexes, suggesting that the HC plays a pivotal role in consolidation of
reward-related memories (Lansink et al., 2009). Furthermore, contextual fear conditioning remaps HC cells
(Moita et al., 2004). These studies indicate that the HC
is critical for forming associations between the context
and appetitive or aversive behaviors. This raises the
intriguing possibility that cocaine might alter hippocampal ensemble activity, and thus spatial memory
processing of the cocaine-associated cues and context.
8. Bed Nucleus of Stria Terminalis and Amygdala
are Involved in Stress-Induced Reinstatement of Cocaine Seeking. The BNST is required for stress-, cue-,
or yohimbine-associated reinstatement of cocaine seeking (Buffalari and See, 2011), based on experimental
inactivation experiments of the BNST with locally
infused GABAA and GABAB agonists. CRF in the BNST,
possibly via projections from the CeA, induces reinstatement (Erb and Stewart, 1999; Erb et al., 2001a,b).
Furthermore, noradrenergic neurotransmission in the
BNST or CeA is required for stress-induced but not
cocaine-induced reinstatement (Leri et al., 2002). Finally, stress-induced reinstatement of cocaine seeking
requires activity in a ventral BNST-VTA pathway,
which releases CRF in the VTA. This release of CRF is
modulated by b2 adrenergic receptors in the BNST
(Vranjkovic et al., 2014). Therefore, the CeA-BNSTVTA system forms a functional circuit involved in

898

Korpi et al.

stress-induced reinstatement of cocaine seeking. The


connections are complicated, as illustrated by reciprocal
connections between the BNST and VTA, with both
GABA and glutamate neurons from the BNST targeting
VTA GABA neurons. When selectively activated by
optogenetics, BNST-VTA GABA projection disinhibits
the VTA DA neurons, whereas BNST-VTA glutamate
projection indirectly inhibits the VTA DA neurons,
producing rewarding and aversive/anxiety-like responding, respectively (Jennings et al., 2013).
9. Effects of Adolescent Cocaine Exposure.
Adolescents are more prone to develop cocaine addiction
than adults; they administer cocaine more readily, are
more sensitive to lower doses, show increased activation
of VTA DA neurons, and show greater escalation of
cocaine intake (reviewed in Gulley and Juraska, 2013).
Indeed, a significant proportion of studies (specifically,
the majority of ex vivo electrophysiology studies) discussed in this review were performed in adolescent
animals, because the experiments are less tedious and
the results more pronounced. However, the effect of
adolescent exposure to cocaine on adult behavior has
not been widely studied.
Of note are two studies, which employed a novel
escalating "binging" cocaine administration protocol in
adolescent rats (12 days between P35 to P46) and its
effects on subsequent behavior during adulthood. This
binging protocol was aimed at recapitulating human
adolescent drug use. Black et al. (2006) demonstrated
that these rats displayed increased responsiveness to the
stimulant effects of cocaine in adulthood. Furthermore,
they exhibited "abnormally" rapid shifts in attention on
the attentional set-shifting task, implicating the PFC in
this dysfunction. A similar cocaine-binging protocol in
adolescents subsequently decreased anxiety and acquisition of contextual fear responses in adulthood (Sillivan
et al., 2011). Furthermore, this protocol perturbed dendritic arborization and synapses, which persisted in
adulthood. Taken together, these studies provide the first
insights into the effects of adolescent cocaine exposure on
subsequent behavior in adulthood. However, because no
adult controls (Black et al., 2006; Sillivan et al., 2011)
with binge administration of cocaine were used, the
possibility that these changes might be due to binge
administration of cocaine per se and not due to exposure
in adolescence cannot be ruled out. Nonetheless, further
studies are required to elucidate these changes.
10. Human Imaging Studies. Imaging studies
attempting to elucidate the neural mechanisms underlying cocaine addiction in humans have been of limited
scope due to a variety of reasons, including poor resolution, difficulty in quantifying the cocaine abuse history of
the participants, and problems with determining causality. However, human studies have the distinct advantage
over animal models of providing information about the
subjective effects of the drugs. These subjective effects,
nonetheless, are prone to a variety of problems. For

instance, it was recently argued that researchers studying the neural correlates of "craving" in drug addicts
might be inadvertently studying a low-level urge (Moeller
et al., 2015; Wilson and Sayette, 2015). Moreover, acute
as well as chronic cocaine intake (even after a 10-day
withdrawal period) reduces cerebral blood flow (possibly
due to vasoconstriction, because it is a sympathomimetic),
which might make the interpretation of experiments
studying glucose metabolism and blood oxygenation
(considered correlates of neural activity) more difficult
(Volkow et al., 1988; London et al., 1990; Wallace et al.,
1996). However, Gollub et al. (1998) argue that regional
changes in neuronal activity using the aforementioned
techniques can be reliably determined despite the global
reduction in blood flow. Nonetheless, one has to be careful
while drawing comparisons across studies. Another
important caveat in human brain imaging studies is the
effect of expectancy. For example, the apparent DA
release in the human NAc is significantly enhanced
[reduced binding potential of raclopride in positron
emission tomography (PET) scanning] by expectation of
receiving caffeine (Kaasinen et al., 2004). Notwithstanding these limitations, significant insights into the mechanisms underlying cocaine addiction have been achieved
using neuroimaging and neurophysiological techniques
including fMRI and PET. In cocaine-dependent individuals, intravenous infusion of cocaine induces increased
fMRI signals in the reward network of the brain,
correlating either with self-reported "rush, high, low, or
craving" (Breiter et al., 1997). Importantly, repeated
sessions reproduced the findings, but activations also
emerged in some saline infusions, which were interpreted
as expectations. The effects were registered within 100
seconds, but still the dependent individuals might have
differentiated the active drug from saline, which might
have altered the responses. However, the pharmacological effects of cocaine in the reward areas might not be
altered by expectation (Kufahl et al., 2008).
A seminal PET study revealed that the DAT occupancy by cocaine correlates with the subjective feeling of
the "high" in cocaine addicts (Volkow et al., 1997a). This
drug-induced high is negatively correlated to activity in
the NAc, OFC, and anterior cingulate cortex (Risinger
et al., 2005). Cocaine addicts have lower levels of DA
receptors in the striatum than controls, when assessed
using PET after 1 week of abstinence. However, 1 month
of abstinence is sufficient to return the receptors to
normal levels (Volkow et al., 1990). Furthermore, acute
exposure to methylphenidate, a DAT blocker like cocaine (and a drug that is reported to produce a similar
high as cocaine in addicts), demonstrates a reduced
release of DA in the striatum in cocaine addicts, even
after extended withdrawal (Volkow et al., 1997b; Martinez
et al., 2004;). Moreover, Gu et al. (2010), using fMRI, found
that resting state functional connectivity between the
VTA and NAc was negatively correlated to years of
cocaine use, further implicating a downregulation in

Drug-Induced Neuroplasticity

DAergic signaling in cocaine addicts. Studies in macaques


extend these findings by demonstrating that chronic
cocaine SA per se reduces D2Rs, which might contribute
to increased drug use (reviewed in Nader and Czoty, 2005;
Nader et al., 2006). In human cocaine addicts, D2Rs are
reduced in several regions of the frontal cortex (FC,
particularly the OFC and cingulate gyri), which is
apparent even after 34 months of abstinence (Volkow
et al., 1993). Similarly, the anterior cingulate cortex, along
with the OFC and dorsolateral PFC, is hypoactive in
response to rewarded drug cue-reactivity task (Goldstein
et al., 2009). This deficit is reversed by oral methylphenidate and is not a consequence of reduced motivation or
impaired behavioral responses, because these were controlled for in the experiments (Goldstein et al., 2009,
2010). Intriguingly, acute exposure to methylphenidate,
presumably by raising DA levels, increases metabolism in
the OFC and mPFC of cocaine addicts but not controls
(Volkow et al., 2005). Nonetheless, cocaine abusers (active
or abstinent) exhibit reduced signaling via D2/3 receptors
compared with controls on exposure to methylphenidate
(Volkow et al., 2014). Taken together, these findings
indicate that cocaine abuse results in persistently reduced
DA release, reduced DA receptor signaling, and hypoactive striatal and frontal circuits.
Strikingly, although chronic cocaine users exhibit
reduced cerebral blood flow on withdrawal as discussed
(Volkow et al., 1988), after 1 week of withdrawal,
increased global brain glucose metabolism and also
regional increases in the basal ganglia and OFC are
observed, which subside in 24 weeks (Volkow et al.,
1991). Further studies are required to understand these
differences.
As highlighted earlier, recent work has challenged the
construct validity of experiments studying craving in
addiction (Moeller et al., 2015; Wilson and Sayette, 2015).
Nonetheless, a study employing PET reported increased
cerebral blood flow to limbic regions (including the
amygdala and anterior cingulate gyri) and decreased
blood flow to the basal ganglia on cue-induced cocaine
craving compared with controls (Childress et al., 1999).
These findings were corroborated by fMRI studies, which
demonstrate increased cingulate and low frontal lobe
activation in cocaine addicts on exposure to cocaineassociated cues (Maas et al., 1998; Wexler et al., 2001).
Risinger et al. (2005) studied cocaine SA by
nontreatment-seeking cocaine addicts. Simultaneously
with SA-associated behavioral ratings they also had the
subjects scanned with fMRI. Craving was positively
correlated with activity in the NAc, OFC, and anterior
cingulate cortices, whereas cocaine-induced high was
negatively correlated with the activity in these regions,
stressing the altered states of the reward circuitry in
cocaine-related subjective states.
11. Limited Evidence of Cocaine Neurotoxicity.
In humans, PET studies have reported decreased blood
flow and frontal cortical glucose metabolism as well as

899

decreased D2 receptor availability in cocaine users


compared with control subjects. These changes were
sustained up to 4 months of abstinence (Volkow et al.,
1988, 1992, 1993). Cocaine users also displayed an
attenuation of the increase in extracellular striatal DA
levels but a greater increase in thalamic DA levels
compared with control subjects in response to a methylphenidate challenge (Volkow et al., 1997b). Studies
employing SPECT and MRI have reported changes in
regional cerebral blood flow and blood oxygen level
dependent (BOLD) signal intensity in cocaine users
while performing cognitive tests measuring memory
and executive functions. Alterations have been observed in cortical regions, the globus pallidus, putamen,
cingulate, hippocampus, and amygdala, but are not
always associated with decreased task performance
(Ernst et al., 2000; Bolla et al., 2003; Hester and
Garavan, 2004; Tomasi et al., 2007; Moeller et al.,
2010). Importantly, these studies are subject to the
same limitations (discussed in more detail in section II.
B on amphetamines) that limit the conclusions that can
be drawn from them.
Unlike amphetamines (see section II.B), the evidence
of cocaine neurotoxicity in animal models is more
limited. In rodents, cocaine does not produce any longterm depletion of DA, 5-HT, or their metabolites in
brain regions normally affected by amphetamines, such
as the striatum, cortex, hypothalamus, and hippocampus (Kleven et al., 1988; Yeh and De Souza, 1991). Other
markers of monoamine system integrity, such as monoamine transporter levels, are likewise unaffected by
binge cocaine treatments. In studies employing silver
staining, even very high-dose, multiday treatments
with cocaine up to 450 mg/kg per day failed to produce
any evidence of axonal degeneration in the striatum,
cortex, or other regions (Ryan et al., 1988; Benmansour
et al., 1992; Goodman and Sloviter, 1993). Even treatment at high ambient temperatures, an exacerbating
factor in amphetamine toxicity, did not result in any
cocaine neurotoxicity (Cappon et al., 1998).
Although widespread toxicity is generally not observed, axonal degeneration after cocaine binge treatments has been reported in one specific region
extending from the Hb toward the VTA (Ellison,
1992). The neurodegeneration in this region is not
specific to cocaine, however, and has been reported
across a wide range of stimulant drugs and also
nicotine. Interestingly, this degeneration has been
implicated to play a role in the loss of forebrain control
circuitry and development of behaviors related to binge
drug use and relapse (Carlson et al., 2000; Ellison,
2002). The activation of LHb activates the "tail" region
of the VTA, which is the rostromedial tegmental nucleus
(rmTg) (Jhou et al., 2009a). Screening of the effects of
many pharmacological agents on the GABAergic cells of
the rmTg has been carried out by monitoring the
induction of c-Fos or DFosB expression in acute

900

Korpi et al.

experiments (Perrotti et al., 2005; Kaufling et al., 2010).


The robust activation of these neurons was only shared
by stimulant drugs, including cocaine, amphetamines,
methylphenidate, and caffeine, and by a DAT blocker,
but not by SERT or NET blockers, indicating that strong
DAergic activity was needed to activate the network
that involves the LHb. Many other groups of drugs
of abuse do not strongly activate the neurons in the
rmTg (Fig. 1), including morphine, ethanol, diazepam,
D9-tetrahydrocannabinol, ketamine, and phencyclidine
(Perrotti et al., 2005; Kaufling et al., 2010). Nicotine
preferentially activates, and degenerates at high doses,
the medial habenula-interpeduncular (MHb-IPN) pathway after subchronic administration (Carlson et al.,
2001; Ciani et al., 2005) (see Fig. 10), but it also acutely
excites the rmTg neurons in anesthetized rats (Lecca
et al., 2011).
12. Potential Mechanisms in Cocaine Neurotoxicity
and Neuroprotection. Some data suggest that, under
some circumstances, cocaine can function as a stressor
or as a neuroprotective agent. In cultured fetal mesencephalic DA neurons, a 5-day cocaine treatment (1025
and 1024 M) produced no effect on levels of TH-positive
cells or high-affinity [3H]DA uptake (Bennett et al.,
1993). However, other studies employing PC12 cells
found that cocaine treatment produced DNA fragmentation, cytotoxicity, decreases in Bcl-2, and increases in
caspase-3 levels (Gramage et al., 2008; Lepsch et al.,
2009). Furthermore, the cocaine metabolite benzoylecgonine has been reported to be cytotoxic (Lin and
Leskawa, 1994). Oxidative stress has been implicated
in cocaine toxicity. In vitro, cocaine toxicity can be
attenuated by application of the superoxide dismutasemimetic tempol (Numa et al., 2011), and in vivo high
doses of cocaine lead to increased levels of oxidative
stress, hydrogen peroxide, and lipid peroxidation in the
striatum and HC of rats for up to 2 days after treatment.
However, these changes coincide with increases in
levels of glutathione peroxidase and superoxide dismutase, suggesting that compensatory mechanisms may
prevent extensive long-term damage (Dietrich et al.,
2005; Numa et al., 2008). Oxidative stress may be further
increased due to cocaines reported effects on mitochondria, which include decreases in levels of complex I
subunit gene expression and membrane depolarization
(Dietrich et al., 2004; Cunha-Oliveira et al., 2006). These
and other factors that have been implicated in cocaine
toxicity, such as effects on the glutamatergic system,
neurogenesis, neuroinflammation, and blood-brain barrier permeability are discussed above and reviewed in
more detail elsewhere (Goncalves et al., 2014).
Cocaine purchased by consumers commonly contains
adulterants such as levamisole that, together with its
major metabolite aminorex, are also pharmacologically
active. Although levamisole is only a rather weak reuptake inhibitor, the metabolite aminorex was found to
exert strong reuptake inhibiting and releasing properties

similar to amphetamine (Hofmaier et al., 2014). Aminorex was previously reported to not produce any longlasting depletion of brain monoamines (Zheng et al.,
1997), but the potential toxicity of these and other cocaine
adulterants, particularly in combination with cocaine, is
a topic of interest for harm reduction requiring further
investigation.
Aside from possible toxic effects, it is interesting to
note that cocaine actually appears to protect against
toxicity of amphetamines. This effect has been linked to
the reduction of cellular uptake of the amphetamines
after cocaine, for instance due to translocation of DAT
from the plasma membrane to endosomes (Daza-Losada
et al., 2008; Peraile et al., 2010).
13. Methylphenidate and Novel Psychoactive Substances.
Aside from cocaine, there are several drugs in clinical
and recreational use that have a very similar mechanism of action. Most important in this regard is
methylphenidate. As with cocaine, there are reports of
increased oxidative stress and lipid peroxidation in
certain brain regions after methylphenidate, whereas
long-term depletion of monoamine levels is not observed
(Yuan et al., 1997; Schmitz et al., 2012; Comim et al.,
2014). Analogous with cocaine, methylphenidate also
protects against toxicity of amphetamines, and it has
been suggested it may also be protective against the
neurodegenerative processes occurring in Parkinsons
disease (Volz, 2008).
During the increase in novel psychoactive substance
use in recent years, one of the compounds that became
very widely used was 3,4-methylenedioxypyrovalerone,
a compound structurally related to amphetamines, but
with more extensive substitutions, including an extended carbon chain instead of a methyl group at the
a-carbon. Hence, the drug does not function as a transporter substrate but, like cocaine, strongly inhibits
monoamine reuptake. It is most potent at the DAT
and NET and to a lesser extent at the SERT (Baumann
et al., 2013; Cameron et al., 2013; Iversen et al., 2014). A
further investigation of the potential neurotoxic and
neuroprotective effects of 3,4-methylenedioxypyrovalerone and similar novel compounds, e.g., mephedrone,
could be of interest.
14. Conclusions. Basic research on cocaine has produced a large amount of detailed neurobiological,
molecular biologic, and gene regulation data (for a summary, see Fig. 7), which help to understand the biologic
background of various phases of cocaine addiction in
different reward/emotion/cognition-related brain regions.
Unfortunately, this vast database has not yet produced
enough translational results to treat cocaine-addicted
patients.
B. Amphetamine-type Psychostimulants
DA plays an important role in reward and motivation,
two aspects of behavior, which are central to goaldirected behavior for both human as well as nonhuman

Drug-Induced Neuroplasticity

animals (Salamone, 1994; Koob, 1996). Reward can be


viewed as the subjective pleasurable sensation associated with fulfilling a physiologic need, such as quenching hunger or thirst with food or water, whereas
motivation is the subsequent willingness to exert effort
in performing goal-directed behavior to achieve these
rewards. The mesolimbic DA system, consisting of
projections from the VTA to the NAc plays an important
role in mediating reward and motivation (Wise, 2006)
(Fig. 1). Recreational drug use is also associated with
pleasurable sensations (Fischman and Foltin, 1991). So
it should come as no surprise that all drugs that produce
these sensations also produce increases in the DA

901

neurotransmission in the NAc in a similar, but more


robust, way as do natural or physiologically important
rewarding stimuli (Wise and Rompre, 1989). Amphetamines such as AMPH and METH are particularly
interesting in this respect as they, unlike other more
indirectly acting drugs, except for cocaine, preferentially increase DA levels in the NAc (Wise and Rompre,
1989) via several mechanisms described below (Fig. 5;
Table 1).
As will be described in more detail in the rest of this
section, the primary effect of amphetamines is to increase monoamine neurotransmission by inhibiting the
reuptake and enhancing the release of the monoamines

Fig. 7. Summary of main neuroplasticity-related changes observed in selected brain regions during different stages of cocaine addiction. The colored
arrows indicate the stages of addiction depicted inside the arrows.

902

Korpi et al.

DA, NE, and 5-HT through the DAT. However, AMPH


also affects secretory vesicles, enzymes involved in DA
biogenesis and metabolism, and many other targets.
AMPH-type stimulants have been employed for a long
time, both for medical and recreational purposes. This
raises the question of what, other than its initial effects
on brain neurochemistry, the long-term effects AMPH
exposure may produce.
1. Many Stimulant Amphetamines for Abuse and
Therapy. The biggest and most diverse group of
psychostimulants is the amphetamines. In its most
basic form, amphetamine (AMPH), a contraction of
a-methyl-phenethylamine, consists of a phenyl ring
connected to an amino group by a two-carbon side chain
with a methyl group on carbon 1. The addition of
a methyl group on the nitrogen atom produces methamphetamine (METH). Various other ring and side
chain substitutions produce other well-known amphetamine analogs such as methylphenidate (MPH) and
3,4-methylenedioxymethamphetamine (MDMA, ecstasy).
We will focus here on the amphetamines in general and
make notes on specific qualities of individual compounds as necessary. The amphetamines all exist as
isomers with subtle differences in their neurochemical
and behavioral properties. For AMPH, the d-isomer
appears to be a more potent DA releaser, whereas the
l-isomer is an equally or more potent releaser of NE
(Heikkila et al., 1975; Holmes and Rutledge, 1976;
Mendelson et al., 2006). Furthermore, the d-isomers
of AMPH and METH appear more potent than the
l-isomers in producing behavioral effects such as locomotor activity, SA, and taste aversion (Balster and
Schuster, 1973; Yokel and Pickens, 1973; Carey and
Goodall, 1974; Segal, 1975). Illicit METH is usually
distributed as either a racemic mixture of d- and
l-isomers or as pure d-isomer, depending on the method
of synthesis most prevalent at the time (Logan, 2002;
Mendelson et al., 2006).
Psychostimulants have been widely used medically
(Iversen, 2006), but presently their therapeutic use is
more limited. AMPH and MPH are the first line treatment of attention deficit hyperactivity disorder (ADHD),
a common disorder primarily characterized by decreased
impulse control and attention, affecting approximately
4% of the population (Surman et al., 2013). They are also
used for management of narcolepsy (De la Herran-Arita
and Garcia-Garcia, 2013). Furthermore, although not
currently approved, recent clinical trials have provided
evidence that MDMA may be effective in the treatment of
otherwise treatment-resistant anxiety disorders (Mithoefer
et al., 2011, 2013). Recreational use of amphetamines,
particularly AMPH, METH, MDMA, and more recently,
various b-ketoamphetamines is also widespread, with
amphetamine-type stimulants being the worlds second
most consumed illegal drugs after cannabis (UNODC,
2011). Like other psychostimulants such as cocaine, the
subjective effects of amphetamines are associated with

increases in DA and include euphoria, increased energy,


insomnia, and decreased appetite. However, excessive
DA stimulation can also lead to amphetamine psychosis,
a condition that shows strong similarity to the psychotic
episodes experienced by patients suffering from schizophrenia and which constitutes one of the primary reasons
for amphetamine-related hospitalization (Snyder, 1973;
McKetin et al., 2013; Medhus et al., 2013).
2. Molecular Targets and Mechanisms of Action of
Amphetamines. Historically, the mechanisms that
have received the most attention are the ability of
amphetamines to inhibit the reuptake and enhance
the release of DA via the DAT, as well as their ability to
promote the release of DA from secretory vesicles
(Sulzer et al., 2005; Fleckenstein et al., 2007)(Fig. 5).
Recently, it has become clear that, aside from these
traditional mechanisms, amphetamines are also capable of modulating action potential-dependent neurotransmitter release (Branch and Beckstead, 2012;
Daberkow et al., 2013). Furthermore, amphetamines
regulate DA neurotransmission via inhibition of monoamine oxidase (MAO) and enhancement of tyrosine
hydroxylase (TH) and have numerous other receptor
targets, including the a2-adrenergic receptor, the sigma
receptor, and the trace amine-associated receptor 1
(TA1) receptor (Fung and Uretsky, 1982; Robinson,
1985; Ritz and Kuhar, 1989; Matsumoto et al., 2014;
Reese et al., 2014). Many of these targets have been
identified only recently, and their importance in mediating the effects of amphetamines on behavior as well as
its therapeutic, reinforcing, and toxic effects are just
beginning to be understood.
3. Comparison of the Mechanisms of Action of
Amphetamine versus Methamphetamine. It is sometimes stated that METH is more addictive and more
powerful than AMPH, but this statement has been hard
to back up empirically. The changes in extracellular DA
levels measured by microdialysis in rats are correlated to
the striatal drug levels and are equal for the d-isomers of
both drugs (Melega et al., 1995). One study confirmed
that d-AMPH and d-METH produce similar increases in
NAc DA levels in rats and actually found that AMPH
increases PFC DA levels more than METH and also
produces a higher peak locomotor activity (Shoblock et al.,
2003b). In line with the effect on PFC DA levels, it was
discovered that AMPH has a stronger effect than METH
on working memory function. The effects of AMPH and
METH on working memory are bimodal, with a small
dose producing an increase in working memory performance and higher doses producing a decrease, an effect
closely reminiscent of the bimodal effects of DA turnover
and D1 receptor activation on working memory (Murphy
et al., 1996; Vijayraghavan et al., 2007). For d-AMPH,
a dose of 0.5 mg/kg produces an increase in working
memory performance, whereas a 2 mg/kg dose produces
a decrease. d-METH on the other hand has little effect at
0.5 mg/kg, producing the highest increase at 2 mg/kg and

Drug-Induced Neuroplasticity

only decreasing at 4 mg/kg (Shoblock et al., 2003a).


Furthermore, d-AMPH fully substitutes for d-METH in
animals trained to discriminate METH from placebo
(Desai and Bergman, 2010) and also causes dosedependent increases in METH-appropriate responding
in humans who have learned to discriminate 10 mg
METH from placebo (Sevak et al., 2009), suggesting that
the subjective effects are also very similar, if not indistinguishable. Recently, it was also shown that varying
doses (between 12 and 50 mg intranasally) of d-AMPH or
METH produce similar subjective effects and did not
differ from each other with regards to the rate at which
participants would opt for a cash reward instead of a dose
of the drug (Kirkpatrick et al., 2012), thereby not
corroborating the notion that METH is more addictive
or reinforcing than d-AMPH.
By tradition, scientific work on the mechanism of
action of amphetamines has been primarily done with
AMPH, whereas studies on neurodegeneration preferentially have employed METH (Sulzer et al., 2005).
4. Effects of Amphetamine at Plasmalemmal Dopamine Transporter. The DAT as well as the other
monoamine transporters, NET and SERT, belong to
the SLC6 gene family of secondary active transporters
consisting of 12 transmembrane domains as well as
intracellular domains with phosphorylation and binding sites vital for its regulation, including by AMPH.
Normally, the transporters depend on pre-existing
concentration gradients of Na+ and Cl2 to cotransport
one molecule of Na+ together with one neurotransmitter
molecule from the extracellular space into the cytoplasm (Robertson et al., 2009).
The ability of AMPH to increase extracellular DA
levels in vivo is well established, and one area particularly sensitive to AMPH-induced increases in extracellular DA is the NAc (Carboni et al., 1989). AMPH blocks the
reuptake of DA via the DAT (Parker and Cubeddu, 1988).
However, intracytoplasmic injections of AMPH also increase extracellular DA, and this effect is blocked by the
DAT blocking drug nomifensine, indicating that AMPH,
aside from blocking the reuptake, also enhances the
release of intracellular DA (Sulzer et al., 1995). The
uptake of radiolabeled AMPH into striatal synaptosomes
is a saturable and high-affinity process that is sensitive to
ouabain and cocaine and dependent on temperature,
suggesting that AMPH is itself a substrate of the DAT
(Sitte et al., 1998; Zaczek et al., 1991). In addition, it is
important to note that AMPH is a lipophilic weak base
(Mack and Bonisch, 1979; Gulaboski et al., 2007), which
will, particularly at higher concentrations, diffuse
through the plasma membrane and enter the cytoplasm
independently of the DAT.
An early model (Fischer and Cho, 1979) known as the
exchange-diffusion model, postulated that AMPH, like
DA, would bind to an extracellular binding site, causing
a conformational change in the DAT protein resulting in
the binding site traversing the membrane and releasing

903

AMPH into the cytoplasm and simultaneously binding


a DA molecule to be transported outward. Later it would
become clear that this exchange-diffusion is not limited
to a 1:1 ratio of AMPH exchange for DA. AMPH also
increases intracellular levels of Na+, and increases in
Na+ are in itself sufficient to induce release of DA via the
DAT; it is likely that it is the influx of Na+ ions via the
DAT rather than AMPH itself that triggers reverse
transport of DA. Furthermore, the DAT can be regulated, as discussed below, via intracellular second
messengers between reluctant and willing states
for AMPH-mediated DA efflux without affecting inward
transport (Khoshbouei et al., 2004). AMPH is also be
capable of producing rapid bursts of DA release via the
DAT, which is indicative of AMPH causing a conformational change to the DAT resulting in a channel-like
state in which DA is rapidly transported outward via
a pore in the transporter protein. The process is
transient, consisting of millisecond bursts and is
inhibited by the presence of DA on both sides of the
plasma membrane and therefore not capable of transporting DA against its concentration gradient (Kahlig
et al., 2005). Finally, recent evidence suggests that
AMPH also affects the DA release associated with
action potentials. One study that examined this demonstrated that AMPH at 10 mg/kg enhanced electrically
evoked DA release measured 10 and 30 minutes after
the initial injection. Additionally, it was shown that
a low dose (1 mg/kg) of AMPH facilitated the DA release
associated with a rewarding stimulus cue, suggesting
that AMPH acts, at least in part, by altering the
characteristics of action potential-dependent exocytotic
DA release (Daberkow et al., 2013). Unlike AMPH, MPH
was recently shown to decrease exocytotic DA release
(Federici et al., 2014), which may help explain the
differing subjective effects of this drug. The exact mechanism by which AMPH increases the exocytotic release of
DA is unknown, but may be related to the ability of
AMPH to induce persistent transporter-mediated ionleakage and excitatory conductance (Ingram et al., 2002;
Branch and Beckstead, 2012; Rodriguez-Menchaca et al.,
2012). It has been suggested that reuptake inhibition is
primarily associated with lower drug concentrations,
possibly corresponding to those in the therapeutic range,
whereas doses in the recreational range shift the importance to AMPH-mediated DA release (Calipari and
Ferris, 2013). Furthermore, there appear to be regionspecific changes in the relative importance of DATdependent and action potential-dependent DA release,
with enhancement of vesicular release being more important in the dorsal striatum compared with the ventral
striatum (Avelar et al., 2013). As discussed further
below, the final increases in cortical and NAc DA levels
mediate both the therapeutic and reinforcing effects of
amphetamines.
5. Regulation of Dopamine Efflux Via the Dopamine
Transporter by Protein Kinase C and Ca2+/Calmodulin

904

Korpi et al.

Kinase II Phosphorylation, and by Reactive Oxygen


Species. The release of DA via the DAT is highly
dependent on DAT expression on the surface of the
plasma membrane and is regulated by several intracellular second-messenger systems. Phosphorylation by
PKC plays an important role in regulation of DAT
function and surface expression. PKC-activating and
phosphatase-inhibiting drugs both decrease the uptake
of DA into striatal synaptosomes in vitro. The effect is
inhibited by the PKC inhibitor staurosporine (Copeland
et al., 1996; Zhang et al., 1997). Studies in striatal
synaptosomes employing [32P]orthophosphate to assess
the amount of DAT phosphorylation confirmed that
phosphorylation at DAT is regulated by PKC and that
the amount of DAT phosphorylation is directly related
to the efficacy of DA uptake, with increased phosphorylation decreasing the uptake efficacy (Vaughan et al.,
1997). PKC activation removes DAT from the membrane to endosomes, whereas PKC inhibition promotes
the insertion of DAT into the membrane, something
that has been confirmed using immunofluorescence
confocal microscopy (Pristupa et al., 1998; Sorkina
et al., 2003). In addition to affecting DA uptake, PKCmediated DAT phosphorylation also affects AMPHinduced DA release. One study found that PKC-inhibiting
drugs almost completely inhibited AMPH-induced DA
efflux, while increasing baseline DA uptake in the
absence of AMPH (Kantor and Gnegy, 1998). The effect
of PKC-mediated DAT phosphorylation on AMPHinduced DA release appears to be mediated specifically
by the PKCbII isoform (Johnson et al., 2005). Thus,
whereas increases in phosphorylation appear to decrease the efficacy of baseline DA uptake and decrease
cell surface DAT expression, the opposite appears to be
the case for AMPH-mediated DA efflux, which is reduced by inhibition of phosphorylation.
METH is capable of inducing DAT phosphorylation
both in vivo and in vitro (Cervinski et al., 2005). A single
in vivo injection of METH rapidly (within 1 hour) and
reversibly decreased plasmalemmal DA uptake,
whereas the total binding of the membrane-permeable
DAT ligand WIN35428 was not affected, indicative of
METH-induced transient transporter internalization or
loss of function (Fleckenstein et al., 1997b; Kokoshka
et al., 1998; Saunders et al., 2000). These data suggest
that AMPH produces transporter internalization in
a similar way as does PKC activation. However, there
appear to be subtle differences between PKC-dependent
and AMPH-dependent DAT internalization. Both AMPH
and PKC activation results in loss of cell surface DAT in
vitro, but contrary to the PKC-induced internalization
that is dependent on DAT ubiquitination and subsequent
sorting to LAMP1-positive endosomes for degradation,
the AMPH-induced internalization sorted the DAT to
Rab11-positive recycling endosomes (Hong and Amara,
2013), likely destined to be quickly reinserted into the
membrane.

AMPH-mediated DA efflux appears to be influenced


also by CaMKIIa. CaMKIIa binds the DAT at the
distant C terminus while inducing phosphorylation at
similar residues as PKC, namely the N-terminal serines. Phosphorylation by this kinase increases AMPHinduced DA efflux via DAT, and mutations at both the
C-terminal binding site as well as the N-terminal
phosphorylation sites inhibit the increase (Fog et al.,
2006). CaMKII appears to influence AMPH-mediated
DA efflux via the DAT through Syntaxin A1, a protein
that, aside from being involved in synaptic vesicle
release, interacts with and regulates transmembrane
proteins, including DAT (Binda et al., 2008; Robertson
et al., 2009). Interestingly, association of Syntaxin A1
with the DAT is also subject to regulation by AMPH
itself (Binda et al., 2008), pointing to another target by
which AMPH itself regulates AMPH-dependent, DATmediated DA efflux.
There is also evidence suggesting that ROS signaling
may be involved in regulating cell surface DAT expression (Fleckenstein et al., 2007). The oxygen radicalgenerating enzyme xanthine oxidase reduces DA
uptake into striatal synaptosomes, an effect inhibited
by coapplication of the free radical scavenger superoxide dismutase (Berman et al., 1996; Fleckenstein et al.,
1997a). As discussed in more detail below, AMPH is also
capable of generating ROS, so this regulatory pathway
is also open to modulation by AMPH itself.
Regulation of DAT function via phosphorylation and
ROS are some mechanisms by which DAT surface
expression and function is regulated. Because these
mechanisms are affected by the presence of AMPH
itself, it appears that the final amount of DA released
into the extracellular space in response to AMPH
depends on a complex interplay between AMPH and
intracellular second-messenger systems, some of which
may not have been discovered yet. To add to this
complexity, AMPH also affects the amount of DA available for release via several mechanisms by affecting DA
sequestration in secretory vesicles.
6. Effects on Secretory Vesicles. In addition to releasing DA from the cytoplasm to the extracellular
space, AMPH also promotes the release of DA from
secretory vesicles. Several mechanisms by which
AMPH may release DA from secretory vesicles into
the cytoplasm have been suggested, including AMPH
weak base effects, competition at the vesicular monoamine transporter 2 (VMAT-2), and redistribution of
VMAT-2.
AMPH is a lipophilic weak base (Gulaboski et al.,
2007; Mack and Bonisch, 1979) and, as such, is capable
of traversing the membranes of synaptic vesicles. The
lumen of secretory vesicles is acidic, with a pH of around
5.5 (Mellman et al., 1986; Njus et al., 1986), and once
inside the vesicle lumen AMPH is protonated, decreasing its membrane permeability and leading to accumulation of AMPH inside the vesicles. The protonation of

Drug-Induced Neuroplasticity

AMPH also leads to a loss of the proton gradient


established by H+-ATPase across the vesicular membrane, which is required by the secondary active transporter VMAT-2 to sequester DA into the vesicles,
eventually leading to DA release into the cytoplasm
(Sulzer and Rayport, 1990; Sulzer et al., 2005). The loss
of DA from secretory vesicles leads to a decrease in
quantal size due to the lower amount of DA in individual
vesicles. However, after prolonged (648 hours) exposure to METH a rebound hyperacidification and subsequent increase in quantal size has been reported
(Markov et al., 2008), indicative of a homeostatic mechanism capable of preventing long-term loss of DA from
secretory vesicles.
In addition to promoting release by interfering with
the membrane proton gradient, AMPH also has affinity
for VMAT-2 itself, suggesting it may also interfere with
the uptake of DA into vesicles and cause a net decrease
of vesicle neurotransmitter content (Sulzer et al., 2005).
Finally, repeated high-dose METH treatment decreases
vesicular DA uptake without affecting total binding of
the VMAT-2 ligand [3H]dihydrotetrabenazine, consistent with removal of VMAT-2 from the vesicular
membrane (Brown et al., 2000; Hogan et al., 2000).
However, due to the relatively low affinity of AMPH for
VMAT-2 (see Table 1), the importance of VMAT-2dependent mechanisms in affecting AMPH-induced
vesicular DA efflux at physiologically relevant concentrations has been questioned (Gonzalez et al., 1994;
Fleckenstein et al., 2007).
7. Other Amphetamine Targets: Monoamine Oxidase,
Tyrosine Hydroxylase, other Monoamine Transporters,
Trace Amine-Associated Receptor 1, and Sigma
Receptors. The ability of AMPH to block MAO has
been repeatedly demonstrated both in vitro and in vivo
(Mantle et al., 1976; Green and el Hait, 1978; Clarke
et al., 1979; Miller et al., 1980; Robinson, 1985). AMPH
also increases DA levels because of enhancement of TH
function (Costa et al., 1972; Kuczenski, 1975), although
high concentrations appear to decrease TH function
(Kogan et al., 1976; Hotchkiss and Gibb, 1980). The
exact mechanism by which AMPH increases TH function is not known (Sulzer et al., 2005). This highlights
that AMPH, at least to a certain extent, compensates for
its enhancement of DA efflux by increasing cellular DA
levels by inhibiting DA breakdown and increasing DA
biosynthesis.
AMPH also increases extracellular levels of 5-HT and
NE (Kuczenski and Segal, 1997) in a similar fashion by
inhibiting the reuptake and enhancing the release of
these neurotransmitters via their respective plasmalemmal transporters (Parada et al., 1988; Rothman
et al., 2001; Hilber et al., 2005). The effects of AMPH
in these other monoamine systems play an important
role in mediating part of its behavioral effects (Xu et al.,
2000; Gingrich and Hen, 2001; Gainetdinov et al., 2002;
Perona et al., 2008).

905

Several other targets for AMPH and METH are


known. Recently the TA1 (Reese et al., 2014) and sigma
receptors (Matsumoto et al., 2014) have received attention. The TA1 is a trace amine-associated GPCR and
colocalizes with DAT in certain brain regions. It exerts
a modulatory effect on monoamine neurotransmission
(Xie and Miller, 2009). Taar1-KO mice show clear
behavioral differences and respond more strongly to
the locomotor activating and rewarding effects of
AMPH (Achat-Mendes et al., 2012a). The sigma receptor, a one-transmembrane domain receptor at the
endoplasmic reticulum (ER) encoded by the SIGMAR1
gene that is involved in cellular Ca2+ and K+ regulation
(Aydar et al., 2002; Monassier and Bousquet, 2002), is
another interesting target in mediating METH neurotoxicity. METH preferentially binds to the s1 over the s2
subtype, and blocking the receptor acutely decreases
the drugs locomotor stimulating effects (Hayashi and
Su, 2007). Sigma receptor blockade also attenuates
striatal DA depletion, cytokine activation, and cognitive
impairment induced by high-dose METH treatment in
rodent models, something that may be related to its
effects on cellular Ca2+ levels or the expression of s
receptors on microglia (Robson et al., 2013a,b; Seminerio et al., 2013).
For AMPH and METH, the DA system is clearly
important in mediating its neurochemical and behavioral effects, although it is clear that some of its effects
are also mediated by different neurotransmitter systems, some of which only recently have been identified.
8. Action of Substituted Amphetamines and
Cathinones. Once the amphetamine molecule is subjected to substitutions, its properties change in important ways. The substituted amphetamines MDMA and
the cathinones mephedrone (4-MMC) and methylone
(MDMC) have a mechanism of action that is reminiscent of AMPH and METH but they differ in the ratio of
their effects at various monoamine systems (Table 1).
AMPH and METH bind to DAT, NET, and SERT, but
they show a strong preference for the catecholamine
transporters DAT and NET and have much lower
affinity for the SERT (Nichols, 1994; Simmler et al.,
2013). Subsequently, their main effect in vivo is to
increase extracellular DA levels in the NAc and FC
without having strong effects on 5-HT release (Carboni
et al., 1989; Kehr et al., 2011; Melega et al., 1995;
Shoblock et al., 2003b).
Substitutions of the phenyl ring on the amphetamine
molecule produce compounds with much higher affinity
for SERT, although they often maintain DA-releasing
properties as well (Nichols, 1994). Thus, MDMA, which
contains a 3,4-methylenedioxy substitution, is a potent
5-HT-releasing agent with relatively weak effects on
DAT (Cozzi et al., 1999; Verrico et al., 2007; Simmler
et al., 2013). Furthermore, like AMPH and METH,
MDMA also induces neurotransmitter efflux from synaptic vesicles (Rudnick and Wall, 1992; Mlinar and

906

Korpi et al.

Corradetti, 2003). In vivo, it produces increases in NAc


levels of 5-HT and DA, with 5-HT levels generally
increasing more than DA (Kehr et al., 2011; Baumann
et al., 2012).
The cathinones 4-MMC and MDMC, also called
b-ketoamphetamines because of the ketone moiety on
the b-carbon, have similar affinity for plasmalemmal
monoamine transporters as their nonketoamphetamine
parent compounds but display a more than 10-fold
lower affinity for the VMAT-2 (Cozzi et al., 1999;
Baumann et al., 2012; Martinez-Clemente et al.,
2012). The exact importance of the lower VMAT-2
affinity is currently unknown, but because 4-MMC is
also a weak base (Santali et al., 2011), it is likely that it
is also capable of inducing neurotransmitter efflux from
secretory vesicles in a similar way as other amphetamines. In vivo, 4-MMC produces an effect somewhere
in between MDMA and AMPH, producing large
increases in both DA and 5-HT levels in the NAc,
whereas MDMC produces a neurotransmitter release
pattern similar to but slightly weaker than MDMA,
increasing primarily 5-HT levels (Kehr et al., 2011;
Baumann et al., 2012; Wright et al., 2012). So, although
the mechanism of action of these ring-substituted
compounds is similar to that of AMPH and METH, they
all share the property of being capable of producing
large increases in extracellular 5-HT levels in addition
to affecting extracellular catecholamine levels to varying extents.
9. Action of Stimulants Common in Clinical Use.
In addition to AMPH and METH, the drug MPH as well
as the AMPH prodrug lisdexamfetamine (LDX) is used
for the treatment of ADHD.
MPH is, because of the loss of its a-methyl moiety, by
chemical definition not an amphetamine. Nonetheless it
is a stimulant drug and has been the mainstay of
ADHD-treatment for several decades (Patrick and
Markowitz, 1997). The action of MPH shows some
similarity to AMPH and also produces an increase of
striatal DA levels in vivo, measured using microdialysis
(Aoyama et al., 1996). Studies in membrane fractions
and synaptosomes show that MPH, like AMPH, inhibits
the reuptake of the catecholamines by DAT and NET
but has little affinity for SERT (Schweri et al., 1985;
Gatley et al., 1996). In contrast to AMPH, however,
MPH appears to function primarily as a reuptake
blocker and not as a releasing agent (Braestrup, 1977;
Patrick et al., 1987).
LDX is a prodrug consisting of AMPH covalently
bound to the amino acid lysine. As a prodrug it is
inactive until it is enzymatically cleaved into lysine and
free AMPH. LDX does not appear to be very lipophilic,
and enzymatic cleavage by hydrolysis occurs only after
peptide carrier-mediated absorption in the small intestine (Pennick, 2010). The prodrug formulation has
been suggested to be a clinical improvement over
regular AMPH because it is longer lasting and thus

allows for once daily dosing as well as possibly limiting


its potential for nonmedical use (Goodman, 2007). A
study investigating the effect of LDX and AMPH (both
1.5 mg/kg AMPH base) indeed shows that the onset of
increases in striatal DA levels is slower with LDX,
corresponding to the plasma concentration of the
AMPH metabolite rather than the prodrug. Moreover,
the plasma area under the curve for AMPH is similar for
both drugs but with LDX showing a 50% lower maximum concentration and significantly longer delay from
administration until the maximum concentration is
reached. Finally, LDX induces significantly less locomotor activity (Rowley et al., 2012).
10. Efficacy and Addiction Liability of Amphetamines in Clinical Use. The ability of amphetamines
to increase synaptic DA levels via various mechanisms
is thought to play a pivotal role in both producing their
therapeutic as well as their rewarding effects.
In ADHD patients, there is a plethora of evidence
pointing toward deficits in DA neurotransmission as
playing a key role in the pathophysiology of the disease.
As reviewed earlier (Swanson et al., 2007), several
imaging studies suggest structural abnormalities in
ADHD patients, such as decreased volumes of the DArich caudate nucleus and its target globus pallidus. This is
consistent with fMRI studies, which have reported hypoactivation of the FC as well as the frontostriatal DA
circuitry while performing tasks requiring sustained
attention. These abnormalities are resolved after treatment with psychostimulants. Functional abnormalities as
well as the symptoms observed in ADHD patients appear
to normalize after stimulant drug treatment, which
increases synaptic DA levels (Lee et al., 2005). However,
the long-term effects of stimulant treatments are still
poorly understood. Although studies generally find sustained symptom improvement, it has been harder to
demonstrate consistent improvement in standardized test
scores and ultimate educational attainment (Hechtman
et al., 2004; Loe and Feldman, 2007; Hazell, 2011).
Aside from efficacy, another interesting question is
how stimulant treatment of ADHD patients affects
addiction liability later in life. A number of studies,
including one meta-analysis, have examined this question and concluded that stimulant treatment generally
appears to lower the risk of substance abuse disorder
later in life (Wilens et al., 2003; Mannuzza et al., 2008).
This effect may be understood by considering that
impulsive behavior is one of the primary risk factors
for initiation of drug use (Kreek et al., 2005) and that
stimulant treatment effectively reduces impulsivity in
ADHD patients. On the other hand, misuse and diversion of stimulant medication has become increasingly common, and it is estimated that in the past year,
5 to 35% of college-aged individuals have used a stimulant drug not prescribed to them (Wilens et al., 2008).
MPH is another drug that elicits subjective effects
of drug-liking similar to AMPH, and the subjective

Drug-Induced Neuroplasticity

rewarding effect appears to be dependent upon


increases in synaptic DA levels produced by the drug
(Rush et al., 2001; Volkow et al., 2002; Arria and Wish,
2006; Botly et al., 2008). LDX also produces similar
feelings of subjective drug-liking on questionnaires in
human volunteers as the 50% lower doses of AMPH
base (Jasinski and Krishnan, 2009). The higher dose
need likely represents the amount needed to achieve
similar peak plasma AMPH concentrations and
increases in the NAc DA levels, which at an equimolar
ratio emerge slower and remain lower due to its
extended absorption time of LDX.
Ring substitution of amphetamines produces substances with a high capability of increasing both 5-HT
and catecholamine levels (Nichols, 1994). Some of these
compounds, most notably MDMA, appear to produce
effects that are somewhere between those of pure
stimulants and typical hallucinogens such as LSD or
psilocybin. The most important effect appears to be that
they, next to creating a feeling of euphoria and wellbeing, produce emotional opening and empathy. This
observation has led to the term empathogen or entactogen being coined to describe these substances, as well
as the suggestion that they may be valuable in the
clinics as an aid to psychotherapy (Nichols, 1986). In
fact, recent clinical trials support the idea that MDMA
may be effective when used as an aid in psychotherapy
for treating treatment-resistant posttraumatic stress
disorder patients (Mithoefer et al., 2011, 2013). The use
of MDMA and related more serotonergic amphetamines
is not usually associated with significant addiction
propensity. This is thought to be due to 5-HT receptor stimulation having antiaddictive properties
something that will also be touched on in section II.I on
hallucinogens.
11. Neuroplasticity Related to Sensitization and
Addiction to Amphetamines. The incentive sensitization theory of drug addiction states that repeated drug
administration can lead to long-lasting adaptations
resulting in hypersensitivity or sensitization in parts
the mesolimbic DA system involved in processing a subcomponent of reward termed incentive salience or drug
wanting, which is thought to mediate the increased
drug seeking and taking characteristic of drug addiction. Repeated treatment with reinforcing drugs produces a gradual increase in locomotor activity after
administration, an effect known as psychomotor sensitization. Because it is assumed that the neural
adaptations responsible for developing psychomotor
sensitization are similar to those involved in producing
incentive sensitization, the former is often used as
a model for the development of addiction (Robinson
and Berridge, 2000).
Treatments with psychostimulants such as AMPH
and METH produce psychomotor sensitization associated with neuroadaptations in the VTA and NAc, but
also in the PFC, amygdala, and HC (Anagnostaras and

907

Robinson, 1996; Nestler, 2001; McDaid et al., 2006;


Perez et al., 2010; van der Veen et al., 2013). A number
of different types of structural and synaptic plasticity
have been described that may be responsible for this
effect. For example, treatment with psychostimulants,
including AMPH, result in upregulation of TH levels in
the VTA and NAc. Exposure to psychostimulants also
causes synaptic changes, such as downregulation of
inhibitory D2Rs but upregulation of excitatory D1Rs in
the VTA. It has also been suggested that NMDARs may
become more permeable to Ca2+, due to secondmessenger-mediated signaling in response to repeated
psychostimulant exposure. The ultimate result of all
these adaptations is enhanced DA neurotransmission in
the mesolimbic system, which is critical for the development of sensitization to psychostimulants (Licata and
Pierce, 2003; Kasanetz et al., 2010; Fernandez-Espejo
and Rodriguez-Espinosa, 2011). VTA NMDARs appear
to play a vital role in psychostimulant-induced neuroplasticity through activation of CaMKII. Increased Ca2+
influx after psychostimulants such as AMPH and cocaine elevate CaMKII levels, and inhibition of CaMKII
blocks both behavioral sensitization and drug-seeking
behavior (Licata et al., 2004; Loweth et al., 2008). The
importance of CaMKII signaling is also evident because
of its vital role in memory formation and because it has
been shown to be directly involved in sensitization via
activation of CREB and regulation of TH transcription
(Lim et al., 2000; Lisman et al., 2002).
Also MDMA affects sensitization and memory. A dose
of 12.5 mg MDMA administered once a week for 3 weeks
increased LTP in CA1 pyramidal neurons (hippocampal
slices) 1 week after in vivo treatment (Morini et al.,
2011), whereas an in vitro study showed that acute
MDMA increases LTP in the CA3-CA1 synapses involving the activation of presynaptic 5-HT2 and postsynaptic D1/5 receptors and was suppressed by a PKA
inhibitor (Rozas et al., 2012). Acute administration of
MDMA increased expression of c-Fos, erg-1 and erg-3 in
the striatum that was suppressed by SL327, a selective
inhibitor of ERK activation, and this effect was selective
for the striatum (Salzmann et al., 2003). Repeated
administration of MDMA, gamma-hydroxybutyrate
(GHB), and their combination results in long-lasting
changes in hippocampal protein expression even after
a washout period (van Nieuwenhuijzen et al., 2010),
whereas MDMA treatment (0.2 and 2 mg/kg) twice
a day for 6 days in young rats caused dose-dependent
impairment in spatial learning and reduced LTP (AriasCavieres et al., 2010). Furthermore, just like AMPH,
MDMA also produces sensitization after repeated exposure (Bradbury et al., 2012).
Repeated stimulant exposure that produces sensitization also results in morphologic changes to neurons in
the NAc and PFC. Both SA and noncontingent injections of cocaine (15 mg/kg) or AMPH (3 mg/kg/day)
produce an increase in the density of and the amount of

908

Korpi et al.

branching of the dendritic spines on NAc MSNs and


PFC pyramidal neurons for more than 3 weeks after the
final treatment (Robinson and Kolb, 1999a; Robinson
et al., 2001). Meanwhile, METH-induced tissue plasminogen activator mRNA in a subgroup of corticostriatal neurons was suggested to contribute to the
development of lasting changes in the specific neuronal
circuits that underlie long-term behavioral modifications (Hashimoto et al., 1998). Repeated, intermittent
exposure to MDMA in rats causes behavioral sensitization and large increases in spine density and number in
the NAc (multiple-headed spines in core and shell) and
prelimbic regions (distal dendrites of layer V pyramidal
neurons) (Ball et al., 2009). Additionally, a reduction in
the basilar dendrites was observed in the anterior
cingulate. These changes accounted for the longlasting locomotor sensitization to MDMA that is accompanied by reorganization of synaptic connectivity in the
limbic-corticostriatal circuit (Ball et al., 2009). The longlasting behavioral (increased locomotion and rearing)
and EEG changes (peak at 34.5 Hz) due to MDMA
treatment were associated with DNA single- and
double-strand breaks, persistent together with longlasting metabolic changes in the hippocampal formation
(Frenzilli et al., 2007).
Protein synthesis plays an important a role in the
induction of these morphologic changes, because proteins such as c-Fos and DFosB, but also Arc, a marker of
neuritic outgrowth, and BDNF, which is involved in
spine formation, are all elevated after AMPH or cocaine
exposure (Nestler et al., 2001; Klebaur et al., 2002;
Meredith et al., 2002; Grimm et al., 2003; Fumagalli
et al., 2006). Particularly the rmTg appears to be an
important target for AMPH-induced modulation of Fos
genes that can also have strong functional and behavioral significance in terms of modulating the mesolimbic
systems through its GABAergic projections (Barrot
et al., 2012; Bourdy and Barrot, 2012) (Fig. 1). MDMA
treatment of young rats increased 2,5-dimethoxy-4iodoamphetamine (DOI)-induced energy metabolism
in the NAc and HC that reflects long-term brain
region-specific changes in plasticity, and such an effect
may result from changes 5-HT2A receptor function (Bull
et al., 2006). Likewise, a dose (12.5 mg/kg) of MDMA
increases basal Arc expression in the cortex and CA1
(Beveridge et al., 2004). DOI treatment (or immobilization stress) regulates neocortical Arc mRNA expression
through BDNF signaling (Benekareddy et al., 2013).
Chronic MDMA treatment increased NT-4 gene expression in the brain stem, cerebellum, and cerebral hemispheres of rats (Hatami et al., 2010).
Aside from changes in a number of well-characterized
proteins, AMPH treatment produces wide-ranging
changes in expression of other genes as well. A recent
study showed increased striatal mRNA expression of 55
genes, whereas 17 genes were downregulated after
a 5-day, twice daily treatment with 5 mg/kg AMPH in

spontaneously hypertensive rats, a model for ADHD.


The genes showing altered expression belonged to
several functional categories, including those involved
in transcription, angiogenesis, cell adhesion, apoptosis,
and neuronal development (Dela Pena et al., 2015).
Furthermore, changes in epigenetic markers such as
methylation status have been reported in several brain
regions, including the NAc, after 2 weeks of amphetamine exposure (Mychasiuk et al., 2013). It is likely that
these changes also have an effect on synaptic and/or
structural plasticity in ways that have yet to be fully
characterized.
12. Glutamate and Amphetamine/Methamphetamine
Reinforcement/Extinction/Reinstatement. Like cocaine,
AMPH and METH produce CPP (Thorn et al., 2012;
Martin et al., 2013; Han et al., 2014) and are reliably
self-administered (Pickens and Harris, 1968; Balster
and Schuster, 1973; Piazza et al., 1989; Kitamura et al.,
2006). Involvement of glutamate-mediated processes
in AMPH reinforcement and SA has been welldocumented. For instance, mGlu5 receptor antagonists
produce a dose-dependent decrease in METH SA without affecting food reinforcement, suggesting that mGlu5
receptors are involved in the reinforcing effects of
METH, but not natural rewards (Osborne and Olive,
2008). Similar effects have been reported with the
mGlu2/3 antagonist LY379268 (1R,4R,5S,6R)-4-amino2-oxabicyclo[3.1.0]hexane-4,6-dicarboxylic acid)
(Crawford et al., 2013; Kufahl et al., 2013). The mGlu5
receptor is also involved in extinction learning, because
administration of an mGlu5-positive allosteric modulator resulted in more rapid extinction of responding on
the active SA lever. However, this effect was not present
during reinstatement tests or a second extinction
session (Kufahl et al., 2012). Modulation of addictive
behavior was also observed in mGlu5-KO mice, which
showed decreased extinction of drug-seeking behavior
and increased cue-induced reinstatement (Chesworth
et al., 2013). Negative allosteric modulation of mGlu5
receptors was similarly found to attenuate both drugand cue-induced reinstatement of METH-seeking behavior, although it produced a similar effect on sucrose
and food seeking (Watterson et al., 2013).
Ionotropic glutamate receptors are also involved in
AMPH-induced addictive behavior (discussed in detail
in section II.A on cocaine; Fig. 6). NMDAR subunit
GluN2A-KO mice fail to display behavioral sensitization and activation of Ca2+-dependent intracellular
signaling pathways in the NAc in response to repeated
METH treatments (Miyazaki et al., 2013). In animals
trained to self-administer METH, striatal AMPAR
GluA2 subunit mRNA was increased 2 hours after
cessation of drug intake but decreased after 24 hours
compared with control animal (Currington et al., 2011).
METH-induced CPP was associated with increased
surface expression of GluA1 and GluA2 subunits in
the BLA (Yu et al., 2013). In the striatum, however,

Drug-Induced Neuroplasticity

a decrease in GluA1 and GluA2 mRNA and protein


levels as well as in those for GluN1 subunits and
increased AMPAR/NMDAR ratio has been reported in
response to repeated METH exposure (Jayanthi et al.,
2014). Furthermore, this treatment decreased enrichment of acetylated histone H4 on GluA1, GluA2, and
GluN1 promotors as well as a number of other epigenetic markers, whereas the expression of these AMPAR
and NMDAR subunits was blocked by the histone
deacetylase inhibitor, valproic acid, suggesting an
epigenetic mechanism, despite valproates limited specificity (Jayanthi et al., 2014). Changes in AMPAR
surface expression may also be mediated by striatalenriched tyrosine phosphatase isoform 61, which internalizes AMPARs, as one study demonstrated that the
decreases in striatal-enriched tyrosine phosphatase
isoform 61 increases in GluA2 surface expression and
behavioral sensitization after repeated METH treatments were blocked by an mGlu5 negative allosteric
modulator (Herrold et al., 2013). Treatment with ceftriaxone, which increased glutamate transporter
mRNA in the mPFC, blocked METH-induced reinstatement of CPP (Abulseoud et al., 2012).
An interesting question is whether observed plasticity phenomena are drug dependent or if they are
primarily associated with development of SA or sensitization. METH-induced glutamatergic neuroplasticity
appears to be at least partially dependent upon concurrent development of addictive behaviors, rather than
the drug itself. One study investigated the effect of
noncontingent METH administration versus METH SA
and found important differences in the glutamatergic
response to a subsequent METH challenge after drug
withdrawal. In noncontingently dosed rats the challenge decreased extracellular glutamate levels in the
NAc, whereas in previously self-administering rats, an
increase in glutamate levels was observed. This increase was also associated with a larger increase in DA
levels (Lominac et al., 2012). Another study found
reduced basal glutamate levels in both the NAc and
PFC, but increased glutamate levels after reinstatement in self-administering rats compared with yoked
saline controls (Parsegian and See, 2014).
How does plasticity differ between METH and cocaine
exposures? One study demonstrating the similarities
between glutamate involvement in METH and cocaine
SA directly compared the effect of a selective mGlu1
receptor antagonist on both METH- and cocaineinduced behavior in squirrel monkeys. It reported that
the antagonist produced a reduction in SA of both drugs
and attenuated the increase in responding for food
rewards produced by both drugs (Achat-Mendes et al.,
2012b). DA receptors play an important role in mediating synaptic plasticity (Wolf et al., 2003; Luscher and
Malenka, 2011), and because both METH and cocaine
produce increases in extracellular DA levels in vivo, it is
not surprising that they produce similar neuroplastic

909

effects that may be related to the development of drug


dependence. On the other hand, differences also exist,
and it has been reported that METH, unlike cocaine, did
not alter NAc or PFC cell surface expression of AMPARs
or NMDARs during withdrawal (Boudreau and Wolf,
2005; Nelson et al., 2009), indicative of important
differences between the two drugs. Cotreatment with
a spin-trapping agent prevented AMPH-induced
increases in glutamate efflux, suggestive of a ROSdependent signaling pathway (Wolf et al., 2000). ROS
signaling may be more important during amphetamine
exposure compared with cocaine because of the increased intracellular ROS after release of DA from
vesicles and cytoplasmic oxidation. Clearly, further
work is needed to outline the similarities and differences between AMPH/METH and cocaine neuroadaptation, as well as its behavioral implications.
13. Long-term Neuroadaptation/Neurotoxicity after
Exposure to High Doses of Psychostimulants. In research on psychostimulant neuroplasticity there has
been a focus on the long-term after-effects of psychostimulants such as METH and MDMA. These
amphetamine-type psychostimulants are sometimes
referred to as club drugs, because they are commonly
consumed recreationally at music festivals and dance
clubs. This pattern of consumption is not generally
associated with the development of addiction, but users
will sometimes redose the drug repeatedly during one
sitting (binge), which has led to concern about the
possible long-term effects of this pattern of consumption
on the brain in terms of long-term neuroadaptation and
cognitive function, such as deficits in memory and
attention.
a. Effects on brain structure. Ricaurte et al. (1982)
demonstrated that high doses of METH (3 injections of
50 mg/kg spaced over 8 hours) to rats caused destruction
of nerve terminals in the striatum as observed 4 days
after the treatments with a silver staining method,
considered the gold standard of neurotoxicity testing
(Switzer, 2000). Furthermore, they reported that similar
regiments also reduced levels of DA and 5-HT both in the
striatum and other brain regions such as the FC and
amygdala (Ricaurte et al., 1980, 1982). The observation
that high doses of METH produced neurotoxicity, visible
as swollen fibers and nerve terminals indicative of
chromatolysis and axonal damage, has also been found
by other investigators (Escalante and Ellinwood, 1970;
Lorez, 1981). Recently, evidence of toxicity, measured by
silver staining, was observed for METH at even lower
doses (3  5 mg/kg) (Ares-Santos et al., 2014). Similar
results were reported by Molliver et al. (1990) for MDMA
(20 mg/kg twice daily for 4 days), who found evidence of
structurally damaged axons that appeared as swollen
stumps in cortical regions after visualization with silver
staining. The damage coincided with loss of fine 5-HT
axons in cortical regions 2 weeks after treatment visualized by 5-HT immunocytochemistry.

910

Korpi et al.

b. Effects on neurochemistry. Quantitative ligand


binding and neurochemical methods suggest alterations in monoamine systems after METH. Numerous
studies have reported decreases in striatal DA and 5-HT
levels after treatment with METH, with at least partial
recovery over time (Wagner et al., 1980; Seiden et al.,
1988; ODell et al., 1991; Cass, 1996; Callahan et al.,
1998; Cass and Manning, 1999). DAT and SERT levels,
as measured by ligand binding or uptake measurements, are decreased after METH (Hirata and Cadet,
1997; Brown et al., 2000; Ladenheim et al., 2000;
Schroder et al., 2003; Armstrong and Noguchi, 2004).
Finally, decreases in the levels of and function of the
synthesizing enzymes TH and tryptophan hydroxylase
(TPH) have been reported repeatedly as well, also with
recovery occurring as the survival interval increases
(Hotchkiss and Gibb, 1980; Bakhit et al., 1981; Haughey
et al., 1999; Schroder et al., 2003).
In rodents, numerous studies demonstrate that
METH can affect 5-HT and DA axons. For instance,
METH produces axons that appear swollen and varicose and display a decrease in 5-HT immunoreactive
boutons (Fukui et al., 1989). Similarly, both METH and
AMPH were reported to produce reduction in THpositive neurons in the striatum and FC, with AMPH
also producing swollen and enlarged axons (Kadota and
Kadota, 2004; Bowyer and Schmued, 2006). MDMA
appears to have an effect on 5-HT, but not on DA axons.
For instance, decreased 5-HT-positive neurons were
observed in the parietal cortex and forebrain regions
after treatment with high doses of MDMA, and similar
to METH, the remaining axons were ablated or thick
and varicose. However, no damage was observed to the
cell bodies in the raphe nucleus or to catecholaminergic
neurons and, furthermore, the loss of 5-HT axons
recovered to a large extent after 1 year (Ohearn et al.,
1988; Molliver et al., 1990; Scanzello et al., 1993).
The available evidence for MDMA suggests that high
doses affect only the 5-HT system in rats, having no longlasting effects on the DA system. In rats, decreased levels
of 5-HT and SERT binding in several brain regions
including the striatum, HC, and hypothalamus have been
reported, as well as decreased TPH activity. The alterations of these markers commonly recovers with time,
and no studies report any alterations in the catecholaminergic systems (Armstrong and Noguchi, 2004; Stone
et al., 1986; Battaglia et al., 1987, 1988; De Souza et al.,
1990; Sprague et al., 1994). In mice, conversely, even very
high doses of MDMA (60 mg/kg) do not affect serotonergic
markers such as brain 5-HT levels or TPH activity (Stone
et al., 1987) but instead decrease the DA levels, particularly in the striatum (Logan et al., 1988; OCallaghan and
Miller, 1994). Thus, there appears to be an important
species difference, with MDMA affecting the 5-HT system
in rats but the DA system in mice.
Studies in nonhuman primates generally support the
neurotoxicity or plasticity profile also observed in rats,

with METH decreasing 5-HT, DA, DAT, TH, and


VMAT-2 binding in various brain regions such as the
striatum and cortex (Woolverton et al., 1989; McCann
et al., 1998b) and MDMA affecting only the 5-HT system
as observed by decreased levels of 5-HT and SERT
binding as well as loss of 5-HT-positive axons in several
brain regions including cortical regions and the HC,
again with at least partial recovery occurring as the
survival interval increases (Insel et al., 1989; De Souza
et al., 1990; Hatzidimitriou et al., 1999).
c. Effects on behavior. In addition to the neurochemical after-effects of high doses of METH and AMPH,
research has also focused on the potential consequences
on behavior, primarily by means of animal models of
cognition, anxiety, and depression and including memory tests such as the novel object recognition test and
the Morris water maze, tests of anxiety-like behavior in
emergence tests, elevated plus maze tests, and social
interaction tests as well as the forced swim test, said to
measure aspects of depression (McGregor et al., 2003b;
Clemens et al., 2004, 2007 Clemens et al., 2004). Overall, however, the results supporting long-term effects of
amphetamines on these behavioral tests are much less
consistent than the data from neurochemical studies,
often showing contradictory effects.
METH was shown to decrease recognition memory in
the novel object recognition test for up to several weeks
after drug treatments (Schroder et al., 2003; Marshall
et al., 2007; Herring et al., 2008). Also decreases in
spatial memory have been observed using tests such as
the radial arm maze and Morris water maze (Nagai
et al., 2007; Camarasa et al., 2010). However, other
studies report no effect on memory using the novel
object recognition test (Clemens et al., 2007), whereas
inconsistent results are obtained for anxiety-related
behavior depending on the type of test used, and no
effect was found on the forced swim test (Clemens et al.,
2004, 2007).
With MDMA, the evidence of persistent long-lasting
effects on rodent behavioral tests is also difficult to
interpret. Decreased recognition memory performance
(Mechan et al., 2002; Piper and Meyer, 2004) and
decreased performance in a delayed nonmatching to
place test (Marston et al., 1999) have been reported.
Evidence of depressive behavior measured as increased
immobility in the forced swim test and evidence of
anxiety-related behavior measured using several tests,
such as the social interaction and emergence test and
the elevated plus maze, have also been described
(McGregor et al., 2003a,b). On the other hand, decreases
in anxiety-like behavior after MDMA have been observed using these same tests (Mechan et al., 2002;
Piper and Meyer, 2004). Also memory deficits after
MDMA are not always observed. For instance, MDMA
has been reported not to affect object recognition
memory (McGregor et al., 2003a). Other studies also
found no evidence of impairment in animals tested on

Drug-Induced Neuroplasticity

a wide range of behavioral tests, including the radial


arm maze and the Morris water maze, despite the fact
that large decreases in cortical 5-HT levels were present
(Seiden et al., 1993; Marston et al., 1999). These results
indicate that the relationship between observed
monoamine system deficits and functional outcome
in terms of behavior and cognitive function is not
straightforward.
A number of limitations must be kept in mind when
interpreting the results from preclinical rodent studies.
Foremost, the use of allometric scaling to adjust for
differences in basal metabolic rate between humans and
rodents has led to rodents commonly being administered
relatively high drug doses, often in the 1020 mg/kg
range or higher. However, the use of allometric scaling
has been criticized for drugs such as MDMA, which
display nonlinear kinetics and are extensively metabolized (Baumann et al., 2007). When employing effects
scalingbased on selecting doses required to achieve
similar amounts of 5-HT release, prolactin secretion,
reinforcement, or drug discriminationthe required
doses are in the 13 mg/kg range, closely corresponding
to the doses used recreationally by humans. More importantly, when administered at these doses (,5 mg/kg)
there is no evidence of long-lasting monoamine depletions (OShea et al., 1998; Baumann et al., 2007).
Furthermore, high doses of the drugs are often given
to animals without any preconditioning to lower doses,
something that would more correctly model human use
and has shown to be protective in animal models (see
below). Finally, a majority of studies report on the effect
on measures such as levels of monoamines and their
transporters and synthesizing enzymes. One can contend whether this reflects actual toxicity or rather longterm adaptation, particularly as these changes are
not always associated with deficits in behavioral tests
and because higher doses are commonly needed to
produce toxicity as measured by methods such as Fluoro
Jade B staining and gliosis marker levels (Baumann
et al., 2007)
14. Limited Evidence for Long-term Effects of
Substituted Cathinones in Rodent Models. Recent
studies on the effects of the substituted cathinones 4MMC and MDMC show much more limited long-term
effects of these drugs on the monoamine systems
compared with METH and MDMA. When 4-MMC is
administered at three injections of up to 10 mg/kg there
was no long-lasting decrease in striatal or cortical
monoamine levels in rats (Baumann et al., 2012).
Interestingly, also intense, high-dose binge treatments
of four injections of up to 40 mg/kg spaced 2 hours apart
failed to produce any decreases in striatal DA levels
(Angoa-Perez et al., 2012). The lack of effect of 4-MMC
on the monoamine system is corroborated by two other
studies employing various dosing regimens such as
30 mg/kg once daily for 10 days (Motbey et al., 2013)
or twice daily for 4 days (den Hollander et al., 2013),

911

which both failed to observe any decreases in monoamine levels in the striatum, cortex, and HC at
a survival interval between 14 and 43 days after the
final 4-MMC treatment. The only studies so far that did
observe an effect of 4-MMC on markers of the monoamine system, such as DAT, SERT, TPH, and TH levels
and functions, were the studies in which the drugs were
administered at elevated (.26C) ambient temperatures (Hadlock et al., 2011; Martinez-Clemente et al.,
2014). High ambient temperatures produce a general
increase in amphetamine neurotoxicity (see below), and
these studies indicate that this also holds true for
cathinones. Furthermore, there is evidence that 4MMC can increase the loss of striatal DA induced by
METH when the two drugs are coadministered (AngoaPerez et al., 2013b), but does not exacerbate the loss of 5HT after METH or MDMA (Angoa-Perez et al., 2013a).
MDMC has so far been less studied, but the available
evidence suggests that it does not have an effect on
monoamine levels at doses of up to three times 10 mg/kg
(Baumann et al., 2012), whereas high doses (30 mg/kg
twice daily for 4 days) produce decreases in 5-HT levels
in the cortex, striatum, and HC at a 2-week survival
interval (den Hollander et al., 2013).
Regarding the effects of these drugs on behavioral
tests of cognition, the data are less clear. Although
a decrease in recognition performance in the novel
object recognition test was observed 1 month after
treatment with 4-MMC (Motbey et al., 2012), as well
as a decrease in working memory measured using the
T-maze spontaneous alternation test, there were no
effects on spatial memory in the Morris water maze test.
Furthermore, both 4-MMC and MDMC have no effect on
anxiety-like behavior measured using the emergence
test or elevated plus maze or on depressive behavior in
the forced swim test and actually appear to improve
aspects of spatial reversal learning (den Hollander
et al., 2013).
15. Mechanisms Involved in Long-term Adaptation
and Neurotoxicity. The four primary mechanisms responsible for producing neurotoxicity after exposure to
amphetamines are oxidative stress, mitochondrial dysfunction, excitotoxicity, and hyperthermia (Fig. 8).
Additionally, a number of other factors including inflammation, microglial activation, blood-brain barrier
disruption, and altered neurogenesis play a role,
whereas neuroplasticity associated with low-dose
AMPH preconditioning may provide protective effects.
A summary of the primary mechanisms is provided
here, whereas more in-depth information can be found
in relevant reviews (Yamamoto et al., 2010; Carvalho
et al., 2012; Goncalves et al., 2014; Halpin et al., 2014)
a. Oxidative stress: DA oxidation and drug metabolites as sources of reactive species. Support for the
involvement of oxidative stress as a causal mechanism
in amphetamine toxicity comes from a number of
studies demonstrating attenuation of METH- and

912

Korpi et al.

MDMA-induced monoamine loss after coadministration of spin-trapping agents or antioxidants such as


N-acetylcysteine, ascorbic acid, vitamin E, and selenium
(Wagner et al., 1985; De Vito and Wagner, 1989; Colado
and Green, 1995; Fukami et al., 2004; Barayuga et al.,
2013). Accordingly, depletion of glutathinone, inhibition
of superoxide dismutase, and selenium-deficient diets
increase toxicity (De Vito and Wagner, 1989; Sanchez
et al., 2003; Chandramani Shivalingappa et al., 2012).
One of the primary sources of oxidative stress after
amphetamines is the large increase in cytoplasmic and
extracellular DA levels and the subsequent metabolism
of DA into toxic metabolites (Yamamoto and Raudensky,
2008). DA can be both enzymatically and nonenzymatically metabolized, with both routes resulting in
production of reactive products. When MAO metabolizes large amounts of DA, the resulting byproduct
hydrogen peroxide may rise to levels exceeding the
detoxification capacity of the glutathione system. Subsequently, the hydrogen peroxide can react with Fe2+,
resulting in the production of the highly reactive
hydroxyl radical. Moreover, both DA and its metabolite
3,4-dihydroxyphenylacetaldehyde can undergo autooxidation to form superoxide and reactive semiquinones
(LaVoie and Hastings, 1999; Sulzer and Zecca, 2000;
Anderson et al., 2011; Munoz et al., 2012). Accordingly,
reducing intracellular DA through blockade of DAT or
inhibition of DA synthesis reduces the METH toxicity

(Schmidt et al., 1985; Marek et al., 1990; Yamamoto and


Zhu, 1998).
Other potential sources of reactive species are amphetamine molecules themselves. Particularly MDMA
has received attention in this respect, because certain
catechol-like hepatic metabolites of MDMA are thought
to be capable of auto-oxidizing and producing reactive
quinones capable of forming thioether conjugates with
sulfur-containing compounds such as glutathione and
N-acetylcysteine. These thioether compounds retain the
ability to redox cycle and are toxic to 5-HT neurons. In
fact, they are more toxic than MDMA itself (Monks
et al., 2004; Ferreira et al., 2013) and have also been
detected in vivo after MDMA exposure (de la Torre and
Farre, 2004; Perfetti et al., 2009). Nonetheless, DA
remains an important factor also in MDMA-induced
5-HT toxicity (Stone et al., 1988; Schmidt et al., 1990;
Shankaran et al., 1999; Falk et al., 2002; Breier et al.,
2006).
b. Mitochondrial dysfunction: ATP deficit, superoxide
leakage, and apoptosis. METH and MDMA are known
to decrease the levels and activity of mitochondrial
complexes I, II, and IV in the striatum (Burrows et al.,
2000; Brown et al., 2005; Klongpanichapak et al., 2006;
Puerta et al., 2010). MDMA was also shown to cause
deletions in mitochondrial DNA coding for subunits of
complexes I and IV (Alves et al., 2007). More recently
it was shown that even relatively low doses of both

Fig. 8. Summary of various mechanisms and factors involved in cellular toxicity by amphetamine-type stimulants.

Drug-Induced Neuroplasticity

AMPH and METH (2 mg/kg or less) decrease complex I,


II, or IV activity in various brain regions, including the
striatum, FC, and amygdala, obtained from animals
killed while still under the influence of the drugs (Feier
et al., 2012). Inhibition of the mitochondrial electron
transport chain (ETC) by amphetamines can result in
reduced ATP levels but also causes excessive leakage of
superoxide, increasing oxidative stress and possibly
triggering apoptosis.
METH decreases ATP levels in the striatum and
cerebral cortex while increasing AMP levels (Shiba
et al., 2011). The importance of decreased ATP levels
in amphetamine toxicity is demonstrated by an experiment where neural glucose uptake was prevented by
administration of 2-deoxyglucose, a manipulation that
enhanced both the decrease in ATP levels observed
directly after METH treatment as well as the loss of
striatal DA 1 week later (Chan et al., 1994). Similarly,
inhibition of metabolism by directly interfering with the
ETC also increases toxicity because the potent complex
II inhibitor malonate exacerbated both METH and
MDMA toxicity (Burrows et al., 2000; Darvesh et al.,
2005), whereas application of the energy substrates
decylubiquinone or nicotinamide 6 hours after METH
treatment reduced toxicity (Stephans et al., 1998).
ETC inhibition can also exacerbate amphetamine
toxicity by increasing oxidative stress. During normal
oxidative phosphorylation, approximately 14% of mitochondrial oxygen will be reduced to superoxide.
However, during inhibition of mitochondrial complexes,
the electron leak and superoxide production will increase dramatically. The resulting increase in oxidative
stress adds to the stress already occurring from DA
oxidation and represents a key factor in mediating
amphetamine toxicity (Brown and Yamamoto, 2003;
Adam-Vizi, 2005).
The mitochondrial-dependent apoptotic pathway has
also been suggested to play a role in amphetamine
toxicity (Cadet et al., 2005). METH increases the
expression of proapoptotic genes BAD, BAX, and BID,
while decreasing the expression of antiapoptotic genes
Bcl-2 and Bcl-XL (Jayanthi et al., 2001; Krasnova et al.,
2005; Beauvais et al., 2011). Indeed, METH has been
shown to trigger the release of apoptosis inducing
factor, cytochrome c, and smac/DIABLO from mitochondria, eventually leading to the recruitment of caspases
and the initiation of apoptotic cell death (Jayanthi et al.,
2004). As expected, both overexpression of Bcl-2 and
inhibition of caspases protect against METH toxicity
(Cadet et al., 1997, 2003; Uemura et al., 2003). However,
the apoptotic effect has primarily been observed after
extremely high doses of METH or in vitro, and its
relevance for human users remains unclear.
c. Excitotoxicity: glutamate receptor involvement and
intracellular excitotoxic processes. Studies employing
in vivo microdialysis have shown that METH increases
glutamate release in the striatum (Mark et al., 2004;

913

Nash and Yamamoto, 1992; Stephans and Yamamoto,


1994) and that blockade of glutamate receptors attenuates METH-induced depletion of striatal DA and other
markers of toxicity (Farfel et al., 1992; Bowyer, 1995;
Battaglia et al., 2002; Shah et al., 2012), indicating
that glutamate receptor-mediated processes are involved in METH toxicity. Conversely, MDMA appears
to reduce glutamate release in certain brain regions,
and the importance of excitotoxicity in mediating
MDMA-induced 5-HT toxicity is still unclear (Capela
et al., 2009).
One event associated with increased Ca2+ levels that
is observed after amphetamine exposure is proteolysis
of the cytoskeletal proteins, such as calpain-mediated
cleavage of spectrin (Lee et al., 1991). Spectrin proteolysis has been observed in the rat striatum after
METH treatments, whereas METH- and MDMAinduced increases in calpain activation and spectrin
proteolysis were blocked by the calpain inhibitor calpastatin in vitro (Staszewski and Yamamoto, 2006;
Warren et al., 2007; Suwanjang et al., 2012).
Excitotoxic processes can also lead to increased production of reactive species. Nitric oxide synthase is
elevated during METH exposure, and NO can readily
react with superoxide to produce peroxynitrite, a highly
reactive anion that produces toxicity through nitrosylation. Direct evidence for its involvement in METH
toxicity comes from data showing that METH-induced
depletion of DA is accompanied by increased levels of
striatal nitrate and nitrotyrosine (Anderson and Itzhak,
2006; Pacher et al., 2007; Wang et al., 2008b). Furthermore, blockade of nitric oxide synthase provides full
protection against METH-induced loss of DA (Itzhak
and Ali, 1996), whereas coadministration of a peroxynitrite decomposition catalyst partially prevents toxicity
(Imam et al., 1999). Specific targets have also been
identified as METH-induced nitrosylation, coupled with
loss of function, and have been observed on the VMAT-2
protein (Eyerman and Yamamoto, 2007). In addition to
nitric oxide synthase, other enzymes may also be
involved. Activation of xanthine oxidase can result in
increased superoxide production, whereas cleavage of
membrane phospholipids and release of arachidonic
acid by phospholipase A2 may also result in production
of superoxide after lipoxygenase- and cyclooxygenasemediated metabolism of arachidonic acid (Battelli et al.,
1995; Yamamoto and Raudensky, 2008).
d. Hyperthermia: a catalyst of other toxic processes.
Hyperthermia plays a well-established and important
role in mediating amphetamine toxicity. Both METH and
MDMA produce hyperthermia, and preventing the hyperthermia provides protection against subsequent neurotoxicity. Furthermore, lower ambient temperatures
attenuate, whereas higher ambient temperatures exacerbate, both METH and MDMA toxicity (Ali et al., 1994;
Bowyer et al., 1994; Broening et al., 1995; Malberg and
Seiden, 1998; Goni-Allo et al., 2008). Hypophysectomy,

914

Korpi et al.

thyroparathyroidectomy, as well as treatment with high


doses of haloperidol or diazepam that reduce body
temperature during amphetamine treatments, also reduce the observed toxicity (Bowyer et al., 1994; Sprague
et al., 2003). A large number of neurotransmitter systems
appears to be involved in producing the hyperthermic
effect, because blockade of receptors, including the
NMDAR; D1R, D2R, and D3R; a1-adrenoreceptors;
5-HT2 receptors; as well as the s and OX1 receptors,
all prevent or reduce METH-induced hyperthermia
while protecting against METH toxicity (Sonsalla
et al., 1991; Farfel et al., 1992; Doyle and Yamamoto,
2010; Seminerio et al., 2011, 2012; Rusyniak et al.,
2012; Ares-Santos et al., 2012; Kikuchi-Utsumi et al.,
2013; Robson et al., 2013a; Sabol et al., 2013; Baladi
et al., 2014).
The effect of body temperature on amphetamine
toxicity appears to depend on a specific interaction,
because hyperthermia on its own does not produce
METH-characteristic toxicity (Bowyer, 1995). The effect
of temperature is most likely due to a general catalytic
effect on the enzymatic and nonenzymatic reactions,
such as those resulting in production of reactive species.
For instance, it has been shown that hypothermia
prevents METH-induced oxidative stress (Fleckenstein
et al., 1997c; Yamamoto et al., 2010), whereas hyperthermia significantly increases oxidative stress and
other measures of toxicity (Silva et al., 2014).
Interestingly, the cathinone 4-MMC produces transient hypothermia rather than hyperthermia in rats
(Shortall et al., 2013), and its neurotoxicity is enhanced
in mice at elevated ambient temperature (26 C) when it
causes hyperthermia (Miller et al., 2013). By using
telemetric body core temperature monitoring in mice, it
was found that high doses of 4-MMC produced complex
changes, first inducing hypothermia that was followed
by hyperthermia (Angoa-Perez et al., 2012).
e. Other factors: inflammation, blood-brain barrier
disruption, protective preconditioning. Recently, there
has been more focus on additional factors that may
mediate amphetamine neurotoxicity. METH- and
MDMA-induced activation of microglia and subsequent
increases in proinflammatory cytokines contribute to
toxicity. In contrast, 4-MMC has failed to induce striatal
microglial activation and astrogliosis even at high doses
in mice, although METH was active in the same
conditions (Angoa-Perez et al., 2012). Similarly,
amphetamines are reported to produce defects in the
ubiquitin proteasomal system and blood-brain barrier
integrity, whereas stress and human immunodeficiency
virus infection also represent modulatory factors
(Yamamoto et al., 2010). Alterations in neurogenesis
from the subventricular zone stem/progenitor cells and
the dentate gyrus have also been reported (Goncalves
et al., 2014), and toxic products such as ammonia,
resulting from METH-induced damage to peripheral
organs may play a further role in mediating neurotoxicity

(Xie and Miller, 2009; Seminerio et al., 2013; Halpin et al.,


2014 ).
There are also factors that mediate neuroprotection,
such as preconditioning with low doses of amphetamines before a toxic challenge. Preconditioning with
escalating doses of METH (0.52.5 mg/kg) for up to 2
weeks before a subsequent toxic METH challenge
provides full protection against METH-induced DA
toxicity (Graham et al., 2008; Cadet et al., 2009; Hodges
et al., 2011). Possible mechanisms have been investigated in an in vitro study that found that low concentrations of METH produced changes in the pERK1/2
level, by decreasing it in basal conditions and increasing
it in response to the toxic 6-hydroxydopamine challenge
(El Ayadi and Zigmond, 2011). METH also upregulated
levels of the antiapoptotic protein Bcl-2. Also voluntary
running wheel endurance-type exercise in mice has
been shown to be protective against METH toxicity,
an effect that appears to be mediated by antioxidant
adaptations in brain microvasculature and tightjunction proteins at the blood-brain barrier (Toborek
et al., 2013). Meanwhile, for MDMA it was shown that
pretreatment with a low doses (3 mg/kg) protected
against toxicity of a subsequent toxic dose of MDMA
(12.5 mg/kg) administered 1 week later. The pretreatment was associated with increased levels of the inflammatory cytokine IL-1b after 6 hours, while slowly
producing a significant increase in the interleukin
receptor antagonist IL-1ra reaching a peak 4 days after
the preconditioning (Cohen et al., 2013). The neuroplasticity that occurs in response to changes in cellular
inflammatory and oxidative stress processes may play
an important role in modulating potential neurotoxicity
in humans, because such preconditioning may happen
naturally during dose escalation that occurs in human
users.
16. Evidence of Long-term Plasticity and Cognitive
Effects after Stimulant Use in Humans.
a. Cognitive behaviors. Deficits in abstinent METH
users compared with control subjects have been observed on tests such as immediate and delayed recall
and digit span tests and measuring performance on
memory-related domains, such as verbal memory, longterm declarative memory, and working memory. Deficits are also observed on tests such as the Wisconsin
card sorting test, Stroop test, and trail-making test,
indicating deficits in the cognitive domains of executive
function and attention (Simon et al., 2000; Kalechstein
et al., 2003; Han et al., 2008; Rendell et al., 2009).
Furthermore, METH users appear to experience higher
levels of anxiety and depression than nonusers (Zweben
et al., 2004; Vik, 2007).
Similar to METH, abstinent MDMA users also show
deficits in certain cognitive domains. The most common
observations include deficits in short-term and longterm memory, verbal reasoning, attention, and executive functions as well as decreased performance in tests

Drug-Induced Neuroplasticity

reflecting aspects of general intelligence (Parrott et al.,


1998; McCann et al., 1999; Gouzoulis-Mayfrank et al.,
2000; Heffernan et al., 2001; von Geusau et al., 2004).
Other studies of abstinent MDMA users have also
reported higher ratings of depression and anxiety in
addition to decreases in memory performance (McCardle
et al., 2004; Thomasius et al., 2006). However, many
studies do not properly match MDMA users and
controls in terms of education, intelligence quotient,
and other variables. When subjects are properly
matched, no consistent detrimental effects of MDMA
are observed (Halpern et al., 2011).
b. Brain imaging of receptors, transporters and
activity/metabolism. PET scans have been used to
assess monoamine-related ligand binding in human
amphetamine users. Similar to what is seen in animals,
decreased DAT and VMAT2 binding has been observed
in the striatum and FC of abstinent METH users, who
also performed lower than controls on measures of
cognitive function, including memory. However, correlations between DAT binding and the time of abstinence
or between DAT binding and measures of cognitive
function were not always observed (McCann et al.,
1998a, 2008; Sekine et al., 2001; Johanson et al., 2006).
PET studies measuring cerebral metabolism have also
reported differences between abstinent METH users and
nonusers. For instance, it was reported that METH users
showed decreased metabolism in the parietal cortex and
thalamus (Volkow et al., 2001), whereas hypometabolism
is observed in the FC of METH users, who also made
more errors of preservation on the Wisconsin card sorting
test than control subjects (Kim et al., 2005, 2009).
Decreased performance of METH users on a vigilance
task and increases in anxiety and depression-like symptoms were associated with lower metabolism in the
cingulate and insula but higher metabolism in the FC.
Moreover, in METH users, task performance was negatively correlated with metabolism in the cingulate and
insula, whereas a positive correlation was observed in
controls (London et al., 2004, 2005).
Abstinent MDMA users show decreases in SERT
binding in several brain regions, including cortical
regions, thalamus, hypothalamus, and striatum, and
the extent to which the binding is decreased is correlated with the amount of previous MDMA use. Additionally, abstinent MDMA users perform worse on tests
of memory function, and their performance on these
tests is correlated with insular and hippocampal SERT
binding (McCann et al., 1998a, 2008; Kish et al., 2010).
However, a previous SPECT study found decreased
SERT binding only in abstinent female, but not male,
subjects. Furthermore, the decreased binding appeared
to recover with protracted abstinence, reaching a level
higher than control subjects (Reneman et al., 2001a).
Decreased metabolism has also been reported in MDMA
users in the cingulate, HC, amygdala, and striatum;
however, no correlation was found between metabolism

915

and the amount of MDMA consumed (Obrocki et al.,


1999; Buchert et al., 2001). Brain 5-HT2A receptor
binding in MDMA users appears to be decreased in all
cortical regions, whereas abstinent users show increased 5-HT2A receptor binding, but only in the
occipital cortex (Reneman et al., 2002). Furthermore,
the abstinent MDMA users scored lower on a test of
verbal memory, and memory performance was positively correlated to the amount of 5-HT2A receptor
binding (Reneman et al., 2000). However, the relationship between memory function and 5-HT2A receptor
binding is hard to interpret, because the abstinent users
had higher levels of 5-HT2A binding than controls but an
overall lower memory performance.
c. Magnetic resonance imaging studies. Both structural and functional MRI studies have reported differences between METH users and nonusers. It has been
reported that abstinent METH users have lower gray
matter density in the FC as well as decreased Wisconsin
card sorting test task performance compared with drugfree control subjects (Kim et al., 2006). Decreased
cingulate and hippocampal volumes that correlated
with performance on a verbal memory test (Thompson
et al., 2004) and correlations between increased METH
exposure and reductions in orbital and medial PFC
volumes have also been noted (Daumann et al., 2011).
Interestingly, striatal volume is reportedly increased in
abstinent METH users in a study that, however, did not
find any evidence of decreased cognitive performance
(Chang et al., 2005). Another study also found increased
putamen volume in METH users that correlated negatively with performance on the go/no-go test for impulsivity (Jan et al., 2012). Decreased volumes of certain
cortical regions and the HC have also been reported in
MDMA users (Cowan et al., 2003; den Hollander et al.,
2012).
Functional MRI studies also report differences between amphetamine users and nonusers. Two studies
reported that METH users showed increased impulsivity as measured by a delay-discounting test and showed
decreased activity in the cingulate and dorsolateral
PFC (Monterosso et al., 2007; Hoffman et al., 2008).
Decreased activation was also observed in the frontal
and insular cortices of METH users in response to
a Stroop task, although task performance was not
consistently different from control subjects (Salo et al.,
2009; Nestor et al., 2011).
In fMRI studies of MDMA users, a weaker BOLD
signal during working memory task performance was
noted in the frontal and temporal cortex in heavy
compared with light users, but when comparing users
to nonusers there was no clear effect on BOLD signal
and, furthermore, task performance was also similar
between MDMA users and controls (Daumann et al.,
2003). Increases in BOLD signal in the frontal gyrus in
response to a memory task on which MDMA users also
were not impaired have been reported as well (Moeller

916

Korpi et al.

et al., 2004). Differences in hippocampal BOLD signal in


response to episodic or associative memory tasks
revealed decreases in MDMA users versus controls
and in continuing versus abstinent MDMA users;
however, here also there were no differences observed
on the actual task performance (Daumann et al., 2005;
Becker et al., 2013). It is possible that differing cognitive
strategies between users and nonusers may account for
the observed variability in BOLD signal intensity
(Honey et al., 2000).
d. Magnetic resonance spectroscopy and analysis of
postmortem tissue. In addition to alterations in structural and functional MRI studies, alterations have also
been reported in cerebral blood flow and white matter
integrity using perfusion and fractional anisotropy MRI
techniques (Iyo et al., 1997; Chang et al., 2002; Alicata
et al., 2009). Furthermore, magnetic resonance spectroscopy has been employed to detect changes in chemical
markers, such as N-acetylaspartate, choline, creatine,
and phosphocreatine, putatively reflecting abnormalities
in neuronal cell membrane integrity or metabolism
(Panenka et al., 2013). Decreases in phosphocreatine
and correlations between N-acetylaspartate/creatine ratio and verbal memory performance have been reported
in METH and MDMA users, respectively (Reneman et al.,
2001b; Sung et al., 2013).
Analyses have also been made of postmortem brain
material, and a decrease in striatal DA, DAT, and TH
levels has been reported (Wilson et al., 1996). Decreases
in D1R-mediated stimulation of adenylyl cyclase (AC) in
response to DA, suggestive of alterations in the level
and function of DA-related proteins, have also been
reported (Tong et al., 2003). However, because brains in
these studies are primarily sourced from individuals
who overdosed on METH, their relevance for normal
recreational users must be questioned.
e. Prospective and experimental studies.
Neurocognitive studies in humans have produced vastly
divergent results, with certain studies reporting decreased performance on various measures of cognitive
performance, whereas others report no difference. Furthermore, specific deficits are not often replicated
between studies. Neuroimaging studies also suffer from
large result variability, with various differences in taskrelated signal intensities, which often occur without any
overt effect on task performance. Another major issue
confounding the interpretation is that causality cannot
be determined from cross-sectional studies, and it is
likely that pre-existing differences between MDMA
users and control subjects account of a large part of
the observed differences.
Prospective studies have been performed to address
some of these issues. In a Dutch study, subjects who
were considered to be at high risk of commencing
MDMA use were recruited for an in-depth set of neuroimaging experiments, both at baseline before any
MDMA and at a follow up within 3 years after the

baseline scans. The individuals who started using


MDMA displayed a slight decrease in regional relative
cerebral blood flow in the globus pallidus and putamen
as well as decreased fractional anisotropy in the
thalamus and frontoparietal white matter. However,
no effect was found on N-acetylaspartate/creatine ratio,
and, importantly, no difference was found on the most
important measure, namely SERT binding between
new MDMA users and those who remained MDMA
naive. A dose-response relationship between cumulative MDMA use and SERT binding was also not present,
suggesting that low-dose MDMA use (mean of six
tablets) has no effect on SERT-binding in novel MDMA
users (de Win et al., 2008).
The same population was also subjected to a wide
array of neurocognitive tests. Subjects who commenced
MDMA use scored significantly lower on immediate and
delayed verbal recall compared with MDMA-naive
subjects; however, no effects were observed on any other
neurocognitive tests, and all scores were within the
normal ranges. Importantly, the verbal memory scores
of MDMA users did not decrease between the baseline
and follow up, and the difference between users and
nonusers was actually due to an increase in performance of the nonusers during the follow up, suggesting
that low-dose MDMA use has no effect on measures of
neurocognitive function, including memory (Schilt
et al., 2007; Krebs and Johansen, 2008).
A second prospective study focused specifically on
users who had a higher MDMA consumption, excluding
subjects who had used less than 10 tablets and thus
comparing heavier users (mean of 34 tablets) to individuals having consumed cannabis only. Like in the
Dutch study (de Win et al., 2008), no effects were found
on verbal memory, and significant differences between
groups being reported only for immediate and delayed
visual memory. Here, the difference was due to a decreased performance in the MDMA users (Wagner et al.,
2013).
Although prospective studies are more informative
than cross-sectional studies, there are still a number of
confounding factors that only experimental studies can
address. Although a number of studies have administered METH or MDMA experimentally to humans (Mas
et al., 1999; Harris et al., 2002; Hart et al., 2012;
Carhart-Harris et al., 2014), long-term cognitive outcomes have usually not been reported. One study that
did reported no effects on cognitive function measured
at 1 year or longer after the drug treatments (Mithoefer
et al., 2013). In summary, although a number of crosssectional studies report cognitive and neurologic deficits
or alterations as a result of the use of amphetamines,
these studies often suffer from methodological limitations that could lead to overestimation of the actual
drug effect. Most results are not replicated in prospective or experimental studies and, furthermore,
although amphetamine users sometimes score lower

Drug-Induced Neuroplasticity

than control subjects, they still score well within the


normal range (Hart et al., 2012), suggesting that any
existing drug effect is very limited and does not produce
any clinically significant impairment.
17. Conclusions. Amphetamines are a class of drugs
that exert powerful stimulatory effects on brain function
by mechanisms that include the inhibition of monoamine
reuptake and enhancement of monoamine release. They
are used medically to treat disorders such as ADHD, but
nonmedical recreational use of amphetamines is also
widespread, because amphetamines are the second most
commonly used illicit drug after cannabis. A large
number of animal studies indicate that amphetamines,
in addition to being addictive, are neurotoxic and even
lethal at high doses. It is still unclear whether similar
neurotoxicity occurs in human recreational users. Some
studies suggest that amphetamine users may display
certain abnormalities in the monoamine systems reminiscent of neurotoxicity observed in animals after highdose treatments. However, because most studies are
cross-sectional, no clear conclusions can be drawn with
regards to causality. Furthermore, amphetamine users
do not show any clear evidence of persistent functional
deficits and consistently score within normal ranges on
neuropsychological tests of memory function.
C. Nicotine
Smoking has been estimated to cause smokers to lose
one decade of their lives compared with never-smokers
(Jha et al., 2013). Early smoking cessation (,40 years of
age) seems to give back most of those years, indicating
some reversibility of smoking-related damage in the
body.
Nicotine is the principal neuroactive alkaloid responsible for tobacco addiction (Jaffe and Kanzler, 1979;
Stolerman and Jarvis, 1995; Pontieri et al., 1996; Pich
et al., 1997). Despite widespread prevalence of tobacco
smoking, it was not until the mid-1990s that there was
a scientific consensus that nicotine is addictive in
a similar manner to other drugs of abuse such as cocaine
(Stolerman and Jarvis, 1995; Pontieri et al., 1996; Pich
et al., 1997). Recent studies have shown that nonnicotine components in tobacco are also reinforcing or
influence nicotine SA. Anabasine dose dependently
increased nicotine SA at 0.02 mg/kg s.c. in rats, whereas
both anabasine and anatabine reduced nicotine SA at
2 mg/kg s.c. (Hall et al., 2014). Another non-nicotine
tobacco component, norharmane, was behaviorally
reinforcing on its own, and these reinforcing effects
were additive with those of nicotine (Arnold et al., 2014).
Norharmane is a MAO inhibitor. It has been suggested
that the inhibition of MAOs contributes to the rewarding effects of nicotine (reviewed in Kapelewski et al.,
2011).
Smoking is well known to be associated with many
adverse effects, including increased risk of cancers and
cardiovascular disease. Although the abuse potential

917

and harmful effects of nicotine were recognized, there


continued to be interest in the potential beneficial acute
effects of nicotine for cognitive enhancement (Everitt,
1997; Robbins et al., 1997; Changeux et al., 1998; Levin,
2013) and neuroprotection (Buckingham et al., 2009;
Hernandez and Dineley, 2012; Kawamata et al., 2012).
Others pointed to cognitive impairment in heavy smokers during abstinence as evidence of adverse effects on
neural plasticity associated with learning and memory
(Abrous et al., 2002). The diversity of nicotinic acetylcholine receptor subtypes contributes to the diverse
functional effects of nicotine.
1. Nicotinic Acetylcholine Receptors. Nicotinic acetylcholine receptors (nAChRs) belong to the "Cys-loop"
superfamily of multisubunit transmembrane ionotropic
neurotransmitter receptors (Albuquerque et al., 2009).
High-resolution structural studies have shown that
nAChRs are assembled from five subunits, which form
to create a central pore (Unwin, 2005; Baenziger and
Corringer, 2011). Sixteen different subunits have been
identified in mammals (Millar and Gotti, 2009). Of the
mammalian nAChR subunits, only a7 and a9 can form
homopentamers and only a7 forms endogenous homopentameric receptors (Vetter et al., 2007; Millar and
Gotti, 2009). The other a subunits usually form heteropentamers containing two a subunits and three b
subunits. Receptor subtypes are named according to
their subunit composition, with an asterisk used to
indicate cases where one or more additional subunit
types may also be present (Lukas et al., 1999; Millar
et al., 2014). nAChRs composed of different subunits are
differentially distributed throughout the brain. a4b2,
a4a5b2, a3b4*, a6b2b3*, and a7 are expressed in the
VTA and SN; a4b2 and a7 in the hypothalamus; a4b2,
a4a5b2, a6b2b3, and a6a4b2b3 in the striatum; a4b2,
a4a5b2, a3b4, and a7 in the HC; and a4b2, a4a5b2, and
a7 are expressed in the cerebral cortex (Gotti et al.,
2006, 2007). Nicotine is an agonist at all nAChRs except
for a9/10 nAChRs, expressed by vestibular and cochlear
mechanosensory hair cells (Elgoyhen et al., 2001; Scholl
et al., 2014) and a subset of dorsal root ganglion sensory
cells (Papadopolou et al., 2004), at which nicotine is an
antagonist (Elgoyhen et al., 2001; Millar et al., 2014).
nAChRs are widely expressed in the midbrain DAergic
neurons projecting to the dorsal striatum, ventral
striatum and NAc, and PFC. A remarkable diversity of
b2* nAChR subtypes modulates DA release in these
projections (Grady et al., 2007; Gotti et al., 2009;
Livingstone and Wonnacott, 2009). These projections,
in particular those to the NAc and ventral striatum, are
likely involved in reward pathways and the reinforcing
properties of nicotine contributing to addiction. Nicotine increases firing of midbrain DA neurons (Grenhoff
et al., 1986; Mereu et al., 1987) and increases DA release
in the NAc (Imperato et al., 1986; Rahman et al., 2003).
Selective blockade of nAChRs in the VTA is sufficient to
prevent nicotine-induced DA release in the NAc,

918

Korpi et al.

nicotine-induced locomotion and nicotine SA (Nisell


et al., 1994). In rodents, when nicotine injections are
repeated, there is a sensitization of the nicotine-induced
locomotor activation (Clarke and Kumar, 1983; Ksir
et al., 1985) and the nicotine-induced NAc DA release
(Benwell and Balfour, 1992; Benwell et al., 1995;
Balfour et al., 1998; Schoffelmeer et al., 2002). Human
studies confirm a role for the DAergic system in the
effects of nicotine but paint a more complex picture. In
human PET studies, reduction of [11C]raclopride binding potential is used as an indirect measure of increase
in DA release. After nicotine administration by nasal
spray, there was no overall change in [11C]raclopride
binding in any striatal regions, but baseline [11C]
raclopride binding potential in the associative striatum
correlated negatively with the Fagerstrm score, an
index of nicotine dependence, and nicotine concentration correlated negatively with binding potential in the
limbic striatum (Montgomery et al., 2007a). However, in
tobacco-dependent individuals, smoking a regular cigarette resulted in greater reduction in ventral striatal
[11C]raclopride binding potential than smoking a denicotinized cigarette (Brody et al., 2009).
At higher doses, nicotine is aversive rather than
rewarding. Nicotine also often has aversive effects on
first exposure. Stronger aversive reactions in first-time
smokers are associated with reduced likelihood of development of habitual tobacco use (Sartor et al., 2010).
There is therefore increasing interest in understanding
the mechanisms of aversive responses to nicotine,
because aversive reactions on initial exposure may
reduce propensity to addiction. The MHb-IPN pathway,
which expresses a unique subset of a5, a3, and b4
nAChR subunits, has been implicated in aversive responses to nicotine (reviewed in Antolin-Fontes et al.,
2015) (see Fig. 10). In particular, a4b2a5* nAChR
subtypes have been implicated (reviewed in Fowler
and Kenny, 2014).
2. Upregulation of Nicotinic Receptors. Sustained
upregulation of nAChRs is a mechanism of nicotineinduced neuroplasticity that is likely of relevance to
nicotine addiction and which may contribute to some of
the sensitization effects observed on repeated nicotine
administration (reviewed in Govind et al., 2009). Generally, chronic exposure to agonists results in receptor
downregulation, whereas, chronic exposure to antagonists results in receptor upregulation. nAChRs are
remarkable in that exposure to the agonist nicotine can
cause long-lasting upregulation of nAChRs (see Fig. 9).
Comparison of brains of smokers and nonsmokers
revealed increased nicotine binding sites in the postmortem brains of smokers (Benwell et al., 1988; Breese
et al., 1997; Perry et al., 1999). Similar increases were
subsequently reproduced in the brains of mice and rats
after repeated nicotine administration (Marks et al.,
1983; Schwartz and Kellar, 1983; Yates et al., 1995;
Pietila et al., 1998). The upregulation does not appear to

involve increases in mRNA expression (Marks et al.,


1992) but rather reflects posttranscriptional mechanisms. The a4b2 nAChR subtype has fairly consistently
been implicated in nicotine-induced upregulation. In
vitro studies in fibroblast and HEK cell culture systems
expressing a4b2 nAChRs showed that nicotine exposure resulted in receptor upregulation as determined by
ligand binding (Peng et al., 1994; Bencherif et al., 1995;
Rothhut et al., 1996; Eilers et al., 1997; Gopalakrishnan
et al., 1997; Warpman et al., 1998; Whiteaker et al.,
1998; Harkness and Millar, 2002). Upregulation of
these subunits was further confirmed by transfection
of a4 and b2 subunits tagged with fluorescent protein
into both HEK cells and neurons from the ventral
midbrain (Nashmi et al., 2003). In vivo, upregulation
of a4 in response to nicotine treatment was shown in
a transgenic mouse expressing a4 tagged with yellow
fluorescent protein (Nashmi et al., 2007).
Replacement of b2 with b4 in receptors containing a4
or a3 reduces nicotine-induced upregulation assayed by
radioligand binding (Wang et al., 1998), whereas addition of a5 or b3 may block upregulation (Perry et al.,
2007; Mao et al., 2008). Association of a4b2 with a5 may
explain why nicotine-induced nicotinic receptor upregulation is not observed in some brain regions (Mao
et al., 2008). Chronic nicotine failed to induce nAChR
upregulation in the adrenal gland, which has little a4b2
nAChRs (Flores et al., 1997; Davila-Garcia et al., 2003).
Comparison of nicotine-induced upregulation of a3b2,
a4b2, and a6b2 nAChRs in transfected cells revealed
subtype differences in sensitivity to nicotine concentration and the time course of upregulation (Walsh et al.,
2008). a4b2 nAChRs were most sensitive, upregulating
at lower nicotine concentration, whereas a6b2 nAChRs
required higher concentrations of nicotine, and a3b2
nAChRs required the highest concentrations of nicotine. a6b2 nAChRs showed the fastest upregulation,
followed by a3b2 nAChRs, whereas a4b2 nAChRs
showed the slowest response. On chronic nicotine
exposure, a6b2* nAChRs are downregulated rather
than upregulated (Lai et al., 2005; Perry et al., 2007;
Doura et al., 2008; Marks et al., 2014). a6b2* nAChR
downregulation was more sensitive to the dose of
chronic nicotine treatment than a4b2* nAChR upregulation (Marks et al., 2014).
Thus, different receptor nAChR receptor subtypes
may respond differently to different doses, durations,
and patterns of nicotine exposure. For example, within
the VTA, it appears that chronic nicotine upregulates
a4* nAChRs on GABAergic neurons but not on DA
neurons (Nashmi et al., 2007). Similarly, in the SN,
chronic nicotine increased functional a4b2* nAChRs on
GABAergic neurons of the SN pars reticulata but not on
DA neurons of the SN pars compacta (Xiao et al., 2009).
It has been proposed that exposure to nicotine leads to
nAChR upregulation in the VTA, increasing excitability
and favoring induction of LTP resulting in sensitization

Drug-Induced Neuroplasticity

919

Fig. 9. Summary of the effects of acute (green) and chronic (red) nicotine on neuronal activities via different subtypes of nACh receptors in the ventral
tegmental area and the cortex/hippocampus. The nicotinic synapse depicts facilitatory actions of nicotine on presynaptic neurotransmitter release and
on postsynaptic LTP.

to the effects of DA leading to behavioral sensitization


and drug addiction (Govind et al., 2009). Further
studies are required to establish how the doses and
patterns of nicotine exposure associated with smoking
differentially regulate the various nAChR subtypes in
different brain regions.
Various mechanisms have been proposed to underlie
nicotine-induced upregulation of nAChRs. Importantly,
it was earlier shown that the mechanisms underlying
upregulation must be posttranslational rather than
transcriptional, because nicotine exposure did not alter
nAChR subunit mRNA levels (Marks et al., 1992).
Proposed mechanisms have included decreased turnover of receptors (Peng et al., 1994), desensitization of
surface receptors (Fenster et al., 1999), enhancement of
receptor maturation (Sallette et al., 2005), and slow
stabilization of receptors in a high-affinity state that is
more easily activated (Vallejo et al., 2005). However,
recent consensus of opinion, at least for upregulation of
a4b2* nAChRs, has converged on the notion that
nicotine acts as a pharmacological chaperone (Kuryatov
et al., 2005; Henderson et al., 2014; Srinivasan et al.,
2011, 2012). Pharmacological chaperoning of intracellular nAChRs is proposed to facilitate subunit assembly
in the ER and enhance export of assembled receptors
from the ER and their transport and insertion into the
membrane.
3. Nicotine-priming Effects and Upregulation of
Catecholamine Synthesis. Nicotine does not only sensitize or prime catecholaminergic systems via upregulation of nicotinic receptors, there is also evidence that
nicotine exposure sensitizes or primes certain catecholaminergic systems more directly by upregulating expression of TH, the rate-limiting enzyme for catecholamine

synthesis. Acute nicotine administration increased catecholamine synthesis in the NAc, hypothalamus, and HC,
whereas chronic administration of nicotine by daily subcutaneous injections in rats for 28 days potentiated this
response in the HC (Mitchell et al., 1989). This chronic
treatment increased TH activity in the HC (Joseph et al.,
1990). Studies of chronic treatment of 7 to 28 days showed
that the first increase in TH activity from 3 days in the
NEergic locus ceruleus (LC) and DAergic nuclei. Subsequently, the TH protein expression spread in the
NEergic projections to the forebrain, reaching the terminals up to 3 weeks later, consistent with axonal transport
of the enzyme from the cell bodies (Smith et al., 1991).
Subsequently, it was shown that 7-day treatment or even
a single injection of nicotine was sufficient to produce an
increase in TH activity in the terminal fields 34 weeks
later (Smith et al., 1991; Mitchell et al., 1993). A single
acute nicotine injection (0.8 mg/kg) was sufficient to
increase mRNA for TH in the LC cell bodies 2 to 6 days
later and, 4 weeks later, to increase TH activity in the
terminals and NE release in the hippocampus on challenge injection of nicotine (0.4 mg/kg) (Mitchell et al.,
1993).
NE is known to induce synaptic plasticity in the
hippocampal dentate gyrus (Neuman and Harley, 1983;
Stanton and Sarvey, 1985; Burgard et al., 1989; Dahl
and Sarvey, 1989; Harley, 1991, 1998, 2007), suggesting the possibility that nicotine-induced priming of
nicotine-induced NE release in the HC could induce
neuroplasticity. After priming with seven daily injections of nicotine, a challenge dose of nicotine produced
a long-lasting potentiation of medial perforant pathdentate gyrus evoked responses (Hamid et al., 1997).
This effect was blocked by both mecamylamine and

920

Korpi et al.

propranolol, indicating involvement of nAChRs and


b-adrenoceptors. Subsequently, it was shown that this
nicotine priming and challenge effect cross-primes with
other mechanisms, increasing TH expression in the LC.
Theta driving at 7.7 Hz, which is suggested to mimic
aspects of anxiety or stress responses, increased TH
activity in rat HC 15 to 33 days later (Graham-Jones
et al., 1985) and 0.4 mg/kg challenge injection of nicotine
induced medial perforant path-dentate gyrus longlasting potentiation in rats primed by theta driving
but not in rats that did not receive theta driving
(Markevich et al., 2007). Chronic clozapine treatment
activated the LC (Souto et al., 1979; Ramirez and Wang,
1986; Nilsson et al., 2005) and increased TH expression
in the LC (Verma et al., 2007). Rats primed with
clozapine (30 mg/kg) daily for 7 days showed nicotine
(0.4 mg/kg) challenge-induced medial perforant pathdentate gyrus long-lasting potentiation 21 to 28 days
later (Rajkumar et al., 2013). These data suggest that
nicotine alone or in conjunction with other stimuli can
induce TH expression in the LC, resulting in delayed
transport of TH to the afferent terminal fields facilitating nicotine-induced NE-mediated synaptic plasticity.
Nicotine may also induce synaptic plasticity within
catecholaminergic nuclei themselves.
4. Nicotine and Neuroplasticity within the Ventral
Tegmental Area. DAergic neurons of the VTA and SN
are tonically active, firing irregularly at low frequency,
but can switch to burst firing (Grace and Bunney,
1984a,b) (Fig. 9). The switch between tonic and burst
firing has been associated with signaling of novel events
and reward (Chergui et al., 1993; Cooper, 2002; Hyland
et al., 2002). A similar switch between tonic and burst
firing occurs in response to nicotine (Grenhoff et al.,
1986; Zhang et al., 2009). This burst firing is thought
to be mediated by activation of NMDARs on DAergic
cell bodies by glutamatergic afferents from the PFC
(Chergui et al., 1993; Tong et al., 1996). It is proposed
that stimulation of presynaptic a7 nAChRs on glutamatergic afferents (Jones and Wonnacott, 2004) releases
glutamate activating NMDARs on DAergic cell bodies to
produce the burst firing (Livingstone and Wonnacott,
2009). The activation of NMDARs not only induces acute
changes in firing but can induce long-lasting presynaptic
facilitation and postsynaptic NMDAR-dependent LTP,
resulting in sustained increases in the sensitivity of
the DAergic neurons to burst firing (Mansvelder and
McGehee, 2000; Mansvelder et al., 2002). Likely, these
mechanisms translate to behavioral nicotine-induced
increases in responsiveness to reward.
However, the essential role of presynaptic a7 nAChRs
in induction of burst firing is questioned by the observation that nicotine still triggers burst firing in a7-KO mice
(for discussion, see Livingstone and Wonnacott, 2009).
Meanwhile, the effect of nicotine on DAergic cell firing
mode is abolished in b2-KO mice (Mameli-Engvall et al.,
2006). The situation is further complicated by the effects

of nicotine on inhibition and excitability of the VTA. a7


nAChRs are also expressed by GABAergic neurons in
the VTA (Klink et al., 2001; Jones and Wonnacott, 2004).
A model integrating these data (McKay et al., 2007;
Livingstone and Wonnacott, 2009) proposes that nicotine
desensitizes a4b2 nAChRs on GABAergic interneurons
rapidly and preferentially (Wooltorton et al., 2003), thus
disinhibiting DAergic neurons while simultaneously activating a7 nAChRs on DAergic cell bodies and glutamatergic afferent terminals (Mansvelder and McGehee,
2000). It is suggested that b2* nAChRs on DAergic cell
bodies contribute by providing depolarization enhancing
excitability and thus bringing the cell closer to the
threshold to change firing patterns or to induce LTP
(Livingstone and Wonnacott, 2009).
The nicotinic control of DAergic firing is further
complicated by essential and complex roles of the
balance between activation and desensitization of
nAChR mechanisms (Picciotto et al., 2008). Although
activation and desensitization likely occur over too fast
a time scale to contribute directly to maintaining
nicotine-induced neuroplasticity, it is likely that this
balance plays a role in determining the conditions
permissive to induction of neuroplasticity. As mentioned earlier, chronic nicotine treatment upregulates
a4* nAChRs on GABAergic interneurons, but not on
DAergic neurons, of the VTA and on glutamatergic
terminals in forebrain. This is suggested to be sufficient
to explain tolerance of DAergic neuron firing in midbrain and sensitization of synaptic transmission in
forebrain (Tapper et al., 2004; Nashmi et al., 2007).
The role of a7 nAChRs remains more controversial.
Although most reports suggest that upregulation of a4*
nAChRs alone is sufficient to explain sensitization and
tolerance, others studying in vitro expression of human
nicotine receptors have suggested that functional inactivation of a4b2 nAChRs and a7 nAChRs but sustained inhibition of a3* nAChRs by chronic exposure to
nicotine contributes to tolerance (Olale et al., 1997).
In addition to the roles of nicotine receptors on
GABAergic interneurons versus DAergic neurons in
the VTA, effects of nicotine on AMPAR/NMDAR ratios
in VTA DA neurons have been implicated in nicotinic
regulation of VTA plasticity. Drugs of abuse, including
nicotine, were found to increase AMPAR/NMDAR ratios
in midbrain DA neurons (Saal et al., 2003). In the case of
nicotine, a single dose of 0.5 mg/kg i.p. was administered
to mice, and 2430 hours later AMPAR/NMDAR ratios
were increased in VTA DA neurons in midbrain slices
(Saal et al., 2003). Subsequent investigations reported
that increases in AMPAR/NMDAR ratios likely occur
via a7 or b2* nAChR-mediated mechanism (Gao et al.,
2010a) and can occur as fast as within 1 hour of a single
nicotine injection.
A recent study by Doyon et al. (2013) on rats adds
further complexity, as a single dose of nicotine dramatically reduced DA release compared with baseline in the

Drug-Induced Neuroplasticity

NAc for at least 15 hours. Furthermore, nicotinetreated animals made more responses for obtaining
alcohol solution during the first operant sessions. These
effects were blocked by dihydro-b-erythroidine (b2*),
but not by methyllycaconitine (a7*). Nicotine pretreatment enhanced bath-applied ethanol-mediated increase
in the frequency of sIPSCs from GABA neurons in the
VTA slices, which was also seen during bath application
of GABAA receptor BZ agonist diazepam. Importantly,
nicotine sensitivity of DA release was not affected at 15
hours after the nicotine pretreatment, and also the bath
applied nicotine similarly enhanced the frequency of
sIPSCs in slices from saline- and nicotine-treated
animals. Therefore, further mechanisms were examined. A glucocorticoid/progesterone receptor antagonist,
mifepristone (RU486), given systemically or intra-VTA
before nicotine, attenuated the increases in sIPSC
frequency and responses for alcohol solution and rescued the impaired ethanol-induced DA release.
Whether the nicotine effect is peripherally or centrally
initiated is not known. Interestingly, both acute and
chronic nicotine as well as acute morphine reduce the
release of GABA in the VTA of mice (Vihavainen et al.,
2008), whereas morphine after nicotine enhanced the
release. In summary, these findings suggest that nicotine can enhance the activity of VTA DA neurons via
multiple mechanisms, including by influencing the
function of GABA neurons.
5. Nicotine and Neuroplasticity in the Lateral
Hypothalamic Orexin System. The lateral hypothalamic orexin (hypocretin) system has also been implicated in mechanisms of addiction and drug-induced
neuroplasticity in the VTA (for discussion, see Baimel
and Borgland, 2012; Baimel et al., 2012). Orexin
signaling to the VTA promotes synaptic plasticity by
potentiating glutamatergic inputs to DA neurons.
Acute nicotine administration excited orexin-expressing
neurons in the lateral hypothalamus/perifornical area
(Huang et al., 2011) and dose dependently increased Fos
expression in these brain regions (Pasumarthi et al., 2006).
Chronic administration of nicotine by repetitive injection
over 1014 days altered expression of mRNAs for preproorexin and orexin receptors (Kane et al., 2001, 2001).
The OX1 (hcrt-R1) receptor antagonist, SB-334867
[N-(2-methyl-6-benzoxazolyl)-N-1,5-naphthyridin-4-yl
urea], reduced nicotine SA in animals on both fixed ratio
and progressive ratio schedules (Hollander et al., 2008).
Fixed ratio schedules are thought to measure drug
reinforcement, whereas progressive ratio schedules
are thought to measure motivation to obtain the drug.
Furthermore, infusion of SB-334867 into the insular
cortex also reduced SA on the fixed ratio schedule
consistent with the observation that tobacco smokers
with damage to the insula are more likely to quit
smoking without relapse and the recurrence of urges
to smoke (Naqvi et al., 2007). Systemic administration
of SB-334867 also prevented reduction of the threshold

921

for intracranial self-stimulation by nicotine (Hollander


et al., 2008). Together these data suggest that nicotine
may act via the orexin system to induce synaptic
plasticity in the VTA, thus modulating responses to
reward. Involvement of the orexin system raises the
question of effects of nicotine on feeding. Overall,
nicotine decreases food intake, but likely this is mediated by actions in the brain stem rather than the VTA,
because it is blocked by mecamylamine infusion into the
fourth ventricle (Guan et al., 2004).
6. Nicotine and Glutamatergic Neuroplasticity.
There are also numerous reports on nicotine-induced
neuroplasticity in the HC and cerebral cortex. For
example, investigations of the effects of nicotine agonists and antagonists on LTD induction in wild-type, a2-,
a7-, and b2-KO mice revealed that LTD-inducing stimulation releases ACh activating a7 nAChRs to suppress
LTD at hippocampal CA3-CA1 synapses, but this effect is
reversed by nicotine via a non-a7, non-a2, and non-b2
receptor mechanism (Nakauchi and Sumikawa, 2014). In
hippocampal slices from animals chronically treated with
nicotine, expression of a4* nAChRs on glutamatergic
terminals is increased, and acute exposure to nicotine
during tetanic stimulation increases induction of LTP in
medial perforant path (Nashmi et al., 2007) (Figs. 9 and
10). Meanwhile, in cultured hippocampal neurons, activation of presynaptic a4b2* nAChRs produced a glutamatergic neurotransmission-dependent enlargement of
postsynaptic dendritic spines (Oda et al., 2014). Chronic
treatment with nicotine (2 mg/kg/day s.c.) for 6 weeks
reversed beta-amyloid (Ab)-induced memory and hippocampal LTP deficits (Srivareerat et al., 2011) and chronic
stress-induced intensification of Ab-induced long-term
memory and synaptic plasticity deficits (Alkadhi, 2011;
Alkadhi et al., 2010, 2011). Although the interaction
between acetylcholine receptors (AChRs) and Ab is
complex, overall activation of AChRs can have protective
effects against Ab toxicity, leading to an interest in
developing a7 agonists for treatment of Alzheimers
disease (reviewed in Lombardo and Maskos, 2015).
In freely moving mice, acute nicotine administration
(1 mg/kg i.p.) induced medial perforant path-dentate gyrus
synaptic potentiation by a mechanism dependent on DA
receptors in the HC and activity of VTA neurons (Tang
and Dani, 2009). Adult rats trained to intravenously selfadminister nicotine 0.04 or 0.08 mg/kg per infusion,
achieving daily nicotine intakes of approximately 200
and 300 mg/kg, respectively, showed decreased cell proliferation and increased cell death in the dentate gyrus of
the HC (Abrous et al., 2002). Nicotine SA under extended
access (0.03 mg/kg/infusion, 21 hours/day for 4 days) with
intermittent periods of deprivation (3 days) for 14 weeks,
but not nicotine intake under limited access (1 hour/day
for 4 days), enhanced proliferation and differentiation
of hippocampal neural progenitors (Cohen et al., 2015).
Together these data represent evidence for effects of
nicotine on plasticity in cortical and hippocampal

922

Korpi et al.

pathways, but it is less clear that this plasticity is of any


functional relevance to the persistent neuroplasticity
underlying behaviors associated with addiction. The
evidence for involvement in addiction is stronger for
midbrain pathways.
7. Nicotine and the Habenula. Among midbrain
pathways involved in nicotine addiction, the habenulointerpeduncular pathway has recently received considerable attention (Velasquez et al., 2014; Antolin-Fontes
et al., 2015). The mammalian habenula is located at the
posterior medial end of the dorsal thalamus next to the
third ventricle. The habenula is phylogenetically highly
conserved across vertebrates (Concha and Wilson, 2001;
Aizawa et al., 2005; Dadda et al., 2010) and connects
limbic forebrain nuclei with midbrain and hindbrain
nuclei (Sutherland, 1982). The MHb receives major
inputs from the septum through the stria medullaris
and has a major output to the IPN through the
fasciculus retroflexus, whereas the LHb makes a minor
contribution to this projection (Herkenham and Nauta,
1979; Swanson and Cowan, 1979; Qin and Luo, 2009).
The MHb-IPN tract highly expresses a-bungarotoxininsensitive nAChRs (Mulle et al., 1991), likely containing a4, and a5, a3, and b4 nAChR subunits encoded by
the CHRNA5-A3-B4 gene cluster (Aizawa et al., 2012;
Shih et al., 2014). This gene cluster has been linked with
increased nicotine consumption and dependence in
human genetics studies (Berrettini et al., 2008; Bierut
et al., 2008; Lips et al., 2010; Liu et al., 2010; Ware et al.,
2011). Consistent with these human genetic studies,
animal models indicate a role for the MHb-IPN pathway
in modulation of nicotine aversion and nicotine withdrawal (Salas et al., 2009; Fowler et al., 2011; Frahm
et al., 2011) (Fig. 10). Mice with null mutation of
Chrna5, the gene encoding for the a5 nAChR subunit,

showed a marked increase in nicotine intake that could


be rescued by expression of a5 subunits in the MHb
(Fowler et al., 2011). The b4 nAChR subunit is important for receptor activation. Asp398Asn mutation in the
a5 nAChR subunit is associated with increased tobacco
usage in humans (Bierut et al., 2008; Saccone et al.,
2009; Bierut, 2010), and a5 Asp398Asn was found to
interact closely with a b4 Ser435 residue important in
increasing nAChR currents (Frahm et al., 2011). In mice
overexpressing Chrnb4, the gene encoding for the b4
nAChR subunit, viral-mediated expression of a5
Asp398Asn subunits reduced nicotine aversion (Frahm
et al., 2011). In mice chronically treated with nicotine,
null mutations of the a2 and a5 subunits highly
expressed in the IPN reduced mecamylamine-induced
withdrawal, and in wild-type mice microinfusion of
mecamylamine into the Hb or IPN precipitated withdrawal (Salas et al., 2009). Moreover, lentiviral expression in the MHb of a gain-of-function Thr374Ile variant
of the b4 nAChR subunit associated with reduced risk of
smoking resulted in strong nicotine aversion in mice,
whereas expression of a loss-of-function Arg348Cys
variant failed to induce nicotine aversion (Slimak
et al., 2014). Nicotine-induced excitation of the MHb
neurons in mice was mediated by tachykinin NK1 and
NK3 receptors (Dao et al., 2014). In chronic nicotinetreated mice, intra-MHb infusions of NK1/3 antagonists
provoked withdrawal symptoms, suggesting that the
tachykinin receptors were in a sensitized state.
Although, as discussed above, the a5, a3, and b4
nAChR subunits that are strongly expressed in the Hb
are likely resistant to nicotine-induced upregulation;
nicotine produces persistent changes in the MHb and
the fasciculus retroflexus projection to the IPN through
selective neurotoxicity. Administration of high doses of

Fig. 10. Summary of the effects of nicotine on various brain regions, with a special reference to adolescent exposure and to neurotoxicity in the medial
habenula. DG, dentate gyrus; fr, fasciculus retroflexus; HC, hippocampus; IPN, interpeduncular nucleus; LH, lateral hypothalamus; MHb, medial
habenula; mPFC, medial prefrontal cortex; MPP, medial perforant path; NAc, nucleus accumbens; TH, tyrosine hydroxylase; VTA, ventral tegmental
area.

Drug-Induced Neuroplasticity

nicotine, continuous infusion of 5.72, 6.44, 7.13, 20.41,


and 43.1 mg/kg/day nicotine tartrate via osmotic minipump or intermittently at 11.32 mg/kg/day via once
daily subcutaneous injection over 5 days resulted in
neurodegeneration of the axons of the MHb neurons
that form the fasciculus retroflexus (Carlson et al.,
2001). More recent neurochemical investigations indicate that cholinergic neurons of the MHb are also
a major target of this nicotine-induced neurodegeneration (Ciani et al., 2005). Previously, rather than affecting the MHb, continuous administration of stimulants
amphetamine or cocaine had been found to cause
neurodegeneration in the LHb, which has a weaker
contribution to the fasciculus retroflexus (Ellison, 1992;
Carlson et al., 2000). It has been suggested that as the
fasciculus retroflexus is a phylogenetically primitive
tract carrying much of the negative feedback from
forebrain back onto midbrain reward circuitry; the
destruction of these descending control pathways by
drug binges could result in loss of forebrain control of
addictive behavior (Ellison, 2002). This is consistent
with the role for the MHb-IPN pathway in modulation of
nicotine aversion and nicotine withdrawal (see above).
It is not known whether human smokers show similar
neurotoxicity in the MHb-IPN pathway. It is also not
known whether there is any subsequent adaptive
neuroplasticity and functional recovery. An important
question that remains to be further addressed is the
extent to which this neurotoxicity in the MHb is reversible. Further investigations are needed to understand the implications for behavior and whether the
effects of the neurotoxicity are permanent. More recently, it has been found that repeated nicotine exposure in adolescent rats reduces MHb activity and
increases nicotine preference, suggesting that repeated
phases of nicotine exposure induce a functional switch
in the activity of MHb neurons in adolescent rats
leading to increased nicotine consumption (Lee et al.,
2015).
8. Long-term Sequelae of Adolescent Exposure to
Nicotine. It has been suggested that nicotine has
especially profound effects on neuronal networks (Fig.
10), in particular in the PFC, when exposure occurs
during adolescence (Goriounova and Mansvelder,
2012a,b). The onset of puberty is a period of significant
development of the emotional brain, including the midbrain DAergic nuclei, the hypothalamus, the NAc, the
striatum, and amygdala. In contrast, development of
the frontal cortical areas responsible for self-control and
decision-making lags behind, maturing throughout
adolescence and into adulthood (Sowell et al., 2003;
Giedd, 2004; Casey et al., 2005) (Fig. 4). Asynchronous
brain development may thus be responsible for the
characteristic traits of adolescence (Goriounova and
Mansvelder, 2012a), such as mood swings, impulsivity,
risk-taking, and susceptibility to peer influence in social
behavior (Orr and Ingersoll, 1995; Spear, 2000; Galvan

923

et al., 2006). Together, impulsivity, risk-taking, and


susceptibility to peer pressure can contribute to increased risk of onset of smoking in adolescents. Once
they smoke, adolescents appear more susceptible to
dependence at lower levels of cigarette consumption
(Colby et al., 2000; Kandel and Chen, 2000). Similarly,
in mice and rats there is increased vulnerability to
nicotine SA during early adolescence (Adriani et al.,
2002; Ahsan et al., 2014). It has been proposed that
adolescents may show this greater vulnerability to
nicotine addiction, because nicotine has greater positive
effects and smaller negative effects in adolescents than
adults (ODell, 2009). Moreover, the adolescent brain
appears to be exceptionally sensitive to nicotine-induce
neuroplasticity.
Acutely, smoking during adolescence is associated with
disruption of working memory and attention consistent
with reduced PFC activation (Jacobsen et al., 2005, 2007;
Musso et al., 2007). Moreover, smoking is a prospective
risk factor for impaired cognitive function in later life
(Cervilla et al., 2000; Richards et al., 2003), and adolescent tobacco use is associated with increased risk of
developing mental and behavioral problems, including
abuse of other drugs, later in life (Brown et al., 1996;
Johnson et al., 2000; McGee et al., 2000; Ellickson et al.,
2001; Brook et al., 2004). In rodents, adolescent animals
are more susceptible to nicotine-induced CPP than adults
(Vastola et al., 2002; Belluzzi et al., 2004; Shram et al.,
2006; Kota et al., 2009; Shram and Le, 2010; Brielmaier
et al., 2012; Ahsan et al., 2014). Moreover, nicotine
administration during, but not after adolescence, has
long-lasting effects on cognitive, emotional, and
addiction-related behaviors (Adriani et al., 2003, 2004;
Counotte et al., 2009, 2011; Iniguez et al., 2009). Together
these data point to the conclusion that adolescent nicotine
exposure has long-term effects on attention, cognition,
and emotional function in adulthood (Adriani et al., 2003;
Counotte et al., 2009, 2011; Goriounova and Mansvelder,
2012a,b).
A model synthesizing the data on effects of adolescent
exposure to nicotine on PFC in rodents (Goriounova and
Mansvelder, 2012a,b) proposes that, in the short-term,
adolescent nicotine exposure increases local inhibition
via nAChRs on GABAergic interneurons, while simultaneously decreasing glutamatergic excitatory inputs to
PFC neurons by increasing presynaptic mGlu2 expression. However, in the long-term, local inhibitory activity
normalizes, but there is a persistent and long-lasting
downregulation of presynaptic mGlu2 on glutamatergic
afferent terminals, resulting in increased excitability of
PFC pyramidal neurons associated with long-lasting
downregulation of synaptic short-term depression and
impaired attention. In rodents, the adolescent PFC is
especially sensitive to nicotine-induced changes in gene
expression (Schochet et al., 2005, 2008; Polesskaya et al.,
2007). In adolescence, exposure to nicotine increases
expression of immediate early genes used as functional

924

Korpi et al.

markers of neuronal activation (Leslie et al., 2004;


Schochet et al., 2005) and dendritic expression of mRNA
for dendrin, a protein associated with synaptic plasticity
(Schochet et al., 2008). Meanwhile, chronic nicotine
exposure increases expression of a wide range of genes
involved in neurotransmission, signal transduction, and
synaptic architecture (Polesskaya et al., 2007), and repeated nicotine exposure leads to increased phosphorylation of ERK and CREB involved in neuroplasticity
(Brunzell et al., 2003). After adolescent nicotine exposure,
nAChR levels in the PFC, likely representing primarily
nAChRs on GABAergic interneurons, are elevated in the
short-term but return to normal by 5 weeks (Counotte
et al., 2012), whereas in contrast, mGlu2 receptors are
strongly downregulated at this time point (Counotte
et al., 2011). In rats, acute nicotine exposure during
adolescence reduced PFC LTP in response to timed
presynaptic and postsynaptic activity (tLTP) but subsequently increased this LTP in adulthood (Goriounova
and Mansvelder, 2012b). These changes in LTP could be
explained by the changes in mGlu receptors, because
activation of mGlu2 receptors mimicked the acute decrease in tLTP and blocking of mGlu2 receptors mimicked
the long-term increase in tLTP. Increases in glutamate
release in the PFC also occur in NMDAR antagonist
models of aspects of schizophrenia. NMDAR antagonistinduced increases in PFC glutamate release are associated with deficits in attention and loss of impulse control,
and these behavioral effects can be reversed by administration of an mGlu2/3 agonist (Pozzi et al., 2011). Administration of mGlu2 agonist also improved attention
performance in animals that had been exposed to nicotine
during adolescence (Counotte et al., 2011).
The effects of adolescent nicotine exposure on mGlu2
are not unique to the PFC. Long-lasting changes in
mGlu2 function after adolescent nicotine exposure have
also been reported in the VTA and NAc, where they
modulate the rewarding properties of nicotine (Helton
et al., 1997; Kenny et al., 2003; Kenny and Markou,
2004; Liechti et al., 2007; Counotte et al., 2011). In these
brain regions, activation of mGlu2/3 receptors reduces
SA of nicotine (Liechti et al., 2007), and these receptors
are thought to play a role in development of drug
dependence and manifestation of affective withdrawal
symptoms (Kenny and Markou, 2004). However, the
precise role of mGlu2 downregulation in the VTA-NAc
system in the mechanisms of increased vulnerability to
nicotine addiction after adolescent nicotine exposure
requires further elucidation.
Recent findings suggest that long-term cigarette
smoking in adulthood is associated with accelerated
cerebral cortical thinning (Karama et al., 2015). Various
degrees of cortical thinning may be associated with
a number of neuropsychiatric diseases, such as schizophrenia, depression, dementias, and fetal alcohol spectrum disorders, and with an increased risk for them
(Narr et al., 2009; Peterson et al., 2009; Zhou et al.,

2011; Byun et al., 2012), and accelerated cortical thinning


may be a normal physiologic phenomenon during adolescence and even associated with better neuropsychological
performance (Squeglia et al., 2013). However, smoking
increases the prevalence of Alzheimers disease (Barnes
and Yaffe, 2011), and, therefore, it is important that
smoking-induced cortical thinning seems to be at least
partially reversible, albeit very slowly. Further studies of
the mechanism and extent of nicotine-induced structural
changes in the adult brain will likely be informative.
There is a strong association between habitual smoking and alcohol dependence, and there may be common
genetic risk factors (Bierut et al., 2000; Grucza and
Bierut, 2006; Mingione et al., 2012; Vrieze et al., 2013;
Anantharaman et al., 2014) (reviewed in Sturgess et al.,
2011). In particular, patients with chronic pain may be
susceptible to tobacco, alcohol, and opioid dependence
(Fishbain et al., 2012), and there is a strong association
between smoking and chronic pain (reviewed in Ditre
et al., 2011; Parkerson et al., 2013). Meanwhile, antidepressants have been used in treatment of alcohol
addiction, and antidepressants and MAO inhibitors
have been used in chronic pain management. It has
also been noted that MAO inhibitors increase tolerance
to alcohol in mice (Popova et al., 2000), and that tobacco
smoke exposure increases alcohol consumption in adolescent C56BL/6 mice (Burns and Proctor, 2013). Acetaldehyde is produced in high concentrations when
cigarettes are burned. There appear to be synergistic
interactions between acetaldehyde and both nicotine
and alcohol. Rats preferred a combination of acetaldehyde and nicotine to either substance alone (Rabinoff
et al., 2007). Moreover, rats could be trained to selfadminister acetaldehyde and subsequently consumed
more alcohol (Amit and Smith, 1985). Although most
alcohol metabolism occurs in the liver, some alcohol is
metabolized to acetaldehyde by microbial action already in the oral cavity. Deficient oral hygiene, heavy
alcohol drinking, and smoking modify oral microbiome
to increase acetaldehyde from local alcohol metabolism
(reviewed in Salaspuro, 2007). The acetaldehyde produced in this way may contribute not only to association
between smoking and alcohol dependence but also to
increased risk of gastrointestinal tract cancers.
9. Conclusions. Nicotine addiction leads to increased risk of severe adverse health effects caused by
tobacco. Distribution of nAChRs with different subunit
combinations in various brain regions contributes
to diverse mechanisms of plasticity. Mechanisms of
nicotine-induced plasticities include upregulation and
downregulation of nAChRs; interactions with catecholaminergic, GABAergic, and glutamatergic neurotransmission; and neurodegenerative structural changes in
fasciculus retroflexus and cortex. The extent to which
the changes induced by nicotine exposure are reversible
remains incompletely understood. There is need for
further investigation of the long-term sequalae of

Drug-Induced Neuroplasticity

nicotine exposure, particularly those of adolescent exposure. These mechanisms and relevant brain pathways
are summarized in Fig. 10. Nicotine addiction is of
particular concern because of the ready availability of
legally available tobacco products in many societies.
There is increasing evidence that early exposure to
nicotine leads to increased risk of later life dependence
not only on nicotine but also to other drugs of abuse
including alcohol. There is need for further investigation
of the mechanisms by which nicotine-induced plasticity
can increase risk of dependence to other drugs of abuse.
D. Neural Adaptations Induced by Ethanol
Ethanol is one of the most widely used psychoactive
substances worldwide, with both reinforcing and
sedative-hypnotic properties. Acutely, these effects
arise from many molecular targets, and they most
prominently involve alterations in synaptic function.
Repeated ethanol exposure leads to homeostatic
changes in synaptic and intracellular functions. Behaviorally, these neuroadaptations are expressed as tolerance, dependence, somatic withdrawal symptoms upon
cessation of chronic ethanol exposure, and so-called
protracted withdrawal characterized by susceptibility
to negative reinforcement that is postulated to contribute to relapse and compulsive features of ethanol
addiction. Figure 11 depicts the cycle of alcohol
intoxication/neural adaptation/withdrawal effects in
the brain, with proposed roles of the neurotransmitter/
modulator systems. In the following, we will focus
primarily on the synaptic effects of ethanol and their
persistent alterations in key brain regions during
chronic ethanol administration.
Adjunct pharmacological means to treat alcoholism
have been expanded from one specific drug, disulfiram,
to opioid receptor antagonists naltrexone and nalmefene and to glutamate receptor antagonists acamprosate (Heilig and Egli, 2006). Benzodiazepine therapy
during alcohol withdrawal is still important, especially
to protect the brain from potential neurotoxicity of
withdrawal-induced seizures (Sellers et al., 1983;
Walker et al., 2013). Recent neurobiological research
on ethanol addiction, relapse, and withdrawal has
produced many other potentially efficient drug targets,
which are mentioned below within a discussion on
mechanism of alcohol action and addiction.
1. Cell Membrane Ion Channels as Primary Targets
of Ethanol. Compared with many other drugs of abuse
that exert their actions through specific neurotransmitter
receptors or transporters, ethanol is a relatively nonspecific and nonpotent pharmacological agent. Ethanol
was believed to fluidize membrane lipids and cause
intoxication by perturbing neuronal function (for reviews,
see Deitrich et al., 1989; Franks and Lieb, 1994; Peoples
et al., 1996; Spanagel, 2009). Then, protein targets
emerged for anesthetics and for alcohols of various carbon
chain lengths (Franks and Lieb, 1984). A general binding

925

pocket for ethanol has not been settled, although in


several ligand-gated ion channels the action of alcohol is
dependent on specific single amino acid residues, as
modeled in a prokaryotic member of ligand-gated ion
channels, GLIC (Howard et al., 2011). Although many
drug targets have been revealed by using high-affinity
(nanomolar) radiolabel binding to brain membranes or
recombinant receptors, with ethanol this approach has
not been possible because of its 1000- or even 100,000-fold
lower affinity. However, there is accumulating evidence
that ethanol has a few well-defined molecular targets
(Trudell et al., 2014). At intoxicating ethanol concentrations, from low to approximately 100 mM, the most
important targets of ethanol include ion channels,
neurotransmitter receptors, and intracellular signaling
molecules. The best characterized of these targets are
ligand-gated ion channels (Lovinger and Roberto, 2013)
that bind extracellular neurotransmitters, which leads to
opening of an intrinsic ion pore.
In short, there are three major classes of ion channels
activated by neurotransmitters. The so-called cys-loop
channels are protein pentamers that have a Cys-Cys
bond in the N-terminal binding domain, four transmembrane domains forming the ion channel, and an
intracellular loop between TM3 and TM4 for modification by phosphorylation/dephosphorylation. This major
class of receptor channels passes either anions, such as
the g-aminobutyric acidA (GABAA) and strychninesensitive glycine receptors, or cations, such as the
nicotinic acetylcholine nACh and serotonin 5-HT3
receptors. The general action of acute ethanol is to
enhance the function of cys-loop ligand-gated ion channels (Lovinger, 1997; Harris, 1999; Perkins et al., 2010),
although ethanol can also produce inhibition in some
subtypes/receptor domains (Johnson et al., 2012). Ethanol exerts its action by potentiating channel opening in
the presence of low agonist concentrations. This is
achieved either by increasing the probability of channel
opening or increasing the agonist affinity (Tonner and
Miller, 1995; Zhou et al., 1998; Welsh et al., 2009).
The potentiation of the GABAA receptor by ethanol
has been investigated widely. The GABAA receptor is
assembled as variable combinations of 19 subunit
proteins (Collingridge et al., 2009), and many subunit
combinations have been characterized in heterologous
expression systems and in native receptor preparations,
including cultured and isolated neurons. In summary,
ethanol potentiates the function of various GABAA
receptor subtypes containing a, b, and g subunits, as
well as those containing a4 and a6 together with b and d
subunits (McCool et al., 2003; Olsen et al., 2007; Lobo
and Harris, 2008). In brain slices from the cerebellum,
HC, and thalamus, ethanol potentiation has also been
observed in extrasynaptic, d subunit-containing highaffinity GABAA receptors mediating a tonic inhibitory
current (Wei et al., 2004; Hanchar et al., 2005; Glykys
et al., 2007; Jia et al., 2008). However, the potentiation

926

Korpi et al.

Fig. 11. The effects of alcohol are shown with a distinction between acute and withdrawal phases, with alcohol drinking during them inducing positive
and negative reinforcement, respectively. Each state is associated with specific effects and symptoms mediated by specific neurobiological mechanisms.
The summary illustrates some of the most important alcohol targets and their adaptations during chronic alcohol exposure. Modified from Littleton
(1998).

has not been universally replicated in all neuron types


or experimental settings (Borghese et al., 2006; Korpi
et al., 2007; Lovinger and Homanics, 2007). Ethanolinduced potentiation of GABAA receptor function may
also depend on protein phosphorylation. For example,
the epsilon subunit of protein kinase C (PKC) appeared
to be necessary for ethanol potentiation of GABAA
receptors containing the g2 subunit (Qi et al., 2007),
whereas PKCd was required for ethanols action on the d
subunit-containing GABAA receptors (Choi et al., 2008).
It is noteworthy that ethanol potentiation of GABAergic
inhibition often results from presynaptic actions of
ethanol. Ethanol enhances GABA release from presynaptic terminals, which contributes to synaptic inhibition
(for reviews, see Siggins et al., 2005; Lovinger and
Roberto, 2013). These data are obtained mostly from
brain slices and isolated neurons from the cerebellum,
HC, VTA, and amygdala (Ariwodola et al., 2003; Roberto
et al., 2003; Kelm et al., 2007; Theile et al., 2008). The
mechanism of ethanol enhancement of GABA release is
not well established, but ethanol may interact with
mechanisms involved in intracellular calcium release,
increasing presynaptic calcium concentrations (Kelm
et al., 2007). The possible role of intracellular signaling
pathways in this ethanol effect is suggested by findings
that GABA release is diminished in cerebellar Purkinje
neurons by PKA inhibitors and in the CeA in mice lacking
PKC (Bajo et al., 2008; Kelm et al., 2008).
Ethanol potentiation of the strychnine-sensitive glycine receptor channels also depends on the receptor
subunit composition. Greatest potentiation is seen in a1

subunit-containing receptors in the Xenopus leavis


oocytes (Valenzuela et al., 1998b), whereas receptors
with the a2 subunit are less sensitive (Mascia et al.,
1996). Expression of the b subunit together with a2
abolishes ethanol potentiation (McCool et al., 2003). As
both the GABAA and glycine receptors are inhibitory,
ethanol enhancement of their function increases neuronal inhibition. The net effect on neural networks is
determined by the primary target, either interneurons
or principal neurons. Intracellular mechanisms for the
glycine receptor potentiation by ethanol have been
revealed, e.g., in isolated VTA neurons (Ye et al., 2001;
Zhu and Ye, 2005). The activated G protein bg-subunits
enhance the ethanol effect in recombinant glycine
receptors by binding to specific intracellular domains
[including two lysines at 385/386) of the a1 subunits
(Yevenes et al., 2003, 2008)]. Knock-in mice with the
LysLys385/386AlaAla mutation in the a1 subunits
exhibited reduced alcohol intoxication (shorter duration
of loss of righting reflex) and almost complete lack of
potentiation by 100 mM ethanol in isolated brain stem
neurons and cultured spinal cord neurons (Aguayo et al.,
2014). Presynaptic glycine receptors can reduce GABA
release in the VTA, which can activate DA neurons (Ye
et al., 2004). In addition to GABA release, ethanol also
augments glycine release via Ca2+-dependent mechanisms (Mariqueo et al., 2014). Strychnine blocks the
ethanol-induced DA release from the nucleus accumbens
and attenuates nicotine- and THC-induced release, but
not the release induced by cocaine and morphine (Jonsson
et al., 2014), indicating the importance of glycine

Drug-Induced Neuroplasticity

receptors in the modulation of mesolimbic DA pathway


activity.
Acute ethanol also potentiates the function of serotonin 5-HT3 receptors that gate an intrinsic cation
channel (Lovinger and White, 1991; Machu and Harris,
1994). As these receptors are expressed both pre- and
postsynaptically, the site of ethanol action remains to be
determined. The roles of the five different 5-HT3 receptor subunits (Thompson, 2013) are not well elucidated. Interestingly, in rat brain, the 5-HT3 receptors
have been localized primarily to cholecystokinincontaining (CCK) GABAergic interneurons (Morales
and Bloom, 1997). Furthermore, they are also expressed
in vagal afferents from the stomach, where they can be
activated by serotonin released from enterochromaffin
cells, which together with the area postrema 5-HT3
receptors participate in nausea and/or emesis reflexes
(Hornby, 2001).
Like the GABAA receptor, the nACh receptor is often
composed of multiple subunits (see section II.C on
nicotine), but homomeric receptors also exist. High
ethanol concentrations (.100 mM) potentiate the majority of nACh receptor subtypes without selectivity
(Narahashi et al., 1999; Zuo et al., 2002). However, the
effects of ethanol concentrations lower than 100 mM are
dependent on the a subunit, with potentiation of a2b4,
a4b4, a2b2, and a4b2 nACh receptors, with no effect on
a3b4 and a3b2 receptors, and inhibition of the a7
receptors. In chimeric recombinant homomeric receptors with the N-terminal ligand-binding domain from
the nACh a7 subunit and the rest from 5-HT3A subunits, nicotine responses were inhibited by ethanol
similarly to a7 receptors, suggesting that the ethanol
inhibition was dependent on the N-terminal ligand
binding domain (Yu et al., 1996).
The second major class of ligand-gated ion channels
consists of the ionotropic glutamate receptors (iGluRs)
that have three major subclasses, the AMPARs, the
NMDARs, and the kainate receptors (Collingridge et al.,
2009). All these subclasses are cation permeable, but
differ in selectivity between Na+, K+, and Ca2+. These
ligand-gated receptor channels are homo- or heterotetramers, with subunits having large N-terminal ligandbinding domains, three TM regions, and a fourth re-entrant
loop as in potassium channels. Although ethanol partially
inhibits all iGluRs, inhibition of NMDARs with intoxicating ethanol concentrations has been most extensively
investigated (Dildy and Leslie, 1989; Hoffman et al., 1989;
Lovinger et al., 1989; Kuner et al., 1993; Moykkynen and
Korpi, 2012). Functional NMDARs are heteromers, containing two glycine-binding GluN1 subunits combined
with two glutamate-binding GluN2 subunits (A-D) or with
two GluN3 subunits. The GluN3 subunit has been poorly
characterized in native receptors, but it may be important
for synaptic plasticity especially during development
(Ulbrich and Isacoff, 2008; Pachernegg et al., 2012; Kehoe
et al., 2013). Ethanol inhibits all NMDAR subtypes, but in

927

recombinant receptors the sensitivity to inhibition


depends on the subunit composition. Ethanol slightly
more efficiently inhibits the GluN1/2A- or GluN1/2Bcontaining receptors than receptors with GluN1/2C subunit composition (Kuner et al., 1993; Masood et al., 1994;
Chu et al., 1995). There are several splice variants of the
GluN1 subunit, and the splicing status of this subunit
may affect ethanol sensitivity, depending on the identity
of GluN2 subunit in the receptor (Jin and Woodward,
2006). Among AMPARs (composed of GluA1-4 subunits),
non-GluA2 receptors are calcium permeable (Burnashev
et al., 1992). Although ethanol also inhibits the AMPARs,
higher concentrations are needed than for the inhibition of
the NMDARs, and their subunit composition does not play
a major role in determining ethanol sensitivity (Lovinger,
1993; Frye and Fincher, 2000). Ethanol produces the
inhibition by promoting desensitization of the AMPARs
(Moykkynen et al., 2003). The kainate receptors (composed of GluK1-5) are inhibited by ethanol at low concentrations, but it is unclear whether this action involves
direct effects on the receptor protein (Valenzuela et al.,
1998a; Lack et al., 2008). The GluK1 receptors are
especially interesting, because the presence of a polymorphism in GRIK1 gene correlated with a robust response to
topiramate in reducing heavy drinking of alcoholdependent subjects (Kranzler et al., 2014). Taken together, inhibition of the iGluRs by ethanol decreases
neuronal excitability and also inhibits synaptic plasticity
that requires activation of iGluRs.
The purinergic trimeric P2X receptors constitute the
third major class of ligand-gated ion channels. At intoxicating concentrations, ethanol has inhibitory effects on
many P2X subtypes, with the P2X4 receptors being the
most sensitive subtype to inhibition (Li et al., 1993;
Davies et al., 2002, 2005). Interestingly, inhibition of
the ATP action on native P2X receptors or recombinant
rat P2X4 receptors by series of alcohols provides evidence
for a cut-off point at butanol, allowing smaller alcohols to
inhibit presumably via binding to a binding pocket of
a restricted size (Li et al., 1994; Asatryan et al., 2010). The
anthelmintic agent ivermectin has a positive allosteric
action at this subtype, among several other ligand-gated
ion channels, such as GABAA receptors (Zemkova et al.,
2014). Ivermectin reduces alcohol drinking in mouse
models (Yardley et al., 2012, 2014). Further functional
significance of ethanol actions on P2X1/3 receptors comes
from findings that ethanol reduces GABA release in
isolated VTA DA neurons via inhibition of ATP effects
on P2X receptors (Xiao et al., 2008).
Also the nonligand-gated ion channels are targets
of ethanol. Ethanol inhibits high-voltage-activated
dihydropyridine-sensitive Ca 2+ channels (L-type,
Cav1.x) (Wang et al., 1994) and potentiates the function
of G protein-activated inwardly rectifying K+ channels
(GIRKs, Kir3.x) (Kobayashi et al., 1999; Lewohl et al.,
1999), which both inhibit neural activity. The
large conductance Ca2+- and voltage-activated K+ (BK)

928

Korpi et al.

channel has emerged as a most interesting molecular


target for ethanol, based on genetic studies on Caenorhabditis elegans and Drosophila, and on recent results
from mouse knockout and human genetic association
studies (Treistman and Martin, 2009; Bettinger and
Davies, 2014; Dopico et al., 2014). BK channels are also
involved in neuroadaptation to alcohol. They are formed
as tetrameric complexes from two a subunits and two b
subunits (usually either b1 or b4), and they are
expressed throughout the body; in the brain they are
present in all areas with presynaptic and extrasynaptic
compartments. Ethanol usually activates these channels
in neurons, which should produce neuronal inhibition
and reduction in neurotransmitter release depending on
the b subunits. Tolerance to ethanol appears in minutes
as reduced channel sensitivity, followed by declustering
and internalization of the channels. The b1 subunits
make the BK channels insensitive to ethanol, while b4
subunits provide ethanol sensitivity (Feinberg-Zadek
et al., 2008). This seems to correlate in mouse models
with quicker escalation of limited access alcohol drinking
after induction of dependence by chronic intermittent
ethanol inhalation in the absence of the b1 subunit
(Kreifeldt et al., 2013) and quicker tolerance development
to motor impairment in the absence of the b4 subunit
(Martin et al., 2008). The search for the site of action of
ethanol and other n-alcohols on BK channel-forming a
subunits is very actively pursued (Bukiya et al., 2014;
Davis et al., 2014). The discovery of this site of action
might allow the rational development of an alcohol
antagonist for BK channels. In addition, establishing
more detailed roles of the accessory b subunits in the
regulation of channel function by phosphorylation (Liu
et al., 2006) might open novel ideas for translational
experiments.
Ethanol may also affect the function of the GPCRs,
but generally these effects by acute ethanol are weak
and the mechanisms are unclear, unlike those of opioids
and cannabinoids acting on their respective GPCRs.
There is early evidence that ethanol can stimulate
cAMP production, perhaps by activating ACs, which
can underlie ethanol-induced neurotransmitter release
(Luthin and Tabakoff, 1984; see above for discussion on
GABA release).
2. Ethanol-Induced Changes in Glutamatergic
Transmission. Repetitive patterned activation of afferent inputs to postsynaptic neurons induces longlasting changes in the efficacy of synaptic transmission.
The most common forms of these changes, LTP and
LTD, have also been examined after acute ethanol
(McCool, 2011). Consistent with the inhibition of
NMDARs at relevant ethanol concentrations, acute
ethanol inhibits induction of LTP at various synapses
and experimental preparations, such as Schaffer collateral inputs to the CA1 pyramidal neurons in the HC
(Zorumski et al., 2014). Acute ethanol also modulates
types of LTD in many brain regions, including LTD in

the HC involving activation of mGluRs (Overstreet


et al., 1997) and the eCB-mediated signaling in the
dorsal striatum (Clarke and Adermark, 2010). Ethanol
inhibition of these types of plasticity is consistent with
acute ethanol-induced impairment of working and
short-term memory (Schweizer and Vogel-Sprott, 2008).
Considering the development of alcoholism in humans,
however, the most relevant neural changes recapitulated
by animal models probably include those produced by
chronic ethanol exposure. Chronic ethanol produces
tolerance, manifested by decreased behavioral response
to ethanol, and dependence, characterized by a specific
symptomology after ethanol withdrawal. In addition,
chronic alcohol exposure can also produce a postdependent state, which is understood as a combination of
behavioral and neuroadaptive consequences that are
produced as the individual becomes dependent and which
remain for extended periods even in the absence of the
drug-specific withdrawal symptoms. Typical behavioral
correlates of the postdependent state include anxiety,
dysphoria, and exaggerated stress responses (Heilig and
Koob, 2007).
Various experimental paradigms have been used for
inducing alcohol dependence, including repeated
intraperitoneal or intragastric administration, ethanolcontaining liquid diets, and either continuous or intermittent ethanol vapor chamber exposure. These
paradigms typically produce peak blood ethanol concentrations of 3569 mM (150300 mg/dl), which cannot
be attained by most voluntary ethanol drinking models.
However, alcohol-preferring mice consume enough ethanol to achieve blood ethanol levels exceeding 23 mM
(100 mg/dl) in the recently described binge-drinking
model (drinking in the dark) (Rhodes et al., 2005), and
the reintroduced intermittent ethanol access model
results in escalation of ethanol drinking in rats, with
intoxicating blood ethanol levels (Wise, 1973; Simms
et al., 2008). Even if induction of alcohol dependence in
these binge models is debatable, long-term binge drinking has been shown to produce a variety of neuroadapations (Stuber et al., 2008; Wilcox et al., 2014).
Chronic ethanol increases NMDAR function at both
synaptic and extrasynaptic compartments in many
brain regions and experimental preparations. Enhanced NMDAR function is seen as an increased intracellular level of NMDAR agonist-induced Ca2+ or
directly as increased ion current through the NMDAR
pore (Gulya et al., 1991; Iorio et al., 1992; Smothers
et al., 1997; Grover et al., 1998; Floyd et al., 2003).
However, the ability of acute ethanol to inhibit
NMDARs is not diminished by chronic ethanol treatment (Roberto et al., 2004b). The mechanism underlying the enhanced NMDAR function is not well known,
but it may involve increases in the levels of the GluN2B
subunit (Floyd et al., 2003; Carpenter-Hyland et al.,
2004; Kash et al., 2009). In the VTA, chronic ethanol has
also been associated with a higher order plasticity,

Drug-Induced Neuroplasticity

described as metaplasticity, that affects synapses of


postsynaptic neurons in a global manner, increasing
their susceptibility to undergo activity-dependent LTP/
LTD (Abraham, 2008). Thus, chronic ethanol enhances
the susceptibility of VTA DA neurons to the induction of
LTP of NMDAR-mediated transmission. This plasticity
results from an increase in the potency of inositol 1,4,5trisphosphate (IP3) in producing amplification of action
potential-evoked Ca2+ signals, lasting for 1 week after
ethanol exposure (Bernier et al., 2011). This kind of
metaplasticity could be an important neuroadaptation
that drives the formation of drug-associated cues both
during drug intoxication and withdrawal.
Acute doses of many drugs of abuse, including
ethanol, increase the AMPAR/NMDAR current ratio
in VTA DA neurons (Saal et al., 2003), reflecting an
enhancement of AMPAR-mediated excitatory synaptic
transmission. The findings of the chronic ethanol effects
on AMPARs are sparse and variable, suggesting
increases in AMPAR expression in the HC (Bruckner
et al., 1997) and increased AMPAR function, as shown
by enhanced AMPAR-mediated synaptic responses in
the BLA (Lack et al., 2007). In the VTA neurons,
voluntary intermittent ethanol consumption enhances
postsynaptic AMPAR function without altering presynaptic glutamate release (Stuber et al., 2008). The
enhanced AMPAR function in the BLA could contribute
to drug seeking, whereas in the VTA, AMPAR activity
could regulate firing of the mesolimbic DA neurons and
thereby contribute to the reinforcing and activating
effects of ethanol.
In the NAc, another key component of the reward
circuitry, glutamatergic innervation, interacts with the
mesolimbic DA system and participates in drug reinforcement, drug seeking, and relapse. The Homer
proteins are integral components of the PSD that
appear to be necessary for alcohol-induced neuroplasticity within the NAc. For example, chronic ethanol
consumption under continuous access produces a robust
increase in the NAc Homer2 protein levels, which was
observable even 2 months after withdrawal from alcohol
drinking (Szumlinski et al., 2008). The increased
Homer2 expression was accompanied by elevation in
total mGlu1 receptor and GluN2B subunit levels during
the first 2 weeks after withdrawal. Furthermore, both
virus-mediated knockdown of Homer2 expression and
intra-NAc antagonism of mGlu5 and PI3K attenuated
ethanol intake, suggesting that upregulation of NAc
mGlu5-Homer2-PI3K signaling constitutes an important neuroadaptation to chronic ethanol (Cozzoli et al.,
2009).
Chronic ethanol produces elevated extracellular glutamate levels in the brain, especially during withdrawal
and repeated cycles of withdrawal, thus contributing to
the hyperglutamatergic state implicated in the development of ethanol dependence. These findings have
been derived from both in vivo microdialysis and

929

magnetic resonance spectroscopy studies (Rossetti and


Carboni, 1995; Hermann et al., 2012), and they do not
directly indicate whether the source of glutamate is
synaptic, extrasynaptic, glial, or metabolic. However,
there is evidence of increased synaptic glutamate release from the CeA after chronic ethanol treatment,
pointing to adaptations in the presynaptic terminal
function (Roberto et al., 2004b).
3. Ethanol-Induced Changes in g-Aminobutyric Acidergic Transmission. GABAergic neurotransmission is
one of the primary targets of ethanol and is hypothesized to be involved in multiple long-lasting neuroadaptations underlying ethanol tolerance, dependence,
and reinforcement. Many early studies focused on
elucidating changes in the expression profiles of specific
GABAA receptor subunits by chronic ethanol, with the
hypothesis that these changes could account for alterations in the functional properties of GABAergic signaling after chronic ethanol exposure (Kumar et al.,
2009). Generally, these studies failed to show dramatic
changes in the number of GABAA receptors across the
preclinical models, but pointed to a number of regionspecific and temporally dependent alterations in
GABAA receptor subunit mRNA and peptide levels
(reviewed in Uusi-Oukari and Korpi, 2010). For example, in the cerebral cortex, chronic ethanol consumption
elicited an increase in a4 subunit mRNA levels and
a decrease in a1 subunit mRNA levels. There was also
an increase in g2S, but not g2L, subunit mRNA levels
after chronic ethanol consumption. In addition, g1
subunit expression was increased, whereas a5, b1, b2,
b3, g3, and d subunit mRNA levels did not change
(Devaud et al., 1995). These changes correlated well
with the receptor subunit peptide levels in ethanoldependent rats, but not necessarily with subunit expression during withdrawal (Devaud et al., 1997). However,
other brain regions did not exhibit similar patterns of
subunit changes, and subregions of the cerebral cortex
differed from the whole cortex (Grobin et al., 2000). Direct
comparison of cerebral cortex and HC revealed that also
in the HC the a4 subunit peptide level was increased,
but a1, a2, a3, b2/3, or g2 subunits were not altered,
suggesting that chronic ethanol changes GABAA receptor
gene expression in the HC but differently from that in the
cerebral cortex (Matthews et al., 1998). Moreover, the
changes were dependent on the length of ethanol exposure. Recent data from human postmortem brain areas
from controls and alcoholics have also revealed brain
region-specific changes in GABAA and glutamate receptor subunit expression in alcoholics (Jin et al., 2011;
Bhandage et al., 2014; Jin et al., 2014a,b).
The hypothesis derived from the adaptive changes in
the GABAA receptor expression by chronic ethanol is that
those adaptations lead to a hypofunction of GABAergic
synaptic functions, expressed as tolerance and increased
neuronal excitability during ethanol withdrawal, which
forms the basis for the treatment of severe alcohol

930

Korpi et al.

withdrawal with benzodiazepines (Sellers et al., 1983).


There is evidence that ethanol-induced alterations in the
expression of GABAA receptor subunits leads to diminished GABAA-mediated transmission. For example,
chronic intermittent ethanol exposure induces a decrease
in a1 and d subunits and an increase in a4 subunit in rat
HC. These changes were reflected in recordings made in
hippocampal slices from ethanol-exposed animals. Thus,
the decay time of GABAA-mediated miniature IPSCs was
decreased and potentiation of these current by positive
modulators was reduced by chronic ethanol, except for
the a4-preferring benzodiazepine ligands, with which
potentiation was either maintained or even increased
(Cagetti et al., 2003). Synaptic and extrasynaptic components of GABAA receptor activation were investigated in
the same animal model and hippocampal preparation.
After chronic ethanol, potentiation of tonic current by the
a1 subunit-preferring agonist zolpidem was lost, whereas
potentiation by the a4 subunit-preferring THIP was only
partially reduced. Also the potentiation of synaptic
GABAA receptor currents by zolpidem was eliminated
by chronic ethanol, whereas THIP enhanced them after
ethanol exposure. These data are consistent with a1
subunit decreases in both synaptic and extrasynaptic
compartments, and with increased a4 subunits in synaptic and decreased a4 subunits in extrasynaptic GABAA
receptors (Liang et al., 2004). The cause for these changes
is not clear, but one possibility is that the putative
primary reduction in a1 expression leaves the g2 subunit
(that is needed for synaptic targeting) available to
assemble with a4 subunits, which increases the synaptic
a4 subunit-containing receptors. In these receptors, the
inverse agonist of the benzodiazepine-site Ro 15-4513
acts as an agonist [like in a6 subunit-containing receptors
of the cerebellar granule neurons (Malminiemi and
Korpi, 1989)] and enhances the synaptic responses only
in the ethanol-dependent rats (Liang et al., 2006).
The CeA, a major component of the extended amygdala, has emerged as an important brain area for both the
acute positive reinforcement of drugs of abuse and
negative reinforcement associated with protracted abstinence. Ethanol drinking in nondependent rats is
inhibited by antagonism of GABAA receptors in the CeA
(Hyyti and Koob, 1995), whereas GABAA activation
modulates ethanol reinforcement only in animals after
chronic ethanol exposure (Roberts et al., 1996), suggesting altered CeA GABAergic transmission by chronic
ethanol. Acute ethanol augments GABAA-mediated inhibition in the CeA, which can be at least partially
mediated by ethanol-induced presynaptic GABA release
(Roberto et al., 2003). In CeA slices from rats treated
chronically with ethanol, the baseline evoked IPSPs and
IPSC amplitudes were increased and the paired-pulse
facilitation ratios were lower than in naive rats, consistent with increased GABAergic transmission involving
postsynaptic effects after chronic ethanol treatment.
Ethanol administered in vitro enhanced IPSPs and

IPSCs similarly in slices from ethanol-naive and


ethanol-dependent rats, suggesting that tolerance did
not develop to this acute effect of ethanol. The frequency
of miniature IPSCs was higher in chronically ethanoltreated rats, in line with the idea that enhanced
GABAergic transmission both by acute and chronic
ethanol could also be ascribed to a presynaptic mechanism involving vesicular GABA release. Verification with
microdialysis showed a fourfold increase in baseline
GABA release in ethanol-treated rats compared with
controls (Roberto et al., 2004a), although this method
cannot necessarily distinguish between neuronal and
glial GABA release. Significant GABA release seems to
take place from glial cells (astrocytes) that may have
a GABAergic phenotype (Lee et al., 2011) and release
GABA via bestrophin 1 anion channels (Lee et al., 2010).
These data show that both acute and chronic ethanol
effects on GABAergic transmission could involve pre- and
postsynaptic alterations. More recently, ethanol-induced
increase in firing rates of subpopulations of CeA neurons
has been shown to be mediated by CRF1 receptors and
tonic GABA currents, with the network effect being
increased output neuron activity to the BNST (Herman
et al., 2013), a brain region involved in sustained anxiety
and relapse. Furthermore, the ethanol-induced increase
in GABA release and IPSC amplitudes in the CeA slices
can be blocked by activation of presynaptic CB1 receptors
(Roberto et al., 2010a), demonstrating multiple presynaptic mechanisms being involved in ethanol modulation
of amygdala activity.
Further evidence for involvement of presynaptic
mechanisms in ethanol effects on GABAergic transmission has been obtained from the VTA. A single
ethanol dose induced a long-lasting facilitation of GABA
transmission, as suggested by paired-pulse depression
in ethanol-treated animals compared with paired-pulse
facilitation in controls (Melis et al., 2002). Also, as the
frequency of spontaneous miniature IPSCs was increased, a presynaptic mechanism can be inferred.
Indeed, a GABAB antagonist shifted paired-pulse depression to paired-pulse facilitation, indicating presynaptic GABAB receptor activation by hypothesized
GABA spillover by ethanol treatment. Similarly in the
VTA, a single ethanol exposure increased GABA release
onto DA neurons (Wanat et al., 2009). The importance of
these cellular, possibly long-lasting signaling events for
ethanol consumption and development of addiction
remains to be investigated. In contrast to GABA
neurotransmission in the amygdala, chronic alcohol
treatment makes the VTA GABA system hyposensitive,
which may be based on persistent changes in gene
expression (Arora et al., 2013).
Dorsal striatum and associated sensorimotor structures have been implicated in the development of addiction, particularly in the mediation of habit formation, and
this region has also emerged as a site of alcohol-induced
neuroadaptations. In cynomolgus monkeys, prolonged

Drug-Induced Neuroplasticity

intermittent alcohol drinking decreased the frequency of


GABAA receptor-mediated miniature IPSCs recorded
from the MSNs in the caudoventral area of the putamen
(Cuzon Carlson et al., 2011). This area corresponds to the
dorsolateral striatum in rodents, and a similar reduction
of IPSC frequency was produced by repeated binge
drinking in mice in this striatal area, but also in the
dorsomedial striatum (Wilcox et al., 2014). These data
suggest that heavy, long-term alcohol exposure causes
a depression of GABAergic transmission in the dorsal
striatum, contributing to an increased output from this
structure.
4. Ethanol-Induced Changes in Neuropeptide
Mechanisms. Various neuropeptide systems have
been hypothesized to be recruited or downregulated
during chronic ethanol exposure. Particularly in the
extended amygdala, neuropeptides may underlie the
negative emotional states that contribute to compulsive
ethanol seeking and drinking via negative reinforcement (Koob, 2008).
CRF plays a role in coordinating hormonal, autonomic, and behavioral stress responses in the body.
In ethanol-dependent animals, ethanol withdrawal
increases CFR levels in the CeA and BNST (Merlo Pich
et al., 1995; Olive et al., 2002). Administered into the
CeA, a CRF antagonist attenuated ethanol withdrawalrelated anxiety (Rassnick et al., 1993). Increased ethanol consumption during protracted abstinence after
induction of ethanol dependence was reduced by selective CRF1 receptor antagonism, but in nondependent
animals CRF antagonist failed to modulate ethanol
intake, suggesting that CRF contributes to increased
ethanol consumption induced by dependence (Funk
et al., 2007). CRF1 and CRF2 receptors mediate their
most physiologic functions in the brain via coupling
to heterotrimeric Gas-proteins and activating AC
(Grammatopoulos, 2012). Increased cAMP levels then
turn on a plethora of cytoplasmic (e.g., activation of
PKA) and nuclear (e.g., nuclear factor k-light-chainenhancer of activated B cells (NFkB), ERK1/2, and
GSK-3B/Wnt signaling) functions. Also, activation of the
PLC-PKC pathway has been implicated in the dose dependently increasing effects of CRF on VTA DA neuron
firing (Wanat et al., 2008). In midbrain slices from bingedrinking mice, CRF-potentiation of NMDAR responses
was found in VTA DA neurons, and the binge drinking
was significantly reduced by intra-VTA infusion of CRF1
receptor antagonist CP-154526 [N-butyl-N-ethyl-2,5dimethyl-7-(2,4,6-trimethylphenyl)-7H-pyrrolo[2,3-d]
pyrimidin-4-amine hydrochloride] (Sparta et al., 2013).
In the CeA, ethanol-induced enhancement of GABAergic
transmission is mediated by CRF1 receptors, perhaps
even by CRF release, as shown by lack of ethanolstimulated GABA release in CRF1 receptor-KO mice
and after CRF1 but not CRF2 receptor antagonists (Nie
et al., 2004). Moreover, CFR1-mediated ethanol effects
on GABA in the CeA involve activation of the PKC

931

pathway (Bajo et al., 2008). In CeA slices of ethanoldependent rats, the sensitivity of GABA release on
CRF and CFR1 antagonists was increased, and in vivo
intra-CeA administration of a CRF1 antagonist reversed dependence-related elevations in baseline extracellular GABA and blocked ethanol-induced increases
in GABA in both dependent and nondependent rats.
Chronic treatment with a CRF1 antagonist blocked
withdrawal-induced increase in alcohol drinking in
ethanol-dependent rats (Roberto et al., 2010b). Collectively, these findings suggest a specific chronic ethanolinduced synaptic neuroadaptation in the interaction of
CRF and GABAergic transmission that could contribute
to the development and maintenance of ethanol dependence, together with other described functional
alterations in the CeA (Hansson et al., 2006; Sommer
et al., 2008). Also the BNST has a rich expression of
CRF, and the region is well connected with many brain
regions, including the midbrain DA neurons and the
CeA. In the juxtacapsular nucleus of the anterior BNST,
a form of LTP in response to HFS of the stria terminalis
was impaired during withdrawal from chronic ethanol,
and this impairment was reversed by CRF1 receptor
antagonism (Francesconi et al., 2009). Therefore, also
the BNST CRF system may undergo persistent changes
during development of ethanol dependence. Novel
modulators have been found to play a role in reinstatement of alcohol drinking, because blockade of the
relaxin-3 receptor RXFP3 by intra-BNST infusion of
an antagonist attenuated cue- and stress-induced reinstatement of alcohol drinking in rats (Ryan et al.,
2013). Importantly, the neurons in the brain stemlocated nucleus incertus, which is the major source of
relaxin-3 in the mammalian brain, are highly sensitive
to stress and CRF (Banerjee et al., 2010; Smith et al.,
2011; Farooq et al., 2013).
Another relevant peptide neurotransmitter system
uses orexin A/hypocretin-1 and orexin B/hypocretin-2
peptides that are synthesized selectively in the lateral
hypothalamus, perifornical area, and dorsomedial hypothalamus (de Lecea et al., 1998; Sakurai et al., 1998),
with the lateral hypothalamus orexin system being
most strongly associated with rewarding behaviors
(Harris et al., 2005, 2007). In the VTA, both orexins
and CRF have similar excitatory effects on DA neurons
(Borgland et al., 2010). Orexin effects are mediated by
OX1 and OX2 receptors, which mainly couple to Gq/11
and PLC-PKC pathways. Both receptor subtypes appear to be expressed in various neuronal populations of
the VTA, and both DA and GABA neurons increased
their firing rates by orexins (Korotkova et al., 2003). The
OX1 receptor is expressed on VTA DA neurons. Its
stimulation induces increased membrane targeting of
the NMDARs, thereby potentiating their currents
(Borgland et al., 2006). Importantly, initial findings on
the effects of OX1 receptor antagonists suggest that they
do reduce alcohol drinking in some, but not all,

932

Korpi et al.

paradigms in rodent models (Lawrence et al., 2006;


Moorman and Aston-Jones, 2009; Dhaher et al., 2010;
Jupp et al., 2011; Voorhees and Cunningham, 2011).
Further work is needed to establish the ways by which
orexin antagonists could be useful in the treatment of
alcohol abuse and other drug addictions. The OX1 and
OX2 receptor antagonists have very strong effects on
arousal and they thus might be emerging as the future
generation of hypnotic drugs for insomnia (Michelson
et al., 2014), provided they pass safety screening, such
as possible induction of narcolepsy, which is related to
deficiency of orexin and its receptors (Thannickal et al.,
2000).
Other neuropeptides that have been shown to modulate the interaction of ethanol with GABAergic neurotransmission in the CeA include neuropeptide Y (NPY)
and nociceptin/orphanin FQ (N/OFQ). NPY mediates
many behavioral responses to ethanol (Thiele et al.,
1998), and NPY injections into the CeA suppress
ethanol drinking, particularly in ethanol-dependent
animals (Gilpin, 2012). Superfusion of NPY decreased
baseline CeA GABAergic transmission and reversed
ethanol-induced enhancement of inhibitory transmission by suppression of GABA release through presynaptic Gi/o-coupled Y2 receptors (Gilpin et al., 2011). NPY
is anxiolytic and may mediate acute ethanol-induced
anxiolysis. By using histone deacetylase inhibitors,
Sakharkar et al. (2012) showed in alcohol-preferring P
rats that a rapid tolerance is produced to this anxiolytic
effect during chronic ethanol treatment via possible
epigenetic changes causing a reduction in NPY expression. It is quite possible that these epigenetic changes
are far more frequent and affect the transcription of
many genes important for plasticity (Nestler, 2014) and
that epigenetic mechanisms can also contribute to
genetic vulnerability to increased alcohol consumption,
at least in animal models (Moonat et al., 2013).
Also the effects of the peptide nociception/orphanin
FQ (N/OFQ) on GABAergic transmission are altered by
chronic ethanol. N/OFQ decreases CeA GABAergic
transmission by reducing GABA release, and this
N/OFQ-induced decrease was more pronounced in
ethanol-dependent rats (Roberto and Siggins, 2006).
N/OFQ is the endogenous ligand of the Gi/o-coupled
N/OFQ (NOP) receptor, also referred to as the opioid
receptor-like 1 receptor. Because N/OFQ has been
shown to attenuate many ethanol-related behaviors,
including cue- and footshock stress-induced ethanol
seeking and anxiety during ethanol withdrawal (Witkin
et al., 2014), there has been a growing interest to test
brain penetrating NOP receptor agonists for their effects
on ethanol drinking. Recently, a novel NOP agonist, MT7716 ((R)-2-{3-[1-(acenaphthen-1-yl)piperidin-4-yl]-2-oxo2,3-dihydro-1H-benzimidazol-1-yl}-N-methylacetamide
hydrochloride hydrate), was shown to reduce ethanol
drinking persistently even for 1 week after discontinuation of the drug and to attenuate ethanol somatic

withdrawal symptoms (Ciccocioppo et al., 2014), presumably by inhibiting presynaptically GABA and glutamate release, as shown in the amygdala slices
(Kallupi et al., 2014a,b). Furthermore, presumably via
presynaptic mechanisms, NOP activation counteracts
the CRF-induced increase in GABAA receptor-mediated
IPSPs in the CeA slices from naive rats and even more
robustly in ethanol-dependent rats (Cruz et al., 2012),
suggesting that dependent animals have developed sensitized responses to NOP activation in the
amygdala.
A large body of evidence suggests that the endogenous
opioid system is involved in the behavioral actions of ethanol. Acute ethanol stimulates the release of the endogenous opioid peptides b-endorphin, enkephalin, and
dynorphin in the brain (Gianoulakis, 2001), and m- and
d-opioid receptor antagonists decrease ethanol SA
and seeking in many experimental models (Koob
et al., 2003), suggesting that the Gi/o-coupled, mainly
presynaptic, m- and d- receptors are involved in the
mediation of the positive reinforcing effects of ethanol.
These results have been translated into the treatment
of alcoholism by using nonselective opioid receptor
antagonists, naltrexone or nalmefene, to block alcohol
reward either chronically or on demand (Volpicelli et al.,
1992; Hyytia and Sinclair, 1993; Mann et al., 2013).
Dynorphin, an endogeneous ligand of the Gi/o-coupled
k-receptors is linked to aversive states, and accumulating data implicate the dynorphin/k-receptor system in
the processes of negative reinforcement in ethanoldependent subjects (Sirohi et al., 2012). In animals
chronically exposed to ethanol, both the predynorphin
mRNA and dynorphin B expression was increased in
the NAc (Przewlocka et al., 1997; Lindholm et al., 2000),
suggesting an upregulation of the k-receptor/dynorphin
system during ethanol dependence, although the
k-receptor mRNA level was decreased (Rosin et al.,
1999). The fact that either intracerebroventricular,
intra-accumbal, or intra-amygdalar administration of
k-receptor antagonists selectively suppressed ethanol
SA in dependent animals with no effect in nondependent subjects (Walker and Koob, 2008; Nealey et al.,
2011; Kissler et al., 2014) suggests that the upregulated
k-receptor/dynorphin system contributed to escalation
of ethanol intake after dependence induction. The functional consequences of the alteration in the k-receptor/
dynorphin system are not known, but increased signaling through the k-receptors was shown to reduce DA
release locally in the NAc (Spanagel et al., 1992) and in
the PFC terminal area via inhibition of VTA DA
neurons (Margolis et al., 2006), possibly leading to
dysphoric and depression-like behavioral states. The
transcriptional processes underlying these events
may include the CREB that mediates the increases in
the dynorphin levels in the mesolimbic DA system
(reviewed in Sirohi et al., 2012). One of the important
targets of CREB is the BDNF that modulates dynorphin

Drug-Induced Neuroplasticity

expression (Lonze and Ginty, 2002). Ethanol exposure


increases BDNF expression in the dorsal striatum,
which could lead to initiation of MAP kinase signaling
and a subsequent increase in the downstream gene
products, including preprodynorphin (McGough et al.,
2004; Logrip et al., 2008). This could be one of the
mechanisms contributing to the hypothesized dysregulation of the dynorphin signaling linked with ethanol
dependence (Walker and Koob, 2008).
Ghrelin is a circulating gut-brain hormone that
promotes food consumption through hypothalamic
growth hormone secretagogue receptors (GHS-R1A)
(Horvath et al., 2001). In addition to the role of ghrelin
in body weight regulation, it has also been suggested to
be involved in mediation of the reinforcing effects of
drugs, including alcohol (Jerlhag et al., 2009). For
example, GHS-R1A antagonists suppress alcohol intake
in many animal models of alcohol SA, and a recent study
demonstrated that no tolerance developed to repeated
antagonist administration (Suchankova et al., 2013).
The GHS-R1A receptors are also widely expressed in
the mesocorticolimbic DA system, and the VTA expression of Ghsr was downregulated in the rats drinking
high amounts of alcohol compared with the lowdrinking individuals (Suchankova et al., 2013). Similar
to ghrelin, other gut-brain hormones have also been
implicated in the regulation alcohol drinking, including
glucagon-like peptide 1, leptin, and galanin (Blednov
et al., 2004; Lewis et al., 2004; Egecioglu et al., 2013).
Also histamine has been as a gut-based mediator of
smooth muscle function and a central neurotransmitter
that participates in regulation of food intake (among
a wide variety of important CNS roles), and recently it
has been shown to modulate alcohol consumption and
alcohol-induced place preference via H3R histamine
receptors (Panula and Nuutinen, 2013).
5. Role of Acetaldehyde in Ethanols Reinforcing
Actions. Although it is widely believed that the actions
of the ethanol molecule itself are responsible for ethanol
reinforcement, there is evidence to suggest that ethanol
metabolism and its products, particularly acetaldehyde,
could have a role in ethanol consumption and abuse
(reviewed in Quertemont, 2004; Deng and Deitrich,
2008; Deehan et al., 2013). The bulk of ethanol is
metabolized in the liver by alcohol dehydrogenase and
aldehyde dehydrogenase (ALDH) enzymes. Alcohol dehydrogenase is not physiologically active in the brain.
However, the first metabolic product, acetaldehyde, can
also be formed in the brain through the peroxidative
activity of catalase and by oxidation via other enzymes,
such as the cytochrome P450 2E1. Significant acetaldehyde concentrations are formed in vitro in brain tissue
at concentrations of ethanol achieved from voluntary
consumption by rodents.
Although there are many demonstrations of the
reinforcing actions of acetaldehyde in the brain, either
by experimenter- or self-administered acetaldehyde

933

(Deehan et al., 2013), the mechanisms of action remain


to be clarified. Converging evidence suggests that
acetaldehyde could contribute to ethanols stimulatory
action on the VTA DA neurons. Acetaldehyde stimulates VTA DAergic activity at remarkably lower concentrations than ethanol in vitro (Brodie and Appel,
1998; Diana et al., 2008), and in vivo a catalase inhibitor
applied in the VTA prevents ethanol-stimulated increase in DAergic activity (Melis et al., 2007; Diana
et al., 2008). Further support for the involvement of
catalase and acetaldehyde in ethanols central actions
comes from studies using viral vectors altering catalase
activity. For example, introduction of an anticatalase
viral vector into the VTA decreased ethanol drinking
and ethanol-induced DA release in the NAc (Karahanian
et al., 2011; Quintanilla et al., 2012). However, the DAreleasing effect of ethanol in the NAc is not mediated by
acetaldehyde (Clarke et al., 2014). It is also to be noted
that acetaldehyde is a highly reactive molecule that
interacts with many endogenous substances in the brain
to form additional biologically active products (Deehan
et al., 2013).
Acetaldehydes putative central reinforcing actions
notwithstanding, peripheral acetaldehyde is aversive.
This is amply demonstrated by the ALDH blocking drug
disulfiram (Antabuse, Odyssey Pharmaceuticals, East
Hanover, NJ) that causes increased accumulation of
acetaldehyde and thereby highly aversive effects
shortly after ethanol consumption, thus curbing further
drinking. In some Asian populations, the ALDH2*2
allele that encodes an inactive form of the mitochondrial
ALDH2 enzyme, has a high prevalence. Individuals
homozygotic for ALDH2*2 exhibit high blood acetaldehyde concentrations even after light to moderate alcohol
drinking, and it has been shown that this allele strongly
protects against alcoholism (Quertemont, 2004).
6. Structural Plasticity in Alcohol Dependence.
Because dendritic spines are both structural and functional units of excitatory synapses and include many
critical postsynaptic components of the synapse, any
changes in spine number or morphology could have
long-reaching consequences for the connectivity of
neural networks. Similar to other drugs of abuse
(reviewed in other sections), chronic ethanol regulates
the spine dynamics in many brain areas important for
learning, memory, reward, executive function, and
stress (reviewed in Cui et al., 2013). The general picture
emerging from these studies is that chronically administered ethanol decreases many measures related to
dendrites and their spines, accompanied by functional
or behavioral alterations. For example, withdrawal
from long-term ethanol exposure decreases spine density in the amydgala, concomitant with increased
anxiety and decreased BDNF-Arc signaling (Pandey
et al., 2008). In the NAc, both chronic alcohol drinking
and alcohol withdrawal decreased spine density, accompanied by reduction in NMDAR-mediated synaptic

934

Korpi et al.

currents and hampered LTD (Zhou et al., 2007; Spiga


et al., 2014). On the other hand, prolonged intermittent
alcohol drinking in cynomolgus monkeys increased the
density of dendritic spines and glutamatergic transmission in the putamen, but not in the caudate nucleus
(Cuzon Carlson et al., 2011). Similarly, chronic intermittent alcohol exposure increased spine density in the
mPFC in mice (Kroener et al., 2012). These findings
suggest that the direction of alcohol-induced alterations
in dendritic spine plasticity is dependent on brain
region. The functional consequences of these structural
alterations in the brain networks remain to be
elucidated.
It is noteworthy that some structural changes seen in
ethanol exposure in adult animals can also be seen if
ethanol treatment has been given during the prenatal
period. Prenatal ethanol treatment retarded the growth
of the SN pars compacta TH-positive neurons and
produced abnormalities in dendritic branching (Shetty
et al., 1993). In the forebrain, ethanol exposure produced a sexually dimorphic effect on dendritic branching in the NAc and reduced spine density and
distribution in the PFC (Lawrence et al., 2012). In
humans, heavy ethanol consumption during pregnancy
produces fetal alcohol spectrum disorders, of which the
most severe form is the fetal alcohol syndrome. Although
it is accepted that high ethanol doses can produce
detrimental effects on brain development, the mechanisms remain elusive. Data from animal models suggest
that mechanisms of ethanol action on the developmental
trajectory of neural cells include alterations in many
neurotransmitter and signaling systems, neural migration, neurogenesis, and synaptic plasticity (Valenzuela
et al., 2012).
7. Conclusions. Ethanol exerts its acute and chronic
effects via a number of molecular targets, altering
synaptic transmission, brain function, and behavior
(Fig. 11). Acutely, ethanol increases synaptic and
extrasynaptic inhibition by enhancing GABAergic
mechanisms and decreases excitation by inhibiting
NMDARs. The net effect of these neurochemical events
is dampening of synaptic excitation, which also either
hampers or modulates various forms of synaptic plasticity. Chronic ethanol exposure leads to compensatory
homeostatic effects in these neurotransmitter systems.
Thus, the function of the NMDARs is enhanced and
extracellular glutamate levels are elevated, contributing to neuronal hyperexcitability implicated in ethanol
dependence and withdrawal. Chronic ethanol effects on
GABAergic transmission lead to a hypofunction of
GABAergic systems. Particularly in the extended amygdala, chronic synaptic neuroadaptations involve interactions of the GABAergic neurotransmission with various
neuropeptides that either mediate or counteract ethanolinduced alterations. The between-systems adaptations
could contribute to the development of ethanol dependence and mediate the interactions between stress,

anxiety, and chronic ethanol effects. In addition, these


neuropeptide systems could be important future targets
for pharmacological treatment of alcoholism.
E. Benzodiazepines and Other GABAergic Drugs
Benzodiazepines (BZs) were introduced to clinical use
as anxiolytics in the 1960s, and soon they displaced
barbiturates and similar compounds as safer and more
efficient anxiolytics and hypnotics. BZs were acutely
well-tolerated and patients are usually compliant with
the therapy. Unfortunately, anxiety did not always
vanish. Soon, problems with their use emerged as some
patients developed tolerance, and in many users
attempts to withdraw from the medication lead to
a return of the initial anxiety symptoms. This apparently led to a strong urge for continued use, with, in
some patients, considerable difficulty in cutting down
chronic BZ medication and possibly a negative effect on
concurrent cognitive therapy (e.g., Westra and Stewart,
1998; Vorma et al., 2002). Whether BZ therapy persistently affects cognitive functions has been difficult to
assess. For example, Puustinen et al. (2007) found no
association of cognitive impairment (assessed with
Mini-Mental State Examination) with long-term use of
BZs in a nonrandomized study comparing patients
admitted to acute hospital wards during 1 month with
and without BZ treatment, whereas a small metaanalysis found significant impairments (Barker et al.,
2004a). The long-term BZ effect was clearer when the
reversal of cognitive impairment in different domains of
neuropsychological tests was monitored during withdrawal from prolonged BZ use (Barker et al., 2004b).
The results showed significant reversal during abstinence, which was not quite complete in comparison with
the scores in controls or normative data. Recently,
several epidemiologic studies reported increased risk
of dementia symptoms in older BZ users (Billioti de
Gage et al., 2015), with the risk increasing with
cumulative dose and treatment time. Causality of the
association is not yet known, because BZ drugs may also
have some protective effect against Alzheimers disease
(Fastbom et al., 1998) or be associated with insomnia
and increased risk for this disease (Chen et al., 2012).
BZs are still used in clinical medicine as anxiolytics,
sedatives, and hypnotics, although guidelines have
suggested that their use in anxiety disorders should
be indicated only in severe cases that do not respond to
certain antidepressants (particularly serotonin selective reuptake inhibitors), gabapentin/pregabalin, or
psychotherapy (Baldwin et al., 2014). BZ-site acting
hypnotics, preferably short-acting ones, should be prescribed for shorter treatment periods with careful
tapering off the dosages. Antihistamines and melatonin
should be tested for insomnia and a2-adrenoceptor
agonists or anesthetics be used for sedating agitated
and aggressive patients before using BZs. Furthermore,
nonpharmacological therapies, such as exposure and

Drug-Induced Neuroplasticity

cognitive psychotherapies, may provide good results for


many patients with less severe insomnia (Hofmann and
Smits, 2008).
Like all other abused drugs, only a portion of patients
will show increase in tolerance and develop troublesome
dependence during BZ therapy. Therefore, these efficacious drugs should not be fully forgotten (e.g., see
Starcevic, 2014). Therefore, it is important to understand their mechanisms of action and their long-term
effects on brain structure and function. In this section,
we will also shortly review the effects of related drugs,
such as GABAB agonists, g-hydroxybutyrate (GHB),
and anesthetetic agents.
1. Molecular Targets for Benzodiazepines. The
GABAA receptor is the main fast-acting inhibitory
neurotransmitter receptor in the mammalian brain
(for reviews for molecular biology and receptor subtypes, genetics, structure, and function, see Sieghart,
1995; Whiting et al., 1995; Olsen and Sieghart, 2008,
2009; Uusi-Oukari and Korpi, 2010). It is a pentameric
integral plasma membrane ligand-gated ion channel,
assembled from 19 different subunits (a1-6, b1-3, g1-3,
d, , u, p, and r1-3), which are products of different genes
residing in several clusters in the human and rodent
genomes (Olsen and Sieghart, 2008). Most GABAA
receptors have the subunit structure of 2a22b2g/d.
Subunits are heterogeneously expressed in various
brain cell types and regions, thus producing pharmacologically and functionally different GABAA receptor
subtypes (Wisden and Seeburg, 1992; Luddens et al.,
1995; Hevers and Luddens, 1998; Korpi et al., 2002;
Sigel and Steinmann, 2012), and some of them are
produced in alternative spliced or RNA edited isoforms.
Positive BZ-site modulators, such as diazepam, midazolam, and zolpidem, typically need the a and g
subunits for their allosteric actions increasing affinity
of GABA binding (Sigel, 2002). The most important site
of the receptor molecule for BZ binding is the 100th
residue of the a subunits, which is histidine (His) in all
diazepam-sensitive receptors with a1, a2, a3, and a5
subunit-containing receptors and arginine in diazepaminsensitive a4 and a6 subunit-containing receptors
(Luddens et al., 1995). This residue was originally
detected in an abnormal subunit a6 (Luddens et al.,
1990; Wieland et al., 1992), with restricted expression to
cerebellar granule neurons. Receptors with this subunit
do not bind diazepam but bind an inverse agonist
imidazobenzodiazepine Ro 15-4513, correlating with
the pharmacological profile of the cerebellar granule
cell layer (Sieghart et al., 1987; Malminiemi and Korpi,
1989; Uusi-Oukari and Korpi, 1990; Turner et al., 1991).
Furthermore, a spontaneous mutation changing arginine (Arg) to glutamine (Gln) was detected in the
cerebellar a6 subunit of alcohol-sensitive, and particularly BZ-sensitive, ANT rat line (Korpi et al., 1993;
reviewed in Korpi et al., 2007). Glutamine behaved
in the receptor similarly as histidine, drastically

935

increasing the receptor sensitivity to benzodiazepines


for binding and positive allosteric modulation (Korpi
et al., 1993). These observations were extended by Uwe
Rudolph, Hanns Mhler, and colleagues by using His to
Arg knock-in mutations for understanding the roles of
various BZ-sensitive GABAA receptors in the behavioral
and physiologic effects of diazepam in mouse models
(Table 2 with the references). Importantly, in mice,
these findings suggested that anxiolytic actions could be
separated from sedative actions, because the former
effects were absent in a2(His101Arg) knock-in mice,
whereas the a1(His101Arg) mice exhibited limited
sedative effects of diazepam (Low et al., 2000) or even
hyperactive phenotype after a sedative dose of diazepam (McKernan et al., 2000). These mouse models have
also indicated that tolerance and dependence-inducing
effects of diazepam are dependent on a1 and a5 subunitcontaining receptors (van Rijnsoever et al., 2004; Tan
et al., 2010). All these effects are also associated with g2
subunit-containing receptors, because this subunit is
obligatory for BZ binding (Pritchett et al., 1989).
Because the g2 subunit gives the subcellular address
for GABAA receptors to be trafficked into inhibitory
synapses (Essrich et al., 1998), all these knock-in mouse
models are related to BZ actions on phasic synaptic
inhibition.
More recently, the focus has moved to extrasynaptic
GABAA receptors, which usually contain the d subunit
instead of g2, with preferential expression together
with either a6 (in cerebellum) or a4 (in forebrain)
subunits (Wisden et al., 1992; Jones et al., 1997). These
receptors do not form high-affinity BZ binding sites (but
see Wallner et al., 2006), but they form a receptor
population with high GABA sensitivity, low conductance, and slow desensitization, and their main function
is to provide tonic extrasynaptic inhibition regulated by
ambient GABA (Mody, 2001; Stell and Mody, 2002;
Farrant and Nusser, 2005). When this kind of tonic
inhibition is abolished from the cerebellar granule cells
by genetically knocking out the a6 subunit, it is compensated by increased expression of two-pore leak K+
channels of TASK-1 type (Brickley et al., 2001). For these
neurons, it has been estimated that the charge transfer
via tonic inhibition is at least 10 times greater than that
via phasic synaptic inhibition (Brickley et al., 1996),
further indicating the physiologic importance of tonic
inhibition. The d subunit-containing receptors have high
affinity to GABA-site agonists muscimol and gaboxadol,
which are less potent in inducing behavioral effects in
GABAA d-KO mice (Chandra et al., 2010). Also, the effects
of neurosteroid agonists, such as ganaxolone, are reduced
in d-KO mice (Mihalek et al., 1999).
The peripheral BZ receptor (PBR) has been detected
in many tissues using the same BZ ligands as for
GABAA receptors. The PBR is now known to be the
mitochondrial translocator protein (TSPO), which is
believed to be particularly involved in cholesterol

936

Korpi et al.

transport into inner mitochondrial matrix for the initial


step of steroidogenesis by P450 side chain-cleavage
enzyme producing pregnenolone (reviewed in Rupprecht
et al., 2010). TSPO-KO from testicular Leydig cells,
however, failed to affect steroidogenesis (Morohaku
et al., 2014). In the brain, the PBR is enriched in glial
cells, astrocytes, and microglia, but it is also found in
some principal neurons. Although most actions of BZs
can be explained by facilitation of GABAA receptors
(see Table 2), it is possible that they also partly act via
stimulation of neurosteroid synthesis (progesterone = .
allopregnanolone; deoxycorticosterone = . 3a,5atetrahydrodeoxycorticosterone) as recently shown
for some brain mechanisms with a neurosteroid synthesis blocker finasteride abolishing the effects of BZs
(Rupprecht et al., 2009; Tokuda et al., 2010). Importantly,

these neurosteroids have rapid effects via allosteric


modulation of GABAA receptors. Novel TSPO ligands
have shown efficacy in reducing anxiety in animal
models and initial human experiments (Rupprecht
et al., 2010), in the absence of affinity to GABAA
receptors. TSPO ligands are being also developed as
PET ligands to monitor glial cell dynamics in vivo in the
brain, including microglial activation as a measure of
neuroinflammation (e.g., Van Camp et al., 2010).
2. Neuroplasticity Induced by Benzodiazepines and
Related Compounds. Sedative compounds like BZs
are generally known to depress brain activity, both in
animal models and humans (Kelly and McCulloch,
1982; Kelly et al., 1986; Wang et al., 1996; Freo et al.,
2008). Ito et al. (1994) compared clonazepam (that acts
selectively on BZ sites of the GABAA receptor, not on

TABLE 2
Summary of the roles of GABAA receptor subunits in benzodiazepine- and GABA-site drug effects and in neuroplasticity-related phenomena in
the brain
Subunit

a1

a2

Subtypesa

Location

a1bg2,

60%, syn and extrasyn

a1bd

all brain regions, incl.


interneurons

a2bg2

syn, limbic areas, cortex,


principal neurons

Behavior/Physiology

Model

References

BZ: sedation, anterograde amnesia,


anticonvulsant effect, addiction,
reward, anxiolysis?
- developmental plasticity
- depression of brain activity?
- conditioned fear
- neuroplasticity of VTA DA neurons

s.d.: zolpidem

(Low et al., 2000)

a1(His101Arg)

(Smith et al., 2012)


(Crestani et al., 2000)
(Tobler et al., 2001)
(Tan et al., 2010)
(Li et al., 2009)
(Huopaniemi et al., 2004)
(Fagiolini et al., 2004)
(Heikkinen et al., 2009)

BZ: anxiolysis, analgesia, anesthesia,


myorelaxation, addiction?
- adult neurogenesis
- CTA to ethanol
- conditioned fear
- reduction of ICSS threshold
(reward)

s.d.:

(Low et al., 2000)

a2(His101Arg)
a2-KO

(Crestani et al., 2001)


(Blednov et al., 2013)
(Smith et al., 2012)
(Reynolds et al., 2012)
(Witschi et al., 2011)
(Knabl et al., 2008)
(Duveau et al., 2011)

a3

a3bg2

syn and extrasyn, limbic


areas, monoamine
neurons, reticular
nucleus of thalamus,
principal neurons

BZ: myorelaxation, anesthesia,


analgesia, anxiolysis, addiction?
-thalamocortical oscillations,
regulated by endozepines?
- reduction of ICSS threshold
(reward)
- sensorimotor gating

s.d.: TP003

(Low et al., 2000)

a3(His126Arg)

(Crestani et al., 2001)

a3-KO

(Straub et al., 2013)


(Christian et al., 2013)
(Reynolds et al., 2012)
(Knabl et al., 2008)
(Yee et al., 2005)
(Sohal et al., 2003)

a5

a5bg2

extrasyn, hippocampus,
cortex, principal
neurons

BZ: learning and memory


-sensorimotor gating
-BZ tolerance

zolpidem-IS
s.d: L-655708
a5(His105Arg)
a5-KO

(Hauser et al., 2005)


(Yee et al., 2004)
(van Rijnsoever et al., 2004)
(Crestani et al., 2002)

a4, a6, d

a4/6bg2,

syn (g2), extrasyn (d),


cerebellum (a6),
forebrain (a4)

BZ-IS

s.d.: gaboxadol,
neurosteroids
a4-KO
a6-KO

(Duveau et al., 2011)

a4/6bd

sedation, sleep, anesthesia


- tonic inhibition
- depression of brain activity
- adult neurogenesis (a4)
- neuroplasticity of VTA DA neurons

(Vashchinkina et al., 2012, 2014a)


(Stell et al., 2003)
(Chandra et al., 2006)
(Jones et al., 1997)
(Brickley et al., 2001)

BZ, benzodiazepine; CTA, conditioned taste aversion; ICSS, intracranial self-stimulation; KO, knockout; IS, insensitive; s.d., selective drug; syn, synaptic; extrasyn,
extrasynaptic; VTA, ventral tegmental area.
a
Receptor subtypes are from the review of Olsen and Sieghart (2008).

Drug-Induced Neuroplasticity

TSPO) and the GABA-site agonist muscimol in rats


using [14C]deoxy-D-glucose autoradiography, and found
reductions of brain glucose uptake by both of them.
However, although clonazepam reduction plateaued at
30% reduction, the effect of muscimol was linearly
related to estimated receptor occupancy at least until
60% reduction of glucose uptake. These results suggest
a major difference in how ligands affecting preferentially synaptic and extrasynaptic GABAA receptors
influence brain metabolism. In addition, these results
generally agree with depressed neuronal activation
after treatments with GABAA agonist drugs using
immediate early gene assays, such as immunohistochemistry of c-Fos expression (see below).
Acute BZs generally depress also gene expression in
the brain (Huopaniemi et al., 2004), except for the CeA
(Beck and Fibiger, 1995; Hitzemann and Hitzemann,
1999; Shaw et al., 2011; Panhelainen and Korpi, 2012).
Induction of immediate early genes in the CeA is
common for various anxiolytics, but not specific for
them (Thompson and Rosen, 2006). Huopaniemi et al.
(2004) used a high dose of diazepam in mice (30 mg/kg)
and found a short reduction in expression of neuroplasticityrelated genes such as BDNF and c-Fos, but, interestingly,
the expression of CaMKIIa remained diminished for
at least 40 hours after a single diazepam administration. These effects on gene expression were not observed
in a1(His101Arg) knock-in mice. Thus, ligands promoting the activity of the main type of GABAA receptor
(a1 subunit-containing) depress brain activity and gene
expression, and by that they might not be expected to
induce neuroplasticity changes in the brain. Acute BZ
treatments might thus strongly prevent neuroplasticity.
In addition, Licata et al. (2013) found that single doses
of triazolam and zolpidem reduced BDNF in the mouse
HC, but unexpectedly a 7-day twice daily treatment with
diazepam or zolpidem failed to change BDNF expression.
However, concurrent 2- to 3-week diazepam treatment
prevents the stimulation of neurogenesis by fluoxetine in
normal rats (Wu and Castren, 2009) and in socially
isolated anxious rats (Sun et al., 2013). It remains to be
tested whether even a single administration of BZs would
prevent antidepressant-induced neuroplasticity in the
rat models.
Acute effects of BZs on DA release and synthesis have
been contradictory, not fully supporting any role for
midbrain DA neurons in reinforcing and dependenceproducing effects of BZs. However, Heikkinen et al.
(2009) observed similar increase in AMPAR/NMDAR
current ratios of VTA DA neurons ex vivo 13 days after
a single administration of diazepam and the GABAA
receptor a1 subunit-preferring zolpidem. This effect
could be blocked by the benzodiazepine-site antagonist
flumazenil and, interestingly, by the NMDAR antagonist dizocilpine. This finding was confirmed and extended by Tan et al. (2010) by showing that midazolam,
either systemically or intra-VTA, induced a persistent

937

change in DA neuron glutamate synapses. AMPARmediated responses showed inward rectification at


positive holding potential in the presence of intracellularly added polyamines, indicating increased targeting
of Ca2+-permeable non-GluA2 AMPA receptor subunits
(Jonas and Burnashev, 1995; Dingledine et al., 1999).
Furthermore, they showed that the primary target of
midazolam was the local GABAergic interneuron
expressing a1 subunit-containing GABAA receptors.
This result was confirmed by using a1(His101Arg)
mutant mice, indicating that the effect by BZs was
mediated by disinhibition of DA neurons. In proof, by
recording spontaneous IPSCs, Tan et al. (2010) observed that midazolam increased the phasic synaptic
inhibition of the VTA GABA neurons, whereas at the
same time the synaptic inhibition of the DA neurons
was diminished. Furthermore, in a1(His101Arg) mouse
midbrain slices, midazolam was ineffective in inhibiting
GABA neurons, which was now accompanied with
greater inhibitory current in DA neurons.
These recent findings clearly indicate that BZ addiction may also have a similar positive reinforcing
component as most other drugs of abuse (e.g., see Ator
et al., 2005; Straub et al., 2010). This component adds to
the better-known negative reinforcement, such as BZinduced relief from punishment or from anxious dysphoria, guilt, and shame. It is also plausible that BZs
have more widespread neuroplasticity effects in other
brain regions such as the HC (for chronic BZ withdrawal
effects, see Shen et al., 2009), although the persistency
and efficacy of single versus chronic dosing are still
poorly known. However, it should be kept in mind that
BZ agonists are especially efficient in reducing conflict
arising from internal or external cues in simple and
more complicated experimental animal tests, such as
Vogel and Geller-Seifter conflict tests (Nazar et al.,
1997; Millan and Brocco, 2003), and conditioned
taste aversion and neophobia tests (Yasoshima and
Yamamoto, 2005; Wislowska-Stanek et al., 2008;
Callaerts-Vegh et al., 2009). Furthermore, the BZ
lorazepam can reverse punishment-suppressed operant
responding for the opioid remifentanil in rats (Panlilio
et al., 2005), indicating its efficacy in reversing the
aversion-induced change in behavior (a conflict situation,
punishment by mild footshocks). However, it was inefficient in reinstating extinguished responding for remifentanil, whereas heroin was effective in both reversing
punishment-suppressed and extinguished responding.
Thus, BZs are efficient in counteracting learned aversion-induced behaviors, by negative reinforcement, and in
addition to their direct reinforcing effect. Most likely,
other related anxiolytic GABAA agonists and other drugs
showing efficacy in conflict situations (Millan and Brocco,
2003) would reinstate punishment-suppressed responding. Whether this effect is a long-lasting one is not known.
Anyway, the possibility that an abstinent alcoholic or
a drug addict might lose the abstinence-promoting effect

938

Korpi et al.

of social control/punishment should be taken into account


when prescribing anxiolytic drugs. One interesting example of the short-term BZ effect is the midazolam-induced
attenuation of conditioned (LiCl, i.p.) taste aversion to
sucrose, which unfortunately returned back to full aversion in the next session (Yasoshima and Yamamoto,
2005).
Does the acute stimulation of tonic extrasynaptic
GABAergic inhibition by selective drugs induce neuroplasticity in VTA DA neurons? This question was studied
by Vashchinkina et al. (2012, 2014a). First, the GABAsite agonist gaboxadol (4,5,6,7-tetrahydroisoxazolo(5,4-c)
pyridin-3-ol), which is a superagonist at d subunitcontaining receptors, whereas GABA is a partial agonist
(Storustovu and Ebert, 2006; Saarelainen et al., 2008),
induced prolonged (.6 days) increase in AMPAR/
NMDAR current ratios and AMPAR current rectification
in VTA DA neurons 24144 hours after a single sedative
dose in a dose-dependent manner (Vashchinkina et al.,
2012). Second, this finding was extended to a neurosteroid
agonist ganaxolone, which does not have genomic effects
like endogenous neurosteroids (Carter et al., 1997). A
single mildly sedative dose of ganaxolone induced a similar long-lasting neuroplasticity in VTA DA neurons ex
vivo as a dose of gaboxadol (Vashchinkina et al., 2014a).
By using a GAD67-GFP knock-in mouse line, in which
GABA neurons are fluorescently labeled, ganaxolone in
vitro provoked a fourfold increase in tonic GABAAmediated conductance in VTA GABA neurons, but not
in DA neurons of the TH-EFGP transgenic mice. In vitro
incubation of midbrain slices with 0.5 mM ganaxolone for
.4 hours also induced increased AMPAR/NMDAR current ratio in DA neurons. In agreement with the preference for d subunit-containing receptors, the
neuroplasticity effects of gaboxadol and ganaxolone were
abolished in d-KO mice (Vashchinkina et al., 2012,
2014a). However, the level of d subunit protein expression
in VTA GABA neurons is not very high, at least compared
with the cerebellar granule neurons, and, therefore,
direct immunohistochemical demonstration of its presence has failed because of background labeling even in the
d-KO mice by available d subunit antibodies (unpublished).These results indicate that drugs potentiating
tonic inhibitory conductance also target primarily
GABAergic interneurons in the VTA and induce neuroplasticity effects in DA neurons by disinhibition.
Surprisingly, both gaboxadol and ganaxolone failed to
induce CPP, and if anything they induced aversion
(Vashchinkina et al., 2012, 2014a). Gaboxadol failed to
sustain the acute nose-poking response in mice using
the yoked control paradigm of intravenous administration over a wide dose range nor was it self-administered
in cocaine-primed baboons, although the BZ agonist
triazolam could replace cocaine (Vashchinkina et al.,
2012). In line with the electrophysiological results,
conditioned aversion to ganaxolone was abolished in
d-KO mice (Vashchinkina et al., 2014a). Because a short

aversive swimming stress induces neuroplasticity in


VTA DA neurons (Saal et al., 2003), it appears possible
that there are several populations of midbrain (and
VTA) DA neurons that may be similarly/differentially
adapted to various rewarding and aversive stimuli
(Matsumoto and Hikosaka, 2009; Bromberg-Martin
et al., 2010; Lammel et al., 2011, 2012; Fiorillo, 2013;
Fiorillo et al., 2013; Jennings et al., 2013; Ilango et al.,
2014). It is also possible that triggering conditioned
aversion takes place primarily in other brain regions,
such as the BNST (Kim et al., 2013b) and/or the LHb
(Matsumoto and Hikosaka, 2007; Hikosaka, 2010;
Stamatakis and Stuber, 2012), which have reciprocal
connections with the VTA.
Ca2+-permeable NMDARs are often involved in LTP,
but in NMDAR-independent forms of LTP, VGCC are
important (Morgan and Teyler, 1999; Bauer et al., 2002;
Schroeder and Shinnick-Gallagher, 2004; Bloodgood
and Sabatini, 2007). Tietz et al. (see below for references) made a series of detailed experiments on the
mechanisms of tolerance to and withdrawal from subchronic flurazepam treatment in the hippocampal CA1
area. They treated rats with a high oral dose of
flurazepam for at least 1 week and studied the brains
under tolerant and 12 day withdrawal conditions.
Several interesting transient changes were observed
in their studies. Downregulation of GABAA receptorassociated ligand binding and GABAA receptor subunit
levels took place in several brain regions of tolerant
animals (Rosenberg et al., 1985; Tietz et al., 1999) (for
a review on GABAA subunit expression after various
GABAA drugs, see Uusi-Oukari and Korpi, 2010).
Surprisingly, the changes in the GABAA system and
tolerant behavior were abolished in the tolerant, withdrawal BZ-free state by a single administration of the
BZ antagonist flumazenil (Tietz et al., 1999), with
the testing being carried out 1 day after the injection
of the short-acting flumazenil. Increased AMPAR currents and binding sites were detected in the hippocampal CA1 and CA3, but not dentate gyrus, region of
withdrawing animals (Van Sickle and Tietz, 2002), the
increased currents correlating with increased withdrawal anxiety (Xiang and Tietz, 2007; Xiang et al.,
2008). GluA1, but not GluA2, subunit protein was
increased synaptically (Das et al., 2008) because of
GluA1 subunit phosphorylation at Ser831 by CaMKII
(Earl et al., 2012). Whereas the increase in AMPAR
function could not be abolished by the NMDAR antagonist dizocipline (MK-801), it was abolished by the
L-type VGCC blocker nimodipine, which also reversed
withdrawal anxiety. At day 2 of withdrawal, the levels
of GluN1 and GluN2B subunits and NMDAR function
were also downregulated, which could be reversed by
blocking AMPARs and their upregulation (Van Sickle
et al., 2004; Xiang and Tietz, 2007), suggesting that the
change in AMPAR was the primary alteration in the
glutamate system in BZ tolerance. These results

Drug-Induced Neuroplasticity

demonstrate an important cascade of transient (14


days) functionally interrelated changes in the hippocampal neurons that are caused by chronic, but not
acute (Van Sickle et al., 2002), BZ administration.
These results suggest that even chronic heavy BZ
treatment induces rather short-lasting effects at least
on behavior and HC, but that they also show some
potential for inducing neuroplasticity.
3. Behavioral After-effects of g-Aminobutyric Acid A
Drugs. Once the single-dose neuroplasticity effect of
the BZ agonists was found (Heikkinen et al., 2009; Tan
et al., 2010), it was of interest to study whether behavioral responses to other drugs of abuse would be affected
by pretreatment (and neuroplasticity effects) of BZs.
Panhelainen et al. (2011) gave a single injection of
diazepam (5 mg/kg i.p.) and tested adolescent and adult
mice for psychomotor stimulation by three daily doses of
morphine and d-AMPH. In contrast to a hypothesis of
enhanced psychomotor activation, based on the sensitizing effects of viral-mediated upregulation of GluA1
subunit in the midbrain (Carlezon and Nestler, 2002),
we found that diazepam pretreatment reduced locomotor responses to morphine (in both sexes) and attenuated behavioral locomotor sensitization to d-AMPH,
with the effects being detectable, although slightly
diminished, also in adult mice (Panhelainen et al.,
2011). The effects on d-AMPH-induced sensitization
could be blocked by a single injection of dizocilpine prior
to diazepam pretreatment. These preliminary results
suggest that occlusion of neuroplasticity attenuates
behavioral effects induced by acute and repeated morphine and amphetamine.
Because gaboxadol also induces neuroplasticity in VTA
DA neurons, which lasted even longer that that induced
by diazepam (Vashchinkina et al., 2012), we repeated
Panhelainens study (Panhelainen et al., 2011) to test
whether the pretreatment with gaboxadol would also affect
the psychomotor stimulation by morphine and d-AMPH.
However, we observed no alterations in the responses in
adolescent or adult mice (unpublished results). This
suggested that rewarding BZs and aversive gaboxadol
might induce neuroplasticity in different populations of
VTA DA neurons, but preliminary examination did not
indicate any regional variation in the VTA of DA neurons
with altered glutamate transmission after gaboxadol
(Vashchinkina et al., 2012). It is also possible that GABAA
receptor drugs persistently affect many brain regions
other than the VTA. To start examining this possibility, we counted c-Fos positive neurons in aversionassociated brain regions 2 hours after a single injection
(6 mg/kg) of gaboxadol (when sedation was gone), and,
interestingly, found a significant activation exclusively
in the oval part of the BNST (E. Vashchinkina et al.,
submitted manuscript). In parallel behavioral tests,
gaboxadol provoked a mild anxiety-like withdrawal
state, which was preceded by transiently increased
plasma corticosterone level.

939

In addition, we examined whether selective blockade


of CRF1 receptors would attenuate gaboxadol-induced
plasticity in VTA DA neurons, because CRF signaling is
often involved in oval BNST-mediated behaviors (Erb
et al., 2001b; Beckerman et al., 2013). Indeed, gaboxadolinduced VTA DA neuron plasticity was totally blocked by
pretreatment with CP-154526, a selective CRF1 receptor
antagonist (Shaham et al., 1998) (E. Vashchinkina et al.,
submitted manuscript). Thus, VTA DA plasticity and
anxiogenic/aversive consequences of activation of extrasynaptic GABAA receptors seem to be linked to activation
of oval BNST and CRF1 signaling.
In contrast to gaboxadol experiments, 2 hours after
diazepam administration, a strong increase in the
number of c-Fos-positive cells was found in the CeA
(Panhelainen and Korpi, 2012). In addition, by using the
Golgi method for counting dendritic spines, we found
that a single dose of gaboxadol increased spine counts in
the HC, whereas diazepam was inactive (Vashchinkina
et al., 2014b). These results suggest that the effects on
brain-wide cellular activity are different in diazepamand gaboxadol-treated mice, making it possible that
synaptic and extrasynaptic GABAA neurotransmissions
modulate different pathways to the VTA DA neurons.
4. Effects of Flumazenil on Benzodiazepine and
Alcohol Tolerance. Flumazenil (Hunkeler et al., 1981)
is acutely very efficient and specific in clinically antagonizing the actions of all BZ-site agonists, whether they
have the BZ structure or other chemical structures, such
as zolpidem and zopiclone. It is available as intravenous
injection and its half-life is short (about 1 hour) in humans
(Klotz et al., 1984; Roncari et al., 1993). It is primarily
used in BZ intoxications or in the treatment and diagnosis
of drug intoxications (Hoffman and Warren, 1993;
Kreshak et al., 2012). As a competitive BZ antagonist,
its acute actions can be well understood as counteracting
the effects of agonists.
Interestingly, flumazenil has shown variable intrinsic activity in preclinical and clinical settings, varying
from weak partial agonism (positive allosteric modulation) to partial inverse agonism (negative modulation)
(e.g., File et al., 1982a,b; Nutt et al., 1990; Da Cunha
et al., 1992; Smolnik et al., 1998; Izumi et al., 1999;
Goulenok et al., 2002). Its intrinsic efficacy has been
explained by antagonism of endogenous positive or
negative modulators. Indeed, there is evidence for
endogenous production of compounds with binding
and functional activity at BZ sites (endozepines), such
as diazepam and its metabolite desmethyldiazepam
(Sangameswaran et al., 1986; Klotz, 1991), and molecules with other structures, including heme derivates
(Ruscito and Harrison, 2003) and the diazepam binding
inhibitor (DBI) and its proteolytic peptide fragments,
such as octadecaneuropeptide (Alho et al., 1985; Costa
and Guidotti, 1991). Recently, DBI and octadecaneuropeptide have been shown to increase neurogenesis in
postnatal mouse subventricular zone by negative

940

Korpi et al.

modulation of GABAA receptors (Alfonso et al., 2012),


and, surprisingly, DBI was described to positively
modulate reticular thalamic a3 subunit-containing
GABAA receptors, an effect that appears to be needed
for antiabsence seizure effect in a mouse model
(Christian et al., 2013). Production of endozepines is
especially high during acute hepatic failure with and
without encephalopathy (Olasmaa et al., 1990; Avallone
et al., 1998). This state is also associated with increased
levels, expression, and function of the DBI peptide and
TSPO protein, which also leads to increased levels of
neurosteroids THDOC and allopregnanolone (Rothstein
et al., 1989; Itzhak et al., 1995; Ahboucha et al., 2006,
2012). Flumazenil has a therapeutic effect in patients
with hepatic encephalopathy and coma (Bansky et al.,
1989; Goulenok et al., 2002), although it does not
counteract the effects of neurosteroids. Although the
setpoint of the GABAA receptors determining positive or
negative BZ modulation might be different between
different subjects or it might change because of illness
or drug treatment, it is possible that at least a part of
the intrinsic activity of flumazenil can be accounted for
by antagonism of various endogenous ligands.
In regard to the present review, a most interesting
finding has been the efficacy of flumazenil to reverse BZ
and alcohol tolerance and dependence in preclinical
models by a single dose or by a pretreatment (Buck
et al., 1991; Flaishon et al., 2003; but see Nutt and
Costello, 1988; Loscher and Rundfeldt, 1990; Tietz
et al., 1999). Although this phenomenon has been tested
in BZ-dependent patients with inconsistent efficacy for
counteracting tolerance in acute settings (Cittadini and
Lader, 1991; Savic et al., 1991; Lader and Morton,
1992), a small clinical study has reported preliminary
results on prolonged subcutaneous flumazenil administration in withdrawing BZ-dependent patients (Hulse
et al., 2013), suggesting that perhaps longer treatments
are needed in human populations. Another study using
twice daily intravenous injections of flumazenil showed
that flumazenil significantly blocked withdrawal symptoms and kept the patients better in abstinence (Gerra
et al., 2002). Because flumazenil increases membrane
targeting of GABAA receptors (Flaishon et al., 2003;
Pericic et al., 2004), its efficacy in reversing BZ tolerance and in alleviating withdrawal in dependent subjects may at least partly be based on increased number
of functional GABAA receptors.
5. Treatment of Addictions by g-Aminobutyric Acid B
Receptor Agonists. Although the ligand-gated anion
channel GABAA receptors mediate fast phasic inhibition
and prolonged tonic inhibition in the mature nervous
system, the GPCR metabotropic GABAB receptors mediate neuronal inhibition by slower actions on neuronal
metabolism and K+ and Ca2+ channels both presynaptically at inhibitory and excitatory terminals and postsynaptically on dendrites and spines (Bowery et al., 2002;
Bettler et al., 2004). GABAB receptors are heterodimers of

two subunits, GABAB1 and GABAB2, that are large


(110 kDa) proteins with seven transmembrane domains
and long extracellular N terminus. GABAB1 seems to be
responsible for GABA binding and GABAB2 in turn for
G protein coupling (Bettler et al., 2004). Knockout of
either subunit eliminates the typical GABAB receptor
responses (Schuler et al., 2001; Gassmann et al., 2004). In
the simplest case, activation of presynaptic GABAB autoand heteroreceptors inhibits, via released Gi/o protein bg
subunits, the P/Q-type (Cav2.1) and N-type (Cav2.2)
currents of high VGCC. Stimulation of postsynaptic
GABAB receptors activates in the same way multiple
types of K+ channels, including inwardly rectifying Kir3
channels. After both presynaptic and postsynaptic activation, there is the release of Gi/oa subunits that inhibits
AC, leading to reduced cAMP production and PKA
activity (Bettler et al., 2004; Koyrakh et al., 2005). The
signaling of GABAB receptors is most likely more complicated and neuronal cell-population specific because of
differences in the receptors cell surface expression and
localization (Becher et al., 2004), isoform diversity (Bettler
et al., 2004), crosstalk with various ionotropic (e.g., GABAA,
NMDA), and metabotropic (e.g., mGlu, TrkB) receptors
(reviewed in Xu et al., 2014) and GABAB receptor interacting proteins. Notably, GABAB receptors interact with
several other proteins, including activating transcription
factor 4, inducing a novel signaling pathway (Nehring et al.,
2000; Vernon et al., 2001; Chen et al., 2003), regulators of G
protein signaling (RGS) (Labouebe et al., 2007), and, most
recently, potassium channel tetramerization domaincontaining proteins that bind to the intracellular carboxy
terminus and alter agonist sensitivity and desensitization
(Schwenk et al., 2010). In relation to neuroplasticity, it is
important to note that GABAB receptor activation reduces
the Ca2+ flux via NMDARs both by producing hyperpolarization due to opening of K+ channels and/or by reducing
the activities of AC and PKA, which reduces the phosphorylation of NMDAR subunits (Morrisett et al., 1991;
Chalifoux and Carter, 2010).
Activation of GABAB receptors inhibits the spontaneous pacemaker-like firing of VTA DA neurons ex vivo
(Seutin et al., 1994; Wu et al., 1999) and attenuates the
firing rate and burst firing of these neurons in vivo
(Erhardt et al., 2002). Therefore, facilitation of the GABA
system is being tested for the treatment of various drug
addictions, with two main lines of approach. Stimulation
of GABAB receptors reduces rewarding effects and
consumption of several addictive drugs in various preclinical models (Roberts, 2005; Vlachou and Markou,
2010), but so far the results from the treatment trials
on human addictions, especially on alcoholism, with the
selective direct sedative agonist baclofen [(3R)-4-amino-3(4-chlorophenyl)butanoic acid] have been inconsistent
(Tyacke et al., 2010; Agabio et al., 2013; Brennan et al.,
2013; Agabio and Colombo, 2014). To date, baclofen is the
only GABAB receptor agonist approved in clinical use, in
addition to g-hydroxybutyrate (see below).

Drug-Induced Neuroplasticity

The research focus is now on positive allosteric


modulators that act on a GABAB receptor site distinct
from GABA and baclofen (Pin and Prezeau, 2007).
Because positive allosteric modulators facilitate the
effects of GABA-site ligands (Jensen and Spalding,
2004), they alter activated receptors only. Hence they
rather enhance the physiologic pattern of GABAB receptor activation, and so may have less adverse effects
than ligands that alter all GABAB receptors. In agreement, allosteric modulators interfere less with locomotor activity (Ong and Kerr, 2005) and do not induce
hypothermia (Jacobson and Cryan, 2008).
It is not known whether the activation of GABAB
receptors would persistently reverse any addictionrelated neuroplasticity. Importantly, these drugs have
not shown any rewarding potential in intracranial selfstimulation or CPP paradigms, although they attenuate
addiction-related behavior such as sensitization, druginduced CPP, and SA of various addictive substances
(for review and references, see Filip et al., 2015).
Furthermore, there are features in GABAB receptor
activation that indicate reduction of neuroplasticity in
adult animals. These are inhibition of neurogenesis
(Giachino et al., 2014), increased expression of activating transcription factor 4 that reduces brain gene
expression and neuroplasticity (Nehring et al., 2000;
Vernon et al., 2001; Chen et al., 2003), reduced expression of BDNF (GABAB antagonists and receptor subunit
knockouts increase BDNF (Heese et al., 2000; Enna
et al., 2006) and, importantly, PKA-mediated blunting
of NMDAR-dependent Ca2+-transients of dendritic
spines in the PFC layer 2/3 (Chalifoux and Carter,
2010). In addition, GABAB receptors retard the synaptic
vesicle priming in glutamate synapses by reducing
cAMP (Sakaba and Neher, 2003). It is still unclear
whether activation or inhibition of GABAB receptors is
antidepressant (Cryan and Slattery, 2010), which may
turn out to be a complicating factor in the therapeutic
use of GABAB ligands.
Another interesting alternative for antiaddiction
drug treatment might be the antiepileptic, irreversible
GABA transaminase (GABA-T) blocker g-vinyl-GABA,
vigabatrin, which increases brain and cerebrospinal
fluid GABA concentration (Halonen et al., 1988; Petroff
et al., 1996). Vigabatrin reduces neurochemical changes
produced by cocaine in the brain reward areas, including
cocaine-induced DA release and cocaine consumption in
preclinical models (Dewey et al., 1997; Kushner et al.,
1997, 1999; Peng et al., 2008). At least some of its actions
can be abolished by GABAB receptor antagonists (Ashby
et al., 1999). Vigabatrin also affects synaptic and
extrasynaptic GABAA transmission (Overstreet and
Westbrook, 2001). Importantly, vigabatrin and a reversible GABA-T blocker ACC (1R,4S-4-amino-cyclopent-2ene-carboxylic acid) do not induce CPP (Paul et al., 2001;
Ashby et al., 2002). Vigabatrin has shown neuroprotective effects in epilepsy models (Halonen et al., 1995;

941

Pitkanen et al., 1996). Initial experiments showed some


efficacy in human cocaine addicts (Brodie et al., 2009), but
a larger study failed to show significant effects, perhaps
due to poor adherence of the addicts to vigabatrin
treatment (Somoza et al., 2013). This drug was originally
developed for epilepsy, in which indication it was efficient,
but permanent, slowly progressing vigabatrin-associated
visual field loss, possibly linked to its cumulative dosing,
has limited its use in epilepsy (Malmgren et al., 2001;
Maguire et al., 2010; Clayton et al., 2013). The mechanism
of retinal damage is not known (Ravindran et al., 2001). All
these efforts have led to the development of a more potent
GABA-T blocker, (1S, 3S)-3-amino-4-difluoromethylenyl1-cyclopentanoic acid (CPP-115) (Pan et al., 2012), which is
being further studied. As with GABAB agonists, there is no
clear information on possible long-term reversal of
addiction-related neuroplasticity by GABA-T blockers.
6. g-Hydroxybutyrate as a Drug of Abuse and
a Therapeutic Compound. GHB was originally synthesized for sedation and anesthesia, and as a GABA
analog it was supposed to enter the brain and act on
GABA receptors easier than exogenous GABA (Laborit,
1964). This short-chain fatty acid compound [half-life in
plasma about 1 hour (Abanades et al., 2006)] was used
for anesthesia (Kam and Yoong, 1998), but now it is
mainly used, with the non-proprietary name sodium
oxybate, in narcolepsy with cataplexy. It improves
consolidation of nonrapid eye movement sleep and
normalizes sleep stage shifts, and through these mechanisms particularly reduces appearance of daytime
cataplexic attacks in narcolepsy patients (Mamelak,
2009; Boscolo-Berto et al., 2012). It is also used in some
European countries in the treatment of alcohol withdrawal symptoms and in the maintenance of alcohol
abstinence (Leone et al., 2010). Interestingly, when
used at low doses in the treatment of narcolepsy, very
little tolerance develops to its sleep-improving effects,
but when abused for its euphoric and intoxicating
effects at high doses, tolerance and dependence develops and withdrawal symptoms appear spontaneously
(or in nonhuman primates after administration of
GABAB receptor antagonists) (Goodwin et al., 2013).
GHB is an endogenous compound, synthesized from
GABA and found in the brain at about 1 mM concentration (Maitre, 1997), but when exogenously given in
preclinical settings for sedative and hypothermic effects,
the doses needed are so high that brain concentrations
are elevated up to 1 mM (Klein et al., 2009), which are in
agreement with EC50 of about 1 mM to activate GABAB
receptors in midbrain slices (Madden and Johnson, 1998).
The preclinical effects of high GHB doses are dependent
on GABAB receptor activation (Kaupmann et al., 2003).
The identity (and significance) of the high-affinity GHB
binding site(s) in the brain is not known. These sites
might represent specific GHB receptor, unaffected in
GABAB knockout mice (Kaupmann et al., 2003), and/or
partly reflect the most sensitive GABAA receptors, the

942

Korpi et al.

extrasynaptic a4b1d receptors (Absalom et al., 2012).


However, these highly GHB-sensitive GABAA receptors
could not be found in specific thalamic, NAc, or HC
dentate gyrus neurons with known tonic inhibitory
currents examined with electrophysiological patchclamp methods in the brain slices from juvenile rats
(Connelly et al., 2013).
GHB has variable effects on firing of midbrain DA
neurons; both inhibition and activation of VTA DA have
been produced in animal models (Pistis et al., 2005).
Interestingly, RGS2 protein has been suggested to be
responsible for the high and low efficacy coupling of the
GABAB receptors to somatodendritic Kir3 K+ channels
in the VTA GABA and DA neurons, respectively
(Labouebe et al., 2007), which seems to explain why
GHB can acutely inhibit the GABA neurons similarly to
baclofen (Cruz et al., 2004). But, when GHB was given
subchronically at a high dose, animals showed aversion
to consuming the initially preferred GHB-containing
sucrose solution. This change coincided with the reduction of RGS2 expression in the VTA DA neurons and
the enhanced inhibition of these neurons by GABAB
activation (Labouebe et al., 2007). Recent experiments
on the effects of acute GHB on brain gene expression
have further indicated that GHB induces neuroplasticity changes in the brain (van Amsterdam et al., 2012).
This is consistent with the development of tolerance
and rewarding effects to repeated GHB administration
(Itzhak and Ali, 2002). Similarly to sedative GABAA
agonists, GHB globally reduces rat brain glucose metabolism (Kuschinsky et al., 1985). However, GHB provoked
surprisingly robust activation of c-Fos protein in many
subcortical brain areas, shared by similar effects by an
equipotent sedative dose of baclofen (van Nieuwenhuijzen
et al., 2009). The activation of thalamic nuclei, including
the LHb in the epithalamus, after high GHB doses that
induce cortical spiking (modeling absence seizures) can
be explained by the oscillating thalamocortical circuits
(Zhang et al., 1991; van Nieuwenhuijzen et al., 2009).
Importantly, the some of the reward-related areas, such
as the VTA and NAc, were clearly not activated by GHB
or baclofen (van Nieuwenhuijzen et al., 2009). Thus,
GHB remains an interesting, although a difficult compound in relation to addictions, with still much to be
learned about basic neurobiology and mechanisms of
action that contribute to its endogenous physiologic
functions, to mechanisms of abuse of exogenous GHB
and to its use in the treatment of alcohol addiction.
Finally, it should be remembered that GHB has similar
abuse liability to sedative drugs and alcohol (Griffiths
and Johnson, 2005; Carter et al., 2006; Johnson and
Griffiths, 2013), with severe delirium presenting in
many cases where use has been extensive and tolerance
has developed. This withdrawal state is hard to treat
with BZs, and GHB may need to be reinstigated and
then tapered slowly (de Jong et al., 2012). Baclofen
might be useful here, too.

7. Anesthetics and Neuroplasticity. Several inhalational (isoflurane, halothane, and sevoflurane) and intravenous (propofol, etomidate, barbiturates, and BZs)
anesthetics facilitate the function of GABAA receptors,
most likely as a part of their clinical mechanisms of
action (Sonner et al., 2003; Franks and Lieb, 2004;
Rudolph and Antkowiak, 2004). In addition to BZs and
barbiturates, these drugs might have addiction potential, but this question has not been much studied.
Propofol abuse has been reported (Wilson et al., 2010;
Bonnet and Scherbaum, 2012). In a recent study,
propofol showed abuse liability and euphoria on the
first exposure in patients undergoing sedation for
endoscopic examination (Kim et al., 2013a).
Because GABAA receptors are, at least in early brain
development, excitatory and essential for the normal
development of neuronal circuits (Akerman and Cline,
2007; Ben-Ari et al., 2007), prolonged facilitation or
activation of these receptor channels may have strong
effects on brain structure and function. Similarly, the
blockade of the NMDARs with ketamine, phencyclidine,
and alcohol affects the developing nervous system
(Olney et al., 1991; Ikonomidou et al., 1999, 2000).
Early studies with chronic halothane administration
(10 ppm halothane/air, 8 hours/day, 5 days/week,
throughout development or in adulthood for about 1.5
months) showed in rats that structural changes evaluated with the Golgi method (retarded synaptogenesis,
impaired dendritic branching, and disruption of organelle structure) were more abundant in the young than
adult rodent nervous system (Quimby et al., 1974; Levin
et al., 1991), the changes generally correlating with
impaired results in tests for cognitive function. On the
other hand, the persistent effects of anesthetics on the
nervous system have recently been studied using prolonged (46 hours) anesthesia at the age when brain
development is still occurring. At that stage, many
clinically used anesthetics and also other neuropsychiatric
drugs such as antiepileptics seem to induce neurotoxicity
(apoptosis) in rodents and nonhuman primates (Vutskits
et al., 2006; Briner et al., 2010, 2011; Creeley and Olney,
2010; 2013; Brambrink et al., 2012a,b; Creeley et al., 2013;
Jevtovic-Todorovic et al., 2013), with often a single administration being neurotoxic. Prolonged pediatric anesthesias for more than 4 hours, which are presently
clinically justified and often unavoidable to provide the
best treatment of the pediatric patient, have putatively
been associated with cognitive impairment later in life
(Jevtovic-Todorovic et al., 2013). Therefore, because
proper anesthesia is important in surgical treatment,
future efforts should be aimed at shortening durations of
anesthesia, finding novel mechanisms of action for anesthetics that would not so globally involve the ligand-gated
receptors of the main neurotransmitter systems, and/or
finding pretreatment options that would limit the
neurotoxicity. For example, lithium and clonidine
might be two compounds with some efficacy to prevent

Drug-Induced Neuroplasticity

anesthetic-induced toxicity (Young et al., 2008; Straiko


et al., 2009; Ponten et al., 2012), whereas hypothermia
may be an option to protect the brain from ischemia/
excitotoxicity and effects of prolonged anesthetics
(Ikonomidou et al., 1989). Ketamine-induced apoptosis is
enhanced by cotreatment with diazepam or midazolam
in a rodent model (Fredriksson et al., 2004; Young et al.,
2005), but diazepam has still been long used to alleviate
psychotomimetic adverse effects of ketamine in humans.
In addition, magnesium loading that has been used in
(pre)eclampsia, may be neurotoxic for the developing
nervous system (Dribben et al., 2009), and at least one
clinical trial with MgSO4 supplementation to pregnant
women with preterm labor had to be preliminarily
stopped due to increased pediatric mortality (Mittendorf
et al., 1997), especially to neonatal intraventricular
hemorrhage (Mittendorf et al., 2002). These last cases
are just two examples of problems recently encountered
in the treatment affecting negatively the developing fetal
or neonatal brain.
In the VTA DA neurons of juvenile mice, a single
sedating dose of diazepam or midazolam induces persistent increases in AMPAR/NMDAR current ratios
(Heikkinen et al., 2009; Tan et al., 2010). Intravenous
anesthetic propofol in vitro at low micromolar concentrations activates AMPARs on VTA DA neurons (Li
et al., 2008). Interestingly, isoflurane anesthesia alone
has been tested in treatment-refractory depressed
patients and found to produce long-lasting antidepressant effect, which was similar to electroconvulsive
therapy (under isoflurane anesthesia) (Langer et al.,
1985, 1995). These results were recently confirmed in
a study on refractory depression, which also showed
that electroconvulsive therapy (on an average ten
15-minute sessions within 3 weeks) had modestly more
efficient antidepressant effect than isoflurane anesthesia alone at 4-week follow up, but produced more severe
neurocognitive adverse effects (Weeks et al., 2013). In
rats, 7 days after a single 2-hour isoflurane anesthesia,
avoidance learning and HC LTP at CA1 synapses were
impaired (Uchimoto et al., 2014). However, a similar
single anesthesia in mice induced only some deficits in
attention tests 1 week afterward, and most results of
an extensive battery of behavioral and physiologic
tests were not altered (Yonezaki et al., 2015). In any
case, further research on possible prolonged neuroplasticity effects of GABAAergic compounds appears
warranted.
8. Conclusions. Although the GABAergic system is
important for neuronal development and circuit functions in the mature nervous system, drugs like benzodiazepines that facilitate the actions of the GABAA
receptors have many effects on the nervous system that
limit neuroplasticity. This may be due to their strong
sedative effect. However, these drugs act by negative
reinforcement, and more recently, benzodiazepines
have been shown to induce similar glutamate synapse

943

neuroplasticity in the DA neurons of the reward system


as stimulants, opioids, and alcohol, suggesting also
potential for positive reinforcement. Prolonged use of
these drugs even in normal therapy may cause cognitive
impairment that may not be fully reversible. Because
GABAA receptors exist in multiple heterogeneous subtypes, subtype-selective BZ ligands have been studied,
but this work has not yet translated into less addictive
drugs. Initial studies on activation of extrasynaptic
GABAA receptors suggest that they might show aversive features, perhaps by inducing glutamate neuroplasticity in a stress/aversion-sensitive subpopulation
of DA neurons and/or by activating other stress/
aversion-related brain areas. GABAB receptors hold
promise as a target for novel addiction therapies with
specific positive allosteric modulators or via increased
GABA levels.
F. N-methyl-D-aspartate Receptor Antagonists
NMDA antagonists are also known as dissociatives or
dissociative hallucinogens because of their subjective
effects. These are described as feelings of sensory detachment or dissociation from the outside world coupled
with perceptual disturbances as well as catatonia-like
motor phenomena. They are clearly distinct from the overt
visual hallucinations produced by serotonin-type hallucinogens (Gouzoulis-Mayfrank et al., 2005; Muetzelfeldt
et al., 2008). The most commonly used dissociatives are
ketamine and phencyclidine (PCP), whereas a third
NMDAR antagonist known as MK-801 or dizocilpine
has been widely used in pharmacological research but
never gained popularity among recreational users. In
recent years there has also been an increase in the use of
structural analogs with similar pharmacological activity
in the so-called legal highs scene (Morris and Wallach,
2014; NIDA, 2015).
As NMDAR antagonists, primarily ketamine, are also
widely used medically, their effects on the nervous
system have been studied extensively, yet they remain
enigmatic because of their widely differing and dosedependent effects. PCP is known to produce psychotomimetic effects in humans and it was first proposed to
be a candidate for a model of drug-induced schizophrenia in the late 1950s (Domino and Luby, 2012; Luby
et al., 1959). This finding later opened the idea of
glutamatergic model of schizophrenia in the past two
decades (Javitt and Zukin, 1991; Javitt et al., 2012).
NMDAR antagonists have neuroprotective effects, but
they also have proven neurotoxic effects in rodents,
although their effects on nonhuman primates and
humans are still unclear (Low and Roland, 2004;
Shirakawa et al., 2013; Wang et al., 2013a). Much effort
has been put in to identify the putative mechanisms of
NMDAR antagonist-induced neuroplasticity (Olney
et al., 1989). What follows is a review of ketamine,
PCP, and MK-801 and their effects on neurogenesis,
neurotoxicity, and behavior.

944

Korpi et al.

1. Ketamine: a Dissociative Anesthetic with Rapid


Antidepressant Effects in Patients. Ketamine is a noncompetitive NMDAR antagonist and a dissociative anesthetic developed in 1962 (Jansen, 2000). Recently, it
has received a lot of attention for its rapid-onset antidepressant effects (Krystal et al., 2013). After the first
report on the surprisingly rapid antidepressant effect
after intravenous infusion of ketamine in depressed
patients (Berman et al., 2000), many reports reproduced
the results and demonstrated that the antidepressant
effects after a single dose (intravenous or intramuscular)
is sustained for 3 to 10 days (Diazgranados et al., 2010;
Zarate et al., 2012b; Chilukuri et al., 2014). Also repeated
infusions of ketamine produced long-term effects (aan het
Rot et al., 2010; Murrough et al., 2013; Diamond et al.,
2014). Because of these data, there is a strong interest to
understand the mechanisms of action of ketamine.
With regards to the neurochemical effects underlying
the antidepressive effects of ketamine, chronic ketamine
intake is associated with increases in serum BDNF in
humans (Ricci et al., 2011). Although one report in adult
rats and mice claimed that ketamine showed a transient
effect (at 50 mg/kg) in forced swim and tail suspension
tests without any persistent antidepressant effect (Popik
et al., 2008), many preclinical studies have recently
focused on understanding the molecular mechanisms
underlying the rapid-onset prolonged antidepressant
effects. These studies show that plasticity-related pathways, especially the BDNF signaling pathway, are
influenced by ketamine treatment (Duman and Voleti,
2012; Kavalali and Monteggia, 2012; Reus et al., 2014;
Yang et al., 2014; Zhou et al., 2014). These effects are
dependent on the mammalian target of rapamycin
(mTOR) signaling pathway (Li et al., 2010a), because
they can be blocked by rapamycin. This effect could be
mimicked by GluN2B subunit-containing NMDAR
antagonist Ro 25-6981 [(aR,bS)-a-(4-hydroxyphenyl)b-methyl-4-(phenylmethyl)-1-piperidinepropanol
maleate], which induced antidepressant-like behavioral effects. mTOR is a serine-threonine protein kinase,
the subtypes of which, mTORC1 and mTORC2, form
active protein complexes with Raptor or Rictor, respectively. mTORC1 via p70S6K promotes cell growth,
mRNA biogenesis, and the translation of ribosomal
proteins. mTORC2 via Rictor activates and phosphorylates Akt to enhance cell survival, relying upon PKCa
for cytoskeleton remodeling, and controlling cell migration through the Rac guanine nucleotide exchange factors
and through Rho signaling (Laplante and Sabatini, 2009;
Buffington et al., 2014; Maiese, 2014). A single dose of
ketamine was reported to upregulate BDNF via activation (phosphorylation) of 59 adenosine monophosphateactivated protein kinase in the rat HC (Xu et al., 2013).
Rats receiving ketamine (75 mg/kg) on postnatal day
7 (P7) developed cognitive impairments in adulthood
(Huang et al., 2012). These effects were associated with
suppressed expression of pPKC, pERK1/2, and Bcl-2 with

unchanged tPKC or tERK levels during postnatal phase.


Long-term ketamine treatment in adult mice may cause
delayed and persistent upregulation of subcortical
DAergic system via a mechanism involving the BDNF
pathway (Tan et al., 2012b). Whereas acute and chronic
treatment of ketamine caused an antidepressant-like
effect in the forced swim test, only chronic treatment
(0.5 mg/kg/day for 10 days) increased AMPAR/NMDAR
density ratio in the HC of the depressed female
Wistar-Kyoto rats (but not in control Wistar rats) at 20
hours after the treatment (Tizabi et al., 2012). Synaptic
potentiation appears to underlie the ketamine efficacy
also in depressive patients, as suggested by increased
cortical excitability measured by means of magnetoencephalography (Cornwell et al., 2012). Importantly,
AMPAR antagonists have been shown to prevent the
antidepressant effects of ketamine in rodents, suggesting
that AMPAR activation is required for more persistent
antidepressant effects that are at least partly mediated
by upregulation of mTOR and BDNF pathways (Koike
et al., 2011; Koike and Chaki, 2014; Zhou et al., 2014).
Stereomers (R,S) of ketamine and its metabolites
norketamine and (2S,6S)-hydroxynorketamine increase
phosphorylation of mTOR and its downstream targets
Akt, ERK1/2, 70S6K, and 4E-BP1 in a cell culture model
(Paul et al., 2014b). As mentioned earlier, the activation
of mTOR pathways is crucial for synaptic plasticity.
Pharmacokinetic studies in treatment-resistant bipolar
patients show that norketamine is the initial but not
the main metabolite and that other metabolites, namely
dehydronorketamine, (2S,6S;2R,6R)-hydroxynorketamine,
and hydroxyketamine are detected during the first
48 hours (Zhao et al., 2012). These metabolites may
play a role in the antidepressant effects. Intriguingly,
higher plasma levels of (2S,5S;2R,5R)-hydroxynorketamine
were related to nonresponse to ketamine in bipolar
disorder patients (Zarate et al., 2012a), some metabolites being perhaps responsible for psychotomimetic
effects of ketamine.
Ketamine and some metabolites have affinity to a7nAChRs at relevant concentrations and behave as negative allosteric modulators (Moaddel et al., 2013).
Furthermore, several neuropharmacological studies have
shown direct and indirect effects of ketamine on other
neurotransmitters. A very recent study showed that
ketamine treatment-associated restoration of plasticity
in the ventral subiculum-NAc shell is caused by activation of D1 receptors (Belujon and Grace, 2014). In vitro
receptor binding studies show that ketamine has very
similar affinity to NMDARs and the high-affinity state
of the D2 receptors (where it works as an antagonist)
and a slightly lower affinity to 5-HT2 receptors (Kapur
and Seeman, 2001, 2002; Seeman and Kapur, 2003;
Seeman et al., 2005). Ketamine (40 mg/kg) pretreatment
prevented bicuculline-induced seizures and reversed
changes in [3H]QNB binding to muscarinic receptors
induced by bicuculline, showing the roles of muscarinic

Drug-Induced Neuroplasticity

and GABAA receptors (Schneider and Rodriguez de


Lores Arnaiz, 2013). Indeed, ketamine, but not PCP or
MK-801, acts as a positive modulator at specific cerebellar extrasynaptic GABAA receptor populations (Hevers
et al., 2008). An earlier study showed that ketamine
inhibits muscarinic signaling in CHO cells as observed
from intracellular Ca2+ release in response to the agonist
acetyl-b-methylcholine (Durieux, 1995). Finally, ketamine has been shown to inhibit the inward pacemaker
current Ih, which is mediated by ketamine binding to
hyperpolarization-activated cyclic nucleotide-gated potassium channel 1 (HCN1) subunits in forebrain neurons
(Chen et al., 2009; Zhou et al., 2013). HCN1-KO mice are
less sensitive to ketamine-induced hypnosis (Chen et al.,
2009), indicating that this mechanism may explain the
high-dose anesthetic effects, but no data are available on
whether this mechanism is involved in other behavioral
or neuroplasticity-inducing effects of ketamine. It is
noteworthy that all these studies draw parallels to the
anesthetic dose levels and not the subanesthetic antidepressant dose levels.
Ketamine has strong and prolonged effects on neuronal synaptic physiology. In conscious rats, ketamine
induces synaptic depression in the Schaffer collateralCA1 pathway sustaining beyond 4 hours but normalizing by 24 hours (Duan et al., 2013). Ketamine effects on
memory impairment were anyway predicted to last for
days or even weeks, because the long-lasting synaptic
change is expected to influence protein synthesis and
gene expression. Ketamine and PCP prevent the LTP in
the CA1 region in response to cornu ammonis stimulation in the HC of anesthetized rats (Stringer and
Guyenet, 1983), and, in agreement, a recent in vitro
study showed that ketamine, at its antidepressant
concentration level, negatively modulated both LTP
and LTD in CA1 of juvenile rats (Izumi and Zorumski,
2014). LTD inhibition and enhancement of somatic
EPSPs occurred shortly after ketamine administration,
whereas modulation of LTP occurred after 2 hours or
later. Clinically relevant micromolar concentrations of
ketamine dose dependently decreased LTP without
affecting paired-pulsed facilitation in Schaffer collateralCA1 pyramidal neuron synapses in a mouse HC slice
preparation, but higher concentrations affected basal
excitatory synaptic transmission and presynaptic
volley amplitude (Ribeiro et al., 2014). In a model for
psychosis, administrations of ketamine (20 mg/kg s.c.
given six times, once every 2 hours.) on P4-7 impaired
LTP at 34 weeks of age when recorded in slices from
the anterior cingulate cortex, which was accompanied
by increased AMPAR-mediated excitatory transmission
and decreased GABAergic inhibitory transmission
(Wang et al., 2014).
Ketamine in high anesthetic doses is thought to be
neurotoxic in developing brain of rodents and nonhuman
primates and in selected regions of the mature nervous
system (Olney et al., 1989; Slikker et al., 2007; Wang

945

et al., 2013a), and chronic ketamine treatment in mice


(6 months of daily 30 mg/kg i.p.) caused persistent
damage in brain function as observed from deterioration
in neuromuscular strength and nociception (Sun et al.,
2011). A case report suggested that chronic ketamine use
could impair memory and synaptic plasticity (Jansen,
1990). Furthermore, one MRI study provided anecdotal
evidence of brain atrophy (affecting both gray and white
matter) in the prefrontal, parietal, occipital, limbic
cortical regions, brain stem, and striatum after 24 years
of use in some ketamine-dependent subjects, although no
clear conclusions can be drawn from this study because of
the lack of statistics and a control group (Wang et al.,
2013b).
The psychedelic effects of ketamine may cause psychologic dependence that is cyclical in nature like cocaine and
amphetamine but has been described to be devoid of any
clear withdrawal syndrome (Jansen and DarracotCankovic, 2001). However, recent studies have shown
withdrawal effects from ketamine use. Female ketamine
users may show more severe withdrawal-associated
cognitive impairment than male users (Fattore et al.,
2009; Chen et al., 2014). Furthermore, in humans the
abuse potential of oral ketamine (65 and 100 mg) is robust
enough to have it being suggested as a positive control for
evaluating abuse potential of other psychedelic compounds (Shram et al., 2011). Ketamine intake is higher
in rats that were transferred to SA chambers during the
test sessions than in rats that also were housed in those
chambers, indicating the significant influence of the
setting on ketamine taking (De Luca and Badiani,
2011). Withdrawal of ketamine after 5 days of high dosing
(twice daily 150 mg/kg oral doses) in male cynomolgus
macaques increased activity and stereotypic behaviors,
including head bobbing, repetitive grooming, repetitive
body/limb or hand movements, and repetitive nail biting
during the withdrawal period (Walgren et al., 2014).
Withdrawal syndrome has also been frequently reported
in rodents. Subchronic ketamine (10 mg/kg twice daily for
10 days) in rats causes persistent working memory
impairment in the delayed spatial win-shift task after
a withdrawal for 10 days (Enomoto and Floresco, 2009).
Chronic administration (100 mg/kg/day i.p. for 10 days) of
ketamine in mice produces hyperactivity, enhanced
immobility period in forced swim test, and reduced
latency period in passive avoidance test that persisted
at least for 10 days after the withdrawal (Chatterjee et al.,
2011). Notably, all these studies used high ketamine
doses and/or models of prolonged abuse, indicating
stronger neuroadaptation in that situation than in lowdose antidepressant treatment studies. In contrast, in
clinical settings ketamine has been shown to reduce
cravings in heroin addicted patients when dispensed in
combination with psychotherapy sessions (Krupitsky
et al., 2002).
In summary, ketamine produces strong dose-dependent
behavioral and neurochemical effects. At low doses, it

946

Korpi et al.

produces significant changes in gene expression, resulting in enhanced synaptic plasticity that is likely to be
related to its rapid-onset antidepressive action in both
animal models and humans. On the other hand, high
doses of ketamine impair neuroplasticity and may cause
neurotoxic effects. Finally, in terms of behavior, both
addictive and antiaddictive properties of ketamine have
been described. It is therefore of interest to compare and
contrast the effects of ketamine with that of the second
commonly used NMDAR antagonist, PCP.
2. Phencyclidine. PCP, an NMDAR antagonist and
dissociative anesthetic, long known for its delirium- and
thought disorder-inducing (cognitive deficits) properties (Javitt and Zukin, 1991; Domino and Luby, 2012)
can acutely dissociate the CNS from peripheral sensory
input and induce metabolic hypofunction in several
brain regions, including the FC, HC, and thalamus.
PCP acts on many mechanisms that have been reported
to underlie its neuropharmacological and neurotoxicological effects. At the receptor level, it has a high affinity
to NMDA, 5-HT2A, and D2 receptors (Yamada et al.,
1999; Kapur and Seeman, 2002; Seeman et al., 2005,
2009). Sigma-1 receptors are involved in cognitive
deficits associated with PCP treatment in rodents
(Hashimoto et al., 2007; Kunitachi et al., 2009). At the
functional level, PCP impairs LTP in both in vivo and in
vitro systems (Stringer et al., 1983), e.g., PCP (and
ketamine) prevents both LTP and LTD in rat perforant
path-dentate gyrus synapses (Tizabi et al., 2012).
NMDAR antagonists, such as 2-amino-5-phosphovaleric
acid, N-acetyl-aspartyl-glutamate, and PCP, prevent
both LTP and primed burst potentiation, produced by
variants of electrical stimulation modeling plasticity, in
the CA1 area of rat HC slices (Wiescholleck and
Manahan-Vaughan, 2013; Xu et al., 2013). Subchronic
PCP treatment to rats caused impaired LTP in the lateral
amygdala and Schaffer collateral to HC-CA1 pathway
and marked increases in the NMDAR and AMPARmediated currents in the lateral amygdala (Pollard
et al., 2012).
Effects of PCP in rodents are especially strong and
persistent when given during brain development. Administration of PCP to pregnant rats during gestational
days 1220 days caused a decrease in PCP binding sites
in the fetal brains (Ali et al., 1988, 1989). The prenatal
administration (embryonic days 1218) of PCP (5
20 mg/kg) caused an increased response to PCP-induced
ataxia and an increased number of [3H]PCP binding
sites in the striatum and cortex of the offspring (Fico
and Vanderwende, 1989). Early oligodendrocyte progenitor cell markers were decreased in rat brains prenatally exposed to PCP (10 mg/kg), suggesting that
surviving cells are arrested at an immature stage,
which might have implications for the role of glutamate
and glutamate receptors in white matter abnormalities
in neurodevelopmental disorders (Lindahl et al., 2008).
These findings were strengthened by another recent

report on prenatally PCP-treated mice, demonstrating


aberrant gene expression (Notch2 and Ntn1) and
consequential decrease in the density of glutamatergic
neurons in the PFC, deficits in cognitive memory, and
sensorimotor gating at adulthood (Toriumi et al., 2012).
Prenatal PCP treatment (20 mg/kg) in mice induced
hypersensitivity to a low dose of PCP (3 mg/kg) in
locomotor activity and impairment of recognition memory in the novel object recognition test at the age of
7 weeks and reduced the phosphorylation of GluN1
subunit, although it increased the expression of this
subunit (Lu et al., 2010b). Prenatal PCP (5 mg/kg)
treatment caused impairment in recognition memory in
a novel object recognition test, enhanced locomotion in
a forced swim test, reduced the extracellular glutamate
level, and increased the expression of a glial glutamate
transporter in the PFC at 8 weeks of age (Lu et al., 2011).
Likewise, prenatal administration of PCP (10 mg/kg)
showed disruption of acquisition of passive avoidance
response and pole-climbing avoidance response
(Nabeshima et al., 1988). 5-Bromo-2-deoxyuridinelabeled newborn cells in the dentate gyrus granule cell
layer were significantly increased in rat offspring that
were subjected to prenatal exposure of PCP (Tanimura
et al., 2009). Prenatal and neonatal exposure of PCP
caused increased Bmax for muscarinic cholinergic receptors in the HC and reduced performance in HC-related
radial arm and water maze tasks (Yanai et al., 1992).
A single administration of PCP caused an increase in
cataleptic freezing and sporadic recurrences of backing
up and weaving behavior for up to 21 days in adult rats
(Haggerty et al., 1984). In rats, PCP initially also
reduces the brain weight, and, interestingly, withdrawal from PCP causes a transitory rebound associated with increased molecular layer depth and number
of synapses in the occipital cortex, suggesting a transitory burst of synaptogenesis (Brooks et al., 1997). After
a 72-hour withdrawal from PCP, sensitized locomotor
activation, a marked increase in GluN1 subunit mRNA
in several brain regions, including the FC and anterior
striatum, were detected (Wang et al., 1999). Furthermore, NMDAR-stimulated DA release was not affected
by PCP treatment, but the inhibition of this release by
NMDAR blockers (PCP, 7-chlorokynurenic acid, and
DL-2-amino-5-phosphovaleric acid) was blunted by
chronic PCP, indicative of tolerance. Most likely, PCP
treatment alters the subunit stoichiometry of NMDARs
that may contribute to behavioral plasticity (Wang
et al., 1999). Although there seems to be a discrepancy
between the persistent behavioral deficits (rotarod and
open field tests) and the amount of cell loss in relation to
the time of maximal prenatal PCP effect, exposure to
PCP increases agoraphilic cells in the entorhinal cortex
and subiculum, but reduces them in the ventromedial
hypothalamus, indicating that neurodegenerative and
antiapoptotic effects are dependent on the brain region
(Jebelli et al., 2002). Postnatal (P515) phencyclidine

Drug-Induced Neuroplasticity

disrupts water maze performance in adult rats, which


was associated with maximal [3H]MK-801 binding
(Bmax) in the HC and FC (Sircar, 2003), although similar
treatment reduced NMDA receptor expression in juvenile rats. Male rats treated with PCP (8.7 mg/kg s.c.) on
P7, 9, and 11 showed slight impairment in performing
a spatial reference memory task and strong impairment
in reversal and spatial working memory tasks (Andersen
and Pouzet, 2004). A similar perinatal treatment
schedule with 10 mg/kg of PCP induced disruption of
cognition as tested in progressive ratio operant schedule
task beyond postnatal 120 days (Wiley and Compton,
2004). PCP administration (10 mg/kg) on P2, 6, 9, and 12
caused a constellation of anatomic changes in rats at
early adulthood (P70), because it reduces densities of
principal neurons in the CA3 and DG of the HC, those
of interneurons in all cortical and HC regions, density
of reelin- and somatostatin-positive cells, and increases
the number of perisomatic inhibitory terminals around
the principal cells in the motor cortex and HC dentate
gyrus (Radonjic et al., 2013). These changes were
correlated with increased expression of neuregulin-1,
a protein involved in synaptic development and schizophrenia (Mei and Xiong, 2008; Shamir et al., 2012), in
the cortex and HC.
Subchronic PCP treatment (twice a day for 7 days)
followed by 7 days of abstinence causes significant
impairment in the reversal phase of a operant reversal
learning paradigm, demonstrating a persistent cognitive deficit (Abdul-Monim et al., 2006). The deficit was
reversed by acute treatments with atypical antipsychotic compounds ziprasidone, olanzapine, and clozapine.
PCP withdrawal impaired performance in object-placecontext task in rats, which was reversed by donepezil
but not by clozapine (Le Cozannet et al., 2010). Subchronic PCP treatment induced a persistent deficit in
novel object recognition that was attenuated by lurasidone (a potent 5-HT1A partial agonist, 5-HT2A antagonist), tandospirone (weak D2 antagonist), and F15599
(selective postsynaptic 5-HT1A agonist), and interaction
studies with these various ligands showed that 5-HT1A
agonism was adequate to attenuate the PCP-induced
effects (Horiguchi and Meltzer, 2012). These deficits
were also reversed by D1R agonism, the effect of which
was facilitated by 5-HT1A and mGlu2/3 receptor signaling (Horiguchi et al., 2013). Recently, Lu AE58054 (2-(6fluoro-1H-indol-3-yl)-N-[[3-(2,2,3,3-tetrafluoropropoxy)
phenyl]methyl]ethanamine), a 5-HT6 receptor antagonist, was found to reverse subchronic PCP-induced
deficits in novel object recognition test (Arnt et al.,
2010). It is not known whether any of the drug treatments can produce persistent reversal of subchronic
PCP effects, but, interestingly, clozapine and haloperidol prevented the metabolic hypofunction and the
reduced Pv expression in several brain regions in rats
treated intermittently with PCP (Cochran et al.,
2003).

947

In the PFC of rats, acute PCP treatment caused


significant changes in about 500 genes, including many
that are associated with neural plasticity (Kaiser et al.,
2004). Single and repeated administrations in adult
rats causes transient BDNF upregulation in the corticolimbic system, and subchronic postnatal administration
of PCP increased BDNF (protein and mRNA) in the HC
and entorhinal cortex that sustained till 8 weeks of age,
being responsible for the persistent decrease in the
prepulse inhibition in adults (Takahashi et al., 2006).
PCP treatment (20 mg/kg/day) for 14 days caused
reduction in the levels of pERK1/2, pCaMKIIa, and
pCREB in the PFC (Molteni et al., 2008). Two injections
of PCP reduced Pv and GAD67 levels in the PFC, and
this effect persisted after 10 days (Amitai et al., 2012). A
single injection of PCP (10 mg/kg) to rats on P7 caused
selective reduction in Pv-containing, not calretinin- or
calbindin-containing, interneurons in cortical areas in
early adulthood (P56) (Wang et al., 2008a). A persistent
decrease in the number of PFC spine synapses, in
parallel to the cognitive deficits, was evident with twice
daily injection of 5 mg/kg of PCP for 7 days (Elsworth
et al., 2011b). These PCP effects on gene regulation are
important examples of the widespread actions of PCP
on neural tissue, which appear to produce long-term
changes in the nervous system. An in vitro study
employing cortical neuronal culture showed that PCP
causes a transient increase within 3 hours followed by
a decrease in the intracellular BDNF, whereas activation of TrkB receptors and downstream cascades
(MAPK/ERK1/2 and PI3K/Akt) were decreased. The
number of synaptic sites and expression of synaptic
proteins were reduced 48 hours after treatment, without affecting the cell viability (Adachi et al., 2013).
Interestingly, exogenous BDNF completely reversed
the downregulating effect of PCP on the TrkB signaling
and synaptic protein expression (Adachi et al., 2013).
Postnatal treatment (P7, 9, and 11) with PCP (10 mg/kg)
reduced Nrg1 and erbB4 expression and phosphorylated Akt during development in the cerebral cortex (du
Bois et al., 2012) and displayed hyperactivity in holeboard and forced swim tests in adulthood (du Bois et al.,
2008).
Cognitive impairment and anatomic and neurochemical effects after PCP administrations have also been
reported in mice. Repeated administration of low doses
(0.52.0 mg/kg) of PCP causes cognitive impairment
associated with alterations in Arc and spinophilin
mRNA expression in the HC, striatum, and retrosplenial cortex (Beraki et al., 2009). Repeated PCP treatment (10 mg/kg for 10 days) decreased the expression of
Pv and synaptophysin mRNA in the mouse PFC and
increased basal Arc mRNA expression (Thomsen et al.,
2009). Lastly, in primates, chronic PCP (0.3 mg/kg twice
a day for 14 days) caused a 40% reduction in Pvcontaining axo-axonic structures in the PFC (Morrow
et al., 2007). PCP treatment in primates also leads to

948

Korpi et al.

persistent deficits of cortical DA (for 9 days) (Jentsch


et al., 1999). Subchronic PCP treatment (0.3 mg/kg
twice a day for 14 days) caused a reduction in the
number of asymmetric spine synapses in layer II/III of
the dorsolateral PFC that may perhaps account for
the cognitive dysfunction (Elsworth et al., 2011a).
Thus, persistent effects of PCP appear to rely at least
in part on its cortical neurotoxicity and neuregulin
downregulation.
In terms of addiction potential, PCP is also abused by
humans (Crider, 1986). Rats self-administer PCP as well
as MK-801 and the competitive NMDAR antagonist
3-((+/2)2-carboxypiperazin-4yl)propyl-1-phosphate into
the NAc and PFC in a DA-independent manner, unaffected by D2 receptor blockade by sulpiride (Carlezon
and Wise, 1996). Many NMDAR antagonists have often,
but not in all studies, dose dependently induced CPP in
rodents (Tzschentke, 2007) and PCP, at least somewhat,
lowered the threshold for lateral hypothalamic selfstimulation (Kornetsky and Esposito, 1979).
PCP, like ketamine, has strong effects on synaptic
plasticity and is capable of altering LTD and LTP as
well as the expression of genes known to be involved in
neuronal plasticity, effects that are often associated
with behavioral abnormalities and changes in cognition
and prepulse inhibition as well as the development of
SA of the drug. This leads to the question of how PCP
and ketamine compare with the prototypical and highly
selective NMDA antagonist MK-801.
3. Dizocilpine, a Prototypic Noncompetitive N-methylD-aspartate Receptor Antagonist. Unlike ketamine
and PCP, the recreational use of dizocilpine (MK-801)
has been very limited, although it has been reported to
have clear psychoactive effects comparable with other
NMDAR antagonists, suggesting that MK-801 has
somehow remained undesirable or unobtainable for
prospective users (Morris and Wallach, 2014). However,
it is the most selective of the noncompetitive inhibitors
of NMDARs and is known to cause behavioral sensitization (Vanderschuren et al., 1998).
The long-lasting and persistent effects of MK-801 on
synaptic plasticity and memory in rodents have been
recently reviewed (Wiescholleck and ManahanVaughan, 2013). Administration of MK-801 to rats causes
facilitation of LTP at the HC CA1-subiculum synapses 24
hour after treatment, suggesting metaplasticity (Buck
et al., 2006). However, a 4-week treatment reversed the
magnitude of LTP to control level. Intrathalamic administration of MK-801 changes local field potentials and
paired-pulse facilitation in rats, indicative of changes in
short-term plasticity (Kiss et al., 2011). After application
of MK-801, homosynaptic LTP of the direct cortical input
to CA1 was abolished for at least 1 week, with partial
recovery after 4 weeks (Whrl et al., 2007). MK-801
treatment causes long-lasting potentiation in HC-mPFC
pathway and impairments in mPFC-dependent cognitive
flexibility and HC-mPFC-dependent working memory in

rats, which were reversed by mGlu2/3 receptor agonist


LY379268
[(1R,4R,5S,6R)-4-amino-2-oxabicyclo[3.1.0]
hexane-4,6-dicarboxylic acid] (Blot et al., 2013). Thus, it
is important to take into consideration the strong effects
of MK-801 treatments themselves on neuroplasticity and
metaplasticity, when one wants to assess the roles of the
NMDARs in plasticity processes by using this antagonist.
Furthermore, in rodents, high neurodegenerative doses
of MK-801 induce prolonged hypothermia, which can
acutely affect cognitive behavior but may not as such have
lasting impairing effects on cognition (Zajaczkowski et al.,
2000).
A small study showed that primates with the history
of PCP use, but not those with a history of cocaine use,
self-administered MK-801 (Beardsley et al., 1990;
Steinpreis et al., 1995). MK-801 also reversed the
withdrawal effects of PCP (Carroll et al., 1994). In
rodents, MK-801 elicits CPP by itself (Layer et al.,
1993). Importantly, it inhibits CPP induced by METH
(Kim and Jang, 1997) and morphine (Tzschentke and
Schmidt, 1995; Kim et al., 1996) and blocks the
persistent neuroplasticity in VTA DA neurons and the
blunting of d-AMPH sensitization after a single dose of
BZs (Heikkinen et al., 2009; Panhelainen et al., 2011).
Interestingly, combination treatment of the rewarding
doses of NMDA and MK-801 did not result in CPP
(Panos et al., 1999). MK-801 significantly reduces fixed
ratio maintained operant responding for cocaine (Pierce
et al., 1997), but it can reinstate extinguished cocaineseeking behavior (De Vries et al., 1998) and disrupt the
reconsolidation of a cocaine-associated memory for CPP
(Brown et al., 2008).
MK-801 produces variable effects on DA release,
because systemic injection of MK-801 significantly suppressed striatal DA release in conscious rats, which was
not related to its behavioral effects (Kashihara et al.,
1990). Infusion of MK-801 into the NAc prevented
intrapallidal morphine-induced NAc DA release
(Anagnostakis et al., 1998). Its systemic administration,
however, increased DA release in the NAc, which could
be prevented by prazosin, an a1 adrenoceptor antagonist (Mathe et al., 1996). Systemic administration or
infusion into the PFC, ventral pallidum, NAc, or
striatum increased local DA release (Wedzony et al.,
1993; Hondo et al., 1994; Kiss et al., 1994; Kashiwa
et al., 1995; Schmidt and Fadayel, 1996; Yan et al.,
1997; Mathe et al., 1999; Kretschmer, 2000). In rodent
striatal slices, MK-801 attenuated ischemia-, hypoxia-,
hypoglycemia-, and L-glutamate-induced DA release
(Milusheva, 1992; Kim et al., 1995; Antonelli et al.,
1997), suggesting neuroprotective effects. MK-801induced behavioral sensitization was associated with
changes in the kinetics of DA release characterized by
a delayed (after 2 hour) release in the PFC (Cui et al.,
2014). But MK-801 failed to alter the striatal DA release
caused by METH (Baldwin et al., 1993), and a few other
reports showed that MK-801 had a weak or no effect on

Drug-Induced Neuroplasticity

pallidal, accumbal, or prefrontal cortical DA release


(Wedzony et al., 1993; Saulskaya and Marsden, 1995;
Wheeler et al., 1995; Druhan et al., 1996; Takahashi
et al., 1996; Anagnostakis et al., 1998; Callado et al.,
2000). MK-801 treatment reversed the decrease in
striatal DA, DOPAC (3,4-dihydroxyphenylacetic acid),
and HVA (homovanillic acid) caused by METH (Boireau
et al., 1995), and d-AMPH pretreatment prevented MK801-induced striatal DA release (Nash and Yamamoto,
1993). In summary, although MK-801 has the potential to
facilitate the release of DA, its effects are complex, most
likely due to different roles of the NMDARs in different
synapses, neuroplasticity, and regulation of neuronal
circuits composed of inhibitory and excitatory neurons
and glial cells, all being responsive to glutamate.
MK-801 has profound effects on gene expression in
the rodent brain. For example, it decreases BDNF
mRNA in the HC and superficial layers of cerebral
cortex and increases BDNF mRNA in the middle layer
(single cells) of cerebral cortex, midline thalamic
nuclei, and retrosplenial and medial entorhinal cortices
(Castren et al., 1993). The treatment increases DNA
binding activity of the activator protein-1 complex,
mRNAs of BDNF (Linden et al., 2000), neuronal Src in
superficial layers of the parietal cortex, tyrosine phosphorylation of GlunN2A subunit (Linden et al., 2001),
c-Fos, DFosB, Fra-2, and JunB expression in the entorhinal cortex (Kontkanen et al., 2000; Vaisanen et al.,
2004), possibly related to psychotic effects. Expression
profiling analysis showed that MK-801 regulates ER
proteins Erp29, a stress-inducible liver protein, and (at
higher doses) RTN1, a reticulon, and ABC transporters
involved in small molecule transport, reflecting possible
toxicity in the cingulate and entorhinal cortices (Toronen
et al., 2002).
The mechanisms involved in MK-801-induced neurotoxicity have been described and include 1) in the adult
brain, occurrence of paradoxical neurodegeneration,
disinhibition of glutamatergic and cholinergic projections of the cerebral cortex due to blockade of NMDARs
on inhibitory neurons in many subcortical regions
(Farber et al., 2002; Li et al., 2002), and 2) in the
developing brain, apoptotic neurodegeneration (Olney
et al., 1989; Dzietko et al., 2004) that is also reported for
ketamine (Zou et al., 2009; Liu et al., 2011) and PCP
(Rahbar et al., 1993; Deutsch et al., 1998). The MK-801induced toxicity in adult rats can be prevented by
injections of scopolamine and NBQX (6-nitro-2,3dioxo-1,4-dihydrobenzo[f]quinoxaline-7-sulfonamide)
(into the cerebral cortex) and clonidine (into basal
forebrain) (Farber et al., 2002) and by systemic scopolamine (Shirakawa et al., 2013), erythropoietin (partial
effect, counteracting the reduction in BDNF and GNDF
expression by NMDAR antagonists)(Dzietko et al.,
2004), and ethanol (Farber et al., 2004), but not by
increased BDNF (Vaisanen et al., 2003). Neurotoxicity
by PCP and ketamine can also be attenuated by

949

diazepam and haloperidol (Nakao et al., 1996; Nakki


et al., 1996). It is noteworthy that ethanol-antagonism
of NMDARs induces more neurotoxicity in developing
than adult rat brain, presumably due to the inhibitory
GABA mimetic effects of ethanol in adult brain (Olney
et al., 2000, 2002). It may also reflect the role of the
NMDARs in neuronal development and excitotoxicity,
both of which can be blocked in cultured neurons by
ethanol (Wegelius and Korpi, 1995).
In summary, MK-801 as a prototypic NMDAR antagonist produces widespread changes in brain gene
expression, which might have prolonged effects on
neuronal function and structure. The high-dose toxicity
of MK-801 and other NMDAR antagonists is associated
with brain region-dependent and time-dependent astrogliosis and microglial activation (Fix et al., 1996; Nakki
et al., 1996). On the other hand, recent work has also
revealed that MK-801 and other NMDAR antagonists
attenuate microglial activation induced by various
drugs and insults (Thomas and Kuhn, 2005; Nair and
Bonneau, 2006), which may at least partly explain the
neuroprotective effects of these drugs. Interestingly,
MK-801 and cocaine have distinct effects on gene
expression, with the interactions between their effects
being complex (Storvik et al., 2006).
4. Conclusions. Potent NMDAR antagonists are
drugs that affect neuroplasticity processes in the brain,
lead to addiction, and suffer from induction of psychotomimetic effects and thought-impairing cognitive deficits
in many subjects, with a high potential to induce
paradoxical neurodegeneration. Ketamine and its analogs are now studied for the rapid, long-lasting antidepressant effect, with early results indicating strong
activation of neuronal plasticity mechanisms. Phencyclidine and dizocilpine are used in preclinical settings to
induce a model of adult psychosis by adolescent exposure or to assess the role of NMDARs in behavior and
physiologic functions, respectively.
G. Opioid-Induced Neural Adaptations
Opioids are widely used as powerful analgesic agents
that are also highly addictive. A sizable proportion of
individuals that either misuse opioids or use them
clinically become addicted. With chronic use, adaptive
neural mechanisms are initiated at various levels of
neural organization, leading to both short-term and
protracted changes in the functioning of opioid-sensitive
neurons and neural networks (Williams et al., 2001;
Christie, 2008; Dacher and Nugent, 2011b; Lutz and
Kieffer, 2013). Some of these adaptations, such as opioid
tolerance, have been intensively investigated. However,
the neural mechanisms underlying protracted features of
withdrawal, such as opioid seeking and relapse contributing to the compulsive features of opioid dependence,
still remain largely unknown. In the following sections,
we will focus on describing the most important findings
related to chronic opioid use, including the mechanisms of

950

Korpi et al.

opioid tolerance, the withdrawal syndrome, and opioidinduced synaptic plasticity.


Opioids exert their actions through three major
receptor subtypes, m, d, and k, opioid receptors that
were first identified by pharmacological tools and later
verified by molecular cloning (Goldstein and Naidu,
1989; Evans et al., 1992; Kieffer et al., 1992; Reisine and
Bell, 1993). Further diversification of opioid receptor
functions may be achieved by alternate splicing, posttranslational regulation, or receptor dimerization. All
opioid receptors are GPCRs, coupled with pertussis
toxin-sensitive G proteins Gi/o (Connor and Christie,
1999). The coupling profile to G proteins of all three
opioid receptors is similar, including inhibition of AC,
activation of potassium conductance, inhibition of calcium conductance, and inhibition of transmitter release
(Williams et al., 2001). Opioid receptors and their
mRNAs display distinct anatomic distribution in neuronal systems, as well as paracrine and exocrine tissues
important for opioid drug actions (Mansour et al., 1987,
1995; Zhu et al., 1998).
The primary target of most opioid drugs is the m
receptor, through which the addictive properties of
opioids are also mediated. The levels of neural adaptations to chronic opioids can be conceptualized as 1) loss
of the m receptor coupling to its cellular effectors, 2)
adaptations in the intracellular signaling cascades, 3)
systems feedback adaptations in neuronal and glial
networks, and 4) changes in synaptic plasticity (Christie,
2008).
1. Desensitization and Internalization in Opioid
Tolerance and Dependence. Development of opioid
tolerance is partly dependent on a partial loss of
capacity of the m receptor to signal to its intracellular
effectors (Kieffer and Evans, 2002; Christie, 2008;
Allouche et al., 2014). This "receptor tolerance" can be
achieved by different mechanisms, primarily by decreased cell surface expression of m receptors and/or
attenuation of coupling efficacy of the surface receptors.
Because chronic morphine administration generally
produces little changes in either m receptor mRNA or
protein expression (Patel et al., 2002), most research
has focused on the reduced coupling of m receptor to its
effectors. However, considering the limited reduction in
potency of opioid agonists in functioning neurons after
prolonged morphine exposure (Williams et al., 2001), it
does not seem likely that behavioral opioid tolerance
can fully be ascribed to decreased receptor coupling.
Tolerance at systems levels also involves interaction of
mechanisms of tolerance at molecular, cellular, and
neural network levels. The magnitude of m receptor
uncoupling from G protein activation has been found to
differ in different neuron types, where m receptors are
coupled with diverse signaling systems (Sim-Selley
et al., 2000).
For receptor uncoupling, the mechanism of opioid
receptor desensitization and internalization are

assumed to resemble the model established for the


b2-adrenergic receptor. In this model, agonist activation
induces m receptor phosphorylation by G proteincoupled receptor kinase 2, which increases the affinity
of the receptor to arrestin 3 (Arr3). The binding of Arr3
uncouples the receptor from G protein signaling (desensitization) and initiates the receptor sequestration
and internalization by an Arr3- and dynamindependent mechanism (Gainetdinov et al., 2004). The
importance of this sequence of events was tested using
Arr3-KO mice (Bohn et al., 2000). In these mice, chronic
morphine treatment did not induce desensitization of
the m receptor and failed to produce either acute or
chronic antinociceptive tolerance. However, these mice
still became physically dependent on morphine and
exhibited upregulation of AC activity that has been
used as a cellular marker of opioid dependence (see
below), suggesting that antinociceptive tolerance and
physical dependence could have different underlying
mechanisms. Although homologous desensitization
thus seems to have a role in morphine tolerance at the
receptor level, the importance of internalization is far
from clear. Compared with other efficacious m receptor
agonists, such as [D-Ala2, NMe-Phe4, Gly-ol5]-enkephalin
(DAMGO), methadone, or sufentanyl, morphine has a low
efficacy for producing receptor internalization (Whistler
et al., 1999; Borgland et al., 2003), consistent with its
profile as a partial agonist. Therefore, the effects of
chronic morphine on desensitization are probably more
important than internalization for the functional uncoupling of the receptor.
2. Cellular Signaling in Opioid Tolerance and
Dependence. Behavioral tolerance can also result from
mechanisms subsequent to receptor activation that
adapt to restore function in the presence of drug.
Continued presence of an inhibitory agonist leads to
homeostatic compensation of the downstream signaling
systems. Subsequent removal of the agonist can then
lead to a rebound activation of the signaling systems
and cause neural excitation. This overshoot may underlie the initial phases of withdrawal, but could also
lead to more persistent effects of signaling systems that
are reflected in synaptic events or structural changes.
Particularly interesting for opioid-induced alterations
are the signaling systems directly downstream from m
receptor activation, such as the AC-cAMP-PKA cascade
and the MAPK cascades (Christie, 2008). Most likely,
however, the mechanisms underlying chronic opioid
effects are not limited to signaling systems directly
downstream of receptor activation, but could reflect
general homeostatic compensations in response to
enhanced neural inhibition by opioid agonists.
The AC-cAMP-PKA cascade is a well-studied example of the tolerance resulting from compensation. Acute
opioids inhibit AC, but depressed concentrations of
cAMP return to normal during chronic opioid exposure
as the AC activity is upregulated. Upon removal of the

Drug-Induced Neuroplasticity

agonist, this compensatory increase persists and has


therefore been regarded as a cellular marker of withdrawal (Sharma et al., 1975; Williams et al., 2001). This
general picture is complicated by the existence of
several isoforms of AC that could be regulated by
distinct mechanisms (Watts and Neve, 2005). Three of
these isoforms, ACI, ACII, and ACV, are neuronal and
they are highly localized to synapses (Mons and Cooper,
1995). In cell lines and expression systems, ACI and
ACV isoforms were supersensitized by chronic opioid
treatment, but at neuronal tissue levels, the magnitude
of upregulation is rather small, which could be due to
the heterogeneity of cell types, presence of many opioid
receptor subtypes, and AC isoforms (Williams et al.,
2001). However, there is abundant evidence that the
opioid-induced upregulation of the AC-cAMP-PKA cascade in specific synapses in different brain areas could
contribute to neural and behavioral adaptations associated with morphine withdrawal.
Adaptations of the NEergic LC have been implicated
in opioid withdrawal behavior. Acute opioids decrease
LC neuronal activity and inhibit the cAMP signaling
pathway (Duman et al., 1988). Chronic opioids produce
tolerance, demonstrated by upregulation of cAMPdependent PKA during chronic morphine treatment
(Ivanov and Aston-Jones, 2001), suggesting that this
signaling pathway is involved in the withdrawalinduced hyperactivity of LC neurons. The LC hyperactivity may influence the activity of several other brain
areas through widespread efferent projections. The
phosphorylation state of CREB exerts transcriptional
control over a large number of neurotransmitter receptors and signaling molecules. In the LC, overexpression
of the constitutively active CREB mutant aggravated
morphine withdrawal symptoms, whereas the dominantnegative CREB mutant suppressed them (Han et al.,
2006). CREB overexpression did not affect neuronal
firing at baseline, but it enhanced the effects of the AC
activator forskolin, suggesting sensitization of the
cAMP pathway. Therefore, morphine-induced increase
in CREB activity could contribute to adaptations underlying morphine dependence and withdrawal. Similar to the LC, physiologic and biochemical studies have
indicated a role for the periaqueductal gray area (PAG)
in the expression of several withdrawal signs. Because
the inhibition of GABAergic transmission by opioids
causes disinhibition of PAG projection neurons, leading
to activation of the descending pathways, adaptations
in the PAG could be associated with the somatic and
aversive withdrawal symptoms (Bandler and Shipley,
1994). In line with this idea, opioid withdrawal increased inhibitory transmission mediated by GABAA
receptors, which was blocked by PKA inhibitors and
occluded by cAMP analogs (Ingram et al., 1998). It
therefore seems that in the PAG, opioid dependence
reflects m receptor coupling to presynaptic inhibition
in GABAergic nerve terminals via AC- and PKA-dependent

951

processes, with the withdrawal being associated with the


abrupt increase in GABA release.
Opioid-induced adaptations in the nuclei of the
mesolimbic DA pathway, such as the VTA and NAc
that lead to alterations in the DAergic tone, could
potentially modulate the motivational aspects of opioid
withdrawal. In the VTA of morphine-treated guinea
pigs, both inhibitory postsynaptic currents and the
potency of forskolin and the cAMP-dependent protein
kinase to increase these currents were augmented
(Bonci and Williams, 1997), suggesting upregulation
of the cAMP-dependent cascade. Similarly, in the NAc
during withdrawal from chronic morphine, forskolin
activation increased inhibitory transmission, suggesting that opioid dependence induced an opioid-sensitive
cAMP-dependent mechanism that regulates transmitter release and could contribute to opioid withdrawal
(Chieng and Williams, 1998).
3. Novel Mechanisms in Between-systems Adaptations.
Because chronic opioid administration produces alterations in the excitability in opioid-sensitive neurons,
these changes can indirectly induce adaptations in
other neurons and synapses in the circuitry. For
example, this is illustrated by opioid-induced adaptations in the VTA DA neurons that are not primary
targets of opioids. It has also been postulated that
specific systems that functionally oppose the m opioid
receptor signaling could modulate opioid tolerance and
dependence. One of these systems is the cholecystokinin
(CCK) system that has been studied especially in
relation to analgesic effects of m-opioid receptor stimulation (reviewed in Wiesenfeld-Hallin et al., 1999). For
example, CCK peptides, G protein-coupled CCK1, and
CCK2 receptors show complex expression patterns
and interact with a number of neurotransmitter/
modulator mechanisms to both oppose and reinforce
opioid-induced analgesia and anxiolysis within the
PAG (Mitchell et al., 2011). CCK is also strongly
expressed in the VTA (Ballaz et al., 2013) and can
influence drug-induced reward processes. Indeed,
CCK infusion into the NAc attenuated intracranial
electrical self-stimulation in the VTA (Vaccarino and
Koob, 1984). Systemic CCK2, but not CCK1, receptor
antagonism attenuated expression of morphineinduced CPP (Lu et al., 2000; Mitchell et al., 2006)
and naloxone-induced withdrawal symptoms (Lu
et al., 2000). Furthermore, the CCK2 antagonism in
the anterior NAc attenuated morphine-induced CPP,
whereas that in the posterior NAc enhanced the CPP
in a manner dependent on D2 receptors (Mitchell
et al., 2006).
Another important one of these modulating systems
is the dynorphin/k-opioid receptor system that has been
suggested to participate in mediating opposing alterations in behavior and brain neurochemistry that occur
in response to repeated drug use, particularly stimulants and alcohol (Shippenberg et al., 2007). Therefore,

952

Korpi et al.

aberrations in this opioid system could also contribute


to the neural dysregulations underlying the development of addiction. In addition, increasing evidence is
accumulating that the opioid receptor-like NOP receptor and its endogenous ligand N/OFQ contribute to
changes in plasticity observed with morphine tolerance
and dependence. In particular, increased antiopioid
activity of the nociceptin system was suggested to
participate in the development of opioid tolerance and
dependence, because either genetic or pharmacological
blockade of NOP signaling prevented them (Ueda et al.,
2000; Chung et al., 2006).
Glial activation is also emerging as a novel modulator
of opioid effects, although glia have mainly been studied
in the context of pain at the spinal level. The glial cells,
particularly astrocytes and microglia in this context,
release several substances that can modulate glianeuron communication and shape neuronal plasticity
(Watkins et al., 2005). Chronic morphine treatment
induces glial activation in many brain areas, including
the spinal cord, cingulate cortex, hippocampus, and
PAG (Song and Zhao, 2001; Hao et al., 2011). Morphine
and other m receptor agonists bind to an accessory
protein of glial toll-like receptor 4 (TLR4), myeloid
differentiation protein 2, and thereby induce TLR4
oligomerization and trigger subsequent glia-mediated
proinflammatory responses. TLR4 activation causes the
release of proinflammatory and neuroexcitatory cytokines, such as tumor necrosis factor-a and interleukin1b (Wang et al., 2012b). Tumor necrosis factor-a has
been shown to modulate neuronal functions in various
ways, including an increase of AMPAR activity, spontaneous and evoked transmitter release, and cell surface expression of AMPARs and NMDARs (reviewed in
Ouyang et al., 2012). Both the opioid-active (2)-naloxone
and (2)-naltrexone and the opioid-inactive isomers
(+)-naloxone and (+)-naltrexone block activation of
TLR4 signaling, providing a means of specifically inhibiting glial activity by the nonopioid isomers (Watkins
et al., 2009). They have been shown to attenuate
morphine analgesic tolerance and withdrawal symptoms
(Hutchinson et al., 2010), as well as opioid addiction-like
behavior, such as morphine-induced CPP, remifentanil
SA (Hutchinson et al., 2012), and abstinence enhancement of cue-induced heroin seeking (Theberge et al.,
2013). However, the potential of (+)-naloxone and
(+)-naltrexone as TLR4 antagonists has been contested
on the basis of in vitro assays (Skolnick et al., 2014).
Further indications of the role of glial cells in morphine
behaviors come from interesting experiments by Schwarz
and coworkers (Schwarz et al., 2011; Schwarz and Bilbo,
2013), who first showed that neonatal handling of rats
(from P2 to P20) strongly decreased the methylation of
the gene encoding the cytokine IL-10 in microgliaenriched brain samples and increased IL-10 expression
in adulthood. This was negatively correlated with
morphine-induced reinstatement of CPP (Schwarz

et al., 2011). On the other hand, subchronic treatment


of rats with morphine in adolescence (P3742), but
not in early adulthood (P5458), strongly facilitated
morphine-induced reinstatement of extinguished CPP
18 days later (Schwarz and Bilbo, 2013). The reinstatement was correlated with increased TLR4 expression in
microglia, suggesting that adolescent morphine exposure produces long-term changes in microglial function
and in sensitivity to drug-induced relapse.
4. Opioid-Induced Changes in Excitatory and Inhibitory Synaptic Plasticity. The compulsiveness of
drug use has been conceptualized to be mediated by
pathologic forms of memory that are based on synaptic
adaptations resembling other forms of activitydependent plasticity, such as LTP and LTD (Kauer
and Malenka, 2007). For example, synaptic plasticity in
the VTA has been suggested to be a common cellular
substrate for abused drugs in the initial phases of drug
addiction. Thus, a single administration of morphine
also induces plasticity in the VTA DA neurons that lasts
long after the administration, as evidenced by increased
ratio of AMPAR/NMDAR currents, which is consistent
with morphine-induced LTP of glutamatergic synapses
onto DA neurons (Saal et al., 2003). In addition, there is
now increasing evidence for GABAergic inhibitory
plasticity after morphine administration. It has long
been known that DA-dependent opiate actions involve
VTA GABA neurons. By inhibiting local GABAergic
interneurons, opiates disinhibit DA neurons (Johnson
and North, 1992), which results in increased DA cell
firing and DA release in the VTA projection areas,
including the NAc.
GABAA receptor-mediated synaptic transmission
onto VTA DA neurons expresses long-term potentiation, known as LTPGABA (Nugent et al., 2007; Niehaus
et al., 2010). This form of LTP has been described as
heterosynaptic, because it requires postsynaptic
NMDAR activation but results from increased GABA
release at neighboring inhibitory nerve terminals.
Activation of NMDARs on postsynaptic DA neurons
produces the retrograde signaling molecule NO, which
in turn activates guanylate cyclase in presynaptic
GABAergic nerve terminals. Increased levels of guanylate cyclase promote potentiation of GABA release,
thereby initiating LTPGABA. Further work showed that
NO-cGMP signaling activates cGMP-dependent kinase
PKG and that NO-cGMP signaling potentiates GABAA
receptor-mediated, but not GABAB receptor-mediated,
synaptic responses (Nugent et al., 2009). Morphine in
vivo blocks LTPGABA by interrupting signaling from NO
to guanylate cyclase and therefore reduces normal
inhibitory control (Nugent et al., 2007).
In addition to LTPGABA, the same GABAergic synapses can also exhibit long-term depression (LTDGABA) in
response to 1-Hz LFS. This form of synaptic downregulation is not dependent on NMDA or CB1 receptors
but on D2-like receptor activation. This D2R activation

Drug-Induced Neuroplasticity

requires a neurochemical signaling via inhibitory


G protein-PKA-inositol phosphate receptor IP3R-Ca2+calcineurin-A-kinase anchoring protein 150 AKAP150
at GABAergic synapses (Dacher et al., 2013). Importantly, LTDGABA is abolished/occluded by a single dose
of morphine 24 hours before the measurements (Dacher
and Nugent, 2011a). Collectively, morphine-induced
changes in LTPGABA and LTDGABA represent forms of
inhibitory plasticity that together with changes in
excitatory synapses provide a mechanism to enhance
mesolimbic DA transmission and thereby increase the
incentive properties of morphine (Dacher and Nugent,
2011b).
More recently, Atwood et al. (2014) found in vitro LTD
of glutamate inputs to the dorsolateral striatum induced by agonists of m, d, and k receptors, with m and d,
but not k, agonists also inducing the LTD in the
dorsomedial striatum. The depression of synaptic responses could be blocked by receptor subtype-specific
antagonists, but once induced and the agonists washed
away, the LTD responses were insensitive to opioid
receptor antagonists. Importantly, a single in vivo dose
of oxycodone occluded the m receptor-LTD and
endocannabinoid-LTD for 2 days. These novel mechanisms may contribute to striatum-dependent behaviors,
such as habit learning.
5. Chronic Opioids and Synaptic Plasticity.
Plasticity associated with chronic opioid exposure could
underlie various aspects of opioid addiction, such as
sensitization, dependence, withdrawal, and opioid seeking. Opioid-induced plasticity in VTA excitatory and
inhibitory synapses could also provide mechanisms for
these chronic opioid effects. Work assessing the alterations in glutamate receptor subunit expression found
that continuously delivered morphine had no influence
on glutamate receptor levels. However, intermittent
morphine administration schedule that has been associated with behavioral sensitization elevated the VTA
GluA1 levels (Fitzgerald et al., 1996). Upregulated
GluA1 subunit levels could contribute to increased
excitability of VTA DA neurons, which was confirmed
by selective viral-mediated increase of GluA1 subunit in
the VTA bringing up a sensitized response to morphine in
locomotion and place preference conditioning (Carlezon
et al., 1997).
Because the pattern of VTA DA neuron firing depends
on excitatory glutamate drive, any adaptive changes in
excitatory synaptic transmission produced by chronic
morphine are reflected in the activity of DA neurons and
could thus also contribute to decrease of mesolimbic
DAergic neuronal activity observed in vivo in morphinewithdrawn rats. Manzoni and Williams (1999) found
that during acute withdrawal from chronic morphine
treatment there was a fundamental change in the
presynaptic regulation of the glutamate release at this
synapse. First, the transduction pathway responsible
for the opioid receptor-mediated inhibition was less

953

sensitive to agents that disrupt the normal inhibition,


including the potassium channels blocker 4-aminopyridine and the lipoxygenase inhibitor baicalein. Second,
they found that in withdrawn animals, the sensitivity of
presynaptic inhibitory control mediated by GABAB and
mGlu2/3 receptors was enhanced. Together, these findings suggested that one of the consequences of withdrawal from chronic morphine could be augmented
presynaptic inhibition of the excitatory inputs to the
VTA DA cells, which would add to the inhibition of these
neurons. Increased GABA release from interneurons
further potentiates the inhibition of VTA DA neurons
during withdrawal (Bonci and Williams, 1997). The
enhanced GABA release appears to be correlated with
physical dependence on morphine, as measured by
somatic withdrawal signs during withdrawal (Madhavan
et al., 2010). As previously discussed, morphine-activated
m receptors do not exhibit substantial endocytosis and
recycling. In a knock-in mouse, in which a genetic change
in the m receptors promotes morphine-induced receptor
desensitization and endocytosis in the VTA GABA interneurons, the increased GABA release onto VTA DA
neurons was abolished during withdrawal. These data
suggest that receptor trafficking is a compensatory cellular mechanism that could modulate synaptic plasticity
produced by chronic opioid administration (Madhavan
et al., 2010).
GABAAergic neurotransmission in the VTA has been
suggested to mediate the switch from the drug-naive to
the dependent and addicted state. A discrete population
of VTA GABAA receptors was shown to undergo a switch
from inhibitory to excitatory in heroin-dependent and
withdrawn rats (Laviolette et al., 2004). The precise
mechanism of this transformation is not known, but it
seems to be mediated by the BDNF, as demonstrated by
promotion of the switch by BDNF injection into the VTA
and attenuation of the aversive motivational states of
heroin withdrawal by local knockdown of the BDNF
receptor TrkB expression (Vargas-Perez et al., 2009,
2014). The cellular phenotypes of the circuit components for this model require still further work in the
VTA and associated areas. More generally, these results
may relate to stress vulnerability and social defeat
stress, which affect drug sensitivities and motivation to
seek reward and drugs (Piazza et al., 1993; Shaham
et al., 1997; Kauer, 2003; Yap et al., 2005; Niehaus et al.,
2010; Ungless et al., 2010; Cabib and Puglisi-Allegra,
2012; McEwen, 2012; Wang et al., 2013c; Walsh et al.,
2014). Recently, BDNF signaling in the NAc has been
shown to be involved in susceptibility to social defeat
stress, actually mediating stress effects, such as social
aversion (Berton et al., 2006). Phasic optogenetic activation of VTA-NAc pathway had no effects on social
interaction and BDNF levels in the NAc in stress-naive
mice, but strongly reduced social interaction and increased BDNF in mice subjected to a subthreshold
social stress (Walsh et al., 2014), confirming that

954

Korpi et al.

stressful events induce prolonged changes in the mesolimbic DA pathway.


In addition to augmentation of VTA presynaptic
mGlu2/3 receptor functions in morphine-withdrawn subjects (Manzoni and Williams, 1999), chronic morphine
also alters mGlu2/3-mediated events in the NAc. Activation of these receptors induces LTD at NAc glutamatergic
synapses. However, this mGlu2/3-dependent LTD was
abolished during withdrawal from chronic morphine
(Robbe et al., 2002). Together with enhancement of VTA
presynaptic mGlu2/3 function and the ability of mGlu2/3
agonists to reduce morphine withdrawal symptoms,
these findings suggest that regulation of mGlu2/3
synaptic plasticity in the mesolimbic system could
be important for the development of opioid addiction.
Other cellular events underlying chronic opioid
actions in the NAc include trafficking of AMPARs in
the D1R-MSNs. Chronic morphine administration
was associated with a decrease in surface trafficking
of the GluA1 subunit, which could be a substrate of
morphine-induced alterations in neuronal activity
within the NAc (Glass et al., 2008).
The involvement of opioid receptors in the modulation
of hippocampal plasticity has been amply documented.
Chronic opioids produce cognitive deficits in human
opioid users and reduce neurogenesis in the rat HC
(Guerra et al., 1987; Eisch et al., 2000), suggesting that
maladaptive alterations in hippocampal plasticity could
partly contribute to opioid addiction, craving, and relapse. In accordance with this idea, chronic morphine or
heroin treatment attenuated the hippocampal CA1 LTP
during withdrawal and produced deficits in spatial
learning (Pu et al., 2002; Salmanzadeh et al., 2003).
Hippocampal LTP could be restored by re-exposure to
opioids, including the m receptor agonist [D-Ala2, NMePhe4, Gly-ol5]-enkephalin (DAMGO) and inhibitors of
PKA, suggesting that upregulation of the cAMP pathway was one of the underlying mechanisms (Pu et al.,
2002; Bao et al., 2007). In addition, chronic morphine
exposure increased hippocampal extracellular adenosine concentrations, which contributed both to the
inhibition of hippocampal LTP and impairment of
spatial memory retrieval, and both deficits could be
reversed by an adenosine antagonist or adenosine
deaminase (Lu et al., 2010a). There is also evidence to
suggest that opioid-induced changes in hippocampal
synaptic plasticity could have a role in drug-context
associations. Morphine-induced CPP was associated
with increased basal synaptic transmission and impaired hippocampal LTP, accompanied with increased
synaptic expression of the NMDAR subunits GluN1 and
GluN2B. After extinction of CPP, impaired LTP and
elevated GluN2B expression persisted, but morphineprimed reinstatement of CPP was associated with
LTP similar to controls (Portugal et al., 2014). These
findings implicate hippocampal LTP and glutamatergic
transmission in context-dependent morphine effects,

consistent with blockade of morphine CPP by GluN2Bselective antagonists (Ma et al., 2006).
In agreement with the hypothesized role of the
AMPARs in drug-induced plasticity, a 12-hour withdrawal from four escalating morphine doses (5, 8, 10, 15
mg/kg at 12-hour intervals) increased the expression of
GluA2-lacking (Ca2+-permeable) glutamate receptors
(containing GluA1 and/or GluA3 subunits) in mouse
hippocampal synaptic and extrasynaptic fractions, correlating with reduced magnitude of LTD in HC neurons
without changes in sensitivity to AMPA (Billa et al.,
2010). When this same morphine dosing was used in
a specific setting (locomotor activity chambers), the
treated mice challenged with 5 mg/kg morphine 1 week
later showed robust context-dependent locomotor sensitization in the same chambers compared with saline
treatment or unpaired morphine challenge (Xia et al.,
2011). Interestingly, this persistent effect correlated
with increased phosphorylation of GluA1 Ser845 and
impaired LTP in the HC. However, earlier work
revealed that GluA1-KO and GluA1-Arg (Ca2+ impermeable) mutant mice show enhanced context-dependent
sensitization to 3 mg/kg morphine in comparison with
wild-type littermates 5 days after 6-day treatment with
10 mg/kg morphine (Vekovischeva et al., 2001), indicating
that GluA1 subunits are not obligatory for contextdependent morphine sensitization and that their impaired function can be compensated. Indeed, although
a single dose of morphine induces persistent increase in
AMPAR/NMDAR current ratios in VTA DA neurons 24
hours after administration in wild-type mice (Saal et al.,
2003; Heikkinen et al., 2009), it fails to alter this ratio in
GluA1-KO mice that had already increased ratio at the
baseline (Aitta-aho et al., 2012). The lack of the VTA DA
neuron plasticity might correlate with impaired statedependent place preference, because the GluA1-KO failed
to show place preference under morphine testing (Aitta-aho
et al., 2012). Taken together, these data support the role
of the AMPARs in adaptation to opioid effects, although
the details of subunit roles are still poorly known and
most likely differ in different brain regions.
In the mPFC of rats with a history of heroin SA,
GluA2 subunit-containing AMPARs are downregulated
in the principal pyramidal cells, leading to reduced
AMPAR/NMDAR current ratios and increased rectification (Van den Oever et al., 2008), which permitted
heroin seeking in abstinent and extinguished rats by
a compound heroin-related cue. Blocking the GluA2
subunit downregulation (and subsequent LTD?) also
prevented relapse to heroin seeking. Furthermore,
exposure to heroin cues after abstinence almost doubled
the frequency of GABAergic spontaneous IPSCs in
mPFC pyramidal neurons in brain slices prepared 30
minutes after cue exposure (Van den Oever et al., 2010).
The heroin seeking was also associated with a reduction
in ECM proteins around the Pv-containing interneurons, suggesting that the PPN was downregulated to

Drug-Induced Neuroplasticity

enhance the inhibition of principal neurons. These data


indicate that mPFC output is reduced after relapseinducing cue exposure and that relapse to heroin
seeking is associated with altered structural and functional aspects of both glutamatergic and GABAergic
systems in the mPFC.
Drug-induced changes in the structural plasticity of
dendrites are potentially an important component
of synaptic plasticity. Basic research into the dynamics
of the dendritic spines suggests that the size and shape of
individual spines correlates with forms of synaptic
plasticity, such as LTP and LTD at glutamate synapses.
First, stabilization of the transient immature spines
into functional spines is activity dependent. Second,
experimental protocols that induce LTD are linked with
shrinkage or retraction of spines, whereas induction of
LTP enhances formation of new spines and enlargement of existing spines (Nagerl et al., 2004). In sharp
contrast to psychomotor stimulants that increase dendritic branching and density, chronic morphine exposure reduces the size of VTA DA neurons (Sklair-Tavron
et al., 1996; Russo et al., 2007; Mazei-Robison et al.,
2011) and the density of dendritic spines in brain areas
considered relevant for drug addiction, such as the NAc,
OFC, and HC (Robinson and Kolb, 1999b; Robinson
et al., 2002). Because stimulants and opioids similarly
induce transcription factors (e.g., CREB, DFosB) that
regulate expression of cytoskeletal proteins, the disparate findings on spine density are perplexing. It is
possible that stimulants and opioids recruit different
signaling pathways regulating synaptogenesis acting
under CREB or DFosB control or that additional pathways are involved (Russo et al., 2010). For example, the
reduced VTA DA neuron soma size induced by chronic
opioids is hypothesized to be due to downregulation of
BDNF signaling, because it was reversed by local
infusion of BDNF (Sklair-Tavron et al., 1996). Further
downstream neurotrophic signaling pathways mediating this effect include the downregulated Akt-mTORC2
signaling that increases the excitability of VTA DA
neurons and could thereby trigger a compensatory decrease in neuron soma size and DA output (MazeiRobison et al., 2011). The results are consistent with the
view that drug state-associated activity in the brain is
also based on rapid changes in the neuronal structures.
6. Conclusions. Opioids are effective in the acute
treatment of pain, but they are also abused due to their
addictive properties. After chronic exposure, opioids
produce a wide range of neuroadaptations that are
thought to underlie opioid tolerance, withdrawal, and
compulsive use. These adaptations can be identified at
different organizational levels, including those at the m
opioid receptors that undergo desensitization and internationalization, depending on the m opioid receptor
agonist. The alterations in intracellular signaling systems are complex, but opioid-induced upregulation of
the AC-cAMP-PKA cascade in specific synapses in

955

different brain areas is suggested to contribute to


neural and behavioral adaptations associated with
morphine withdrawal. Several neuropeptide systems
have been described to functionally oppose the m opioid
receptor signaling, thus inducing development of opioid
tolerance and dependence. In accordance with the idea
of aberrant forms of synaptic plasticity driving addictive
behavior, opioids influence plasticity both at glutamatergic and GABAergic synapses in many brain regions,
including the mesolimbic DA system. However, further
work is needed to understand how these adaptations
together produce the addicted state of compulsive opioid
use and relapse.
H. Cannabinoids: Multiple Mechanisms and
Possible Indications
Cannabis is one of the oldest and most used illegal
drugs in the world. There are also many potential
therapeutic indications for cannabinoid-based drugs
(Izzo et al., 2009; Pertwee, 2012), with the treatment
of spasticity and pain in multiple sclerosis and of
nausea, lack of appetite, and subsequent wasting in
AIDS having already been translated into clinical
medicine. The importance of normal endocannabinoid
function for brain health was dramatically revealed
during chronic treatment of extreme obesity with
rimonabant, the cannabinoid type 1 (CB1) receptor
inverse agonist. Although rimonabant (SR141716) was
as efficient as any other pharmacological treatment in
producing sustained reduction in weight and normalizing lipid and glucose metabolism in clinical trials
(Despres et al., 2005; Van Gaal et al., 2005), it was
withdrawn from the market rapidly after its launch
because of treatment-associated psychiatric adverse
effects, such as anxiety, depression, and suicidality in
a minority of the population studied (depressed mood
was an exclusion criterium for the clinical studies)
(Christensen et al., 2007; Van Gaal et al., 2008).
Cannabis intoxication is rarely considered lethal for
otherwise healthy subjects (Weissenborn and Nutt,
2012; Hall, 2015). Pediatric cases of cannabis-induced
coma and young- to middle-age adult cases of cannabisinduced cardiovascular fatalities and arrhythmias were
recently reported (Bachs and Morland, 2001; Tormey,
2012; Le Garrec et al., 2014). The low toxicity may be
related to low abundance of CB1 receptors in vital brain
stem nuclei regulating cardiovascular and respiratory
functions (Herkenham et al., 1990), although the receptors are present in the PAG and rostroventral medulla,
in which activation of CB1 receptors may produce
analgesia by disinhibition of descending pain-relieving
neurons (Vaughan et al., 1999, 2000). Another mechanism related to the inherent limits of acute cannabinoid
toxicity may be that treatment with 9-tetrahydrocannabinol (9-THC) activates the ERK1/2-MAPK pathway
via CB1 receptors producing an increase in the cytochrome P450 side-chain-cleavage enzyme (P450scc),

956

Korpi et al.

rapidly increasing the brain synthesis of pregnenolone


from cholesterol (Vallee et al., 2014). Among other
actions, pregnenolone is an allosteric inhibitor of CB1
receptors, thereby significantly reducing a wide range of
pharmacological effects of 9-THC, the main active
substance in cannabis. Moreover, 9-THC is only a partial agonist at cannabinoid receptors (Howlett et al.,
2002), which may limit its toxicity. It should be noted
that synthetic cannabinoids of various chemical structures and receptor binding profiles are much more
potent and efficient than 9-THC, and an alert on their
potential toxicity and lethality recently appeared
(Trecki et al., 2015).
1. Endocannabinoid System as an Endogenous Lipid
Messenger System. The cannabinoid system mediates
its main effects via the CB1 and CB2 receptors (Pertwee
and Ross, 2002; Pertwee et al., 2010). They are modulated by the endocannabinoids (eCB) N-arachidonoylethanolamine (anandamide) and 2-arachidonoylglycerol
(2-AG), plant cannabinoids (phytocannabinoids) such as
9-THC, as well as synthetic cannabinoid agonists {e.g.,
WIN 55,212-2 [R)-(+)-[2,3-dihydro-5-methyl-3-(4morpholinylmethyl)pyrrolo[1,2,3-de]-1,4-benzoxazin-6yl]-1-naphthalenylmethanone mesylate], CP55,940
[(2)-cis-3-[2-hydroxy-4-(1,1-dimethylheptyl)phenyl]trans-4-(3-hydroxypropyl)cyclohexanol], HU-210
[(6aR)-trans-3-(1,1-dimethylheptyl)-6a,7,10,10atetrahydro-1-hydroxy-6,6-dimethyl-6H-dibenzo[b,d]
pyran-9-methanol], and JWH-081 [(1-pentyl-1H-indol3-yl)-1-naphthalenylmethanone]} and antagonists (often
inverse agonists) such as the synthetic CB1 receptorselective rimonabant. The CB1 receptor is predominantly
expressed in the brain and plays a role in regulation of
mood and appetite, in processes of pain, in addiction, and
in synaptic plasticity in response to eCB-induced activation (Fig. 12) (for reviews, see Witkin et al., 2005; Katona
et al., 2006; Katona and Freund, 2008; Heifets and
Castillo, 2009; Hoffman and Lupica, 2013). CB2 receptors
are found predominantly in peripheral tissues, such as T
and B cells, keratinocytes, and peripheral nervous system,
and have so far remained a target for preclinical investigations related to immunomodulation (Dhopeshwarkar
and Mackie, 2014; Malfitano et al., 2014).
The cannabinoid system has turned out to be very
complex, because various endogenous and exogenous
ligands are not particularly specific for the cannabinoid
receptors (Pertwee, 2008). For example, the eCB anandamide also targets TRPV1 receptors, cannabidiol targets 5-HT1A receptors, whereas 9-THC, anandamide,
2-AG, and rimonabant all target 5-HT3, glycine, and
GABAA receptor channels. Despite this apparent nonselectivity, the G protein-coupled CB1 and CB2 receptors
mediate most of the cannabis effects, with most behavioral effects of 9-THC being absent in CB1-knockout mice
(Ledent et al., 1999; Zimmer et al., 1999). More recently,
however, CB2 receptors have also been found expressed in
various neuronal populations, as well as in microglia

(Van Sickle et al., 2005). Specifically, CB2 receptors are


present in the VTA DA neurons, and the selective agonists
JWH133 [(6aR,10aR)-3-(1,1-dimethylbutyl)-6a,7,10,10atetrahydro-6,6,9-trimethyl-6H-dibenzo[b,d]pyran]
and GW405833 [1-(2,3-dichlorobenzoyl)-5-methoxy-2methyl-3-[2-(4-morpholinyl)ethyl]-1H-indole] (antagonized
by AM630) were found to reduce DA neuron firing and
cocaine SA in the wild-type mice but not in CB2 receptor
partial knockout mice (Xi et al., 2011; Zhang et al., 2015a).
Most of the brain CB1 receptors are localized in axons
and presynaptic terminals of GABAergic and glutamatergic neurons (Howlett et al., 2002). This presynaptic
localization is illustrated by the mismatch of CB1
receptor mRNA (enriched in the soma) and the autoradiographic signal for CB1 receptor-specific ligand (e.g.,
[3H]CP55940) in the rodent and human brain. A clear
mismatch is also seen in glutamatergic cerebellar
granule neurons displaying CB1 mRNA in the cerebellar granule cell layer and CB1 ligand binding in the
molecular layer, where the axons impinge on dendritic
trees of Purkinje neurons. Another example is the
mismatch in the GABAergic neurons of the basal
ganglia, showing CB1 mRNA in the caudate-putamen
and presynaptic CB1 ligand binding in the external
segment of the globus pallidus and the substantia nigra
reticulata. Yet the definitive proof for presynaptic
localization of CB1 receptor comes from direct visualization using antibodies at the electron microscopy level
especially in hippocampal GABAergic CCK-expressing
basket cells (Hajos et al., 2000; Katona et al., 2000). In
conclusion, the CB1 receptors are predominantly presynaptic in both glutamate and GABA neurons, and
thus have the possibility to modulate synaptic efficacy
and neuronal circuit functions by controlling the release
of the main neurotransmitters (Freund et al., 2003). As
an example, activation of CB1 receptors on striatopallidal axon terminals leads to presynaptic inhibition of
GABAergic transmission to globus pallidus neurons
(Engler et al., 2006), whereas activation of CB1 receptors on subthalamic neuron axon terminals leads to
presynaptic inhibition of glutamatergic excitation of
globus pallidus neurons (Freiman and Szabo, 2005).
The eCB system is tightly coupled to regulation of
synaptic efficacy. The main eCBs, 2-AG and anandamide,
are formed from membrane phospholipid precursor molecules in the postsynaptic neuron in response to increased
intracellular Ca2+ after intense activation, after which
they are retrogradely released onto presynaptic CB1
receptors (Figs. 2, 6, and 12). The most abundant eCB
2-AG (Mechoulam et al., 1995) is formed from diacylglycerol by the action of specific postsynaptic membraneassociated diacylglycerol lipase (DGLa) (Bisogno et al.,
2003; Matyas et al., 2008), and degraded by presynaptic
monoacylglycerol lipase (MGL) that colocalizes with the
CB1 receptor (Dinh et al., 2002; Gulyas et al., 2004).
Although MGL is responsible for about 85% of 2-AG
hydrolysis, two other serine hydrolase family enzymes

Drug-Induced Neuroplasticity

957

Fig. 12. Endocannabinoids such as 2-arachidonoylglycerol (2-AG) and anandamide (not shown) are synthesized on demand and attenuate excessive
neuronal firing. When a postsynaptic neuron (shown here as a glutamatergic terminal) fires repeatedly, NMDA receptors and voltage-gated Ca2+
channels (VGCC) are activated, leading to influx of Ca2+ and activation of phospholipase Cb (PLCb) and production of diacylglycerol (DAG), which is
subsequently converted to 2-AG by diacylglycerol lipase (DGL). The newly synthesized 2-AG then diffuses to the presynaptically located cannabinoid
CB1 receptor, a Gi protein-coupled receptor that, when activated, deactivates presynaptic VGCCs, leading to an attenuation of neuronal firing and
glutamate release. Finally, the enzymes monoacylglycerol lipase and ABHD6 and 12 hydrolyze 2-AG to arachidonic acid (AA) and other metabolites.

have been discovered with potential roles in hydrolyzing


excess 2-AG, namely postsynaptic ABHD6 and microgliaenriched ABHD12 (Savinainen et al., 2012). Anandamide
(Devane et al., 1992) is synthesized by several enzymes
and degraded by the fatty acid amide hydrolase localized
mainly in the postsynaptic neuron (Gulyas et al., 2004).
Genetic inactivation of DGLa, but not that of DGLb, in
mice abolished retrograde eCB-mediated short-term
synaptic plasticity processes in the cerebellar, hippocampal, and striatal slices using various electrical and
pharmacological ways to induce eCB release (Gao et al.,
2010b; Tanimura et al., 2010).
2. Short-term Plasticity Involving Retrograde Endocannabinoid Signaling. In hippocampal slices, a train
of action potentials of CA1 neurons induces a suppression of GABAA receptor-mediated inhibition (Pitler and
Alger, 1994) with reduced frequency of spontaneous
IPSPs and reduced amplitudes of IPSCs but with
unchanged sensitivity to iontophoretically applied
GABA. This reproducible phenomenon was called DSI
(Fig. 2), and it was further shown that there a was a lag
of about 1 second between the train of action potentials
and maximal DSI and that it was prevented by pertussis toxin (Pitler and Alger, 1994). Therefore, presynaptic mechanisms were hypothesized, and a few years
later, it was found that this transient DSI in pyramidal
neurons was produced by the short-distance retrograde release of eCBs that inhibited GABA release
from interneurons via an action on presynaptic CB1
receptors (Ohno-Shosaku et al., 2001; Wilson and
Nicoll, 2001). Retrograde eCB signaling is very widespread and affects both inhibitory and excitatory synapses. DSE was also observed in the cerebellar cortex,

where a short depolarization of Purkinje neurons


induces the inhibition of glutamatergic inputs from
parallel and climbing fibers (Kreitzer and Regehr,
2001). Similar Purkinje neuron depolarization also
suppresses the GABAergic synapses from cerebellar
cortical interneurons (Tanimura et al., 2010), such as
the GABAergic basket cells (Szabo et al., 2004). In
addition to the HC and cerebellum, DSI and/or DSE is
also observed, e.g., in the VTA DA neurons (Melis et al.,
2004), NAc MSNs (Tanimura et al., 2010; Shonesy et al.,
2013), and in the BLA neurons (Zhu and Lovinger, 2005;
Patel et al., 2009; Yoshida et al., 2011; Shonesy et al.,
2014). Together these findings suggest that the retrograde eCB signaling is a very general local transient
process that tunes neurotransmission (reviewed in
Wilson and Nicoll, 2002; Kano et al., 2009).
Upon firing of a neuron containing the eCB synthesis
machinery, its Ca2+ level increases to micromolar levels
via VGCCs and NMDAR channels, maybe also from
release from intracellular Ca2+ stores (Kano et al.,
2009). Through as yet poorly known mechanisms,
2-AG synthesis is then promoted, part of that being
dependent on the phospholipase PLCb. For a single
activated neuron, diffusion of endocannabinoids
extends to a sphere of about 40 mm (Wilson and Nicoll,
2001), the effect thus remaining very local. Other
important activators of PLCb are all GPCRs that
signal through Gq/11 proteins (Kano et al., 2009), such
as group I mGlu receptors and M1/M3 muscarinic ACh
receptors (Katona and Freund, 2012). The eCB-CB1
receptor system is thus an important part of a versatile mechanism in presynaptic regulation of brain
circuits.

958

Korpi et al.

Signal transduction of CB1 (and CB2) receptors is


mediated via Gi/o proteins (reviewed in Heifets and
Castillo, 2009)). Thus, the activation of CB1 receptors
generally leads to inhibition of AC (leading to reduced
phosphorylation of PKA targets) and P/Q-type calcium
channels and stimulation of inwardly rectifying potassium channels. Many other signaling alternatives have
been proposed for these receptors, which may also be
present as homodimers or heterodimers.
3. Long-term Plasticity and CB1 Receptors. It is
thus clear that the activation of CB1 receptor by
endogenous or exogenous ligands has the potential to
influence synaptic plasticity. An early in vitro study
showed that a low concentration (10 pM) of 9-THC
prolonged LTP, whereas higher concentrations (100 and
1000 pM) reduced the duration of LTP in rat CA1hippocampal slice preparation (Nowicky et al., 1987).
Cannabinoid receptor activation in vitro impairs LTP
and LTD by reducing presynaptic glutamate release to
a level below that what is required to depolarize the
postsynaptic membrane to relieve the Mg2+-blockade of
NMDARs (Misner and Sullivan, 1999). Cannabinoid
receptor agonists (WIN 55,212-2 and CP55,940) and an
antagonist (rimonabant) favor LTD and LTP, respectively, in the glutamatergic synapses of rat PFC slice
preparation (Auclair et al., 2000). WIN 55,212-2 blocks
the induction of LTP and suppresses paired-pulse depression in rat CA1 slice preparation, an effect that is
blocked by rimonabant (Paton et al., 1998). Striatal LTD
at glutamate synapses is produced by postsynaptic release of eCBs, because it is not detected in CB1-KO mice,
it is blocked by rimonabant and activated by postsynaptic
anandamine (Gerdeman et al., 2002). Also, the synthetic
CB1 ligand, JWH-081, impairs LTP in mouse CA1 slice
preparation (Basavarajappa and Subbanna, 2014).
A single administration of 9-THC (3 mg/kg) to mice
blocked the eCB-dependent LTD in the NAc and HC for
at least 1 day and reduced the hippocampal DSI after 20
hours by almost 60% (Mato et al., 2004). A single dose of
9-THC ($5 mg/kg) impairs the LFS-induced LTD in
the HC (Chen et al., 2013b), which is absent in
cyclooxygenase-2 (COX-2) knockout mice. Seven-day
treatment of mice with 9-THC (10 mg/kg) impaired
fear memory (reduced footshock-induced freezing 24
hours after the last dose) and spatial learning and
memory in a water maze test (Chen et al., 2013b). These
CB1 receptor-mediated effects were dependent on increased COX-2 activity and production of PGE2. Furthermore, in the HC, this treatment downregulated the
expression, phosphorylation, and surface trafficking of
GluA1, GluN2A, and GluN2B subunits and consequently reduced LTP at CA3-to-CA1 synapses, which
was apparently correlated with reduced dendritic spine
density of HC-CA1 neurons. Interestingly, 9-THCinduced CB1 receptor activation upregulates COX-2
expression, whereas that by 2-AG suppresses the
COX-2 expression in cell culture models (Chen et al.,

2013b). This suggests that exogenous agonists and


eCBs may have different signaling routes, with
9-THC acting via Gbg subunits and 2-AG via the Gai
subunit. In cultured hippocampal neurons, synaptic
release sites are greatly reduced by incubation with
9-THC, which is mediated by decreased cAMP (Kim
and Thayer, 2001). It should be further noted that the
COX-2 enzyme is known to oxidize eCBs, anandamide
and 2-AG, and to produce derivatives with prostaglandins, the selective inhibition of which has been proposed
to be used in drug development to enhance eCB actions
(Hermanson et al., 2014).
In rats, acute treatment with 9-THC increased ERK,
pCREB, and c-Fos in the caudate-putamen and cerebellum, whereas chronic 9-THC exposure induced
increases in ERK, pCREB, and FosB in the PFC and
HC (Rubino et al., 2004). Blocking ERK activation in
Ras-GRF1-KO mice prevented tolerance to 9-THC in
sedation and analgesia (Rubino et al., 2004). In mice, it
has been reported that a very low dose (0.002 mg/kg) of
9-THC induced long-lasting (7 weeks) modifications of
ERK activity in the HC, FC, and cerebellum and
elevated pCREB in the HC and BDNF in the FC
(Fishbein et al., 2012). The effects of cannabinoids on
the ERK1/2 signaling in vivo were regulated by both D1
and NMDARs (Daigle et al., 2011). Chronic exposure of
rats to 9-THC (1.5 mg/kg for 7 days) upregulated
BDNF in the NAc, posterior VTA, hypothalamic paraventricular nucleus, and mPFC (Butovsky et al., 2005).
Chronic exposure to 9-THC prevents synaptic plasticity in the NAc and reduces sensitivity of GABAergic and
glutamatergic synapses to cannabinoids and opioids
(Hoffman et al., 2003), indicating a possible tolerance
and cross-tolerance development. It is noteworthy that
subchronic treatment of mice with 9-THC induced CB1
receptor desensitization in [35S]GTPgS autoradiography in several brain regions, but not in the striatal
regions, in which the treatment induced the greatest
expression of DFosB (Lazenka et al., 2014). This
suggests that gene expression regulated by DFosB
counteracted the CB1 receptor desensitization.
4. Cannabinoid Effects on Brain Development.
Recent rodent studies have indicated that cannabinoid
exposure to the developing brain during pregnancy or
early development may cause alterations in offspring
behavior and physiology via CB1 receptor-mediated
mechanisms (reviewed in Calvigioni et al., 2014). One
study revealed that maternal exposure to 9-THC affected the stability of cortical neurites and connections
and disrupted CB1 receptor-positive perisomatic baskets
in the cortex (Tortoriello et al., 2014). This was mediated
by a reduction in a microtubule-binding protein Superior
Cervical Ganglion 10 (stathmin-2) via phosphorylation by
CB1 receptor-mediated activation of JNK. Importantly,
reduced expression of Superior Cervical Ganglion
10 mRNA and protein was also found in human hippocampal tissue from cannabis-exposed fetuses (Tortoriello

Drug-Induced Neuroplasticity

et al., 2014). Behavioral effects are also evident, because


CB1 receptor blockade in the prenatal period decreases
immobility behavior, demonstrating that the cannabinoid
system plays a crucial role in ontogeny of psychomotor
behaviors (Moreno et al., 2005).
Cannabinoid receptors in the white matter tracts are
abundant in rats during early brain development but
dramatically reduce by P30 (Romero et al., 1997). In
prenatal brains, WIN55,212 activated [35S]GTPgS
binding in white matter areas, which could be antagonized by the CB1 receptor antagonist rimonabant.
Indeed, various components of the brain eCB system
are present at early development (Fernandez-Ruiz
et al., 2000), suggesting a possible involvement of
cannabinoid signaling in neurogenesis and gliogenesis
(Berghuis et al., 2005; Arevalo-Martin et al., 2007;
Garcia-Ovejero et al., 2013; Zhou et al., 2015), development and myelination of neuronal pathways, and thus
the macro- and microstructure of the brain during
adolescence and adulthood (Keimpema et al., 2011).
Adult neurogenesis in the HC is compromised in DGLaKO and DGLb-KO mice (Gao et al., 2010b), indicating
a role for eCBs. Importantly, CB1 receptors are present
in myelin-forming oligodendrocytes and CB2 receptors
in microglia, with both being increased in white matter
plaques of multiple sclerosis (Benito et al., 2007). As in
rodents, CB1 receptor binding is remarkable in the
white matter tracts of human fetal brains but hardly
detectable in postmortem brains from juvenile or adult
ages (Mato et al., 2003). In heavy cannabis users, axonal
connectivity, as assessed by diffusion-weighted magnetic resonance imaging, was impaired in comparison
with matched controls (Zalesky et al., 2012), with the
impairment correlating with age of starting regular
cannabis use. In another recent study, marijuana use
was associated with reduced gray matter volume of the
OFC and with increased functional and structural
connectivities in the forceps minor pathway (Filbey
et al., 2014). The connectivity changes in the latter
study were associated with early onset of use, being first
increased and then decreased during protracted use.
Thus, longitudinal studies are obviously needed, especially because the reversibility of the changes is not
known. Although the effects of prenatal 9-THC are not
as dramatic as those produced by alcohol abuse during
brain development in pregnancy, it is possible that early
cannabis use could affect brain development during
critical periods and, therefore, be a contributing factor
to later cognitive impairment, substance abuse/
addictions, and symptoms related to mental illnesses.
5. Cannabinoid-Induced Cognitive Impairment.
Cannabis acutely impairs various cognitive, perceptual,
and executive functions (see Ameri, 1999; Iversen,
2000). In one study, memory impairment after acute
cannabis in occasional cannabis users could be counteracted by the acetylcholinesterase inhibitor rivastigmine
(Theunissen et al., 2015), which might also activate the

959

same presynaptic G proteins as 9-THC via indirect


stimulation of mACh receptors. However, similarly to
prolonged use of benzodiazepines that might produce
persistent cognitive impairment, but which has been
hard to prove (c.f., section II.E), the cognitive deficiencies after chronic, heavy cannabis abuse have not been
proven to be persistent (Pope et al., 2002) nor purely
transient (Solowij et al., 2002), although at least one
report showed heavy users still performing worse than
light users after a 4-week abstinence (Bolla et al., 2002).
It should be noted that 9-THC levels remain at
a significant level for a few weeks after stopping the
use of cannabis (Bergamaschi et al., 2013), which could
explain the results of the latter report on cognitive
impairment during the first few abstinent weeks (Bolla
et al., 2002). Although adolescents might be more
sensitive to cognitive impairment by cannabis use, any
persistent impairment is still difficult to assess, and
there appears to be a need for larger longitudinal
studies (James et al., 2013).
Longitudinal studies would also be needed to establish possible gross abnormalities in human brain structural features produced during cannabis/marijuana
use. Some of the recent studies using MRI on subjects
with a history of cannabis use compared with nonusers,
have suggested statistically significant alterations in
gray matter density and/or volumes of brain regions
involved in reward and cognition, such as the HC, NAc,
and amygdala (Yucel et al., 2008; Demirakca et al.,
2011; Gilman et al., 2014; Lorenzetti et al., 2015). Most
recently, the report by Weiland et al. (2015) controlled
for the alcohol use between the daily marijuana user/
nonuser groups by matching the scores on the Alcohol
Use Disorders Identification Test (AUDIT). This study
found no significant effects of marijuana use on gray
matter density and volumes of the bilateral NAc and
amygdala, the HC, and the cerebellum. Although the
cannabis-use associated fine structural alterations cannot be excluded by this result, alcohol use may have
been a significant confounding factor in the earlier
studies in which the groups might have been more
heterogeneous in relation to alcohol use. Even in the
study of Weiland et al., alcohol drinking averaged five
drinks a day in all study groups.
In rats, cannabinoid treatments impair learning and
memory processes, but again the persistence of effects is
not known. Acute pharmacological blockade of MGL,
which increases 2-AG levels, also impairs learning and
memory in rats (Griebel et al., 2015). Chronic 15-day
treatment with a potent cannabinoid receptor agonist
HU-210 (100 mg/kg i.p.) caused impaired learning in
Morris water maze with longer intertrial delays and
reduced LTP in the CA1 region of HC tested 18 hours
after the last dose (Hill et al., 2004). Both 9-THC and
WIN 55,212-2 decrease PFC DA turnover for up to 2
weeks, suggesting that persistent cognitive disturbances after cannabinoid exposure are possible and perhaps

960

Korpi et al.

at least partly related to the reduction in the DA


transmission in the PFC (Verrico et al., 2003).
Repeated 9 -THC administration in adolescent
rhesus monkeys impaired age- and practice-related
improvements in accuracy on a spatial working memory
task and neither tolerance nor sensitization developed
to the acute effects after 6 months (Verrico et al., 2014).
This result is consistent with the acute impairment of
cognition by cannabis in adolescent humans.
6. Rewarding and Aversive Behaviors, Specific
Actions on VTA DA Neurons. Cannabis produces
euphoria and also aversive effects, such as anxiety, in
humans. 9-THC induced striatal DA release in
humans in some experiments (Bossong et al., 2009)
but not in others (Barkus et al., 2011). Similarly, in
rodent microdialysis or firing rate studies, 9-THC
induced DA release in some experiments (French,
1997; Tanda et al., 1997) but failed to do so in others
(Castaneda et al., 1991). Importantly, high-resolution in
vivo voltammetric analyses suggest that CB1 receptor
blockade by rimonabant can block DA transients in the
NAc induced by ethanol, nicotine, and cocaine in freely
moving rats (Cheer et al., 2007). Indeed, CB1 receptors
are present in the VTA GABA neurons and in afferent
GABAergic MSNs from the NAc (Matsuda et al., 1993;
Marsicano and Lutz, 1999), which is in keeping with
cannabinoid-induced decrease in GABA release in the
VTA (Szabo et al., 2002). There are no CB1 receptors in
the VTA DA neurons, but they contain the machinery
for eCB synthesis and thus for retrograde signaling.
These results indicate that the CB1 receptor activation
can lead to increased mesolimbic DA release, although
this effect seems to be in several cases more modest than
that induced by other drugs of abuse. Direct infusion of
9-THC to the posterior VTA and the shell of the NAc in
rats induces rewarding behavior, such as lever pressing
for the infusion and CPP (Zangen et al., 2006). Also,
9-THC increases GluA2-lacking AMPA receptors and
causes LTD in the pedunculopontine nucleus, a subcortical nucleus that influences reward and motivational functions (Good and Lupica, 2010). Early studies
on medial forebrain bundle self-stimulation at the
hypothalamus in rats indicated that, like other drugs
of abuse (Wise, 1996), 9-THC also facilitated the ICSS,
especially in Lewis rats (Lepore et al., 1996; Gardner
et al., 1988). However, like the effect on DA release, this
effect of 9-THC has been difficult to replicate in rats or
mice (Vlachou et al., 2007; Fokos and Panagis, 2010;
Wiebelhaus et al., 2015). A recent study with SpragueDawley rats suggested that a lower dose of 0.1 mg/kg of
9-THC reduced the threshold for self-stimulation,
whereas the often-used 1.0 mg/kg failed to do so, both
effects being antagonized by rimonabant (Kwilasz and
Negus, 2012).
There are two important behaviors that appear to be
under the control of eCB system: habit formation by
modulation of dorsolateral striatal neurotransmission

and extinction of aversive memories by modulation of


BLA neurotransmission. Recent development of behavioral assays for goal-directed versus habitual drugseeking behaviors have established specific roles for
the ventromedial striatum in goal-directed behavior
and for the dorsolateral striatum for habitual instrumental responding (Hilario and Costa, 2008; Everitt,
2014;) (Fig. 1). Experiments comparing responses to
immediate rewarding effects versus responding to
devalued rewards, e.g., after a preceding drug or food
reward, have been important to distinguish the goaldirected and habitual instrumental responding
(Balleine and Dickinson, 1998). Repeated responding
for drugs of abuse starts with activation of ventral pathways from the VTA to the NAc, but later overtraining
preferentially activates the dorsolateral nigrostriatal
pathways, previously suggested as a spiraling activation of midbrain-basal ganglia-cortical loops (Haber
et al., 2000; Haber and Knutson, 2010). There exists
a steeply increasing gradient of CB1 receptors from the
medial to lateral striatum (Herkenham et al., 1991),
which correlates with the presynaptic CB1 receptor
actions and retrograde actions of eCBs on striatal
excitatory transmission (Gerdeman and Lovinger,
2001; Gerdeman et al., 2002). Furthermore, pharmacologically isolated EPSCs and IPSCs were found to be
attenuated by neuronal firing-induced eCB release
(Adermark et al., 2009). Thus, both LTD and LTP
processes can be produced in the striatal MSNs by
activity-driven eCB retrograde signaling, which is more
prominent in the dorsolateral than dorsomedial striatum and can promote more habitual than goal-directed
responding. Habit formation is impaired under cannabinoid receptor blockade and in CB1-KO mice (Hilario
et al., 2007). Stimulant drugs initially induce goaldirected drug-seeking behavior, which after overtraining and prolonged drug exposure alters to more habitual
instrumental behavior (Everitt and Robbins, 2005).
This correlates with drug-induced changes in synaptic
functions (LTP, LTD) in the dorsal striatum (Gerdeman
et al., 2003) and possibly agrees with chronic stimulant
treatment-induced changes in the sensitization of behavioral stereotypy and striosomal early gene activation in the lateral striatum (Canales and Graybiel,
2000; Vanderschuren and Everitt, 2005). In THCdependent mice that show no difference in responding
to valued versus devalued rewards, bilateral infusion of
apamin (a SKCa channel blocker) into the dorsolateral
striatum rescued the eCB-dependent LTD (lost during
tolerance development) for up to 5 days and changed the
behavior from habitual responding to goal-directed
responding with a significantly greater responding for
valued than devalued rewards (Nazzaro et al., 2012).
D1R activation (D1R-like) has been linked to LTP and
D2R activation (D2R-like) to LTD in MSN glutamate
synapses (Calabresi et al., 2007). This is correlated with
DA inducing increased NMDAR currents in D1R-MSNs

Drug-Induced Neuroplasticity

and decreased currents in D2R-MSNs (Andre et al.,


2010), with both DA responses being modulated/partly
mediated the eCB system. The role of DA and DA
receptors in MSN plasticity has been restudied using
mouse models with selectively labeled MSNs of the
direct (D1R) or indirect (D2R) pathways (Surmeier et al.,
2009), revealing a complex interplay between timed
firing and dendritic activity and activation of a number
of GPCRs and intracellular signaling cascades (Shen
et al., 2008). In the striatum, CB1 receptors are present
in both direct (striatonigral) and indirect (striatopallidal) MSNs populations (Marsicano and Lutz, 1999;
Hohmann and Herkenham, 2000), presynaptically in
both GABAergic and glutamatergic axon terminals, but
not in TH-positive axon terminals (Fitzgerald et al.,
2012). The differential DA receptor distribution between the direct and indirect GABAergic striatal projection neurons (MSNs) is well known (Gerfen et al.,
1990). D2R activation stimulates anandamide production (Giuffrida et al., 1999), and this DA receptor
subtype has been shown to functionally interact with
CB1 (Kearn et al., 2005) and adenosine A2A receptors
(Ferre et al., 2010). Cocaine decreases striatal GABA
transmission, enhances anandamide synthesis, and
inhibits its breakdown via D2R-like receptor activation
(Centonze et al., 2002, 2004). However, the HFS-eCBdependent LTD at glutamate synapses of striatal MSNs
can be induced by opening of L-type VGCCs (Adermark
and Lovinger, 2007), suggesting that the action of D2R
activation on indirect pathway LTD is not obligatory.
Thus, the effect of D2R activation here is mediated
perhaps via suppression of adenosine A2A receptor
signaling that enhances retrograde eCB signaling (Fuxe
et al., 2007; Lerner et al., 2010). In the direct pathway,
the activation of D1R induces LTP (Calabresi et al.,
2007), after activation PKA-mediated phosphorylation
of DA- and cAMP-regulated phosphoprotein DARPP-32
and other plasticity-related signaling cascades
(Greengard, 2001; Nairn et al., 2004; Svenningsson
et al., 2004). Activation of this cascade has been linked
to the effects of many drugs of abuse (Nairn et al., 2004;
Svenningsson et al., 2004), and some behaviors induced
by 9-THC are altered in DARPP-32 knockout mice
(Lazenka et al., 2015). In line, subchronic 9-THC
selectively induces accumulation of DFosB in D1RMSNs (Lobo et al., 2013). This induction by 9-THC
can be blocked by D1R antagonist, but not by genetic
deletion of DARPP-32 (Lazenka et al., 2015). In conclusion, the present data indicate a central role for CB1
receptors and eCB signaling in the regulation of shortand long-term plasticities in the main striatal projection neuron populations. Further studies will be
needed to better understand the roles of these plasticity
processes in behaving animals.
Cocaine SA in rhesus monkeys was initially associated with reduced [14C]deoxyglucose uptake in the
ventral striatum and PFC, which, after prolonged SA,

961

shifted to the dorsolateral striatum (Porrino et al.,


2004). It is not known whether the cocaine-induced
reduction in striatal activation involves the CB1/2 receptor mechanisms or some other signaling pathways
that are affected by chronic cocaine. In rodents at least,
antagonism of CB1 receptors can abolish the rewarding
effects of cocaine (Xi et al., 2008), whereas the effects of
CB2 receptor blockade may show species differences
between mice and rats (Zhang et al., 2014). The cocaine
SA experiments in rhesus monkeys stress the importance of serial engagement of medial versus lateral
midbrain-striatal-cortical loops. However, these changes
in neuronal activity may be also explained by prolonged
changes in DA release, receptors, and transporters
(Beveridge et al., 2009), not necessarily by eCB-induced
LTD.
Aversive memories induced by drugs might be an
important deterrent of drug abuse. Marsicano et al.
(2002) found that eCB activation promotes extinction of
aversive fear memories by inhibiting GABAergic circuits in the amygdala. By using brain slices, HFS of
lateral amygdala induces LTD of GABAergic transmission in the BLA, which then increased amplitudes of
EPSCs in the neurons of the CeA, the final output nuclei
from the amygdala (Azad et al., 2004). This activation
might then induce a novel memory trace for extinction
of fear conditioning. Notably, CB1-KO animals failed to
show extinction of freezing response to an auditory tone
previously coupled to footshocks. Extinction of appetitively motivated responding was not affected in CB1-KO
animals (Holter et al., 2005). More recently, infusion of
the CB1 receptor antagonist AM251 [N-(piperidin-1-yl)5-(4-iodophenyl)-1-(2,4-dichlorophenyl)-4-methyl-1Hpyrazole-3-carboxamide] into the BLA was shown to
abolish the freezing response if infused after the
reactivation of fear memory (Ratano et al., 2014),
suggesting that reconsolidation of fear memory is
dependent on CB1 receptor activation. This effect of
AM251 could in turn be reversed by infusion of
bicuculline, a GABAA receptor antagonist, indicating
that CB1 receptor blockade enhanced GABA transmission. Altogether these results suggest that cannabinoid mechanisms affect the amygdaloid circuitry
and produce long-term alterations in responding to
aversively conditioned stimuli.
7. Cannabinoids during Adolescence: Increased Risk
for Schizophrenia?. Older literature on cannabis
abundantly recognizes that heavy use is associated
with sensory modality alterations and hallucinations
(Iversen, 2000). However, an association between cannabis use and the development of schizophrenia has
been more difficult to demonstrate. A study following
Swedish conscripts from adolescence found a relationship between cannabis use and subsequent development of schizophrenia, with an odds ratio (OR) of about
6 when cannabis had been used chronically more than
50 times (Zammit et al., 2002). This hints toward severe

962

Korpi et al.

persistent alterations in brain functions due to cannabis


use, even if the risk appears to decline with time in the
original study cohort (Manrique-Garcia et al., 2012),
and it has been calculated that thousands of users
would need to be intervened to prevent one case of
schizophrenia (Hickman et al., 2009). Meta-analyses
have also found increased risk for different mental
illnesses (OR about 1.4) for those who have ever used
cannabis, especially for those with frequent use (OR
about 2) (Moore et al., 2007). Furthermore, a recent
report showed no increase in cases of first-episode
psychosis with a history of milder hash-resin use but
an increased number of cases with a history of using
stronger skunk-like cannabis (Di Forti et al., 2015).
Cannabinoids may provoke greater pharmacological
effects on cognition and other behavioral domains in
schizophrenic patients (DSouza et al., 2005). Indeed,
a study of young students suggested that individual
variability in propensity to experience psychotomimetic
effects during the first use of cannabis may depend on
polymorphisms of the CB1 gene (CNR1) (and not on
COMT, AKT1, or BDNF polymorphisms) (Krebs et al.,
2014). Although it is possible that cannabis abuse may
accelerate the start of psychosis and schizophrenia in
susceptible individuals, a causal role of cannabis use in
the development of schizophrenia is not established.
Several studies have assessed the effects of cannabinoid exposure on adolescent rodents (reviewed in
Renard et al., 2014). Adolescent female rats treated
(from P35 to P45) with 9-THC develop depressive-like
behavior in adulthood (Rubino et al., 2008, 2009a),
which could be reversed by chronic URB597 (fatty acid
amide hydrolase inhibitor) treatment during adulthood,
indicating a role for eCB-CB1 receptor signaling
(Realini et al., 2011). In rats treated with 9-THC in
adolescence, impaired performance in the radial arm
maze was observed, and this was associated with
decreases in synaptic markers and synaptic efficiency
in the HC and PFC (Rubino et al., 2009a,b). Adolescent
exposure to the CB1 receptor agonist CP 55,940 impaired short-term memory (30- and 120-minute delay)
in a similar fashion in Wistar and Lister Hooded rats.
However, in an object location task, the impairment in
Wistar rats was observed after 30-minute delay but only
after a 120-minute delay in Lister Hooded rats (Renard
et al., 2013). Although the deficits induced by the potent
synthetic cannabinoid agonist HU 210 in learning and
motor activity were reversed in 3 days of abstinence, it
produced more persistent disruption of acquisition and
vocalization that was still evident after a 7-day abstinence (Ferrari et al., 1999). On the contrary, one report
shows that chronic treatment of periadolescent rats
with the cannabinoid agonist CP 55,940 had only
marginal effects on cognition and anxiety during
adulthood with the exception of an elevated hippocampal PSA-NCAM expression (Higuera-Matas et al.,
2009). Taken together, cannabinoid exposure during

adolescence may cause long-lasting effects on behavior,


especially in female rats. Importantly, the experiments
by Schneider and Koch (2003) indicate that the time of
cannabinoid treatment is crucial. Chronic WIN 55,2122 treatment during puberty (from P40 to P65), but not
during prepubertal or adult age, induced prolonged
effects on sensorimotor gating (loss of prepulse inhibition of acoustic startle, which could be normalized by
acute haloperidol 0.1 mg/kg), object recognition memory, and a progressive ratio task in adulthood (P75
P120). Thus, as in humans, the significance of the
timing of cannabinoid exposure for persistent effects is
important (Schneider, 2008), which may be related to
the increased activity of the brain cannabinoid system
during adolescence in humans and experimental animals (Mato et al., 2003; Rodriguez de Fonseca et al.,
1993).
8. Conclusions. Cannabis is a relatively safe drug,
with low potential for lethality in cannibis intoxications
of young people. But because the endocannabinoid
system is wired into mechanisms of brain development
and synaptic plasticity, they have the potential for
producing long-term changes in the nervous system
structure and function. Particularly the heavy cannabis
exposure during adolescence may induce early psychosis and other changes in cognitive and emotional
behaviors. Furthermore, more longitudinal research
is needed for several aspects of cannabinoid-induced
long-term alterations in brain structure and function,
especially with an advent of more potent synthetic
cannabinoids and their entering the illegal markets.
I. Hallucinogens
The study of the persistent effects of 5-HT, lysergic
acid diethylamide (LSD), and other serotonergic hallucinogens on perception, cognition, and mood started more
than four decades ago but has long been hampered by
legislative hurdles (Geyer and Vollenweider, 2008; Nutt
et al., 2013). Therefore, it is only recently that proper
guidelines for psychedelic drug research have been
established (Johnson et al., 2008) and studies have begun
to produce some understanding of the neurochemical
events directly responsible for producing hallucinogeninduced neuroplasticity (Catlow et al., 2013; Lepack
et al., 2015), although many questions still remain.
In this section, the basic pharmacology of hallucinogens is discussed, followed by a review of their acute
effects on neurochemistry and behavior. Finally, the
neuroplasticity of hallucinogens will be discussed in
terms of their long-term effects on sensory processing,
mood, and anxiety.
1. Effects of Serotonergic Hallucinogens Are Mediated
by the 5-HT2A Receptor. The typical serotonergic hallucinogens can be divided into several chemical classes
including tryptamines, phenethylamines, ergolines, and
amphetamines. They all display high affinity for the
5-HT2 receptors but also show varying degrees of affinity

Drug-Induced Neuroplasticity

for a range of other 5-HT receptors including the 5HT1


5HT4, 5HT5, 5HT6, and 5HT7 receptors (Nichols, 2004).
The ergoline hallucinogen LSD, for instance, has the
highest affinity to, and (partial) agonistic effects on,
5-HT2A, 5-HT2C, 5-HT1A, and 5-HT6 receptors in the low
nanomolar range, but also shows intrinsic activity at the
DA D2 and a-adrenergic receptors (Nichols, 2004; Passie
et al., 2008). The tryptamine hallucinogen, psilocybin
(4-phosphoryloxy-N,N-dimethyltryptamine), is synthesized de novo in several species of hallucinogenic mushrooms and, together with its active metabolite psilocin,
it also displays a low nanomolar affinity for 5-HT2A
receptors, and a somewhat lower affinity for 5-HT1A and
5-HT2B receptors (McKenna et al., 1990). With a Ki of
0.7 nM, the 5-HT2A receptor is also the primary target of
the synthetic amphetamine hallucinogen DOI, although
it also targets 5HT2B and 5HT2C receptors with slightly
lower affinities (Canal and Morgan, 2012). Also other
hallucinogens, such as N,N-dimethyltryptamine and
the phenethylamine hallucinogen mescaline have significant affinity mainly to the Gq/11-coupled 5-HT2A/C
receptors (Aghajanian and Marek, 1999; Ray, 2010).
Moreover, efavirenz [(4S)-6-chloro-4-(2-cyclopropylethynyl)4-(trifluoromethyl)-2,4-dihydro-1H,3,1-benoxazin-2-one], a
nonnucleoside reverse-transcriptase inhibitor used for
treatment of HIV-1 infection, shows a risk for neuropsychiatric adverse effects such as night terrors and
hallucinations and is a potent 5-HT2A agonist (Gatch
et al., 2013). MDMA, however, a stimulant with secondary hallucinogenic properties, produces its hallucinogenic effects mainly indirectly, by increasing synaptic
5-HT levels through SERT-mediated 5-HT release
(Liechti et al., 2000).
Although hallucinogens have a number of 5-HT
receptor targets, the hallucinogenic effects in fact
appear to be mediated primarily by one receptor subtype (Fig. 13). The hallucinogen-induced head shaking
response can be restored in transgenic mice lacking the
5-HT2A receptor by selectively restoring this receptor in
cortical pyramidal neurons (Gonzalez-Maeso et al.,
2007), although the discriminative stimulus properties
of LSD and subjective effects of hallucinogens can be
fully blocked by 5-HT2A antagonists in both experimental animals and humans (Vollenweider et al., 1998;
Gresch et al., 2007). Furthermore, the 5-HT2A/C antagonist ketanserin blocks the effect of psilocybin on visual
EEG and event-related potentials and on prepulse
inhibition of the acoustic startle in healthy volunteers
(Quednow et al., 2012; Kometer et al., 2013). This
indicates that the hallucinogenic effects of serotonergic
hallucinogens are primarily mediated by the 5-HT2A
receptor. This receptor subtype is enriched in the deeper
cortical layers (Andrade, 2011).
Aside from serotonergic hallucinogens, other drugs
that produce hallucinogenic effects include substances
such as ketamine, nitrous oxide, and phencyclidine,
which belong to a subtype of hallucinogens often

963

referred to as dissociatives, acting primarily as NMDAR


antagonists. In recent years, a number of new analogs,
such as methoxetamine and 3-MeO-PCP have become
available on the recreational drug market. They have
been shown to have a similar mechanism of action
(Morris and Wallach, 2014; Roth et al., 2013). Additionally, drugs such as dextromethorphan and salvinorin A
that act primarily as m-, d- or k-opioid receptor agonists
or as monoamine reuptake inhibitors, are also included
in the class of dissociative hallucinogens (Codd et al.,
1995; Burns and Boyer, 2013; Morris and Wallach,
2014). So, although the 5-HT2A receptor is the most
important for mediating the hallucinogenic effects of
serotonergic hallucinogens, many other receptors mediate the hallucinogenic effects of other classes of
psychedelics. As the receptor targets of serotonergic
hallucinogens have been clearly identified, it raises the
question of what downstream affects result after modulation of receptor activity by these substances.
2. Hallucinogen ActionA Perturbation of Sensory
Gating or Desynchronization of Cortical Rhythms?.
The involvement of 5-HT2A receptors in the apical
dendrites of cortical pyramidal neurons in producing
hallucinogenic effects (Gonzalez-Maeso et al., 2007)
suggests involvement of cortical circuit mechanisms.
Evidence of thalamocortical glutamate pathway involvement comes from a study demonstrating hallucinogeninduced changes in glutamate release (Aghajanian and
Marek, 1999). These results were supported by human
imaging studies performed in subjects under acute
exposure to hallucinogens (psilocybin, mescaline, and
ketamine) that indicate activation of the frontal, cingulate, and insular cortices, with reduced activation in
the striatum, thalamus, and parietal cortex (Geyer and
Vollenweider, 2008). Interestingly, a more recent experiment using intravenous psilocybin (2 mg) indicated
reduced blood flow and blood oxygen level signals in
the anterior and posterior cingulate cortex and thalamus
of psilocybin-experienced healthy volunteers reporting
strong subjective acute effects of the drug (CarhartHarris et al., 2012). This was replicated by a subsequent
magnetoencephalography study showing large decreases
in spontaneous oscillary power after psilocybin, especially in cortical association areas and default-mode
network (Muthukumaraswamy et al., 2013). These
changes were of neuronal origin, consistent with excitation of 5-HT2A receptors on deep layer 5 pyramidal
neurons, which results in activation of inhibitory mechanisms in more superficial layers (Andrade, 2011;
Bastos et al., 2012).
The sensory gating theory of hallucinogenic action
states that hallucinations occur because of drug-induced
perturbations of the limbic cortico-striato-thalamiccortical feedback loop and subsequent reduction of thalamic sensory gating, leading to sensory overload that
manifests as hallucinations (Vollenweider, 2001). The
action of other drugs capable of inducing hallucinations,

964

Korpi et al.

such as NMDAR antagonists and even amphetamine,


may be explained using the same model, suggesting that
alterations of the cortico-striato-thalamic-cortical loops
may be a final common pathway for all hallucinogenic
drugs. Moreover, the drug-induced hallucinatory state
also shows strong phenomenological similarity to acute
psychosis in patients suffering from schizophrenia, and
both states are reversed by antipsychotic drugs, which
has led to use of the hallucinogen-induced state as
a pharmacological model of psychosis (Vollenweider et al.,
1997a,b). More recently, however, the cortical desyncronization theory of hallucinogen action (see the previous
paragraph) has replaced the sensory gating theory,
because the drug effects can be explained by direct actions
on cortical 5-HT2A receptors. Moreover, pharmacological
models of psychosis can be induced by many drugs
activating or inhibiting the main cortical neurotransmitter systems (reviewed in Steeds et al., 2015).

Fig. 13. The hallucinogenic effects of serotonergic hallucinogens such as


mescaline (MESC), psilocybin (PSILO), LSD, and DOI are primarily
mediated by the 5-HT2A receptor. However, most hallucinogens also show
affinity for a variety of other 5-HT receptors, including both postsynaptic
receptors as well as presynaptic inhibitory autoreceptors. Drugs such as
MDMA and related substances, such as methylone, are not primarily
hallucinogens but do produce some hallucinogen-like effects in users.
These effects are not primarily due to direct interaction with the 5-HT2A
receptor but are instead mediated predominantly due to the fact that
MDMA enhances synaptic serotonin levels through competitive reuptake
inhibition and enhancement of 5-HT release. The downstream effects of
5-HT receptor stimulation include activation of phospholipase C (PLC),
adenylate cyclase (AC), and subsequent modulations of postsynaptic
diacylglycerol (DAG), inositol trisphosphate (IP3), and cyclic adenosine
monophosphate (cAMP) levels.

3. Behavioral Effects and Addiction Potential.


Serotonergic hallucinogens, and also some other hallucinogens such as salvinorin A, do not have any clear
rewarding potential in animal models (Passie et al., 2002;
Nichols, 2004; Cunningham et al., 2011). Similarly, although these substances are used frequently by humans,
addiction appears to be very uncommon, and indeed they
have been used to treat alcohol addiction (Gable, 1993;
Studerus et al., 2011; van Amsterdam et al., 2011; Krebs
and Johansen, 2013). This is consistent with the rather
low-affinity, mixed and delayed efficacy of LSD on various
types of DA receptors (Passie et al., 2008); with a very low
affinity of psilocybin/psilocin to DA receptors (Passie et al.,
2002) and with salvinorin A having selective high affinity
to and agonistic efficacy on k-opioid receptors (Roth et al.,
2002; Chavkin et al., 2004). Although the k-opioid receptor
agonists may also influence DAergic mechanisms, they are
generally aversive (decreasing DA release) and in fact
produce conditioned place aversion (Gable, 1993; Zhang
et al., 2005; Cunningham et al., 2011). Interestingly, dosedependent rewarding/aversive effects in place conditioning
with salvinorin A via k-opioid and CB1 receptors have been
reported in zebrafish (Braida et al., 2007).
On the other hand, a 3-month LSD treatment of rats
(0.16 mg/kg i.p. every 2nd day) induced hyperlocomotion
for at least 3 months after discontinuation of the
treatment (Marona-Lewicka et al., 2011), which suggests an action on DA receptors. Indeed, using a drug
discrimination paradigm, Marona-Lewicka and coworkers showed that LSD produces initial generalization
to a 5-HT2A agonist (and suppression of locomotion),
followed at a later phase of cue responding, and
generalization to a D2 agonist (and locomotor activation) (Marona-Lewicka et al., 2005; Marona-Lewicka
and Nichols, 2007). The prolonged LSD treatment
shows development of persistent psychosis-like behavioral and neurochemical features in rats (MaronaLewicka et al., 2011), with at least the hyperlocomotion
being antagonized by antipsychotic drugs blocking D2
and/or 5-HT2A receptors.
The picture is somewhat different with regard to
NMDAR antagonism-based psychedelics, because there
is some evidence that dissociative psychedelics such as
ketamine and PCP are self-administered, at least under
some circumstances (Broadbear et al., 2004; Carroll
et al., 2005; De Luca and Badiani, 2011). These effects
are likely related to the ability of these substances to
increase accumbal DA release and turnover (Carboni
et al., 1989; Irifune et al., 1991) (see section II.F).
4. Long-term Residual Effects of Serotonergic Hallucinogens: the Role of Neuroplasticity. Hallucinogens produce clear neurochemical and acute behavioral and
subjective effects. However, they also produce more
long-term, persistent effects on behavior and sensory
processing, as well as on mood and attitude.
a. Sensory processing. Hallucinogen-induced longterm alterations of sensory processing have been

Drug-Induced Neuroplasticity

reported, initially as LSD-like recurrences or "flashbacks" in the 1950s. Later, diagnostic criteria were
proposed for flashbacks in the DSM-III-R that were
further refined and termed hallucinogen persisting
perception disorder (HPPD) in the DSM-IV (American
Psychiatric Association, 1994). Although this disorder
has a very low prevalence (Halpern and Pope, 2003),
HPPD is characterized by perceptual symptoms, most
often visual in nature, that manifest themselves in
patients previously intoxicated with hallucinogens,
even long after the cessation of drug intake (and beyond
the accepted time of the drugs acute pharmacological
actions), associated with significant distress, and without being attributed to any other medical condition or
psychiatric disorder (Horowitz, 1969; Abraham et al.,
1996; Halpern and Pope, 2003). Several discrete case
studies report flashback phenomenon days to months
after the use of hallucinogens such as LSD (Duncan,
1974) and psilocybin (Espiard et al., 2005) alone or in
combination with alcohol or other drugs of abuse, but
there are hardly any controlled studies on the subject. It
is noteworthy that pharmacological treatment, e.g.,
with antipsychotics, benzodiazepines, and clonidine,
has proved useful in controlling the symptoms in most
of the few cases published (Halpern and Pope, 2003).
Alterations in visual processing may be better described without reference to HPPD, because this diagnosis requires the presence of significant distress
associated with the symptoms. Persistent effects such
as perception disorders have also been observed in
healthy individuals after experimentally administered
doses of psilocybin (Studerus et al., 2011), and these
could all be managed by interpersonal discussion and
support. Furthermore, drug-free visual experiences
were observed in 62% of the respondents in a webbased study that analyzed a population of 2455
participants for abnormal experiences associated with
a history of hallucinogen use, especially LSD (Baggott
et al., 2011). The persistent perceptual alterations in
subjects with past hallucinogen use is clear evidence of
neuroplasticity. Even in the absence of recent LSD use,
patients with HPPD display an augmented neural
coherence in the occipital regions (isolated from anterior
regions) that is proposed to underlie their illusions and
hallucinations (Abraham and Duffy, 2001). However,
little is known about the exact mechanisms responsible
for these changes or about the factors involved in
producing susceptibility to long-lasting sensory alterations or HPPD.
b. Mood and anxiety. It has been known for a long
time that serotonergic hallucinogens can produce longlasting effects on mood, anxiety, and other aspects of
neuropsychiatric functioning, thereby providing indirect evidence of neuroplasticity. For instance, LSD was
studied early on as a drug or an adjunct to other
treatment modalities for neuropsychiatric illnesses,
but research collapsed in the late 1960s after LSD and

965

other psychedelics were made illegal. Subsequently,


a long period largely devoid of any psychedelic research
followed, something that has been attributed largely to
the fact that their illegality made it excessively difficult
for researchers to work with these substances (Nutt
et al., 2013). However, after a long hiatus, there now
appears to be a renewed interest in the potential clinical
applications of psychedelic drugs, such as psilocybin
and LSD and also other psychedelic drugs including
ketamine and MDMA, for treating disorders ranging
from alcohol and nicotine addiction to depression,
anxiety disorders, and cluster headaches (Sewell
et al., 2006; Chabrol, 2013; Grob et al., 2011; Krebs
and Johansen, 2012; Mithoefer et al., 2013; Gasser
et al., 2014; Moaddel et al., 2015). Evidence of longterm effects on behavior comes also from a set of studies
where psilocybin was administered to healthy subjects.
They reported that doses of psilocybin that resulted in
strong visual and perceptual effects, including those
described as mystical-type experiences, subsequently
also produced persistent positive changes in mood,
attitude, and behavior that were sustained for more
than 1 year after the drug was administered (Griffiths
et al., 2006, 2011; MacLean et al., 2011). This suggests
that serotonergic hallucinogens may possess anxiolytic,
antiaddictive, and antidepressive properties, but the
question remains what mechanisms mediate the neuroplasticity required to produce these long-lasting effects.
One recent study attempted to shed light on the
relationship between the therapeutic effects of serotonergic hallucinogens and their effect of neuroplasticity by measuring the effects of varying doses of
psilocybin on extinction of fear conditioning and neurogenesis in mice (Catlow et al., 2013). They demonstrated that psilocybin significantly accelerated the
extinction of fear conditioning, and this was associated
with a trend toward increased neurogenesis in the
dentate gyrus. Although further work is needed, one
could speculate that the effects of psilocybin on neurogenesis are due to 5HT2A receptor-mediated upregulation of BDNF (Vaidya et al., 1997), suggesting a mechanism
strongly resembling that of SSRIs (Martinowich and Lu,
2008; Catlow et al., 2013; Quesseveur et al., 2013) and
possibly the antidepressive mechanism of ketamine
(Lepack et al., 2015).
Considering the vital role of the 5HT2A receptor, it is
important to note that although the 5-HT2A receptor is
the primary target for hallucinogens, there are other
5-HT2A agonists, such as lisuride, that do not produce
any hallucinogenic effects. The exact reason for this
difference is not known, but may be related to differing
recruitment of intracellular second messengers; although both LSD and lisuride activate PLC, only LSD
also recruits Gi/o proteins and Src (Gonzalez-Maeso
et al., 2007). Interestingly, the specific 5HT2A receptor
activation caused by hallucinogens seems to be of key
importance in producing neuroplasticity. Hallucinogens

966

Korpi et al.

such as LSD induce brain-region dependent mRNA


expression of various immediate early genes, such as
c-Fos, arc, and homer1a splice isoform ania3 (Nichols
and Sanders-Bush, 2002), that influence regulation of
gene expression and synaptic responses, whereas DOI
has been shown to produce changes in dendritic spine
morphology on cortical neurons (Jones et al., 2009).
Chronic LSD administration to rats causes a persistent
increase in synthesis and turnover of 5-HT as well as
downregulation of cortical 5HT2A receptors (Diaz and
Huttunen, 1971; Lee and Geyer, 1980; Gresch et al.,
2005), and a 2-week LSD treatment followed by 14 days
of abstinence still showed an augmented 5-HT level in
the midbrain and cerebral cortex (Peters and Tang,
1977). Continuous delivery, but not daily injections, of
LSD in rats caused long-lasting increases (measured
after 30 days) in [3H]LSD binding in the FC, amygdala,
septum, and NAc that coincided with behavioral alterations (King and Ellison, 1989). LSD (1 mg/kg) causes
a sustained increase in Nor1 gene expression in the
PFC, which could play a role in synaptic dysregulation
(Nichols et al., 2003). Similar effects of LSD and DOI on
transcriptional responses in mouse somatosensory cortex
were dependent on 5-HT2A receptors, because they were
absent in 5-HT2A-KO mice (Gonzalez-Maeso et al., 2003).
Importantly, lisuride, a 5-HT2A agonist not producing the
hallucinogenic activity-correlating head-twitching behavior in rodents, failed to produce similar transcriptomics
changes as LSD and DOI. These results are consistent
with 5-HT2A receptor antagonists abolishing the behavioral effects of hallucinogenic drugs (Fiorella et al., 1995;
Vollenweider, 1998; Winter et al., 1999).
5. Scopolamine: Another Hallucinogen Revisited for
Depression. Scopolamine (hyoscine or burundanga) is
a nonselective muscarinic antagonist well known for its
use in preclinical animal models to induce cognitive
impairment and for its use to treat, e.g., motion
sickness. Scopolamine disrupts working memory and
other parameters in rodents, canines, and primates
(Rupniak et al., 1989; Araujo et al., 2004; Winsauer
et al., 2004; Klinkenberg and Blokland, 2010). Some
case reports have revealed hallucinogenic effects of
scopolamine, largely from patients of different age
groups in whom hallucinations unexpectedly followed
peripheral administrations (transdermal patch or
eye drops) of scopolamine (Warburton et al., 1985;
Sennhauser and Schwarz, 1986). A sparse amount of
literature predicts that scopolamine could be addictive
and it is likely to cause CNS effects upon withdrawal
(Luetje and Wooten, 1996; Patel and Ezzo, 2009), but its
recreational or predatory use has not been convincingly
identified (King et al., 2014). On the other hand,
scopolamine might have a beneficial effect in opioid
and cocaine addiction (Gambelunghe et al., 2014; Liu
et al., 2013). With the recent interest in exploration for
quick-onset antidepressants, scopolamine, like ketamine, was reported to be a prospective candidate and

interestingly the clinical improvement lasted for more


than 2 weeks (Furey and Drevets, 2006). These reports
show that although the cognitive impairments are due
to mAChR antagonism, the persistent effects of scopolamine that might be mediated via NMDAR gene
expression and mTOR signaling-mediated synaptic
plasticity (Drevets et al., 2013); these ideas need to be
investigated in the future.
6. Conclusions. The hallucinogenic effects of serotonergic hallucinogens are mediated primarily by the
5HT2A receptor, activation of which produces hallucinations by a range of functional changes thought to
result in deficient thalamic sensory gating and/or
cortical desyncronization. Furthermore, 5HT2A receptor
activation by hallucinogens also mediates many
changes in gene expression, receptor regulation, and
synaptic morphology likely to be associated with the
long-term, persistent effects of hallucinogens on perception, mood, and anxiety. However, the molecular
mechanisms linking the neuropharmacological actions,
human brain functional changes, and behavioral outcomes need to be further studied.
III. General Discussion
We have reviewed neuroplasticity induced by cocaine,
amphetamine, and related psychostimulants, nicotine,
ethanol, benzodiazepines and other GABA ligands,
NMDAR antagonists, opioids, cannabinoids, and hallucinogens. Brain regions implicated across the various
classes of drugs of abuse were diverse but again certain
commonalities emerged. As might be expected, the VTA
was recurrently implicated, but other brain regions that
featured repeatedly across classes of drug of abuse were
the striatum, amygdala, BNST, Hb, mPFC, and HC.
Although there are many diverse mechanisms of neuroplasticity specific to the pharmacology of each of these
classes of drug of abuse, many common themes
surfaced.
Among the drugs of abuse that might be characterized as having stimulant and/or disinhibitory effects
cocaine, amphetamine and related psychostimulants,
nicotine and ethanol, and benzodiazepinesa recurrent
theme was the role of neuroplasticity of glutamate
synapses in the VTA DA neurons. However, the specific
mechanisms of early induction of the persistent changes
in the AMPAR/NMDAR ratios varied between the
different drugs. Additional investigations are required
to further elucidate whether different drugs induce
glutamate neuroplasticity in different populations of
VTA DA neurons, perhaps one of them linking to
reinforcing mechanisms and another one to aversive
mechanisms.
Another commonality was the recurrent role of
GABAergic mechanisms, by disinhibition of the DA
neurons, in the VTA in the neuroplasticity induced by
cocaine, nicotine, ethanol, benzodiazepines and other

Drug-Induced Neuroplasticity

GABA ligands, opioids, and cannabinoids. The specific


mechanisms of induction of changes in GABAergic
function in the VTA varied between the various classes
of drugs, and further studies are required to investigate
whether there would be possibilities in controlling
GABAergic function at specific neuronal populations
in the midbrain to modulate addictive behaviors across
these different classes of drugs.
An overarching theme that has repeatedly emerged is
concern over the possibility that neuroplasticity induced by one drug of abuse may increase the risk of
dependence on other drugs of abuse. In particular, there
is concern that relatively easily available drugs of abuse
such as nicotine and alcohol may trigger mechanism of
neuroplasticity that predispose to increased risk of
abuse of other drugs. However, further studies are
required to determine whether the underlying mechanisms of neuroplasticity induced by nicotine and alcohol
predispose risk to addiction to all other drugs of abuse
equally or preferentially to specific classes of drug of
abuse. Moreover, the question of the reversibility of the
neuroplastic changes induced has seldom been exhaustively addressed. For many of the drugs of abuse the
question of whether there is long-lasting and irreversible alteration of function has not been fully resolved.
Further studies are required to determine how long
these changes continue to increase the risk of drug
abuse.
Another overarching theme that has repeatedly
emerged across our discussion of the various classes of
drugs of abuse is the particular vulnerability of the
adolescent brain to drug-induced neuroplasticity. Evidence is accumulating for concern over significantly
stronger and more insidious induction of plasticity on
adolescent drug exposure compared with adult drug
exposure. Adolescent exposure to drugs of abuse is
a major concern because the relatively easy availability
of alcohol, nicotine, and cannabis in many societies
could lead to risk of exposure. Even limited exposure
may be sufficient to induce neuroplastic changes in the
highly sensitive adolescent brain, leading to increased
risk of later life addiction not only to nicotine and alcohol
but also to other drugs of abuse (Table 3). The adolescent/
early adulthood period in rodents is also sensitive to
nondrug interventions, such as effects of environmental
enrichment (reviewed in Nithianantharajah and
Hannan, 2006).
A. Novel Methods for Future Studies
Further investigations of the mechanisms of action
and persistent neuroplasticity induced by drugs of
abuse are clearly warranted. Recent methodological
advances in systems neurobiology may provide new
tools for such investigations. In the past decade or so,
a variety of techniques with unprecedented resolution
has been developed. Research on long-term neuroplasticity in addiction can benefit tremendously from the

967

continued use of these techniques and incorporation of


others. In this regard, advances in optical techniques
have enabled us to manipulate the activity of certain
classes of neurons and determine their causal role in
brain function. As such manipulations (e.g., optogenetics) have high temporal resolution, their effect on the
firing characteristics of neurons, using optical and
electrical readouts, can be simultaneously monitored
along with their effect on behavior. This has allowed us
to observe how changes in behavior correlate with
changes in the "neural code" (the firing pattern of
populations of neurons).
Closed-loop optogenetics (i.e., optical stimulation of
the brain when a certain criteria is met, e.g., on or before
SA or when a certain pattern of activity is observed in
the brain) are helping us further understand addiction.
Indeed, it is conceivable that in the future we might be
able to manipulate, using optogenetics and chemogenetics, increasingly more specific classes of neurons
(e.g., only the projections from and to certain regions in
the brain that selectively express three or more genes).
Furthermore, recent developments in optogenetic technologies including red-shifted opsins (Chuong et al.,
2014) will enable simultaneous optogenetic stimulation
of two genetically distinct cell types or projections and
hence, elucidate the causal role of different populations
of neurons with unprecedented temporal precision (e.g.,
the behavioral effects of simultaneous manipulation of
D2R- and D1R-MSNs). Additionally, direct optical control of intracellular pathways (reviewed in Stuber and
Mason, 2013) can provide an understanding of the
role of specific intracellular signaling pathways in
a specific population of neurons with similar temporal
resolution. At the other end of the spectrum, optofMRI,
which involves the combination of optogenetics and
fMRI, can elucidate changes in functional connectivity
at the level of the brain (Lee, 2011). Similarly, recent
advances in optical imaging techniques, including development of genetically encoded calcium indicators (for
instance, Gcamp6; Chen et al., 2013c) with the necessary temporal resolution to detect action potentialassociated calcium ion changes and suprathreshold
subcellular activity and development of red-shifted
genetically encoded calcium indicators (Rcamp;
Akerboom et al., 2013), will soon enable simultaneous
optical imaging from different populations of neurons in
vivo (using implantable imaging devices, albeit at
a lower resolution) or ex vivo (similar to the ex vivo
brain slice patch-clamping techniques currently in
widespread use but allowing the visualization of
multiple neurons simultaneously) (see Deisseroth and
Schnitzer, 2013).
Monitoring calcium ion dynamics associated with
neuronal activity is advantageous, because calcium ions
play a fundamental role in induction of synaptic plasticity
and hence, possibly in the long-term neuroplasticity
induced by drugs of abuse. On the other hand, recent

968

Korpi et al.
TABLE 3
Examples of the long-term effects of postnatal/adolescent drug exposures on the brain and behavior in adulthood

Drug/ Intervention

Alcohol
Nicotine

Period

Effects in Adulthood

P2545
P3043
P3648
P3443

Increased social anxiety


Increased alcohol drinking
Disruption of locomotor habituation
Increase in nicotine intake, disturbed
attentional behavior, increased impulsiveness
Decreases reward sensitivity, increased stress
sensitivity, vulnerability to depression
Increased stimulant effects of cocaine,
abnormal shift in attention, decreased
anxiety, deficient contextual fear response
Increased locomotor response to morphine
Increased reinstatement of morphine CPP
Reduced social interaction and novel object recognition
Increased heroin SA
Impaired cognitive behavior, model for psychosis
Reduced exploration, normal social interaction
A great number of neuroplasticity changes in
neuropsychiatric animal models

P30-44
Cocaine
Morphine
-THC
9

Phencyclidine
Environmental
enrichment

P3546
P3032
P3742
P3051
P2849
P515
P3542
Prolonged, at
variable times
after weaning
P6096
P3060

Multiple structural and gene expression


changes in the brain
Transient decrease in cocaine seeking

advancements in genetically encoded voltage indicators


enable the simultaneous detection of voltage fluctuations
in multiple neurons and combined with optogenetics
enable "opto-patching" of neurons ex vivo, if not in vivo
yet (Hochbaum et al., 2014). These advancements can
greatly benefit the drug research and addiction fields
because ex vivo techniques have been very successfully
used thus far.
Similarly, the burgeoning field of chemogenetics
provides the opportunity to specifically modulate (via
GPCRs or ion channels) a certain genetically defined
and/or spatially defined population of neurons using
systemic administration of synthetic ligands, which do
not affect other neurons. This provides the advantage
over optogenetics of performing noninvasive chronic
interventions even across spatially nonrestricted population of neurons (for instance, D2R-expressing neurons across the brain) and possibly simultaneous
activation or inhibition of multiple pathways across
the brain (see Sternson and Roth, 2014).
Furthermore, optrodes with high-density silicon
microelectrodes coupled with multiple microoptogenetic stimulation points (for optical tagging and
stimulation of small regions) allow chronic acquisition
of neuronal activity from large-scale "identified" neuronal ensembles. In addition, novel techniques have been
developed to allow visualization of receptor expression
and the anatomic pathways in unprecedented detail
without sectioning the brain (e.g., Hama et al., 2011;
Chung and Deisseroth, 2013). These techniques will
allow for an improved visualization of structural and
biochemical changes induced across the brain during
addiction.
Finally, recent development of genome engineering
technologies based on CRISPR-associated RNA-guided
endonuclease Cas9 (CRISPR-Cas9) enable easy editing

Reference

(reviewed in Spear, 2014)


(Adriani et al., 2004)
(Adriani et al., 2003;
Counotte et al., 2011)
(Iniguez et al., 2009)
(Black et al., 2006; Sillivan et al., 2011)
(White and Holtzman, 2005)
(Schwarz and Bilbo, 2013)
(reviewed in Rubino and Parolaro, 2008)
See the section II/F.
(Metaxas et al., 2014)
(reviewed in Nithianantharajah
and Hannan, 2006)
(Mychasiuk et al., 2014)
(Chauvet et al., 2012)

or removal of any DNA sequence within the endogenous


genome at any age. Indeed, multiple CRISPR-Cas9
mediated genetic manipulations can be performed
simultaneously (Hsu et al., 2014). This technique holds
potential to elucidate the roles of different genes,
at different time points, in the development of addiction.
Targeted knockdown of multiple genes will allow us
to determine the role of non-Mendelian genes in addiction. Furthermore, this technique will enable us to
understand the complex interplay of genes in such
disorders, which was not possible with knockdown of
a single gene.
The techniques discussed in this section allow unprecedented access to probe brain form and function.
Therefore, great insights in to the neural mechanisms
underlying addiction can be expected in the years to
come.
Understanding addiction will also require elucidation
of the genetic and epigenetic changes, the alterations
in the intracellular biochemical pathways, changes
in neural excitability and information integration,
changes in release of neurotransmitters and receptor
expression levels, systems-level neural processing in
different brain structures and the resultant change in
information processing across structures, and, finally,
understanding how these changes translate into altered
behavior. This is by no means a small feat. The issue is
further complicated by the fact that different drugs
affect the brain differently and the role of each of the
aforementioned factors changes as the disease progresses. One of the principal goals of addiction research
is to understand and develop therapies for addiction by
finding commonalities and differences between various
drugs of abuse. Our review of neuroplasticity induced by
cocaine, amphetamine, and related psychostimulants,
nicotine, ethanol, benzodiazepines and other GABA

Drug-Induced Neuroplasticity

ligands, NMDAR antagonists, opioids, cannabinoids,


and hallucinogens offers a step toward consolidating the
current status of knowledge of such commonalities and
the identification of directions for future research.
Authorship Contributions

Wrote or contributed to the writing of the manuscript: Korpi, den


Hollander, Farooq, Vashchinkina, Rajkumar, Nutt, Hyyti, Dawe.
References
aan het Rot M, Collins KA, Murrough JW, Perez AM, Reich DL, Charney DS,
and Mathew SJ (2010) Safety and efficacy of repeated-dose intravenous ketamine
for treatment-resistant depression. Biol Psychiatry 67:139145.
Abanades S, Farr M, Segura M, Pichini S, Barral D, Pacifici R, Pellegrini M,
Fonseca F, Langohr K, and De La Torre R (2006) Gamma-hydroxybutyrate
(GHB) in humans: pharmacodynamics and pharmacokinetics. Ann N Y Acad Sci
1074:559576.
Abdul-Monim Z, Reynolds GP, and Neill JC (2006) The effect of atypical and classical
antipsychotics on sub-chronic PCP-induced cognitive deficits in a reversal-learning
paradigm. Behav Brain Res 169:263273.
Abraham HD, Aldridge AM, and Gogia P (1996) The psychopharmacology of hallucinogens. Neuropsychopharmacology 14:285298.
Abraham HD and Duffy FH (2001) EEG coherence in post-LSD visual hallucinations.
Psychiatry Res 107:151163.
Abraham WC (2008) Metaplasticity: tuning synapses and networks for plasticity. Nat
Rev Neurosci 9:387399.
Abraham WC and Bear MF (1996) Metaplasticity: the plasticity of synaptic plasticity.
Trends Neurosci 19:126130.
Abrous DN, Adriani W, Montaron MF, Aurousseau C, Rougon G, Le Moal M,
and Piazza PV (2002) Nicotine self-administration impairs hippocampal plasticity.
J Neurosci 22:36563662.
Absalom N, Eghorn LF, Villumsen IS, Karim N, Bay T, Olsen JV, Knudsen GM,
Bruner-Osborne H, Frlund B, and Clausen RP, et al. (2012) a4bd GABA(A)
receptors are high-affinity targets for g-hydroxybutyric acid (GHB). Proc Natl Acad
Sci USA 109:1340413409.
Abulseoud OA, Miller JD, Wu J, Choi D-S, and Holschneider DP (2012) Ceftriaxone
upregulates the glutamate transporter in medial prefrontal cortex and blocks reinstatement of methamphetamine seeking in a condition place preference paradigm. Brain Res 1456:1421.
Achat-Mendes C, Lynch LJ, Sullivan KA, Vallender EJ, and Miller GM (2012a)
Augmentation of methamphetamine-induced behaviors in transgenic mice lacking
the trace amine-associated receptor 1. Pharmacol Biochem Behav 101:201207.
Achat-Mendes C, Platt DM, and Spealman RD (2012b) Antagonism of metabotropic
glutamate 1 receptors attenuates behavioral effects of cocaine and methamphetamine in squirrel monkeys. J Pharmacol Exp Ther 343:214224.
Adachi N, Numakawa T, Kumamaru E, Itami C, Chiba S, Iijima Y, Richards M,
Katoh-Semba R, and Kunugi H (2013) Phencyclidine-induced decrease of synaptic
connectivity via inhibition of BDNF secretion in cultured cortical neurons. Cereb
Cortex 23:847858.
Adam-Vizi V (2005) Production of reactive oxygen species in brain mitochondria:
contribution by electron transport chain and non-electron transport chain sources.
Antioxid Redox Signal 7:11401149.
Adermark L and Lovinger DM (2007) Combined activation of L-type Ca2+ channels
and synaptic transmission is sufficient to induce striatal long-term depression.
J Neurosci 27:67816787.
Adermark L, Talani G, and Lovinger DM (2009) Endocannabinoid-dependent plasticity at GABAergic and glutamatergic synapses in the striatum is regulated by
synaptic activity. Eur J Neurosci 29:3241.
Adriani W, Granstrem O, Macri S, Izykenova G, Dambinova S, and Laviola G (2004)
Behavioral and neurochemical vulnerability during adolescence in mice: studies
with nicotine. Neuropsychopharmacology 29:869878.
Adriani W, Macr S, Pacifici R, and Laviola G (2002) Peculiar vulnerability to
nicotine oral self-administration in mice during early adolescence. Neuropsychopharmacology 27:212224.
Adriani W, Spijker S, Deroche-Gamonet V, Laviola G, Le Moal M, Smit AB,
and Piazza PV (2003) Evidence for enhanced neurobehavioral vulnerability to
nicotine during periadolescence in rats. J Neurosci 23:47124716.
Agabio R and Colombo G (2014) GABAB receptor ligands for the treatment of alcohol
use disorder: preclinical and clinical evidence. Front Neurosci 8:140.
Agabio R, Preti A, and Gessa GL (2013) Efficacy and tolerability of baclofen in substance use disorders: a systematic review. Eur Addict Res 19:325345.
Aghajanian GK and Marek GJ (1999) Serotonin and hallucinogens. Neuropsychopharmacology 21(2, Suppl)16S23S.
Aguayo LG, Castro P, Mariqueo T, Munoz B, Xiong W, Zhang L, Lovinger DM,
and Homanics GE (2014) Altered sedative effects of ethanol in mice with alpha1
glycine receptor subunits that are insensitive to Gbg modulation. Neuropsychopharmacology. 39:25382548
Ahboucha S, Gamrani H, and Baker G (2012) GABAergic neurosteroids: the endogenous benzodiazepines of acute liver failure. Neurochem Int 60:707714.
Ahboucha S, Pomier-Layrargues G, Mamer O, and Butterworth RF (2006) Increased
levels of pregnenolone and its neuroactive metabolite allopregnanolone in autopsied brain tissue from cirrhotic patients who died in hepatic coma. Neurochem Int
49:372378.
Ahmed SH and Koob GF (1998) Transition from moderate to excessive drug intake:
change in hedonic set point. Science 282:298300.

969

Ahsan HM, de la Pena JB, Botanas CJ, Kim HJ, Yu GY, and Cheong JH (2014)
Conditioned place preference and self-administration induced by nicotine in adolescent and adult rats. Biomol Ther (Seoul) 22:460466.
Aitta-aho T, Mykkynen TP, Panhelainen AE, Vekovischeva OY, Bckstrm P,
and Korpi ER (2012) Importance of GluA1 subunit-containing AMPA glutamate
receptors for morphine state-dependency. PLoS One 7:e38325.
Aizawa H, Bianco IH, Hamaoka T, Miyashita T, Uemura O, Concha ML, Russell C,
Wilson SW, and Okamoto H (2005) Laterotopic representation of left-right information onto the dorso-ventral axis of a zebrafish midbrain target nucleus. Curr
Biol 15:238243.
Aizawa H, Kobayashi M, Tanaka S, Fukai T, and Okamoto H (2012) Molecular
characterization of the subnuclei in rat habenula. J Comp Neurol 520:40514066.
Akbik FV, Bhagat SM, Patel PR, Cafferty WB, and Strittmatter SM (2013) Anatomical plasticity of adult brain is titrated by Nogo Receptor 1. Neuron 77:859866.
Akerboom J, Carreras Caldern N, Tian L, Wabnig S, Prigge M, Tol J, Gordus A,
Orger MB, Severi KE, and Macklin JJ, et al. (2013) Genetically encoded calcium
indicators for multi-color neural activity imaging and combination with optogenetics. Front Mol Neurosci 6:2.
Akerman CJ and Cline HT (2007) Refining the roles of GABAergic signaling during
neural circuit formation. Trends Neurosci 30:382389.
Albertson DN, Pruetz B, Schmidt CJ, Kuhn DM, Kapatos G, and Bannon MJ (2004)
Gene expression profile of the nucleus accumbens of human cocaine abusers: evidence for dysregulation of myelin. J Neurochem 88:12111219.
Albertson DN, Schmidt CJ, Kapatos G, and Bannon MJ (2006) Distinctive profiles of
gene expression in the human nucleus accumbens associated with cocaine and
heroin abuse. Neuropsychopharmacology 31:23042312.
Albuquerque EX, Pereira EF, Alkondon M, and Rogers SW (2009) Mammalian nicotinic acetylcholine receptors: from structure to function. Physiol Rev 89:73120.
Alfonso J, Le Magueresse C, Zuccotti A, Khodosevich K, and Monyer H (2012) Diazepam binding inhibitor promotes progenitor proliferation in the postnatal SVZ
by reducing GABA signaling. Cell Stem Cell 10:7687.
Alho H, Costa E, Ferrero P, Fujimoto M, Cosenza-Murphy D, and Guidotti A (1985)
Diazepam-binding inhibitor: a neuropeptide located in selected neuronal populations of rat brain. Science 229:179182.
Ali SF, Ahmad G, Slikker W Jr, and Body SC (1988) Gestational exposure to phencyclidine (PCP) in rats decreases PCP binding sites in term fetal brain. Int J Dev
Neurosci 6:547552.
Ali SF, Ahmad G, Slikker W Jr, and Bondy SC (1989) Effects of gestational exposure
to phencyclidine: distribution and neurochemical alterations in maternal and fetal
brain. Neurotoxicology 10:383392.
Ali SF, Newport GD, Holson RR, Slikker W Jr, and Bowyer JF (1994) Low environmental temperatures or pharmacologic agents that produce hypothermia decrease methamphetamine neurotoxicity in mice. Brain Res 658:3338.
Alicata D, Chang L, Cloak C, Abe K, and Ernst T (2009) Higher diffusion in striatum
and lower fractional anisotropy in white matter of methamphetamine users. Psychiatry Res 174:18.
Alkadhi KA (2011) Chronic stress and Alzheimers disease-like pathogenesis in a rat
model: prevention by nicotine. Curr Neuropharmacol 9:587597.
Alkadhi KA, Alzoubi KH, Srivareerat M, and Tran TT (2011) Chronic psychosocial
stress exacerbates impairment of synaptic plasticity in b-amyloid rat model of
Alzheimers disease: prevention by nicotine. Curr Alzheimer Res 8:718731.
Alkadhi KA, Srivareerat M, and Tran TT (2010) Intensification of long-term memory
deficit by chronic stress and prevention by nicotine in a rat model of Alzheimers
disease. Mol Cell Neurosci 45:289296.
Allen CP and Leri F (2014) Perseveration in the presence of punishment: the effects
of chronic cocaine exposure and lesions to the prefrontal cortex. Behav Brain Res
261:185192.
Allouche S, Noble F, and Marie N (2014) Opioid receptor desensitization: mechanisms and its link to tolerance. Front Pharmacol 5:280.
Alves E, Summavielle T, Alves CJ, Gomes-da-Silva J, Barata JC, Fernandes E,
Bastos MdeL, Tavares MA, and Carvalho F (2007) Monoamine oxidase-B mediates
ecstasy-induced neurotoxic effects to adolescent rat brain mitochondria. J Neurosci
27:1020310210.
Ameri A (1999) The effects of cannabinoids on the brain. Prog Neurobiol 58:315348.
American Psychiatric Association (1994) Diagnostic and Statistical Manual of Mental
Disorders, 4th ed, American Psychiatric Association, Washington, DC.
Amit Z and Smith BR (1985) A multi-dimensional examination of the positive reinforcing properties of acetaldehyde. Alcohol 2:367370.
Amitai N, Kuczenski R, Behrens MM, and Markou A (2012) Repeated phencyclidine
administration alters glutamate release and decreases GABA markers in the
prefrontal cortex of rats. Neuropharmacology 62:14221431.
Anagnostakis Y, Kastellakis A, and Spyraki C (1998) Dizocilpine (MK-801) and tetrodotoxin influence accumbal dopamine release evoked by intrapallidal morphine.
Eur Neuropsychopharmacol 8:4753.
Anagnostaras SG and Robinson TE (1996) Sensitization to the psychomotor stimulant effects of amphetamine: modulation by associative learning. Behav Neurosci
110:13971414.
Anantharaman D, Chabrier A, Gaborieau V, Franceschi S, Herrero R, Rajkumar T,
Samant T, Mahimkar MB, Brennan P, and McKay JD (2014) Genetic variants in
nicotine addiction and alcohol metabolism genes, oral cancer risk and the propensity to
smoke and drink alcohol: a replication study in India. PLoS One 9:e88240.
Andersen JD and Pouzet B (2004) Spatial memory deficits induced by perinatal treatment of rats with PCP and reversal effect of D-serine. Neuropsychopharmacology 29:
10801090.
Anderson DG, Mariappan SV, Buettner GR, and Doorn JA (2011) Oxidation of
3,4-dihydroxyphenylacetaldehyde, a toxic dopaminergic metabolite, to a semiquinone radical and an ortho-quinone. J Biol Chem 286:2697826986.
Anderson KL and Itzhak Y (2006) Methamphetamine-induced selective dopaminergic neurotoxicity is accompanied by an increase in striatal nitrate in the mouse.
Ann N Y Acad Sci 1074:225233.

970

Korpi et al.

Anderson SM, Famous KR, Sadri-Vakili G, Kumaresan V, Schmidt HD, Bass CE,
Terwilliger EF, Cha JH, and Pierce RC (2008) CaMKII: a biochemical bridge
linking accumbens dopamine and glutamate systems in cocaine seeking. Nat
Neurosci 11:344353.
Andrade R (2011) Serotonergic regulation of neuronal excitability in the prefrontal
cortex. Neuropharmacology 61:382386.
Andr VM, Cepeda C, Cummings DM, Jocoy EL, Fisher YE, William Yang X,
and Levine MS (2010) Dopamine modulation of excitatory currents in the striatum
is dictated by the expression of D1 or D2 receptors and modified by endocannabinoids. Eur J Neurosci 31:1428.
Angoa-Prez M, Kane MJ, Briggs DI, Francescutti DM, Sykes CE, Shah MM, Thomas
DM, and Kuhn DM (2013a) Mephedrone does not damage dopamine nerve endings
of the striatum, but enhances the neurotoxicity of methamphetamine, amphetamine, and MDMA. J Neurochem 125:102110.
Angoa-Prez M, Kane MJ, Francescutti DM, Sykes KE, Shah MM, Mohammed AM,
Thomas DM, and Kuhn DM (2012) Mephedrone, an abused psychoactive component of bath salts and methamphetamine congener, does not cause neurotoxicity
to dopamine nerve endings of the striatum. J Neurochem 120:10971107.
Angoa-Perez M, Kane MJ, Herrera-Mundo N, Francescutti DM, and Kuhn DM
(2013b) Effects of combined treatment with mephedrone and methamphetamine or
3,4-methylenedioxymethamphetamine on serotonin nerve endings of the hippocampus. Life Sci 97:3136
Antolin-Fontes B, Ables JL, Gorlich A, and Ibanez-Tallon I (2015) The habenulointerpeduncular pathway in nicotine aversion and withdrawal. Neuropharmacology 96 (Pt B):213222.
Antonelli T, Govoni BM, Bianchi C, and Beani L (1997) Glutamate regulation of
dopamine release in guinea pig striatal slices. Neurochem Int 30:203209.
Aoyama T, Kotaki H, Sawada Y, and Iga T (1996) Pharmacokinetics and pharmacodynamics of methylphenidate enantiomers in rats. Psychopharmacology (Berl)
127:117122.
Araque A, Carmignoto G, and Haydon PG (2001) Dynamic signaling between
astrocytes and neurons. Annu Rev Physiol 63:795813.
Araujo JA, Chan AD, Winka LL, Seymour PA, and Milgram NW (2004) Dose-specific
effects of scopolamine on canine cognition: impairment of visuospatial memory, but
not visuospatial discrimination. Psychopharmacology (Berl) 175:9298.
Ares-Santos S, Granado N, Espadas I, Martinez-Murillo R, and Moratalla R (2014)
Methamphetamine causes degeneration of dopamine cell bodies and terminals of
the nigrostriatal pathway evidenced by silver staining. Neuropsychopharmacology
39:10661080.
Ares-Santos S, Granado N, Oliva I, OShea E, Martin ED, Colado MI, and Moratalla
R (2012) Dopamine D(1) receptor deletion strongly reduces neurotoxic effects of
methamphetamine. Neurobiol Dis 45:810820.
Arvalo-Martn A, Garca-Ovejero D, Rubio-Araiz A, Gmez O, Molina-Holgado F,
and Molina-Holgado E (2007) Cannabinoids modulate Olig2 and polysialylated
neural cell adhesion molecule expression in the subventricular zone of post-natal
rats through cannabinoid receptor 1 and cannabinoid receptor 2. Eur J Neurosci
26:15481559.
Arias-Cavieres A, Rozas C, Reyes-Parada M, Barrera N, Pancetti F, Loyola S, Lorca
RA, Zeise ML, and Morales B (2010) MDMA (ecstasy) impairs learning in the
Morris Water Maze and reduces hippocampal LTP in young rats. Neurosci Lett
469:375379.
Ariwodola OJ, Crowder TL, Grant KA, Daunais JB, Friedman DP, and Weiner JL
(2003) Ethanol modulation of excitatory and inhibitory synaptic transmission in
rat and monkey dentate granule neurons. Alcohol Clin Exp Res 27:16321639.
Armstrong BD and Noguchi KK (2004) The neurotoxic effects of 3,4-methylenedioxymethamphetamine (MDMA) and methamphetamine on serotonin, dopamine, and GABA-ergic terminals: an in-vitro autoradiographic study in rats.
Neurotoxicology 25:905914.
Arnold MM, Loughlin SE, Belluzzi JD, and Leslie FM (2014) Reinforcing and neural
activating effects of norharmane, a non-nicotine tobacco constituent, alone and in
combination with nicotine. Neuropharmacology 85:293304.
Arnt J, Bang-Andersen B, Grayson B, Bymaster FP, Cohen MP, DeLapp NW, Giethlen
B, Kreilgaard M, McKinzie DL, and Neill JC, et al. (2010) Lu AE58054, a 5-HT6
antagonist, reverses cognitive impairment induced by subchronic phencyclidine in
a novel object recognition test in rats. Int J Neuropsychopharmacol 13:10211033.
Arora DS, Nimitvilai S, Teppen TL, McElvain MA, Sakharkar AJ, You C, Pandey SC,
and Brodie MS (2013) Hyposensitivity to gamma-aminobutyric acid in the ventral
tegmental area during alcohol withdrawal: reversal by histone deacetylase inhibitors. Neuropsychopharmacology 38:16741684.
Arria AM and Wish ED (2006) Nonmedical use of prescription stimulants among
students. Pediatr Ann 35:565571.
Asato MR, Terwilliger R, Woo J, and Luna B (2010) White matter development in
adolescence: a DTI study. Cereb Cortex 20:21222131.
Asatryan L, Popova M, Perkins D, Trudell JR, Alkana RL, and Davies DL (2010)
Ivermectin antagonizes ethanol inhibition in purinergic P2X4 receptors. J Pharmacol Exp Ther 334:720728.
Ascoli GA, Alonso-Nanclares L, Anderson SA, Barrionuevo G, Benavides-Piccione R,
Burkhalter A, Buzski G, Cauli B, Defelipe J, and Fairn A, et al.; Petilla Interneuron Nomenclature Group (2008) Petilla terminology: nomenclature of features
of GABAergic interneurons of the cerebral cortex. Nat Rev Neurosci 9:557568.
Ashby CR Jr, Paul M, Gardner EL, Gerasimov MR, Dewey SL, Lennon IC, and Taylor
SJ (2002) Systemic administration of 1R,4S-4-amino-cyclopent-2-ene-carboxylic
acid, a reversible inhibitor of GABA transaminase, blocks expression of conditioned
place preference to cocaine and nicotine in rats. Synapse 44:6163.
Ashby CR Jr, Rohatgi R, Ngosuwan J, Borda T, Gerasimov MR, Morgan AE, Kushner
S, Brodie JD, and Dewey SL (1999) Implication of the GABA(B) receptor in gamma
vinyl-GABAs inhibition of cocaine-induced increases in nucleus accumbens dopamine. Synapse 31:151153.

Ator NA, Griffiths RR, and Weerts EM (2005) Self-injection of flunitrazepam alone
and in the context of methadone maintenance in baboons. Drug Alcohol Depend 78:
113123.
Atwood BK, Kupferschmidt DA, and Lovinger DM (2014) Opioids induce dissociable
forms of long-term depression of excitatory inputs to the dorsal striatum. Nat
Neurosci 17:540548.
Auclair N, Otani S, Soubrie P, and Crepel F (2000) Cannabinoids modulate synaptic
strength and plasticity at glutamatergic synapses of rat prefrontal cortex pyramidal neurons. J Neurophysiol 83:32873293.
Avallone R, Zeneroli ML, Venturini I, Corsi L, Schreier P, Kleinschnitz M, Ferrarese
C, Farina F, Baraldi C, and Pecora N, et al. (1998) Endogenous benzodiazepine-like
compounds and diazepam binding inhibitor in serum of patients with liver cirrhosis with and without overt encephalopathy. Gut 42:861867.
Avelar AJ, Juliano SA, and Garris PA (2013) Amphetamine augments vesicular dopamine release in the dorsal and ventral striatum through different mechanisms.
J Neurochem 125:373385.
Aydar E, Palmer CP, Klyachko VA, and Jackson MB (2002) The sigma receptor as
a ligand-regulated auxiliary potassium channel subunit. Neuron 34:399410.
Azad SC, Monory K, Marsicano G, Cravatt BF, Lutz B, Zieglgnsberger W,
and Rammes G (2004) Circuitry for associative plasticity in the amygdala involves
endocannabinoid signaling. J Neurosci 24:99539961.
Bachs L and Mrland H (2001) Acute cardiovascular fatalities following cannabis use.
Forensic Sci Int 124:200203.
Badiani A (2013) Substance-specific environmental influences on drug use and drug
preference in animals and humans. Curr Opin Neurobiol 23:588596.
Badiani A, Belin D, Epstein D, Calu D, and Shaham Y (2011) Opiate versus psychostimulant addiction: the differences do matter. Nat Rev Neurosci 12:685700.
Baenziger JE and Corringer PJ (2011) 3D structure and allosteric modulation of the
transmembrane domain of pentameric ligand-gated ion channels. Neuropharmacology 60:116125.
Baggott MJ, Coyle JR, Erowid E, Erowid F, and Robertson LC (2011) Abnormal
visual experiences in individuals with histories of hallucinogen use: a Web-based
questionnaire. Drug Alcohol Depend 114:6167.
Baimel C and Borgland SL (2012) Hypocretin modulation of drug-induced synaptic
plasticity. Prog Brain Res 198:123131.
Baimel C, Borgland SL, and Corrigall W (2012) Cocaine and nicotine research
illustrates a range of hypocretin mechanisms in addiction. Vitam Horm 89:
291313.
Bajo M, Cruz MT, Siggins GR, Messing R, and Roberto M (2008) Protein kinase C
epsilon mediation of CRF- and ethanol-induced GABA release in central amygdala.
Proc Natl Acad Sci USA 105:84108415.
Baker DA, McFarland K, Lake RW, Shen H, Tang XC, Toda S, and Kalivas PW (2003)
Neuroadaptations in cystine-glutamate exchange underlie cocaine relapse. Nat
Neurosci 6:743749.
Baker-Andresen D, Flavell CR, Li X, and Bredy TW (2013) Activation of BDNF
signaling prevents the return of fear in female mice. Learn Mem 20:237240.
Bakhit C, Morgan ME, Peat MA, and Gibb JW (1981) Long-term effects of methamphetamine on the synthesis and metabolism of 5-hydroxytryptamine in various
regions of the rat brain. Neuropharmacology 20 (12A):11351140.
Baladi MG, Newman AH, Nielsen SM, Hanson GR, and Fleckenstein AE (2014)
Dopamine D(3) receptors contribute to methamphetamine-induced alterations in
dopaminergic neuronal function: role of hyperthermia. Eur J Pharmacol 732:
105110.
Balcita-Pedicino JJ, Omelchenko N, Bell R, and Sesack SR (2011) The inhibitory
influence of the lateral habenula on midbrain dopamine cells: ultrastructural evidence for indirect mediation via the rostromedial mesopontine tegmental nucleus.
J Comp Neurol 519:11431164.
Baldwin DS, Anderson IM, Nutt DJ, Allgulander C, Bandelow B, den Boer JA,
Christmas DM, Davies S, Fineberg N, and Lidbetter N, et al. (2014) Evidencebased pharmacological treatment of anxiety disorders, post-traumatic stress disorder and obsessive-compulsive disorder: a revision of the 2005 guidelines from the
British Association for Psychopharmacology. J Psychopharmacol 28:403439.
Baldwin HA, Colado MI, Murray TK, De Souza RJ, and Green AR (1993) Striatal
dopamine release in vivo following neurotoxic doses of methamphetamine and effect of the neuroprotective drugs, chlormethiazole and dizocilpine. Br J Pharmacol
108:590596.
Balfour DJ, Benwell ME, Birrell CE, Kelly RJ, and Al-Aloul M (1998) Sensitization of
the mesoaccumbens dopamine response to nicotine. Pharmacol Biochem Behav 59:
10211030.
Ball KT, Wellman CL, Fortenberry E, and Rebec GV (2009) Sensitizing regimens of (+/-)
3, 4-methylenedioxymethamphetamine (ecstasy) elicit enduring and differential
structural alterations in the brain motive circuit of the rat. Neuroscience 160:264274.
Ballaz SJ, Perez J, Waselus M, Akil H, and Watson SJ (2013) Interaction between
cholecystokinin and the fibroblast growth factor system in the ventral tegmental
area of selectively bred high- and low-responder rats. Neuroscience 255:6875.
Balleine BW and Dickinson A (1998) Goal-directed instrumental action: contingency and
incentive learning and their cortical substrates. Neuropharmacology 37:407419.
Balster RL and Schuster CR (1973) A comparison of d-amphetamine, l-amphetamine,
and methamphetamine self-administration in rhesus monkeys. Pharmacol Biochem Behav 1:6771.
Bandeira F, Lent R, and Herculano-Houzel S (2009) Changing numbers of neuronal
and non-neuronal cells underlie postnatal brain growth in the rat. Proc Natl Acad
Sci USA 106:1410814113.
Bandler R and Shipley MT (1994) Columnar organization in the midbrain periaqueductal gray: modules for emotional expression? Trends Neurosci 17:379389.
Banerjee A, Shen PJ, Ma S, Bathgate RA, and Gundlach AL (2010) Swim stress
excitation of nucleus incertus and rapid induction of relaxin-3 expression via CRF1
activation. Neuropharmacology 58:145155.
Banke TG and McBain CJ (2006) GABAergic input onto CA3 hippocampal interneurons remains shunting throughout development. J Neurosci 26:1172011725.

Drug-Induced Neuroplasticity
Bansky G, Meier PJ, Riederer E, Walser H, Ziegler WH, and Schmid M (1989) Effects
of the benzodiazepine receptor antagonist flumazenil in hepatic encephalopathy in
humans. Gastroenterology 97:744750.
Bao G, Kang L, Li H, Li Y, Pu L, Xia P, Ma L, and Pei G (2007) Morphine and heroin
differentially modulate in vivo hippocampal LTP in opiate-dependent rat. Neuropsychopharmacology 32:17381749.
Barayuga SM, Pang X, Andres MA, Panee J, and Bellinger FP (2013) Methamphetamine decreases levels of glutathione peroxidases 1 and 4 in SH-SY5Y neuronal cells: protective effects of selenium. Neurotoxicology 37:240246.
Barker JM, Torregrossa MM, and Taylor JR (2012) Low prefrontal PSA-NCAM
confers risk for alcoholism-related behavior. Nat Neurosci 15:13561358.
Barker MJ, Greenwood KM, Jackson M, and Crowe SF (2004a) Cognitive effects of
long-term benzodiazepine use: a meta-analysis. CNS Drugs 18:3748.
Barker MJ, Greenwood KM, Jackson M, and Crowe SF (2004b) Persistence of cognitive
effects after withdrawal from long-term benzodiazepine use: a meta-analysis. Arch
Clin Neuropsychol 19:437454.
Barkus E, Morrison PD, Vuletic D, Dickson JC, Ell PJ, Pilowsky LS, Brenneisen R, Holt
DW, Powell J, and Kapur S, et al. (2011) Does intravenous D9-tetrahydrocannabinol
increase dopamine release? A SPET study. J Psychopharmacol 25:14621468.
Barnes DE and Yaffe K (2011) The projected effect of risk factor reduction on Alzheimers disease prevalence. Lancet Neurol 10:819828.
Barr AM, Markou A, and Phillips AG (2002) A crash course on psychostimulant
withdrawal as a model of depression. Trends Pharmacol Sci 23:475482.
Barria A and Malinow R (2002) Subunit-specific NMDA receptor trafficking to synapses. Neuron 35:345353.
Barrot M, Sesack SR, Georges F, Pistis M, Hong S, and Jhou TC (2012) Braking
dopamine systems: a new GABA master structure for mesolimbic and nigrostriatal
functions. J Neurosci 32:1409414101.
Basavarajappa BS and Subbanna S (2014) CB1 receptor-mediated signaling underlies the hippocampal synaptic, learning, and memory deficits following treatment
with JWH-081, a new component of spice/K2 preparations. Hippocampus 24:
178188.
Bastos AM, Usrey WM, Adams RA, Mangun GR, Fries P, and Friston KJ (2012)
Canonical microcircuits for predictive coding. Neuron 76:695711.
Batista-Brito R, Machold R, Klein C, and Fishell G (2008) Gene expression in cortical
interneuron precursors is prescient of their mature function. Cereb Cortex 18:
23062317.
Battaglia G, Fornai F, Busceti CL, Aloisi G, Cerrito F, De Blasi A, Melchiorri D,
and Nicoletti F (2002) Selective blockade of mGlu5 metabotropic glutamate
receptors is protective against methamphetamine neurotoxicity. J Neurosci 22:
21352141.
Battaglia G, Yeh SY, and De Souza EB (1988) MDMA-induced neurotoxicity:
parameters of degeneration and recovery of brain serotonin neurons. Pharmacol
Biochem Behav 29:269274.
Battaglia G, Yeh SY, OHearn E, Molliver ME, Kuhar MJ, and De Souza EB (1987)
3,4-Methylenedioxymethamphetamine and 3,4-methylenedioxyamphetamine destroy serotonin terminals in rat brain: quantification of neurodegeneration by
measurement of [3H]paroxetine-labeled serotonin uptake sites. J Pharmacol Exp
Ther 242:911916.
Battelli MG, Buonamici L, Abbondanza A, Virgili M, Contestabile A, and Stirpe F
(1995) Excitotoxic increase of xanthine dehydrogenase and xanthine oxidase in the
rat olfactory cortex. Brain Res Dev Brain Res 86:340344.
Bauer EP, Schafe GE, and LeDoux JE (2002) NMDA receptors and L-type voltagegated calcium channels contribute to long-term potentiation and different components of fear memory formation in the lateral amygdala. J Neurosci 22:52395249.
Baumann MH, Ayestas MA Jr, Partilla JS, Sink JR, Shulgin AT, Daley PF, Brandt
SD, Rothman RB, Ruoho AE, and Cozzi NV (2012) The designer methcathinone
analogs, mephedrone and methylone, are substrates for monoamine transporters
in brain tissue. Neuropsychopharmacology 37:11921203.
Baumann MH, Partilla JS, Lehner KR, Thorndike EB, Hoffman AF, Holy M,
Rothman RB, Goldberg SR, Lupica CR, and Sitte HH, et al. (2013) Powerful
cocaine-like actions of 3,4-methylenedioxypyrovalerone (MDPV), a principal
constituent of psychoactive bath salts products. Neuropsychopharmacology 38:
552562.
Baumann MH, Wang X, and Rothman RB (2007) 3,4-Methylenedioxymethamphetamine (MDMA) neurotoxicity in rats: a reappraisal of past and present findings.
Psychopharmacology (Berl) 189:407424.
Bavelier D, Levi DM, Li RW, Dan Y, and Hensch TK (2010) Removing brakes on
adult brain plasticity: from molecular to behavioral interventions. J Neurosci 30:
1496414971.
Beardsley PM, Hayes BA, and Balster RL (1990) The self-administration of MK-801
can depend upon drug-reinforcement history, and its discriminative stimulus
properties are phencyclidine-like in rhesus monkeys. J Pharmacol Exp Ther 252:
953959.
Beaulieu JM and Gainetdinov RR (2011) The physiology, signaling, and pharmacology of dopamine receptors. Pharmacol Rev 63:182217.
Beauvais G, Atwell K, Jayanthi S, Ladenheim B, and Cadet JL (2011) Involvement of
dopamine receptors in binge methamphetamine-induced activation of endoplasmic
reticulum and mitochondrial stress pathways. PLoS One 6:e28946.
Becher A, Green A, Ige AO, Wise A, White JH, and McIlhinney RA (2004) Ectopically
expressed gamma-aminobutyric acid receptor B is functionally down-regulated in
isolated lipid raft-enriched membranes. Biochem Biophys Res Commun 321:
981987.
Beck CH and Fibiger HC (1995) Conditioned fear-induced changes in behavior and in
the expression of the immediate early gene c-fos: with and without diazepam
pretreatment. J Neurosci 15:709720.
Becker B, Wagner D, Koester P, Bender K, Kabbasch C, Gouzoulis-Mayfrank E,
and Daumann J (2013) Memory-related hippocampal functioning in ecstasy and
amphetamine users: a prospective fMRI study. Psychopharmacology (Berl) 225:
923934.

971

Beckerman MA, Van Kempen TA, Justice NJ, Milner TA, and Glass MJ (2013)
Corticotropin-releasing factor in the mouse central nucleus of the amygdala: ultrastructural distribution in NMDA-NR1 receptor subunit expressing neurons as
well as projection neurons to the bed nucleus of the stria terminalis. Exp Neurol
239:120132.
Belin D and Everitt BJ (2008) Cocaine seeking habits depend upon dopaminedependent serial connectivity linking the ventral with the dorsal striatum. Neuron
57:432441.
Bellone C and Lscher C (2006) Cocaine triggered AMPA receptor redistribution is
reversed in vivo by mGluR-dependent long-term depression. Nat Neurosci 9:
636641.
Belluzzi JD, Lee AG, Oliff HS, and Leslie FM (2004) Age-dependent effects of nicotine
on locomotor activity and conditioned place preference in rats. Psychopharmacology (Berl) 174:389395.
Belujon P and Grace AA (2014) Restoring mood balance in depression: ketamine
reverses deficit in dopamine-dependent synaptic plasticity. Biol Psychiatry 76:
927936.
Ben-Ari Y, Gaiarsa JL, Tyzio R, and Khazipov R (2007) GABA: a pioneer transmitter
that excites immature neurons and generates primitive oscillations. Physiol Rev
87:12151284.
Bencherif M, Fowler K, Lukas RJ, and Lippiello PM (1995) Mechanisms of upregulation of neuronal nicotinic acetylcholine receptors in clonal cell lines and
primary cultures of fetal rat brain. J Pharmacol Exp Ther 275:987994.
Benekareddy M, Nair AR, Dias BG, Suri D, Autry AE, Monteggia LM, and Vaidya VA
(2013) Induction of the plasticity-associated immediate early gene Arc by stress
and hallucinogens: role of brain-derived neurotrophic factor. Int J Neuropsychopharmacol 16:405415.
Benito C, Romero JP, Toln RM, Clemente D, Docagne F, Hillard CJ, Guaza C,
and Romero J (2007) Cannabinoid CB1 and CB2 receptors and fatty acid amide
hydrolase are specific markers of plaque cell subtypes in human multiple sclerosis.
J Neurosci 27:23962402.
Benmansour S, Tejani-Butt SM, Hauptmann M, and Brunswick DJ (1992) Lack of
effect of high-dose cocaine on monoamine uptake sites in rat brain measured by
quantitative autoradiography. Psychopharmacology (Berl) 106:459462.
Bennett BA, Hyde CE, Pecora JR, and Clodfelter JE (1993) Differing neurotoxic
potencies of methamphetamine, mazindol, and cocaine in mesencephalic cultures.
J Neurochem 60:14441452.
Bennett BA, Wichems CH, Hollingsworth CK, Davies HM, Thornley C, Sexton T,
and Childers SR (1995) Novel 2-substituted cocaine analogs: uptake and ligand
binding studies at dopamine, serotonin and norepinephrine transport sites in the
rat brain. J Pharmacol Exp Ther 272:11761186.
Benwell ME and Balfour DJ (1992) The effects of acute and repeated nicotine
treatment on nucleus accumbens dopamine and locomotor activity. Br J Pharmacol
105:849856.
Benwell ME, Balfour DJ, and Anderson JM (1988) Evidence that tobacco smoking
increases the density of (-)-[3H]nicotine binding sites in human brain. J Neurochem
50:12431247.
Benwell ME, Balfour DJ, and Birrell CE (1995) Desensitization of the nicotineinduced mesolimbic dopamine responses during constant infusion with nicotine. Br
J Pharmacol 114:454460.
Beraki S, Diaz-Heijtz R, Tai F, and Ogren SO (2009) Effects of repeated treatment of
phencyclidine on cognition and gene expression in C57BL/6 mice. Int J Neuropsychopharmacol 12:243255.
Bergamaschi MM, Karschner EL, Goodwin RS, Scheidweiler KB, Hirvonen J,
Queiroz RH, and Huestis MA (2013) Impact of prolonged cannabinoid excretion in
chronic daily cannabis smokers blood on per se drugged driving laws. Clin Chem
59:519526.
Berghuis P, Dobszay MB, Wang X, Spano S, Ledda F, Sousa KM, Schulte G, Ernfors
P, Mackie K, and Paratcha G, et al. (2005) Endocannabinoids regulate interneuron
migration and morphogenesis by transactivating the TrkB receptor. Proc Natl
Acad Sci USA 102:1911519120.
Bergles DE, Diamond JS, and Jahr CE (1999) Clearance of glutamate inside the
synapse and beyond. Curr Opin Neurobiol 9:293298.
Berman RM, Cappiello A, Anand A, Oren DA, Heninger GR, Charney DS,
and Krystal JH (2000) Antidepressant effects of ketamine in depressed patients.
Biol Psychiatry 47:351354.
Berman SB, Zigmond MJ, and Hastings TG (1996) Modification of dopamine transporter function: effect of reactive oxygen species and dopamine. J Neurochem 67:
593600.
Bernier BE, Whitaker LR, and Morikawa H (2011) Previous ethanol experience
enhances synaptic plasticity of NMDA receptors in the ventral tegmental area. J
Neurosci 31:52055212.
Berrettini W, Yuan X, Tozzi F, Song K, Francks C, Chilcoat H, Waterworth D, Muglia
P, and Mooser V (2008) Alpha-5/alpha-3 nicotinic receptor subunit alleles increase
risk for heavy smoking. Mol Psychiatry 13:368373.
Berton O, McClung CA, Dileone RJ, Krishnan V, Renthal W, Russo SJ, Graham D,
Tsankova NM, Bolanos CA, and Rios M, et al. (2006) Essential role of BDNF in the
mesolimbic dopamine pathway in social defeat stress. Science 311:864868.
Bettinger JC and Davies AG (2014) The role of the BK channel in ethanol response
behaviors: evidence from model organism and human studies. Front Physiol 5:346.
Bettler B, Kaupmann K, Mosbacher J, and Gassmann M (2004) Molecular structure
and physiological functions of GABA(B) receptors. Physiol Rev 84:835867.
Beurdeley M, Spatazza J, Lee HH, Sugiyama S, Bernard C, Di Nardo AA, Hensch
TK, and Prochiantz A (2012) Otx2 binding to perineuronal nets persistently regulates plasticity in the mature visual cortex. J Neurosci 32:94299437.
Beurrier C and Malenka RC (2002) Enhanced inhibition of synaptic transmission by
dopamine in the nucleus accumbens during behavioral sensitization to cocaine.
J Neurosci 22:58175822.

972

Korpi et al.

Beveridge TJ, Mechan AO, Sprakes M, Pei Q, Zetterstrom TS, Green AR, and Elliott
JM (2004) Effect of 5-HT depletion by MDMA on hyperthermia and Arc mRNA
induction in rat brain. Psychopharmacology (Berl) 173:346352.
Beveridge TJ, Smith HR, Nader MA, and Porrino LJ (2009) Abstinence from chronic
cocaine self-administration alters striatal dopamine systems in rhesus monkeys.
Neuropsychopharmacology 34:11621171.
Bhandage AK, Jin Z, Bazov I, Kononenko O, Bakalkin G, Korpi ER, and Birnir B
(2014) GABA-A and NMDA receptor subunit mRNA expression is altered in the
caudate but not the putamen of the postmortem brains of alcoholics. Front Cell
Neurosci 8:415. DOI 10.3389/fncel.2014.00415
Bi G and Poo M (2001) Synaptic modification by correlated activity: Hebbs postulate
revisited. Annu Rev Neurosci 24:139166.
Bierut LJ (2010) Convergence of genetic findings for nicotine dependence and smoking
related diseases with chromosome 15q24-25. Trends Pharmacol Sci 31:4651.
Bierut LJ, Schuckit MA, Hesselbrock V, and Reich T (2000) Co-occurring risk factors
for alcohol dependence and habitual smoking. Alcohol Res Health 24:233241.
Bierut LJ, Stitzel JA, Wang JC, Hinrichs AL, Grucza RA, Xuei X, Saccone NL,
Saccone SF, Bertelsen S, and Fox L, et al. (2008) Variants in nicotinic receptors and
risk for nicotine dependence. Am J Psychiatry 165:11631171.
Billa SK, Liu J, Bjorklund NL, Sinha N, Fu Y, Shinnick-Gallagher P, and Morn JA
(2010) Increased insertion of glutamate receptor 2-lacking alpha-amino-3-hydroxy5-methyl-4-isoxazole propionic acid (AMPA) receptors at hippocampal synapses
upon repeated morphine administration. Mol Pharmacol 77:874883.
Billioti de Gage S, Pariente A, and Bgaud B (2015) Is there really a link between
benzodiazepine use and the risk of dementia? Expert Opin Drug Saf 14:733747.
Binda F, Dipace C, Bowton E, Robertson SD, Lute BJ, Fog JU, Zhang M, Sen N,
Colbran RJ, and Gnegy ME, et al. (2008) Syntaxin 1A interaction with the dopamine transporter promotes amphetamine-induced dopamine efflux. Mol Pharmacol 74:11011108.
Bisogno T, Howell F, Williams G, Minassi A, Cascio MG, Ligresti A, Matias I,
Schiano-Moriello A, Paul P, and Williams EJ, et al. (2003) Cloning of the first sn1DAG lipases points to the spatial and temporal regulation of endocannabinoid
signaling in the brain. J Cell Biol 163:463468.
Bjrklund A and Dunnett SB (2007) Dopamine neuron systems in the brain: an
update. Trends Neurosci 30:194202.
Black YD, Maclaren FR, Naydenov AV, Carlezon WA Jr, Baxter MG, and Konradi C
(2006) Altered attention and prefrontal cortex gene expression in rats after bingelike exposure to cocaine during adolescence. J Neurosci 26:96569665.
Blacktop JM, Seubert C, Baker DA, Ferda N, Lee G, Graf EN, and Mantsch JR (2011)
Augmented cocaine seeking in response to stress or CRF delivered into the ventral
tegmental area following long-access self-administration is mediated by CRF receptor type 1 but not CRF receptor type 2. J Neurosci 31:1139611403.
Blaesse P, Airaksinen MS, Rivera C, and Kaila K (2009) Cation-chloride cotransporters and neuronal function. Neuron 61:820838.
Blednov YA, Benavidez JM, Black M, Chandra D, Homanics GE, Rudolph U,
and Harris RA (2013) Linking GABA(A) receptor subunits to alcohol-induced
conditioned taste aversion and recovery from acute alcohol intoxication. Neuropharmacology 67:4656.
Blednov YA, Walker D, and Harris RA (2004) Blockade of the leptin-sensitive
pathway markedly reduces alcohol consumption in mice. Alcohol Clin Exp Res 28:
16831692.
Bliss TV and Lomo T (1973) Long-lasting potentiation of synaptic transmission in the
dentate area of the anaesthetized rabbit following stimulation of the perforant
path. J Physiol 232:331356.
Bloodgood BL and Sabatini BL (2007) Ca(2+) signaling in dendritic spines. Curr Opin
Neurobiol 17:345351.
Blot K, Bai J, and Otani S (2013) The effect of non-competitive NMDA receptor
antagonist MK-801 on neuronal activity in rodent prefrontal cortex: an animal
model for cognitive symptoms of schizophrenia. J Physiol Paris 107:448451.
Bock R, Shin JH, Kaplan AR, Dobi A, Markey E, Kramer PF, Gremel CM,
Christensen CH, Adrover MF, and Alvarez VA (2013) Strengthening the
accumbal indirect pathway promotes resilience to compulsive cocaine use. Nat
Neurosci 16:632638.
Bocklisch C, Pascoli V, Wong JC, House DR, Yvon C, de Roo M, Tan KR, and Lscher
C (2013) Cocaine disinhibits dopamine neurons by potentiation of GABA transmission in the ventral tegmental area. Science 341:15211525.
Bohn LM, Gainetdinov RR, Lin FT, Lefkowitz RJ, and Caron MG (2000) Mu-opioid
receptor desensitization by beta-arrestin-2 determines morphine tolerance but not
dependence. Nature 408:720723.
Boireau A, Bordier F, Dubdat P, and Doble A (1995) Methamphetamine and dopamine neurotoxicity: differential effects of agents interfering with glutamatergic
transmission. Neurosci Lett 195:912.
Bolla KI, Brown K, Eldreth D, Tate K, and Cadet JL (2002) Dose-related neurocognitive effects of marijuana use. Neurology 59:13371343.
Bolla KI, Eldreth DA, London ED, Kiehl KA, Mouratidis M, Contoreggi C, Matochik JA,
Kurian V, Cadet JL, and Kimes AS, et al. (2003) Orbitofrontal cortex dysfunction in
abstinent cocaine abusers performing a decision-making task. Neuroimage 19:10851094.
Bonci A and Williams JT (1997) Increased probability of GABA release during
withdrawal from morphine. J Neurosci 17:796803.
Bonnet U and Scherbaum N (2012) Craving dominates propofol addiction of an affected physician. J Psychoactive Drugs 44:186190.
Borghese CM, Strustovu S, Ebert B, Herd MB, Belelli D, Lambert JJ, Marshall G,
Wafford KA, and Harris RA (2006) The delta subunit of gamma-aminobutyric acid
type A receptors does not confer sensitivity to low concentrations of ethanol.
J Pharmacol Exp Ther 316:13601368.
Borgland SL, Connor M, Osborne PB, Furness JB, and Christie MJ (2003) Opioid
agonists have different efficacy profiles for G protein activation, rapid desensitization, and endocytosis of mu-opioid receptors. J Biol Chem 278:
1877618784.

Borgland SL, Malenka RC, and Bonci A (2004) Acute and chronic cocaine-induced
potentiation of synaptic strength in the ventral tegmental area: electrophysiological and behavioral correlates in individual rats. J Neurosci 24:74827490.
Borgland SL, Storm E, and Bonci A (2008) Orexin B/hypocretin 2 increases glutamatergic transmission to ventral tegmental area neurons. Eur J Neurosci 28:
15451556.
Borgland SL, Taha SA, Sarti F, Fields HL, and Bonci A (2006) Orexin A in the VTA is
critical for the induction of synaptic plasticity and behavioral sensitization to cocaine. Neuron 49:589601.
Borgland SL, Ungless MA, and Bonci A (2010) Convergent actions of orexin/
hypocretin and CRF on dopamine neurons: Emerging players in addiction. Brain
Res 1314:139144.
Boscolo-Berto R, Viel G, Montagnese S, Raduazzo DI, Ferrara SD, and Dauvilliers Y
(2012) Narcolepsy and effectiveness of gamma-hydroxybutyrate (GHB): a systematic
review and meta-analysis of randomized controlled trials. Sleep Med Rev 16:431443.
Bosman FT and Stamenkovic I (2003) Functional structure and composition of the
extracellular matrix. J Pathol 200:423428.
Bossong MG and Niesink RJ (2010) Adolescent brain maturation, the endogenous
cannabinoid system and the neurobiology of cannabis-induced schizophrenia. Prog
Neurobiol 92:370385.
Bossong MG, van Berckel BN, Boellaard R, Zuurman L, Schuit RC, Windhorst AD,
van Gerven JM, Ramsey NF, Lammertsma AA, and Kahn RS (2009) Delta 9tetrahydrocannabinol induces dopamine release in the human striatum. Neuropsychopharmacology 34:759766.
Botly LC, Burton CL, Rizos Z, and Fletcher PJ (2008) Characterization of methylphenidate self-administration and reinstatement in the rat. Psychopharmacology
(Berl) 199:5566.
Boudreau AC, Reimers JM, Milovanovic M, and Wolf ME (2007) Cell surface AMPA
receptors in the rat nucleus accumbens increase during cocaine withdrawal but
internalize after cocaine challenge in association with altered activation of
mitogen-activated protein kinases. J Neurosci 27:1062110635.
Boudreau AC and Wolf ME (2005) Behavioral sensitization to cocaine is associated
with increased AMPA receptor surface expression in the nucleus accumbens.
J Neurosci 25:91449151.
Bourdy R and Barrot M (2012) A new control center for dopaminergic systems:
pulling the VTA by the tail. Trends Neurosci 35:681690.
Bouwmeester H, Smits K, and Van Ree JM (2002) Neonatal development of projections to the basolateral amygdala from prefrontal and thalamic structures in rat. J
Comp Neurol 450:241255.
Bowery NG, Bettler B, Froestl W, Gallagher JP, Marshall F, Raiteri M, Bonner TI,
and Enna SJ (2002) International Union of Pharmacology. XXXIII. Mammalian
gamma-aminobutyric acid(B) receptors: structure and function. Pharmacol Rev 54:
247264.
Bowyer JF (1995) The role of hyperthermia in amphetamines interactions with
NMDA receptors, nitric oxide, and age to produce neurotoxicity. Ann N Y Acad Sci
765:309310.
Bowyer JF, Davies DL, Schmued L, Broening HW, Newport GD, Slikker W Jr,
and Holson RR (1994) Further studies of the role of hyperthermia in methamphetamine neurotoxicity. J Pharmacol Exp Ther 268:15711580.
Bowyer JF and Schmued LC (2006) Fluoro-Ruby labeling prior to an amphetamine
neurotoxic insult shows a definitive massive loss of dopaminergic terminals and
axons in the caudate-putamen. Brain Res 1075:236239.
Bradbury S, Gittings D, and Schenk S (2012) Repeated exposure to MDMA and
amphetamine: sensitization, cross-sensitization, and response to dopamine D- and
D-like agonists. Psychopharmacology (Berl) 223:389399.
Braestrup C (1977) Biochemical differentiation of amphetamine vs methylphenidate
and nomifensine in rats. J Pharm Pharmacol 29:463470.
Braida D, Limonta V, Pegorini S, Zani A, Guerini-Rocco C, Gori E, and Sala M (2007)
Hallucinatory and rewarding effect of salvinorin A in zebrafish: kappa-opioid and
CB1-cannabinoid receptor involvement. Psychopharmacology (Berl) 190:441448.
Brambrink AM, Back SA, Riddle A, Gong X, Moravec MD, Dissen GA, Creeley CE,
Dikranian KT, and Olney JW (2012a) Isoflurane-induced apoptosis of oligodendrocytes in the neonatal primate brain. Ann Neurol 72:525535.
Brambrink AM, Evers AS, Avidan MS, Farber NB, Smith DJ, Martin LD, Dissen GA,
Creeley CE, and Olney JW (2012b) Ketamine-induced neuroapoptosis in the fetal
and neonatal rhesus macaque brain. Anesthesiology 116:372384.
Branch SY and Beckstead MJ (2012) Methamphetamine produces bidirectional,
concentration-dependent effects on dopamine neuron excitability and dopaminemediated synaptic currents. J Neurophysiol 108:802809.
Breese CR, Marks MJ, Logel J, Adams CE, Sullivan B, Collins AC, and Leonard S
(1997) Effect of smoking history on [3H]nicotine binding in human postmortem
brain. J Pharmacol Exp Ther 282:713.
Breier JM, Bankson MG, and Yamamoto BK (2006) L-tyrosine contributes to (+)-3,4methylenedioxymethamphetamine-induced serotonin depletions. J Neurosci 26:
290299.
Breiter HC, Gollub RL, Weisskoff RM, Kennedy DN, Makris N, Berke JD, Goodman
JM, Kantor HL, Gastfriend DR, and Riorden JP, et al. (1997) Acute effects of
cocaine on human brain activity and emotion. Neuron 19:591611.
Brenhouse HC and Andersen SL (2011) Developmental trajectories during adolescence in males and females: a cross-species understanding of underlying brain
changes. Neurosci Biobehav Rev 35:16871703.
Brennan JL, Leung JG, Gagliardi JP, Rivelli SK, and Muzyk AJ (2013) Clinical
effectiveness of baclofen for the treatment of alcohol dependence: a review. Clin
Pharmacol 5:99107.
Brickley SG, Cull-Candy SG, and Farrant M (1996) Development of a tonic form of
synaptic inhibition in rat cerebellar granule cells resulting from persistent activation of GABAA receptors. J Physiol 497:753759.
Brickley SG, Revilla V, Cull-Candy SG, Wisden W, and Farrant M (2001) Adaptive
regulation of neuronal excitability by a voltage-independent potassium conductance. Nature 409:8892.

Drug-Induced Neuroplasticity
Brielmaier J, McDonald CG, and Smith RF (2012) Effects of acute stress on acquisition of nicotine conditioned place preference in adolescent rats: a role for
corticotropin-releasing factor 1 receptors. Psychopharmacology (Berl) 219:7382.
Briner A, De Roo M, Dayer A, Muller D, Habre W, and Vutskits L (2010) Volatile
anesthetics rapidly increase dendritic spine density in the rat medial prefrontal
cortex during synaptogenesis. Anesthesiology 112:546556.
Briner A, Nikonenko I, De Roo M, Dayer A, Muller D, and Vutskits L (2011) Developmental Stage-dependent persistent impact of propofol anesthesia on dendritic
spines in the rat medial prefrontal cortex. Anesthesiology 115:282293.
Brischoux F, Chakraborty S, Brierley DI, and Ungless MA (2009) Phasic excitation of
dopamine neurons in ventral VTA by noxious stimuli. Proc Natl Acad Sci USA 106:
48944899.
Britt JP, Benaliouad F, McDevitt RA, Stuber GD, Wise RA, and Bonci A (2012)
Synaptic and behavioral profile of multiple glutamatergic inputs to the nucleus
accumbens. Neuron 76:790803.
Broadbear JH, Winger G, and Woods JH (2004) Self-administration of fentanyl, cocaine and ketamine: effects on the pituitary-adrenal axis in rhesus monkeys.
Psychopharmacology (Berl) 176:398406.
Brodie JD, Case BG, Figueroa E, Dewey SL, Robinson JA, Wanderling JA, and Laska
EM (2009) Randomized, double-blind, placebo-controlled trial of vigabatrin for the
treatment of cocaine dependence in Mexican parolees. Am J Psychiatry 166:
12691277.
Brodie MS and Appel SB (1998) The effects of ethanol on dopaminergic neurons of the
ventral tegmental area studied with intracellular recording in brain slices. Alcohol
Clin Exp Res 22:236244.
Brody AL, Mandelkern MA, Olmstead RE, Allen-Martinez Z, Scheibal D, Abrams AL,
Costello MR, Farahi J, Saxena S, and Monterosso J, et al. (2009) Ventral striatal
dopamine release in response to smoking a regular vs a denicotinized cigarette.
Neuropsychopharmacology 34:282289.
Broening HW, Bowyer JF, and Slikker W Jr (1995) Age-dependent sensitivity of
rats to the long-term effects of the serotonergic neurotoxicant (+/-)-3,4methylenedioxymethamphetamine (MDMA) correlates with the magnitude of
the MDMA-induced thermal response. J Pharmacol Exp Ther 275:325333.
Bromberg-Martin ES, Matsumoto M, and Hikosaka O (2010) Dopamine in motivational control: rewarding, aversive, and alerting. Neuron 68:815834.
Brook JS, Schuster E, and Zhang C (2004) Cigarette smoking and depressive symptoms:
a longitudinal study of adolescents and young adults. Psychol Rep 95:159166.
Brooks WJ, Petit TL, and LeBoutillier JC (1997) Effect of chronic administration of
NMDA antagonists on synaptic development. Synapse 26:104113.
Brown JM, Hanson GR, and Fleckenstein AE (2000) Methamphetamine rapidly
decreases vesicular dopamine uptake. J Neurochem 74:22212223.
Brown JM, Quinton MS, and Yamamoto BK (2005) Methamphetamine-induced inhibition of mitochondrial complex II: roles of glutamate and peroxynitrite.
J Neurochem 95:429436.
Brown JM and Yamamoto BK (2003) Effects of amphetamines on mitochondrial
function: role of free radicals and oxidative stress. Pharmacol Ther 99:4553.
Brown RA, Lewinsohn PM, Seeley JR, and Wagner EF (1996) Cigarette smoking,
major depression, and other psychiatric disorders among adolescents. J Am Acad
Child Adolesc Psychiatry 35:16021610.
Brown TE, Lee BR, Mu P, Ferguson D, Dietz D, Ohnishi YN, Lin Y, Suska A, Ishikawa M, and Huang YH, et al. (2011) A silent synapse-based mechanism for
cocaine-induced locomotor sensitization. J Neurosci 31:81638174.
Brown TE, Lee BR, and Sorg BA (2008) The NMDA antagonist MK-801 disrupts
reconsolidation of a cocaine-associated memory for conditioned place preference
but not for self-administration in rats. Learn Mem 15:857865.
Brckner MK, Rossner S, and Arendt T (1997) Differential changes in the expression
of AMPA receptors genes in rat brain after chronic exposure to ethanol: an in situ
hybridization study. J Hirnforsch 38:369376.
Brunzell DH, Russell DS, and Picciotto MR (2003) In vivo nicotine treatment regulates mesocorticolimbic CREB and ERK signaling in C57Bl/6J mice. J Neurochem
84:14311441.
Buchert R, Obrocki J, Thomasius R, Vterlein O, Petersen K, Jenicke L, Bohuslavizki
KH, and Clausen M (2001) Long-term effects of ecstasy abuse on the human brain
studied by FDG PET. Nucl Med Commun 22:889897.
Buck KJ, Heim H, and Harris RA (1991) Reversal of alcohol dependence and tolerance by a single administration of flumazenil. J Pharmacol Exp Ther 257:984989.
Buck N, Cali S, and Behr J (2006) Enhancement of long-term potentiation at CA1subiculum synapses in MK-801-treated rats. Neurosci Lett 392:59.
Buckingham SD, Jones AK, Brown LA, and Sattelle DB (2009) Nicotinic acetylcholine receptor signalling: roles in Alzheimers disease and amyloid neuroprotection.
Pharmacol Rev 61:3961.
Buffalari DM and See RE (2011) Inactivation of the bed nucleus of the stria terminalis in an animal model of relapse: effects on conditioned cue-induced reinstatement and its enhancement by yohimbine. Psychopharmacology (Berl) 213:
1927.
Buffington SA, Huang W, and Costa-Mattioli M (2014) Translational control in
synaptic plasticity and cognitive dysfunction. Annu Rev Neurosci 37:1738.
Bukiya AN, Kuntamallappanavar G, Edwards J, Singh AK, Shivakumar B,
and Dopico AM (2014) An alcohol-sensing site in the calcium- and voltage-gated,
large conductance potassium (BK) channel. Proc Natl Acad Sci USA 111:
93139318.
Bull EJ, Porkess V, Rigby M, Hutson PH, and Fone KC (2006) Pre-treatment with
3,4-methylenedioxymethamphetamine (MDMA) causes long-lasting changes in
5-HT2A receptor-mediated glucose utilization in the rat brain. J Psychopharmacol
20:272280.
Burgard EC, Decker G, and Sarvey JM (1989) NMDA receptor antagonists block
norepinephrine-induced long-lasting potentiation and long-term potentiation in rat
dentate gyrus. Brain Res 482:351355.

973

Burnashev N, Schoepfer R, Monyer H, Ruppersberg JP, Gnther W, Seeburg PH,


and Sakmann B (1992) Control by asparagine residues of calcium permeability and
magnesium blockade in the NMDA receptor. Science 257:14151419.
Burns BE and Proctor WR (2013) Cigarette smoke exposure greatly increases alcohol
consumption in adolescent C57BL/6 mice. Alcohol Clin Exp Res 37 (Suppl 1):
E364E372.
Burns JM and Boyer EW (2013) Antitussives and substance abuse. Subst Abuse
Rehabil 4:7582.
Burrows KB, Gudelsky G, and Yamamoto BK (2000) Rapid and transient inhibition
of mitochondrial function following methamphetamine or 3,4-methylenedioxymethamphetamine administration. Eur J Pharmacol 398:1118.
Butovsky E, Juknat A, Goncharov I, Elbaz J, Eilam R, Zangen A, and Vogel Z (2005)
In vivo up-regulation of brain-derived neurotrophic factor in specific brain areas by
chronic exposure to Delta-tetrahydrocannabinol. J Neurochem 93:802811.
Bymaster FP, Katner JS, Nelson DL, Hemrick-Luecke SK, Threlkeld PG, Heiligenstein JH, Morin SM, Gehlert DR, and Perry KW (2002) Atomoxetine increases
extracellular levels of norepinephrine and dopamine in prefrontal cortex of rat:
a potential mechanism for efficacy in attention deficit/hyperactivity disorder.
Neuropsychopharmacology 27:699711.
Byun MS, Kim JS, Jung WH, Jang JH, Choi JS, Kim SN, Choi CH, Chung CK, An
SK, and Kwon JS (2012) Regional cortical thinning in subjects with high genetic
loading for schizophrenia. Schizophr Res 141:197203.
Cabib S and Puglisi-Allegra S (2012) The mesoaccumbens dopamine in coping with
stress. Neurosci Biobehav Rev 36:7989.
Cadet JL, Jayanthi S, and Deng X (2003) Speed kills: cellular and molecular bases of
methamphetamine-induced nerve terminal degeneration and neuronal apoptosis.
FASEB J 17:17751788.
Cadet JL, Jayanthi S, and Deng X (2005) Methamphetamine-induced neuronal apoptosis involves the activation of multiple death pathways. Review. Neurotox Res 8:
199206.
Cadet JL, Krasnova IN, Ladenheim B, Cai NS, McCoy MT, and Atianjoh FE (2009)
Methamphetamine preconditioning: differential protective effects on monoaminergic systems in the rat brain. Neurotox Res 15:252259.
Cadet JL, Ordonez SV, and Ordonez JV (1997) Methamphetamine induces apoptosis in
immortalized neural cells: protection by the proto-oncogene, bcl-2. Synapse 25:176184.
Cagetti E, Liang J, Spigelman I, and Olsen RW (2003) Withdrawal from chronic
intermittent ethanol treatment changes subunit composition, reduces synaptic
function, and decreases behavioral responses to positive allosteric modulators of
GABAA receptors. Mol Pharmacol 63:5364.
Calabresi P, Picconi B, Tozzi A, and Di Filippo M (2007) Dopamine-mediated regulation of corticostriatal synaptic plasticity. Trends Neurosci 30:211219.
Caldji C, Tannenbaum B, Sharma S, Francis D, Plotsky PM, and Meaney MJ (1998)
Maternal care during infancy regulates the development of neural systems mediating the expression of fearfulness in the rat. Proc Natl Acad Sci USA 95:
53355340.
Calipari ES and Ferris MJ (2013) Amphetamine mechanisms and actions at the
dopamine terminal revisited. J Neurosci 33:89238925.
Callado LF, Hopwood SE, Hancock PJ, and Stamford JA (2000) Effects of dizocilpine
(MK 801) on noradrenaline, serotonin and dopamine release and uptake. Neuroreport 11:173176.
Callaerts-Vegh Z, Hoyer D, and Kelly PH (2009) Selective effects of benzodiazepines
on the acquisition of conditioned taste aversion compared to attenuation of neophobia in C57BL/6 mice. Psychopharmacology (Berl) 206:389401.
Callahan B, Yuan J, Stover G, Hatzidimitriou G, and Ricaurte G (1998) Effects of
2-deoxy-D-glucose on methamphetamine-induced dopamine and serotonin neurotoxicity. J Neurochem 70:190197.
Calvigioni D, Hurd YL, Harkany T, and Keimpema E (2014) Neuronal substrates and
functional consequences of prenatal cannabis exposure. Eur Child Adolesc Psychiatry 23:931941.
Camarasa J, Rodrigo T, Pubill D, and Escubedo E (2010) Memantine is a useful drug
to prevent the spatial and non-spatial memory deficits induced by methamphetamine in rats. Pharmacol Res 62:450456.
Cameron K, Kolanos R, Vekariya R, De Felice L, and Glennon RA (2013) Mephedrone
and methylenedioxypyrovalerone (MDPV), major constituents of bath salts,
produce opposite effects at the human dopamine transporter. Psychopharmacology
(Berl) 227:493499.
Canal CE and Morgan D (2012) Head-twitch response in rodents induced by the
hallucinogen 2,5-dimethoxy-4-iodoamphetamine: a comprehensive history, a reevaluation of mechanisms, and its utility as a model. Drug Test Anal 4:556576.
Canales JJ and Graybiel AM (2000) A measure of striatal function predicts motor
stereotypy. Nat Neurosci 3:377383.
Canals S, Beyerlein M, Merkle H, and Logothetis NK (2009) Functional MRI evidence for LTP-induced neural network reorganization. Curr Biol 19:398403.
Capela JP, Carmo H, Remio F, Bastos ML, Meisel A, and Carvalho F (2009) Molecular and cellular mechanisms of ecstasy-induced neurotoxicity: an overview. Mol
Neurobiol 39:210271.
Cappon GD, Morford LL, and Vorhees CV (1998) Enhancement of cocaine-induced
hyperthermia fails to elicit neurotoxicity. Neurotoxicol Teratol 20:531535.
Carboni E, Imperato A, Perezzani L, and Di Chiara G (1989) Amphetamine, cocaine,
phencyclidine and nomifensine increase extracellular dopamine concentrations
preferentially in the nucleus accumbens of freely moving rats. Neuroscience 28:
653661.
Carey RJ and Goodall EB (1974) Amphetamine-induced taste aversion: a comparison
of d- versus l-amphetamine. Pharmacol Biochem Behav 2:325330.
Carhart-Harris RL, Erritzoe D, Williams T, Stone JM, Reed LJ, Colasanti A, Tyacke
RJ, Leech R, Malizia AL, and Murphy K, et al. (2012) Neural correlates of the
psychedelic state as determined by fMRI studies with psilocybin. Proc Natl Acad
Sci USA 109:21382143.
Carhart-Harris RL, Murphy K, Leech R, Erritzoe D, Wall MB, Ferguson B, Williams
LT, Roseman L, Brugger S, and De Meer I, et al. (2014) The effects of acutely

974

Korpi et al.

administered 3,4-methylenedioxymethamphetamine on spontaneous brain function in healthy volunteers measured with arterial spin labeling and blood oxygen
level-dependent resting state functional connectivity. Biol Psychiatry S0006-3223
(14)00005-5.
Carlezon WA Jr, Boundy VA, Haile CN, Lane SB, Kalb RG, Neve RL, and Nestler EJ
(1997) Sensitization to morphine induced by viral-mediated gene transfer. Science
277:812814.
Carlezon WA Jr and Nestler EJ (2002) Elevated levels of GluR1 in the midbrain:
a trigger for sensitization to drugs of abuse? Trends Neurosci 25:610615.
Carlezon WA Jr, Thome J, Olson VG, Lane-Ladd SB, Brodkin ES, Hiroi N, Duman
RS, Neve RL, and Nestler EJ (1998) Regulation of cocaine reward by CREB. Science 282:22722275.
Carlezon WA Jr and Wise RA (1996) Rewarding actions of phencyclidine and related
drugs in nucleus accumbens shell and frontal cortex. J Neurosci 16:31123122.
Carlson J, Armstrong B, Switzer RC 3rd, and Ellison G (2000) Selective neurotoxic
effects of nicotine on axons in fasciculus retroflexus further support evidence that
this a weak link in brain across multiple drugs of abuse. Neuropharmacology 39:
27922798.
Carlson J, Noguchi K, and Ellison G (2001) Nicotine produces selective degeneration
in the medial habenula and fasciculus retroflexus. Brain Res 906:127134.
Caroni P, Donato F, and Muller D (2012) Structural plasticity upon learning: regulation and functions. Nat Rev Neurosci 13:478490.
Carpenter-Hyland EP, Woodward JJ, and Chandler LJ (2004) Chronic ethanol
induces synaptic but not extrasynaptic targeting of NMDA receptors. J Neurosci
24:78597868.
Carroll FI, Kotian P, Dehghani A, Gray JL, Kuzemko MA, Parham KA, Abraham P,
Lewin AH, Boja JW, and Kuhar MJ (1995) Cocaine and 3 beta-(49-substituted
phenyl)tropane-2 beta-carboxylic acid ester and amide analogues. New highaffinity and selective compounds for the dopamine transporter. J Med Chem 38:
379388.
Carroll ME, Batulis DK, Landry KL, and Morgan AD (2005) Sex differences in the
escalation of oral phencyclidine (PCP) self-administration under FR and PR
schedules in rhesus monkeys. Psychopharmacology (Berl) 180:414426.
Carroll ME, Carmona GN, and Rodefer JS (1994) Phencyclidine (PCP) selfadministration and withdrawal in rhesus monkeys: effects of buprenorphine and
dizocilpine (MK-801) pretreatment. Pharmacol Biochem Behav 48:723732.
Carter LP, Richards BD, Mintzer MZ, and Griffiths RR (2006) Relative abuse liability
of GHB in humans: A comparison of psychomotor, subjective, and cognitive effects
of supratherapeutic doses of triazolam, pentobarbital, and GHB. Neuropsychopharmacology 31:25372551.
Carter RB, Wood PL, Wieland S, Hawkinson JE, Belelli D, Lambert JJ, White HS,
Wolf HH, Mirsadeghi S, and Tahir SH, et al. (1997) Characterization of the
anticonvulsant properties of ganaxolone (CCD 1042; 3alpha-hydroxy-3betamethyl-5alpha-pregnan-20-one), a selective, high-affinity, steroid modulator of
the gamma-aminobutyric acid(A) receptor. J Pharmacol Exp Ther 280:12841295.
Carulli D, Pizzorusso T, Kwok JC, Putignano E, Poli A, Forostyak S, Andrews MR,
Deepa SS, Glant TT, and Fawcett JW (2010) Animals lacking link protein have
attenuated perineuronal nets and persistent plasticity. Brain 133:23312347.
Carvalho M, Carmo H, Costa VM, Capela JP, Pontes H, Remio F, Carvalho F,
and Bastos MdeL (2012) Toxicity of amphetamines: an update. Arch Toxicol 86:
11671231.
Casey BJ, Galvan A, and Hare TA (2005) Changes in cerebral functional organization
during cognitive development. Curr Opin Neurobiol 15:239244.
Cass WA (1996) GDNF selectively protects dopamine neurons over serotonin neurons
against the neurotoxic effects of methamphetamine. J Neurosci 16:81328139.
Cass WA and Manning MW (1999) Recovery of presynaptic dopaminergic functioning
in rats treated with neurotoxic doses of methamphetamine. J Neurosci 19:
76537660.
Castaeda E, Moss DE, Oddie SD, and Whishaw IQ (1991) THC does not affect
striatal dopamine release: microdialysis in freely moving rats. Pharmacol Biochem
Behav 40:587591.
Castillo PE (2012) Presynaptic LTP and LTD of excitatory and inhibitory synapses.
Cold Spring Harb Perspect Biol 4:2. DOI 10.1101/cshperspect.a005728
Castillo PE, Chiu CQ, and Carroll RC (2011) Long-term plasticity at inhibitory
synapses. Curr Opin Neurobiol 21:328338.
Castillo-Gmez E, Gmez-Climent MA, Varea E, Guirado R, Blasco-Ibez JM,
Crespo C, Martnez-Guijarro FJ, and Ncher J (2008) Dopamine acting through D2
receptors modulates the expression of PSA-NCAM, a molecule related to neuronal
structural plasticity, in the medial prefrontal cortex of adult rats. Exp Neurol 214:
97111.
Castrn E, da Penha Berzaghi M, Lindholm D, and Thoenen H (1993) Differential
effects of MK-801 on brain-derived neurotrophic factor mRNA levels in different
regions of the rat brain. Exp Neurol 122:244252.
Castrn E, Elgersma Y, Maffei L, and Hagerman R (2012) Treatment of neurodevelopmental disorders in adulthood. J Neurosci 32:1407414079.
Castro-Alamancos MA and Calcagnotto ME (1999) Presynaptic long-term potentiation in corticothalamic synapses. J Neurosci 19:90909097.
Catlow BJ, Song S, Paredes DA, Kirstein CL, and Sanchez-Ramos J (2013) Effects of
psilocybin on hippocampal neurogenesis and extinction of trace fear conditioning.
Exp Brain Res 228:481491.
Celio MR, Spreafico R, De Biasi S, and Vitellaro-Zuccarello L (1998) Perineuronal
nets: past and present. Trends Neurosci 21:510515.
Centonze D, Battista N, Rossi S, Mercuri NB, Finazzi-Agr A, Bernardi G, Calabresi
P, and Maccarrone M (2004) A critical interaction between dopamine D2 receptors
and endocannabinoids mediates the effects of cocaine on striatal gabaergic
Transmission. Neuropsychopharmacology 29:14881497.
Centonze D, Picconi B, Baunez C, Borrelli E, Pisani A, Bernardi G, and Calabresi P
(2002) Cocaine and amphetamine depress striatal GABAergic synaptic transmission through D2 dopamine receptors. Neuropsychopharmacology 26:164175.

Cepeda C, Andr VM, Yamazaki I, Wu N, Kleiman-Weiner M, and Levine MS (2008)


Differential electrophysiological properties of dopamine D1 and D2 receptorcontaining striatal medium-sized spiny neurons. Eur J Neurosci 27:671682.
Cervilla JA, Prince M, and Mann A (2000) Smoking, drinking, and incident cognitive
impairment: a cohort community based study included in the Gospel Oak project.
J Neurol Neurosurg Psychiatry 68:622626.
Cervinski MA, Foster JD, and Vaughan RA (2005) Psychoactive substrates stimulate
dopamine transporter phosphorylation and down-regulation by cocaine-sensitive
and protein kinase C-dependent mechanisms. J Biol Chem 280:4044240449.
Chabrol H (2013) MDMA assisted psychotherapy found to have a large effect for
chronic post-traumatic stress disorder. J Psychopharmacol 27:865866.
Chalifoux JR and Carter AG (2010) GABAB receptors modulate NMDA receptor
calcium signals in dendritic spines. Neuron 66:101113.
Chan P, Di Monte DA, Luo JJ, DeLanney LE, Irwin I, and Langston JW (1994) Rapid
ATP loss caused by methamphetamine in the mouse striatum: relationship between energy impairment and dopaminergic neurotoxicity. J Neurochem 62:
24842487.
Chandra D, Halonen LM, Linden AM, Procaccini C, Hellsten K, Homanics GE,
and Korpi ER (2010) Prototypic GABA(A) receptor agonist muscimol acts preferentially through forebrain high-affinity binding sites. Neuropsychopharmacology
35:9991007.
Chandra D, Jia F, Liang J, Peng Z, Suryanarayanan A, Werner DF, Spigelman I,
Houser CR, Olsen RW, and Harrison NL, et al. (2006) GABAA receptor alpha 4
subunits mediate extrasynaptic inhibition in thalamus and dentate gyrus and the
action of gaboxadol. Proc Natl Acad Sci USA 103:1523015235.
Chandramani Shivalingappa P, Jin H, Anantharam V, Kanthasamy A,
and Kanthasamy A (2012) N-Acetyl Cysteine Protects against MethamphetamineInduced Dopaminergic Neurodegeneration via Modulation of Redox Status and
Autophagy in Dopaminergic Cells. Parkinsons Dis 2012:424285.
Chang L, Cloak C, Patterson K, Grob C, Miller EN, and Ernst T (2005) Enlarged
striatum in abstinent methamphetamine abusers: a possible compensatory response. Biol Psychiatry 57:967974.
Chang L, Ernst T, Speck O, Patel H, DeSilva M, Leonido-Yee M, and Miller EN
(2002) Perfusion MRI and computerized cognitive test abnormalities in abstinent
methamphetamine users. Psychiatry Res 114:6579.
Changeux JP, Bertrand D, Corringer PJ, Dehaene S, Edelstein S, Lna C, Le Novre
N, Marubio L, Picciotto M, and Zoli M (1998) Brain nicotinic receptors: structure
and regulation, role in learning and reinforcement. Brain Res Brain Res Rev 26:
198216.
Changeux JP and Danchin A (1976) Selective stabilisation of developing synapses as
a mechanism for the specification of neuronal networks. Nature 264:705712.
Chatterjee M, Ganguly S, Srivastava M, and Palit G (2011) Effect of chronic versus
acute ketamine administration and its withdrawal effect on behavioural alterations in mice: implications for experimental psychosis. Behav Brain Res 216:
247254.
Chauvet C, Goldberg SR, Jaber M, and Solinas M (2012) Effects of environmental
enrichment on the incubation of cocaine craving. Neuropharmacology 63:635641.
Chavkin C, Sud S, Jin W, Stewart J, Zjawiony JK, Siebert DJ, Toth BA, Hufeisen SJ,
and Roth BL (2004) Salvinorin A, an active component of the hallucinogenic sage
salvia divinorum is a highly efficacious kappa-opioid receptor agonist: structural
and functional considerations. J Pharmacol Exp Ther 308:11971203.
Cheer JF, Wassum KM, Sombers LA, Heien ML, Ariansen JL, Aragona BJ, Phillips
PE, and Wightman RM (2007) Phasic dopamine release evoked by abused substances requires cannabinoid receptor activation. J Neurosci 27:791795.
Chen A, Muzzio IA, Malleret G, Bartsch D, Verbitsky M, Pavlidis P, Yonan AL,
Vronskaya S, Grody MB, and Cepeda I, et al. (2003) Inducible enhancement of
memory storage and synaptic plasticity in transgenic mice expressing an inhibitor
of ATF4 (CREB-2) and C/EBP proteins. Neuron 39:655669.
Chen BT, Bowers MS, Martin M, Hopf FW, Guillory AM, Carelli RM, Chou JK,
and Bonci A (2008) Cocaine but not natural reward self-administration nor passive
cocaine infusion produces persistent LTP in the VTA. Neuron 59:288297.
Chen BT, Yau HJ, Hatch C, Kusumoto-Yoshida I, Cho SL, Hopf FW, and Bonci A
(2013a) Rescuing cocaine-induced prefrontal cortex hypoactivity prevents compulsive cocaine seeking. Nature 496:359362.
Chen PL, Lee WJ, Sun WZ, Oyang YJ, and Fuh JL (2012) Risk of dementia in
patients with insomnia and long-term use of hypnotics: a population-based retrospective cohort study. PLoS One 7:e49113.
Chen Q, Zhu X, Zhang Y, Wetsel WC, Lee TH, and Zhang X (2010) Integrin-linked
kinase is involved in cocaine sensitization by regulating PSD-95 and synapsin I
expression and GluR1 Ser845 phosphorylation. J Mol Neurosci 40:284294.
Chen R, Zhang J, Fan N, Teng ZQ, Wu Y, Yang H, Tang YP, Sun H, Song Y, and Chen
C (2013b) D9-THC-caused synaptic and memory impairments are mediated
through COX-2 signaling. Cell 155:11541165.
Chen SX, Kim AN, Peters AJ, and Komiyama T (2015) Subtype-specific plasticity of
inhibitory circuits in motor cortex during motor learning. Nat Neurosci 18:11091115.
Chen TW, Wardill TJ, Sun Y, Pulver SR, Renninger SL, Baohan A, Schreiter ER,
Kerr RA, Orger MB, and Jayaraman V, et al. (2013c) Ultrasensitive fluorescent
proteins for imaging neuronal activity. Nature 499:295300.
Chen WY, Huang MC, and Lin SK (2014) Gender differences in subjective discontinuation symptoms associated with ketamine use. Subst Abuse Treat Prev Policy
9:39.
Chen X, Shu S, and Bayliss DA (2009) HCN1 channel subunits are a molecular
substrate for hypnotic actions of ketamine. J Neurosci 29:600609.
Chergui K, Charlty PJ, Akaoka H, Saunier CF, Brunet JL, Buda M, Svensson TH,
and Chouvet G (1993) Tonic activation of NMDA receptors causes spontaneous
burst discharge of rat midbrain dopamine neurons in vivo. Eur J Neurosci 5:
137144.
Chesworth R, Brown RM, Kim JH, and Lawrence AJ (2013) The metabotropic glutamate 5 receptor modulates extinction and reinstatement of methamphetamineseeking in mice. PLoS One 8:e68371.

Drug-Induced Neuroplasticity
Chevaleyre V and Castillo PE (2003) Heterosynaptic LTD of hippocampal GABAergic
synapses: a novel role of endocannabinoids in regulating excitability. Neuron 38:
461472.
Chevaleyre V, Takahashi KA, and Castillo PE (2006) Endocannabinoid-mediated
synaptic plasticity in the CNS. Annu Rev Neurosci 29:3776.
Chieng B and Williams JT (1998) Increased opioid inhibition of GABA release in
nucleus accumbens during morphine withdrawal. J Neurosci 18:70337039.
Childress AR, Mozley PD, McElgin W, Fitzgerald J, Reivich M, and OBrien CP
(1999) Limbic activation during cue-induced cocaine craving. Am J Psychiatry 156:
1118.
Chilukuri H, Reddy NP, Pathapati RM, Manu AN, Jollu S, and Shaik AB (2014)
Acute antidepressant effects of intramuscular versus intravenous ketamine. Indian J Psychol Med 36:7176.
Choi DS, Wei W, Deitchman JK, Kharazia VN, Lesscher HM, McMahon T, Wang D,
Qi ZH, Sieghart W, and Zhang C, et al. (2008) Protein kinase Cdelta regulates
ethanol intoxication and enhancement of GABA-stimulated tonic current. J Neurosci 28:1189011899.
Christensen R, Kristensen PK, Bartels EM, Bliddal H, and Astrup A (2007) Efficacy
and safety of the weight-loss drug rimonabant: a meta-analysis of randomised
trials. Lancet 370:17061713.
Christian CA, Herbert AG, Holt RL, Peng K, Sherwood KD, Pangratz-Fuehrer S,
Rudolph U, and Huguenard JR (2013) Endogenous positive allosteric modulation of
GABA(A) receptors by diazepam binding inhibitor. Neuron 78:10631074.
Christie BR, Schexnayder LK, and Johnston D (1997) Contribution of voltage-gated
Ca2+ channels to homosynaptic long-term depression in the CA1 region in vitro.
J Neurophysiol 77:16511655.
Christie MJ (2008) Cellular neuroadaptations to chronic opioids: tolerance, withdrawal and addiction. Br J Pharmacol 154:384396.
Chu B, Anantharam V, and Treistman SN (1995) Ethanol inhibition of recombinant
heteromeric NMDA channels in the presence and absence of modulators. J Neurochem 65:140148.
Chung HJ, Xia J, Scannevin RH, Zhang X, and Huganir RL (2000) Phosphorylation of
the AMPA receptor subunit GluR2 differentially regulates its interaction with PDZ
domain-containing proteins. J Neurosci 20:72587267.
Chung K and Deisseroth K (2013) CLARITY for mapping the nervous system. Nat
Methods 10:508513.
Chung S, Pohl S, Zeng J, Civelli O, and Reinscheid RK (2006) Endogenous orphanin
FQ/nociceptin is involved in the development of morphine tolerance. J Pharmacol
Exp Ther 318:262267.
Chuong AS, Miri ML, Busskamp V, Matthews GA, Acker LC, Srensen AT, Young A,
Klapoetke NC, Henninger MA, and Kodandaramaiah SB, et al. (2014) Noninvasive
optical inhibition with a red-shifted microbial rhodopsin. Nat Neurosci 17:11231129.
Ciani E, Severi S, Bartesaghi R, and Contestabile A (2005) Neurochemical correlates
of nicotine neurotoxicity on rat habenulo-interpeduncular cholinergic neurons.
Neurotoxicology 26:467474.
Ciccocioppo R, Stopponi S, Economidou D, Kuriyama M, Kinoshita H, Heilig M,
Roberto M, Weiss F, and Teshima K (2014) Chronic treatment with novel brainpenetrating selective NOP receptor agonist MT-7716 reduces alcohol drinking and
seeking in the rat. Neuropsychopharmacology 39:26012610.
Cittadini A and Lader M (1991) Lack of effect of a small dose of flumazenil in reversing short-term tolerance to benzodiazepines in normal subjects. J Psychopharmacol 5:220227.
Clarke DE, Miller HH, and Shore PA (1979) Monoamine oxidase inhibition by
(+)-amphetamine in vivo [proceedings]. Br J Pharmacol 66:435P436P.
Clarke PB and Kumar R (1983) Characterization of the locomotor stimulant action of
nicotine in tolerant rats. Br J Pharmacol 80:587594.
Clarke RB and Adermark L (2010) Acute ethanol treatment prevents
endocannabinoid-mediated long-lasting disinhibition of striatal output. Neuropharmacology 58:799805.
Clarke RB, Adermark L, Chau P, Sderpalm B, and Ericson M (2014) Increase in
nucleus accumbens dopamine levels following local ethanol administration is not
mediated by acetaldehyde. Alcohol Alcohol 49:498504.
Clayton LM, Stern WM, Newman WD, Sander JW, Acheson J, and Sisodiya SM
(2013) Evolution of visual field loss over ten years in individuals taking vigabatrin.
Epilepsy Res 105:262271.
Clemens KJ, Cornish JL, Hunt GE, and McGregor IS (2007) Repeated weekly exposure to MDMA, methamphetamine or their combination: long-term behavioural
and neurochemical effects in rats. Drug Alcohol Depend 86:183190.
Clemens KJ, Van Nieuwenhuyzen PS, Li KM, Cornish JL, Hunt GE, and McGregor
IS (2004) MDMA (ecstasy), methamphetamine and their combination: long-term
changes in social interaction and neurochemistry in the rat. Psychopharmacology
(Berl) 173:318325.
Cochran SM, Kennedy M, McKerchar CE, Steward LJ, Pratt JA, and Morris BJ
(2003) Induction of metabolic hypofunction and neurochemical deficits after
chronic intermittent exposure to phencyclidine: differential modulation by antipsychotic drugs. Neuropsychopharmacology 28:265275.
Codd EE, Shank RP, Schupsky JJ, and Raffa RB (1995) Serotonin and norepinephrine uptake inhibiting activity of centrally acting analgesics: structural determinants and role in antinociception. J Pharmacol Exp Ther 274:12631270.
Cohen A, Soleiman MT, Talia R, Koob GF, George O, and Mandyam CD (2015)
Extended access nicotine self-administration with periodic deprivation increases
immature neurons in the hippocampus. Psychopharmacology (Berl) 232:453463.
Cohen LM, Collins FL, Young A, McChargue DE, Leffingwell TR, and Cook KL (2013)
Pharmacology and Treatment of Substance Abuse: Evidence and Outcome Based
Perspectives, New York, Taylor & Francis.
Colado MI and Green AR (1995) The spin trap reagent alpha-phenyl-N-tert-butyl
nitrone prevents ecstasy-induced neurodegeneration of 5-hydroxytryptamine
neurones. Eur J Pharmacol 280:343346.

975

Colby SM, Tiffany ST, Shiffman S, and Niaura RS (2000) Are adolescent smokers
dependent on nicotine? A review of the evidence. Drug Alcohol Depend 59 (Suppl 1):
S83S95.
Colgin LL, Moser EI, and Moser MB (2008) Understanding memory through hippocampal remapping. Trends Neurosci 31:469477.
Collingridge GL, Olsen RW, Peters J, and Spedding M (2009) A nomenclature for
ligand-gated ion channels. Neuropharmacology 56:25.
Comim CM, Gomes KM, Rus GZ, Petronilho F, Ferreira GK, Streck EL, Dal-Pizzol
F, and Quevedo J (2014) Methylphenidate treatment causes oxidative stress and
alters energetic metabolism in an animal model of attention-deficit hyperactivity
disorder. Acta Neuropsychiatr 26:96103.
Concha ML and Wilson SW (2001) Asymmetry in the epithalamus of vertebrates.
J Anat 199:6384.
Connelly WM, Errington AC, and Crunelli V (2013) g-Hydroxybutyric acid (GHB) is
not an agonist of extrasynaptic GABAA receptors. PLoS One 8:e79062.
Connor M and Christie MD (1999) Opioid receptor signalling mechanisms. Clin Exp
Pharmacol Physiol 26:493499.
Conrad KL, Tseng KY, Uejima JL, Reimers JM, Heng LJ, Shaham Y, Marinelli M,
and Wolf ME (2008) Formation of accumbens GluR2-lacking AMPA receptors
mediates incubation of cocaine craving. Nature 454:118121.
Cooper DC (2002) The significance of action potential bursting in the brain reward
circuit. Neurochem Int 41:333340.
Copeland BJ, Vogelsberg V, Neff NH, and Hadjiconstantinou M (1996) Protein kinase
C activators decrease dopamine uptake into striatal synaptosomes. J Pharmacol
Exp Ther 277:15271532.
Corbit LH, Chieng BC, and Balleine BW (2014) Effects of repeated cocaine exposure
on habit learning and reversal by N-acetylcysteine. Neuropsychopharmacology 39:
18931901.
Corbit LH and Janak PH (2010) Posterior dorsomedial striatum is critical for both
selective instrumental and Pavlovian reward learning. Eur J Neurosci 31:
13121321.
Corbit LH, Nie H, and Janak PH (2012) Habitual alcohol seeking: time course and
the contribution of subregions of the dorsal striatum. Biol Psychiatry 72:389395.
Cornwell BR, Salvadore G, Furey M, Marquardt CA, Brutsche NE, Grillon C,
and Zarate CA Jr (2012) Synaptic potentiation is critical for rapid antidepressant
response to ketamine in treatment-resistant major depression. Biol Psychiatry 72:
555561.
Costa E, Groppetti A, and Naimzada MK (1972) Effects of amphetamine on the
turnover rate of brain catecholamines and motor activity. Br J Pharmacol 44:
742751.
Costa E and Guidotti A (1991) Diazepam binding inhibitor (DBI): a peptide with
multiple biological actions. Life Sci 49:325344.
Coull JA, Beggs S, Boudreau D, Boivin D, Tsuda M, Inoue K, Gravel C, Salter MW,
and De Koninck Y (2005) BDNF from microglia causes the shift in neuronal anion
gradient underlying neuropathic pain. Nature 438:10171021.
Counotte DS, Goriounova NA, Li KW, Loos M, van der Schors RC, Schetters D,
Schoffelmeer AN, Smit AB, Mansvelder HD, and Pattij T, et al. (2011) Lasting
synaptic changes underlie attention deficits caused by nicotine exposure during
adolescence. Nat Neurosci 14:417419.
Counotte DS, Goriounova NA, Moretti M, Smoluch MT, Irth H, Clementi F,
Schoffelmeer AN, Mansvelder HD, Smit AB, and Gotti C, et al. (2012) Adolescent
nicotine exposure transiently increases high-affinity nicotinic receptors and modulates inhibitory synaptic transmission in rat medial prefrontal cortex. FASEB J
26:18101820.
Counotte DS, Spijker S, Van de Burgwal LH, Hogenboom F, Schoffelmeer AN, De
Vries TJ, Smit AB, and Pattij T (2009) Long-lasting cognitive deficits resulting
from adolescent nicotine exposure in rats. Neuropsychopharmacology 34:299306.
Covington HE 3rd, Maze I, Sun H, Bomze HM, DeMaio KD, Wu EY, Dietz DM, Lobo
MK, Ghose S, and Mouzon E, et al. (2011) A role for repressive histone methylation
in cocaine-induced vulnerability to stress. Neuron 71:656670.
Cowan RL, Lyoo IK, Sung SM, Ahn KH, Kim MJ, Hwang J, Haga E, Vimal RL, Lukas
SE, and Renshaw PF (2003) Reduced cortical gray matter density in human
MDMA (Ecstasy) users: a voxel-based morphometry study. Drug Alcohol Depend
72:225235.
Cozzi NV, Sievert MK, Shulgin AT, Jacob P 3rd, and Ruoho AE (1999) Inhibition of
plasma membrane monoamine transporters by beta-ketoamphetamines. Eur J
Pharmacol 381:6369.
Cozzoli DK, Goulding SP, Zhang PW, Xiao B, Hu JH, Ary AW, Obara I, Rahn A,
Abou-Ziab H, and Tyrrel B, et al. (2009) Binge drinking upregulates accumbens
mGluR5-Homer2-PI3K signaling: functional implications for alcoholism. J Neurosci 29:86558668.
Crawford JT, Roberts DC, and Beveridge TJ (2013) The group II metabotropic
glutamate receptor agonist, LY379268, decreases methamphetamine selfadministration in rats. Drug Alcohol Depend 132:414419.
Creeley C, Dikranian K, Dissen G, Martin L, Olney J, and Brambrink A (2013)
Propofol-induced apoptosis of neurones and oligodendrocytes in fetal and neonatal
rhesus macaque brain. Br J Anaesth 110 (Suppl 1):i29i38.
Creeley CE and Olney JW (2010) The young: neuroapoptosis induced by anesthetics
and what to do about it. Anesth Analg 110:442448.
Creeley CE and Olney JW (2013) Drug-induced apoptosis: mechanism by which alcohol and many other drugs can disrupt brain development. Brain Sci 3:
11531181.
Crestani F, Keist R, Fritschy JM, Benke D, Vogt K, Prut L, Blthmann H, Mhler H,
and Rudolph U (2002) Trace fear conditioning involves hippocampal alpha5 GABA
(A) receptors. Proc Natl Acad Sci USA 99:89808985.
Crestani F, Lw K, Keist R, Mandelli M, Mhler H, and Rudolph U (2001) Molecular
targets for the myorelaxant action of diazepam. Mol Pharmacol 59:442445.
Crestani F, Martin JR, Mhler H, and Rudolph U (2000) Mechanism of action of the
hypnotic zolpidem in vivo. Br J Pharmacol 131:12511254.

976

Korpi et al.

Crider R (1986) Phencyclidine: changing abuse patterns. NIDA Res Monogr 64:
163173.
Cruz HG, Ivanova T, Lunn ML, Stoffel M, Slesinger PA, and Lscher C (2004) Bidirectional effects of GABA(B) receptor agonists on the mesolimbic dopamine
system. Nat Neurosci 7:153159.
Cruz MT, Herman MA, Kallupi M, and Roberto M (2012) Nociceptin/orphanin FQ
blockade of corticotropin-releasing factor-induced gamma-aminobutyric acid release in central amygdala is enhanced after chronic ethanol exposure. Biol Psychiatry 71:666676.
Cryan JF and Slattery DA (2010) GABAB receptors and depression. Current status.
Adv Pharmacol 58:427451.
Cui C, Noronha A, Morikawa H, Alvarez VA, Stuber GD, Szumlinski KK, Kash TL,
Roberto M, and Wilcox MV (2013) New insights on neurobiological mechanisms
underlying alcohol addiction. Neuropharmacology 67:223232.
Cui X, Lefevre E, Turner KM, Coelho CM, Alexander S, Burne TH, and Eyles DW
(2014) MK-801-induced behavioural sensitisation alters dopamine release and
turnover in rat prefrontal cortex. Psychopharmacology (Berl) 232:509517.
Cunha-Oliveira T, Rego AC, Cardoso SM, Borges F, Swerdlow RH, Macedo T, and de
Oliveira CR (2006) Mitochondrial dysfunction and caspase activation in rat cortical
neurons treated with cocaine or amphetamine. Brain Res 1089:4454.
Cunningham CW, Rothman RB, and Prisinzano TE (2011) Neuropharmacology of the
naturally occurring kappa-opioid hallucinogen salvinorin A. Pharmacol Rev 63:
316347.
Cunningham MG, Bhattacharyya S, and Benes FM (2002) Amygdalo-cortical
sprouting continues into early adulthood: implications for the development of
normal and abnormal function during adolescence. J Comp Neurol 453:116130.
Currington R, Justinova Z, Ladenheim B, Goldberg SR, Krasnova IN, and Cadet JL
(2011) Methamphetamine self-administration is associated with changes in the
expression of glutamate AMPA receptors in the rat striatum. FASEB J 25:857.853.
Cuzon Carlson VC, Seabold GK, Helms CM, Garg N, Odagiri M, Rau AR, Daunais J,
Alvarez VA, Lovinger DM, and Grant KA (2011) Synaptic and morphological
neuroadaptations in the putamen associated with long-term, relapsing alcohol
drinking in primates. Neuropsychopharmacology 36:25132528.
Daberkow DP, Brown HD, Bunner KD, Kraniotis SA, Doellman MA, Ragozzino ME,
Garris PA, and Roitman MF (2013) Amphetamine paradoxically augments exocytotic dopamine release and phasic dopamine signals. J Neurosci 33:452463.
Dacher M, Gouty S, Dash S, Cox BM, and Nugent FS (2013) A-kinase anchoring
protein-calcineurin signaling in long-term depression of GABAergic synapses.
J Neurosci 33:26502660.
Dacher M and Nugent FS (2011a) Morphine-induced modulation of LTD at
GABAergic synapses in the ventral tegmental area. Neuropharmacology 61:
11661171.
Dacher M and Nugent FS (2011b) Opiates and plasticity. Neuropharmacology 61:
10881096.
Da Cunha C, Wolfman C, Levi de Stein M, Ruschel AC, Izquierdo I, and Medina JH
(1992) Anxiogenic effects of the intraamygdala injection of flumazenil, a benzodiazepine receptor antagonist. Funct Neurol 7:401405.
Dadda M, Domenichini A, Piffer L, Argenton F, and Bisazza A (2010) Early differences in epithalamic left-right asymmetry influence lateralization and personality
of adult zebrafish. Behav Brain Res 206:208215.
Dahl D and Sarvey JM (1989) Norepinephrine induces pathway-specific long-lasting
potentiation and depression in the hippocampal dentate gyrus. Proc Natl Acad Sci
USA 86:47764780.
Daigle TL, Wetsel WC, and Caron MG (2011) Opposite function of dopamine D1 and
N-methyl-D-aspartate receptors in striatal cannabinoid-mediated signaling. Eur
J Neurosci 34:13781389.
Damez-Werno D, LaPlant Q, Sun H, Scobie KN, Dietz DM, Walker IM, Koo JW,
Vialou VF, Mouzon E, and Russo SJ, et al. (2012) Drug experience epigenetically
primes Fosb gene inducibility in rat nucleus accumbens. J Neurosci 32:
1026710272.
Dao DQ, Perez EE, Teng Y, Dani JA, and De Biasi M (2014) Nicotine enhances
excitability of medial habenular neurons via facilitation of neurokinin signaling. J
Neurosci 34:42734284.
Darvesh AS, Yamamoto BK, and Gudelsky GA (2005) Evidence for the involvement of
nitric oxide in 3,4-methylenedioxymethamphetamine-induced serotonin depletion
in the rat brain. J Pharmacol Exp Ther 312:694701.
Das P, Lilly SM, Zerda R, Gunning WT 3rd, Alvarez FJ, and Tietz EI (2008) Increased AMPA receptor GluR1 subunit incorporation in rat hippocampal CA1
synapses during benzodiazepine withdrawal. J Comp Neurol 511:832846.
Daumann J, Fimm B, Willmes K, Thron A, and Gouzoulis-Mayfrank E (2003) Cerebral activation in abstinent ecstasy (MDMA) users during a working memory
task: a functional magnetic resonance imaging (fMRI) study. Brain Res Cogn Brain
Res 16:479487.
Daumann J, Fischermann T, Heekeren K, Henke K, Thron A, and GouzoulisMayfrank E (2005) Memory-related hippocampal dysfunction in poly-drug ecstasy
(3,4-methylenedioxymethamphetamine) users. Psychopharmacology (Berl) 180:
607611.
Daumann J, Koester P, Becker B, Wagner D, Imperati D, Gouzoulis-Mayfrank E,
and Tittgemeyer M (2011) Medial prefrontal gray matter volume reductions in
users of amphetamine-type stimulants revealed by combined tract-based spatial
statistics and voxel-based morphometry. Neuroimage 54:794801.
Davidson TJ, Kloosterman F, and Wilson MA (2009) Hippocampal replay of extended
experience. Neuron 63:497507.
Davies DL, Kochegarov AA, Kuo ST, Kulkarni AA, Woodward JJ, King BF,
and Alkana RL (2005) Ethanol differentially affects ATP-gated P2X(3) and P2X(4)
receptor subtypes expressed in Xenopus oocytes. Neuropharmacology 49:243253.
Davies DL, Machu TK, Guo Y, and Alkana RL (2002) Ethanol sensitivity in ATPgated P2X receptors is subunit dependent. Alcohol Clin Exp Res 26:773778.

Dvila-Garca MI, Musachio JL, and Kellar KJ (2003) Chronic nicotine administration does not increase nicotinic receptors labeled by [125I]epibatidine in adrenal
gland, superior cervical ganglia, pineal or retina. J Neurochem 85:12371246.
Davis SJ, Scott LL, Hu K, and Pierce-Shimomura JT (2014) Conserved single residue
in the BK potassium channel required for activation by alcohol and intoxication in
C. elegans. J Neurosci 34:95629573.
Daza-Losada M, Rodrguez-Arias M, Maldonado C, Aguilar MA, and Miarro J (2008)
Behavioural and neurotoxic long-lasting effects of MDMA plus cocaine in adolescent mice. Eur J Pharmacol 590:204211.
Deehan GA Jr, Hauser SR, Wilden JA, Truitt WA, and Rodd ZA (2013) Elucidating
the biological basis for the reinforcing actions of alcohol in the mesolimbic dopamine system: the role of active metabolites of alcohol. Front Behav Neurosci 7:104.
de Jong CA, Kamal R, Dijkstra BA, and de Haan HA (2012) Gamma-hydroxybutyrate
detoxification by titration and tapering. Eur Addict Res 18:4045.
Deisseroth K and Schnitzer MJ (2013) Engineering approaches to illuminating brain
structure and dynamics. Neuron 80:568577.
Deitrich RA, Dunwiddie TV, Harris RA, and Erwin VG (1989) Mechanism of action of
ethanol: initial central nervous system actions. Pharmacol Rev 41:489537.
De la Herrn-Arita AK and Garca-Garca F (2013) Current and emerging options for
the drug treatment of narcolepsy. Drugs 73:17711781.
Dela Pena I, de la Pena JB, Kim BN, Han DH, Noh M, and Cheong JH (2015) Gene
expression profiling in the striatum of amphetamine-treated spontaneously hypertensive rats which showed amphetamine conditioned place preference and selfadministration. Arch Pharm Res 38:865875.
de la Torre R and Farr M (2004) Neurotoxicity of MDMA (ecstasy): the limitations of
scaling from animals to humans. Trends Pharmacol Sci 25:505508.
de Lecea L, Kilduff TS, Peyron C, Gao X, Foye PE, Danielson PE, Fukuhara C,
Battenberg EL, Gautvik VT, and Bartlett FS 2nd, et al. (1998) The hypocretins:
hypothalamus-specific peptides with neuroexcitatory activity. Proc Natl Acad Sci
USA 95:322327.
De Luca MT and Badiani A (2011) Ketamine self-administration in the rat: evidence
for a critical role of setting. Psychopharmacology (Berl) 214:549556.
Demirakca T, Sartorius A, Ende G, Meyer N, Welzel H, Skopp G, Mann K,
and Hermann D (2011) Diminished gray matter in the hippocampus of cannabis
users: possible protective effects of cannabidiol. Drug Alcohol Depend 114:242245.
den Hollander B, Rozov S, Linden AM, Uusi-Oukari M, Ojanper I, and Korpi ER
(2013) Long-term cognitive and neurochemical effects of bath salt designer drugs
methylone and mephedrone. Pharmacol Biochem Behav 103:501509.
den Hollander B, Schouw M, Groot P, Huisman H, Caan M, Barkhof F, and Reneman
L (2012) Preliminary evidence of hippocampal damage in chronic users of ecstasy.
J Neurol Neurosurg Psychiatry 83:8385.
De Souza EB, Battaglia G, and Insel TR (1990) Neurotoxic effect of MDMA on brain
serotonin neurons: evidence from neurochemical and radioligand binding studies.
Ann N Y Acad Sci 600:682697, discussion 697698.
De Vito MJ and Wagner GC (1989) Methamphetamine-induced neuronal damage:
a possible role for free radicals. Neuropharmacology 28:11451150.
De Vries TJ, Schoffelmeer AN, Binnekade R, Mulder AH, and Vanderschuren LJ
(1998) MK-801 reinstates drug-seeking behaviour in cocaine-trained rats. Neuroreport 9:637640.
de Win MM, Jager G, Booij J, Reneman L, Schilt T, Lavini C, Olabarriaga SD, den
Heeten GJ, and van den Brink W (2008) Sustained effects of ecstasy on the human
brain: a prospective neuroimaging study in novel users. Brain 131:29362945.
Deng XS and Deitrich RA (2008) Putative role of brain acetaldehyde in ethanol
addiction. Curr Drug Abuse Rev 1:38.
Dennis EL and Thompson PM (2013) Typical and atypical brain development: a review of neuroimaging studies. Dialogues Clin Neurosci 15:359384.
Deroche-Gamonet V, Belin D, and Piazza PV (2004) Evidence for addiction-like behavior in the rat. Science 305:10141017.
Desai RI and Bergman J (2010) Drug discrimination in methamphetamine-trained
rats: effects of cholinergic nicotinic compounds. J Pharmacol Exp Ther 335:
807816.
Deschaux O, Vendruscolo LF, Schlosburg JE, Diaz-Aguilar L, Yuan CJ, Sobieraj JC,
George O, Koob GF, and Mandyam CD (2014) Hippocampal neurogenesis protects
against cocaine-primed relapse. Addict Biol 19:562574.
Desprs JP, Golay A, and Sjstrm L; Rimonabant in Obesity-Lipids Study Group
(2005) Effects of rimonabant on metabolic risk factors in overweight patients with
dyslipidemia. N Engl J Med 353:21212134.
Deutsch SI, Mastropaolo J, and Rosse RB (1998) Neurodevelopmental consequences
of early exposure to phencyclidine and related drugs. Clin Neuropharmacol 21:
320332.
Devane WA, Hanus L, Breuer A, Pertwee RG, Stevenson LA, Griffin G, Gibson D,
Mandelbaum A, Etinger A, and Mechoulam R (1992) Isolation and structure of
a brain constituent that binds to the cannabinoid receptor. Science 258:19461949.
Devaud LL, Fritschy JM, Sieghart W, and Morrow AL (1997) Bidirectional alterations of GABA(A) receptor subunit peptide levels in rat cortex during chronic
ethanol consumption and withdrawal. J Neurochem 69:126130.
Devaud LL, Smith FD, Grayson DR, and Morrow AL (1995) Chronic ethanol consumption differentially alters the expression of gamma-aminobutyric acidA receptor subunit mRNAs in rat cerebral cortex: competitive, quantitative reverse
transcriptase-polymerase chain reaction analysis. Mol Pharmacol 48:861868.
Dewey SL, Chaurasia CS, Chen CE, Volkow ND, Clarkson FA, Porter SP, StraughterMoore RM, Alexoff DL, Tedeschi D, and Russo NB, et al. (1997) GABAergic
attenuation of cocaine-induced dopamine release and locomotor activity. Synapse
25:393398.
Dhaher R, Hauser SR, Getachew B, Bell RL, McBride WJ, McKinzie DL, and Rodd ZA
(2010) The orexin-1 receptor antagonist SB-334867 reduces alcohol relapse drinking,
but not alcohol-seeking, in alcohol-preferring (P) rats. J Addict Med 4:153159.
Dhopeshwarkar A and Mackie K (2014) CB2 Cannabinoid receptors as a therapeutic
target-what does the future hold? Mol Pharmacol 86:430437.

Drug-Induced Neuroplasticity
Diamond PR, Farmery AD, Atkinson S, Haldar J, Williams N, Cowen PJ, Geddes JR,
and McShane R (2014) Ketamine infusions for treatment resistant depression:
a series of 28 patients treated weekly or twice weekly in an ECT clinic. J Psychopharmacol 28:536544.
Diana M, Peana AT, Sirca D, Lintas A, Melis M, and Enrico P (2008) Crucial role of
acetaldehyde in alcohol activation of the mesolimbic dopamine system. Ann N Y
Acad Sci 1139:307317.
Daz JL and Huttunen MO (1971) Persistent increase in brain serotonin turnover
after chronic administration of LSD in the rat. Science 174:6264.
Diazgranados N, Ibrahim L, Brutsche NE, Newberg A, Kronstein P, Khalife S,
Kammerer WA, Quezado Z, Luckenbaugh DA, and Salvadore G, et al. (2010) A
randomized add-on trial of an N-methyl-D-aspartate antagonist in treatmentresistant bipolar depression. Arch Gen Psychiatry 67:793802.
Dietrich JB, Mangeol A, Revel MO, Burgun C, Aunis D, and Zwiller J (2005) Acute or
repeated cocaine administration generates reactive oxygen species and induces
antioxidant enzyme activity in dopaminergic rat brain structures. Neuropharmacology 48:965974.
Dietrich JB, Poirier R, Aunis D, and Zwiller J (2004) Cocaine downregulates the
expression of the mitochondrial genome in rat brain. Ann N Y Acad Sci 1025:
345350.
Dietz DM, Sun H, Lobo MK, Cahill ME, Chadwick B, Gao V, Koo JW, Mazei-Robison
MS, Dias C, and Maze I, et al. (2012) Rac1 is essential in cocaine-induced structural plasticity of nucleus accumbens neurons. Nat Neurosci 15:891896.
Di Forti M, Marconi A, Carra E, Fraietta S, Trotta A, Bonomo M, Bianconi F,
Gardner-Sood P, OConnor J, and Russo M, et al. (2015) Proportion of patients in
south London with first-episode psychosis attributable to use of high potency
cannabis: a case-control study. Lancet Psychiatry 2:233238.
Dildy JE and Leslie SW (1989) Ethanol inhibits NMDA-induced increases in free
intracellular Ca2+ in dissociated brain cells. Brain Res 499:383387.
DiLuca M and Olesen J (2014) The cost of brain diseases: a burden or a challenge?
Neuron 82:12051208.
Ding X, Cai J, Li S, Liu XD, Wan Y, and Xing GG (2014) BDNF contributes to the
development of neuropathic pain by induction of spinal long-term potentiation via
SHP2 associated GluN2B-containing NMDA receptors activation in rats with
spinal nerve ligation. Neurobiol Dis 73C:428451.
Dingledine R, Borges K, Bowie D, and Traynelis SF (1999) The glutamate receptor
ion channels. Pharmacol Rev 51:761.
Dinh TP, Carpenter D, Leslie FM, Freund TF, Katona I, Sensi SL, Kathuria S,
and Piomelli D (2002) Brain monoglyceride lipase participating in endocannabinoid inactivation. Proc Natl Acad Sci USA 99:1081910824.
Ditre JW, Brandon TH, Zale EL, and Meagher MM (2011) Pain, nicotine, and
smoking: research findings and mechanistic considerations. Psychol Bull 137:
10651093.
Domino EF and Luby ED (2012) Phencyclidine/schizophrenia: one view toward the
past, the other to the future. Schizophr Bull 38:914919.
Dopico AM, Bukiya AN, and Martin GE (2014) Ethanol modulation of mammalian
BK channels in excitable tissues: molecular targets and their possible contribution
to alcohol-induced altered behavior. Front Physiol 5:466.
Dosenbach NU, Nardos B, Cohen AL, Fair DA, Power JD, Church JA, Nelson SM,
Wig GS, Vogel AC, and Lessov-Schlaggar CN, et al. (2010) Prediction of individual
brain maturity using fMRI. Science 329:13581361.
Doura MB, Gold AB, Keller AB, and Perry DC (2008) Adult and periadolescent rats
differ in expression of nicotinic cholinergic receptor subtypes and in the response of
these subtypes to chronic nicotine exposure. Brain Res 1215:4052.
Doyle JR and Yamamoto BK (2010) Serotonin 2 receptor modulation of hyperthermia, corticosterone, and hippocampal serotonin depletions following serial exposure to chronic stress and methamphetamine. Psychoneuroendocrinology 35:
629633.
Doyon WM, Dong Y, Ostroumov A, Thomas AM, Zhang TA, and Dani JA (2013)
Nicotine decreases ethanol-induced dopamine signaling and increases selfadministration via stress hormones. Neuron 79:530540.
Drevets WC, Zarate CA Jr, and Furey ML (2013) Antidepressant effects of the
muscarinic cholinergic receptor antagonist scopolamine: a review. Biol Psychiatry
73:11561163.
Dribben WH, Creeley CE, Wang HH, Smith DJ, Farber NB, and Olney JW (2009)
High dose magnesium sulfate exposure induces apoptotic cell death in the developing neonatal mouse brain. Neonatology 96:2332.
Druhan JP, Rajabi H, and Stewart J (1996) MK-801 increases locomotor activity
without elevating extracellular dopamine levels in the nucleus accumbens. Synapse 24:135146.
DSouza DC, Abi-Saab WM, Madonick S, Forselius-Bielen K, Doersch A, Braley G,
Gueorguieva R, Cooper TB, and Krystal JH (2005) Delta-9-tetrahydrocannabinol
effects in schizophrenia: implications for cognition, psychosis, and addiction. Biol
Psychiatry 57:594608.
Duan TT, Tan JW, Yuan Q, Cao J, Zhou QX, and Xu L (2013) Acute ketamine induces
hippocampal synaptic depression and spatial memory impairment through dopamine D1/D5 receptors. Psychopharmacology (Berl) 228:451461.
du Bois TM, Huang XF, and Deng C (2008) Perinatal administration of PCP alters
adult behaviour in female Sprague-Dawley rats. Behav Brain Res 188:416419.
du Bois TM, Newell KA, and Huang XF (2012) Perinatal phencyclidine treatment
alters neuregulin 1/erbB4 expression and activation in later life. Eur Neuropsychopharmacol 22:356363.
Duman RS, Tallman JF, and Nestler EJ (1988) Acute and chronic opiate-regulation
of adenylate cyclase in brain: specific effects in locus coeruleus. J Pharmacol Exp
Ther 246:10331039.
Duman RS and Voleti B (2012) Signaling pathways underlying the pathophysiology
and treatment of depression: novel mechanisms for rapid-acting agents. Trends
Neurosci 35:4756.
Duncan JW (1974) Persisting psychotic states in adolescent drug users. Child Psychiatry Hum Dev 5:5162.

977

Durieux ME (1995) Inhibition by ketamine of muscarinic acetylcholine receptor


function. Anesth Analg 81:5762.
Duveau V, Laustela S, Barth L, Gianolini F, Vogt KE, Keist R, Chandra D, Homanics
GE, Rudolph U, and Fritschy JM (2011) Spatiotemporal specificity of GABAA
receptor-mediated regulation of adult hippocampal neurogenesis. Eur J Neurosci
34:362373.
Dzietko M, Felderhoff-Mueser U, Sifringer M, Krutz B, Bittigau P, Thor F, Heumann
R, Bhrer C, Ikonomidou C, and Hansen HH (2004) Erythropoietin protects the
developing brain against N-methyl-D-aspartate receptor antagonist neurotoxicity.
Neurobiol Dis 15:177187.
Earl DE, Das P, Gunning WT 3rd, and Tietz EI (2012) Regulation of Ca/calmodulindependent protein kinase II signaling within hippocampal glutamatergic postsynapses during flurazepam withdrawal. Neural Plast 2012:405926.
Egecioglu E, Steensland P, Fredriksson I, Feltmann K, Engel JA, and Jerlhag E
(2013) The glucagon-like peptide 1 analogue Exendin-4 attenuates alcohol mediated behaviors in rodents. Psychoneuroendocrinology 38:12591270.
Ehrich JM, Phillips PE, and Chavkin C (2014) Kappa opioid receptor activation
potentiates the cocaine-induced increase in evoked dopamine release recorded in
vivo in the mouse nucleus accumbens. Neuropsychopharmacology 39:30363048.
Eilers H, Schaeffer E, Bickler PE, and Forsayeth JR (1997) Functional deactivation of
the major neuronal nicotinic receptor caused by nicotine and a protein kinase
C-dependent mechanism. Mol Pharmacol 52:11051112.
Eisch AJ, Barrot M, Schad CA, Self DW, and Nestler EJ (2000) Opiates inhibit
neurogenesis in the adult rat hippocampus. Proc Natl Acad Sci USA 97:75797584.
El Ayadi A and Zigmond MJ (2011) Low concentrations of methamphetamine can
protect dopaminergic cells against a larger oxidative stress injury: mechanistic
study. PLoS One 6:e24722.
Elgoyhen AB, Vetter DE, Katz E, Rothlin CV, Heinemann SF, and Boulter J (2001)
alpha10: a determinant of nicotinic cholinergic receptor function in mammalian
vestibular and cochlear mechanosensory hair cells. Proc Natl Acad Sci USA 98:
35013506.
Ellickson PL, Tucker JS, and Klein DJ (2001) High-risk behaviors associated with
early smoking: results from a 5-year follow-up. J Adoles Health 28:465473.
Ellison G (1992) Continuous amphetamine and cocaine have similar neurotoxic effects in
lateral habenular nucleus and fasciculus retroflexus. Brain Res 598:353356.
Ellison G (2002) Neural degeneration following chronic stimulant abuse reveals
a weak link in brain, fasciculus retroflexus, implying the loss of forebrain control
circuitry. Eur Neuropsychopharmacol 12:287297.
Elsworth JD, Hajszan T, Leranth C, and Roth RH (2011a) Loss of asymmetric spine
synapses in dorsolateral prefrontal cortex of cognitively impaired phencyclidinetreated monkeys. Int J Neuropsychopharmacol 14:14111415.
Elsworth JD, Morrow BA, Hajszan T, Leranth C, and Roth RH (2011b)
Phencyclidine-induced loss of asymmetric spine synapses in rodent prefrontal
cortex is reversed by acute and chronic treatment with olanzapine. Neuropsychopharmacology 36:20542061.
Engler B, Freiman I, Urbanski M, and Szabo B (2006) Effects of exogenous and
endogenous cannabinoids on GABAergic neurotransmission between the caudateputamen and the globus pallidus in the mouse. J Pharmacol Exp Ther 316:
608617.
Enna SJ, Reisman SA, and Stanford JA (2006) CGP 56999A, a GABA(B) receptor
antagonist, enhances expression of brain-derived neurotrophic factor and attenuates dopamine depletion in the rat corpus striatum following a 6-hydroxydopamine
lesion of the nigrostriatal pathway. Neurosci Lett 406:102106.
Enomoto T and Floresco SB (2009) Disruptions in spatial working memory, but not
short-term memory, induced by repeated ketamine exposure. Prog Neuropsychopharmacol Biol Psychiatry 33:668675.
Erb S, Salmaso N, Rodaros D, and Stewart J (2001a) A role for the CRF-containing
pathway from central nucleus of the amygdala to bed nucleus of the stria terminalis in the stress-induced reinstatement of cocaine seeking in rats. Psychopharmacology (Berl) 158:360365.
Erb S, Shaham Y, and Stewart J (2001b) Stress-induced relapse to drug seeking in
the rat: role of the bed nucleus of the stria terminalis and amygdala. Stress 4:
289303.
Erb S and Stewart J (1999) A role for the bed nucleus of the stria terminalis, but not
the amygdala, in the effects of corticotropin-releasing factor on stress-induced reinstatement of cocaine seeking. J Neurosci 19:RC35 16.
Erhardt S, Math JM, Chergui K, Engberg G, and Svensson TH (2002) GABA(B)
receptor-mediated modulation of the firing pattern of ventral tegmental area dopamine neurons in vivo. Naunyn Schmiedebergs Arch Pharmacol 365:173180.
Ernst T, Chang L, Oropilla G, Gustavson A, and Speck O (2000) Cerebral perfusion
abnormalities in abstinent cocaine abusers: a perfusion MRI and SPECT study.
Psychiatry Res 99:6374.
Escalante OD and Ellinwood EH Jr (1970) Central nervous system cytopathological
changes in cats with chronic methedrine intoxication. Brain Res 21:151155.
Eshleman AJ, Wolfrum KM, Hatfield MG, Johnson RA, Murphy KV, and Janowsky A
(2013) Substituted methcathinones differ in transporter and receptor interactions.
Biochem Pharmacol 85:18031815.
Espiard ML, Lecardeur L, Abadie P, Halbecq I, and Dollfus S (2005) Hallucinogen
persisting perception disorder after psilocybin consumption: a case study. Eur
Psychiatry 20:458460.
Essrich C, Lorez M, Benson JA, Fritschy JM, and Lscher B (1998) Postsynaptic
clustering of major GABAA receptor subtypes requires the gamma 2 subunit and
gephyrin. Nat Neurosci 1:563571.
Evans CJ, Keith DE Jr, Morrison H, Magendzo K, and Edwards RH (1992) Cloning of
a delta opioid receptor by functional expression. Science 258:19521955.
Everitt B (1997) Craving cocaine cues: cognitive neuroscience meets drug addiction
research. Trends Cogn Sci 1:12.
Everitt BJ (2014) Neural and psychological mechanisms underlying compulsive drug
seeking habits and drug memoriesindications for novel treatments of addiction.
Eur J Neurosci 40:21632182.

978

Korpi et al.

Everitt BJ and Robbins TW (2005) Neural systems of reinforcement for drug addiction: from actions to habits to compulsion. Nat Neurosci 8:14811489.
Everitt BJ and Robbins TW (2013) From the ventral to the dorsal striatum: devolving
views of their roles in drug addiction. Neurosci Biobehav Rev 37 (9 Pt A):
19461954.
Eyerman DJ and Yamamoto BK (2007) A rapid oxidation and persistent decrease in
the vesicular monoamine transporter 2 after methamphetamine. J Neurochem
103:12191227.
Fagiolini M, Fritschy JM, Lw K, Mhler H, Rudolph U, and Hensch TK (2004)
Specific GABAA circuits for visual cortical plasticity. Science 303:16811683.
Fagiolini M and Hensch TK (2000) Inhibitory threshold for critical-period activation
in primary visual cortex. Nature 404:183186.
Falk EM, Cook VJ, Nichols DE, and Sprague JE (2002) An antisense oligonucleotide
targeted at MAO-B attenuates rat striatal serotonergic neurotoxicity induced by
MDMA. Pharmacol Biochem Behav 72:617622.
Farber NB, Heinkel C, Dribben WH, Nemmers B, and Jiang X (2004) In the adult
CNS, ethanol prevents rather than produces NMDA antagonist-induced neurotoxicity. Brain Res 1028:6674.
Farber NB, Kim SH, Dikranian K, Jiang XP, and Heinkel C (2002) Receptor mechanisms and circuitry underlying NMDA antagonist neurotoxicity. Mol Psychiatry
7:3243.
Farfel GM, Vosmer GL, and Seiden LS (1992) The N-methyl-D-aspartate antagonist
MK-801 protects against serotonin depletions induced by methamphetamine, 3,4methylenedioxymethamphetamine and p-chloroamphetamine. Brain Res 595:
121127.
Farooq U, Rajkumar R, Sukumaran S, Wu Y, Tan WH, and Dawe GS (2013)
Corticotropin-releasing factor infusion into nucleus incertus suppresses medial
prefrontal cortical activity and hippocampo-medial prefrontal cortical long-term
potentiation. Eur J Neurosci 38:25162525.
Farrant M and Nusser Z (2005) Variations on an inhibitory theme: phasic and tonic
activation of GABA(A) receptors. Nat Rev Neurosci 6:215229.
Fastbom J, Forsell Y, and Winblad B (1998) Benzodiazepines may have protective
effects against Alzheimer disease. Alzheimer Dis Assoc Disord 12:1417.
Fattore L, Fadda P, and Fratta W (2009) Sex differences in the self-administration of
cannabinoids and other drugs of abuse. Psychoneuroendocrinology 34 (Suppl 1):
S227S236.
Federici M, Latagliata EC, Ledonne A, Rizzo FR, Feligioni M, Sulzer D, Dunn M, Sames
D, Gu H, and Nistic R, et al. (2014) Paradoxical abatement of striatal dopaminergic
transmission by cocaine and methylphenidate. J Biol Chem 289:264274.
Feier G, Valvassori SS, Lopes-Borges J, Varela RB, Bavaresco DV, Scaini G, Morais
MO, Andersen ML, Streck EL, and Quevedo J (2012) Behavioral changes and brain
energy metabolism dysfunction in rats treated with methamphetamine or dextroamphetamine. Neurosci Lett 530:7579.
Feinberg-Zadek PL, Martin G, and Treistman SN (2008) BK channel subunit composition modulates molecular tolerance to ethanol. Alcohol Clin Exp Res 32:
12071216.
Fenster CP, Whitworth TL, Sheffield EB, Quick MW, and Lester RA (1999) Upregulation of surface alpha4beta2 nicotinic receptors is initiated by receptor desensitization after chronic exposure to nicotine. J Neurosci 19:48044814.
Fernndez-Ruiz J, Berrendero F, Hernndez ML, and Ramos JA (2000) The endogenous cannabinoid system and brain development. Trends Neurosci 23:1420.
Ferrari F, Ottani A, Vivoli R, and Giuliani D (1999) Learning impairment produced
in rats by the cannabinoid agonist HU 210 in a water-maze task. Pharmacol
Biochem Behav 64:555561.
Ferrario CR, Goussakov I, Stutzmann GE, and Wolf ME (2012) Withdrawal from
cocaine self-administration alters NMDA receptor-mediated Ca2+ entry in nucleus
accumbens dendritic spines. PLoS One 7:e40898.
Ferr S, Llus C, Justinova Z, Quiroz C, Orru M, Navarro G, Canela EI, Franco R,
and Goldberg SR (2010) Adenosine-cannabinoid receptor interactions. Implications
for striatal function. Br J Pharmacol 160:443453.
Ferreira PS, Nogueira TB, Costa VM, Branco PS, Ferreira LM, Fernandes E, Bastos
ML, Meisel A, Carvalho F, and Capela JP (2013) Neurotoxicity of ecstasy and its
metabolites in human dopaminergic differentiated SH-SY5Y cells. Toxicol Lett 216:
159170.
Fico TA and Vanderwende C (1989) Effects of prenatal phencyclidine on 3H-PCP
binding and PCP-induced motor activity and ataxia. Neurotoxicol Teratol 11:
373376.
Fields RD and Itoh K (1996) Neural cell adhesion molecules in activity-dependent
development and synaptic plasticity. Trends Neurosci 19:473480.
Filbey FM, Aslan S, Calhoun VD, Spence JS, Damaraju E, Caprihan A, and Segall J
(2014) Long-term effects of marijuana use on the brain. Proc Natl Acad Sci USA
111:1691316918.
File SE, Lister RG, and Nutt DJ (1982a) The anxiogenic action of benzodiazepine
antagonists. Neuropharmacology 21:10331037.
File SE, Lister RG, and Nutt DJ (1982b) Intrinsic actions of benzodiazepine antagonists. Neurosci Lett 32:165168.
Filip M, Frankowska M, Sadakierska-Chudy A, Suder A, Szumiec L, Mierzejewski P,
Bienkowski P, Przegali
nski E, and Cryan JF (2015) GABAB receptors as a therapeutic strategy in substance use disorders: focus on positive allosteric modulators.
Neuropharmacology 88:3647.
Fioravante D and Regehr WG (2011) Short-term forms of presynaptic plasticity. Curr
Opin Neurobiol 21:269274.
Fiorella D, Rabin RA, and Winter JC (1995) The role of the 5-HT2A and 5-HT2C
receptors in the stimulus effects of hallucinogenic drugs. I: Antagonist correlation
analysis. Psychopharmacology (Berl) 121:347356.
Fiorillo CD (2013) Two dimensions of value: dopamine neurons represent reward but
not aversiveness. Science 341:546549.
Fiorillo CD, Yun SR, and Song MR (2013) Diversity and homogeneity in responses of
midbrain dopamine neurons. J Neurosci 33:46934709.

Fischer JF and Cho AK (1979) Chemical release of dopamine from striatal homogenates:
evidence for an exchange diffusion model. J Pharmacol Exp Ther 208:203209.
Fischman MW and Foltin RW (1991) Utility of subjective-effects measurements in
assessing abuse liability of drugs in humans. Br J Addict 86:15631570.
Fishbain DA, Cole B, Lewis JE, and Gao J (2012) Is smoking associated with alcoholdrug dependence in patients with pain and chronic pain patients? An evidencebased structured review. Pain Med 13:12121226.
Fishbein M, Gov S, Assaf F, Gafni M, Keren O, and Sarne Y (2012) Long-term behavioral and biochemical effects of an ultra-low dose of D9-tetrahydrocannabinol
(THC): neuroprotection and ERK signaling. Exp Brain Res 221:437448.
Fitzgerald LW, Ortiz J, Hamedani AG, and Nestler EJ (1996) Drugs of abuse and
stress increase the expression of GluR1 and NMDAR1 glutamate receptor subunits
in the rat ventral tegmental area: common adaptations among cross-sensitizing
agents. J Neurosci 16:274282.
Fitzgerald ML, Shobin E, and Pickel VM (2012) Cannabinoid modulation of the dopaminergic circuitry: implications for limbic and striatal output. Prog Neuropsychopharmacol Biol Psychiatry 38:2129.
Fix AS, Ross JF, Stitzel SR, and Switzer RC (1996) Integrated evaluation of central
nervous system lesions: stains for neurons, astrocytes, and microglia reveal the
spatial and temporal features of MK-801-induced neuronal necrosis in the rat cerebral cortex. Toxicol Pathol 24:291304.
Flaishon R, Weinbroum AA, Veenman L, Leschiner S, Rudick V, and Gavish M (2003)
Flumazenil attenuates development of tolerance to diazepam after chronic treatment of mice with either isoflurane or diazepam. Anesth Analg 97:10461052.
Flatscher-Bader T, Harrison E, Matsumoto I, and Wilce PA (2010) Genes associated
with alcohol abuse and tobacco smoking in the human nucleus accumbens and
ventral tegmental area. Alcohol Clin Exp Res 34:12911302.
Flatscher-Bader T, van der Brug M, Hwang JW, Gochee PA, Matsumoto I, Niwa S,
and Wilce PA (2005) Alcohol-responsive genes in the frontal cortex and nucleus
accumbens of human alcoholics. J Neurochem 93:359370.
Flatscher-Bader T, van der Brug MP, Landis N, Hwang JW, Harrison E, and Wilce
PA (2006) Comparative gene expression in brain regions of human alcoholics.
Genes Brain Behav 5 (Suppl 1):7884.
Fleckenstein AE, Haughey HM, Metzger RR, Kokoshka JM, Riddle EL, Hanson JE,
Gibb JW, and Hanson GR (1999) Differential effects of psychostimulants and related agents on dopaminergic and serotonergic transporter function. Eur J Pharmacol 382:4549.
Fleckenstein AE, Metzger RR, Beyeler ML, Gibb JW, and Hanson GR (1997a) Oxygen radicals diminish dopamine transporter function in rat striatum. Eur J
Pharmacol 334:111114.
Fleckenstein AE, Metzger RR, Wilkins DG, Gibb JW, and Hanson GR (1997b) Rapid
and reversible effects of methamphetamine on dopamine transporters. J Pharmacol Exp Ther 282:834838.
Fleckenstein AE, Volz TJ, Riddle EL, Gibb JW, and Hanson GR (2007) New insights
into the mechanism of action of amphetamines. Annu Rev Pharmacol Toxicol 47:
681698.
Fleckenstein AE, Wilkins DG, Gibb JW, and Hanson GR (1997c) Interaction between
hyperthermia and oxygen radical formation in the 5-hydroxytryptaminergic response to a single methamphetamine administration. J Pharmacol Exp Ther 283:
281285.
Flores CM, Dvila-Garca MI, Ulrich YM, and Kellar KJ (1997) Differential regulation of neuronal nicotinic receptor binding sites following chronic nicotine administration. J Neurochem 69:22162219.
Floyd DW, Jung KY, and McCool BA (2003) Chronic ethanol ingestion facilitates
N-methyl-D-aspartate receptor function and expression in rat lateral/basolateral
amygdala neurons. J Pharmacol Exp Ther 307:10201029.
Flurkey K and Currer JM (2004) Pitfalls of animal model systems in ageing research.
Best Pract Res Clin Endocrinol Metab 18:407421.
Fog JU, Khoshbouei H, Holy M, Owens WA, Vaegter CB, Sen N, Nikandrova Y,
Bowton E, McMahon DG, and Colbran RJ, et al. (2006) Calmodulin kinase II
interacts with the dopamine transporter C terminus to regulate amphetamineinduced reverse transport. Neuron 51:417429.
Fokos S and Panagis G (2010) Effects of delta9-tetrahydrocannabinol on reward and
anxiety in rats exposed to chronic unpredictable stress. J Psychopharmacol 24:
767777.
Fone KC and Nutt DJ (2005) Stimulants: use and abuse in the treatment of attention
deficit hyperactivity disorder. Curr Opin Pharmacol 5:8793.
Foscarin S, Rossi F, and Carulli D (2012) Influence of the environment on adult CNS
plasticity and repair. Cell Tissue Res 349:161167.
Fowler CD and Kenny PJ (2014) Nicotine aversion: Neurobiological mechanisms and
relevance to tobacco dependence vulnerability. Neuropharmacology 76 (Pt B):
533544.
Fowler CD, Lu Q, Johnson PM, Marks MJ, and Kenny PJ (2011) Habenular a5
nicotinic receptor subunit signalling controls nicotine intake. Nature 471:597601.
Frahm S, Slimak MA, Ferrarese L, Santos-Torres J, Antolin-Fontes B, Auer S, Filkin
S, Pons S, Fontaine JF, and Tsetlin V, et al. (2011) Aversion to nicotine is regulated
by the balanced activity of b4 and a5 nicotinic receptor subunits in the medial
habenula. Neuron 70:522535.
Francesconi W, Berton F, Repunte-Canonigo V, Hagihara K, Thurbon D, Lekic D,
Specio SE, Greenwell TN, Chen SA, and Rice KC, et al. (2009) Protracted withdrawal from alcohol and drugs of abuse impairs long-term potentiation of intrinsic
excitability in the juxtacapsular bed nucleus of the stria terminalis. J Neurosci 29:
53895401.
Franks NP and Lieb WR (1984) Do general anaesthetics act by competitive binding to
specific receptors? Nature 310:599601.
Franks NP and Lieb WR (1994) Molecular and cellular mechanisms of general anaesthesia. Nature 367:607614.
Franks NP and Lieb WR (2004) Seeing the light: protein theories of general anesthesia. 1984. Anesthesiology 101:235237.

Drug-Induced Neuroplasticity
Fredriksson A, Archer T, Alm H, Gordh T, and Eriksson P (2004) Neurofunctional
deficits and potentiated apoptosis by neonatal NMDA antagonist administration.
Behav Brain Res 153:367376.
Freiman I and Szabo B (2005) Cannabinoids depress excitatory neurotransmission
between the subthalamic nucleus and the globus pallidus. Neuroscience 133:
305313.
French ED (1997) delta9-Tetrahydrocannabinol excites rat VTA dopamine neurons
through activation of cannabinoid CB1 but not opioid receptors. Neurosci Lett 226:
159162.
Frenzilli G, Ferrucci M, Giorgi FS, Blandini F, Nigro M, Ruggieri S, Murri L, Paparelli A,
and Fornai F (2007) DNA fragmentation and oxidative stress in the hippocampal
formation: a bridge between 3,4-methylenedioxymethamphetamine (ecstasy) intake
and long-lasting behavioral alterations. Behav Pharmacol 18:471481.
Freo U, Dam M, and Ori C (2008) The time-dependent effects of midazolam on regional cerebral glucose metabolism in rats. Anesth Analg 106:15161523.
Freund TF, Katona I, and Piomelli D (2003) Role of endogenous cannabinoids in
synaptic signaling. Physiol Rev 83:10171066.
Frey U and Morris RG (1997) Synaptic tagging and long-term potentiation. Nature
385:533536.
Frye GD and Fincher A (2000) Sustained ethanol inhibition of native AMPA receptors on
medial septum/diagonal band (MS/DB) neurons. Br J Pharmacol 129:8794.
Fukami G, Hashimoto K, Koike K, Okamura N, Shimizu E, and Iyo M (2004) Effect of
antioxidant N-acetyl-L-cysteine on behavioral changes and neurotoxicity in rats
after administration of methamphetamine. Brain Res 1016:9095.
Fukui K, Nakajima T, Kariyama H, Kashiba A, Kato N, Tohyama I, and Kimura H
(1989) Selective reduction of serotonin immunoreactivity in some forebrain regions
of rats induced by acute methamphetamine treatment; quantitative morphometric
analysis by serotonin immunocytochemistry. Brain Res 482:198203.
Fumagalli F, Bedogni F, Frasca A, Di Pasquale L, Racagni G, and Riva MA (2006)
Corticostriatal up-regulation of activity-regulated cytoskeletal-associated protein
expression after repeated exposure to cocaine. Mol Pharmacol 70:17261734.
Fung YK and Uretsky NJ (1982) The importance of calcium in the amphetamineinduced stimulation of dopamine synthesis in mouse striata in vivo. J Pharmacol
Exp Ther 223:477482.
Funk CK, Zorrilla EP, Lee MJ, Rice KC, and Koob GF (2007) Corticotropin-releasing
factor 1 antagonists selectively reduce ethanol self-administration in ethanoldependent rats. Biol Psychiatry 61:7886.
Furey ML and Drevets WC (2006) Antidepressant efficacy of the antimuscarinic drug
scopolamine: a randomized, placebo-controlled clinical trial. Arch Gen Psychiatry
63:11211129.
Fuxe K, Ferr S, Genedani S, Franco R, and Agnati LF (2007) Adenosine receptordopamine receptor interactions in the basal ganglia and their relevance for brain
function. Physiol Behav 92:210217.
Gable RS (1993) Toward a comparative overview of dependence potential and acute
toxicity of psychoactive substances used nonmedically. Am J Drug Alcohol Abuse
19:263281.
Gagnon M, Bergeron MJ, Lavertu G, Castonguay A, Tripathy S, Bonin RP, PerezSanchez J, Boudreau D, Wang B, and Dumas L, et al. (2013) Chloride extrusion
enhancers as novel therapeutics for neurological diseases. Nat Med 19:15241528.
Gainetdinov RR, Premont RT, Bohn LM, Lefkowitz RJ, and Caron MG (2004) Desensitization of G protein-coupled receptors and neuronal functions. Annu Rev
Neurosci 27:107144.
Gainetdinov RR, Sotnikova TD, and Caron MG (2002) Monoamine transporter
pharmacology and mutant mice. Trends Pharmacol Sci 23:367373.
Galvan A, Hare TA, Parra CE, Penn J, Voss H, Glover G, and Casey BJ (2006) Earlier
development of the accumbens relative to orbitofrontal cortex might underlie risktaking behavior in adolescents. J Neurosci 26:68856892.
Gambelunghe C, Bacci M, Aroni K, De Falco F, and Ayroldi EM (2014) Cocaine
addiction treatment and home remedies: use of the scopolamine transdermal
patch. Subst Use Misuse 49:12. DOI 10.3109/10826084.2013.824477
Gao M, Jin Y, Yang K, Zhang D, Lukas RJ, and Wu J (2010a) Mechanisms involved in
systemic nicotine-induced glutamatergic synaptic plasticity on dopamine neurons
in the ventral tegmental area. J Neurosci 30:1381413825.
Gao Y, Vasilyev DV, Goncalves MB, Howell FV, Hobbs C, Reisenberg M, Shen R,
Zhang MY, Strassle BW, and Lu P, et al. (2010b) Loss of retrograde endocannabinoid signaling and reduced adult neurogenesis in diacylglycerol lipase knock-out
mice. J Neurosci 30:20172024.
Garca-Ovejero D, Arvalo-Martn , Navarro-Galve B, Pinteaux E, Molina-Holgado
E, and Molina-Holgado F (2013) Neuroimmmune interactions of cannabinoids in
neurogenesis: focus on interleukin-1b (IL-1b) signalling. Biochem Soc Trans 41:
15771582.
Gardner EL, Paredes W, Smith D, Donner A, Milling C, Cohen D, and Morrison D
(1988) Facilitation of brain stimulation reward by delta 9-tetrahydrocannabinol.
Psychopharmacology (Berl) 96:142144.
Gasser P, Holstein D, Michel Y, Doblin R, Yazar-Klosinski B, Passie T,
and Brenneisen R (2014) Safety and efficacy of lysergic acid diethylamide-assisted
psychotherapy for anxiety associated with life-threatening diseases. J Nerv Ment
Dis 202:513520.
Gassmann M, Shaban H, Vigot R, Sansig G, Haller C, Barbieri S, Humeau Y, Schuler
V, Mller M, and Kinzel B, et al. (2004) Redistribution of GABAB(1) protein and
atypical GABAB responses in GABAB(2)-deficient mice. J Neurosci 24:60866097.
Gatch MB, Kozlenkov A, Huang RQ, Yang W, Nguyen JD, Gonzlez-Maeso J, Rice
KC, France CP, Dillon GH, and Forster MJ, et al. (2013) The HIV antiretroviral
drug efavirenz has LSD-like properties. Neuropsychopharmacology 38:23732384.
Gatley SJ, Pan D, Chen R, Chaturvedi G, and Ding YS (1996) Affinities of methylphenidate derivatives for dopamine, norepinephrine and serotonin transporters.
Life Sci 58:231239.
Gerdeman G and Lovinger DM (2001) CB1 cannabinoid receptor inhibits synaptic
release of glutamate in rat dorsolateral striatum. J Neurophysiol 85:468471.

979

Gerdeman GL, Partridge JG, Lupica CR, and Lovinger DM (2003) It could be habit
forming: drugs of abuse and striatal synaptic plasticity. Trends Neurosci 26:
184192.
Gerdeman GL, Ronesi J, and Lovinger DM (2002) Postsynaptic endocannabinoid
release is critical to long-term depression in the striatum. Nat Neurosci 5:446451.
Gerfen CR, Engber TM, Mahan LC, Susel Z, Chase TN, Monsma FJ Jr, and Sibley
DR (1990) D1 and D2 dopamine receptor-regulated gene expression of striatonigral
and striatopallidal neurons. Science 250:14291432.
Gerra G, Zaimovic A, Giusti F, Moi G, and Brewer C (2002) Intravenous flumazenil
versus oxazepam tapering in the treatment of benzodiazepine withdrawal: a randomized, placebo-controlled study. Addict Biol 7:385395.
Geyer MA and Vollenweider FX (2008) Serotonin research: contributions to understanding psychoses. Trends Pharmacol Sci 29:445453.
Giachino C, Barz M, Tchorz JS, Tome M, Gassmann M, Bischofberger J, Bettler B,
and Taylor V (2014) GABA suppresses neurogenesis in the adult hippocampus
through GABAB receptors. Development 141:8390.
Gianoulakis C (2001) Influence of the endogenous opioid system on high alcohol
consumption and genetic predisposition to alcoholism. J Psychiatry Neurosci 26:
304318.
Giedd JN (2004) Structural magnetic resonance imaging of the adolescent brain. Ann
N Y Acad Sci 1021:7785.
Giese KP, Fedorov NB, Filipkowski RK, and Silva AJ (1998) Autophosphorylation at
Thr286 of the alpha calcium-calmodulin kinase II in LTP and learning. Science
279:870873.
Gilman JM, Kuster JK, Lee S, Lee MJ, Kim BW, Makris N, van der Kouwe A, Blood
AJ, and Breiter HC (2014) Cannabis use is quantitatively associated with nucleus
accumbens and amygdala abnormalities in young adult recreational users.
J Neurosci 34:55295538.
Gilpin NW (2012) Neuropeptide Y (NPY) in the extended amygdala is recruited
during the transition to alcohol dependence. Neuropeptides 46:253259.
Gilpin NW, Misra K, Herman MA, Cruz MT, Koob GF, and Roberto M (2011) Neuropeptide Y opposes alcohol effects on gamma-aminobutyric acid release in
amygdala and blocks the transition to alcohol dependence. Biol Psychiatry 69:
10911099.
Gingrich JA and Hen R (2001) Dissecting the role of the serotonin system in neuropsychiatric disorders using knockout mice. Psychopharmacology (Berl) 155:110.
Gipson CD, Kupchik YM, Shen H, Reissner KJ, Thomas CA, and Kalivas PW (2013)
Relapse induced by cues predicting cocaine depends on rapid, transient synaptic
potentiation. Neuron 77:867872.
Giuffrida A, Parsons LH, Kerr TM, Rodrguez de Fonseca F, Navarro M, and Piomelli
D (1999) Dopamine activation of endogenous cannabinoid signaling in dorsal
striatum. Nat Neurosci 2:358363.
Glass MJ, Lane DA, Colago EE, Chan J, Schlussman SD, Zhou Y, Kreek MJ,
and Pickel VM (2008) Chronic administration of morphine is associated with
a decrease in surface AMPA GluR1 receptor subunit in dopamine D1 receptor
expressing neurons in the shell and non-D1 receptor expressing neurons in the core
of the rat nucleus accumbens. Exp Neurol 210:750761.
Glykys J, Peng Z, Chandra D, Homanics GE, Houser CR, and Mody I (2007) A new
naturally occurring GABA(A) receptor subunit partnership with high sensitivity to
ethanol. Nat Neurosci 10:4048.
Gogolla N, Caroni P, Lthi A, and Herry C (2009) Perineuronal nets protect fear
memories from erasure. Science 325:12581261.
Goldstein A (2001) Addiction: From Biology to Drug Policy, Oxford University Press,
New York.
Goldstein A and Naidu A (1989) Multiple opioid receptors: ligand selectivity profiles
and binding site signatures. Mol Pharmacol 36:265272.
Goldstein RZ, Alia-Klein N, Tomasi D, Carrillo JH, Maloney T, Woicik PA, Wang R,
Telang F, and Volkow ND (2009) Anterior cingulate cortex hypoactivations to an
emotionally salient task in cocaine addiction. Proc Natl Acad Sci USA 106:
94539458.
Goldstein RZ, Woicik PA, Maloney T, Tomasi D, Alia-Klein N, Shan J, Honorio J,
Samaras D, Wang R, and Telang F, et al. (2010) Oral methylphenidate normalizes
cingulate activity in cocaine addiction during a salient cognitive task. Proc Natl
Acad Sci USA 107:1666716672.
Gollub RL, Breiter HC, Kantor H, Kennedy D, Gastfriend D, Mathew RT, Makris N,
Guimaraes A, Riorden J, and Campbell T, et al. (1998) Cocaine decreases cortical
cerebral blood flow but does not obscure regional activation in functional magnetic
resonance imaging in human subjects. J Cereb Blood Flow Metab 18:724734.
Gmez-Climent MA, Guirado R, Castillo-Gmez E, Varea E, Gutierrez-Mecinas M,
Gilabert-Juan J, Garca-Momp C, Vidueira S, Sanchez-Mataredona D,
and Hernndez S, et al. (2011) The polysialylated form of the neural cell adhesion
molecule (PSA-NCAM) is expressed in a subpopulation of mature cortical interneurons characterized by reduced structural features and connectivity. Cereb
Cortex 21:10281041.
Gonalves J, Baptista S, and Silva AP (2014) Psychostimulants and brain dysfunction: a review of the relevant neurotoxic effects. Neuropharmacology 87:135149.
Goni-Allo B, O Mathna B, Segura M, Puerta E, Lasheras B, de la Torre R,
and Aguirre N (2008) The relationship between core body temperature and
3,4-methylenedioxymethamphetamine metabolism in rats: implications for neurotoxicity. Psychopharmacology (Berl) 197:263278.
Gonzalez AM, Walther D, Pazos A, and Uhl GR (1994) Synaptic vesicular monoamine
transporter expression: distribution and pharmacologic profile. Brain Res Mol
Brain Res 22:219226.
Gonzlez-Maeso J, Weisstaub NV, Zhou M, Chan P, Ivic L, Ang R, Lira A, BradleyMoore M, Ge Y, and Zhou Q, et al. (2007) Hallucinogens recruit specific cortical 5HT(2A) receptor-mediated signaling pathways to affect behavior. Neuron 53:
439452.
Gonzlez-Maeso J, Yuen T, Ebersole BJ, Wurmbach E, Lira A, Zhou M, Weisstaub N,
Hen R, Gingrich JA, and Sealfon SC (2003) Transcriptome fingerprints distinguish

980

Korpi et al.

hallucinogenic and nonhallucinogenic 5-hydroxytryptamine 2A receptor agonist


effects in mouse somatosensory cortex. J Neurosci 23:88368843.
Good CH and Lupica CR (2010) Afferent-specific AMPA receptor subunit composition
and regulation of synaptic plasticity in midbrain dopamine neurons by abused
drugs. J Neurosci 30:79007909.
Goodman DW (2007) Lisdexamfetamine dimesylate: the first prodrug stimulant.
Psychiatry (Edgmont) 4:3945.
Goodman JH and Sloviter RS (1993) Cocaine neurotoxicity and altered neuropeptide
Y immunoreactivity in the rat hippocampus; a silver degeneration and immunocytochemical study. Brain Res 616:263272.
Goodwin AK, Gibson KM, and Weerts EM (2013) Physical dependence on gammahydroxybutrate (GHB) prodrug 1,4-butanediol (1,4-BD): time course and severity of
withdrawal in baboons. Drug Alcohol Depend 132:427433.
Gopalakrishnan M, Molinari EJ, and Sullivan JP (1997) Regulation of human
alpha4beta2 neuronal nicotinic acetylcholine receptors by cholinergic channel
ligands and second messenger pathways. Mol Pharmacol 52:524534.
Goriounova NA and Mansvelder HD (2012a) Nicotine exposure during adolescence
alters the rules for prefrontal cortical synaptic plasticity during adulthood. Front
Synaptic Neurosci 4:3.
Goriounova NA and Mansvelder HD (2012b) Nicotine exposure during adolescence
leads to short- and long-term changes in spike timing-dependent plasticity in rat
prefrontal cortex. J Neurosci 32:1048410493.
Gotti C, Clementi F, Fornari A, Gaimarri A, Guiducci S, Manfredi I, Moretti M,
Pedrazzi P, Pucci L, and Zoli M (2009) Structural and functional diversity of native
brain neuronal nicotinic receptors. Biochem Pharmacol 78:703711.
Gotti C, Moretti M, Gaimarri A, Zanardi A, Clementi F, and Zoli M (2007) Heterogeneity and complexity of native brain nicotinic receptors. Biochem Pharmacol 74:
11021111.
Gotti C, Zoli M, and Clementi F (2006) Brain nicotinic acetylcholine receptors: native
subtypes and their relevance. Trends Pharmacol Sci 27:482491.
Goulenok C, Bernard B, Cadranel JF, Thabut D, Di Martino V, Opolon P,
and Poynard T (2002) Flumazenil vs. placebo in hepatic encephalopathy in patients
with cirrhosis: a meta-analysis. Aliment Pharmacol Ther 16:361372.
Gouzoulis-Mayfrank E, Daumann J, Tuchtenhagen F, Pelz S, Becker S, Kunert HJ,
Fimm B, and Sass H (2000) Impaired cognitive performance in drug free users of
recreational ecstasy (MDMA). J Neurol Neurosurg Psychiatry 68:719725.
Gouzoulis-Mayfrank E, Heekeren K, Neukirch A, Stoll M, Stock C, Obradovic M,
and Kovar KA (2005) Psychological effects of (S)-ketamine and N,N-dimethyltryptamine
(DMT): a double-blind, cross-over study in healthy volunteers. Pharmacopsychiatry 38:
301311.
Govind AP, Vezina P, and Green WN (2009) Nicotine-induced upregulation of nicotinic receptors: underlying mechanisms and relevance to nicotine addiction. Biochem Pharmacol 78:756765.
Grabowski J (Ed.) (1984) Cocaine: pharmacology, effects, and treatment of abuse.
NIDA Res Monogr 50. U.S. Government Printing Office, Washington, D.C.
Grace AA and Bunney BS (1984a) The control of firing pattern in nigral dopamine
neurons: burst firing. J Neurosci 4:28772890.
Grace AA and Bunney BS (1984b) The control of firing pattern in nigral dopamine
neurons: single spike firing. J Neurosci 4:28662876.
Grady SR, Salminen O, Laverty DC, Whiteaker P, McIntosh JM, Collins AC,
and Marks MJ (2007) The subtypes of nicotinic acetylcholine receptors on dopaminergic terminals of mouse striatum. Biochem Pharmacol 74:12351246.
Graham DL, Noailles PA, and Cadet JL (2008) Differential neurochemical consequences of an escalating dose-binge regimen followed by single-day multiple-dose
methamphetamine challenges. J Neurochem 105:18731885.
Graham-Jones S, Holt L, Gray JA, and Fillenz M (1985) Low-frequency septal
stimulation increases tyrosine hydroxylase activity in the hippocampus. Pharmacol Biochem Behav 23:489493.
Gramage E, Alguacil LF, and Herradon G (2008) Pleiotrophin prevents cocaineinduced toxicity in vitro. Eur J Pharmacol 595:3538.
Grammatopoulos DK (2012) Insights into mechanisms of corticotropin-releasing
hormone receptor signal transduction. Br J Pharmacol 166:8597.
Green AL and el Hait MA (1978) Inhibition of mouse brain monoamine oxidase by
(+)-amphetamine in vivo. J Pharm Pharmacol 30:262263.
Greengard P (2001) The neurobiology of slow synaptic transmission. Science 294:
10241030.
Grenhoff J, Aston-Jones G, and Svensson TH (1986) Nicotinic effects on the firing
pattern of midbrain dopamine neurons. Acta Physiol Scand 128:351358.
Gresch PJ, Barrett RJ, Sanders-Bush E, and Smith RL (2007) 5-Hydroxytryptamine
(serotonin)2A receptors in rat anterior cingulate cortex mediate the discriminative
stimulus properties of d-lysergic acid diethylamide. J Pharmacol Exp Ther 320:662669.
Gresch PJ, Smith RL, Barrett RJ, and Sanders-Bush E (2005) Behavioral tolerance to
lysergic acid diethylamide is associated with reduced serotonin-2A receptor signaling in rat cortex. Neuropsychopharmacology 30:16931702.
Griebel G, Pichat P, Beesk S, Leroy T, Redon N, Jacquet A, Franon D, Bert L, Even
L, and Lopez-Grancha M, et al. (2015) Selective blockade of the hydrolysis of the
endocannabinoid 2-arachidonoylglycerol impairs learning and memory performance while producing antinociceptive activity in rodents. Sci Rep 5:7642.
Griffen TC, Wang L, Fontanini A, and Maffei A (2012) Developmental regulation of
spatio-temporal patterns of cortical circuit activation. Front Cell Neurosci 6:65.
Griffiths RR and Johnson MW (2005) Relative abuse liability of hypnotic drugs:
a conceptual framework and algorithm for differentiating among compounds.
J Clin Psychiatry 66 (Suppl 9):3141.
Griffiths RR, Johnson MW, Richards WA, Richards BD, McCann U, and Jesse R
(2011) Psilocybin occasioned mystical-type experiences: immediate and persisting
dose-related effects. Psychopharmacology (Berl) 218:649665.
Griffiths RR, Richards WA, McCann U, and Jesse R (2006) Psilocybin can occasion
mystical-type experiences having substantial and sustained personal meaning and
spiritual significance. Psychopharmacology (Berl) 187:268283, discussion 284
292.

Grimm JW, Lu L, Hayashi T, Hope BT, Su TP, and Shaham Y (2003) Time-dependent
increases in brain-derived neurotrophic factor protein levels within the mesolimbic
dopamine system after withdrawal from cocaine: implications for incubation of
cocaine craving. J Neurosci 23:742747.
Grob CS, Danforth AL, Chopra GS, Hagerty M, McKay CR, Halberstadt AL,
and Greer GR (2011) Pilot study of psilocybin treatment for anxiety in patients
with advanced-stage cancer. Arch Gen Psychiatry 68:7178.
Grobin AC, Papadeas ST, and Morrow AL (2000) Regional variations in the effects of
chronic ethanol administration on GABA(A) receptor expression: potential mechanisms. Neurochem Int 37:453461.
Grover CA, Wallace KA, Lindberg SA, and Frye GD (1998) Ethanol inhibition of
NMDA currents in acutely dissociated medial septum/diagonal band neurons from
ethanol dependent rats. Brain Res 782:4352.
Grucza RA and Bierut LJ (2006) Co-occurring risk factors for alcohol dependence and
habitual smoking: update on findings from the Collaborative Study on the Genetics
of Alcoholism. Alcohol Res Health 29:172178.
Grueter BA, Robison AJ, Neve RL, Nestler EJ, and Malenka RC (2013) FosB differentially modulates nucleus accumbens direct and indirect pathway function.
Proc Natl Acad Sci USA 110:19231928.
Gu H, Salmeron BJ, Ross TJ, Geng X, Zhan W, Stein EA, and Yang Y (2010) Mesocorticolimbic circuits are impaired in chronic cocaine users as demonstrated by
resting-state functional connectivity. Neuroimage 53:593601.
Gu X, Lohrenz T, Salas R, Baldwin PR, Soltani A, Kirk U, Cinciripini PM,
and Montague PR (2015) Belief about nicotine selectively modulates value and
reward prediction error signals in smokers. Proc Natl Acad Sci USA 112:
25392544.
Guan G, Kramer SF, Bellinger LL, Wellman PJ, and Kramer PR (2004) Intermittent
nicotine administration modulates food intake in rats by acting on nicotine
receptors localized to the brainstem. Life Sci 74:27252737.
Guerra D, Sol A, Cam J, and Tobea A (1987) Neuropsychological performance in
opiate addicts after rapid detoxification. Drug Alcohol Depend 20:261270.
Gulaboski R, Cordeiro MN, Milhazes N, Garrido J, Borges F, Jorge M, Pereira CM,
Bogeski I, Morales AH, and Naumoski B, et al. (2007) Evaluation of the lipophilic
properties of opioids, amphetamine-like drugs, and metabolites through electrochemical studies at the interface between two immiscible solutions. Anal Biochem
361:236243.
Gulley JM and Juraska JM (2013) The effects of abused drugs on adolescent development of corticolimbic circuitry and behavior. Neuroscience 249:320.
Gulya K, Grant KA, Valverius P, Hoffman PL, and Tabakoff B (1991) Brain regional
specificity and time-course of changes in the NMDA receptor-ionophore complex
during ethanol withdrawal. Brain Res 547:129134.
Gulyas AI, Cravatt BF, Bracey MH, Dinh TP, Piomelli D, Boscia F, and Freund TF
(2004) Segregation of two endocannabinoid-hydrolyzing enzymes into pre- and
postsynaptic compartments in the rat hippocampus, cerebellum and amygdala.
Eur J Neurosci 20:441458.
Gupta AS, van der Meer MA, Touretzky DS, and Redish AD (2010) Hippocampal
replay is not a simple function of experience. Neuron 65:695705.
Haber SN, Fudge JL, and McFarland NR (2000) Striatonigrostriatal pathways in
primates form an ascending spiral from the shell to the dorsolateral striatum.
J Neurosci 20:23692382.
Haber SN and Knutson B (2010) The reward circuit: linking primate anatomy and
human imaging. Neuropsychopharmacology 35:426.
Hadlock GC, Webb KM, McFadden LM, Chu PW, Ellis JD, Allen SC, Andrenyak DM,
Vieira-Brock PL, German CL, and Conrad KM, et al. (2011) 4-Methylmethcathinone (mephedrone): neuropharmacological effects of a designer stimulant of abuse. J Pharmacol Exp Ther 339:530536.
Haggerty GC, Forney RB, and Johnson JM (1984) The effect of a single administration of phencyclidine on behavior in the rat over a 21-day period. Toxicol Appl
Pharmacol 75:444453.
Hahn J, Hopf FW, and Bonci A (2009) Chronic cocaine enhances corticotropinreleasing factor-dependent potentiation of excitatory transmission in ventral tegmental area dopamine neurons. J Neurosci 29:65356544.
Hjos N, Katona I, Naiem SS, MacKie K, Ledent C, Mody I, and Freund TF (2000)
Cannabinoids inhibit hippocampal GABAergic transmission and network oscillations. Eur J Neurosci 12:32393249.
Hall BJ, Wells C, Allenby C, Lin MY, Hao I, Marshall L, Rose JE, and Levin ED
(2014) Differential effects of non-nicotine tobacco constituent compounds on nicotine self-administration in rats. Pharmacol Biochem Behav 120:103108.
Hall W (2015) What has research over the past two decades revealed about the
adverse health effects of recreational cannabis use? Addiction 110:1935.
Halonen T, Lehtinen M, Pitknen A, Ylinen A, and Riekkinen PJ (1988) Inhibitory
and excitatory amino acids in CSF of patients suffering from complex partial
seizures during chronic treatment with gamma-vinyl GABA (vigabatrin). Epilepsy
Res 2:246252.
Halonen T, Miettinen R, Toppinen A, Tuunanen J, Kotti T, and Riekkinen PJ Sr
(1995) Vigabatrin protects against kainic acid-induced neuronal damage in the rat
hippocampus. Neurosci Lett 195:1316.
Halpern JH and Pope HG Jr (2003) Hallucinogen persisting perception disorder:
what do we know after 50 years? Drug Alcohol Depend 69:109119.
Halpern JH, Sherwood AR, Hudson JI, Gruber S, Kozin D, and Pope HG Jr (2011)
Residual neurocognitive features of long-term ecstasy users with minimal exposure
to other drugs. Addiction 106:777786.
Halpin LE, Collins SA, and Yamamoto BK (2014) Neurotoxicity of methamphetamine
and 3,4-methylenedioxymethamphetamine. Life Sci 97:3744.
Hama H, Kurokawa H, Kawano H, Ando R, Shimogori T, Noda H, Fukami K, SakaueSawano A, and Miyawaki A (2011) Scale: a chemical approach for fluorescence
imaging and reconstruction of transparent mouse brain. Nat Neurosci 14:
14811488.

Drug-Induced Neuroplasticity
Hamid S, Dawe GS, Gray JA, and Stephenson JD (1997) Nicotine induces longlasting potentiation in the dentate gyrus of nicotine-primed rats. Neurosci Res 29:
8185.
Han DD and Gu HH (2006) Comparison of the monoamine transporters from human
and mouse in their sensitivities to psychostimulant drugs. BMC Pharmacol 6:6.
Han DH, Yoon SJ, Sung YH, Lee YS, Kee BS, Lyoo IK, Renshaw PF, and Cho SC
(2008) A preliminary study: novelty seeking, frontal executive function, and dopamine receptor (D2) TaqI A gene polymorphism in patients with methamphetamine dependence. Compr Psychiatry 49:387392.
Han MH, Bolaos CA, Green TA, Olson VG, Neve RL, Liu RJ, Aghajanian GK,
and Nestler EJ (2006) Role of cAMP response element-binding protein in the rat
locus ceruleus: regulation of neuronal activity and opiate withdrawal behaviors.
J Neurosci 26:46244629.
Han WY, Du P, Fu SY, Wang F, Song M, Wu CF, and Yang JY (2014) Oxytocin via its
receptor affects restraint stress-induced methamphetamine CPP reinstatement in
mice: Involvement of the medial prefrontal cortex and dorsal hippocampus glutamatergic system. Pharmacol Biochem Behav 119:8087.
Hanchar HJ, Dodson PD, Olsen RW, Otis TS, and Wallner M (2005) Alcohol-induced
motor impairment caused by increased extrasynaptic GABA(A) receptor activity.
Nat Neurosci 8:339345.
Hansson AC, Cippitelli A, Sommer WH, Fedeli A, Bjrk K, Soverchia L, Terasmaa A,
Massi M, Heilig M, and Ciccocioppo R (2006) Variation at the rat Crhr1 locus and
sensitivity to relapse into alcohol seeking induced by environmental stress. Proc
Natl Acad Sci USA 103:1523615241.
Hao S, Liu S, Zheng X, Zheng W, Ouyang H, Mata M, and Fink DJ (2011) The role of
TNFa in the periaqueductal gray during naloxone-precipitated morphine withdrawal in rats. Neuropsychopharmacology 36:664676.
Harkness PC and Millar NS (2002) Changes in conformation and subcellular distribution of alpha4beta2 nicotinic acetylcholine receptors revealed by chronic nicotine treatment and expression of subunit chimeras. J Neurosci 22:1017210181.
Harley C (1991) Noradrenergic and locus coeruleus modulation of the perforant pathevoked potential in rat dentate gyrus supports a role for the locus coeruleus in
attentional and memorial processes. Prog Brain Res 88:307321.
Harley CW (1998) Noradrenergic long-term potentiation in the dentate gyrus. Adv
Pharmacol 42:952956.
Harley CW (2007) Norepinephrine and the dentate gyrus. Prog Brain Res 163:
299318.
Harris DS, Baggott M, Mendelson JH, Mendelson JE, and Jones RT (2002) Subjective
and hormonal effects of 3,4-methylenedioxymethamphetamine (MDMA) in
humans. Psychopharmacology (Berl) 162:396405.
Harris GC, Wimmer M, and Aston-Jones G (2005) A role for lateral hypothalamic
orexin neurons in reward seeking. Nature 437:556559.
Harris GC, Wimmer M, Randall-Thompson JF, and Aston-Jones G (2007) Lateral
hypothalamic orexin neurons are critically involved in learning to associate an
environment with morphine reward. Behav Brain Res 183:4351.
Harris RA (1999) Ethanol actions on multiple ion channels: which are important?
Alcohol Clin Exp Res 23:15631570.
Hart CL, Marvin CB, Silver R, and Smith EE (2012) Is cognitive functioning impaired in
methamphetamine users? A critical review. Neuropsychopharmacology 37:586608.
Hashimoto K, Fujita Y, and Iyo M (2007) Phencyclidine-induced cognitive deficits in
mice are improved by subsequent subchronic administration of fluvoxamine: role of
sigma-1 receptors. Neuropsychopharmacology 32:514521.
Hashimoto T, Kajii Y, and Nishikawa T (1998) Psychotomimetic-induction of tissue
plasminogen activator mRNA in corticostriatal neurons in rat brain. Eur J Neurosci 10:33873399.
Hashimoto T, Nguyen QL, Rotaru D, Keenan T, Arion D, Beneyto M, GonzalezBurgos G, and Lewis DA (2009) Protracted developmental trajectories of GABAA
receptor alpha1 and alpha2 subunit expression in primate prefrontal cortex. Biol
Psychiatry 65:10151023.
Hatami H, Hossainpour-Faizi MA, Azarfarin M, and Azarfam P(2010) Chronic ecstasy use increases neurotrophin-4 gene expression and protein levels in the rat
brain. Pharmacol Rep 62:9981004.
Hatzidimitriou G, McCann UD, and Ricaurte GA (1999) Altered serotonin
innervation patterns in the forebrain of monkeys treated with (+/-)3,4methylenedioxymethamphetamine seven years previously: factors influencing abnormal recovery. J Neurosci 19:50965107.
Haughey HM, Fleckenstein AE, and Hanson GR (1999) Differential regional effects of
methamphetamine on the activities of tryptophan and tyrosine hydroxylase.
J Neurochem 72:661668.
Hauser J, Rudolph U, Keist R, Mhler H, Feldon J, and Yee BK (2005) Hippocampal
alpha5 subunit-containing GABAA receptors modulate the expression of prepulse
inhibition. Mol Psychiatry 10:201207.
Hayashi T and Su TP (2007) Sigma-1 receptor chaperones at the ER-mitochondrion
interface regulate Ca(2+) signaling and cell survival. Cell 131:596610.
Hayashi Y, Shi SH, Esteban JA, Piccini A, Poncer JC, and Malinow R (2000) Driving
AMPA receptors into synapses by LTP and CaMKII: requirement for GluR1 and
PDZ domain interaction. Science 287:22622267.
Hazell P (2011) The challenges to demonstrating long-term effects of psychostimulant treatment for attention-deficit/hyperactivity disorder. Curr Opin Psychiatry
24:286290.
Hebb D (1949) The Organization of Behavior, Wiley, New York.
Hechtman L, Abikoff H, Klein RG, Weiss G, Respitz C, Kouri J, Blum C, Greenfield
B, Etcovitch J, and Fleiss K, et al. (2004) Academic achievement and emotional
status of children with ADHD treated with long-term methylphenidate and multimodal psychosocial treatment. J Am Acad Child Adolesc Psychiatry 43:812819.
Heese K, Otten U, Mathivet P, Raiteri M, Marescaux C, and Bernasconi R (2000)
GABA(B) receptor antagonists elevate both mRNA and protein levels of the neurotrophins nerve growth factor (NGF) and brain-derived neurotrophic factor
(BDNF) but not neurotrophin-3 (NT-3) in brain and spinal cord of rats. Neuropharmacology 39:449462.

981

Heffernan TM, Jarvis H, Rodgers J, Scholey AB, and Ling J (2001) Prospective
memory, everyday cognitive failure and central executive function in recreational
users of Ecstasy. Hum Psychopharmacol 16:607612.
Heifets BD and Castillo PE (2009) Endocannabinoid signaling and long-term synaptic plasticity. Annu Rev Physiol 71:283306.
Heikkila RE, Orlansky H, Mytilineou C, and Cohen G (1975) Amphetamine: evaluation of d- and l-isomers as releasing agents and uptake inhibitors for 3Hdopamine and 3H-norepinephrine in slices of rat neostriatum and cerebral cortex.
J Pharmacol Exp Ther 194:4756.
Heikkinen AE, Mykkynen TP, and Korpi ER (2009) Long-lasting modulation of
glutamatergic transmission in VTA dopamine neurons after a single dose of benzodiazepine agonists. Neuropsychopharmacology 34:290298.
Heilig M and Egli M (2006) Pharmacological treatment of alcohol dependence: target
symptoms and target mechanisms. Pharmacol Ther 111:855876.
Heilig M and Koob GF (2007) A key role for corticotropin-releasing factor in alcohol
dependence. Trends Neurosci 30:399406.
Helton DR, Tizzano JP, Monn JA, Schoepp DD, and Kallman MJ (1997) LY354740:
a metabotropic glutamate receptor agonist which ameliorates symptoms of nicotine
withdrawal in rats. Neuropharmacology 36:15111516.
Henderson BJ, Srinivasan R, Nichols WA, Dilworth CN, Gutierrez DF, Mackey ED,
McKinney S, Drenan RM, Richards CI, and Lester HA (2014) Nicotine exploits
a COPI-mediated process for chaperone-mediated up-regulation of its receptors.
J Gen Physiol 143:5166.
Hensch TK (2005) Critical period plasticity in local cortical circuits. Nat Rev Neurosci
6:877888.
Hensch TK (2014) Bistable parvalbumin circuits pivotal for brain plasticity. Cell 156:
1719.
Hensch TK and Bilimoria PM (2012) Re-opening windows: manipulating critical
periods for brain development. Cerebrum 2012:11. http://dana.org/news/cerebrum/
detail.aspx?id=39360
Herkenham M, Lynn AB, Johnson MR, Melvin LS, de Costa BR, and Rice KC (1991)
Characterization and localization of cannabinoid receptors in rat brain: a quantitative in vitro autoradiographic study. J Neurosci 11:563583.
Herkenham M, Lynn AB, Little MD, Johnson MR, Melvin LS, de Costa BR, and Rice KC
(1990) Cannabinoid receptor localization in brain. Proc Natl Acad Sci USA 87:19321936.
Herkenham M and Nauta WJ (1979) Efferent connections of the habenular nuclei in
the rat. J Comp Neurol 187:1947.
Herman MA, Contet C, Justice NJ, Vale W, and Roberto M (2013) Novel subunitspecific tonic GABA currents and differential effects of ethanol in the central
amygdala of CRF receptor-1 reporter mice. J Neurosci 33:32843298.
Hermann D, Weber-Fahr W, Sartorius A, Hoerst M, Frischknecht U, Tunc-Skarka N,
Perreau-Lenz S, Hansson AC, Krumm B, and Kiefer F, et al. (2012) Translational
magnetic resonance spectroscopy reveals excessive central glutamate levels during
alcohol withdrawal in humans and rats. Biol Psychiatry 71:10151021.
Hermanson DJ, Gamble-George JC, Marnett LJ, and Patel S (2014) Substrateselective COX-2 inhibition as a novel strategy for therapeutic endocannabinoid
augmentation. Trends Pharmacol Sci 35:358367.
Hernandez CM and Dineley KT (2012) a7 nicotinic acetylcholine receptors in Alzheimers disease: neuroprotective, neurotrophic or both? Curr Drug Targets 13:
613622.
Herring NR, Schaefer TL, Gudelsky GA, Vorhees CV, and Williams MT (2008) Effect
of +-methamphetamine on path integration learning, novel object recognition, and
neurotoxicity in rats. Psychopharmacology (Berl) 199:637650.
Herrold AA, Voigt RM, and Napier TC (2013) mGluR5 is necessary for maintenance
of methamphetamine-induced associative learning. Eur Neuropsychopharmacol
23:691696.
Hester R and Garavan H (2004) Executive dysfunction in cocaine addiction: evidence
for discordant frontal, cingulate, and cerebellar activity. J Neurosci 24:
1101711022.
Hevers W, Hadley SH, Lddens H, and Amin J (2008) Ketamine, but not phencyclidine, selectively modulates cerebellar GABA(A) receptors containing alpha6 and
delta subunits. J Neurosci 28:53835393.
Hevers W and Lddens H (1998) The diversity of GABAA receptors. Pharmacological
and electrophysiological properties of GABAA channel subtypes. Mol Neurobiol 18:
3586.
Hickman M, Vickerman P, Macleod J, Lewis G, Zammit S, Kirkbride J, and Jones P
(2009) If cannabis caused schizophreniahow many cannabis users may need to be
prevented in order to prevent one case of schizophrenia? England and Wales calculations. Addiction 104:18561861.
Higuera-Matas A, Botreau F, Miguns M, Del Olmo N, Borcel E, Prez-Alvarez L,
Garca-Lecumberri C, and Ambrosio E (2009) Chronic periadolescent cannabinoid
treatment enhances adult hippocampal PSA-NCAM expression in male Wistar rats
but only has marginal effects on anxiety, learning and memory. Pharmacol Biochem Behav 93:482490.
Hikosaka O (2010) The habenula: from stress evasion to value-based decision-making.
Nat Rev Neurosci 11:503513.
Hilrio MR, Clouse E, Yin HH, and Costa RM (2007) Endocannabinoid signaling is
critical for habit formation. Front Integr Neurosci 1:6.
Hilrio MR and Costa RM (2008) High on habits. Front Neurosci 2:208217.
Hilber B, Scholze P, Dorostkar MM, Sandtner W, Holy M, Boehm S, Singer EA,
and Sitte HH (2005) Serotonin-transporter mediated efflux: a pharmacological
analysis of amphetamines and non-amphetamines. Neuropharmacology 49:
811819.
Hill MN, Froc DJ, Fox CJ, Gorzalka BB, and Christie BR (2004) Prolonged cannabinoid treatment results in spatial working memory deficits and impaired longterm potentiation in the CA1 region of the hippocampus in vivo. Eur J Neurosci 20:
859863.
Hirata H and Cadet JL (1997) Methamphetamine-induced serotonin neurotoxicity is
attenuated in p53-knockout mice. Brain Res 768:345348.

982

Korpi et al.

Hitzemann B and Hitzemann R (1999) Chlordiazepoxide-induced expression of c-Fos


in the central extended amygdala and other brain regions of the C57BL/6J and
DBA/2J inbred mouse strains: relationships to mechanisms of ethanol action. Alcohol Clin Exp Res 23:11581172.
Hnasko TS and Edwards RH (2012) Neurotransmitter corelease: mechanism and
physiological role. Annu Rev Physiol 74:225243.
Ho SY, Chen CH, Liu TH, Chang HF, and Liou JC (2012) Protein kinase mz is
necessary for cocaine-induced synaptic potentiation in the ventral tegmental area.
Biol Psychiatry 71:706713.
Hochbaum DR, Zhao Y, Farhi SL, Klapoetke N, Werley CA, Kapoor V, Zou P, Kralj
JM, Maclaurin D, and Smedemark-Margulies N, et al. (2014) All-optical electrophysiology in mammalian neurons using engineered microbial rhodopsins. Nat
Methods 11:825833.
Hodges AB, Ladenheim B, McCoy MT, Beauvais G, Cai N, Krasnova IN, and Cadet
JL (2011) Long-term protective effects of methamphetamine preconditioning
against single-day methamphetamine toxic challenges. Curr Neuropharmacol 9:
3539.
Hoffman AF and Lupica CR (2013) Synaptic targets of D9-tetrahydrocannabinol in
the central nervous system. Cold Spring Harb Perspect Med 3:3. DOI 10.1101/
cshperspect.a012237
Hoffman AF, Oz M, Caulder T, and Lupica CR (2003) Functional tolerance and
blockade of long-term depression at synapses in the nucleus accumbens after
chronic cannabinoid exposure. J Neurosci 23:48154820.
Hoffman EJ and Warren EW (1993) Flumazenil: a benzodiazepine antagonist. Clin
Pharm 12:641656, quiz 699701.
Hoffman PL, Rabe CS, Moses F, and Tabakoff B (1989) N-methyl-D-aspartate
receptors and ethanol: inhibition of calcium flux and cyclic GMP production.
J Neurochem 52:19371940.
Hoffman WF, Schwartz DL, Huckans MS, McFarland BH, Meiri G, Stevens AA,
and Mitchell SH (2008) Cortical activation during delay discounting in abstinent
methamphetamine dependent individuals. Psychopharmacology (Berl) 201:
183193.
Hofmaier T, Luf A, Seddik A, Stockner T, Holy M, Freissmuth M, Ecker GF, Schmid
R, Sitte HH, and Kudlacek O (2014) Aminorex, a metabolite of the cocaine adulterant levamisole, exerts amphetamine like actions at monoamine transporters.
Neurochem Int 73:3241.
Hofmann SG and Smits JA (2008) Cognitive-behavioral therapy for adult anxiety
disorders: a meta-analysis of randomized placebo-controlled trials. J Clin Psychiatry 69:621632.
Hogan KA, Staal RG, and Sonsalla PK (2000) Analysis of VMAT2 binding after
methamphetamine or MPTP treatment: disparity between homogenates and vesicle preparations. J Neurochem 74:22172220.
Hohmann AG and Herkenham M (2000) Localization of cannabinoid CB(1) receptor
mRNA in neuronal subpopulations of rat striatum: a double-label in situ hybridization study. Synapse 37:7180.
Holbrook BD and Rayburn WF (2014) Teratogenic risks from exposure to illicit drugs.
Obstet Gynecol Clin North Am 41:229239.
Hollander JA, Lu Q, Cameron MD, Kamenecka TM, and Kenny PJ (2008) Insular
hypocretin transmission regulates nicotine reward. Proc Natl Acad Sci USA 105:
1948019485.
Holmes JC and Rutledge CO (1976) Effects of the d- and l-isomers of amphetamine on
uptake, release and catabolism of norepinephrine, dopamine and 5-hydroxytryptamine
in several regions of rat brain. Biochem Pharmacol 25:447451.
Hlscher C (1999) Synaptic plasticity and learning and memory: LTP and beyond.
J Neurosci Res 58:6275.
Hlter SM, Kallnik M, Wurst W, Marsicano G, Lutz B, and Wotjak CT (2005) Cannabinoid CB1 receptor is dispensable for memory extinction in an appetitivelymotivated learning task. Eur J Pharmacol 510:6974.
Holtmaat A and Svoboda K (2009) Experience-dependent structural synaptic plasticity in the mammalian brain. Nat Rev Neurosci 10:647658.
Hondo H, Yonezawa Y, Nakahara T, Nakamura K, Hirano M, Uchimura H,
and Tashiro N (1994) Effect of phencyclidine on dopamine release in the rat prefrontal cortex; an in vivo microdialysis study. Brain Res 633:337342.
Honey GD, Bullmore ET, and Sharma T (2000) Prolonged reaction time to a verbal
working memory task predicts increased power of posterior parietal cortical activation. Neuroimage 12:495503.
Hong WC and Amara SG (2013) Differential targeting of the dopamine transporter to
recycling or degradative pathways during amphetamine- or PKC-regulated endocytosis in dopamine neurons. FASEB J 27:29953007.
Hope BT, Nye HE, Kelz MB, Self DW, Iadarola MJ, Nakabeppu Y, Duman RS,
and Nestler EJ (1994) Induction of a long-lasting AP-1 complex composed of altered
Fos-like proteins in brain by chronic cocaine and other chronic treatments. Neuron
13:12351244.
Horiguchi M, Hannaway KE, Adelekun AE, Huang M, Jayathilake K, and Meltzer
HY (2013) D(1) receptor agonists reverse the subchronic phencyclidine (PCP)induced novel object recognition (NOR) deficit in female rats. Behav Brain Res 238:
3643.
Horiguchi M and Meltzer HY (2012) The role of 5-HT1A receptors in phencyclidine
(PCP)-induced novel object recognition (NOR) deficit in rats. Psychopharmacology
(Berl) 221:205215.
Hornby PJ (2001) Central neurocircuitry associated with emesis. Am J Med 111
(Suppl 8A):106S112S.
Horowitz MJ (1969) Flashbacks: recurrent intrusive images after the use of LSD. Am
J Psychiatry 126:565569.
Horvath TL, Diano S, Sotonyi P, Heiman M, and Tschp M (2001) Minireview:
ghrelin and the regulation of energy balancea hypothalamic perspective. Endocrinology 142:41634169.
Hotchkiss AJ and Gibb JW (1980) Long-term effects of multiple doses of methamphetamine on tryptophan hydroxylase and tyrosine hydroxylase activity in rat
brain. J Pharmacol Exp Ther 214:257262.

Howard RJ, Murail S, Ondricek KE, Corringer PJ, Lindahl E, Trudell JR, and Harris
RA (2011) Structural basis for alcohol modulation of a pentameric ligand-gated ion
channel. Proc Natl Acad Sci USA 108:1214912154.
Howlett AC, Barth F, Bonner TI, Cabral G, Casellas P, Devane WA, Felder CC,
Herkenham M, Mackie K, and Martin BR, et al. (2002) International Union of
Pharmacology. XXVII. Classification of cannabinoid receptors. Pharmacol Rev 54:
161202.
Hsiang HL, Epp JR, van den Oever MC, Yan C, Rashid AJ, Insel N, Ye L, Niibori Y,
Deisseroth K, and Frankland PW, et al. (2014) Manipulating a cocaine engram in
mice. J Neurosci 34:1411514127.
Hsu PD, Lander ES, and Zhang F (2014) Development and applications of CRISPRCas9 for genome engineering. Cell 157:12621278.
Huang H, Xu Y, and van den Pol AN (2011) Nicotine excites hypothalamic arcuate
anorexigenic proopiomelanocortin neurons and orexigenic neuropeptide Y neurons:
similarities and differences. J Neurophysiol 106:11911202.
Huang L, Liu Y, Jin W, Ji X, and Dong Z (2012) Ketamine potentiates hippocampal
neurodegeneration and persistent learning and memory impairment through the
PKCg-ERK signaling pathway in the developing brain. Brain Res 1476:164171.
Huang YH, Lin Y, Mu P, Lee BR, Brown TE, Wayman G, Marie H, Liu W, Yan Z,
and Sorg BA, et al. (2009) In vivo cocaine experience generates silent synapses.
Neuron 63:4047.
Huang YY and Kandel ER (1996) Modulation of both the early and the late phase of
mossy fiber LTP by the activation of beta-adrenergic receptors. Neuron 16:
611617.
Huang ZJ, Kirkwood A, Pizzorusso T, Porciatti V, Morales B, Bear MF, Maffei L,
and Tonegawa S (1999) BDNF regulates the maturation of inhibition and the
critical period of plasticity in mouse visual cortex. Cell 98:739755.
Huganir RL and Nicoll RA (2013) AMPARs and synaptic plasticity: the last 25 years.
Neuron 80:704717.
Hulse G, ONeil G, Morris N, Bennett K, Norman A, and Hood S (2013) Withdrawal
and psychological sequelae, and patient satisfaction associated with subcutaneous
flumazenil infusion for the management of benzodiazepine withdrawal: a case
series. J Psychopharmacol 27:222227.
Hunkeler W, Mhler H, Pieri L, Polc P, Bonetti EP, Cumin R, Schaffner R,
and Haefely W (1981) Selective antagonists of benzodiazepines. Nature 290:
514516.
Huopaniemi L, Keist R, Randolph A, Certa U, and Rudolph U (2004) Diazepaminduced adaptive plasticity revealed by alpha1 GABAA receptor-specific expression
profiling. J Neurochem 88:10591067.
Hutchinson MR, Northcutt AL, Hiranita T, Wang X, Lewis SS, Thomas J, van Steeg
K, Kopajtic TA, Loram LC, and Sfregola C, et al. (2012) Opioid activation of toll-like
receptor 4 contributes to drug reinforcement. J Neurosci 32:1118711200.
Hutchinson MR, Zhang Y, Shridhar M, Evans JH, Buchanan MM, Zhao TX, Slivka
PF, Coats BD, Rezvani N, and Wieseler J, et al. (2010) Evidence that opioids may
have toll-like receptor 4 and MD-2 effects. Brain Behav Immun 24:8395.
Huttenlocher PR (1979) Synaptic density in human frontal cortex - developmental
changes and effects of aging. Brain Res 163:195205.
Huttenlocher PR (1990) Morphometric study of human cerebral cortex development.
Neuropsychologia 28:517527.
Huttenlocher PR and Dabholkar AS (1997) Regional differences in synaptogenesis in
human cerebral cortex. J Comp Neurol 387:167178.
Hyland BI, Reynolds JN, Hay J, Perk CG, and Miller R (2002) Firing modes of
midbrain dopamine cells in the freely moving rat. Neuroscience 114:475492.
Hyyti P and Koob GF (1995) GABAA receptor antagonism in the extended amygdala
decreases ethanol self-administration in rats. Eur J Pharmacol 283:151159.
Hyyti P and Sinclair JD (1993) Responding for oral ethanol after naloxone treatment by alcohol-preferring AA rats. Alcohol Clin Exp Res 17:631636.
Ikonomidou C, Bittigau P, Ishimaru MJ, Wozniak DF, Koch C, Genz K, Price MT,
Stefovska V, Hrster F, and Tenkova T, et al. (2000) Ethanol-induced apoptotic
neurodegeneration and fetal alcohol syndrome. Science 287:10561060.
Ikonomidou C, Bosch F, Miksa M, Bittigau P, Vckler J, Dikranian K, Tenkova TI,
Stefovska V, Turski L, and Olney JW (1999) Blockade of NMDA receptors and
apoptotic neurodegeneration in the developing brain. Science 283:7074.
Ikonomidou C, Mosinger JL, and Olney JW (1989) Hypothermia enhances protective
effect of MK-801 against hypoxic/ischemic brain damage in infant rats. Brain Res
487:184187.
Ilango A, Kesner AJ, Keller KL, Stuber GD, Bonci A, and Ikemoto S (2014) Similar
roles of substantia nigra and ventral tegmental dopamine neurons in reward and
aversion. J Neurosci 34:817822.
Imam SZ, Crow JP, Newport GD, Islam F, Slikker W Jr, and Ali SF (1999) Methamphetamine generates peroxynitrite and produces dopaminergic neurotoxicity in mice:
protective effects of peroxynitrite decomposition catalyst. Brain Res 837:1521.
Imperato A, Mulas A, and Di Chiara G (1986) Nicotine preferentially stimulates
dopamine release in the limbic system of freely moving rats. Eur J Pharmacol 132:
337338.
Ingram SL, Prasad BM, and Amara SG (2002) Dopamine transporter-mediated
conductances increase excitability of midbrain dopamine neurons. Nat Neurosci 5:
971978.
Ingram SL, Vaughan CW, Bagley EE, Connor M, and Christie MJ (1998) Enhanced
opioid efficacy in opioid dependence is caused by an altered signal transduction
pathway. J Neurosci 18:1026910276.
Iiguez SD, Warren BL, Parise EM, Alcantara LF, Schuh B, Maffeo ML, Manojlovic
Z, and Bolaos-Guzmn CA (2009) Nicotine exposure during adolescence induces
a depression-like state in adulthood. Neuropsychopharmacology 34:16091624.
Insel TR, Battaglia G, Johannessen JN, Marra S, and De Souza EB (1989) 3,4Methylenedioxymethamphetamine (ecstasy) selectively destroys brain serotonin
terminals in rhesus monkeys. J Pharmacol Exp Ther 249:713720.
Iorio KR, Reinlib L, Tabakoff B, and Hoffman PL (1992) Chronic exposure of cerebellar granule cells to ethanol results in increased N-methyl-D-aspartate receptor
function. Mol Pharmacol 41:11421148.

Drug-Induced Neuroplasticity
Irifune M, Shimizu T, and Nomoto M (1991) Ketamine-induced hyperlocomotion
associated with alteration of presynaptic components of dopamine neurons in the
nucleus accumbens of mice. Pharmacol Biochem Behav 40:399407.
Ito K, Sawada Y, Sugiyama Y, Suzuki H, Hanano M, and Iga T (1994) Linear relationship between GABAA receptor occupancy of muscimol and glucose metabolic
response in the conscious mouse brain. Clinical implication based on comparison
with benzodiazepine receptor agonist. Drug Metab Dispos 22:5054.
Ito R, Dalley JW, Howes SR, Robbins TW, and Everitt BJ (2000) Dissociation in
conditioned dopamine release in the nucleus accumbens core and shell in response
to cocaine cues and during cocaine-seeking behavior in rats. J Neurosci 20:
74897495.
Ito R, Dalley JW, Robbins TW, and Everitt BJ (2002) Dopamine release in the dorsal
striatum during cocaine-seeking behavior under the control of a drug-associated
cue. J Neurosci 22:62476253.
Itzhak Y and Ali SF (1996) The neuronal nitric oxide synthase inhibitor,
7-nitroindazole, protects against methamphetamine-induced neurotoxicity in
vivo. J Neurochem 67:17701773.
Itzhak Y and Ali SF (2002) Repeated administration of gamma-hydroxybutyric acid
(GHB) to mice: assessment of the sedative and rewarding effects of GHB. Ann N Y
Acad Sci 965:451460.
Itzhak Y, Roig-Cantisano A, Dombro RS, and Norenberg MD (1995) Acute liver
failure and hyperammonemia increase peripheral-type benzodiazepine receptor
binding and pregnenolone synthesis in mouse brain. Brain Res 705:345348.
Ivanov A and Aston-Jones G (2001) Local opiate withdrawal in locus coeruleus
neurons in vitro. J Neurophysiol 85:23882397.
Iversen L (2000) The Science of Marijuana, Oxford University Press, New York.
Iversen L, Gibbons S, Treble R, Setola V, Huang XP, and Roth BL (2013) Neurochemical profiles of some novel psychoactive substances. Eur J Pharmacol 700:
147151.
Iversen L, White M, and Treble R (2014) Designer psychostimulants: pharmacology
and differences. Neuropharmacology 87:5965.
Iversen LL (2006) Speed, Ecstasy, Ritalin: The Science of Amphetamines, Oxford, UK,
Oxford University Press.
Iyo M, Namba H, Yanagisawa M, Hirai S, Yui N, and Fukui S (1997) Abnormal
cerebral perfusion in chronic methamphetamine abusers: a study using 99MTcHMPAO and SPECT. Prog Neuropsychopharmacol Biol Psychiatry 21:789796.
Izumi T, Inoue T, Tsuchiya K, Hashimoto S, Ohmori T, and Koyama T (1999) Effects
of the benzodiazepine antagonist flumazenil on conditioned fear stress in rats. Prog
Neuropsychopharmacol Biol Psychiatry 23:12471258.
Izumi Y and Zorumski CF (2014) Metaplastic effects of subanesthetic ketamine on
CA1 hippocampal function. Neuropharmacology 86:273281.
Izzo AA, Borrelli F, Capasso R, Di Marzo V, and Mechoulam R (2009) Nonpsychotropic plant cannabinoids: new therapeutic opportunities from an ancient
herb. Trends Pharmacol Sci 30:515527.
Jacobsen LK, Krystal JH, Mencl WE, Westerveld M, Frost SJ, and Pugh KR (2005)
Effects of smoking and smoking abstinence on cognition in adolescent tobacco
smokers. Biol Psychiatry 57:5666.
Jacobsen LK, Mencl WE, Constable RT, Westerveld M, and Pugh KR (2007) Impact of
smoking abstinence on working memory neurocircuitry in adolescent daily tobacco
smokers. Psychopharmacology (Berl) 193:557566.
Jacobson LH and Cryan JF (2008) Evaluation of the anxiolytic-like profile of the
GABAB receptor positive modulator CGP7930 in rodents. Neuropharmacology 54:
854862.
Jaffe JH and Kanzler M (1979) Smoking as an addictive disorder. NIDA Res Monogr
23:423.
Jalil SJ, Sacktor TC, and Shouval HZ (2015) Atypical PKCs in memory maintenance:
the roles of feedback and redundancy. Learn Mem 22:344353.
James A, James C, and Thwaites T (2013) The brain effects of cannabis in healthy
adolescents and in adolescents with schizophrenia: a systematic review. Psychiatry
Res 214:181189.
Jan RK, Lin JC, Miles SW, Kydd RR, and Russell BR (2012) Striatal volume
increases in active methamphetamine-dependent individuals and correlation with
cognitive performance. Brain Sci 2:553572.
Jansen KL (1990) Ketaminecan chronic use impair memory? Int J Addict 25:
133139.
Jansen KL (2000) A review of the nonmedical use of ketamine: use, users and consequences. J Psychoactive Drugs 32:419433.
Jansen KL and Darracot-Cankovic R (2001) The nonmedical use of ketamine, part
two: A review of problem use and dependence. J Psychoactive Drugs 33:151158.
Jasinski DR and Krishnan S (2009) Abuse liability and safety of oral lisdexamfetamine dimesylate in individuals with a history of stimulant abuse. J Psychopharmacol 23:419427.
Javitt DC and Zukin SR (1991) Recent advances in the phencyclidine model of
schizophrenia. Am J Psychiatry 148:13011308.
Javitt DC, Zukin SR, Heresco-Levy U, and Umbricht D (2012) Has an angel shown
the way? Etiological and therapeutic implications of the PCP/NMDA model of
schizophrenia. Schizophr Bull 38:958966.
Jayanthi S, Deng X, Bordelon M, McCoy MT, and Cadet JL (2001) Methamphetamine
causes differential regulation of pro-death and anti-death Bcl-2 genes in the mouse
neocortex. FASEB J 15:17451752.
Jayanthi S, Deng X, Noailles PA, Ladenheim B, and Cadet JL (2004) Methamphetamine induces neuronal apoptosis via cross-talks between endoplasmic reticulum
and mitochondria-dependent death cascades. FASEB J 18:238251.
Jayanthi S, McCoy MT, Chen B, Britt JP, Kourrich S, Yau HJ, Ladenheim B,
Krasnova IN, Bonci A, and Cadet JL (2014) Methamphetamine downregulates
striatal glutamate receptors via diverse epigenetic mechanisms. Biol Psychiatry
76:4756.
Jebelli AK, Doan N, and Ellison G (2002) Prenatal phencyclidine induces heightened
neurodegeneration in rats in some brain regions, especially during 2nd trimester,
but possible anti-apoptotic effects in others. Pharmacol Toxicol 90:2025.

983

Jennings JH, Sparta DR, Stamatakis AM, Ung RL, Pleil KE, Kash TL, and Stuber
GD (2013) Distinct extended amygdala circuits for divergent motivational states.
Nature 496:224228.
Jensen AA and Spalding TA (2004) Allosteric modulation of G-protein coupled
receptors. Eur J Pharm Sci 21:407420.
Jentsch JD, Taylor JR, Elsworth JD, Redmond DE Jr, and Roth RH (1999) Altered
frontal cortical dopaminergic transmission in monkeys after subchronic phencyclidine exposure: involvement in frontostriatal cognitive deficits. Neuroscience 90:
823832.
Jerlhag E, Egecioglu E, Landgren S, Salom N, Heilig M, Moechars D, Datta R,
Perrissoud D, Dickson SL, and Engel JA (2009) Requirement of central ghrelin
signaling for alcohol reward. Proc Natl Acad Sci USA 106:1131811323.
Jevtovic-Todorovic V, Absalom AR, Blomgren K, Brambrink A, Crosby G, Culley DJ,
Fiskum G, Giffard RG, Herold KF, and Loepke AW, et al. (2013) Anaesthetic
neurotoxicity and neuroplasticity: an expert group report and statement based on
the BJA Salzburg Seminar. Br J Anaesth 111:143151.
Jha P, Ramasundarahettige C, Landsman V, Rostron B, Thun M, Anderson RN,
McAfee T, and Peto R (2013) 21st-century hazards of smoking and benefits of
cessation in the United States. N Engl J Med 368:341350.
Jhou TC, Fields HL, Baxter MG, Saper CB, and Holland PC (2009a) The rostromedial
tegmental nucleus (RMTg), a GABAergic afferent to midbrain dopamine neurons,
encodes aversive stimuli and inhibits motor responses. Neuron 61:786800.
Jhou TC, Geisler S, Marinelli M, Degarmo BA, and Zahm DS (2009b) The mesopontine rostromedial tegmental nucleus: A structure targeted by the lateral
habenula that projects to the ventral tegmental area of Tsai and substantia nigra
compacta. J Comp Neurol 513:566596.
Jia F, Chandra D, Homanics GE, and Harrison NL (2008) Ethanol modulates synaptic and extrasynaptic GABAA receptors in the thalamus. J Pharmacol Exp Ther
326:475482.
Jin C and Woodward JJ (2006) Effects of 8 different NR1 splice variants on the
ethanol inhibition of recombinant NMDA receptors. Alcohol Clin Exp Res 30:
673679.
Jin Z, Bazov I, Kononenko O, Korpi ER, Bakalkin G, and Birnir B (2011) Selective
Changes of GABA(A) Channel Subunit mRNAs in the Hippocampus and Orbitofrontal Cortex but not in Prefrontal Cortex of Human Alcoholics. Front Cell Neurosci 5:30. DOI10.3389/fncel.2011.00030
Jin Z, Bhandage AK, Bazov I, Kononenko O, Bakalkin G, Korpi ER, and Birnir B
(2014a) Expression of specific ionotropic glutamate and GABA-A receptor subunits
is decreased in central amygdala of alcoholics. Front Cell Neurosci 8:288.
DOI 10.3389/fncel.2014.00288
Jin Z, Bhandage AK, Bazov I, Kononenko O, Bakalkin G, Korpi ER, and Birnir B
(2014b) Selective increases of AMPA, NMDA, and kainate receptor subunit
mRNAs in the hippocampus and orbitofrontal cortex but not in prefrontal cortex of
human alcoholics. Front Cell Neurosci 8:11. DOI 10.3389/fncel.2014.00011
Johanson CE, Frey KA, Lundahl LH, Keenan P, Lockhart N, Roll J, Galloway GP,
Koeppe RA, Kilbourn MR, and Robbins T, et al. (2006) Cognitive function and
nigrostriatal markers in abstinent methamphetamine abusers. Psychopharmacology (Berl) 185:327338.
John CE and Jones SR (2007) Voltammetric characterization of the effect of monoamine uptake inhibitors and releasers on dopamine and serotonin uptake in mouse
caudate-putamen and substantia nigra slices. Neuropharmacology 52:15961605.
Johnson JG, Cohen P, Pine DS, Klein DF, Kasen S, and Brook JS (2000) Association
between cigarette smoking and anxiety disorders during adolescence and early
adulthood. JAMA 284:23482351.
Johnson LA, Guptaroy B, Lund D, Shamban S, and Gnegy ME (2005) Regulation of
amphetamine-stimulated dopamine efflux by protein kinase C beta. J Biol Chem
280:1091410919.
Johnson M, Richards W, and Griffiths R (2008) Human hallucinogen research:
guidelines for safety. J Psychopharmacol 22:603620.
Johnson MW and Griffiths RR (2013) Comparative abuse liability of GHB and ethanol in humans. Exp Clin Psychopharmacol 21:112123.
Johnson SW and North RA (1992) Opioids excite dopamine neurons by hyperpolarization of local interneurons. J Neurosci 12:483488.
Johnson WD 2nd, Howard RJ, Trudell JR, and Harris RA (2012) The TM2 69 position
of GABA(A) receptors mediates alcohol inhibition. J Pharmacol Exp Ther 340:
445456.
Jonas P and Burnashev N (1995) Molecular mechanisms controlling calcium entry
through AMPA-type glutamate receptor channels. Neuron 15:987990.
Jones A, Korpi ER, McKernan RM, Pelz R, Nusser Z, Mkel R, Mellor JR, Pollard S,
Bahn S, and Stephenson FA, et al. (1997) Ligand-gated ion channel subunit
partnerships: GABAA receptor alpha6 subunit gene inactivation inhibits delta
subunit expression. J Neurosci 17:13501362.
Jones IW and Wonnacott S (2004) Precise localization of alpha7 nicotinic acetylcholine receptors on glutamatergic axon terminals in the rat ventral tegmental
area. J Neurosci 24:1124411252.
Jones KA, Srivastava DP, Allen JA, Strachan RT, Roth BL, and Penzes P (2009)
Rapid modulation of spine morphology by the 5-HT2A serotonin receptor through
kalirin-7 signaling. Proc Natl Acad Sci USA 106:1957519580.
Jonsson S, Adermark L, Ericson M, and Sderpalm B (2014) The involvement of
accumbal glycine receptors in the dopamine-elevating effects of addictive drugs.
Neuropharmacology 82:6975.
Joseph MH, Peters SL, Prior A, Mitchell SN, Brazell MP, and Gray JA (1990) Chronic
nicotine administration increases tyrosine hydroxylase selectivity in the rat hippocampus. Neurochem Int 16:269273.
Jupp B, Krivdic B, Krstew E, and Lawrence AJ (2011) The orexin receptor antagonist SB-334867 dissociates the motivational properties of alcohol and sucrose in
rats. Brain Res 1391:5459.
Jutras-Aswad D, DiNieri JA, Harkany T, and Hurd YL (2009) Neurobiological consequences of maternal cannabis on human fetal development and its neuropsychiatric outcome. Eur Arch Psychiatry Clin Neurosci 259:395412.

984

Korpi et al.

Kaasinen V, Aalto S, Ngren K, and Rinne JO (2004) Expectation of caffeine induces


dopaminergic responses in humans. Eur J Neurosci 19:23522356.
Kadota T and Kadota K (2004) Neurotoxic morphological changes induced in the
medial prefrontal cortex of rats behaviorally sensitized to methamphetamine. Arch
Histol Cytol 67:241251.
Kahlig KM, Binda F, Khoshbouei H, Blakely RD, McMahon DG, Javitch JA,
and Galli A (2005) Amphetamine induces dopamine efflux through a dopamine
transporter channel. Proc Natl Acad Sci USA 102:34953500.
Kaila K, Price TJ, Payne JA, Puskarjov M, and Voipio J (2014) Cation-chloride
cotransporters in neuronal development, plasticity and disease. Nat Rev Neurosci
15:637654.
Kaiser S, Foltz LA, George CA, Kirkwood SC, Bemis KG, Lin X, Gelbert LM,
and Nisenbaum LK (2004) Phencyclidine-induced changes in rat cortical gene expression identified by microarray analysis: implications for schizophrenia. Neurobiol Dis 16:220235.
Kalechstein AD, Newton TF, and Green M (2003) Methamphetamine dependence is
associated with neurocognitive impairment in the initial phases of abstinence.
J Neuropsychiatry Clin Neurosci 15:215220.
Kalivas PW (2009) The glutamate homeostasis hypothesis of addiction. Nat Rev
Neurosci 10:561572.
Kallupi M, Oleata CS, Luu G, Teshima K, Ciccocioppo R, and Roberto M (2014a) MT7716, a novel selective nonpeptidergic NOP receptor agonist, effectively blocks
ethanol-induced increase in GABAergic transmission in the rat central amygdala.
Front Integr Neurosci 8:18. DOI 10.3389/fnint.2014.00018
Kallupi M, Varodayan FP, Oleata CS, Correia D, Luu G, and Roberto M (2014b)
Nociceptin/orphanin FQ decreases glutamate transmission and blocks ethanolinduced effects in the central amygdala of naive and ethanol-dependent rats.
Neuropsychopharmacology 39:10811092.
Kalsbeek A, Voorn P, Buijs RM, Pool CW, and Uylings HB (1988) Development of the
dopaminergic innervation in the prefrontal cortex of the rat. J Comp Neurol 269:5872.
Kam PC and Yoong FF (1998) Gamma-hydroxybutyric acid: an emerging recreational
drug. Anaesthesia 53:11951198.
Kandel DB and Chen K (2000) Extent of smoking and nicotine dependence in the
United States: 1991-1993. Nicotine Tob Res 2:263274.
Kane JK, Parker SL, and Li MD (2001) Hypothalamic orexin-A binding sites are
downregulated by chronic nicotine treatment in the rat. Neurosci Lett 298:14.
Kane JK, Parker SL, Matta SG, Fu Y, Sharp BM, and Li MD (2000) Nicotine upregulates expression of orexin and its receptors in rat brain. Endocrinology 141:
36233629.
Kano M, Ohno-Shosaku T, Hashimotodani Y, Uchigashima M, and Watanabe M
(2009) Endocannabinoid-mediated control of synaptic transmission. Physiol Rev
89:309380.
Kantor L and Gnegy ME (1998) Protein kinase C inhibitors block amphetaminemediated dopamine release in rat striatal slices. J Pharmacol Exp Ther 284:592598.
Kapelewski CH, Vandenbergh DJ, and Klein LC (2011) Effect of the monoamine
oxidase inhibition on rewarding effects of nicotine in rodents. Curr Drug Abuse Rev
4:110121.
Kapur S and Seeman P (2001) Ketamine has equal affinity for NMDA receptors and
the high-affinity state of the dopamine D2 receptor. Biol Psychiatry 49:954957.
Kapur S and Seeman P (2002) NMDA receptor antagonists ketamine and PCP have
direct effects on the dopamine D(2) and serotonin 5-HT(2)receptors-implications for
models of schizophrenia. Mol Psychiatry 7:837844.
Karahanian E, Quintanilla ME, Tampier L, Rivera-Meza M, Bustamante D,
Gonzalez-Lira V, Morales P, Herrera-Marschitz M, and Israel Y (2011) Ethanol
as a prodrug: brain metabolism of ethanol mediates its reinforcing effects. Alcohol Clin Exp Res 35:606612.
Karama S, Ducharme S, Corley J, Chouinard-Decorte F, Starr JM, Wardlaw JM,
Bastin ME, and Deary IJ (2015) Cigarette smoking and thinning of the brains
cortex. Mol Psychiatry 20:778785.
Karpova NN, Pickenhagen A, Lindholm J, Tiraboschi E, Kulesskaya N, Agstsdttir
A, Antila H, Popova D, Akamine Y, and Bahi A, et al. (2011) Fear erasure in mice
requires synergy between antidepressant drugs and extinction training. Science
334:17311734.
Kasai H, Fukuda M, Watanabe S, Hayashi-Takagi A, and Noguchi J (2010) Structural dynamics of dendritic spines in memory and cognition. Trends Neurosci 33:
121129.
Kasanetz F, Deroche-Gamonet V, Berson N, Balado E, Lafourcade M, Manzoni O,
and Piazza PV (2010) Transition to addiction is associated with a persistent impairment in synaptic plasticity. Science 328:17091712.
Kash TL, Baucum AJ 2nd, Conrad KL, Colbran RJ, and Winder DG (2009) Alcohol
exposure alters NMDAR function in the bed nucleus of the stria terminalis. Neuropsychopharmacology 34:24202429.
Kashihara K, Hamamura T, Okumura K, and Otsuki S (1990) Effect of MK-801 on
endogenous dopamine release in vivo. Brain Res 528:8082.
Kashiwa A, Nishikawa T, Nishijima K, Umino A, and Takahashi K (1995) Dizocilpine
(MK-801) elicits a tetrodotoxin-sensitive increase in extracellular release of dopamine in rat medial frontal cortex. Neurochem Int 26:269279.
Kato HK, Watabe AM, and Manabe T (2009) Non-Hebbian synaptic plasticity induced by repetitive postsynaptic action potentials. J Neurosci 29:1115311160.
Katona I and Freund TF (2008) Endocannabinoid signaling as a synaptic circuit
breaker in neurological disease. Nat Med 14:923930.
Katona I and Freund TF (2012) Multiple functions of endocannabinoid signaling in
the brain. Annu Rev Neurosci 35:529558.
Katona I, Sperlgh B, Maglczky Z, Sntha E, Kfalvi A, Czirjk S, Mackie K, Vizi
ES, and Freund TF (2000) GABAergic interneurons are the targets of cannabinoid
actions in the human hippocampus. Neuroscience 100:797804.
Katona I, Urbn GM, Wallace M, Ledent C, Jung KM, Piomelli D, Mackie K,
and Freund TF (2006) Molecular composition of the endocannabinoid system at
glutamatergic synapses. J Neurosci 26:56285637.

Kauer JA (2003) Addictive drugs and stress trigger a common change at VTA synapses. Neuron 37:549550.
Kauer JA and Malenka RC (2007) Synaptic plasticity and addiction. Nat Rev Neurosci 8:844858.
Kaufling J, Veinante P, Pawlowski SA, Freund-Mercier MJ, and Barrot M (2009)
Afferents to the GABAergic tail of the ventral tegmental area in the rat. J Comp
Neurol 513:597621.
Kaufling J, Waltisperger E, Bourdy R, Valera A, Veinante P, Freund-Mercier MJ,
and Barrot M (2010) Pharmacological recruitment of the GABAergic tail of the
ventral tegmental area by acute drug exposure. Br J Pharmacol 161:16771691.
Kaufman MJ, Levin JM, Maas LC, Rose SL, Lukas SE, Mendelson JH, Cohen BM,
and Renshaw PF (1998) Cocaine decreases relative cerebral blood volume in
humans: a dynamic susceptibility contrast magnetic resonance imaging study.
Psychopharmacology (Berl) 138:7681.
Kaupmann K, Cryan JF, Wellendorph P, Mombereau C, Sansig G, Klebs K, Schmutz
M, Froestl W, van der Putten H, and Mosbacher J, et al. (2003) Specific gammahydroxybutyrate-binding sites but loss of pharmacological effects of gammahydroxybutyrate in GABA(B)(1)-deficient mice. Eur J Neurosci 18:27222730.
Kavalali ET and Monteggia LM (2012) Synaptic mechanisms underlying rapid antidepressant action of ketamine. Am J Psychiatry 169:11501156.
Kawamata J, Suzuki S, and Shimohama S (2012) a7 nicotinic acetylcholine receptor
mediated neuroprotection in Parkinsons disease. Curr Drug Targets 13:623630.
Kearn CS, Blake-Palmer K, Daniel E, Mackie K, and Glass M (2005) Concurrent
stimulation of cannabinoid CB1 and dopamine D2 receptors enhances heterodimer
formation: a mechanism for receptor cross-talk? Mol Pharmacol 67:16971704.
Kehoe LA, Bernardinelli Y, and Muller D (2013) GluN3A: an NMDA receptor subunit
with exquisite properties and functions. Neural Plast 2013:145387.
Kehr J, Ichinose F, Yoshitake S, Goiny M, Sievertsson T, Nyberg F, and Yoshitake T
(2011) Mephedrone, compared with MDMA (ecstasy) and amphetamine, rapidly
increases both dopamine and 5-HT levels in nucleus accumbens of awake rats. Br J
Pharmacol 164:19491958.
Keimpema E, Mackie K, and Harkany T (2011) Molecular model of cannabis sensitivity in developing neuronal circuits. Trends Pharmacol Sci 32:551561.
Kelly PA, Ford I, and McCulloch J (1986) The effect of diazepam upon local cerebral
glucose use in the conscious rat. Neuroscience 19:257265.
Kelly PA and McCulloch J (1982) Effects of the putative GABAergic agonists, muscimol and THIP, upon local cerebral glucose utilisation. J Neurochem 39:613624.
Kelm MK, Criswell HE, and Breese GR (2007) Calcium release from presynaptic
internal stores is required for ethanol to increase spontaneous gammaaminobutyric acid release onto cerebellum Purkinje neurons. J Pharmacol Exp
Ther 323:356364.
Kelm MK, Criswell HE, and Breese GR (2008) The role of protein kinase A in the
ethanol-induced increase in spontaneous GABA release onto cerebellar Purkinje
neurons. J Neurophysiol 100:34173428.
Kelz MB, Chen J, Carlezon WA Jr, Whisler K, Gilden L, Beckmann AM, Steffen C,
Zhang YJ, Marotti L, and Self DW, et al. (1999) Expression of the transcription
factor deltaFosB in the brain controls sensitivity to cocaine. Nature 401:272276.
Kennedy PJ, Feng J, Robison AJ, Maze I, Badimon A, Mouzon E, Chaudhury D,
Damez-Werno DM, Haggarty SJ, and Han MH, et al. (2013) Class I HDAC inhibition blocks cocaine-induced plasticity by targeted changes in histone methylation. Nat Neurosci 16:434440.
Kenny PJ, Gasparini F, and Markou A (2003) Group II metabotropic and alphaamino-3-hydroxy-5-methyl-4-isoxazole propionate (AMPA)/kainate glutamate
receptors regulate the deficit in brain reward function associated with nicotine
withdrawal in rats. J Pharmacol Exp Ther 306:10681076.
Kenny PJ and Markou A (2004) The ups and downs of addiction: role of metabotropic
glutamate receptors. Trends Pharmacol Sci 25:265272.
Khoshbouei H, Sen N, Guptaroy B, Johnson L, Lund D, Gnegy ME, Galli A,
and Javitch JA (2004) N-terminal phosphorylation of the dopamine transporter is
required for amphetamine-induced efflux. PLoS Biol 2:E78.
Kieffer BL, Befort K, Gaveriaux-Ruff C, and Hirth CG (1992) The delta-opioid receptor: isolation of a cDNA by expression cloning and pharmacological characterization. Proc Natl Acad Sci USA 89:1204812052.
Kieffer BL and Evans CJ (2002) Opioid tolerance-in search of the holy grail. Cell 108:
587590.
Kikuchi-Utsumi K, Ishizaka M, Matsumura N, Watabe M, Aoyama K, Sasakawa N,
and Nakaki T (2013) Involvement of the a(1D)-adrenergic receptor in methamphetamineinduced hyperthermia and neurotoxicity in rats. Neurotox Res 24:130138.
Kilb W (2012) Development of the GABAergic system from birth to adolescence.
Neuroscientist 18:613630.
Kim D and Thayer SA (2001) Cannabinoids inhibit the formation of new synapses
between hippocampal neurons in culture. J Neurosci 21:RC146.
Kim HS and Jang CG (1997) MK-801 inhibits methamphetamine-induced conditioned place preference and behavioral sensitization to apomorphine in mice. Brain
Res Bull 44:221227.
Kim HS, Jang CG, and Park WK (1996) Inhibition by MK-801 of morphine-induced
conditioned place preference and postsynaptic dopamine receptor supersensitivity
in mice. Pharmacol Biochem Behav 55:1117.
Kim J, Park BH, Lee JH, Park SK, and Kim JH (2011) Cell type-specific alterations
in the nucleus accumbens by repeated exposures to cocaine. Biol Psychiatry 69:
10261034.
Kim JH, Byun H, and Kim JH (2013a) Abuse potential of propofol used for sedation
in gastric endoscopy and its correlation with subject characteristics. Korean J
Anesthesiol 65:403409.
Kim KW, Kim DC, Kim YH, Eun YA, Kim HI, and Cho KP (1995) Ca2+-dependent
and -independent mechanisms of ischaemia-evoked release of [3H]-dopamine from
rat striatal slices. Clin Exp Pharmacol Physiol 22:301302.
Kim SJ, Lyoo IK, Hwang J, Chung A, Hoon Sung Y, Kim J, Kwon DH, Chang KH,
and Renshaw PF (2006) Prefrontal grey-matter changes in short-term and

Drug-Induced Neuroplasticity
long-term abstinent methamphetamine abusers. Int J Neuropsychopharmacol 9:
221228.
Kim SJ, Lyoo IK, Hwang J, Sung YH, Lee HY, Lee DS, Jeong DU, and Renshaw PF
(2005) Frontal glucose hypometabolism in abstinent methamphetamine users.
Neuropsychopharmacology 30:13831391.
Kim SY, Adhikari A, Lee SY, Marshel JH, Kim CK, Mallory CS, Lo M, Pak S, Mattis
J, and Lim BK, et al. (2013b) Diverging neural pathways assemble a behavioural
state from separable features in anxiety. Nature 496:219223.
Kim YT, Lee SW, Kwon DH, Seo JH, Ahn BC, and Lee J (2009) Dose-dependent
frontal hypometabolism on FDG-PET in methamphetamine abusers. J Psychiatr
Res 43:11661170.
King LA, Fortson QCR, Ujvary I, Ramsey J, and Nutt DJ (2014) Scopolamine: useful
medicine or dangerous drug? Sci Justice 54:321322.
King W Jr and Ellison G (1989) Long-lasting alterations in behavior and brain
neurochemistry following continuous low-level LSD administration. Pharmacol
Biochem Behav 33:6973.
Kiraly DD, Nemirovsky NE, LaRese TP, Tomek SE, Yahn SL, Olive MF, Eipper BA,
and Mains RE (2013) Constitutive knockout of kalirin-7 leads to increased rates of
cocaine self-administration. Mol Pharmacol 84:582590.
Kirkpatrick MG, Gunderson EW, Johanson CE, Levin FR, Foltin RW, and Hart CL
(2012) Comparison of intranasal methamphetamine and d-amphetamine selfadministration by humans. Addiction 107:783791.
Kish SJ, Lerch J, Furukawa Y, Tong J, McCluskey T, Wilkins D, Houle S, Meyer J,
Mundo E, and Wilson AA, et al. (2010) Decreased cerebral cortical serotonin
transporter binding in ecstasy users: a positron emission tomography/[(11)C]DASB
and structural brain imaging study. Brain 133:17791797.
Kiss JP, Tth E, Lajtha A, and Vizi ES (1994) NMDA receptors are not involved in
the MK-801-induced increase of striatal dopamine release in rat: a microdialysis
study. Brain Res 641:145148.
Kiss T, Hoffmann WE, Scott L, Kawabe TT, Milici AJ, Nilsen EA, and Hajs M (2011)
Role of thalamic projection in NMDA receptor-induced disruption of cortical slow
oscillation and short-term plasticity. Front Psychiatry 2:14. DOI 10.3389/
fpsyt.2011.00014
Kissler JL, Sirohi S, Reis DJ, Jansen HT, Quock RM, Smith DG, and Walker BM
(2014) The one-two punch of alcoholism: role of central amygdala dynorphins/
kappa-opioid receptors. Biol Psychiatry 75:774782.
Kitamura O, Wee S, Specio SE, Koob GF, and Pulvirenti L (2006) Escalation of
methamphetamine self-administration in rats: a dose-effect function. Psychopharmacology (Berl) 186:4853.
Klausberger T and Somogyi P (2008) Neuronal diversity and temporal dynamics: the
unity of hippocampal circuit operations. Science 321:5357.
Klebaur JE, Ostrander MM, Norton CS, Watson SJ, Akil H, and Robinson TE (2002)
The ability of amphetamine to evoke arc (Arg 3.1) mRNA expression in the caudate, nucleus accumbens and neocortex is modulated by environmental context.
Brain Res 930:3036.
Klein C, Kemmel V, Taleb O, Aunis D, and Maitre M (2009) Pharmacological doses of
gamma-hydroxybutyrate (GHB) potentiate histone acetylation in the rat brain by
histone deacetylase inhibition. Neuropharmacology 57:137147.
Kleven MS, Woolverton WL, and Seiden LS (1988) Lack of long-term monoamine
depletions following repeated or continuous exposure to cocaine. Brain Res Bull 21:
233237.
Klink R, de Kerchove dExaerde A, Zoli M, and Changeux JP (2001) Molecular and
physiological diversity of nicotinic acetylcholine receptors in the midbrain dopaminergic nuclei. J Neurosci 21:14521463.
Klinkenberg I and Blokland A (2010) The validity of scopolamine as a pharmacological model for cognitive impairment: a review of animal behavioral studies. Neurosci Biobehav Rev 34:13071350.
Klongpanichapak S, Govitrapong P, Sharma SK, and Ebadi M (2006) Attenuation of
cocaine and methamphetamine neurotoxicity by coenzyme Q10. Neurochem Res 31:
303311.
Klotz U (1991) Occurrence of natural benzodiazepines. Life Sci 48:209215.
Klotz U, Ziegler G, and Reimann IW (1984) Pharmacokinetics of the selective benzodiazepine antagonist Ro 15-1788 in man. Eur J Clin Pharmacol 27:115117.
Knabl J, Witschi R, Hsl K, Reinold H, Zeilhofer UB, Ahmadi S, Brockhaus J,
Sergejeva M, Hess A, and Brune K, et al. (2008) Reversal of pathological pain
through specific spinal GABAA receptor subtypes. Nature 451:330334.
Knackstedt LA, Melendez RI, and Kalivas PW (2010) Ceftriaxone restores glutamate
homeostasis and prevents relapse to cocaine seeking. Biol Psychiatry 67:8184.
Knickmeyer RC, Gouttard S, Kang C, Evans D, Wilber K, Smith JK, Hamer RM, Lin
W, Gerig G, and Gilmore JH (2008) A structural MRI study of human brain development from birth to 2 years. J Neurosci 28:1217612182.
Kobayashi T, Ikeda K, Kojima H, Niki H, Yano R, Yoshioka T, and Kumanishi T
(1999) Ethanol opens G-protein-activated inwardly rectifying K+ channels. Nat
Neurosci 2:10911097.
Kogan FJ, Nichols WK, and Gibb JW (1976) Influence of methamphetamine on nigral
and striatal tyrosine hydroxylase activity and on striatal dopamine levels. Eur J
Pharmacol 36:363371.
Koike H and Chaki S (2014) Requirement of AMPA receptor stimulation for the
sustained antidepressant activity of ketamine and LY341495 during the forced
swim test in rats. Behav Brain Res 271:111115.
Koike H, Iijima M, and Chaki S (2011) Involvement of AMPA receptor in both the
rapid and sustained antidepressant-like effects of ketamine in animal models of
depression. Behav Brain Res 224:107111.
Kokoshka JM, Vaughan RA, Hanson GR, and Fleckenstein AE (1998) Nature of
methamphetamine-induced rapid and reversible changes in dopamine transporters. Eur J Pharmacol 361:269275.
Kolb B and Gibb R (2015) Plasticity in the prefrontal cortex of adult rats. Front Cell
Neurosci 9:15. DOI 10.3389/fncel.2015.00015

985

Kometer M, Schmidt A, Jncke L, and Vollenweider FX (2013) Activation of serotonin


2A receptors underlies the psilocybin-induced effects on a oscillations, N170 visualevoked potentials, and visual hallucinations. J Neurosci 33:1054410551.
Kontkanen O, Lakso M, Koponen E, Wong G, and Castrn E (2000) Molecular effects
of the psychotropic NMDA receptor antagonist MK-801 in the rat entorhinal cortex: increases in AP-1 DNA binding activity and expression of Fos and Jun family
members. Ann N Y Acad Sci 911:7382.
Koob GF (1996) Hedonic valence, dopamine and motivation. Mol Psychiatry 1:
186189.
Koob GF (2008) A role for brain stress systems in addiction. Neuron 59:1134.
Koob GF and Le Moal M (2008) Addiction and the brain antireward system. Annu
Rev Psychol 59:2953.
Koob GF, Roberts AJ, Kieffer BL, Heyser CJ, Katner SN, Ciccocioppo R, and Weiss F
(2003) Animal models of motivation for drinking in rodents with a focus on opioid
receptor neuropharmacology. Recent Dev Alcohol 16:263281.
Koob GF and Volkow ND (2010) Neurocircuitry of addiction. Neuropsychopharmacology 35:217238.
Koob GF and Zorrilla EP (2010) Neurobiological mechanisms of addiction: focus on
corticotropin-releasing factor. Curr Opin Investig Drugs 11:6371.
Kornetsky C and Esposito RU (1979) Euphorigenic drugs: effects on the reward
pathways of the brain. Fed Proc 38:24732476.
Korotkova TM, Sergeeva OA, Eriksson KS, Haas HL, and Brown RE (2003) Excitation of ventral tegmental area dopaminergic and nondopaminergic neurons by
orexins/hypocretins. J Neurosci 23:711.
Korpi ER, Debus F, Linden AM, Malcot C, Lepp E, Vekovischeva O, Rabe H, Bhme I,
Aller MI, and Wisden W, et al. (2007) Does ethanol act preferentially via selected brain
GABAA receptor subtypes? the current evidence is ambiguous. Alcohol 41:163176.
Korpi ER, Grnder G, and Lddens H (2002) Drug interactions at GABA(A) receptors. Prog Neurobiol 67:113159.
Korpi ER, Kleingoor C, Kettenmann H, and Seeburg PH (1993) Benzodiazepineinduced motor impairment linked to point mutation in cerebellar GABAA receptor.
Nature 361:356359.
Kota D, Robinson SE, and Imad Damaj M (2009) Enhanced nicotine reward in
adulthood after exposure to nicotine during early adolescence in mice. Biochem
Pharmacol 78:873879.
Kourrich S, Rothwell PE, Klug JR, and Thomas MJ (2007) Cocaine experience controls bidirectional synaptic plasticity in the nucleus accumbens. J Neurosci 27:
79217928.
Koya E, Cruz FC, Ator R, Golden SA, Hoffman AF, Lupica CR, and Hope BT (2012)
Silent synapses in selectively activated nucleus accumbens neurons following cocaine sensitization. Nat Neurosci 15:15561562.
Koya E, Golden SA, Harvey BK, Guez-Barber DH, Berkow A, Simmons DE, Bossert
JM, Nair SG, Uejima JL, and Marin MT, et al. (2009) Targeted disruption of
cocaine-activated nucleus accumbens neurons prevents context-specific sensitization. Nat Neurosci 12:10691073.
Koyrakh L, Lujn R, Coln J, Karschin C, Kurachi Y, Karschin A, and Wickman K
(2005) Molecular and cellular diversity of neuronal G-protein-gated potassium
channels. J Neurosci 25:1146811478.
Kranzler HR, Covault J, Feinn R, Armeli S, Tennen H, Arias AJ, Gelernter J, Pond T,
Oncken C, and Kampman KM (2014) Topiramate treatment for heavy drinkers:
moderation by a GRIK1 polymorphism. Am J Psychiatry 171:445452.
Krasnova IN, Ladenheim B, and Cadet JL (2005) Amphetamine induces apoptosis of
medium spiny striatal projection neurons via the mitochondria-dependent pathway. FASEB J 19:851853.
Krebs MO, Morvan Y, Jay T, Gaillard R, and Kebir O (2014) Psychotomimetic effects
at initiation of cannabis use are associated with cannabinoid receptor 1 (CNR1)
variants in healthy students. Mol Psychiatry 19:402403.
Krebs TS and Johansen PO (2008) No evidence of decrease in cognitive function in
users of low-dose Ecstasy. Arch Gen Psychiatry 65:236237.
Krebs TS and Johansen PO (2012) Lysergic acid diethylamide (LSD) for alcoholism:
meta-analysis of randomized controlled trials. J Psychopharmacol 26:9941002.
Krebs TS and Johansen PO (2013) Psychedelics and mental health: a population
study. PLoS One 8:e63972.
Kreek MJ, Nielsen DA, Butelman ER, and LaForge KS (2005) Genetic influences on
impulsivity, risk taking, stress responsivity and vulnerability to drug abuse and
addiction. Nat Neurosci 8:14501457.
Kreifeldt M, Le D, Treistman SN, Koob GF, and Contet C (2013) BK channel b1 and
b4 auxiliary subunits exert opposite influences on escalated ethanol drinking in
dependent mice. Front Integr Neurosci 7:105. DOI 10.3389/fnint.2013.00105
Kreitzer AC and Regehr WG (2001) Retrograde inhibition of presynaptic calcium
influx by endogenous cannabinoids at excitatory synapses onto Purkinje cells.
Neuron 29:717727.
Kreshak AA, Cantrell FL, Clark RF, and Tomaszewski CA (2012) A poison centers
ten-year experience with flumazenil administration to acutely poisoned adults.
J Emerg Med 43:677682.
Kretschmer BD (2000) NMDA receptor antagonist-induced dopamine release in the
ventral pallidum does not correlate with motor activation. Brain Res 859:147156.
Kroener S, Mulholland PJ, New NN, Gass JT, Becker HC, and Chandler LJ (2012)
Chronic alcohol exposure alters behavioral and synaptic plasticity of the rodent
prefrontal cortex. PLoS One 7:e37541.
Krupitsky E, Burakov A, Romanova T, Dunaevsky I, Strassman R, and Grinenko A
(2002) Ketamine psychotherapy for heroin addiction: immediate effects and twoyear follow-up. J Subst Abuse Treat 23:273283.
Krystal JH, Sanacora G, and Duman RS (2013) Rapid-acting glutamatergic antidepressants: the path to ketamine and beyond. Biol Psychiatry 73:11331141.
Ksir C, Hakan R, Hall DP Jr, and Kellar KJ (1985) Exposure to nicotine enhances the
behavioral stimulant effect of nicotine and increases binding of [3H]acetylcholine to
nicotinic receptors. Neuropharmacology 24:527531.

986

Korpi et al.

Kuczenski R (1975) Effects of catecholamine releasing agents on synaptosomal dopamine biosynthesis: multiple pools of dopamine or multiple forms of tyrosine
hydroxylase. Neuropharmacology 14:110.
Kuczenski R and Segal DS (1997) Effects of methylphenidate on extracellular dopamine, serotonin, and norepinephrine: comparison with amphetamine. J Neurochem 68:20322037.
Kuczewski N, Porcher C, and Gaiarsa JL (2010) Activity-dependent dendritic secretion of brain-derived neurotrophic factor modulates synaptic plasticity. Eur J
Neurosci 32:12391244.
Kufahl P, Li Z, Risinger R, Rainey C, Piacentine L, Wu G, Bloom A, Yang Z, and Li SJ
(2008) Expectation modulates human brain responses to acute cocaine: a functional
magnetic resonance imaging study. Biol Psychiatry 63:222230.
Kufahl PR, Hood LE, Nemirovsky NE, Barabas P, Halstengard C, Villa A, Moore E,
Watterson LR, and Olive MF (2012) Positive allosteric modulation of mglur5
accelerates extinction learning but not relearning following methamphetamine
self-administration. Front Pharmacol 3:194. DOI 10.3389/fphar.2012.00194
Kufahl PR, Watterson LR, Nemirovsky NE, Hood LE, Villa A, Halstengard C, Zautra
N, and Olive MF (2013) Attenuation of methamphetamine seeking by the mGluR2/
3 agonist LY379268 in rats with histories of restricted and escalated selfadministration. Neuropharmacology 66:290301.
Kuhlman SJ, Lu J, Lazarus MS, and Huang ZJ (2010) Maturation of GABAergic
inhibition promotes strengthening of temporally coherent inputs among convergent pathways. PLOS Comput Biol 6:e1000797.
Kullmann DM and Lamsa KP (2007) Long-term synaptic plasticity in hippocampal
interneurons. Nat Rev Neurosci 8:687699.
Kullmann DM and Lamsa KP (2011) LTP and LTD in cortical GABAergic interneurons: emerging rules and roles. Neuropharmacology 60:712719.
Kumar A, Choi KH, Renthal W, Tsankova NM, Theobald DE, Truong HT, Russo SJ,
Laplant Q, Sasaki TS, and Whistler KN, et al. (2005) Chromatin remodeling is
a key mechanism underlying cocaine-induced plasticity in striatum. Neuron 48:
303314.
Kumar S, Porcu P, Werner DF, Matthews DB, Diaz-Granados JL, Helfand RS,
and Morrow AL (2009) The role of GABA(A) receptors in the acute and chronic
effects of ethanol: a decade of progress. Psychopharmacology (Berl) 205:529564.
Kuner T, Schoepfer R, and Korpi ER (1993) Ethanol inhibits glutamate-induced
currents in heteromeric NMDA receptor subtypes. Neuroreport 5:297300.
Kunitachi S, Fujita Y, Ishima T, Kohno M, Horio M, Tanibuchi Y, Shirayama Y, Iyo
M, and Hashimoto K (2009) Phencyclidine-induced cognitive deficits in mice are
ameliorated by subsequent subchronic administration of donepezil: role of sigma-1
receptors. Brain Res 1279:189196.
Kuryatov A, Luo J, Cooper J, and Lindstrom J (2005) Nicotine acts as a pharmacological chaperone to up-regulate human alpha4beta2 acetylcholine receptors. Mol
Pharmacol 68:18391851.
Kuschinsky W, Suda S, and Sokoloff L (1985) Influence of gamma-hydroxybutyrate
on the relationship between local cerebral glucose utilization and local cerebral
blood flow in the rat brain. J Cereb Blood Flow Metab 5:5864.
Kushner SA, Dewey SL, and Kornetsky C (1997) Gamma-vinyl GABA attenuates
cocaine-induced lowering of brain stimulation reward thresholds. Psychopharmacology (Berl) 133:383388.
Kushner SA, Dewey SL, and Kornetsky C (1999) The irreversible gammaaminobutyric acid (GABA) transaminase inhibitor gamma-vinyl-GABA blocks cocaine self-administration in rats. J Pharmacol Exp Ther 290:797802.
Kwilasz AJ and Negus SS (2012) Dissociable effects of the cannabinoid receptor
agonists D9-tetrahydrocannabinol and CP55940 on pain-stimulated versus paindepressed behavior in rats. J Pharmacol Exp Ther 343:389400.
Kwok JC, Dick G, Wang D, and Fawcett JW (2011) Extracellular matrix and perineuronal nets in CNS repair. Dev Neurobiol 71:10731089.
Kyosseva SV, Owens SM, Elbein AD, and Karson CN (2001) Differential and regionspecific activation of mitogen-activated protein kinases following chronic administration of phencyclidine in rat brain. Neuropsychopharmacology 24:267277.
Laborit H (1964) Sodium 4-Hydroxybutyrate. Int J Neuropharmacol 3:433451.
Laboube G, Lomazzi M, Cruz HG, Creton C, Lujn R, Li M, Yanagawa Y, Obata K,
Watanabe M, and Wickman K, et al. (2007) RGS2 modulates coupling between
GABAB receptors and GIRK channels in dopamine neurons of the ventral tegmental area. Nat Neurosci 10:15591568.
Lck AK, Ariwodola OJ, Chappell AM, Weiner JL, and McCool BA (2008) Ethanol
inhibition of kainate receptor-mediated excitatory neurotransmission in the rat
basolateral nucleus of the amygdala. Neuropharmacology 55:661668.
Lck AK, Diaz MR, Chappell A, DuBois DW, and McCool BA (2007) Chronic ethanol
and withdrawal differentially modulate pre- and postsynaptic function at glutamatergic synapses in rat basolateral amygdala. J Neurophysiol 98:31853196.
Ladenheim B, Krasnova IN, Deng X, Oyler JM, Polettini A, Moran TH, Huestis MA,
and Cadet JL (2000) Methamphetamine-induced neurotoxicity is attenuated in
transgenic mice with a null mutation for interleukin-6. Mol Pharmacol 58:12471256.
Lader MH and Morton SV (1992) A pilot study of the effects of flumazenil on
symptoms persisting after benzodiazepine withdrawal. J Psychopharmacol 6:
357363.
Lai A, Parameswaran N, Khwaja M, Whiteaker P, Lindstrom JM, Fan H, McIntosh
JM, Grady SR, and Quik M (2005) Long-term nicotine treatment decreases striatal
alpha 6* nicotinic acetylcholine receptor sites and function in mice. Mol Pharmacol
67:16391647.
Lammel S, Ion DI, Roeper J, and Malenka RC (2011) Projection-specific modulation
of dopamine neuron synapses by aversive and rewarding stimuli. Neuron 70:
855862.
Lammel S, Lim BK, Ran C, Huang KW, Betley MJ, Tye KM, Deisseroth K,
and Malenka RC (2012) Input-specific control of reward and aversion in the ventral
tegmental area. Nature 491:212217.
Lamsa KP, Heeroma JH, Somogyi P, Rusakov DA, and Kullmann DM (2007) AntiHebbian long-term potentiation in the hippocampal feedback inhibitory circuit.
Science 315:12621266.

Langer G, Karazman R, Neumark J, Saletu B, Schnbeck G, Grnberger J, Dittrich


R, Petricek W, Hoffmann P, and Linzmayer L, et al. (1995) Isoflurane narcotherapy
in depressive patients refractory to conventional antidepressant drug treatment. A
double-blind comparison with electroconvulsive treatment. Neuropsychobiology 31:
182194.
Langer G, Neumark J, Koinig G, Graf M, and Schnbeck G (1985) Rapid psychotherapeutic effects of anesthesia with isoflurane (ES narcotherapy) in treatmentrefractory depressed patients. Neuropsychobiology 14:118120.
Lansink CS, Goltstein PM, Lankelma JV, Joosten RN, McNaughton BL,
and Pennartz CM (2008) Preferential reactivation of motivationally relevant information in the ventral striatum. J Neurosci 28:63726382.
Lansink CS, Goltstein PM, Lankelma JV, McNaughton BL, and Pennartz CM (2009)
Hippocampus leads ventral striatum in replay of place-reward information. PLoS
Biol 7:e1000173.
Laplante M and Sabatini DM (2009) mTOR signaling at a glance. J Cell Sci 122:
35893594.
Lapray D, Lasztoczi B, Lagler M, Viney TJ, Katona L, Valenti O, Hartwich K,
Borhegyi Z, Somogyi P, and Klausberger T (2012) Behavior-dependent specialization of identified hippocampal interneurons. Nat Neurosci 15:12651271.
LaRowe SD, Kalivas PW, Nicholas JS, Randall PK, Mardikian PN, and Malcolm RJ
(2013) A double-blind placebo-controlled trial of N-acetylcysteine in the treatment
of cocaine dependence. Am J Addict 22:443452.
Larson EB, Graham DL, Arzaga RR, Buzin N, Webb J, Green TA, Bass CE, Neve RL,
Terwilliger EF, and Nestler EJ, et al. (2011) Overexpression of CREB in the nucleus accumbens shell increases cocaine reinforcement in self-administering rats. J
Neurosci 31:1644716457.
Lasseter HC, Wells AM, Xie X, and Fuchs RA (2011) Interaction of the basolateral
amygdala and orbitofrontal cortex is critical for drug context-induced reinstatement of cocaine-seeking behavior in rats. Neuropsychopharmacology 36:
711720.
Lasseter HC, Xie X, Arguello AA, Wells AM, Hodges MA, and Fuchs RA (2014)
Contribution of a mesocorticolimbic subcircuit to drug context-induced reinstatement of cocaine-seeking behavior in rats. Neuropsychopharmacology 39:
660669.
Laviolette SR, Gallegos RA, Henriksen SJ, and van der Kooy D (2004) Opiate state
controls bi-directional reward signaling via GABAA receptors in the ventral tegmental area. Nat Neurosci 7:160169.
LaVoie MJ and Hastings TG (1999) Dopamine quinone formation and protein modification associated with the striatal neurotoxicity of methamphetamine: evidence
against a role for extracellular dopamine. J Neurosci 19:14841491.
Lawrence AJ, Cowen MS, Yang HJ, Chen F, and Oldfield B (2006) The orexin system
regulates alcohol-seeking in rats. Br J Pharmacol 148:752759.
Lawrence RC, Otero NK, and Kelly SJ (2012) Selective effects of perinatal ethanol
exposure in medial prefrontal cortex and nucleus accumbens. Neurotoxicol Teratol
34:128135.
Layer RT, Kaddis FG, and Wallace LJ (1993) The NMDA receptor antagonist MK801 elicits conditioned place preference in rats. Pharmacol Biochem Behav 44:
245247.
Lazenka MF, Selley DE, and Sim-Selley LJ (2014) DFosB induction correlates inversely with CB receptor desensitization in a brain region-dependent manner
following repeated D-THC administration. Neuropharmacology 77:224233.
Lazenka MF, Tomarchio A, Lichtman AH, Greengard P, Flajolet M, Selley DE,
and Sim-Selley LJ (2015) Role of dopamine type 1 receptors and DARPP-32 in
Delta9-THC-mediated induction of DeltaFosB in the mouse forebrain. J Pharmacol
Exp Ther 354:316327.
Le Cozannet R, Fone KC, and Moran PM (2010) Phencyclidine withdrawal disrupts
episodic-like memory in rats: reversal by donepezil but not clozapine. Int J Neuropsychopharmacol 13:10111020.
Le Garrec S, Dauger S, and Sachs P (2014) Cannabis poisoning in children. Intensive
Care Med 40:13941395.
Lecca S, Melis M, Luchicchi A, Ennas MG, Castelli MP, Muntoni AL, and Pistis M
(2011) Effects of drugs of abuse on putative rostromedial tegmental neurons, inhibitory afferents to midbrain dopamine cells. Neuropsychopharmacology 36:
589602.
Lecca S, Meye FJ, and Mameli M (2014) The lateral habenula in addiction and
depression: an anatomical, synaptic and behavioral overview. Eur J Neurosci 39:
11701178.
Ledent C, Valverde O, Cossu G, Petitet F, Aubert JF, Beslot F, Bhme GA, Imperato
A, Pedrazzini T, and Roques BP, et al. (1999) Unresponsiveness to cannabinoids
and reduced addictive effects of opiates in CB1 receptor knockout mice. Science
283:401404.
Lee AM, Kanter BR, Wang D, Lim JP, Zou ME, Qiu C, McMahon T, Dadgar J,
Fischbach-Weiss SC, and Messing RO (2013a) Prkcz null mice show normal
learning and memory. Nature 493:416419.
Lee BR and Dong Y (2011) Cocaine-induced metaplasticity in the nucleus accumbens:
silent synapse and beyond. Neuropharmacology 61:10601069.
Lee BR, Ma YY, Huang YH, Wang X, Otaka M, Ishikawa M, Neumann PA, Graziane
NM, Brown TE, and Suska A, et al. (2013b) Maturation of silent synapses in
amygdala-accumbens projection contributes to incubation of cocaine craving. Nat
Neurosci 16:16441651.
Lee EH and Geyer MA (1980) Persistent effects of chronic administration of LSD on
intracellular serotonin content in rat midbrain. Neuropharmacology 19:10051007.
Lee H, Kang MS, Chung JM, and Noh J (2015) Repeated nicotine exposure in adolescent rats: Reduction of medial habenular activity and augmentation of nicotine
preference. Physiol Behav 138:345350.
Lee HK, Takamiya K, Han JS, Man H, Kim CH, Rumbaugh G, Yu S, Ding L, He C,
and Petralia RS, et al. (2003) Phosphorylation of the AMPA receptor GluR1 subunit is required for synaptic plasticity and retention of spatial memory. Cell 112:
631643.

Drug-Induced Neuroplasticity
Lee JH (2011) Tracing activity across the whole brain neural network with optogenetic functional magnetic resonance imaging. Front Neuroinform 5:21.
Lee JS, Kim BN, Kang E, Lee DS, Kim YK, Chung JK, Lee MC, and Cho SC (2005)
Regional cerebral blood flow in children with attention deficit hyperactivity disorder: comparison before and after methylphenidate treatment. Hum Brain Mapp
24:157164.
Lee KS, Frank S, Vanderklish P, Arai A, and Lynch G (1991) Inhibition of proteolysis
protects hippocampal neurons from ischemia. Proc Natl Acad Sci USA 88:
72337237.
Lee KW, Kim Y, Kim AM, Helmin K, Nairn AC, and Greengard P (2006) Cocaineinduced dendritic spine formation in D1 and D2 dopamine receptor-containing
medium spiny neurons in nucleus accumbens. Proc Natl Acad Sci USA 103:
33993404.
Lee M, Schwab C, and McGeer PL (2011) Astrocytes are GABAergic cells that
modulate microglial activity. Glia 59:152165.
Lee S, Yoon BE, Berglund K, Oh SJ, Park H, Shin HS, Augustine GJ, and Lee CJ
(2010) Channel-mediated tonic GABA release from glia. Science 330:790796.
Lehrmann E, Colantuoni C, Deep-Soboslay A, Becker KG, Lowe R, Huestis MA, Hyde
TM, Kleinman JE, and Freed WJ (2006) Transcriptional changes common to human cocaine, cannabis and phencyclidine abuse. PLoS One 1:e114.
Leone MA, Vigna-Taglianti F, Avanzi G, Brambilla R, and Faggiano F (2010)
Gamma-hydroxybutyrate (GHB) for treatment of alcohol withdrawal and prevention of relapses. Cochrane Database Syst Rev (2):CD006266.
Lepack AE, Fuchikami M, Dwyer JM, Banasr M, and Duman RS (2015) BDNF release is required for the behavioral actions of ketamine. Int J Neuropsychopharmacol 18:18. DOI 10.1093/ijnp/pyu033
Lepore M, Liu X, Savage V, Matalon D, and Gardner EL (1996) Genetic differences in
delta 9-tetrahydrocannabinol-induced facilitation of brain stimulation reward as
measured by a rate-frequency curve-shift electrical brain stimulation paradigm in
three different rat strains. Life Sci 58:PL365PL372.
Lepsch LB, Munhoz CD, Kawamoto EM, Yshii LM, Lima LS, Curi-Boaventura MF,
Salgado TM, Curi R, Planeta CS, and Scavone C (2009) Cocaine induces cell death
and activates the transcription nuclear factor kappa-B in PC12 cells. Mol Brain
2:3.
Leri F, Flores J, Rodaros D, and Stewart J (2002) Blockade of stress-induced but not
cocaine-induced reinstatement by infusion of noradrenergic antagonists into the
bed nucleus of the stria terminalis or the central nucleus of the amygdala.
J Neurosci 22:57135718.
Lerner TN, Horne EA, Stella N, and Kreitzer AC (2010) Endocannabinoid signaling
mediates psychomotor activation by adenosine A2A antagonists. J Neurosci 30:
21602164.
Leslie FM, Loughlin SE, Wang R, Perez L, Lotfipour S, and Belluzzia JD (2004)
Adolescent development of forebrain stimulant responsiveness: insights from animal studies. Ann N Y Acad Sci 1021:148159.
Levin ED (2013) Complex relationships of nicotinic receptor actions and cognitive
functions. Biochem Pharmacol 86:11451152.
Levin ED, Uemura E, and Bowman RE (1991) Neurobehavioral toxicology of halothane in rats. Neurotoxicol Teratol 13:461470.
Lewis MJ, Johnson DF, Waldman D, Leibowitz SF, and Hoebel BG (2004) Galanin
microinjection in the third ventricle increases voluntary ethanol intake. Alcohol
Clin Exp Res 28:18221828.
Lewohl JM, Wilson WR, Mayfield RD, Brozowski SJ, Morrisett RA, and Harris RA
(1999) G-protein-coupled inwardly rectifying potassium channels are targets of
alcohol action. Nat Neurosci 2:10841090.
Li B, Piriz J, Mirrione M, Chung C, Proulx CD, Schulz D, Henn F, and Malinow R
(2011) Synaptic potentiation onto habenula neurons in the learned helplessness
model of depression. Nature 470:535539.
Li C, Aguayo L, Peoples RW, and Weight FF (1993) Ethanol inhibits a neuronal ATPgated ion channel. Mol Pharmacol 44:871875.
Li C, Peoples RW, and Weight FF (1994) Alcohol action on a neuronal membrane
receptor: evidence for a direct interaction with the receptor protein. Proc Natl Acad
Sci USA 91:82008204.
Li J, Liu N, Lu K, Zhang L, Gu J, Guo F, An S, Zhang L, and Zhang L (2012) Cocaineinduced dendritic remodeling occurs in both D1 and D2 dopamine receptorexpressing neurons in the nucleus accumbens. Neurosci Lett 517:118122.
Li KY, Xiao C, Xiong M, Delphin E, and Ye JH (2008) Nanomolar propofol stimulates
glutamate transmission to dopamine neurons: a possible mechanism of abuse potential? J Pharmacol Exp Ther 325:165174.
Li N, Lee B, Liu RJ, Banasr M, Dwyer JM, Iwata M, Li XY, Aghajanian G,
and Duman RS (2010a) mTOR-dependent synapse formation underlies the rapid
antidepressant effects of NMDA antagonists. Science 329:959964.
Li P, Rudolph U, and Huntsman MM (2009) Long-term sensory deprivation selectively rearranges functional inhibitory circuits in mouse barrel cortex. Proc Natl
Acad Sci USA 106:1215612161.
Li Q, Clark S, Lewis DV, and Wilson WA (2002) NMDA receptor antagonists disinhibit rat posterior cingulate and retrosplenial cortices: a potential mechanism of
neurotoxicity. J Neurosci 22:30703080.
Li S and Carmichael ST (2006) Growth-associated gene and protein expression in the
region of axonal sprouting in the aged brain after stroke. Neurobiol Dis 23:
362373.
Li S, Overman JJ, Katsman D, Kozlov SV, Donnelly CJ, Twiss JL, Giger RJ, Coppola
G, Geschwind DH, and Carmichael ST (2010b) An age-related sprouting transcriptome provides molecular control of axonal sprouting after stroke. Nat Neurosci
13:14961504.
Liang J, Cagetti E, Olsen RW, and Spigelman I (2004) Altered pharmacology of
synaptic and extrasynaptic GABAA receptors on CA1 hippocampal neurons is
consistent with subunit changes in a model of alcohol withdrawal and dependence.
J Pharmacol Exp Ther 310:12341245.

987

Liang J, Zhang N, Cagetti E, Houser CR, Olsen RW, and Spigelman I (2006) Chronic
intermittent ethanol-induced switch of ethanol actions from extrasynaptic to synaptic hippocampal GABAA receptors. J Neurosci 26:17491758.
Licata SC and Pierce RC (2003) The roles of calcium/calmodulin-dependent and Ras/
mitogen-activated protein kinases in the development of psychostimulant-induced
behavioral sensitization. J Neurochem 85:1422.
Licata SC, Schmidt HD, and Pierce RC (2004) Suppressing calcium/calmodulindependent protein kinase II activity in the ventral tegmental area enhances the
acute behavioural response to cocaine but attenuates the initiation of cocaineinduced behavioural sensitization in rats. Eur J Neurosci 19:405414.
Licata SC, Shinday NM, Huizenga MN, Darnell SB, Sangrey GR, Rudolph U, Rowlett
JK, and Sadri-Vakili G (2013) Alterations in brain-derived neurotrophic factor in
the mouse hippocampus following acute but not repeated benzodiazepine treatment. PLoS One 8:e84806.
Liechti ME, Baumann C, Gamma A, and Vollenweider FX (2000) Acute psychological
effects of 3,4-methylenedioxymethamphetamine (MDMA, Ecstasy) are attenuated by the serotonin uptake inhibitor citalopram. Neuropsychopharmacology 22:
513521.
Liechti ME, Lhuillier L, Kaupmann K, and Markou A (2007) Metabotropic glutamate
2/3 receptors in the ventral tegmental area and the nucleus accumbens shell are
involved in behaviors relating to nicotine dependence. J Neurosci 27:90779085.
Lim J, Yang C, Hong SJ, and Kim KS (2000) Regulation of tyrosine hydroxylase gene
transcription by the cAMP-signaling pathway: involvement of multiple transcription factors. Mol Cell Biochem 212:5160.
Lin Y and Leskawa KC (1994) Cytotoxicity of the cocaine metabolite benzoylecgonine.
Brain Res 643:108114.
Lindahl JS, Kjellsen BR, Tigert J, and Miskimins R (2008) In utero PCP exposure
alters oligodendrocyte differentiation and myelination in developing rat frontal
cortex. Brain Res 1234:137147.
Lindn A, Storvik M, Lakso M, Haapasalo A, Lee D, Witkin JM, Sei Y, Castrn E,
and Wong G (2001) Increased expression of neuronal Src and tyrosine phosphorylation of NMDA receptors in rat brain after systemic treatment with MK-801.
Neuropharmacology 40:469481.
Lindn AM, Visnen J, Lakso M, Nawa H, Wong G, and Castrn E (2000) Expression of neurotrophins BDNF and NT-3, and their receptors in rat brain after
administration of antipsychotic and psychotrophic agents. J Mol Neurosci 14:
2737.
Lindholm S, Ploj K, Franck J, and Nylander I (2000) Repeated ethanol administration induces short- and long-term changes in enkephalin and dynorphin tissue
concentrations in rat brain. Alcohol 22:165171.
Ling DS, Benardo LS, and Sacktor TC (2006) Protein kinase Mzeta enhances excitatory synaptic transmission by increasing the number of active postsynaptic AMPA
receptors. Hippocampus 16:443452.
Lips EH, Gaborieau V, McKay JD, Chabrier A, Hung RJ, Boffetta P, Hashibe M,
Zaridze D, Szeszenia-Dabrowska N, and Lissowska J, et al. (2010) Association
between a 15q25 gene variant, smoking quantity and tobacco-related cancers
among 17 000 individuals. Int J Epidemiol 39:563577.
Lisman J, Schulman H, and Cline H (2002) The molecular basis of CaMKII function
in synaptic and behavioural memory. Nat Rev Neurosci 3:175190.
Littleton J (1998) Neurochemical mechanisms underlying alcohol withdrawal. Alcohol Health Res World 22:1324.
Liu D, Diorio J, Tannenbaum B, Caldji C, Francis D, Freedman A, Sharma S,
Pearson D, Plotsky PM, and Meaney MJ (1997) Maternal care, hippocampal glucocorticoid receptors, and hypothalamic-pituitary-adrenal responses to stress.
Science 277:16591662.
Liu F, Paule MG, Ali S, and Wang C (2011) Ketamine-induced neurotoxicity and
changes in gene expression in the developing rat brain. Curr Neuropharmacol 9:
256261.
Liu J, Asuncion-Chin M, Liu P, and Dopico AM (2006) CaM kinase II phosphorylation
of slo Thr107 regulates activity and ethanol responses of BK channels. Nat Neurosci 9:4149.
Liu JZ, Tozzi F, Waterworth DM, Pillai SG, Muglia P, Middleton L, Berrettini W,
Knouff CW, Yuan X, and Waeber G, et al.; Wellcome Trust Case Control Consortium (2010) Meta-analysis and imputation refines the association of 15q25 with
smoking quantity. Nat Genet 42:436440.
Liu S, Li L, Shen W, Shen X, Yang G, and Zhou W (2013) Scopolamine detoxification
technique for heroin dependence: a randomized trial. CNS Drugs 27:10931102.
Livingstone PD and Wonnacott S (2009) Nicotinic acetylcholine receptors and the
ascending dopamine pathways. Biochem Pharmacol 78:744755.
Lobo IA and Harris RA (2008) GABA(A) receptors and alcohol. Pharmacol Biochem
Behav 90:9094.
Lobo MK, Covington HE 3rd, Chaudhury D, Friedman AK, Sun H, Damez-Werno D,
Dietz DM, Zaman S, Koo JW, and Kennedy PJ, et al. (2010) Cell type-specific loss of
BDNF signaling mimics optogenetic control of cocaine reward. Science 330:
385390.
Lobo MK, Zaman S, Damez-Werno DM, Koo JW, Bagot RC, DiNieri JA, Nugent A,
Finkel E, Chaudhury D, and Chandra R, et al. (2013) DFosB induction in striatal
medium spiny neuron subtypes in response to chronic pharmacological, emotional,
and optogenetic stimuli. J Neurosci 33:1838118395.
Loe IM and Feldman HM (2007) Academic and educational outcomes of children with
ADHD. Ambul Pediatr 7:8290.
Logan BJ, Laverty R, Sanderson WD, and Yee YB (1988) Differences between rats
and mice in MDMA (methylenedioxymethylamphetamine) neurotoxicity. Eur J
Pharmacol 152:227234.
Logan BK (2002) Methamphetamine-effects on human performance and behavior.
Forensic Sci Rev 14:133151.
Logrip ML, Janak PH, and Ron D (2008) Dynorphin is a downstream effector of
striatal BDNF regulation of ethanol intake. FASEB J 22:23932404.
Lombardo S and Maskos U (2015) Role of the nicotinic acetylcholine receptor in
Alzheimers disease pathology and treatment. Neuropharmacology 96:255262.

988

Korpi et al.

Lominac KD, Sacramento AD, Szumlinski KK, and Kippin TE (2012) Distinct neurochemical adaptations within the nucleus accumbens produced by a history of selfadministered vs non-contingently administered intravenous methamphetamine.
Neuropsychopharmacology 37:707722.
London ED, Berman SM, Voytek B, Simon SL, Mandelkern MA, Monterosso J,
Thompson PM, Brody AL, Geaga JA, and Hong MS, et al. (2005) Cerebral metabolic dysfunction and impaired vigilance in recently abstinent methamphetamine
abusers. Biol Psychiatry 58:770778.
London ED, Cascella NG, Wong DF, Phillips RL, Dannals RF, Links JM, Herning R,
Grayson R, Jaffe JH, and Wagner HN Jr (1990) Cocaine-induced reduction of
glucose utilization in human brain. A study using positron emission tomography
and [fluorine 18]-fluorodeoxyglucose. Arch Gen Psychiatry 47:567574.
London ED, Simon SL, Berman SM, Mandelkern MA, Lichtman AM, Bramen J,
Shinn AK, Miotto K, Learn J, and Dong Y, et al. (2004) Mood disturbances and
regional cerebral metabolic abnormalities in recently abstinent methamphetamine
abusers. Arch Gen Psychiatry 61:7384.
Lonze BE and Ginty DD (2002) Function and regulation of CREB family transcription factors in the nervous system. Neuron 35:605623.
Lpez-Arnau R, Martnez-Clemente J, Pubill D, Escubedo E, and Camarasa J (2012)
Comparative neuropharmacology of three psychostimulant cathinone derivatives:
butylone, mephedrone and methylone. Br J Pharmacol 167:407420.
Lpez de Armentia M and Sah P (2007) Bidirectional synaptic plasticity at nociceptive afferents in the rat central amygdala. J Physiol 581:961970.
Lorenzetti V, Solowij N, Whittle S, Fornito A, Lubman DI, Pantelis C, and Yucel M
(2015) Gross morphological brain changes with chronic, heavy cannabis use. Br J
Psychiatry 206:7778.
Lorez H (1981) Fluorescence histochemistry indicates damage of striatal dopamine
nerve terminals in rats after multiple doses of methamphetamine. Life Sci 28:
911916.
Lscher W and Rundfeldt C (1990) Development of tolerance to clobazam in fully
kindled rats: effects of intermittent flumazenil administration. Eur J Pharmacol
180:255271.
Lovinger DM (1993) High ethanol sensitivity of recombinant AMPA-type glutamate
receptors expressed in mammalian cells. Neurosci Lett 159:8387.
Lovinger DM (1997) Alcohols and neurotransmitter gated ion channels: past, present
and future. Naunyn Schmiedebergs Arch Pharmacol 356:267282.
Lovinger DM and Homanics GE (2007) Tonic for what ails us? high-affinity GABAA
receptors and alcohol. Alcohol 41:139143.
Lovinger DM and Roberto M (2013) Synaptic effects induced by alcohol. Curr Top
Behav Neurosci 13:3186.
Lovinger DM and White G (1991) Ethanol potentiation of 5-hydroxytryptamine3
receptor-mediated ion current in neuroblastoma cells and isolated adult mammalian neurons. Mol Pharmacol 40:263270.
Lovinger DM, White G, and Weight FF (1989) Ethanol inhibits NMDA-activated ion
current in hippocampal neurons. Science 243:17211724.
Lw K, Crestani F, Keist R, Benke D, Brnig I, Benson JA, Fritschy JM, Rlicke T,
Bluethmann H, and Mhler H, et al. (2000) Molecular and neuronal substrate for
the selective attenuation of anxiety. Science 290:131134.
Low SJ and Roland CL (2004) Review of NMDA antagonist-induced neurotoxicity
and implications for clinical development. Int J Clin Pharmacol Ther 42:114.
Loweth JA, Baker LK, Guptaa T, Guillory AM, and Vezina P (2008) Inhibition of
CaMKII in the nucleus accumbens shell decreases enhanced amphetamine intake
in sensitized rats. Neurosci Lett 444:157160.
Loweth JA, Scheyer AF, Milovanovic M, LaCrosse AL, Flores-Barrera E, Werner CT,
Li X, Ford KA, Le T, and Olive MF, et al. (2014) Synaptic depression via mGluR1
positive allosteric modulation suppresses cue-induced cocaine craving. Nat Neurosci 17:7380.
Lu B, Pang PT, and Woo NH (2005) The yin and yang of neurotrophin action. Nat Rev
Neurosci 6:603614.
Lu G, Zhou QX, Kang S, Li QL, Zhao LC, Chen JD, Sun JF, Cao J, Wang YJ,
and Chen J, et al. (2010a) Chronic morphine treatment impaired hippocampal
long-term potentiation and spatial memory via accumulation of extracellular
adenosine acting on adenosine A1 receptors. J Neurosci 30:50585070.
Lu L, Grimm JW, Hope BT, and Shaham Y (2004) Incubation of cocaine craving after
withdrawal: a review of preclinical data. Neuropharmacology 47 (Suppl 1):214226.
Lu L, Huang M, Liu Z, and Ma L (2000) Cholecystokinin-B receptor antagonists
attenuate morphine dependence and withdrawal in rats. Neuroreport 11:829832.
Lu L, Mamiya T, Lu P, Toriumi K, Mouri A, Hiramatsu M, Kim HC, Zou LB, Nagai T,
and Nabeshima T (2010b) Prenatal exposure to phencyclidine produces abnormal
behaviour and NMDA receptor expression in postpubertal mice. Int J Neuropsychopharmacol 13:877889.
Lu L, Mamiya T, Lu P, Toriumi K, Mouri A, Hiramatsu M, Zou LB, and Nabeshima T
(2011) Prenatal exposure to PCP produces behavioral deficits accompanied by the
overexpression of GLAST in the prefrontal cortex of postpubertal mice. Behav
Brain Res 220:132139.
Luby ED, Cohen BD, Rosenbaum G, Gottlieb JS, and Kelley R (1959) Study of a new
schizophrenomimetic drug; sernyl. AMA Arch Neurol Psychiatry 81:363369.
Lucantonio F, Takahashi YK, Hoffman AF, Chang CY, Bali-Chaudhary S, Shaham Y,
Lupica CR, and Schoenbaum G (2014) Orbitofrontal activation restores insight lost
after cocaine use. Nat Neurosci 17:10921099.
Lddens H, Korpi ER, and Seeburg PH (1995) GABAA/benzodiazepine receptor
heterogeneity: neurophysiological implications. Neuropharmacology 34:245254.
Lddens H, Pritchett DB, Khler M, Killisch I, Keinnen K, Monyer H, Sprengel R,
and Seeburg PH (1990) Cerebellar GABAA receptor selective for a behavioural
alcohol antagonist. Nature 346:648651.
Luetje CM and Wooten J (1996) Clinical manifestations of transdermal scopolamine
addiction. Ear Nose Throat J 75:210214.
Lukas RJ, Changeux JP, Le Novre N, Albuquerque EX, Balfour DJ, Berg DK,
Bertrand D, Chiappinelli VA, Clarke PB, and Collins AC, et al. (1999)

International Union of Pharmacology. XX. Current status of the nomenclature for


nicotinic acetylcholine receptors and their subunits. Pharmacol Rev 51:397401.
Luo AH, Tahsili-Fahadan P, Wise RA, Lupica CR, and Aston-Jones G (2011) Linking
context with reward: a functional circuit from hippocampal CA3 to ventral tegmental area. Science 333:353357.
Lscher C (2013) Drug-evoked synaptic plasticity causing addictive behavior.
J Neurosci 33:1764117646.
Lscher C and Malenka RC (2011) Drug-evoked synaptic plasticity in addiction: from
molecular changes to circuit remodeling. Neuron 69:650663.
Lscher C and Malenka RC (2012) NMDA receptor-dependent long-term potentiation
and long-term depression (LTP/LTD). Cold Spring Harb Perspect Biol 4:4.
DOI 10.1101/cshperspect.a005710
Luthin GR and Tabakoff B (1984) Activation of adenylate cyclase by alcohols requires
the nucleotide-binding protein. J Pharmacol Exp Ther 228:579587.
Lutz PE and Kieffer BL (2013) The multiple facets of opioid receptor function:
implications for addiction. Curr Opin Neurobiol 23:473479.
Lyss PJ, Andersen SL, LeBlanc CJ, and Teicher MH (1999) Degree of neuronal
activation following FG-7142 changes across regions during development. Brain
Res Dev Brain Res 116:201203.
Ma YY, Guo CY, Yu P, Lee DY, Han JS, and Cui CL (2006) The role of NR2B
containing NMDA receptor in place preference conditioned with morphine and
natural reinforcers in rats. Exp Neurol 200:343355.
Ma YY, Lee BR, Wang X, Guo C, Liu L, Cui R, Lan Y, Balcita-Pedicino JJ, Wolf ME,
and Sesack SR, et al. (2014) Bidirectional modulation of incubation of cocaine
craving by silent synapse-based remodeling of prefrontal cortex to accumbens
projections. Neuron 83:14531467.
Maas LC, Lukas SE, Kaufman MJ, Weiss RD, Daniels SL, Rogers VW, Kukes TJ,
and Renshaw PF (1998) Functional magnetic resonance imaging of human brain
activation during cue-induced cocaine craving. Am J Psychiatry 155:124126.
Machu TK and Harris RA (1994) Alcohols and anesthetics enhance the function of
5-hydroxytryptamine3 receptors expressed in Xenopus laevis oocytes. J Pharmacol
Exp Ther 271:898905.
Mack F and Bnisch H (1979) Dissociation constants and lipophilicity of catecholamines and related compounds. Naunyn Schmiedebergs Arch Pharmacol 310:19.
MacLean KA, Johnson MW, and Griffiths RR (2011) Mystical experiences occasioned
by the hallucinogen psilocybin lead to increases in the personality domain of
openness. J Psychopharmacol 25:14531461.
Madden TE and Johnson SW (1998) Gamma-hydroxybutyrate is a GABAB receptor
agonist that increases a potassium conductance in rat ventral tegmental dopamine
neurons. J Pharmacol Exp Ther 287:261265.
Madhavan A, He L, Stuber GD, Bonci A, and Whistler JL (2010) micro-Opioid receptor endocytosis prevents adaptations in ventral tegmental area GABA transmission induced during naloxone-precipitated morphine withdrawal. J Neurosci
30:32763286.
Maguire MJ, Hemming K, Wild JM, Hutton JL, and Marson AG (2010) Prevalence of
visual field loss following exposure to vigabatrin therapy: a systematic review.
Epilepsia 51:24232431.
Maiese K (2014) Driving neural regeneration through the mammalian target of
rapamycin. Neural Regen Res 9:14131417.
Maitre M (1997) The gamma-hydroxybutyrate signalling system in brain: organization and functional implications. Prog Neurobiol 51:337361.
Malberg JE and Seiden LS (1998) Small changes in ambient temperature cause large
changes in 3,4-methylenedioxymethamphetamine (MDMA)-induced serotonin
neurotoxicity and core body temperature in the rat. J Neurosci 18:50865094.
Malenka RC and Bear MF (2004) LTP and LTD: an embarrassment of riches. Neuron
44:521.
Malfitano AM, Basu S, Maresz K, Bifulco M, and Dittel BN (2014) What we know and
do not know about the cannabinoid receptor 2 (CB2). Semin Immunol 26:369379.
Malinow R and Malenka RC (2002) AMPA receptor trafficking and synaptic plasticity. Annu Rev Neurosci 25:103126.
Malmgren K, Ben-Menachem E, and Frisn L (2001) Vigabatrin visual toxicity:
evolution and dose dependence. Epilepsia 42:609615.
Malminiemi O and Korpi ER (1989) Diazepam-insensitive [3H]Ro 15-4513 binding in
intact cultured cerebellar granule cells. Eur J Pharmacol 169:5360.
Mamelak M (2009) Narcolepsy and depression and the neurobiology of gammahydroxybutyrate. Prog Neurobiol 89:193219.
Mameli M, Bellone C, Brown MT, and Lscher C (2011) Cocaine inverts rules for
synaptic plasticity of glutamate transmission in the ventral tegmental area. Nat
Neurosci 14:414416.
Mameli M, Halbout B, Creton C, Engblom D, Parkitna JR, Spanagel R, and Lscher
C (2009) Cocaine-evoked synaptic plasticity: persistence in the VTA triggers
adaptations in the NAc. Nat Neurosci 12:10361041.
Mameli-Engvall M, Evrard A, Pons S, Maskos U, Svensson TH, Changeux JP,
and Faure P (2006) Hierarchical control of dopamine neuron-firing patterns by
nicotinic receptors. Neuron 50:911921.
Mammen AL, Kameyama K, Roche KW, and Huganir RL (1997) Phosphorylation of
the alpha-amino-3-hydroxy-5-methylisoxazole4-propionic acid receptor GluR1
subunit by calcium/calmodulin-dependent kinase II. J Biol Chem 272:
3252832533.
Manent JB, Jorquera I, Franco V, Ben-Ari Y, Perucca E, and Represa A (2008)
Antiepileptic drugs and brain maturation: fetal exposure to lamotrigine generates
cortical malformations in rats. Epilepsy Res 78:131139.
Manitt C, Eng C, Pokinko M, Ryan RT, Torres-Berro A, Lopez JP, Yogendran SV,
Daubaras MJ, Grant A, and Schmidt ER, et al. (2013) dcc orchestrates the development of the prefrontal cortex during adolescence and is altered in psychiatric
patients. Transl Psychiatry 3:e338.
Mann K, Bladstrm A, Torup L, Gual A, and van den Brink W (2013) Extending the
treatment options in alcohol dependence: a randomized controlled study of asneeded nalmefene. Biol Psychiatry 73:706713.

Drug-Induced Neuroplasticity
Mannuzza S, Klein RG, Truong NL, Moulton JL 3rd, Roizen ER, Howell KH,
and Castellanos FX (2008) Age of methylphenidate treatment initiation in children
with ADHD and later substance abuse: prospective follow-up into adulthood. Am J
Psychiatry 165:604609.
Manrique-Garcia E, Zammit S, Dalman C, Hemmingsson T, Andreasson S,
and Allebeck P (2012) Cannabis, schizophrenia and other non-affective psychoses:
35 years of follow-up of a population-based cohort. Psychol Med 42:13211328.
Mansour A, Fox CA, Akil H, and Watson SJ (1995) Opioid-receptor mRNA expression
in the rat CNS: anatomical and functional implications. Trends Neurosci 18:2229.
Mansour A, Khachaturian H, Lewis ME, Akil H, and Watson SJ (1987) Autoradiographic differentiation of mu, delta, and kappa opioid receptors in the rat forebrain
and midbrain. J Neurosci 7:24452464.
Mansvelder HD, Keath JR, and McGehee DS (2002) Synaptic mechanisms underlie
nicotine-induced excitability of brain reward areas. Neuron 33:905919.
Mansvelder HD and McGehee DS (2000) Long-term potentiation of excitatory inputs
to brain reward areas by nicotine. Neuron 27:349357.
Mantle TJ, Tipton KF, and Garrett NJ (1976) Inhibition of monoamine oxidase by
amphetamine and related compounds. Biochem Pharmacol 25:20732077.
Manzoni OJ and Williams JT (1999) Presynaptic regulation of glutamate release in
the ventral tegmental area during morphine withdrawal. J Neurosci 19:
66296636.
Mao D, Perry DC, Yasuda RP, Wolfe BB, and Kellar KJ (2008) The alpha4beta2alpha5 nicotinic cholinergic receptor in rat brain is resistant to up-regulation by
nicotine in vivo. J Neurochem 104:446456.
Marek GJ, Vosmer G, and Seiden LS (1990) Dopamine uptake inhibitors block longterm neurotoxic effects of methamphetamine upon dopaminergic neurons. Brain
Res 513:274279.
Margolis EB, Lock H, Chefer VI, Shippenberg TS, Hjelmstad GO, and Fields HL
(2006) Kappa opioids selectively control dopaminergic neurons projecting to the
prefrontal cortex. Proc Natl Acad Sci USA 103:29382942.
Mariqueo TA, Agurto A, Muoz B, San Martin L, Coronado C, Fernndez-Prez EJ,
Murath P, Snchez A, Homanics GE, and Aguayo LG (2014) Effects of ethanol on glycinergic synaptic currents in mouse spinal cord neurons. J Neurophysiol 111:19401948.
Mark KA, Soghomonian JJ, and Yamamoto BK (2004) High-dose methamphetamine
acutely activates the striatonigral pathway to increase striatal glutamate and
mediate long-term dopamine toxicity. J Neurosci 24:1144911456.
Markevich VA, Grigoryan GA, Dawe GS, and Stephenson JD (2007) Theta driving
both inhibits and potentiates the effects of nicotine on dentate gyrus responses.
Neurosci Behav Physiol 37:403409.
Markov D, Mosharov EV, Setlik W, Gershon MD, and Sulzer D (2008) Secretory
vesicle rebound hyperacidification and increased quantal size resulting from prolonged methamphetamine exposure. J Neurochem 107:17091721.
Markowitz JS, DeVane CL, Pestreich LK, Patrick KS, and Muniz R (2006) A comprehensive in vitro screening of d-, l-, and dl-threo-methylphenidate: an exploratory study. J Child Adolesc Psychopharmacol 16:687698.
Marks MJ, Burch JB, and Collins AC (1983) Effects of chronic nicotine infusion on
tolerance development and nicotinic receptors. J Pharmacol Exp Ther 226:
817825.
Marks MJ, Grady SR, Salminen O, Paley MA, Wageman CR, McIntosh JM,
and Whiteaker P (2014) a6b2*-subtype nicotinic acetylcholine receptors are more
sensitive than a4b2*-subtype receptors to regulation by chronic nicotine administration. J Neurochem 130:185198.
Marks MJ, Pauly JR, Gross SD, Deneris ES, Hermans-Borgmeyer I, Heinemann SF,
and Collins AC (1992) Nicotine binding and nicotinic receptor subunit RNA after
chronic nicotine treatment. J Neurosci 12:27652784.
Marona-Lewicka D, Nichols CD, and Nichols DE (2011) An animal model of schizophrenia based on chronic LSD administration: old idea, new results. Neuropharmacology 61:503512.
Marona-Lewicka D and Nichols DE (2007) Further evidence that the delayed temporal dopaminergic effects of LSD are mediated by a mechanism different than the
first temporal phase of action. Pharmacol Biochem Behav 87:453461.
Marona-Lewicka D, Thisted RA, and Nichols DE (2005) Distinct temporal phases in
the behavioral pharmacology of LSD: dopamine D2 receptor-mediated effects in the
rat and implications for psychosis. Psychopharmacology (Berl) 180:427435.
Maroteaux M and Mameli M (2012) Cocaine evokes projection-specific synaptic
plasticity of lateral habenula neurons. J Neurosci 32:1264112646.
Marshall JF, Belcher AM, Feinstein EM, and ODell SJ (2007) Methamphetamineinduced neural and cognitive changes in rodents. Addiction 102 (Suppl 1):6169.
Marsicano G and Lutz B (1999) Expression of the cannabinoid receptor CB1 in distinct neuronal subpopulations in the adult mouse forebrain. Eur J Neurosci 11:
42134225.
Marsicano G, Wotjak CT, Azad SC, Bisogno T, Rammes G, Cascio MG, Hermann H,
Tang J, Hofmann C, and Zieglgnsberger W, et al. (2002) The endogenous cannabinoid system controls extinction of aversive memories. Nature 418:530534.
Marston HM, Reid ME, Lawrence JA, Olverman HJ, and Butcher SP (1999)
Behavioural analysis of the acute and chronic effects of MDMA treatment in the
rat. Psychopharmacology (Berl) 144:6776.
Martin GE, Hendrickson LM, Penta KL, Friesen RM, Pietrzykowski AZ, Tapper AR,
and Treistman SN (2008) Identification of a BK channel auxiliary protein controlling molecular and behavioral tolerance to alcohol. Proc Natl Acad Sci USA
105:1754317548.
Martn YB, Gramage E, and Herradn G (2013) Maintenance of amphetamineinduced place preference does not correlate with astrocytosis. Eur J Pharmacol
699:258263.
Martinez D, Broft A, Foltin RW, Slifstein M, Hwang DR, Huang Y, Perez A, Frankle
WG, Cooper T, and Kleber HD, et al. (2004) Cocaine dependence and d2 receptor
availability in the functional subdivisions of the striatum: relationship with
cocaine-seeking behavior. Neuropsychopharmacology 29:11901202.

989

Martnez-Clemente J, Escubedo E, Pubill D, and Camarasa J (2012) Interaction of


mephedrone with dopamine and serotonin targets in rats. Eur Neuropsychopharmacol 22:231236.
Martnez-Clemente J, Lpez-Arnau R, Abad S, Pubill D, Escubedo E, and Camarasa
J (2014) Dose and time-dependent selective neurotoxicity induced by mephedrone
in mice. PLoS One 9:e99002.
Martinowich K and Lu B (2008) Interaction between BDNF and serotonin: role in
mood disorders. Neuropsychopharmacology 33:7383.
Mas M, Farr M, de la Torre R, Roset PN, Ortuo J, Segura J, and Cam J (1999)
Cardiovascular and neuroendocrine effects and pharmacokinetics of 3, 4-methylenedioxymethamphetamine in humans. J Pharmacol Exp Ther 290:136145.
Mascia MP, Mihic SJ, Valenzuela CF, Schofield PR, and Harris RA (1996) A single
amino acid determines differences in ethanol actions on strychnine-sensitive glycine receptors. Mol Pharmacol 50:402406.
Masood K, Wu C, Brauneis U, and Weight FF (1994) Differential ethanol sensitivity
of recombinant N-methyl-D-aspartate receptor subunits. Mol Pharmacol 45:
324329.
Math JM, Nomikos GG, Blakeman KH, and Svensson TH (1999) Differential actions
of dizocilpine (MK-801) on the mesolimbic and mesocortical dopamine systems: role
of neuronal activity. Neuropharmacology 38:121128.
Math JM, Nomikos GG, Hildebrand BE, Hertel P, and Svensson TH (1996) Prazosin
inhibits MK-801-induced hyperlocomotion and dopamine release in the nucleus
accumbens. Eur J Pharmacol 309:111.
Mato S, Chevaleyre V, Robbe D, Pazos A, Castillo PE, and Manzoni OJ (2004) A
single in-vivo exposure to delta 9THC blocks endocannabinoid-mediated synaptic
plasticity. Nat Neurosci 7:585586.
Mato S, Del Olmo E, and Pazos A (2003) Ontogenetic development of cannabinoid
receptor expression and signal transduction functionality in the human brain. Eur
J Neurosci 17:17471754.
Matsuda LA, Bonner TI, and Lolait SJ (1993) Localization of cannabinoid receptor
mRNA in rat brain. J Comp Neurol 327:535550.
Matsumoto M and Hikosaka O (2007) Lateral habenula as a source of negative reward signals in dopamine neurons. Nature 447:11111115.
Matsumoto M and Hikosaka O (2009) Two types of dopamine neuron distinctly
convey positive and negative motivational signals. Nature 459:837841.
Matsumoto RR, Nguyen L, Kaushal N, and Robson MJ (2014) Sigma (s) receptors as
potential therapeutic targets to mitigate psychostimulant effects. Adv Pharmacol
69:323386.
Matsuzaki M, Honkura N, Ellis-Davies GC, and Kasai H (2004) Structural basis of
long-term potentiation in single dendritic spines. Nature 429:761766.
Matthews DB, Devaud LL, Fritschy JM, Sieghart W, and Morrow AL (1998) Differential regulation of GABA(A) receptor gene expression by ethanol in the rat hippocampus versus cerebral cortex. J Neurochem 70:11601166.
Mtys F, Urbn GM, Watanabe M, Mackie K, Zimmer A, Freund TF, and Katona I
(2008) Identification of the sites of 2-arachidonoylglycerol synthesis and action
imply retrograde endocannabinoid signaling at both GABAergic and glutamatergic
synapses in the ventral tegmental area. Neuropharmacology 54:95107.
Maze I, Chaudhury D, Dietz DM, Von Schimmelmann M, Kennedy PJ, Lobo MK,
Sillivan SE, Miller ML, Bagot RC, and Sun H, et al. (2014) G9a influences neuronal
subtype specification in striatum. Nat Neurosci 17:533539.
Maze I, Covington HE 3rd, Dietz DM, LaPlant Q, Renthal W, Russo SJ, Mechanic M,
Mouzon E, Neve RL, and Haggarty SJ, et al. (2010) Essential role of the histone
methyltransferase G9a in cocaine-induced plasticity. Science 327:213216.
Mazei-Robison MS, Koo JW, Friedman AK, Lansink CS, Robison AJ, Vinish M,
Krishnan V, Kim S, Siuta MA, and Galli A, et al. (2011) Role for mTOR signaling
and neuronal activity in morphine-induced adaptations in ventral tegmental area
dopamine neurons. Neuron 72:977990.
McCann UD, Mertl M, Eligulashvili V, and Ricaurte GA (1999) Cognitive performance in (+/-) 3,4-methylenedioxymethamphetamine (MDMA, ecstasy) users:
a controlled study. Psychopharmacology (Berl) 143:417425.
McCann UD, Szabo Z, Scheffel U, Dannals RF, and Ricaurte GA (1998a) Positron
emission tomographic evidence of toxic effect of MDMA (Ecstasy) on brain serotonin neurons in human beings. Lancet 352:14331437.
McCann UD, Szabo Z, Vranesic M, Palermo M, Mathews WB, Ravert HT,
Dannals RF, and Ricaurte GA (2008) Positron emission tomographic studies
of brain dopamine and serotonin transporters in abstinent (+/-)3,4methylenedioxymethamphetamine (ecstasy) users: relationship to cognitive performance. Psychopharmacology (Berl) 200:439450.
McCann UD, Wong DF, Yokoi F, Villemagne V, Dannals RF, and Ricaurte GA
(1998b) Reduced striatal dopamine transporter density in abstinent methamphetamine and methcathinone users: evidence from positron emission tomography
studies with [11C]WIN-35,428. J Neurosci 18:84178422.
McCardle K, Luebbers S, Carter JD, Croft RJ, and Stough C (2004) Chronic MDMA
(ecstasy) use, cognition and mood. Psychopharmacology (Berl) 173:434439.
McClung CA and Nestler EJ (2003) Regulation of gene expression and cocaine reward
by CREB and DeltaFosB. Nat Neurosci 6:12081215.
McCool BA (2011) Ethanol modulation of synaptic plasticity. Neuropharmacology 61:
10971108.
McCool BA, Frye GD, Pulido MD, and Botting SK (2003) Effects of chronic ethanol
consumption on rat GABA(A) and strychnine-sensitive glycine receptors expressed
by lateral/basolateral amygdala neurons. Brain Res 963:165177.
McCutcheon JE, Loweth JA, Ford KA, Marinelli M, Wolf ME, and Tseng KY (2011a)
Group I mGluR activation reverses cocaine-induced accumulation of calciumpermeable AMPA receptors in nucleus accumbens synapses via a protein kinase
C-dependent mechanism. J Neurosci 31:1453614541.
McCutcheon JE, Wang X, Tseng KY, Wolf ME, and Marinelli M (2011b) Calciumpermeable AMPA receptors are present in nucleus accumbens synapses after
prolonged withdrawal from cocaine self-administration but not experimenteradministered cocaine. J Neurosci 31:57375743.

990

Korpi et al.

McDaid J, Graham MP, and Napier TC (2006) Methamphetamine-induced sensitization differentially alters pCREB and DeltaFosB throughout the limbic circuit of
the mammalian brain. Mol Pharmacol 70:20642074.
McEwen BS (2012) Brain on stress: how the social environment gets under the skin.
Proc Natl Acad Sci USA 109 (Suppl 2):1718017185.
McGee R, Williams S, Poulton R, and Moffitt T (2000) A longitudinal study of cannabis use and mental health from adolescence to early adulthood. Addiction 95:
491503.
McGough NN, He DY, Logrip ML, Jeanblanc J, Phamluong K, Luong K, Kharazia V,
Janak PH, and Ron D (2004) RACK1 and brain-derived neurotrophic factor: a homeostatic pathway that regulates alcohol addiction. J Neurosci 24:1054210552.
McGregor IS, Clemens KJ, Van der Plasse G, Li KM, Hunt GE, Chen F,
and Lawrence AJ (2003a) Increased anxiety 3 months after brief exposure to
MDMA (Ecstasy) in rats: association with altered 5-HT transporter and receptor
density. Neuropsychopharmacology 28:14721484.
McGregor IS, Gurtman CG, Morley KC, Clemens KJ, Blokland A, Li KM, Cornish JL,
and Hunt GE (2003b) Increased anxiety and depressive symptoms months after
MDMA (ecstasy) in rats: drug-induced hyperthermia does not predict long-term
outcomes. Psychopharmacology (Berl) 168:465474.
McKay DB, Chang C, Gonzlez-Cestari TF, McKay SB, El-Hajj RA, Bryant DL, Zhu
MX, Swaan PW, Arason KM, and Pulipaka AB, et al. (2007) Analogs of methyllycaconitine as novel noncompetitive inhibitors of nicotinic receptors: pharmacological
characterization, computational modeling, and pharmacophore development. Mol
Pharmacol 71:12881297.
McKenna DJ, Repke DB, Lo L, and Peroutka SJ (1990) Differential interactions of
indolealkylamines with 5-hydroxytryptamine receptor subtypes. Neuropharmacology 29:193198.
McKernan RM, Rosahl TW, Reynolds DS, Sur C, Wafford KA, Atack JR, Farrar S,
Myers J, Cook G, and Ferris P, et al. (2000) Sedative but not anxiolytic properties
of benzodiazepines are mediated by the GABA(A) receptor alpha1 subtype. Nat
Neurosci 3:587592.
McKetin R, Lubman DI, Baker AL, Dawe S, and Ali RL (2013) Dose-related psychotic
symptoms in chronic methamphetamine users: evidence from a prospective longitudinal study. JAMA Psychiatry 70:319324.
McNamara CG, Tejero-Cantero , Trouche S, Campo-Urriza N, and Dupret D (2014)
Dopaminergic neurons promote hippocampal reactivation and spatial memory
persistence. Nat Neurosci 17:16581660.
Meaney MJ, Aitken DH, van Berkel C, Bhatnagar S, and Sapolsky RM (1988) Effect
of neonatal handling on age-related impairments associated with the hippocampus.
Science 239:766768.
Mechan AO, Moran PM, Elliott M, Young AJ, Joseph MH, and Green R (2002) A
study of the effect of a single neurotoxic dose of 3,4-methylenedioxymethamphetamine (MDMA; ecstasy) on the subsequent long-term behaviour of rats in the
plus maze and open field. Psychopharmacology (Berl) 159:167175.
Mechoulam R, Ben-Shabat S, Hanus L, Ligumsky M, Kaminski NE, Schatz AR,
Gopher A, Almog S, Martin BR, and Compton DR, et al. (1995) Identification of an
endogenous 2-monoglyceride, present in canine gut, that binds to cannabinoid
receptors. Biochem Pharmacol 50:8390.
Medhus S, Mordal J, Holm B, Mrland J, and Bramness JG (2013) A comparison of
symptoms and drug use between patients with methamphetamine associated
psychoses and patients diagnosed with schizophrenia in two acute psychiatric
wards. Psychiatry Res 206:1721.
Medina I, Friedel P, Rivera C, Kahle KT, Kourdougli N, Uvarov P, and Pellegrino C
(2014) Current view on the functional regulation of the neuronal K(+)-Cl(-)
cotransporter KCC2. Front Cell Neurosci 8:27.
Mei L and Xiong WC (2008) Neuregulin 1 in neural development, synaptic plasticity
and schizophrenia. Nat Rev Neurosci 9:437452.
Melega WP, Williams AE, Schmitz DA, DiStefano EW, and Cho AK (1995) Pharmacokinetic and pharmacodynamic analysis of the actions of D-amphetamine and
D-methamphetamine on the dopamine terminal. J Pharmacol Exp Ther 274:9096.
Melis M, Camarini R, Ungless MA, and Bonci A (2002) Long-lasting potentiation of
GABAergic synapses in dopamine neurons after a single in vivo ethanol exposure.
J Neurosci 22:20742082.
Melis M, Enrico P, Peana AT, and Diana M (2007) Acetaldehyde mediates alcohol
activation of the mesolimbic dopamine system. Eur J Neurosci 26:28242833.
Melis M, Pistis M, Perra S, Muntoni AL, Pillolla G, and Gessa GL (2004) Endocannabinoids mediate presynaptic inhibition of glutamatergic transmission in rat
ventral tegmental area dopamine neurons through activation of CB1 receptors. J
Neurosci 24:5362.
Mellman I, Fuchs R, and Helenius A (1986) Acidification of the endocytic and exocytic
pathways. Annu Rev Biochem 55:663700.
Meltzer PC, Butler D, Deschamps JR, and Madras BK (2006) 1-(4-Methylphenyl)-2pyrrolidin-1-yl-pentan-1-one (Pyrovalerone) analogues: a promising class of
monoamine uptake inhibitors. J Med Chem 49:14201432.
Mendelson J, Uemura N, Harris D, Nath RP, Fernandez E, Jacob P 3rd, Everhart
ET, and Jones RT (2006) Human pharmacology of the methamphetamine stereoisomers. Clin Pharmacol Ther 80:403420.
Meredith GE, Callen S, and Scheuer DA (2002) Brain-derived neurotrophic factor
expression is increased in the rat amygdala, piriform cortex and hypothalamus
following repeated amphetamine administration. Brain Res 949:218227.
Mereu G, Yoon KW, Boi V, Gessa GL, Naes L, and Westfall TC (1987) Preferential
stimulation of ventral tegmental area dopaminergic neurons by nicotine. Eur J
Pharmacol 141:395399.
Merlo Pich E, Lorang M, Yeganeh M, Rodriguez de Fonseca F, Raber J, Koob GF,
and Weiss F (1995) Increase of extracellular corticotropin-releasing factor-like
immunoreactivity levels in the amygdala of awake rats during restraint stress and
ethanol withdrawal as measured by microdialysis. J Neurosci 15:54395447.
Metaxas A, Willems R, Kooijman EJ, Renjan VA, Klein PJ, Windhorst AD, Donck
LV, Leysen JE, and Berckel BN (2014) Subchronic treatment with phencyclidine in
adolescence leads to impaired exploratory behavior in adult rats without altering

social interaction or N-methyl-D-aspartate receptor binding levels. J Neurosci Res


92:15991607.
Meye FJ, Lecca S, Valentinova K, and Mameli M (2013) Synaptic and cellular profile
of neurons in the lateral habenula. Front Hum Neurosci 7:860. DOI 10.3389/
fnhum.2013.00860
Meye FJ, Valentinova K, Lecca S, Marion-Poll L, Maroteaux MJ, Musardo S,
Moutkine I, Gardoni F, Huganir RL, and Georges F, et al. (2015) Cocaine-evoked
negative symptoms require AMPA receptor trafficking in the lateral habenula. Nat
Neurosci 18:376378.
Michels G and Moss SJ (2007) GABAA receptors: properties and trafficking. Crit Rev
Biochem Mol Biol 42:314.
Michelson D, Snyder E, Paradis E, Chengan-Liu M, Snavely DB, Hutzelmann J,
Walsh JK, Krystal AD, Benca RM, and Cohn M, et al. (2014) Safety and efficacy of
suvorexant during 1-year treatment of insomnia with subsequent abrupt treatment discontinuation: a phase 3 randomised, double-blind, placebo-controlled trial.
Lancet Neurol 13:461471.
Mihalek RM, Banerjee PK, Korpi ER, Quinlan JJ, Firestone LL, Mi ZP, Lagenaur C,
Tretter V, Sieghart W, and Anagnostaras SG, et al. (1999) Attenuated sensitivity to
neuroactive steroids in gamma-aminobutyrate type A receptor delta subunit
knockout mice. Proc Natl Acad Sci USA 96:1290512910.
Millan MJ and Brocco M (2003) The Vogel conflict test: procedural aspects, gammaaminobutyric acid, glutamate and monoamines. Eur J Pharmacol 463:6796.
Millar NS and Gotti C (2009) Diversity of vertebrate nicotinic acetylcholine receptors.
Neuropharmacology 56:237246.
Millar NS, Gotti C, Marks MJ, and Wonnacott S(2014) Nicotinic acetylcholine
receptors, introduction, in IUPHAR/BPS Guide to PHARMACOLOGY, http://www.
guidetopharmacology.org/GRAC/FamilyIntroductionForward?familyId=76.
Miller HH, Shore PA, and Clarke DE (1980) In vivo monoamine oxidase inhibition by
d-amphetamine. Biochem Pharmacol 29:13471354.
Miller KD (1996) Synaptic economics: competition and cooperation in synaptic
plasticity. Neuron 17:371374.
Miller ML, Creehan KM, Angrish D, Barlow DJ, Houseknecht KL, Dickerson TJ,
and Taffe MA (2013) Changes in ambient temperature differentially alter the
thermoregulatory, cardiac and locomotor stimulant effects of 4-methylmethcathinone (mephedrone). Drug Alcohol Depend 127:248253.
Milusheva E (1992) Effect of MK-801 on dopamine release evoked by hypoxia combined with hypoglycemia. Acta Physiol Hung 79:347354.
Mingione CJ, Heffner JL, Blom TJ, and Anthenelli RM (2012) Childhood adversity,
serotonin transporter (5-HTTLPR) genotype, and risk for cigarette smoking and
nicotine dependence in alcohol dependent adults. Drug Alcohol Depend 123:
201206.
Minichiello L (2009) TrkB signalling pathways in LTP and learning. Nat Rev Neurosci 10:850860.
Mirenowicz J and Schultz W (1996) Preferential activation of midbrain dopamine
neurons by appetitive rather than aversive stimuli. Nature 379:449451.
Misner DL and Sullivan JM (1999) Mechanism of cannabinoid effects on long-term
potentiation and depression in hippocampal CA1 neurons. J Neurosci 19:
67956805.
Mitchell JM, Bergren LJ, Chen KS, and Fields HL (2006) Cholecystokinin is necessary for the expression of morphine conditioned place preference. Pharmacol Biochem Behav 85:787795.
Mitchell SN, Brazell MP, Joseph MH, Alavijeh MS, and Gray JA (1989) Regionally
specific effects of acute and chronic nicotine on rates of catecholamine and
5-hydroxytryptamine synthesis in rat brain. Eur J Pharmacol 167:311322.
Mitchell SN, Smith KM, Joseph MH, and Gray JA (1993) Increases in tyrosine hydroxylase messenger RNA in the locus coeruleus after a single dose of nicotine are
followed by time-dependent increases in enzyme activity and noradrenaline release. Neuroscience 56:989997.
Mitchell VA, Jeong HJ, Drew GM, and Vaughan CW (2011) Cholecystokinin exerts
an effect via the endocannabinoid system to inhibit GABAergic transmission in
midbrain periaqueductal gray. Neuropsychopharmacology 36:18011810.
Mithoefer MC, Wagner MT, Mithoefer AT, Jerome L, and Doblin R (2011) The safety
and efficacy of +/-3,4-methylenedioxymethamphetamine-assisted psychotherapy in
subjects with chronic, treatment-resistant posttraumatic stress disorder: the first
randomized controlled pilot study. J Psychopharmacol 25:439452.
Mithoefer MC, Wagner MT, Mithoefer AT, Jerome L, Martin SF, Yazar-Klosinski B,
Michel Y, Brewerton TD, and Doblin R (2013) Durability of improvement in posttraumatic stress disorder symptoms and absence of harmful effects or drug dependency after 3,4-methylenedioxymethamphetamine-assisted psychotherapy:
a prospective long-term follow-up study. J Psychopharmacol 27:2839.
Mittendorf R, Covert R, Boman J, Khoshnood B, Lee KS, and Siegler M (1997) Is
tocolytic magnesium sulphate associated with increased total paediatric mortality?
Lancet 350:15171518.
Mittendorf R, Dambrosia J, Pryde PG, Lee KS, Gianopoulos JG, Besinger RE, and Tomich
PG (2002) Association between the use of antenatal magnesium sulfate in preterm
labor and adverse health outcomes in infants. Am J Obstet Gynecol 186:11111118.
Miyazaki M, Noda Y, Mouri A, Kobayashi K, Mishina M, Nabeshima T, and Yamada
K (2013) Role of convergent activation of glutamatergic and dopaminergic systems
in the nucleus accumbens in the development of methamphetamine psychosis and
dependence. Int J Neuropsychopharmacol 16:13411350.
Mlinar B and Corradetti R (2003) Endogenous 5-HT, released by MDMA through
serotonin transporter- and secretory vesicle-dependent mechanisms, reduces hippocampal excitatory synaptic transmission by preferential activation of 5-HT1B
receptors located on CA1 pyramidal neurons. Eur J Neurosci 18:15591571.
Moaddel R, Abdrakhmanova G, Kozak J, Jozwiak K, Toll L, Jimenez L, Rosenberg A,
Tran T, Xiao Y, and Zarate CA, et al. (2013) Sub-anesthetic concentrations of (R,S)ketamine metabolites inhibit acetylcholine-evoked currents in a7 nicotinic acetylcholine receptors. Eur J Pharmacol 698:228234.
Moaddel R, Luckenbaugh DA, Xie Y, Villaseor A, Brutsche NE, Machado-Vieira R,
Ramamoorthy A, Lorenzo MP, Garcia A, and Bernier M, et al. (2015) D-serine

Drug-Induced Neuroplasticity
plasma concentration is a potential biomarker of (R,S)-ketamine antidepressant
response in subjects with treatment-resistant depression. Psychopharmacology
(Berl) 232:399409.
Mody I (2001) Distinguishing between GABA(A) receptors responsible for tonic and
phasic conductances. Neurochem Res 26:907913.
Moeller FG, Steinberg JL, Dougherty DM, Narayana PA, Kramer LA, and Renshaw
PF (2004) Functional MRI study of working memory in MDMA users. Psychopharmacology (Berl) 177:185194.
Moeller FG, Steinberg JL, Schmitz JM, Ma L, Liu S, Kjome KL, Rathnayaka N, Kramer
LA, and Narayana PA (2010) Working memory fMRI activation in cocaine-dependent
subjects: association with treatment response. Psychiatry Res 181:174182.
Moeller SJ, Konova AB, and Goldstein RZ (2015) Multiple ambiguities in the measurement of drug craving. Addiction 110:205206.
Moita MA, Rosis S, Zhou Y, LeDoux JE, and Blair HT (2004) Putting fear in its place:
remapping of hippocampal place cells during fear conditioning. J Neurosci 24:
70157023.
Molliver ME, Berger UV, Mamounas LA, Molliver DC, OHearn E, and Wilson MA
(1990) Neurotoxicity of MDMA and related compounds: anatomic studies. Ann NY
Acad Sci 600:649664.
Molteni R, Pasini M, Moraschi S, Gennarelli M, Drago F, Racagni G, and Riva MA
(2008) Reduced activation of intracellular signaling pathways in rat prefrontal
cortex after chronic phencyclidine administration. Pharmacol Res 57:296302.
Monassier L and Bousquet P (2002) Sigma receptors: from discovery to highlights of
their implications in the cardiovascular system. Fundam Clin Pharmacol 16:18.
Monks TJ, Jones DC, Bai F, and Lau SS (2004) The role of metabolism in 3,4(+)-methylenedioxyamphetamine and 3,4-(+)-methylenedioxymethamphetamine
(ecstasy) toxicity. Ther Drug Monit 26:132136.
Mons N and Cooper DM (1995) Adenylate cyclases: critical foci in neuronal signaling.
Trends Neurosci 18:536542.
Monterosso JR, Ainslie G, Xu J, Cordova X, Domier CP, and London ED (2007)
Frontoparietal cortical activity of methamphetamine-dependent and comparison
subjects performing a delay discounting task. Hum Brain Mapp 28:383393.
Montgomery AJ, Lingford-Hughes AR, Egerton A, Nutt DJ, and Grasby PM (2007a)
The effect of nicotine on striatal dopamine release in man: A [11C]raclopride PET
study. Synapse 61:637645.
Montgomery T, Buon C, Eibauer S, Guiry PJ, Keenan AK, and McBean GJ (2007b)
Comparative potencies of 3,4-methylenedioxymethamphetamine (MDMA) analogues as inhibitors of [3H]noradrenaline and [3H]5-HT transport in mammalian
cell lines. Br J Pharmacol 152:11211130.
Moonat S, Sakharkar AJ, Zhang H, Tang L, and Pandey SC (2013) Aberrant histone
deacetylase2-mediated histone modifications and synaptic plasticity in the amygdala predisposes to anxiety and alcoholism. Biol Psychiatry 73:763773.
Moore TH, Zammit S, Lingford-Hughes A, Barnes TR, Jones PB, Burke M, and Lewis
G (2007) Cannabis use and risk of psychotic or affective mental health outcomes:
a systematic review. Lancet 370:319328.
Moorman DE and Aston-Jones G (2009) Orexin-1 receptor antagonism decreases
ethanol consumption and preference selectively in high-ethanolpreferring
SpragueDawley rats. Alcohol 43:379386.
Morales M and Bloom FE (1997) The 5-HT3 receptor is present in different subpopulations of GABAergic neurons in the rat telencephalon. J Neurosci 17:
31573167.
Moratalla R, Xu M, Tonegawa S, and Graybiel AM (1996) Cellular responses to
psychomotor stimulant and neuroleptic drugs are abnormal in mice lacking the D1
dopamine receptor. Proc Natl Acad Sci USA 93:1492814933.
Moreno M, Escuredo L, Muoz R, Rodriguez de Fonseca F, and Navarro M (2005)
Long-term behavioural and neuroendocrine effects of perinatal activation or
blockade of CB1 cannabinoid receptors. Behav Pharmacol 16:423430.
Morgan SL and Teyler TJ (1999) VDCCs and NMDARs underlie two forms of LTP in
CA1 hippocampus in vivo. J Neurophysiol 82:736740.
Morini R, Mlinar B, Baccini G, and Corradetti R (2011) Enhanced hippocampal longterm potentiation following repeated MDMA treatment in Dark-Agouti rats. Eur
Neuropsychopharmacol 21:8091.
Morohaku K, Pelton SH, Daugherty DJ, Butler WR, Deng W, and Selvaraj V (2014)
Translocator protein/peripheral benzodiazepine receptor is not required for steroid
hormone biosynthesis. Endocrinology 155:8997.
Morris H and Wallach J (2014) From PCP to MXE: a comprehensive review of the
non-medical use of dissociative drugs. Drug Test Anal 6:614632.
Morrisett RA, Mott DD, Lewis DV, Swartzwelder HS, and Wilson WA (1991) GABABreceptor-mediated inhibition of the N-methyl-D-aspartate component of synaptic
transmission in the rat hippocampus. J Neurosci 11:203209.
Morrow BA, Elsworth JD, and Roth RH (2007) Repeated phencyclidine in monkeys
results in loss of parvalbumin-containing axo-axonic projections in the prefrontal
cortex. Psychopharmacology (Berl) 192:283290.
Morter P and Herculano-Houzel S (2012) Age-related neuronal loss in the rat brain
starts at the end of adolescence. Front Neuroanat 6:45.
Motbey CP, Clemens KJ, Apetz N, Winstock AR, Ramsey J, Li KM, Wyatt N, Callaghan
PD, Bowen MT, and Cornish JL, et al. (2013) High levels of intravenous mephedrone
(4-methylmethcathinone) self-administration in rats: neural consequences and comparison with methamphetamine. J Psychopharmacol 27:823836.
Motbey CP, Karanges E, Li KM, Wilkinson S, Winstock AR, Ramsay J, Hicks C,
Kendig MD, Wyatt N, and Callaghan PD, et al. (2012) Mephedrone in adolescent
rats: residual memory impairment and acute but not lasting 5-HT depletion. PLoS
One 7:e45473.
Mykkynen T and Korpi ER (2012) Acute effects of ethanol on glutamate receptors.
Basic Clin Pharmacol Toxicol 111:413.
Mykkynen T, Korpi ER, and Lovinger DM (2003) Ethanol inhibits alpha-amino-3hydyroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptor function in central
nervous system neurons by stabilizing desensitization. J Pharmacol Exp Ther 306:
546555.

991

Muetzelfeldt L, Kamboj SK, Rees H, Taylor J, Morgan CJ, and Curran HV (2008)
Journey through the K-hole: phenomenological aspects of ketamine use. Drug Alcohol Depend 95:219229.
Mulle C, Vidal C, Benoit P, and Changeux JP (1991) Existence of different subtypes
of nicotinic acetylcholine receptors in the rat habenulo-interpeduncular system.
J Neurosci 11:25882597.
Muoz P, Huenchuguala S, Paris I, and Segura-Aguilar J (2012) Dopamine oxidation
and autophagy. Parkinsons Dis 2012: DOI 10.1155/2012/920953
Murphy BL, Arnsten AF, Goldman-Rakic PS, and Roth RH (1996) Increased dopamine turnover in the prefrontal cortex impairs spatial working memory performance in rats and monkeys. Proc Natl Acad Sci USA 93:13251329.
Murrough JW, Perez AM, Pillemer S, Stern J, Parides MK, aan het Rot M, Collins
KA, Mathew SJ, Charney DS, and Iosifescu DV (2013) Rapid and longer-term
antidepressant effects of repeated ketamine infusions in treatment-resistant major
depression. Biol Psychiatry 74:250256.
Muschamp JW, Vant Veer A, Parsegian A, Gallo MS, Chen M, Neve RL, Meloni EG,
and Carlezon WA Jr (2011) Activation of CREB in the nucleus accumbens shell
produces anhedonia and resistance to extinction of fear in rats. J Neurosci 31:
30953103.
Musso F, Bettermann F, Vucurevic G, Stoeter P, Konrad A, and Winterer G (2007)
Smoking impacts on prefrontal attentional network function in young adult brains.
Psychopharmacology (Berl) 191:159169.
Muthukumaraswamy SD, Carhart-Harris RL, Moran RJ, Brookes MJ, Williams TM,
Errtizoe D, Sessa B, Papadopoulos A, Bolstridge M, and Singh KD, et al. (2013)
Broadband cortical desynchronization underlies the human psychedelic state.
J Neurosci 33:1517115183.
Mychasiuk R, Muhammad A, Ilnytskyy S, and Kolb B (2013) Persistent gene expression changes in NAc, mPFC, and OFC associated with previous nicotine or
amphetamine exposure. Behav Brain Res 256:655661.
Mychasiuk R, Muhammad A, and Kolb B (2014) Environmental enrichment alters
structural plasticity of the adolescent brain but does not remediate the effects of
prenatal nicotine exposure. Synapse 68:293305.
Nabeshima T, Hiramatsu M, Yamaguchi K, Kasugai M, Ishizaki K, Kawashima K,
Itoh K, Ogawa S, Katoh A, and Furukawa H, et al. (1988) Effects of prenatal
administration of phencyclidine on the learning and memory processes of rat offspring. J Pharmacobiodyn 11:816823.
Nacher J, Guirado R, and Castillo-Gmez E (2013) Structural plasticity of interneurons in the adult brain: role of PSA-NCAM and implications for psychiatric
disorders. Neurochem Res 38:11221133.
Nader MA and Czoty PW (2005) PET imaging of dopamine D2 receptors in monkey
models of cocaine abuse: genetic predisposition versus environmental modulation.
Am J Psychiatry 162:14731482.
Nader MA, Morgan D, Gage HD, Nader SH, Calhoun TL, Buchheimer N, Ehrenkaufer R, and Mach RH (2006) PET imaging of dopamine D2 receptors during
chronic cocaine self-administration in monkeys. Nat Neurosci 9:10501056.
Nagai T, Takuma K, Dohniwa M, Ibi D, Mizoguchi H, Kamei H, Nabeshima T,
and Yamada K (2007) Repeated methamphetamine treatment impairs spatial
working memory in rats: reversal by clozapine but not haloperidol. Psychopharmacology (Berl) 194:2132.
Ngerl UV, Eberhorn N, Cambridge SB, and Bonhoeffer T (2004) Bidirectional
activity-dependent morphological plasticity in hippocampal neurons. Neuron 44:
759767.
Nair A and Bonneau RH (2006) Stress-induced elevation of glucocorticoids increases
microglia proliferation through NMDA receptor activation. J Neuroimmunol 171:
7285.
Nair-Roberts RG, Chatelain-Badie SD, Benson E, White-Cooper H, Bolam JP,
and Ungless MA (2008) Stereological estimates of dopaminergic, GABAergic and
glutamatergic neurons in the ventral tegmental area, substantia nigra and retrorubral field in the rat. Neuroscience 152:10241031.
Nairn AC, Svenningsson P, Nishi A, Fisone G, Girault JA, and Greengard P (2004)
The role of DARPP-32 in the actions of drugs of abuse. Neuropharmacology 47
(Suppl 1):1423.
Nakamura Y, Darnieder LM, Deeb TZ, and Moss SJ (2015) Regulation of GABAARs
by phosphorylation. Adv Pharmacol 72:97146.
Nakao S, Adachi T, Murakawa M, Shinomura T, Kurata J, Shichino T, Shibata M,
Tocyama I, Kimura H, and Mori K (1996) Halothane and diazepam inhibit
ketamine-induced c-fos expression in the rat cingulate cortex. Anesthesiology 85:
874882.
Nakauchi S and Sumikawa K (2014) Endogenous ACh suppresses LTD induction and
nicotine relieves the suppression via different nicotinic ACh receptor subtypes in
the mouse hippocampus. Life Sci 111:6268.
Nakki R, Nickolenko J, Chang J, Sagar SM, and Sharp FR (1996) Haloperidol prevents ketamine- and phencyclidine-induced HSP70 protein expression but not
microglial activation. Exp Neurol 137:234241.
Naqvi NH, Rudrauf D, Damasio H, and Bechara A (2007) Damage to the insula
disrupts addiction to cigarette smoking. Science 315:531534.
Narahashi T, Aistrup GL, Marszalec W, and Nagata K (1999) Neuronal nicotinic
acetylcholine receptors: a new target site of ethanol. Neurochem Int 35:131141.
Narr KL, Woods RP, Lin J, Kim J, Phillips OR, DelHomme M, Caplan R, Toga AW,
McCracken JT, and Levitt JG (2009) Widespread cortical thinning is a robust
anatomical marker for attention-deficit/hyperactivity disorder. J Am Acad Child
Adolesc Psychiatry 48:10141022.
Nash JF and Yamamoto BK (1992) Methamphetamine neurotoxicity and striatal
glutamate release: comparison to 3,4-methylenedioxymethamphetamine. Brain
Res 581:237243.
Nash JF and Yamamoto BK (1993) Effect of D-amphetamine on the extracellular concentrations of glutamate and dopamine in iprindole-treated rats. Brain Res 627:18.
Nashmi R, Dickinson ME, McKinney S, Jareb M, Labarca C, Fraser SE, and Lester
HA (2003) Assembly of alpha4beta2 nicotinic acetylcholine receptors assessed with
functional fluorescently labeled subunits: effects of localization, trafficking, and

992

Korpi et al.

nicotine-induced upregulation in clonal mammalian cells and in cultured midbrain


neurons. J Neurosci 23:1155411567.
Nashmi R, Xiao C, Deshpande P, McKinney S, Grady SR, Whiteaker P, Huang Q,
McClure-Begley T, Lindstrom JM, and Labarca C, et al. (2007) Chronic nicotine
cell specifically upregulates functional alpha 4* nicotinic receptors: basis for both
tolerance in midbrain and enhanced long-term potentiation in perforant path.
J Neurosci 27:82028218.
Nazar M, Jessa M, and Paznik A (1997) Benzodiazepine-GABAA receptor complex
ligands in two models of anxiety. J Neural Transm 104:733746.
Nazzaro C, Greco B, Cerovic M, Baxter P, Rubino T, Trusel M, Parolaro D, Tkatch T,
Benfenati F, and Pedarzani P, et al. (2012) SK channel modulation rescues striatal
plasticity and control over habit in cannabinoid tolerance. Nat Neurosci 15:
284293.
Nealey KA, Smith AW, Davis SM, Smith DG, and Walker BM (2011) k-opioid
receptors are implicated in the increased potency of intra-accumbens nalmefene in
ethanol-dependent rats. Neuropharmacology 61:3542.
Neher E and Sakaba T (2008) Multiple roles of calcium ions in the regulation of
neurotransmitter release. Neuron 59:861872.
Nehring RB, Horikawa HP, El Far O, Kneussel M, Brandsttter JH, Stamm S, Wischmeyer
E, Betz H, and Karschin A (2000) The metabotropic GABAB receptor directly
interacts with the activating transcription factor 4. J Biol Chem 275:3518535191.
Nelson CL, Milovanovic M, Wetter JB, Ford KA, and Wolf ME (2009) Behavioral
sensitization to amphetamine is not accompanied by changes in glutamate receptor
surface expression in the rat nucleus accumbens. J Neurochem 109:3551.
Nelson SB (2004) Hebb and anti-Hebb meet in the brainstem. Nat Neurosci 7:
687688.
Nestler EJ (2001) Molecular basis of long-term plasticity underlying addiction. Nat
Rev Neurosci 2:119128.
Nestler EJ (2014) Epigenetic mechanisms of drug addiction. Neuropharmacology 76
(Pt B):259268.
Nestler EJ, Barrot M, and Self DW (2001) DeltaFosB: a sustained molecular switch
for addiction. Proc Natl Acad Sci USA 98:1104211046.
Nestor LJ, Ghahremani DG, Monterosso J, and London ED (2011) Prefrontal hypoactivation during cognitive control in early abstinent methamphetaminedependent subjects. Psychiatry Res 194:287295.
Neuman RS and Harley CW (1983) Long-lasting potentiation of the dentate gyrus
population spike by norepinephrine. Brain Res 273:162165.
Nichols DE(1994) Medicinal chemistry and structure-activity telationships, in Amphetamine and Its Analogs: Psychopharmacology, Toxicology, And Abuse (Cho AK
and Segal DS eds), p 342, Academic Press, San Diego, CA.
Nichols CD, Garcia EE, and Sanders-Bush E (2003) Dynamic changes in prefrontal
cortex gene expression following lysergic acid diethylamide administration. Brain
Res Mol Brain Res 111:182188.
Nichols CD and Sanders-Bush E (2002) A single dose of lysergic acid diethylamide
influences gene expression patterns within the mammalian brain. Neuropsychopharmacology 26:634642.
Nichols DE (1986) Differences between the mechanism of action of MDMA, MBDB,
and the classic hallucinogens. Identification of a new therapeutic class: entactogens. J Psychoactive Drugs 18:305313.
Nichols DE (2004) Hallucinogens. Pharmacol Ther 101:131181.
Nickell JR, Krishnamurthy S, Norrholm S, Deaciuc G, Siripurapu KB, Zheng G,
Crooks PA, and Dwoskin LP (2010) Lobelane inhibits methamphetamine-evoked
dopamine release via inhibition of the vesicular monoamine transporter-2.
J Pharmacol Exp Ther 332:612621.
Nicoll RA and Schmitz D (2005) Synaptic plasticity at hippocampal mossy fibre
synapses. Nat Rev Neurosci 6:863876.
NIDA(2015) National Institute on Drug Abuse. Trends & Statistics. http://www.
drugabuse.gov/related-topics/trends-statistics [20 Feb 2015].
Nie Z, Schweitzer P, Roberts AJ, Madamba SG, Moore SD, and Siggins GR (2004)
Ethanol augments GABAergic transmission in the central amygdala via CRF1
receptors. Science 303:15121514.
Niehaus JL, Murali M, and Kauer JA (2010) Drugs of abuse and stress impair LTP at
inhibitory synapses in the ventral tegmental area. Eur J Neurosci 32:108117.
Nilsson LK, Schwieler L, Engberg G, Linderholm KR, and Erhardt S (2005) Activation of noradrenergic locus coeruleus neurons by clozapine and haloperidol: involvement of glutamatergic mechanisms. Int J Neuropsychopharmacol 8:329339.
Nisell M, Nomikos GG, and Svensson TH (1994) Systemic nicotine-induced dopamine
release in the rat nucleus accumbens is regulated by nicotinic receptors in the
ventral tegmental area. Synapse 16:3644.
Nithianantharajah J and Hannan AJ (2006) Enriched environments, experiencedependent plasticity and disorders of the nervous system. Nat Rev Neurosci 7:
697709.
Njus D, Kelley PM, and Harnadek GJ (1986) Bioenergetics of secretory vesicles.
Biochim Biophys Acta 853:237265.
Noonan MA, Bulin SE, Fuller DC, and Eisch AJ (2010) Reduction of adult hippocampal neurogenesis confers vulnerability in an animal model of cocaine addiction.
J Neurosci 30:304315.
Noonan MA, Choi KH, Self DW, and Eisch AJ (2008) Withdrawal from cocaine selfadministration normalizes deficits in proliferation and enhances maturity of adultgenerated hippocampal neurons. J Neurosci 28:25162526.
Nowicky AV, Teyler TJ, and Vardaris RM (1987) The modulation of long-term potentiation by delta-9-tetrahydrocannabinol in the rat hippocampus, in vitro. Brain
Res Bull 19:663672.
Nugent FS and Kauer JA (2008) LTP of GABAergic synapses in the ventral tegmental area and beyond. J Physiol 586:14871493.
Nugent FS, Niehaus JL, and Kauer JA (2009) PKG and PKA signaling in LTP at
GABAergic synapses. Neuropsychopharmacology 34:18291842.
Nugent FS, Penick EC, and Kauer JA (2007) Opioids block long-term potentiation of
inhibitory synapses. Nature 446:10861090.

Numa R, Baron M, Kohen R, and Yaka R (2011) Tempol attenuates cocaine-induced


death of PC12 cells through decreased oxidative damage. Eur J Pharmacol 650:
157162.
Numa R, Kohen R, Poltyrev T, and Yaka R (2008) Tempol diminishes cocaine-induced
oxidative damage and attenuates the development and expression of behavioral
sensitization. Neuroscience 155:649658.
Nutt DJ and Costello MJ (1988) Rapid induction of lorazepam dependence and reversal with flumazenil. Life Sci 43:10451053.
Nutt DJ, Glue P, Lawson C, and Wilson S (1990) Flumazenil provocation of panic
attacks. Evidence for altered benzodiazepine receptor sensitivity in panic disorder.
Arch Gen Psychiatry 47:917925.
Nutt DJ, King LA, and Nichols DE (2013) Effects of Schedule I drug laws on neuroscience research and treatment innovation. Nat Rev Neurosci 14:577585.
Nutt DJ, Lingford-Hughes A, Erritzoe D, and Stokes PR (2015) The dopamine theory
of addiction: 40 years of highs and lows. Nat Rev Neurosci 16:305312.
OCallaghan JP and Miller DB (1994) Neurotoxicity profiles of substituted amphetamines in the C57BL/6J mouse. J Pharmacol Exp Ther 270:741751.
ODell LE (2009) A psychobiological framework of the substrates that mediate nicotine use during adolescence. Neuropharmacology 56 (Suppl 1):263278.
ODell SJ, Weihmuller FB, and Marshall JF (1991) Multiple methamphetamine
injections induce marked increases in extracellular striatal dopamine which correlate with subsequent neurotoxicity. Brain Res 564:256260.
OHearn E, Battaglia G, De Souza EB, Kuhar MJ, and Molliver ME (1988) Methylenedioxyamphetamine (MDA) and methylenedioxymethamphetamine (MDMA)
cause selective ablation of serotonergic axon terminals in forebrain: immunocytochemical evidence for neurotoxicity. J Neurosci 8:27882803.
OKeefe J and Dostrovsky J (1971) The hippocampus as a spatial map. Preliminary
evidence from unit activity in the freely-moving rat. Brain Res 34:171175.
OShea E, Granados R, Esteban B, Colado MI, and Green AR (1998) The relationship
between the degree of neurodegeneration of rat brain 5-HT nerve terminals and
the dose and frequency of administration of MDMA (ecstasy). Neuropharmacology
37:919926.
Obrocki J, Buchert R, Vterlein O, Thomasius R, Beyer W, and Schiemann T (1999)
Ecstasylong-term effects on the human central nervous system revealed by positron emission tomography. Br J Psychiatry 175:186188.
Oda A, Yamagata K, Nakagomi S, Uejima H, Wiriyasermkul P, Ohgaki R, Nagamori
S, Kanai Y, and Tanaka H (2014) Nicotine induces dendritic spine remodeling in
cultured hippocampal neurons. J Neurochem 128:246255.
Ohno-Shosaku T, Maejima T, and Kano M (2001) Endogenous cannabinoids mediate
retrograde signals from depolarized postsynaptic neurons to presynaptic terminals. Neuron 29:729738.
Olale F, Gerzanich V, Kuryatov A, Wang F, and Lindstrom J (1997) Chronic nicotine
exposure differentially affects the function of human alpha3, alpha4, and alpha7
neuronal nicotinic receptor subtypes. J Pharmacol Exp Ther 283:675683.
Olasmaa M, Rothstein JD, Guidotti A, Weber RJ, Paul SM, Spector S, Zeneroli ML,
Baraldi M, and Costa E (1990) Endogenous benzodiazepine receptor ligands in
human and animal hepatic encephalopathy. J Neurochem 55:20152023.
Olesen J, Gustavsson A, Svensson M, Wittchen HU, and Jnsson B; CDBE2010 study
group; ; European Brain Council (2012) The economic cost of brain disorders in
Europe. Eur J Neurol 19:155162.
Olive MF, Koenig HN, Nannini MA, and Hodge CW (2002) Elevated extracellular
CRF levels in the bed nucleus of the stria terminalis during ethanol withdrawal
and reduction by subsequent ethanol intake. Pharmacol Biochem Behav 72:
213220.
Olivito L, Saccone P, Perri V, Bachman JL, Fragapane P, Mele A, Huganir RL,
and De Leonibus E (2014) Phosphorylation of the AMPA receptor GluA1 subunit
regulates memory load capacity. Brain Struct Funct DOI 10.1007/s00429-0140927-1
Olney JW, Ishimaru MJ, Bittigau P, and Ikonomidou C (2000) Ethanol-induced apoptotic neurodegeneration in the developing brain. Apoptosis 5:515521.
Olney JW, Labruyere J, and Price MT (1989) Pathological changes induced in cerebrocortical neurons by phencyclidine and related drugs. Science 244:13601362.
Olney JW, Labruyere J, Wang G, Wozniak DF, Price MT, and Sesma MA (1991)
NMDA antagonist neurotoxicity: mechanism and prevention. Science 254:
15151518.
Olney JW, Wozniak DF, Farber NB, Jevtovic-Todorovic V, Bittigau P,
and Ikonomidou C (2002) The enigma of fetal alcohol neurotoxicity. Ann Med 34:
109119.
Olsen RW, Hanchar HJ, Meera P, and Wallner M (2007) GABAA receptor subtypes:
the one glass of wine receptors. Alcohol 41:201209.
Olsen RW and Sieghart W (2008) International Union of Pharmacology. LXX. Subtypes of gamma-aminobutyric acid(A) receptors: classification on the basis of subunit composition, pharmacology, and function. Update. Pharmacol Rev 60:
243260.
Olsen RW and Sieghart W (2009) GABA A receptors: subtypes provide diversity of
function and pharmacology. Neuropharmacology 56:141148.
Olson VG and Nestler EJ (2007) Topographical organization of GABAergic neurons
within the ventral tegmental area of the rat. Synapse 61:8795.
Ong J and Kerr DI (2005) Clinical potential of GABAB receptor modulators. CNS
Drug Rev 11:317334.
Orr DP and Ingersoll GM (1995) The contribution of level of cognitive complexity and
pubertal timing to behavioral risk in young adolescents. Pediatrics 95:528533.
Osborne MP and Olive MF (2008) A role for mGluR5 receptors in intravenous
methamphetamine self-administration. Ann N Y Acad Sci 1139:206211.
Otis JM, Fitzgerald MK, and Mueller D (2014) Infralimbic BDNF/TrkB enhancement
of GluN2B currents facilitates extinction of a cocaine-conditioned place preference.
J Neurosci 34:60576064.
Ouellet L and de Villers-Sidani E (2014) Trajectory of the main GABAergic interneuron populations from early development to old age in the rat primary auditory
cortex. Front Neuroanat 8:40.

Drug-Induced Neuroplasticity
Ouyang H, Liu S, Zeng W, Levitt RC, Candiotti KA, and Hao S (2012) An emerging
new paradigm in opioid withdrawal: a critical role for glia-neuron signaling in the
periaqueductal gray. ScientificWorldJournal 2012:940613.
Overstreet LS, Pasternak JF, Colley PA, Slater NT, and Trommer BL (1997)
Metabotropic glutamate receptor mediated long-term depression in developing
hippocampus. Neuropharmacology 36:831844.
Overstreet LS and Westbrook GL (2001) Paradoxical reduction of synaptic inhibition
by vigabatrin. J Neurophysiol 86:596603.
Pacher P, Beckman JS, and Liaudet L (2007) Nitric oxide and peroxynitrite in health
and disease. Physiol Rev 87:315424.
Pachernegg S, Strutz-Seebohm N, and Hollmann M (2012) GluN3 subunit-containing
NMDA receptors: not just one-trick ponies. Trends Neurosci 35:240249.
Pan Y, Gerasimov MR, Kvist T, Wellendorph P, Madsen KK, Pera E, Lee H,
Schousboe A, Chebib M, and Bruner-Osborne H, et al. (2012) (1S, 3S)-3-amino-4difluoromethylenyl-1-cyclopentanoic acid (CPP-115), a potent g-aminobutyric acid
aminotransferase inactivator for the treatment of cocaine addiction. J Med Chem
55:357366.
Pandey SC, Zhang H, Ugale R, Prakash A, Xu T, and Misra K (2008) Effector
immediate-early gene arc in the amygdala plays a critical role in alcoholism.
J Neurosci 28:25892600.
Panenka WJ, Procyshyn RM, Lecomte T, MacEwan GW, Flynn SW, Honer WG,
and Barr AM (2013) Methamphetamine use: a comprehensive review of molecular,
preclinical and clinical findings. Drug Alcohol Depend 129:167179.
Panhelainen AE and Korpi ER (2012) Evidence for a role of inhibition of orexinergic
neurons in the anxiolytic and sedative effects of diazepam: A c-Fos study. Pharmacol Biochem Behav 101:115124.
Panhelainen AE, Vekovischeva OY, Aitta-Aho T, Rasanen I, Ojanper I, and Korpi
ER (2011) Diazepam-induced neuronal plasticity attenuates locomotor responses
to morphine and amphetamine challenges in mice. Neuroscience 192:312321.
Panlilio LV, Thorndike EB, and Schindler CW (2005) Lorazepam reinstates
punishment-suppressed remifentanil self-administration in rats. Psychopharmacology (Berl) 179:374382.
Panos JJ, Rademacher DJ, Renner SL, and Steinpreis RE (1999) The rewarding
properties of NMDA and MK-801 (dizocilpine) as indexed by the conditioned place
preference paradigm. Pharmacol Biochem Behav 64:591595.
Panula P and Nuutinen S (2013) The histaminergic network in the brain: basic
organization and role in disease. Nat Rev Neurosci 14:472487.
Papadopolou S, Hartmann P, Lips KS, Kummer W, and Haberberger RV (2004)
Nicotinic receptor mediated stimulation of NO-generation in neurons of rat thoracic dorsal root ganglia. Neurosci Lett 361:3235.
Parada M, Hernandez L, Schwartz D, and Hoebel BG (1988) Hypothalamic infusion
of amphetamine increases serotonin, dopamine and norepinephrine. Physiol Behav
44:607610.
Park H and Poo MM (2013) Neurotrophin regulation of neural circuit development
and function. Nat Rev Neurosci 14:723.
Parker EM and Cubeddu LX (1988) Comparative effects of amphetamine, phenylethylamine and related drugs on dopamine efflux, dopamine uptake and mazindol
binding. J Pharmacol Exp Ther 245:199210.
Parkerson HA, Zvolensky MJ, and Asmundson GJ (2013) Understanding the relationship between smoking and pain. Expert Rev Neurother 13:14071414.
Parrott AC, Lees A, Garnham NJ, Jones M, and Wesnes K (1998) Cognitive performance in recreational users of MDMA of ecstasy: evidence for memory deficits.
J Psychopharmacol 12:7983.
Parsegian A and See RE (2014) Dysregulation of dopamine and glutamate release in
the prefrontal cortex and nucleus accumbens following methamphetamine selfadministration and during reinstatement in rats. Neuropsychopharmacology 39:
811822.
Pascoli V, Terrier J, Espallergues J, Valjent E, OConnor EC, and Lscher C (2014)
Contrasting forms of cocaine-evoked plasticity control components of relapse. Nature 509:459464.
Pascoli V, Turiault M, and Lscher C (2012) Reversal of cocaine-evoked synaptic
potentiation resets drug-induced adaptive behaviour. Nature 481:7175.
Passie T, Halpern JH, Stichtenoth DO, Emrich HM, and Hintzen A (2008) The
pharmacology of lysergic acid diethylamide: a review. CNS Neurosci Ther 14:
295314.
Passie T, Seifert J, Schneider U, and Emrich HM (2002) The pharmacology of psilocybin. Addict Biol 7:357364.
Pasumarthi RK, Reznikov LR, and Fadel J (2006) Activation of orexin neurons by
acute nicotine. Eur J Pharmacol 535:172176.
Patel MB, Patel CN, Rajashekara V, and Yoburn BC (2002) Opioid agonists differentially regulate mu-opioid receptors and trafficking proteins in vivo. Mol Pharmacol 62:14641470.
Patel PN and Ezzo DC (2009) Withdrawal symptoms after discontinuation of
transdermal scopolamine therapy: treatment with meclizine. Am J Health Syst
Pharm 66:20242026.
Patel S, Kingsley PJ, Mackie K, Marnett LJ, and Winder DG (2009) Repeated
homotypic stress elevates 2-arachidonoylglycerol levels and enhances short-term
endocannabinoid signaling at inhibitory synapses in basolateral amygdala. Neuropsychopharmacology 34:26992709.
Paton GS, Pertwee RG, and Davies SN (1998) Correlation between cannabinoid
mediated effects on paired pulse depression and induction of long term potentiation in the rat hippocampal slice. Neuropharmacology 37:11231130.
Patrick KS, Caldwell RW, Ferris RM, and Breese GR (1987) Pharmacology of the
enantiomers of threo-methylphenidate. J Pharmacol Exp Ther 241:152158.
Patrick KS and Markowitz JS (1997) Pharmacology of methylphenidate, amphetamine enantiomers and pemoline in attention-deficit hyperactivity disorder. Hum
Psychopharmacol Clin Exp 12:527546.
Paul J, Yvenes GE, Benke D, Di Lio A, Ralvenius WT, Witschi R, Scheurer L, Cook
JM, Rudolph U, and Fritschy JM, et al. (2014a) Antihyperalgesia by a2-GABAA

993

receptors occurs via a genuine spinal action and does not involve supraspinal sites.
Neuropsychopharmacology 39:477487.
Paul M, Dewey SL, Gardner EL, Brodie JD, and Ashby CR Jr (2001) Gamma-vinyl
GABA (GVG) blocks expression of the conditioned place preference response to
heroin in rats. Synapse 41:219220.
Paul RK, Singh NS, Khadeer M, Moaddel R, Sanghvi M, Green CE, OLoughlin K,
Torjman MC, Bernier M, and Wainer IW (2014b) (R,S)-Ketamine metabolites (R,
S)-norketamine and (2S,6S)-hydroxynorketamine increase the mammalian target
of rapamycin function. Anesthesiology 121:149159.
Pavlov I, Lauri S, Taira T, and Rauvala H (2004) The role of ECM molecules in
activity-dependent synaptic development and plasticity. Birth Defects Res C Embryo Today 72:1224.
Peng X, Gerzanich V, Anand R, Whiting PJ, and Lindstrom J (1994) Nicotine-induced
increase in neuronal nicotinic receptors results from a decrease in the rate of
receptor turnover. Mol Pharmacol 46:523530.
Peng XQ, Li X, Gilbert JG, Pak AC, Ashby CR Jr, Brodie JD, Dewey SL, Gardner EL,
and Xi ZX (2008) Gamma-vinyl GABA inhibits cocaine-triggered reinstatement of
drug-seeking behavior in rats by a non-dopaminergic mechanism. Drug Alcohol
Depend 97:216225.
Pennick M (2010) Absorption of lisdexamfetamine dimesylate and its enzymatic
conversion to d-amphetamine. Neuropsychiatr Dis Treat 6:317327.
Peoples RW, Li C, and Weight FF (1996) Lipid vs protein theories of alcohol action in
the nervous system. Annu Rev Pharmacol Toxicol 36:185201.
Peraile I, Torres E, Mayado A, Izco M, Lopez-Jimenez A, Lopez-Moreno JA, Colado
MI, and OShea E (2010) Dopamine transporter down-regulation following repeated cocaine: implications for 3,4-methylenedioxymethamphetamine-induced
acute effects and long-term neurotoxicity in mice. Br J Pharmacol 159:201211.
Perez MF, Gabach LA, Almiron RS, Carlini VP, De Barioglio SR, and Ramirez OA
(2010) Different chronic cocaine administration protocols induce changes on dentate gyrus plasticity and hippocampal dependent behavior. Synapse 64:742753.
Perfetti X, OMathna B, Pizarro N, Cuys E, Khymenets O, Almeida B, Pellegrini
M, Pichini S, Lau SS, and Monks TJ, et al. (2009) Neurotoxic thioether adducts of
3,4-methylenedioxymethamphetamine identified in human urine after ecstasy ingestion. Drug Metab Dispos 37:14481455.
Pericic D, Lazic J, Jembrek MJ, Strac DS, and Rajcan I (2004) Chronic exposure of
cells expressing recombinant GABAA receptors to benzodiazepine antagonist flumazenil enhances the maximum number of benzodiazepine binding sites. Life Sci
76:303317.
Perkins DI, Trudell JR, Crawford DK, Alkana RL, and Davies DL (2010) Molecular
targets and mechanisms for ethanol action in glycine receptors. Pharmacol Ther
127:5365.
Perona MT, Waters S, Hall FS, Sora I, Lesch KP, Murphy DL, Caron M, and Uhl GR
(2008) Animal models of depression in dopamine, serotonin, and norepinephrine
transporter knockout mice: prominent effects of dopamine transporter deletions.
Behav Pharmacol 19:566574.
Perovic M, Tesic V, Mladenovic Djordjevic A, Smiljanic K, Loncarevic-Vasiljkovic N,
Ruzdijic S, and Kanazir S (2013) BDNF transcripts, proBDNF and proNGF, in the
cortex and hippocampus throughout the life span of the rat. Age (Dordr) 35:
20572070.
Perrotti LI, Bolaos CA, Choi KH, Russo SJ, Edwards S, Ulery PG, Wallace DL, Self
DW, Nestler EJ, and Barrot M (2005) DeltaFosB accumulates in a GABAergic cell
population in the posterior tail of the ventral tegmental area after psychostimulant
treatment. Eur J Neurosci 21:28172824.
Perry DC, Dvila-Garca MI, Stockmeier CA, and Kellar KJ (1999) Increased nicotinic receptors in brains from smokers: membrane binding and autoradiography
studies. J Pharmacol Exp Ther 289:15451552.
Perry DC, Mao D, Gold AB, McIntosh JM, Pezzullo JC, and Kellar KJ (2007) Chronic
nicotine differentially regulates alpha6- and beta3-containing nicotinic cholinergic
receptors in rat brain. J Pharmacol Exp Ther 322:306315.
Pertwee RG (2008) Ligands that target cannabinoid receptors in the brain: from THC
to anandamide and beyond. Addict Biol 13:147159.
Pertwee RG (2012) Targeting the endocannabinoid system with cannabinoid receptor
agonists: pharmacological strategies and therapeutic possibilities. Philos Trans R
Soc Lond B Biol Sci 367:33533363.
Pertwee RG, Howlett AC, Abood ME, Alexander SP, Di Marzo V, Elphick MR,
Greasley PJ, Hansen HS, Kunos G, and Mackie K, et al. (2010) International Union
of Basic and Clinical Pharmacology. LXXIX. Cannabinoid receptors and their
ligands: beyond CB and CB. Pharmacol Rev 62:588631.
Pertwee RG and Ross RA (2002) Cannabinoid receptors and their ligands. Prostaglandins Leukot Essent Fatty Acids 66:101121.

Petanjek Z, Judas M, Simic
G, Rasin MR, Uylings HB, Rakic P, and Kostovic I (2011)
Extraordinary neoteny of synaptic spines in the human prefrontal cortex. Proc Natl
Acad Sci USA 108:1328113286.
Peters DA and Tang S (1977) Persistent effects of repeated injections of D-lysergic
acid diethylamide on rat brain 5-hydroxytryptamine and 5-hydroxyindoleacetic
acid levels. Biochem Pharmacol 26:10851086.
Peterson BS, Warner V, Bansal R, Zhu H, Hao X, Liu J, Durkin K, Adams PB,
Wickramaratne P, and Weissman MM (2009) Cortical thinning in persons at increased familial risk for major depression. Proc Natl Acad Sci USA 106:62736278.
Petroff OA, Rothman DL, Behar KL, and Mattson RH (1996) Human brain GABA
levels rise after initiation of vigabatrin therapy but fail to rise further with increasing dose. Neurology 46:14591463.
Piazza PV, Deminire JM, Le Moal M, and Simon H (1989) Factors that predict
individual vulnerability to amphetamine self-administration. Science 245:
15111513.
Piazza PV, Deroche V, Deminire JM, Maccari S, Le Moal M, and Simon H (1993)
Corticosterone in the range of stress-induced levels possesses reinforcing properties: implications for sensation-seeking behaviors. Proc Natl Acad Sci USA 90:
1173811742.

994

Korpi et al.

Picciotto MR, Addy NA, Mineur YS, and Brunzell DH (2008) It is not either/or:
activation and desensitization of nicotinic acetylcholine receptors both contribute
to behaviors related to nicotine addiction and mood. Prog Neurobiol 84:329342.
Pich EM, Pagliusi SR, Tessari M, Talabot-Ayer D, Hooft van Huijsduijnen R,
and Chiamulera C (1997) Common neural substrates for the addictive properties of
nicotine and cocaine. Science 275:8386.
Pickens R and Harris WC (1968) Self-administration of d-amphetamine by rats.
Psychopharmacology (Berl) 12:158163.
Pierce RC, Meil WM, and Kalivas PW (1997) The NMDA antagonist, dizocilpine,
enhances cocaine reinforcement without influencing mesoaccumbens dopamine
transmission. Psychopharmacology (Berl) 133:188195.
Pietil K, Lhde T, Attila M, Ahtee L, and Nordberg A (1998) Regulation of nicotinic
receptors in the brain of mice withdrawn from chronic oral nicotine treatment.
Naunyn Schmiedebergs Arch Pharmacol 357:176182.
Pin JP and Przeau L (2007) Allosteric modulators of GABA(B) receptors: mechanism
of action and therapeutic perspective. Curr Neuropharmacol 5:195201.
Piper BJ and Meyer JS (2004) Memory deficit and reduced anxiety in young adult
rats given repeated intermittent MDMA treatment during the periadolescent period. Pharmacol Biochem Behav 79:723731.
Pistis M, Muntoni AL, Pillolla G, Perra S, Cignarella G, Melis M, and Gessa GL
(2005) Gamma-hydroxybutyric acid (GHB) and the mesoaccumbens reward circuit:
evidence for GABA(B) receptor-mediated effects. Neuroscience 131:465474.
Pitknen A, Tuunanen J, and Halonen T (1996) Vigabatrin and carbamazepine have
different efficacies in the prevention of status epilepticus induced neuronal damage
in the hippocampus and amygdala. Epilepsy Res 24:2945.
Pitler TA and Alger BE (1994) Depolarization-induced suppression of GABAergic
inhibition in rat hippocampal pyramidal cells: G protein involvement in a presynaptic mechanism. Neuron 13:14471455.
Pizzorusso T, Medini P, Berardi N, Chierzi S, Fawcett JW, and Maffei L (2002)
Reactivation of ocular dominance plasticity in the adult visual cortex. Science 298:
12481251.
Planert H, Berger TK, and Silberberg G (2013) Membrane properties of striatal direct
and indirect pathway neurons in mouse and rat slices and their modulation by
dopamine. PLoS One 8:e57054.
Polesskaya OO, Fryxell KJ, Merchant AD, Locklear LL, Ker KF, McDonald CG,
Eppolito AK, Smith LN, Wheeler TL, and Smith RF (2007) Nicotine causes agedependent changes in gene expression in the adolescent female rat brain. Neurotoxicol Teratol 29:126140.
Pollard M, Varin C, Hrupka B, Pemberton DJ, Steckler T, and Shaban H (2012)
Synaptic transmission changes in fear memory circuits underlie key features of an
animal model of schizophrenia. Behav Brain Res 227:184193.
Pontn E, Viberg H, Gordh T, Eriksson P, and Fredriksson A (2012) Clonidine
abolishes the adverse effects on apoptosis and behaviour after neonatal ketamine
exposure in mice. Acta Anaesthesiol Scand 56:10581065.
Pontieri FE, Tanda G, Orzi F, and Di Chiara G (1996) Effects of nicotine on the
nucleus accumbens and similarity to those of addictive drugs. see comments Nature 382:255257.
Pope HG Jr, Gruber AJ, Hudson JI, Huestis MA, and Yurgelun-Todd D (2002)
Cognitive measures in long-term cannabis users. J Clin Pharmacol 42(11, Suppl)
41S47S.
Popik P, Kos T, Sowa-Kucma M, and Nowak G (2008) Lack of persistent effects of
ketamine in rodent models of depression. Psychopharmacology (Berl) 198:421430.
Popova NK, Vishnivetskaya GB, Ivanova EA, Skrinskaya JA, and Seif I (2000) Altered behavior and alcohol tolerance in transgenic mice lacking MAO A: a comparison with effects of MAO A inhibitor clorgyline. Pharmacol Biochem Behav 67:
719727.
Porrino LJ, Esposito RU, Seeger TF, Crane AM, Pert A, and Sokoloff L (1984) Metabolic mapping of the brain during rewarding self-stimulation. Science 224:
306309.
Porrino LJ, Lyons D, Smith HR, Daunais JB, and Nader MA (2004) Cocaine selfadministration produces a progressive involvement of limbic, association, and
sensorimotor striatal domains. J Neurosci 24:35543562.
Portugal GS, Al-Hasani R, Fakira AK, Gonzalez-Romero JL, Melyan Z, McCall JG,
Bruchas MR, and Morn JA (2014) Hippocampal long-term potentiation is disrupted during expression and extinction but is restored after reinstatement of
morphine place preference. J Neurosci 34:527538.
Pozzi L, Baviera M, Sacchetti G, Calcagno E, Balducci C, Invernizzi RW, and Carli M
(2011) Attention deficit induced by blockade of N-methyl D-aspartate receptors in
the prefrontal cortex is associated with enhanced glutamate release and cAMP
response element binding protein phosphorylation: role of metabotropic glutamate
receptors 2/3. Neuroscience 176:336348.
Pressler R and Auvin S (2013) Comparison of brain maturation among species: an
example in translational research suggesting the possible use of bumetanide in
newborn. Front Neurol 4:36. DOI 10.3389/fneur.2013.00036
Pristupa ZB, McConkey F, Liu F, Man HY, Lee FJ, Wang YT, and Niznik HB (1998)
Protein kinase-mediated bidirectional trafficking and functional regulation of the
human dopamine transporter. Synapse 30:7987.
Pritchett DB, Sontheimer H, Shivers BD, Ymer S, Kettenmann H, Schofield PR,
and Seeburg PH (1989) Importance of a novel GABAA receptor subunit for benzodiazepine pharmacology. Nature 338:582585.
Przewocka B, Turchan J, Laso
n W, and Przewocki R (1997) Ethanol withdrawal
enhances the prodynorphin system activity in the rat nucleus accumbens. Neurosci
Lett 238:1316.
Pu L, Bao GB, Xu NJ, Ma L, and Pei G (2002) Hippocampal long-term potentiation is
reduced by chronic opiate treatment and can be restored by re-exposure to opiates.
J Neurosci 22:19141921.
Puerta E, Hervias I, Goi-Allo B, Zhang SF, Jordn J, Starkov AA, and Aguirre N
(2010) Methylenedioxymethamphetamine inhibits mitochondrial complex I activity in mice: a possible mechanism underlying neurotoxicity. Br J Pharmacol 160:
233245.

Pulipparacharuvil S, Renthal W, Hale CF, Taniguchi M, Xiao G, Kumar A, Russo SJ,


Sikder D, Dewey CM, and Davis MM, et al. (2008) Cocaine regulates MEF2 to
control synaptic and behavioral plasticity. Neuron 59:621633.
Purgianto A, Scheyer AF, Loweth JA, Ford KA, Tseng KY, and Wolf ME (2013)
Different adaptations in AMPA receptor transmission in the nucleus accumbens
after short vs long access cocaine self-administration regimens. Neuropsychopharmacology 38:17891797.
Puustinen J, Nurminen J, Kukola M, Vahlberg T, Laine K, and Kivel SL (2007)
Associations between use of benzodiazepines or related drugs and health, physical
abilities and cognitive function: a non-randomised clinical study in the elderly.
Drugs Aging 24:10451059.
Qi ZH, Song M, Wallace MJ, Wang D, Newton PM, McMahon T, Chou WH, Zhang C,
Shokat KM, and Messing RO (2007) Protein kinase C epsilon regulates gammaaminobutyrate type A receptor sensitivity to ethanol and benzodiazepines through
phosphorylation of gamma2 subunits. J Biol Chem 282:3305233063.
Qin C and Luo M (2009) Neurochemical phenotypes of the afferent and efferent
projections of the mouse medial habenula. Neuroscience 161:827837.
Quednow BB, Kometer M, Geyer MA, and Vollenweider FX (2012) Psilocybin-induced
deficits in automatic and controlled inhibition are attenuated by ketanserin in
healthy human volunteers. Neuropsychopharmacology 37:630640.
Quertemont E (2004) Genetic polymorphism in ethanol metabolism: acetaldehyde
contribution to alcohol abuse and alcoholism. Mol Psychiatry 9:570581.
Quesseveur G, David DJ, Gaillard MC, Pla P, Wu MV, Nguyen HT, Nicolas V,
Auregan G, David I, and Dranovsky A, et al. (2013) BDNF overexpression in mouse
hippocampal astrocytes promotes local neurogenesis and elicits anxiolytic-like activities. Transl Psychiatry 3:e253.
Quimby KL, Aschkenase LJ, Bowman RE, Katz J, and Chang LW (1974) Enduring
learning deficits and cerebral synaptic malformation from exposure to 10 parts of
halothane per million. Science 185:625627.
Quintanilla ME, Tampier L, Karahanian E, Rivera-Meza M, Herrera-Marschitz M,
and Israel Y (2012) Reward and relapse: complete gene-induced dissociation in an
animal model of alcohol dependence. Alcohol Clin Exp Res 36:517522.
Rabinoff M, Caskey N, Rissling A, and Park C (2007) Pharmacological and chemical
effects of cigarette additives. Am J Public Health 97:19811991.
Radonjic NV, Jakovcevski I, Bumbasirevic V, and Petronijevic ND (2013) Perinatal
phencyclidine administration decreases the density of cortical interneurons and
increases the expression of neuregulin-1. Psychopharmacology (Berl) 227:673683.
Rahbar F, Fomufod A, White D, and Westney LS (1993) Impact of intrauterine exposure to phencyclidine (PCP) and cocaine on neonates. J Natl Med Assoc 85:
349352.
Rahman S, Zhang J, and Corrigall WA (2003) Effects of acute and chronic nicotine on
somatodendritic dopamine release of the rat ventral tegmental area: in vivo
microdialysis study. Neurosci Lett 348:6164.
Rajkumar R, Suri S, Deng HM, and Dawe GS (2013) Nicotine and clozapine crossprime the locus coeruleus noradrenergic system to induce long-lasting potentiation
in the rat hippocampus. Hippocampus 23:616624.
Ramirez OA and Wang RY (1986) Locus coeruleus norepinephrine-containing neurons: effects produced by acute and subchronic treatment with antipsychotic drugs
and amphetamine. Brain Res 362:165170.
Rassnick S, Heinrichs SC, Britton KT, and Koob GF (1993) Microinjection of a corticotropin-releasing factor antagonist into the central nucleus of the amygdala
reverses anxiogenic-like effects of ethanol withdrawal. Brain Res 605:2532.
Ratano P, Everitt BJ, and Milton AL (2014) The CB1 receptor antagonist AM251
impairs reconsolidation of pavlovian fear memory in the rat basolateral amygdala.
Neuropsychopharmacology 39:25292537.
Ravindran J, Blumbergs P, Crompton J, Pietris G, and Waddy H (2001) Visual field
loss associated with vigabatrin: pathological correlations. J Neurol Neurosurg
Psychiatry 70:787789.
Ray TS (2010) Psychedelics and the human receptorome. PLoS One 5:e9019.
Realini N, Vigano D, Guidali C, Zamberletti E, Rubino T, and Parolaro D (2011)
Chronic URB597 treatment at adulthood reverted most depressive-like symptoms
induced by adolescent exposure to THC in female rats. Neuropharmacology 60:
235243.
Redondo RL, Kim J, Arons AL, Ramirez S, Liu X, and Tonegawa S (2014) Bidirectional switch of the valence associated with a hippocampal contextual memory
engram. Nature 513:426430.
Reese EA, Norimatsu Y, Grandy MS, Suchland KL, Bunzow JR, and Grandy DK (2014)
Exploring the determinants of trace amine-associated receptor 1s functional selectivity
for the stereoisomers of amphetamine and methamphetamine. J Med Chem 57:378390.
Reisine T and Bell GI (1993) Molecular biology of opioid receptors. Trends Neurosci
16:506510.
Reith ME, Meisler BE, and Lajtha A (1985) Locomotor effects of cocaine, cocaine
congeners, and local anesthetics in mice. Pharmacol Biochem Behav 23:831836.
Renard J, Krebs MO, Jay TM, and Le Pen G (2013) Long-term cognitive impairments
induced by chronic cannabinoid exposure during adolescence in rats: a strain
comparison. Psychopharmacology (Berl) 225:781790.
Renard J, Krebs MO, Le Pen G, and Jay TM (2014) Long-term consequences of
adolescent cannabinoid exposure in adult psychopathology. Front Neurosci 8:361.
DOI 10.3389/fnins.2014.00361
Rendell PG, Mazur M, and Henry JD (2009) Prospective memory impairment in
former users of methamphetamine. Psychopharmacology (Berl) 203:609616.
Reneman L, Booij J, de Bruin K, Reitsma JB, de Wolff FA, Gunning WB, den Heeten
GJ, and van den Brink W (2001a) Effects of dose, sex, and long-term abstention
from use on toxic effects of MDMA (ecstasy) on brain serotonin neurons. Lancet
358:18641869.
Reneman L, Booij J, Schmand B, van den Brink W, and Gunning B (2000) Memory
disturbances in Ecstasy users are correlated with an altered brain serotonin
neurotransmission. Psychopharmacology (Berl) 148:322324.

Drug-Induced Neuroplasticity
Reneman L, Endert E, de Bruin K, Lavalaye J, Feenstra MG, de Wolff FA, and Booij
J (2002) The acute and chronic effects of MDMA (ecstasy) on cortical 5-HT2A
receptors in rat and human brain. Neuropsychopharmacology 26:387396.
Reneman L, Majoie CB, Schmand B, van den Brink W, and den Heeten GJ (2001b)
Prefrontal N-acetylaspartate is strongly associated with memory performance in
(abstinent) ecstasy users: preliminary report. Biol Psychiatry 50:550554.
Renthal W, Maze I, Krishnan V, Covington HE 3rd, Xiao G, Kumar A, Russo SJ,
Graham A, Tsankova N, and Kippin TE, et al. (2007) Histone deacetylase 5 epigenetically controls behavioral adaptations to chronic emotional stimuli. Neuron
56:517529.
Rus GZ, Vieira FG, Abelaira HM, Michels M, Tomaz DB, dos Santos MA, Carlessi
AS, Neotti MV, Matias BI, and Luz JR, et al. (2014) MAPK signaling correlates
with the antidepressant effects of ketamine. J Psychiatr Res 55:1521.
Reynolds LM, Engin E, Tantillo G, Lau HM, Muschamp JW, Carlezon WA Jr,
and Rudolph U (2012) Differential roles of GABA(A) receptor subtypes in
benzodiazepine-induced enhancement of brain-stimulation reward. Neuropsychopharmacology 37:25312540.
Rhodes JS, Best K, Belknap JK, Finn DA, and Crabbe JC (2005) Evaluation of
a simple model of ethanol drinking to intoxication in C57BL/6J mice. Physiol Behav
84:5363.
Ribeiro PO, Tom AR, Silva HB, Cunha RA, and Antunes LM (2014) Clinically relevant concentrations of ketamine mainly affect long-term potentiation rather than
basal excitatory synaptic transmission and do not change paired-pulse facilitation
in mouse hippocampal slices. Brain Res 1560:1017.
Ricaurte GA, Guillery RW, Seiden LS, Schuster CR, and Moore RY (1982) Dopamine
nerve terminal degeneration produced by high doses of methylamphetamine in the
rat brain. Brain Res 235:93103.
Ricaurte GA, Schuster CR, and Seiden LS (1980) Long-term effects of repeated
methylamphetamine administration on dopamine and serotonin neurons in the rat
brain: a regional study. Brain Res 193:153163.
Ricci V, Martinotti G, Gelfo F, Tonioni F, Caltagirone C, Bria P, and Angelucci F
(2011) Chronic ketamine use increases serum levels of brain-derived neurotrophic
factor. Psychopharmacology (Berl) 215:143148.
Richards M, Jarvis MJ, Thompson N, and Wadsworth ME (2003) Cigarette smoking
and cognitive decline in midlife: evidence from a prospective birth cohort study. Am
J Public Health 93:994998.
Risinger RC, Salmeron BJ, Ross TJ, Amen SL, Sanfilipo M, Hoffmann RG, Bloom AS,
Garavan H, and Stein EA (2005) Neural correlates of high and craving during
cocaine self-administration using BOLD fMRI. Neuroimage 26:10971108.
Ritz MC and Kuhar MJ (1989) Relationship between self-administration of amphetamine and monoamine receptors in brain: comparison with cocaine. J Pharmacol Exp Ther 248:10101017.
Robbe D, Bockaert J, and Manzoni OJ (2002) Metabotropic glutamate receptor 2/3dependent long-term depression in the nucleus accumbens is blocked in morphine
withdrawn mice. Eur J Neurosci 16:22312235.
Robbins TW and Everitt BJ (1999) Drug addiction: bad habits add up. Nature 398:
567570.
Robbins TW, McAlonan G, Muir JL, and Everitt BJ (1997) Cognitive enhancers in
theory and practice: studies of the cholinergic hypothesis of cognitive deficits in
Alzheimers disease. Behav Brain Res 83:1523.
Roberto M, Cruz M, Bajo M, Siggins GR, Parsons LH, and Schweitzer P (2010a) The
endocannabinoid system tonically regulates inhibitory transmission and depresses
the effect of ethanol in central amygdala. Neuropsychopharmacology 35:
19621972.
Roberto M, Cruz MT, Gilpin NW, Sabino V, Schweitzer P, Bajo M, Cottone P,
Madamba SG, Stouffer DG, and Zorrilla EP, et al. (2010b) Corticotropin releasing
factor-induced amygdala gamma-aminobutyric Acid release plays a key role in
alcohol dependence. Biol Psychiatry 67:831839.
Roberto M, Madamba SG, Moore SD, Tallent MK, and Siggins GR (2003) Ethanol
increases GABAergic transmission at both pre- and postsynaptic sites in rat central amygdala neurons. Proc Natl Acad Sci USA 100:20532058.
Roberto M, Madamba SG, Stouffer DG, Parsons LH, and Siggins GR (2004a) Increased GABA release in the central amygdala of ethanol-dependent rats. J Neurosci 24:1015910166.
Roberto M, Schweitzer P, Madamba SG, Stouffer DG, Parsons LH, and Siggins GR
(2004b) Acute and chronic ethanol alter glutamatergic transmission in rat central
amygdala: an in vitro and in vivo analysis. J Neurosci 24:15941603.
Roberto M and Siggins GR (2006) Nociceptin/orphanin FQ presynaptically decreases
GABAergic transmission and blocks the ethanol-induced increase of GABA release
in central amygdala. Proc Natl Acad Sci USA 103:97159720.
Roberts AJ, Cole M, and Koob GF (1996) Intra-amygdala muscimol decreases operant
ethanol self-administration in dependent rats. Alcohol Clin Exp Res 20:12891298.
Roberts DC (2005) Preclinical evidence for GABAB agonists as a pharmacotherapy
for cocaine addiction. Physiol Behav 86:1820.
Robertson SD, Matthies HJ, and Galli A (2009) A closer look at amphetamineinduced reverse transport and trafficking of the dopamine and norepinephrine
transporters. Mol Neurobiol 39:7380.
Robinson JB (1985) Stereoselectivity and isoenzyme selectivity of monoamine oxidase inhibitors. Enantiomers of amphetamine, N-methylamphetamine and deprenyl. Biochem Pharmacol 34:41054108.
Robinson TE and Berridge KC (1993) The neural basis of drug craving: an incentivesensitization theory of addiction. Brain Res Brain Res Rev 18:247291.
Robinson TE and Berridge KC (2000) The psychology and neurobiology of addiction:
an incentive-sensitization view. Addiction 95 (Suppl 2):S91S117.
Robinson TE, Gorny G, Mitton E, and Kolb B (2001) Cocaine self-administration
alters the morphology of dendrites and dendritic spines in the nucleus accumbens
and neocortex. Synapse 39:257266.
Robinson TE, Gorny G, Savage VR, and Kolb B (2002) Widespread but regionally
specific effects of experimenter- versus self-administered morphine on dendritic

995

spines in the nucleus accumbens, hippocampus, and neocortex of adult rats. Synapse 46:271279.
Robinson TE and Kolb B (1999a) Alterations in the morphology of dendrites and
dendritic spines in the nucleus accumbens and prefrontal cortex following repeated
treatment with amphetamine or cocaine. Eur J Neurosci 11:15981604.
Robinson TE and Kolb B (1999b) Morphine alters the structure of neurons in the
nucleus accumbens and neocortex of rats. Synapse 33:160162.
Robison AJ and Nestler EJ (2011) Transcriptional and epigenetic mechanisms of
addiction. Nat Rev Neurosci 12:623637.
Robison AJ, Vialou V, Mazei-Robison M, Feng J, Kourrich S, Collins M, Wee S, Koob
G, Turecki G, and Neve R, et al. (2013) Behavioral and structural responses to
chronic cocaine require a feedforward loop involving DFosB and calcium/
calmodulin-dependent protein kinase II in the nucleus accumbens shell. J Neurosci 33:42954307.
Robson MJ, Seminerio MJ, McCurdy CR, Coop A, and Matsumoto RR(2013a) s Receptor antagonist attenuation of methamphetamine-induced neurotoxicity is correlated to body temperature modulation. Pharmacol Rep 65:343349.
Robson MJ, Turner RC, Naser ZJ, McCurdy CR, Huber JD, and Matsumoto RR
(2013b) SN79, a sigma receptor ligand, blocks methamphetamine-induced microglial activation and cytokine upregulation. Exp Neurol 247:134142.
Roche KW, OBrien RJ, Mammen AL, Bernhardt J, and Huganir RL (1996) Characterization of multiple phosphorylation sites on the AMPA receptor GluR1 subunit. Neuron 16:11791188.
Rodrguez de Fonseca F, Ramos JA, Bonnin A, and Fernndez-Ruiz JJ (1993) Presence of cannabinoid binding sites in the brain from early postnatal ages. Neuroreport 4:135138.
Fernandez-Espejo E and Rodriguez-Espinosa N (2011) Psychostimulant Drugs and
Neuroplasticity. Pharmaceuticals 4:976991.
Rodriguez-Menchaca AA, Solis E Jr, Cameron K, and De Felice LJ (2012)
S(+)amphetamine induces a persistent leak in the human dopamine transporter: molecular stent hypothesis. Br J Pharmacol 165:27492757.
Rogge GA, Singh H, Dang R, and Wood MA (2013) HDAC3 is a negative regulator of
cocaine-context-associated memory formation. J Neurosci 33:66236632.
Romero J, Garcia-Palomero E, Berrendero F, Garcia-Gil L, Hernandez ML, Ramos
JA, and Fernndez-Ruiz JJ (1997) Atypical location of cannabinoid receptors in
white matter areas during rat brain development. Synapse 26:317323.
Romieu P, Deschatrettes E, Host L, Gobaille S, Sandner G, and Zwiller J (2011) The
inhibition of histone deacetylases reduces the reinstatement of cocaine-seeking
behavior in rats. Curr Neuropharmacol 9:2125.
Romieu P, Host L, Gobaille S, Sandner G, Aunis D, and Zwiller J (2008) Histone
deacetylase inhibitors decrease cocaine but not sucrose self-administration in rats.
J Neurosci 28:93429348.
Roncari G, Timm U, Zell M, Zumbrunnen R, and Weber W (1993) Flumazenil kinetics
in the elderly. Eur J Clin Pharmacol 45:585587.
Rosenberg HC, Tietz EI, and Chiu TH (1985) Tolerance to the anticonvulsant action
of benzodiazepines. Relationship to decreased receptor density. Neuropharmacology 24:639644.
Rosin A, Lindholm S, Franck J, and Georgieva J (1999) Downregulation of kappa
opioid receptor mRNA levels by chronic ethanol and repetitive cocaine in rat
ventral tegmentum and nucleus accumbens. Neurosci Lett 275:14.
Ross EJ, Graham DL, Money KM, and Stanwood GD (2015) Developmental consequences of fetal exposure to drugs: what we know and what we still must learn.
Neuropsychopharmacology 40:6187.
Rossetti ZL and Carboni S (1995) Ethanol withdrawal is associated with increased
extracellular glutamate in the rat striatum. Eur J Pharmacol 283:177183.
Roth BL, Baner K, Westkaemper R, Siebert D, Rice KC, Steinberg S, Ernsberger P,
and Rothman RB (2002) Salvinorin A: a potent naturally occurring nonnitrogenous
kappa opioid selective agonist. Proc Natl Acad Sci USA 99:1193411939.
Roth BL, Gibbons S, Arunotayanun W, Huang XP, Setola V, Treble R, and Iversen L
(2013) The ketamine analogue methoxetamine and 3- and 4-methoxy analogues of
phencyclidine are high affinity and selective ligands for the glutamate NMDA
receptor. PLoS One 8:e59334.
Rothhut B, Romano SJ, Vijayaraghavan S, and Berg DK (1996) Post-translational
regulation of neuronal acetylcholine receptors stably expressed in a mouse fibroblast cell line. J Neurobiol 29:115125.
Rothman RB, Baumann MH, Dersch CM, Romero DV, Rice KC, Carroll FI,
and Partilla JS (2001) Amphetamine-type central nervous system stimulants release norepinephrine more potently than they release dopamine and serotonin.
Synapse 39:3241.
Rothstein JD, McKhann G, Guarneri P, Barbaccia ML, Guidotti A, and Costa E
(1989) Cerebrospinal fluid content of diazepam binding inhibitor in chronic hepatic
encephalopathy. Ann Neurol 26:5762.
Rowley HL, Kulkarni R, Gosden J, Brammer R, Hackett D, and Heal DJ (2012)
Lisdexamfetamine and immediate release d-amfetamine - differences in
pharmacokinetic/pharmacodynamic relationships revealed by striatal microdialysis in freely-moving rats with simultaneous determination of plasma drug
concentrations and locomotor activity. Neuropharmacology 63:10641074.
Rozas C, Loyola S, Ugarte G, Zeise ML, Reyes-Parada M, Pancetti F, Rojas P,
and Morales B (2012) Acutely applied MDMA enhances long-term potentiation in
rat hippocampus involving D1/D5 and 5-HT2 receptors through a polysynaptic
mechanism. Eur Neuropsychopharmacol 22:584595.
Rubino T, Forlani G, Vigan D, Zippel R, and Parolaro D (2004) Modulation of extracellular signal-regulated kinases cascade by chronic delta 9-tetrahydrocannabinol treatment. Mol Cell Neurosci 25:355362.
Rubino T, Forlani G, Vigan D, Zippel R, and Parolaro D (2005) Ras/ERK signalling in
cannabinoid tolerance: from behaviour to cellular aspects. J Neurochem 93:984991.
Rubino T and Parolaro D (2008) Long lasting consequences of cannabis exposure in
adolescence. Mol Cell Endocrinol 286(1-2, Suppl 1)S108S113.
Rubino T, Realini N, Braida D, Alberio T, Capurro V, Vigan D, Guidali C, Sala M,
Fasano M, and Parolaro D (2009a) The depressive phenotype induced in adult

996

Korpi et al.

female rats by adolescent exposure to THC is associated with cognitive impairment


and altered neuroplasticity in the prefrontal cortex. Neurotox Res 15:291302.
Rubino T, Realini N, Braida D, Guidi S, Capurro V, Vigan D, Guidali C, Pinter M,
Sala M, and Bartesaghi R, et al. (2009b) Changes in hippocampal morphology and
neuroplasticity induced by adolescent THC treatment are associated with cognitive
impairment in adulthood. Hippocampus 19:763772.
Rubino T, Vigano D, Realini N, Guidali C, Braida D, Capurro V, Castiglioni C,
Cherubino F, Romualdi P, and Candeletti S, et al. (2008) Chronic delta 9-tetrahydrocannabinol during adolescence provokes sex-dependent changes in the emotional
profile in adult rats: behavioral and biochemical correlates. Neuropsychopharmacology
33:27602771.
Rudnick G and Wall SC (1992) The molecular mechanism of ecstasy [3,4methylenedioxy-methamphetamine (MDMA)]: serotonin transporters are targets
for MDMA-induced serotonin release. Proc Natl Acad Sci USA 89:18171821.
Rudolph U and Antkowiak B (2004) Molecular and neuronal substrates for general
anaesthetics. Nat Rev Neurosci 5:709720.
Rupniak NM, Steventon MJ, Field MJ, Jennings CA, and Iversen SD (1989) Comparison of the effects of four cholinomimetic agents on cognition in primates following disruption by scopolamine or by lists of objects. Psychopharmacology (Berl)
99:189195.
Rupprecht R, Papadopoulos V, Rammes G, Baghai TC, Fan J, Akula N, Groyer G,
Adams D, and Schumacher M (2010) Translocator protein (18 kDa) (TSPO) as
a therapeutic target for neurological and psychiatric disorders. Nat Rev Drug
Discov 9:971988.
Rupprecht R, Rammes G, Eser D, Baghai TC, Schle C, Nothdurfter C, Troxler T,
Gentsch C, Kalkman HO, and Chaperon F, et al. (2009) Translocator protein (18
kD) as target for anxiolytics without benzodiazepine-like side effects. Science 325:
490493.
Ruscito BJ and Harrison NL (2003) Hemoglobin metabolites mimic benzodiazepines
and are possible mediators of hepatic encephalopathy. Blood 102:15251528.
Rush CR, Essman WD, Simpson CA, and Baker RW (2001) Reinforcing and subjectrated effects of methylphenidate and d-amphetamine in non-drug-abusing
humans. J Clin Psychopharmacol 21:273286.
Russo SJ, Bolanos CA, Theobald DE, DeCarolis NA, Renthal W, Kumar A,
Winstanley CA, Renthal NE, Wiley MD, and Self DW, et al. (2007) IRS2-Akt
pathway in midbrain dopamine neurons regulates behavioral and cellular
responses to opiates. Nat Neurosci 10:9399.
Russo SJ, Dietz DM, Dumitriu D, Morrison JH, Malenka RC, and Nestler EJ (2010)
The addicted synapse: mechanisms of synaptic and structural plasticity in nucleus
accumbens. Trends Neurosci 33:267276.
Russo SJ, Wilkinson MB, Mazei-Robison MS, Dietz DM, Maze I, Krishnan V, Renthal
W, Graham A, Birnbaum SG, and Green TA, et al. (2009) Nuclear factor kappa B
signaling regulates neuronal morphology and cocaine reward. J Neurosci 29:
35293537.
Rusyniak DE, Zaretsky DV, Zaretskaia MV, Durant PJ, and DiMicco JA (2012) The
orexin-1 receptor antagonist SB-334867 decreases sympathetic responses to
a moderate dose of methamphetamine and stress. Physiol Behav 107:743750.
Ryan LJ, Martone ME, Linder JC, and Groves PM (1988) Cocaine, in contrast to
D-amphetamine, does not cause axonal terminal degeneration in neostriatum and
agranular frontal cortex of Long-Evans rats. Life Sci 43:14031409.
Ryan PJ, Kastman HE, Krstew EV, Rosengren KJ, Hossain MA, Churilov L, Wade
JD, Gundlach AL, and Lawrence AJ (2013) Relaxin-3/RXFP3 system regulates
alcohol-seeking. Proc Natl Acad Sci USA 110:2078920794.
Saal D, Dong Y, Bonci A, and Malenka RC (2003) Drugs of abuse and stress trigger
a common synaptic adaptation in dopamine neurons. Neuron 37:577582.
Saarelainen KS, Ranna M, Rabe H, Sinkkonen ST, Mykkynen T, Uusi-Oukari M,
Linden AM, Lddens H, and Korpi ER (2008) Enhanced behavioral sensitivity to
the competitive GABA agonist, gaboxadol, in transgenic mice over-expressing
hippocampal extrasynaptic alpha6beta GABA(A) receptors. J Neurochem 105:
338350.
Sabol KE, Yancey DM, Speaker HA, and Mitchell SL (2013) Methamphetamine and
core temperature in the rat: ambient temperature, dose, and the effect of a D2
receptor blocker. Psychopharmacology (Berl) 228:551561.
Saccone NL, Wang JC, Breslau N, Johnson EO, Hatsukami D, Saccone SF, Grucza
RA, Sun L, Duan W, and Budde J, et al. (2009) The CHRNA5-CHRNA3-CHRNB4
nicotinic receptor subunit gene cluster affects risk for nicotine dependence in
African-Americans and in European-Americans. Cancer Res 69:68486856.
Sacktor TC (2008) PKMzeta, LTP maintenance, and the dynamic molecular biology of
memory storage. Prog Brain Res 169:2740.
Sacktor TC (2011) How does PKMz maintain long-term memory? Nat Rev Neurosci
12:915.
Sajikumar S and Frey JU (2004) Late-associativity, synaptic tagging, and the role of
dopamine during LTP and LTD. Neurobiol Learn Mem 82:1225.
Sajikumar S and Korte M (2011) Metaplasticity governs compartmentalization of
synaptic tagging and capture through brain-derived neurotrophic factor (BDNF)
and protein kinase Mzeta (PKMzeta). Proc Natl Acad Sci USA 108:25512556.
Sakaba T and Neher E (2003) Direct modulation of synaptic vesicle priming by GABA
(B) receptor activation at a glutamatergic synapse. Nature 424:775778.
Sakharkar AJ, Zhang H, Tang L, Shi G, and Pandey SC (2012) Histone deacetylases
(HDAC)-induced histone modifications in the amygdala: a role in rapid tolerance to
the anxiolytic effects of ethanol. Alcohol Clin Exp Res 36:6171.
Sakurai T, Amemiya A, Ishii M, Matsuzaki I, Chemelli RM, Tanaka H, Williams SC,
Richardson JA, Kozlowski GP, and Wilson S, et al. (1998) Orexins and orexin
receptors: a family of hypothalamic neuropeptides and G protein-coupled receptors
that regulate feeding behavior. Cell 92:573585.
Salamone JD (1994) The involvement of nucleus accumbens dopamine in appetitive
and aversive motivation. Behav Brain Res 61:117133.
Salamone JD and Correa M (2012) The mysterious motivational functions of mesolimbic dopamine. Neuron 76:470485.

Salas R, Sturm R, Boulter J, and De Biasi M (2009) Nicotinic receptors in the


habenulo-interpeduncular system are necessary for nicotine withdrawal in mice.
J Neurosci 29:30143018.
Salaspuro M (2007) Interrelationship between alcohol, smoking, acetaldehyde and
cancer. Novartis Found Symp 285:8089.
Sale A, Maya Vetencourt JF, Medini P, Cenni MC, Baroncelli L, De Pasquale R,
and Maffei L (2007) Environmental enrichment in adulthood promotes amblyopia
recovery through a reduction of intracortical inhibition. Nat Neurosci 10:679681.
Salin PA, Malenka RC, and Nicoll RA (1996) Cyclic AMP mediates a presynaptic
form of LTP at cerebellar parallel fiber synapses. Neuron 16:797803.
Sallette J, Pons S, Devillers-Thiery A, Soudant M, Prado de Carvalho L, Changeux
JP, and Corringer PJ (2005) Nicotine upregulates its own receptors through enhanced intracellular maturation. Neuron 46:595607.
Salmanzadeh F, Fathollahi Y, Semnanian S, and Shafizadeh M (2003) Dependence
on morphine impairs the induction of long-term potentiation in the CA1 region of
rat hippocampal slices. Brain Res 965:108113.
Salo R, Ursu S, Buonocore MH, Leamon MH, and Carter C (2009) Impaired prefrontal
cortical function and disrupted adaptive cognitive control in methamphetamine
abusers: a functional magnetic resonance imaging study. Biol Psychiatry 65:706709.
Salzmann J, Marie-Claire C, Le Guen S, Roques BP, and Noble F (2003) Importance
of ERK activation in behavioral and biochemical effects induced by MDMA in mice.
Br J Pharmacol 140:831838.
Sammut S, Threlfell S, and West AR (2010) Nitric oxide-soluble guanylyl cyclase
signaling regulates corticostriatal transmission and short-term synaptic plasticity
of striatal projection neurons recorded in vivo. Neuropharmacology 58:624631.
SAMSHA(2014a) Substance Abuse and Mental Heath Services Administration, Center
for Behavioral Health Stistics and Quality. (September 4, 2014). The NSDUH Report: Substance Use and Mental Health Estimates from 2013 National Survey on
Drug Use and Health: Overview of Findings,Substance Abuse and Mental Health
Services Administration, Rockville, MD.
SAMSHA (2014b) Substance Abuse and Mental Heath Services Administration,
Results from the 2013 National Survey on Drug Use and Health: Summary of
National Findings, NSDUH Series H-48, HHS Publication No. (SMA) 14-4863,
Substance Abuse and Mental Health Services Administration, Rockville, MD.
Sanchez V, Camarero J, OShea E, Green AR, and Colado MI (2003) Differential
effect of dietary selenium on the long-term neurotoxicity induced by MDMA in mice
and rats. Neuropharmacology 44:449461.
Sanchis-Segura C and Spanagel R (2006) Behavioural assessment of drug reinforcement and addictive features in rodents: an overview. Addict Biol 11:238.
Sangameswaran L, Fales HM, Friedrich P, and De Blas AL (1986) Purification of
a benzodiazepine from bovine brain and detection of benzodiazepine-like immunoreactivity in human brain. Proc Natl Acad Sci USA 83:92369240.
Santali EY, Cadogan AK, Daeid NN, Savage KA, and Sutcliffe OB (2011) Synthesis,
full chemical characterisation and development of validated methods for the
quantification of (6)-49-methylmethcathinone (mephedrone): a new legal high.
J Pharm Biomed Anal 56:246255.
Sarkar J, Wakefield S, MacKenzie G, Moss SJ, and Maguire J (2011) Neurosteroidogenesis is required for the physiological response to stress: role of
neurosteroid-sensitive GABAA receptors. J Neurosci 31:1819818210.
Sarti F, Borgland SL, Kharazia VN, and Bonci A (2007) Acute cocaine exposure alters
spine density and long-term potentiation in the ventral tegmental area. Eur J
Neurosci 26:749756.
Sartor CE, Lessov-Schlaggar CN, Scherrer JF, Bucholz KK, Madden PA, Pergadia
ML, Grant JD, Jacob T, and Xian H (2010) Initial response to cigarettes predicts
rate of progression to regular smoking: findings from an offspring-of-twins design.
Addict Behav 35:771778.
Saulskaya N and Marsden CA (1995) Conditioned dopamine release: dependence
upon N-methyl-D-aspartate receptors. Neuroscience 67:5763.
Saunders C, Ferrer JV, Shi L, Chen J, Merrill G, Lamb ME, Leeb-Lundberg LM,
Carvelli L, Javitch JA, and Galli A (2000) Amphetamine-induced loss of human
dopamine transporter activity: an internalization-dependent and cocaine-sensitive
mechanism. Proc Natl Acad Sci USA 97:68506855.
Savic I, Widn L, and Stone-Elander S (1991) Feasibility of reversing benzodiazepine
tolerance with flumazenil. Lancet 337:133137.
Savinainen JR, Saario SM, and Laitinen JT (2012) The serine hydrolases MAGL,
ABHD6 and ABHD12 as guardians of 2-arachidonoylglycerol signalling through
cannabinoid receptors. Acta Physiol (Oxf) 204:267276.
Scanzello CR, Hatzidimitriou G, Martello AL, Katz JL, and Ricaurte GA (1993) Serotonergic recovery after (+/-)3,4-(methylenedioxy) methamphetamine injury:
observations in rats. J Pharmacol Exp Ther 264:14841491.
Schilt T, de Win MM, Koeter M, Jager G, Korf DJ, van den Brink W, and Schmand B
(2007) Cognition in novice ecstasy users with minimal exposure to other drugs:
a prospective cohort study. Arch Gen Psychiatry 64:728736.
Schmaal L, Veltman DJ, Nederveen A, van den Brink W, and Goudriaan AE (2012)
N-acetylcysteine normalizes glutamate levels in cocaine-dependent patients:
a randomized crossover magnetic resonance spectroscopy study. Neuropsychopharmacology 37:21432152.
Schmidt CJ, Black CK, and Taylor VL (1990) Antagonism of the neurotoxicity due to
a single administration of methylenedioxymethamphetamine. Eur J Pharmacol
181:5970.
Schmidt CJ and Fadayel GM (1996) Regional effects of MK-801 on dopamine release:
effects of competitive NMDA or 5-HT2A receptor blockade. J Pharmacol Exp Ther
277:15411549.
Schmidt CJ, Ritter JK, Sonsalla PK, Hanson GR, and Gibb JW (1985) Role of dopamine in the neurotoxic effects of methamphetamine. J Pharmacol Exp Ther 233:
539544.
Schmitz F, Scherer EB, Machado FR, da Cunha AA, Tagliari B, Netto CA, and Wyse
AT (2012) Methylphenidate induces lipid and protein damage in prefrontal cortex,
but not in cerebellum, striatum and hippocampus of juvenile rats. Metab Brain Dis
27:605612.

Drug-Induced Neuroplasticity
Schneider M (2008) Puberty as a highly vulnerable developmental period for the
consequences of cannabis exposure. Addict Biol 13:253263.
Schneider M and Koch M (2003) Chronic pubertal, but not adult chronic cannabinoid
treatment impairs sensorimotor gating, recognition memory, and the performance
in a progressive ratio task in adult rats. Neuropsychopharmacology 28:17601769.
Schneider PG and Rodrguez de Lores Arnaiz G (2013) Ketamine prevents seizures
and reverses changes in muscarinic receptor induced by bicuculline in rats. Neurochem Int 62:258264.
Schochet TL, Bremer QZ, Brownfield MS, Kelley AE, and Landry CF (2008) The
dendritically targeted protein Dendrin is induced by acute nicotine in cortical
regions of adolescent rat brain. Eur J Neurosci 28:19671979.
Schochet TL, Kelley AE, and Landry CF (2005) Differential expression of arc mRNA
and other plasticity-related genes induced by nicotine in adolescent rat forebrain.
Neuroscience 135:285297.
Schoffelmeer AN, De Vries TJ, Wardeh G, van de Ven HW, and Vanderschuren LJ
(2002) Psychostimulant-induced behavioral sensitization depends on nicotinic receptor activation. J Neurosci 22:32693276.
Scholl ES, Pirone A, Cox DH, Duncan RK, and Jacob MH (2014) Alternative splice
isoforms of small conductance calcium-activated SK2 channels differ in molecular
interactions and surface levels. Channels (Austin) 8:6275.
Schramm NL, Egli RE, and Winder DG (2002) LTP in the mouse nucleus accumbens
is developmentally regulated. Synapse 45:213219.
Schrder N, ODell SJ, and Marshall JF (2003) Neurotoxic methamphetamine regimen severely impairs recognition memory in rats. Synapse 49:8996.
Schroeder BW and Shinnick-Gallagher P (2004) Fear memories induce a switch in
stimulus response and signaling mechanisms for long-term potentiation in the
lateral amygdala. Eur J Neurosci 20:549556.
Schroeder JP, Spanos M, Stevenson JR, Besheer J, Salling M, and Hodge CW (2008)
Cue-induced reinstatement of alcohol-seeking behavior is associated with increased ERK1/2 phosphorylation in specific limbic brain regions: blockade by the
mGluR5 antagonist MPEP. Neuropharmacology 55:546554.
Schuler V, Lscher C, Blanchet C, Klix N, Sansig G, Klebs K, Schmutz M, Heid J,
Gentry C, and Urban L, et al. (2001) Epilepsy, hyperalgesia, impaired memory, and
loss of pre- and postsynaptic GABA(B) responses in mice lacking GABA(B(1)).
Neuron 31:4758.
Schwartz BG, Rezkalla S, and Kloner RA (2010) Cardiovascular effects of cocaine.
Circulation 122:25582569.
Schwartz RD and Kellar KJ (1983) Nicotinic cholinergic receptor binding sites in the
brain: regulation in vivo. Science 220:214216.
Schwarz JM and Bilbo SD (2013) Adolescent morphine exposure affects long-term
microglial function and later-life relapse liability in a model of addiction. J Neurosci 33:961971.
Schwarz JM, Hutchinson MR, and Bilbo SD (2011) Early-life experience decreases
drug-induced reinstatement of morphine CPP in adulthood via microglial-specific
epigenetic programming of anti-inflammatory IL-10 expression. J Neurosci 31:
1783517847.
Schweizer TA and Vogel-Sprott M (2008) Alcohol-impaired speed and accuracy of
cognitive functions: a review of acute tolerance and recovery of cognitive performance. Exp Clin Psychopharmacol 16:240250.
Schwenk J, Metz M, Zolles G, Turecek R, Fritzius T, Bildl W, Tarusawa E, Kulik A,
Unger A, and Ivankova K, et al. (2010) Native GABA(B) receptors are heteromultimers with a family of auxiliary subunits. Nature 465:231235.
Schweri MM, Skolnick P, Rafferty MF, Rice KC, Janowsky AJ, and Paul SM (1985)
[3H]Threo-(+/-)-methylphenidate binding to 3,4-dihydroxyphenylethylamine uptake sites in corpus striatum: correlation with the stimulant properties of ritalinic
acid esters. J Neurochem 45:10621070.
Seeman P, Guan HC, and Hirbec H (2009) Dopamine D2High receptors stimulated by
phencyclidines, lysergic acid diethylamide, salvinorin A, and modafinil. Synapse
63:698704.
Seeman P and Kapur S (2003) Anesthetics inhibit high-affinity states of dopamine D2
and other G-linked receptors. Synapse 50:3540.
Seeman P, Ko F, and Tallerico T (2005) Dopamine receptor contribution to the action
of PCP, LSD and ketamine psychotomimetics. Mol Psychiatry 10:877883.
Segal DS (1975) Behavioral characterization of d- and l-amphetamine: neurochemical
implications. Science 190:475477.
Seiden LS, Woolverton WL, Lorens SA, Williams J, Corwin RL, Hata N, and Olimski
M (1993) Behavioral consequences of partial monoamine depletion in the CNS
after methamphetamine-like drugs: the conflict between pharmacology and toxicology. NIDA Res Monogr 136:3446.
Seiden LS, Commins DL, Vosmer G, Axt K, and Marek G (1988) Neurotoxicity in
dopamine and 5-hydroxytryptamine terminal fields: a regional analysis in
nigrostriatal and mesolimbic projections. Ann N Y Acad Sci 537:161172.
Sekine Y, Iyo M, Ouchi Y, Matsunaga T, Tsukada H, Okada H, Yoshikawa E,
Futatsubashi M, Takei N, and Mori N (2001) Methamphetamine-related psychiatric symptoms and reduced brain dopamine transporters studied with PET. Am J
Psychiatry 158:12061214.
Sellers EM, Naranjo CA, Harrison M, Devenyi P, Roach C, and Sykora K (1983)
Diazepam loading: simplified treatment of alcohol withdrawal. Clin Pharmacol
Ther 34:822826.
Seminerio MJ, Hansen R, Kaushal N, Zhang HT, McCurdy CR, and Matsumoto RR
(2013) The evaluation of AZ66, an optimized sigma receptor antagonist, against
methamphetamine-induced dopaminergic neurotoxicity and memory impairment
in mice. Int J Neuropsychopharmacol 16:10331044.
Seminerio MJ, Kaushal N, Shaikh J, Huber JD, Coop A, and Matsumoto RR (2011)
Sigma (s) receptor ligand, AC927 (N-phenethylpiperidine oxalate), attenuates
methamphetamine-induced hyperthermia and serotonin damage in mice. Pharmacol Biochem Behav 98:1220.
Seminerio MJ, Robson MJ, McCurdy CR, and Matsumoto RR (2012) Sigma receptor antagonists attenuate acute methamphetamine-induced hyperthermia

997

by a mechanism independent of IL-1b mRNA expression in the hypothalamus. Eur


J Pharmacol 691:103109.
Semple BD, Blomgren K, Gimlin K, Ferriero DM, and Noble-Haeusslein LJ (2013)
Brain development in rodents and humans: Identifying benchmarks of maturation
and vulnerability to injury across species. Prog Neurobiol 106-107:116.
Sennhauser FH and Schwarz HP (1986) Toxic psychosis from transdermal scopolamine in a child. Lancet 2:1033.
Sesack SR and Grace AA (2010) Cortico-Basal Ganglia reward network: microcircuitry. Neuropsychopharmacology 35:2747.
Seutin V, Johnson SW, and North RA (1994) Effect of dopamine and baclofen on
N-methyl-D-aspartate-induced burst firing in rat ventral tegmental neurons.
Neuroscience 58:201206.
Sevak RJ, Stoops WW, Hays LR, and Rush CR (2009) Discriminative stimulus and
subject-rated effects of methamphetamine, d-amphetamine, methylphenidate, and
triazolam in methamphetamine-trained humans. J Pharmacol Exp Ther 328:
10071018.
Sewell RA, Halpern JH, and Pope HG Jr (2006) Response of cluster headache to
psilocybin and LSD. Neurology 66:19201922.
Shah A, Silverstein PS, Singh DP, and Kumar A (2012) Involvement of metabotropic
glutamate receptor 5, AKT/PI3K signaling and NF-kB pathway in
methamphetamine-mediated increase in IL-6 and IL-8 expression in astrocytes.
J Neuroinflammation 9:52.
Shaham Y, Erb S, Leung S, Buczek Y, and Stewart J (1998) CP-154,526, a selective,
non-peptide antagonist of the corticotropin-releasing factor1 receptor attenuates
stress-induced relapse to drug seeking in cocaine- and heroin-trained rats. Psychopharmacology (Berl) 137:184190.
Shaham Y, Funk D, Erb S, Brown TJ, Walker CD, and Stewart J (1997)
Corticotropin-releasing factor, but not corticosterone, is involved in stress-induced
relapse to heroin-seeking in rats. J Neurosci 17:26052614.
Shamir A, Kwon OB, Karavanova I, Vullhorst D, Leiva-Salcedo E, Janssen MJ,
and Buonanno A (2012) The importance of the NRG-1/ErbB4 pathway for synaptic
plasticity and behaviors associated with psychiatric disorders. J Neurosci 32:
29882997.
Shankaran M, Yamamoto BK, and Gudelsky GA (1999) Mazindol attenuates the 3,4methylenedioxymethamphetamine-induced formation of hydroxyl radicals and
long-term depletion of serotonin in the striatum. J Neurochem 72:25162522.
Sharma SK, Klee WA, and Nirenberg M (1975) Dual regulation of adenylate cyclase
accounts for narcotic dependence and tolerance. Proc Natl Acad Sci USA 72:
30923096.
Shaw D, Norwood K, and Leslie JC (2011) Chlordiazepoxide and lavender oil alter
unconditioned anxiety-induced c-fos expression in the rat brain. Behav Brain Res
224:17.
Shen G, Mohamed MS, Das P, and Tietz EI (2009) Positive allosteric activation of
GABAA receptors bi-directionally modulates hippocampal glutamate plasticity and
behaviour. Biochem Soc Trans 37:13941398.
Shen HW, Gipson CD, Huits M, and Kalivas PW (2014) Prelimbic cortex and ventral
tegmental area modulate synaptic plasticity differentially in nucleus accumbens
during cocaine-reinstated drug seeking. Neuropsychopharmacology 39:11691177.
Shen W, Flajolet M, Greengard P, and Surmeier DJ (2008) Dichotomous dopaminergic control of striatal synaptic plasticity. Science 321:848851.
Shen X, Ruan X, and Zhao H (2012) Stimulation of midbrain dopaminergic structures
modifies firing rates of rat lateral habenula neurons. PLoS One 7:e34323.
Shepherd JD and Huganir RL (2007) The cell biology of synaptic plasticity: AMPA
receptor trafficking. Annu Rev Cell Dev Biol 23:613643.
Shetty AK, Burrows RC, and Phillips DE (1993) Alterations in neuronal development
in the substantia nigra pars compacta following in utero ethanol exposure: immunohistochemical and Golgi studies. Neuroscience 52:311322.
Shiba T, Yamato M, Kudo W, Watanabe T, Utsumi H, and Yamada K (2011) In vivo
imaging of mitochondrial function in methamphetamine-treated rats. Neuroimage
57:866872.
Shih PY, Engle SE, Oh G, Deshpande P, Puskar NL, Lester HA, and Drenan RM
(2014) Differential expression and function of nicotinic acetylcholine receptors in
subdivisions of medial habenula. J Neurosci 34:97899802.
Shippenberg TS, Zapata A, and Chefer VI (2007) Dynorphin and the pathophysiology
of drug addiction. Pharmacol Ther 116:306321.
Shipton OA and Paulsen O (2014) GluN2A and GluN2B subunit-containing NMDA
receptors in hippocampal plasticity. Philos Trans R Soc Lond B Biol Sci 369:
20130163.
Shirakawa T, Mitsuoka K, Kuroda K, Miyoshi S, Shiraki K, Naraoka H, Noda A,
Fujikawa A, and Fujiwara M (2013) [18F]FDG-PET as an imaging biomarker to
NMDA receptor antagonist-induced neurotoxicity. Toxicol Sci 133:1321.
Shoblock JR, Maisonneuve IM, and Glick SD (2003a) Differences between
d-methamphetamine and d-amphetamine in rats: working memory, tolerance, and
extinction. Psychopharmacology (Berl) 170:150156.
Shoblock JR, Sullivan EB, Maisonneuve IM, and Glick SD (2003b) Neurochemical
and behavioral differences between d-methamphetamine and d-amphetamine in
rats. Psychopharmacology (Berl) 165:359369.
Shonesy BC, Bluett RJ, Ramikie TS, Bldi R, Hermanson DJ, Kingsley PJ, Marnett
LJ, Winder DG, Colbran RJ, and Patel S (2014) Genetic disruption of 2-arachidonoylglycerol synthesis reveals a key role for endocannabinoid signaling in
anxiety modulation. Cell Reports 9:16441653.
Shonesy BC, Wang X, Rose KL, Ramikie TS, Cavener VS, Rentz T, Baucum AJ 2nd,
Jalan-Sakrikar N, Mackie K, and Winder DG, et al. (2013) CaMKII regulates
diacylglycerol lipase-a and striatal endocannabinoid signaling. Nat Neurosci 16:
456463.
Shortall SE, Green AR, Swift KM, Fone KC, and King MV (2013) Differential effects
of cathinone compounds and MDMA on body temperature in the rat, and pharmacological characterization of mephedrone-induced hypothermia. Br J Pharmacol
168:966977.

998

Korpi et al.

Shram MJ, Funk D, Li Z, and L AD (2006) Periadolescent and adult rats respond
differently in tests measuring the rewarding and aversive effects of nicotine. Psychopharmacology (Berl) 186:201208.
Shram MJ and L AD (2010) Adolescent male Wistar rats are more responsive than
adult rats to the conditioned rewarding effects of intravenously administered
nicotine in the place conditioning procedure. Behav Brain Res 206:240244.
Shram MJ, Sellers EM, and Romach MK (2011) Oral ketamine as a positive control in
human abuse potential studies. Drug Alcohol Depend 114:185193.
Sieghart W (1995) Structure and pharmacology of gamma-aminobutyric acidA receptor subtypes. Pharmacol Rev 47:181234.
Sieghart W, Eichinger A, Richards JG, and Mhler H (1987) Photoaffinity labeling of
benzodiazepine receptor proteins with the partial inverse agonist [3H]Ro 15-4513:
a biochemical and autoradiographic study. J Neurochem 48:4652.
Sigel E (2002) Mapping of the benzodiazepine recognition site on GABA(A) receptors.
Curr Top Med Chem 2:833839.
Sigel E and Steinmann ME (2012) Structure, function, and modulation of GABA(A)
receptors. J Biol Chem 287:4022440231.
Siggins GR, Roberto M, and Nie Z (2005) The tipsy terminal: presynaptic effects of
ethanol. Pharmacol Ther 107:8098.
Sillivan SE, Black YD, Naydenov AV, Vassoler FR, Hanlin RP, and Konradi C (2011)
Binge cocaine administration in adolescent rats affects amygdalar gene expression
patterns and alters anxiety-related behavior in adulthood. Biol Psychiatry 70:
583592.
Silva DD, Silva E, and Carmo H (2014) Combination effects of amphetamines under
hyperthermia - the role played by oxidative stress. J Appl Toxicol 34:637650.
Sim-Selley LJ, Selley DE, Vogt LJ, Childers SR, and Martin TJ (2000) Chronic heroin
self-administration desensitizes mu opioid receptor-activated G-proteins in specific
regions of rat brain. J Neurosci 20:45554562.
Simmler LD, Buser TA, Donzelli M, Schramm Y, Dieu LH, Huwyler J, Chaboz S,
Hoener MC, and Liechti ME (2013) Pharmacological characterization of designer
cathinones in vitro. Br J Pharmacol 168:458470.
Simms JA, Steensland P, Medina B, Abernathy KE, Chandler LJ, Wise R,
and Bartlett SE (2008) Intermittent access to 20% ethanol induces high ethanol
consumption in Long-Evans and Wistar rats. Alcohol Clin Exp Res 32:18161823.
Simon SL, Domier C, Carnell J, Brethen P, Rawson R, and Ling W (2000) Cognitive
impairment in individuals currently using methamphetamine. Am J Addict 9:
222231.
Sinclair JD (1972) The alcohol-deprivation effect. Influence of various factors. Q J
Stud Alcohol 33:769782.
Sircar R (2003) Postnatal phencyclidine-induced deficit in adult water maze performance is associated with N-methyl-D-aspartate receptor upregulation. Int J Dev
Neurosci 21:159167.
Sirohi S, Bakalkin G, and Walker BM (2012) Alcohol-induced plasticity in the
dynorphin/kappa-opioid receptor system. Front Mol Neurosci 5:95.
Sitte HH, Huck S, Reither H, Boehm S, Singer EA, and Pifl C (1998) Carriermediated release, transport rates, and charge transfer induced by amphetamine,
tyramine, and dopamine in mammalian cells transfected with the human dopamine transporter. J Neurochem 71:12891297.
Skaggs WE and McNaughton BL (1996) Replay of neuronal firing sequences in rat
hippocampus during sleep following spatial experience. Science 271:18701873.
Sklair-Tavron L, Shi WX, Lane SB, Harris HW, Bunney BS, and Nestler EJ (1996)
Chronic morphine induces visible changes in the morphology of mesolimbic dopamine neurons. Proc Natl Acad Sci USA 93:1120211207.
Skolnick P, Davis H, Arnelle D, and Deaver D (2014) Translational potential of
naloxone and naltrexone as TLR4 antagonists. Trends Pharmacol Sci 35:431432.
Slikker W Jr, Zou X, Hotchkiss CE, Divine RL, Sadovova N, Twaddle NC, Doerge DR,
Scallet AC, Patterson TA, and Hanig JP, et al. (2007) Ketamine-induced neuronal
cell death in the perinatal rhesus monkey. Toxicol Sci 98:145158.
Slimak MA, Ables JL, Frahm S, Antolin-Fontes B, Santos-Torres J, Moretti M, Gotti
C, and Ibaez-Tallon I (2014) Habenular expression of rare missense variants of
the b4 nicotinic receptor subunit alters nicotine consumption. Front Hum Neurosci
8:12.
Smith CM, Ryan PJ, Hosken IT, Ma S, and Gundlach AL (2011) Relaxin-3 systems in
the brainthe first 10 years. J Chem Neuroanat 42:262275.
Smith KM, Mitchell SN, and Joseph MH (1991) Effects of chronic and subchronic
nicotine on tyrosine hydroxylase activity in noradrenergic and dopaminergic neurones in the rat brain. J Neurochem 57:17501756.
Smith KS, Engin E, Meloni EG, and Rudolph U (2012) Benzodiazepine-induced
anxiolysis and reduction of conditioned fear are mediated by distinct GABAA receptor subtypes in mice. Neuropharmacology 63:250258.
Smolnik R, Pietrowsky R, Fehm HL, and Born J (1998) Enhanced selective attention
after low-dose administration of the benzodiazepine antagonist flumazenil. J Clin
Psychopharmacol 18:241247.
Smothers CT, Mrotek JJ, and Lovinger DM (1997) Chronic ethanol exposure leads to
a selective enhancement of N-methyl-D-aspartate receptor function in cultured
hippocampal neurons. J Pharmacol Exp Ther 283:12141222.
Snyder SH (1973) Amphetamine psychosis: a model schizophrenia mediated by
catecholamines. Am J Psychiatry 130:6167.
Sohal VS, Keist R, Rudolph U, and Huguenard JR (2003) Dynamic GABA(A) receptor
subtype-specific modulation of the synchrony and duration of thalamic oscillations.
J Neurosci 23:36493657.
Soleman S, Filippov MA, Dityatev A, and Fawcett JW (2013) Targeting the neural
extracellular matrix in neurological disorders. Neuroscience 253:194213.
Solowij N, Stephens RS, Roffman RA, Babor T, Kadden R, Miller M, Christiansen K,
McRee B, and Vendetti J; Marijuana Treatment Project Research Group (2002)
Cognitive functioning of long-term heavy cannabis users seeking treatment. JAMA
287:11231131.
Sommer WH, Rimondini R, Hansson AC, Hipskind PA, Gehlert DR, Barr CS,
and Heilig MA (2008) Upregulation of voluntary alcohol intake, behavioral

sensitivity to stress, and amygdala crhr1 expression following a history of dependence. Biol Psychiatry 63:139145.
Somoza EC, Winship D, Gorodetzky CW, Lewis D, Ciraulo DA, Galloway GP, Segal
SD, Sheehan M, Roache JD, and Bickel WK, et al. (2013) A multisite, double-blind,
placebo-controlled clinical trial to evaluate the safety and efficacy of vigabatrin for
treating cocaine dependence. JAMA Psychiatry 70:630637.
Song P and Zhao ZQ (2001) The involvement of glial cells in the development of
morphine tolerance. Neurosci Res 39:281286.
Sonner JM, Antognini JF, Dutton RC, Flood P, Gray AT, Harris RA, Homanics GE,
Kendig J, Orser B, and Raines DE, et al. (2003) Inhaled anesthetics and immobility: mechanisms, mysteries, and minimum alveolar anesthetic concentration.
Anesth Analg 97:718740.
Sonsalla PK, Riordan DE, and Heikkila RE (1991) Competitive and noncompetitive
antagonists at N-methyl-D-aspartate receptors protect against methamphetamineinduced dopaminergic damage in mice. J Pharmacol Exp Ther 256:506512.
Sorkina T, Doolen S, Galperin E, Zahniser NR, and Sorkin A (2003) Oligomerization
of dopamine transporters visualized in living cells by fluorescence resonance energy transfer microscopy. J Biol Chem 278:2827428283.
Souto M, Monti JM, and Altier H (1979) Effects of clozapine on the activity of central
dopaminergic and noradrenergic neurons. Pharmacol Biochem Behav 10:59.
Sowell ER, Peterson BS, Thompson PM, Welcome SE, Henkenius AL, and Toga AW
(2003) Mapping cortical change across the human life span. Nat Neurosci 6:
309315.
Spanagel R (2009) Alcoholism: a systems approach from molecular physiology to
addictive behavior. Physiol Rev 89:649705.
Spanagel R, Herz A, and Shippenberg TS (1992) Opposing tonically active endogenous opioid systems modulate the mesolimbic dopaminergic pathway. Proc Natl
Acad Sci USA 89:20462050.
Sparta DR, Hopf FW, Gibb SL, Cho SL, Stuber GD, Messing RO, Ron D, and Bonci A
(2013) Binge ethanol-drinking potentiates corticotropin releasing factor R1 receptor activity in the ventral tegmental area. Alcohol Clin Exp Res 37:16801687.
Spear LP (2000) The adolescent brain and age-related behavioral manifestations.
Neurosci Biobehav Rev 24:417463.
Spear LP (2014) Adolescents and alcohol: acute sensitivities, enhanced intake, and
later consequences. Neurotoxicol Teratol 41:5159.
Spiga S, Talani G, Mulas G, Licheri V, Fois GR, Muggironi G, Masala N, Cannizzaro
C, Biggio G, and Sanna E, et al. (2014) Hampered long-term depression and thin
spine loss in the nucleus accumbens of ethanol-dependent rats. Proc Natl Acad Sci
USA 111:E3745E3754.
Sprague JE, Banks ML, Cook VJ, and Mills EM (2003) Hypothalamic-pituitarythyroid axis and sympathetic nervous system involvement in hyperthermia induced by 3,4-methylenedioxymethamphetamine (Ecstasy). J Pharmacol Exp Ther
305:159166.
Sprague JE, Huang X, Kanthasamy A, and Nichols DE (1994) Attenuation of 3,4methylenedioxymethamphetamine (MDMA) induced neurotoxicity with the serotonin precursors tryptophan and 5-hydroxytryptophan. Life Sci 55:11931198.
Squeglia LM, Jacobus J, Sorg SF, Jernigan TL, and Tapert SF (2013) Early adolescent cortical thinning is related to better neuropsychological performance. J Int
Neuropsychol Soc 19:962970.
Srinivasan R, Pantoja R, Moss FJ, Mackey ED, Son CD, Miwa J, and Lester HA
(2011) Nicotine up-regulates alpha4beta2 nicotinic receptors and ER exit sites via
stoichiometry-dependent chaperoning. J Gen Physiol 137:5979.
Srinivasan R, Richards CI, Xiao C, Rhee D, Pantoja R, Dougherty DA, Miwa JM,
and Lester HA (2012) Pharmacological chaperoning of nicotinic acetylcholine
receptors reduces the endoplasmic reticulum stress response. Mol Pharmacol 81:
759769.
Srivareerat M, Tran TT, Salim S, Aleisa AM, and Alkadhi KA (2011) Chronic nicotine
restores normal Ab levels and prevents short-term memory and E-LTP impairment in Ab rat model of Alzheimers disease. Neurobiol Aging 32:834844.
Stamatakis AM and Stuber GD (2012) Activation of lateral habenula inputs to the
ventral midbrain promotes behavioral avoidance. Nat Neurosci 15:11051107.
Stanton PK and Sarvey JM (1985) Depletion of norepinephrine, but not serotonin,
reduces long-term potentiation in the dentate gyrus of rat hippocampal slices.
J Neurosci 5:21692176.
Stanwood GD and Levitt P (2004) Drug exposure early in life: functional repercussions of changing neuropharmacology during sensitive periods of brain development. Curr Opin Pharmacol 4:6571.
Starcevic V (2014) The reappraisal of benzodiazepines in the treatment of anxiety
and related disorders. Expert Rev Neurother 14:12751286.
Staszewski RD and Yamamoto BK (2006) Methamphetamine-induced spectrin proteolysis in the rat striatum. J Neurochem 96:12671276.
Steeds H, Carhart-Harris RL, and Stone JM (2015) Drug models of schizophrenia.
Ther Adv Psychopharmacol 5:4358.
Stefanik MT, Kupchik YM, Brown RM, and Kalivas PW (2013a) Optogenetic evidence
that pallidal projections, not nigral projections, from the nucleus accumbens core
are necessary for reinstating cocaine seeking. J Neurosci 33:1365413662.
Stefanik MT, Moussawi K, Kupchik YM, Smith KC, Miller RL, Huff ML, Deisseroth
K, Kalivas PW, and LaLumiere RT (2013b) Optogenetic inhibition of cocaine
seeking in rats. Addict Biol 18:5053.
Steinpreis RE, Kramer MA, Mix KS, and Piwowarczyk MC (1995) The effects of
MK801 on place conditioning. Neurosci Res 22:427430.
Stell BM, Brickley SG, Tang CY, Farrant M, and Mody I (2003) Neuroactive steroids
reduce neuronal excitability by selectively enhancing tonic inhibition mediated by
delta subunit-containing GABAA receptors. Proc Natl Acad Sci USA 100:
1443914444.
Stell BM and Mody I (2002) Receptors with different affinities mediate phasic and
tonic GABA(A) conductances in hippocampal neurons. J Neurosci 22:RC223.
Stephans SE, Whittingham TS, Douglas AJ, Lust WD, and Yamamoto BK (1998)
Substrates of energy metabolism attenuate methamphetamine-induced neurotoxicity in striatum. J Neurochem 71:613621.

Drug-Induced Neuroplasticity
Stephans SE and Yamamoto BK (1994) Methamphetamine-induced neurotoxicity:
roles for glutamate and dopamine efflux. Synapse 17:203209.
Sternson SM and Roth BL (2014) Chemogenetic tools to interrogate brain functions.
Annu Rev Neurosci 37:387407.
Stoenica L, Senkov O, Gerardy-Schahn R, Weinhold B, Schachner M, and Dityatev A
(2006) In vivo synaptic plasticity in the dentate gyrus of mice deficient in the
neural cell adhesion molecule NCAM or its polysialic acid. Eur J Neurosci 23:
22552264.
Stolerman IP and Jarvis MJ (1995) The scientific case that nicotine is addictive.
Psychopharmacology (Berl) 117:210, discussion 1420.
Stone DM, Hanson GR, and Gibb JW (1987) Differences in the central serotonergic
effects of methylenedioxymethamphetamine (MDMA) in mice and rats. Neuropharmacology 26:16571661.
Stone DM, Johnson M, Hanson GR, and Gibb JW (1988) Role of endogenous dopamine in the central serotonergic deficits induced by 3,4-methylenedioxymethamphetamine. J Pharmacol Exp Ther 247:7987.
Stone DM, Stahl DC, Hanson GR, and Gibb JW (1986) The effects of 3,4-methylenedioxymethamphetamine (MDMA) and 3,4-methylenedioxyamphetamine
(MDA) on monoaminergic systems in the rat brain. Eur J Pharmacol 128:4148.
Strustovu SI and Ebert B (2006) Pharmacological characterization of agonists at
delta-containing GABAA receptors: Functional selectivity for extrasynaptic
receptors is dependent on the absence of gamma2. J Pharmacol Exp Ther 316:
13511359.
Storvik M, Tiikkainen P, van Iersel M, and Wong G (2006) Distinct gene expression
profiles in adult rat brains after acute MK-801 and cocaine treatments. Eur Neuropsychopharmacol 16:211219.
Straiko MM, Young C, Cattano D, Creeley CE, Wang H, Smith DJ, Johnson SA, Li
ES, and Olney JW (2009) Lithium protects against anesthesia-induced developmental neuroapoptosis. Anesthesiology 110:862868.
Straub CJ, Carlezon WA Jr, and Rudolph U (2010) Diazepam and cocaine potentiate
brain stimulation reward in C57BL/6J mice. Behav Brain Res 206:1720.
Straub CJ, Lau HM, Parlato R, Schuetz G, Fritschy JM, and Rudolph U (2013)
Bidirectional regulation of intravenous general anesthetic actions by a3-containing
g-aminobutyric acid A receptors. Anesthesiology 118:562576.
Stringer JL, Greenfield LJ, Hackett JT, and Guyenet PG (1983) Blockade of longterm potentiation by phencyclidine and sigma opiates in the hippocampus in vivo
and in vitro. Brain Res 280:127138.
Stringer JL and Guyenet PG (1983) Elimination of long-term potentiation in the
hippocampus by phencyclidine and ketamine. Brain Res 258:159164.
Stuber GD, Hnasko TS, Britt JP, Edwards RH, and Bonci A (2010) Dopaminergic
terminals in the nucleus accumbens but not the dorsal striatum corelease glutamate. J Neurosci 30:82298233.
Stuber GD, Hopf FW, Hahn J, Cho SL, Guillory A, and Bonci A (2008) Voluntary
ethanol intake enhances excitatory synaptic strength in the ventral tegmental
area. Alcohol Clin Exp Res 32:17141720.
Stuber GD and Mason AO (2013) Integrating optogenetic and pharmacological
approaches to study neural circuit function: current applications and future
directions. Pharmacol Rev 65:156170.
Stuber GD, Sparta DR, Stamatakis AM, van Leeuwen WA, Hardjoprajitno JE, Cho S,
Tye KM, Kempadoo KA, Zhang F, and Deisseroth K, et al. (2011) Excitatory
transmission from the amygdala to nucleus accumbens facilitates reward seeking.
Nature 475:377380.
Studerus E, Kometer M, Hasler F, and Vollenweider FX (2011) Acute, subacute and
long-term subjective effects of psilocybin in healthy humans: a pooled analysis of
experimental studies. J Psychopharmacol 25:14341452.
Sturgess JE, George TP, Kennedy JL, Heinz A, and Mller DJ (2011) Pharmacogenetics of alcohol, nicotine and drug addiction treatments. Addict Biol 16:357376.
Suchankova P, Steensland P, Fredriksson I, Engel JA, and Jerlhag E (2013) Ghrelin
receptor (GHS-R1A) antagonism suppresses both alcohol consumption and the
alcohol deprivation effect in rats following long-term voluntary alcohol consumption. PLoS One 8:e71284.
Sugiyama S, Di Nardo AA, Aizawa S, Matsuo I, Volovitch M, Prochiantz A,
and Hensch TK (2008) Experience-dependent transfer of Otx2 homeoprotein into
the visual cortex activates postnatal plasticity. Cell 134:508520.
Sugiyama S, Prochiantz A, and Hensch TK (2009) From brain formation to plasticity:
insights on Otx2 homeoprotein. Dev Growth Differ 51:369377.
Sulzer D (2011) How addictive drugs disrupt presynaptic dopamine neurotransmission. Neuron 69:628649.
Sulzer D, Chen TK, Lau YY, Kristensen H, Rayport S, and Ewing A (1995) Amphetamine redistributes dopamine from synaptic vesicles to the cytosol and promotes reverse transport. J Neurosci 15:41024108.
Sulzer D and Rayport S (1990) Amphetamine and other psychostimulants reduce pH
gradients in midbrain dopaminergic neurons and chromaffin granules: a mechanism of action. Neuron 5:797808.
Sulzer D, Sonders MS, Poulsen NW, and Galli A (2005) Mechanisms of neurotransmitter release by amphetamines: a review. Prog Neurobiol 75:406433.
Sulzer D and Zecca L (2000) Intraneuronal dopamine-quinone synthesis: a review.
Neurotox Res 1:181195.
Sun L, Lam WP, Wong YW, Lam LH, Tang HC, Wai MS, Mak YT, Pan F, and Yew
DT (2011) Permanent deficits in brain functions caused by long-term ketamine
treatment in mice. Hum Exp Toxicol 30:12871296.
Sun Y, Evans J, Russell B, Kydd R, and Connor B (2013) A benzodiazepine impairs
the neurogenic and behavioural effects of fluoxetine in a rodent model of chronic
stress. Neuropharmacology 72:2028.
Sung YH, Yurgelun-Todd DA, Shi XF, Kondo DG, Lundberg KJ, McGlade EC, Hellem
TL, Huber RS, Fiedler KK, and Harrell RE, et al. (2013) Decreased frontal lobe
phosphocreatine levels in methamphetamine users. Drug Alcohol Depend 129:
102109.

999

Surman CB, Hammerness PG, Pion K, and Faraone SV (2013) Do stimulants improve functioning in adults with ADHD? A review of the literature. Eur Neuropsychopharmacol 23:528533.
Surmeier DJ, Plotkin J, and Shen W (2009) Dopamine and synaptic plasticity in
dorsal striatal circuits controlling action selection. Curr Opin Neurobiol 19:
621628.
Sutherland RJ (1982) The dorsal diencephalic conduction system: a review of the
anatomy and functions of the habenular complex. Neurosci Biobehav Rev 6:113.
Suwanjang W, Phansuwan-Pujito P, Govitrapong P, and Chetsawang B (2012) Calpastatin reduces calpain and caspase activation in methamphetamine-induced
toxicity in human neuroblastoma SH-SY5Y cultured cells. Neurosci Lett 526:4953.
Svenningsson P, Nishi A, Fisone G, Girault JA, Nairn AC, and Greengard P (2004)
DARPP-32: an integrator of neurotransmission. Annu Rev Pharmacol Toxicol 44:
269296.
Swanson JM, Kinsbourne M, Nigg J, Lanphear B, Stefanatos GA, Volkow N, Taylor
E, Casey BJ, Castellanos FX, and Wadhwa PD (2007) Etiologic subtypes of
attention-deficit/hyperactivity disorder: brain imaging, molecular genetic and environmental factors and the dopamine hypothesis. Neuropsychol Rev 17:3959.
Swanson LW and Cowan WM (1979) The connections of the septal region in the rat.
J Comp Neurol 186:621655.
Switzer RC 3rd (2000) Application of silver degeneration stains for neurotoxicity
testing. Toxicol Pathol 28:7083.
Szabo B, Siemes S, and Wallmichrath I (2002) Inhibition of GABAergic neurotransmission in the ventral tegmental area by cannabinoids. Eur J Neurosci 15:
20572061.
Szabo B, Than M, Thorn D, and Wallmichrath I (2004) Analysis of the effects of
cannabinoids on synaptic transmission between basket and Purkinje cells in the
cerebellar cortex of the rat. J Pharmacol Exp Ther 310:915925.
Szumlinski KK, Ary AW, Lominac KD, Klugmann M, and Kippin TE (2008)
Accumbens Homer2 overexpression facilitates alcohol-induced neuroplasticity in
C57BL/6J mice. Neuropsychopharmacology 33:13651378.
Takahashi M, Kakita A, Futamura T, Watanabe Y, Mizuno M, Sakimura K, Castren
E, Nabeshima T, Someya T, and Nawa H (2006) Sustained brain-derived neurotrophic factor up-regulation and sensorimotor gating abnormality induced by
postnatal exposure to phencyclidine: comparison with adult treatment. J Neurochem 99:770780.
Takahashi T, Yamashita H, Zhang YX, and Nakamura S (1996) Inhibitory effect of
MK-801 on amantadine-induced dopamine release in the rat striatum. Brain Res
Bull 41:363367.
Tan KR, Brown M, Laboube G, Yvon C, Creton C, Fritschy JM, Rudolph U,
and Lscher C (2010) Neural bases for addictive properties of benzodiazepines.
Nature 463:769774.
Tan KR, Yvon C, Turiault M, Mirzabekov JJ, Doehner J, Laboube G, Deisseroth K,
Tye KM, and Lscher C (2012a) GABA neurons of the VTA drive conditioned place
aversion. Neuron 73:11731183.
Tan S, Lam WP, Wai MS, Yu WH, and Yew DT (2012b) Chronic ketamine administration modulates midbrain dopamine system in mice. PLoS One 7:e43947.
Tanda G, Pontieri FE, and Di Chiara G (1997) Cannabinoid and heroin activation of
mesolimbic dopamine transmission by a common mu1 opioid receptor mechanism.
Science 276:20482050.
Tang J and Dani JA (2009) Dopamine enables in vivo synaptic plasticity associated
with the addictive drug nicotine. Neuron 63:673682.
Tanimura A, Liu J, Namba T, Seki T, Matsubara Y, Itoh M, Suzuki T, and Arai H
(2009) Prenatal phencyclidine exposure alters hippocampal cell proliferation in
offspring rats. Synapse 63:729736.
Tanimura A, Yamazaki M, Hashimotodani Y, Uchigashima M, Kawata S, Abe M,
Kita Y, Hashimoto K, Shimizu T, and Watanabe M, et al. (2010) The endocannabinoid 2-arachidonoylglycerol produced by diacylglycerol lipase alpha mediates
retrograde suppression of synaptic transmission. Neuron 65:320327.
Tapper AR, McKinney SL, Nashmi R, Schwarz J, Deshpande P, Labarca C,
Whiteaker P, Marks MJ, Collins AC, and Lester HA (2004) Nicotine activation
of alpha4* receptors: sufficient for reward, tolerance, and sensitization. Science
306:10291032.
Teng L, Crooks PA, and Dwoskin LP (1998) Lobeline displaces [3H]dihydrotetrabenazine binding and releases [3H]dopamine from rat striatal synaptic vesicles:
comparison with d-amphetamine. J Neurochem 71:258265.
Thalhammer A and Cingolani LA (2014) Cell adhesion and homeostatic synaptic
plasticity. Neuropharmacology 78:2330.
Thannickal TC, Moore RY, Nienhuis R, Ramanathan L, Gulyani S, Aldrich M,
Cornford M, and Siegel JM (2000) Reduced number of hypocretin neurons in human narcolepsy. Neuron 27:469474.
Theberge FR, Li X, Kambhampati S, Pickens CL, St Laurent R, Bossert JM,
Baumann MH, Hutchinson MR, Rice KC, and Watkins LR, et al. (2013) Effect of
chronic delivery of the Toll-like receptor 4 antagonist (+)-naltrexone on incubation of heroin craving. Biol Psychiatry 73:729737.
Theile JW, Morikawa H, Gonzales RA, and Morrisett RA (2008) Ethanol enhances
GABAergic transmission onto dopamine neurons in the ventral tegmental area of
the rat. Alcohol Clin Exp Res 32:10401048.
Theunissen EL, Heckman P, de Sousa Fernandes Perna EB, Kuypers KP, Sambeth
A, Blokland A, Prickaerts J, Toennes SW, and Ramaekers JG (2015) Rivastigmine
but not vardenafil reverses cannabis-induced impairment of verbal memory in
healthy humans. Psychopharmacology (Berl) 232:343353.
Thiele TE, Marsh DJ, Ste Marie L, Bernstein IL, and Palmiter RD (1998) Ethanol
consumption and resistance are inversely related to neuropeptide Y levels. Nature
396:366369.
Thomas DM and Kuhn DM (2005) MK-801 and dextromethorphan block microglial
activation and protect against methamphetamine-induced neurotoxicity. Brain Res
1050:190198.

1000

Korpi et al.

Thomas MJ, Beurrier C, Bonci A, and Malenka RC (2001) Long-term depression in


the nucleus accumbens: a neural correlate of behavioral sensitization to cocaine.
Nat Neurosci 4:12171223.
Thomasius R, Zapletalova P, Petersen K, Buchert R, Andresen B, Wartberg L,
Nebeling B, and Schmoldt A (2006) Mood, cognition and serotonin transporter
availability in current and former ecstasy (MDMA) users: the longitudinal perspective. J Psychopharmacol 20:211225.
Thompson AJ (2013) Recent developments in 5-HT3 receptor pharmacology. Trends
Pharmacol Sci 34:100109.
Thompson BL and Rosen JB (2006) Immediate-early gene expression in the central
nucleus of the amygdala is not specific for anxiolytic or anxiogenic drugs. Neuropharmacology 50:5768.
Thompson PM, Cannon TD, Narr KL, van Erp T, Poutanen VP, Huttunen M,
Lnnqvist J, Standertskjld-Nordenstam CG, Kaprio J, and Khaledy M, et al.
(2001) Genetic influences on brain structure. Nat Neurosci 4:12531258.
Thompson PM, Hayashi KM, Simon SL, Geaga JA, Hong MS, Sui Y, Lee JY, Toga
AW, Ling W, and London ED (2004) Structural abnormalities in the brains of
human subjects who use methamphetamine. J Neurosci 24:60286036.
Thompson PM, Sowell ER, Gogtay N, Giedd JN, Vidal CN, Hayashi KM, Leow A,
Nicolson R, Rapoport JL, and Toga AW (2005) Structural MRI and brain development. Int Rev Neurobiol 67:285323.
Thomsen MS, Christensen DZ, Hansen HH, Redrobe JP, and Mikkelsen JD (2009) alpha(7)
Nicotinic acetylcholine receptor activation prevents behavioral and molecular changes
induced by repeated phencyclidine treatment. Neuropharmacology 56:10011009.
Thorn DA, Winter JC, and Li JX (2012) Agmatine attenuates methamphetamineinduced conditioned place preference in rats. Eur J Pharmacol 680:6972.
Tietz EI, Zeng XJ, Chen S, Lilly SM, Rosenberg HC, and Kometiani P (1999)
Antagonist-induced reversal of functional and structural measures of hippocampal
benzodiazepine tolerance. J Pharmacol Exp Ther 291:932942.
Tizabi Y, Bhatti BH, Manaye KF, Das JR, and Akinfiresoye L (2012) Antidepressantlike effects of low ketamine dose is associated with increased hippocampal AMPA/
NMDA receptor density ratio in female Wistar-Kyoto rats. Neuroscience 213:7280.
Tobler I, Kopp C, Deboer T, and Rudolph U (2001) Diazepam-induced changes in
sleep: role of the alpha 1 GABA(A) receptor subtype. Proc Natl Acad Sci USA 98:
64646469.
Toborek M, Seelbach MJ, Rashid CS, Andrs IE, Chen L, Park M, and Esser KA
(2013) Voluntary exercise protects against methamphetamine-induced oxidative
stress in brain microvasculature and disruption of the blood-brain barrier. Mol
Neurodegener 8:22. DOI 10.1186/1750-1326-8-22
Toda S, Shen H, and Kalivas PW (2010) Inhibition of actin polymerization prevents
cocaine-induced changes in spine morphology in the nucleus accumbens. Neurotox
Res 18:410415.
Tokuda K, ODell KA, Izumi Y, and Zorumski CF (2010) Midazolam inhibits hippocampal long-term potentiation and learning through dual central and peripheral
benzodiazepine receptor activation and neurosteroidogenesis. J Neurosci 30:
1678816795.
Tomasi D, Goldstein RZ, Telang F, Maloney T, Alia-Klein N, Caparelli EC,
and Volkow ND (2007) Widespread disruption in brain activation patterns to
a working memory task during cocaine abstinence. Brain Res 1171:8392.
Tong J, Ross BM, Schmunk GA, Peretti FJ, Kalasinsky KS, Furukawa Y, Ang LC,
Aiken SS, Wickham DJ, and Kish SJ (2003) Decreased striatal dopamine D1
receptor-stimulated adenylyl cyclase activity in human methamphetamine users.
Am J Psychiatry 160:896903.
Tong ZY, Overton PG, and Clark D (1996) Antagonism of NMDA receptors but not
AMPA/kainate receptors blocks bursting in dopaminergic neurons induced by
electrical stimulation of the prefrontal cortex. J Neural Transm 103:889904.
Tonini R, Ciardo S, Cerovic M, Rubino T, Parolaro D, Mazzanti M, and Zippel R
(2006) ERK-dependent modulation of cerebellar synaptic plasticity after chronic
Delta9-tetrahydrocannabinol exposure. J Neurosci 26:58105818.
Tonner PH and Miller KW (1995) Molecular sites of general anaesthetic action on
acetylcholine receptors. Eur J Anaesthesiol 12:2130.
Toriumi K, Mouri A, Narusawa S, Aoyama Y, Ikawa N, Lu L, Nagai T, Mamiya T,
Kim HC, and Nabeshima T (2012) Prenatal NMDA receptor antagonism impaired
proliferation of neuronal progenitor, leading to fewer glutamatergic neurons in the
prefrontal cortex. Neuropsychopharmacology 37:13871396.
Tormey WP (2012) Cannabis, possible cardiac deaths and the coroner in Ireland. Ir J
Med Sci 181:479482.
Tornberg J, Segerstrle M, Kulesskaya N, Voikar V, Taira T, and Airaksinen MS
(2007) KCC2-deficient mice show reduced sensitivity to diazepam, but normal
alcohol-induced motor impairment, gaboxadol-induced sedation, and neurosteroidinduced hypnosis. Neuropsychopharmacology 32:911918.
Trnen P, Storvik M, Lindn AM, Kontkane O, Marvanov M, Lakso M, Castrn E,
and Wong G (2002) Expression profiling to understand actions of NMDA/glutamate
receptor antagonists in rat brain. Neurochem Res 27:12091220.
Tortoriello G, Morris CV, Alpar A, Fuzik J, Shirran SL, Calvigioni D, Keimpema E,
Botting CH, Reinecke K, and Herdegen T, et al. (2014) Miswiring the brain: D9tetrahydrocannabinol disrupts cortical development by inducing an SCG10/
stathmin-2 degradation pathway. EMBO J 33:668685.
Trecki J, Gerona RR, and Schwartz MD (2015) Synthetic cannabinoid-related illnesses and deaths. N Engl J Med 373:103107.
Treistman SN and Martin GE (2009) BK Channels: mediators and models for alcohol
tolerance. Trends Neurosci 32:629637.
Trudell JR, Messing RO, Mayfield J, and Harris RA (2014) Alcohol dependence:
molecular and behavioral evidence. Trends Pharmacol Sci 35:317323.
Tseng KY, Chambers RA, and Lipska BK (2009) The neonatal ventral hippocampal
lesion as a heuristic neurodevelopmental model of schizophrenia. Behav Brain Res
204:295305.
Tseng KY, Lewis BL, Lipska BK, and ODonnell P (2007) Post-pubertal disruption of
medial prefrontal cortical dopamine-glutamate interactions in a developmental
animal model of schizophrenia. Biol Psychiatry 62:730738.

Turner DM, Sapp DW, and Olsen RW (1991) The benzodiazepine/alcohol antagonist
Ro 15-4513: binding to a GABAA receptor subtype that is insensitive to diazepam.
J Pharmacol Exp Ther 257:12361242.
Turrigiano GG (1999) Homeostatic plasticity in neuronal networks: the more things
change, the more they stay the same. Trends Neurosci 22:221227.
Turrigiano GG and Nelson SB (2004) Homeostatic plasticity in the developing nervous system. Nat Rev Neurosci 5:97107.
Tyacke RJ, Lingford-Hughes A, Reed LJ, and Nutt DJ (2010) GABAB receptors in
addiction and its treatment. Adv Pharmacol 58:373396.
Tzschentke TM (2007) Measuring reward with the conditioned place preference
(CPP) paradigm: update of the last decade. Addict Biol 12:227462.
Tzschentke TM and Schmidt WJ (1995) N-methyl-D-aspartic acid-receptor antagonists block morphine-induced conditioned place preference in rats. Neurosci Lett
193:3740.
Uchimoto K, Miyazaki T, Kamiya Y, Mihara T, Koyama Y, Taguri M, Inagawa G,
Takahashi T, and Goto T (2014) Isoflurane impairs learning and hippocampal longterm potentiation via the saturation of synaptic plasticity. Anesthesiology 121:
302310.
Ueda H, Inoue M, Takeshima H, and Iwasawa Y (2000) Enhanced spinal nociceptin
receptor expression develops morphine tolerance and dependence. J Neurosci 20:
76407647.
Uemura K, Aki T, Yamaguchi K, and Yoshida Ki (2003) Protein kinase C-epsilon
protects PC12 cells against methamphetamine-induced death: possible involvement of suppression of glutamate receptor. Life Sci 72:15951607.
Ulbrich MH and Isacoff EY (2008) Rules of engagement for NMDA receptor subunits.
Proc Natl Acad Sci USA 105:1416314168.
Ungless MA, Argilli E, and Bonci A (2010) Effects of stress and aversion on dopamine
neurons: implications for addiction. Neurosci Biobehav Rev 35:151156.
Ungless MA, Magill PJ, and Bolam JP (2004) Uniform inhibition of dopamine neurons in the ventral tegmental area by aversive stimuli. Science 303:20402042.
Ungless MA, Whistler JL, Malenka RC, and Bonci A (2001) Single cocaine exposure
in vivo induces long-term potentiation in dopamine neurons. Nature 411:583587.
UNODC (2011) 2011 Global ATS Assessment, Vienna, United Nations Publication.
Unwin N (2005) Refined structure of the nicotinic acetylcholine receptor at 4A resolution. J Mol Biol 346:967989.
Uusi-Oukari M and Korpi ER (1990) Diazepam sensitivity of the binding of an imidazobenzodiazepine, [3H]Ro 15-4513, in cerebellar membranes from two rat lines
developed for high and low alcohol sensitivity. J Neurochem 54:19801987.
Uusi-Oukari M and Korpi ER (2010) Regulation of GABA(A) receptor subunit expression by pharmacological agents. Pharmacol Rev 62:97135.
Vaccarino FJ and Koob GF (1984) Microinjections of nanogram amounts of sulfated
cholecystokinin octapeptide into the rat nucleus accumbens attenuates brain
stimulation reward. Neurosci Lett 52:6166.
Vaidya VA, Marek GJ, Aghajanian GK, and Duman RS (1997) 5-HT2A receptormediated regulation of brain-derived neurotrophic factor mRNA in the hippocampus and the neocortex. J Neurosci 17:27852795.
Visnen J, Ihalainen J, Tanila H, and Castrn E (2004) Effects of NMDA-receptor
antagonist treatment on c-fos expression in rat brain areas implicated in schizophrenia. Cell Mol Neurobiol 24:769780.
Visnen J, Saarelainen T, Koponen E, and Castrn E (2003) Altered trkB neurotrophin
receptor activation does not influence the N-methyl-D-aspartate receptor antagonistmediated neurotoxicity in mouse posterior cingulate cortex. Neurosci Lett 350:14.
Valenzuela CF, Bhave S, Hoffman P, and Harris RA (1998a) Acute effects of ethanol
on pharmacologically isolated kainate receptors in cerebellar granule neurons:
comparison with NMDA and AMPA receptors. J Neurochem 71:17771780.
Valenzuela CF, Cardoso RA, Wick MJ, Weiner JL, Dunwiddie TV, and Harris RA
(1998b) Effects of ethanol on recombinant glycine receptors expressed in mammalian cell lines. Alcohol Clin Exp Res 22:11321136.
Valenzuela CF, Morton RA, Diaz MR, and Topper L (2012) Does moderate drinking
harm the fetal brain? Insights from animal models. Trends Neurosci 35:284292.
Valjent E, Pags C, Herv D, Girault JA, and Caboche J (2004) Addictive and nonaddictive drugs induce distinct and specific patterns of ERK activation in mouse
brain. Eur J Neurosci 19:18261836.
Valle M, Vitiello S, Bellocchio L, Hbert-Chatelain E, Monlezun S, Martin-Garcia E,
Kasanetz F, Baillie GL, Panin F, and Cathala A, et al. (2014) Pregnenolone can
protect the brain from cannabis intoxication. Science 343:9498.
Vallejo YF, Buisson B, Bertrand D, and Green WN (2005) Chronic nicotine exposure
upregulates nicotinic receptors by a novel mechanism. J Neurosci 25:55635572.
van Amsterdam J, Opperhuizen A, and van den Brink W (2011) Harm potential of
magic mushroom use: a review. Regul Toxicol Pharmacol 59:423429.
van Amsterdam JG, Brunt TM, McMaster MT, and Niesink RJ (2012) Possible longterm effects of g-hydroxybutyric acid (GHB) due to neurotoxicity and overdose.
Neurosci Biobehav Rev 36:12171227.
Van Camp N, Boisgard R, Kuhnast B, Thz B, Viel T, Grgoire MC, Chauveau F,
Boutin H, Katsifis A, and Doll F, et al. (2010) In vivo imaging of neuroinflammation: a comparative study between [(18)F]PBR111, [ (11)C]CLINME and
[(11)C]PK11195 in an acute rodent model. Eur J Nucl Med Mol Imaging 37:
962972.
Van den Oever MC, Goriounova NA, Li KW, Van der Schors RC, Binnekade R,
Schoffelmeer AN, Mansvelder HD, Smit AB, Spijker S, and De Vries TJ (2008)
Prefrontal cortex AMPA receptor plasticity is crucial for cue-induced relapse to
heroin-seeking. Nat Neurosci 11:10531058.
Van den Oever MC, Lubbers BR, Goriounova NA, Li KW, Van der Schors RC, Loos M,
Riga D, Wiskerke J, Binnekade R, and Stegeman M, et al. (2010) Extracellular
matrix plasticity and GABAergic inhibition of prefrontal cortex pyramidal cells
facilitates relapse to heroin seeking. Neuropsychopharmacology 35:21202133.
Vanderschuren LJ, Di Ciano P, and Everitt BJ (2005) Involvement of the dorsal
striatum in cue-controlled cocaine seeking. J Neurosci 25:86658670.
Vanderschuren LJ and Everitt BJ (2004) Drug seeking becomes compulsive after
prolonged cocaine self-administration. Science 305:10171019.

Drug-Induced Neuroplasticity
Vanderschuren LJ and Everitt BJ (2005) Behavioral and neural mechanisms of
compulsive drug seeking. Eur J Pharmacol 526:7788.
Vanderschuren LJ, Schoffelmeer AN, Mulder AH, and De Vries TJ (1998) Dizocilpine
(MK801): use or abuse? Trends Pharmacol Sci 19:7981.
van der Veen R, Boshuizen MC, and de Kloet ER (2013) Mifepristone treatment
affects the response to repeated amphetamine injections, but does not attenuate
the expression of sensitization. Psychopharmacology (Berl) 230:547556.
Van Gaal L, Pi-Sunyer X, Desprs JP, McCarthy C, and Scheen A (2008) Efficacy and
safety of rimonabant for improvement of multiple cardiometabolic risk factors in
overweight/obese patients: pooled 1-year data from the Rimonabant in Obesity
(RIO) program. Diabetes Care 31 (Suppl 2):S229S240.
Van Gaal LF, Rissanen AM, Scheen AJ, Ziegler O, and Rssner S; RIO-Europe Study
Group (2005) Effects of the cannabinoid-1 receptor blocker rimonabant on weight
reduction and cardiovascular risk factors in overweight patients: 1-year experience
from the RIO-Europe study. Lancet 365:13891397.
van Nieuwenhuijzen PS, Kashem MA, Matsumoto I, Hunt GE, and McGregor IS
(2010) A long hangover from party drugs: residual proteomic changes in the hippocampus of rats 8 weeks after g-hydroxybutyrate (GHB), 3,4-methylenedioxymethamphetamine (MDMA) or their combination. Neurochem Int 56:871877.
van Nieuwenhuijzen PS, McGregor IS, and Hunt GE (2009) The distribution of
gamma-hydroxybutyrate-induced Fos expression in rat brain: comparison with
baclofen. Neuroscience 158:441455.
van Rijnsoever C, Tuber M, Choulli MK, Keist R, Rudolph U, Mohler H, Fritschy
JM, and Crestani F (2004) Requirement of alpha5-GABAA receptors for the development of tolerance to the sedative action of diazepam in mice. J Neurosci 24:
67856790.
Van Sickle BJ, Cox AS, Schak K, Greenfield LJ Jr, and Tietz EI (2002) Chronic
benzodiazepine administration alters hippocampal CA1 neuron excitability:
NMDA receptor function and expression(1). Neuropharmacology 43:595606.
Van Sickle BJ and Tietz EI (2002) Selective enhancement of AMPA receptormediated function in hippocampal CA1 neurons from chronic benzodiazepinetreated rats. Neuropharmacology 43:1127.
Van Sickle BJ, Xiang K, and Tietz EI (2004) Transient plasticity of hippocampal CA1
neuron glutamate receptors contributes to benzodiazepine withdrawal-anxiety.
Neuropsychopharmacology 29:19942006.
Van Sickle MD, Duncan M, Kingsley PJ, Mouihate A, Urbani P, Mackie K, Stella N,
Makriyannis A, Piomelli D, and Davison JS, et al. (2005) Identification and functional characterization of brainstem cannabinoid CB2 receptors. Science 310:
329332.
van Zessen R, Phillips JL, Budygin EA, and Stuber GD (2012) Activation of VTA
GABA neurons disrupts reward consumption. Neuron 73:11841194.
Vargas-Perez H, Bahi A, Bufalino MR, Ting-A-Kee R, Maal-Bared G, Lam J, Fahmy
A, Clarke L, Blanchard JK, and Larsen BR, et al. (2014) BDNF signaling in the
VTA links the drug-dependent state to drug withdrawal aversions. J Neurosci 34:
78997909.
Vargas-Perez H, Ting-A Kee R, Walton CH, Hansen DM, Razavi R, Clarke L,
Bufalino MR, Allison DW, Steffensen SC, and van der Kooy D (2009) Ventral
tegmental area BDNF induces an opiate-dependent-like reward state in naive rats.
Science 324:17321734.
Vashchinkina E, Manner AK, Vekovischeva O, den Hollander B, Uusi-Oukari M,
Aitta-Aho T, and Korpi ER (2014a) Neurosteroid Agonist at GABAA receptor
induces persistent neuroplasticity in VTA dopamine neurons. Neuropsychopharmacology 39:727737.
Vashchinkina E, Panhelainen A, Aitta-Aho T, and Korpi ER (2014b) GABAA receptor
drugs and neuronal plasticity in reward and aversion: focus on the ventral tegmental area. Front Pharmacol 5:256. DOI 10.3389/fphar.2014.00256
Vashchinkina E, Panhelainen A, Vekovischeva OY, Aitta-aho T, Ebert B, Ator NA,
and Korpi ER (2012) GABA site agonist gaboxadol induces addiction-predicting
persistent changes in ventral tegmental area dopamine neurons but is not rewarding in mice or baboons. J Neurosci 32:53105320.
Vasileiadis GT, Thompson RT, Han VK, and Gelman N (2009) Females follow a more
compact early human brain development model than males. A case-control study
of preterm neonates. Pediatr Res 66:551555.
Vastola BJ, Douglas LA, Varlinskaya EI, and Spear LP (2002) Nicotine-induced
conditioned place preference in adolescent and adult rats. Physiol Behav 77:
107114.
Vaughan CW, Connor M, Bagley EE, and Christie MJ (2000) Actions of cannabinoids
on membrane properties and synaptic transmission in rat periaqueductal gray
neurons in vitro. Mol Pharmacol 57:288295.
Vaughan CW, McGregor IS, and Christie MJ (1999) Cannabinoid receptor activation
inhibits GABAergic neurotransmission in rostral ventromedial medulla neurons in
vitro. Br J Pharmacol 127:935940.
Vaughan RA, Huff RA, Uhl GR, and Kuhar MJ (1997) Protein kinase C-mediated
phosphorylation and functional regulation of dopamine transporters in striatal
synaptosomes. J Biol Chem 272:1554115546.
Vekovischeva OY, Zamanillo D, Echenko O, Seppl T, Uusi-Oukari M, Honkanen A,
Seeburg PH, Sprengel R, and Korpi ER (2001) Morphine-induced dependence and
sensitization are altered in mice deficient in AMPA-type glutamate receptor-A
subunits. J Neurosci 21:44514459.
Velasquez KM, Molfese DL, and Salas R (2014) The role of the habenula in drug
addiction. Front Hum Neurosci 8:174. DOI 10.3389/fnhum.2014.00174
Vlez-Hernndez ME, Vzquez-Torres R, Velasquez-Martinez MC, Jimnez L, Bez
F, Sacktor TC, and Jimnez-Rivera CA (2013) Inhibition of protein kinase Mzeta
(PKMz) in the mesolimbic system alters cocaine sensitization in rats. J Drug Alcohol Res 2:235669.
Verma V, Lim EP, Han SP, Nagarajah R, and Dawe GS (2007) Chronic high-dose
haloperidol has qualitatively similar effects to risperidone and clozapine on
immediate-early gene and tyrosine hydroxylase expression in the rat locus coeruleus but not medial prefrontal cortex. Neurosci Res 57:1728.

1001

Vernon E, Meyer G, Pickard L, Dev K, Molnar E, Collingridge GL, and Henley JM


(2001) GABA(B) receptors couple directly to the transcription factor ATF4. Mol Cell
Neurosci 17:637645.
Verrico CD, Gu H, Peterson ML, Sampson AR, and Lewis DA (2014) Repeated D9tetrahydrocannabinol exposure in adolescent monkeys: persistent effects selective
for spatial working memory. Am J Psychiatry 171:416425.
Verrico CD, Jentsch JD, and Roth RH (2003) Persistent and anatomically selective
reduction in prefrontal cortical dopamine metabolism after repeated, intermittent
cannabinoid administration to rats. Synapse 49:6166.
Verrico CD, Miller GM, and Madras BK (2007) MDMA (Ecstasy) and human dopamine, norepinephrine, and serotonin transporters: implications for MDMA-induced
neurotoxicity and treatment. Psychopharmacology (Berl) 189:489503.
Vetter DE, Katz E, Maison SF, Taranda J, Turcan S, Ballestero J, Liberman MC,
Elgoyhen AB, and Boulter J (2007) The alpha10 nicotinic acetylcholine receptor
subunit is required for normal synaptic function and integrity of the olivocochlear
system. Proc Natl Acad Sci USA 104:2059420599.
Vialou V, Feng J, Robison AJ, Ku SM, Ferguson D, Scobie KN, Mazei-Robison MS,
Mouzon E, and Nestler EJ (2012) Serum response factor and cAMP response element binding protein are both required for cocaine induction of DFosB. J Neurosci
32:75777584.
Vihavainen T, Relander TR, Leivisk R, Airavaara M, Tuominen RK, Ahtee L,
and Piepponen TP (2008) Chronic nicotine modifies the effects of morphine on
extracellular striatal dopamine and ventral tegmental GABA. J Neurochem 107:
844854.
Vijayraghavan S, Wang M, Birnbaum SG, Williams GV, and Arnsten AF (2007)
Inverted-U dopamine D1 receptor actions on prefrontal neurons engaged in
working memory. Nat Neurosci 10:376384.
Vik PW (2007) Methamphetamine use by incarcerated women: comorbid mood and
anxiety problems. Womens Health Issues 17:256263.
Vlachou S and Markou A (2010) GABAB receptors in reward processes. Adv Pharmacol 58:315371.
Vlachou S, Nomikos GG, Stephens DN, and Panagis G (2007) Lack of evidence for
appetitive effects of Delta 9-tetrahydrocannabinol in the intracranial selfstimulation and conditioned place preference procedures in rodents. Behav Pharmacol 18:311319.
Volk LJ, Bachman JL, Johnson R, Yu Y, and Huganir RL (2013) PKM-z is not required for hippocampal synaptic plasticity, learning and memory. Nature 493:
420423.
Volkow ND, Chang L, Wang GJ, Fowler JS, Franceschi D, Sedler MJ, Gatley SJ,
Hitzemann R, Ding YS, and Wong C, et al. (2001) Higher cortical and lower subcortical metabolism in detoxified methamphetamine abusers. Am J Psychiatry 158:
383389.
Volkow ND, Fowler JS, and Ding YS (1996) Cardiotoxic properties of cocaine: studies
with positron emission tomography. NIDA Res Monogr 163:159174.
Volkow ND, Fowler JS, Wang GJ, Hitzemann R, Logan J, Schlyer DJ, Dewey SL,
and Wolf AP (1993) Decreased dopamine D2 receptor availability is associated with
reduced frontal metabolism in cocaine abusers. Synapse 14:169177.
Volkow ND, Fowler JS, Wolf AP, Hitzemann R, Dewey S, Bendriem B, Alpert R,
and Hoff A (1991) Changes in brain glucose metabolism in cocaine dependence and
withdrawal. Am J Psychiatry 148:621626.
Volkow ND, Fowler JS, Wolf AP, Schlyer D, Shiue CY, Alpert R, Dewey SL, Logan J,
Bendriem B, and Christman D, et al. (1990) Effects of chronic cocaine abuse on
postsynaptic dopamine receptors. Am J Psychiatry 147:719724.
Volkow ND, Hitzemann R, Wang GJ, Fowler JS, Wolf AP, Dewey SL,
and Handlesman L (1992) Long-term frontal brain metabolic changes in cocaine
abusers. Synapse 11:184190.
Volkow ND, Mullani N, Gould KL, Adler S, and Krajewski K (1988) Cerebral blood
flow in chronic cocaine users: a study with positron emission tomography. Br J
Psychiatry 152:641648.
Volkow ND, Tomasi D, Wang GJ, Logan J, Alexoff DL, Jayne M, Fowler JS, Wong C,
Yin P, and Du C (2014) Stimulant-induced dopamine increases are markedly
blunted in active cocaine abusers. Mol Psychiatry 19:10371043.
Volkow ND, Wang GJ, Fischman MW, Foltin RW, Fowler JS, Abumrad NN, Vitkun
S, Logan J, Gatley SJ, and Pappas N, et al. (1997a) Relationship between subjective effects of cocaine and dopamine transporter occupancy. Nature 386:
827830.
Volkow ND, Wang GJ, Fowler JS, Logan J, Gatley SJ, Hitzemann R, Chen AD,
Dewey SL, and Pappas N (1997b) Decreased striatal dopaminergic responsiveness
in detoxified cocaine-dependent subjects. Nature 386:830833.
Volkow ND, Wang GJ, Fowler JS, Thanos PP, Logan J, Gatley SJ, Gifford A, Ding
YS, Wong C, and Pappas N (2002) Brain DA D2 receptors predict reinforcing effects
of stimulants in humans: replication study. Synapse 46:7982.
Volkow ND, Wang GJ, Fowler JS, Tomasi D, Telang F, and Baler R (2010) Addiction:
decreased reward sensitivity and increased expectation sensitivity conspire to
overwhelm the brains control circuit. Bioessays 32:748755.
Volkow ND, Wang GJ, Ma Y, Fowler JS, Wong C, Ding YS, Hitzemann R, Swanson
JM, and Kalivas P (2005) Activation of orbital and medial prefrontal cortex by
methylphenidate in cocaine-addicted subjects but not in controls: relevance to
addiction. J Neurosci 25:39323939.
Vollenweider FX (1998) Advances and pathophysiological models of hallucinogenic
drug actions in humans: a preamble to schizophrenia research. Pharmacopsychiatry 31 (Suppl 2):92103.
Vollenweider FX (2001) Brain mechanisms of hallucinogens and entactogens. Dialogues Clin Neurosci 3:265279.
Vollenweider FX, Leenders KL, Scharfetter C, Antonini A, Maguire P, Missimer J,
and Angst J (1997a) Metabolic hyperfrontality and psychopathology in the ketamine model of psychosis using positron emission tomography (PET) and [18F]
fluorodeoxyglucose (FDG). Eur Neuropsychopharmacol 7:924.
Vollenweider FX, Leenders KL, Scharfetter C, Maguire P, Stadelmann O, and Angst
J (1997b) Positron emission tomography and fluorodeoxyglucose studies of

1002

Korpi et al.

metabolic hyperfrontality and psychopathology in the psilocybin model of psychosis. Neuropsychopharmacology 16:357372.
Vollenweider FX, Vollenweider-Scherpenhuyzen MF, Bbler A, Vogel H, and Hell D
(1998) Psilocybin induces schizophrenia-like psychosis in humans via a serotonin-2
agonist action. Neuroreport 9:38973902.
Volpicelli JR, Alterman AI, Hayashida M, and OBrien CP (1992) Naltrexone in the
treatment of alcohol dependence. Arch Gen Psychiatry 49:876880.
Volz TJ (2008) Neuropharmacological mechanisms underlying the neuroprotective
effects of methylphenidate. Curr Neuropharmacol 6:379385.
von Geusau NA, Stalenhoef P, Huizinga M, Snel J, and Ridderinkhof KR (2004)
Impaired executive function in male MDMA (ecstasy) users. Psychopharmacology
(Berl) 175:331341.
Voorhees CM and Cunningham CL (2011) Involvement of the orexin/hypocretin system
in ethanol conditioned place preference. Psychopharmacology (Berl) 214:805818.
Vorel SR, Liu X, Hayes RJ, Spector JA, and Gardner EL (2001) Relapse to cocaineseeking after hippocampal theta burst stimulation. Science 292:11751178.
Vorma H, Naukkarinen H, Sarna S, and Kuoppasalmi K (2002) Treatment of outpatients with complicated benzodiazepine dependence: comparison of two
approaches. Addiction 97:851859.
Vranjkovic O, Gasser PJ, Gerndt CH, Baker DA, and Mantsch JR (2014) Stressinduced cocaine seeking requires a beta-2 adrenergic receptor-regulated pathway
from the ventral bed nucleus of the stria terminalis that regulates CRF actions in
the ventral tegmental area. J Neurosci 34:1250412514.
Vrieze SI, McGue M, Miller MB, Hicks BM, and Iacono WG (2013) Three mutually
informative ways to understand the genetic relationships among behavioral disinhibition, alcohol use, drug use, nicotine use/dependence, and their co-occurrence:
twin biometry, GCTA, and genome-wide scoring. Behav Genet 43:97107.
Vutskits L, Gascon E, Tassonyi E, and Kiss JZ (2006) Effect of ketamine on dendritic
arbor development and survival of immature GABAergic neurons in vitro. Toxicol
Sci 91:540549.
Wagner D, Becker B, Koester P, Gouzoulis-Mayfrank E, and Daumann J (2013) A
prospective study of learning, memory, and executive function in new MDMA
users. Addiction 108:136145.
Wagner GC, Carelli RM, and Jarvis MF (1985) Pretreatment with ascorbic acid
attenuates the neurotoxic effects of methamphetamine in rats. Res Commun Chem
Pathol Pharmacol 47:221228.
Wagner GC, Ricaurte GA, Seiden LS, Schuster CR, Miller RJ, and Westley J (1980)
Long-lasting depletions of striatal dopamine and loss of dopamine uptake sites
following repeated administration of methamphetamine. Brain Res 181:151160.
Walgren JL, Carfagna MA, Koger D, Sgro M, and Kallman MJ (2014) Withdrawal
assessment following subchronic oral ketamine administration in Cynomolgus
macaques. Drug Dev Res 75:162171.
Walker B, Anderson M, Hauser L, and Werchan I (2013) Ethanol for alcohol withdrawal: the end of an era. J Trauma Acute Care Surg 74:926931.
Walker BM and Koob GF (2008) Pharmacological evidence for a motivational role of
kappa-opioid systems in ethanol dependence. Neuropsychopharmacology 33:
643652.
Wallace EA, Wisniewski G, Zubal G, vanDyck CH, Pfau SE, Smith EO, Rosen MI,
Sullivan MC, Woods SW, and Kosten TR (1996) Acute cocaine effects on absolute
cerebral blood flow. Psychopharmacology (Berl) 128:1720.
Wallner M, Hanchar HJ, and Olsen RW (2006) Low-dose alcohol actions on alpha4beta3delta GABAA receptors are reversed by the behavioral alcohol antagonist
Ro15-4513. Proc Natl Acad Sci USA 103:85408545.
Walsh H, Govind AP, Mastro R, Hoda JC, Bertrand D, Vallejo Y, and Green WN
(2008) Up-regulation of nicotinic receptors by nicotine varies with receptor subtype. J Biol Chem 283:60226032.
Walsh JJ, Friedman AK, Sun H, Heller EA, Ku SM, Juarez B, Burnham VL, MazeiRobison MS, Ferguson D, and Golden SA, et al. (2014) Stress and CRF gate neural
activation of BDNF in the mesolimbic reward pathway. Nat Neurosci 17:2729.
Walters CL and Blendy JA (2001) Different requirements for cAMP response element
binding protein in positive and negative reinforcing properties of drugs of abuse.
J Neurosci 21:94389444.
Wanat MJ, Hopf FW, Stuber GD, Phillips PE, and Bonci A (2008) Corticotropinreleasing factor increases mouse ventral tegmental area dopamine neuron firing
through a protein kinase C-dependent enhancement of Ih. J Physiol 586:
21572170.
Wanat MJ, Sparta DR, Hopf FW, Bowers MS, Melis M, and Bonci A (2009) Strain
specific synaptic modifications on ventral tegmental area dopamine neurons after
ethanol exposure. Biol Psychiatry 65:646653.
Wang B, Shaham Y, Zitzman D, Azari S, Wise RA, and You ZB (2005) Cocaine
experience establishes control of midbrain glutamate and dopamine by
corticotropin-releasing factor: a role in stress-induced relapse to drug seeking. J
Neurosci 25:53895396.
Wang B, You ZB, Rice KC, and Wise RA (2007) Stress-induced relapse to cocaine
seeking: roles for the CRF(2) receptor and CRF-binding protein in the ventral
tegmental area of the rat. Psychopharmacology (Berl) 193:283294.
Wang C, Liu F, Patterson TA, Paule MG, and Slikker W Jr (2013a) Preclinical assessment of ketamine. CNS Neurosci Ther 19:448453.
Wang C, Showalter VM, Hillman GR, and Johnson KM (1999) Chronic phencyclidine
increases NMDA receptor NR1 subunit mRNA in rat forebrain. J Neurosci Res 55:
762769.
Wang C, Zheng D, Xu J, Lam W, and Yew DT (2013b) Brain damages in ketamine
addicts as revealed by magnetic resonance imaging. Front Neuroanat 7:23.
DOI 10.3389/fnana.2013.00023
Wang CZ, Yang SF, Xia Y, and Johnson KM (2008a) Postnatal phencyclidine administration selectively reduces adult cortical parvalbumin-containing interneurons. Neuropsychopharmacology 33:24422455.
Wang F, Nelson ME, Kuryatov A, Olale F, Cooper J, Keyser K, and Lindstrom J
(1998) Chronic nicotine treatment up-regulates human alpha3 beta2 but not

alpha3 beta4 acetylcholine receptors stably transfected in human embryonic kidney cells. J Biol Chem 273:2872128732.
Wang GJ, Volkow ND, Overall J, Hitzemann RJ, Pappas N, Pascani K, and Fowler
JS (1996) Reproducibility of regional brain metabolic responses to lorazepam.
J Nucl Med 37:16091613.
Wang J, Fanous S, Terwilliger EF, Bass CE, Hammer RP Jr, and Nikulina EM
(2013c) BDNF overexpression in the ventral tegmental area prolongs social defeat
stress-induced cross-sensitization to amphetamine and increases DFosB expression in mesocorticolimbic regions of rats. Neuropsychopharmacology 38:22862296.
Wang JC, Kapoor M, and Goate AM (2012a) The genetics of substance dependence.
Annu Rev Genomics Hum Genet 13:241261.
Wang RR, Jin JH, Womack AW, Lyu D, Kokane SS, Tang N, Zou X, Lin Q, and Chen
J (2014) Neonatal ketamine exposure causes impairment of long-term synaptic
plasticity in the anterior cingulate cortex of rats. Neuroscience 268:309317.
Wang SF, Yen JC, Yin PH, Chi CW, and Lee HC (2008b) Involvement of oxidative
stress-activated JNK signaling in the methamphetamine-induced cell death of
human SH-SY5Y cells. Toxicology 246:234241.
Wang X, Cahill ME, Werner CT, Christoffel DJ, Golden SA, Xie Z, Loweth JA,
Marinelli M, Russo SJ, and Penzes P, et al. (2013d) Kalirin-7 mediates cocaineinduced AMPA receptor and spine plasticity, enabling incentive sensitization. J
Neurosci 33:1101211022.
Wang X, Loram LC, Ramos K, de Jesus AJ, Thomas J, Cheng K, Reddy A, Somogyi
AA, Hutchinson MR, and Watkins LR, et al. (2012b) Morphine activates neuroinflammation in a manner parallel to endotoxin. Proc Natl Acad Sci USA 109:
63256330.
Wang X, Wang G, Lemos JR, and Treistman SN (1994) Ethanol directly modulates
gating of a dihydropyridine-sensitive Ca2+ channel in neurohypophysial terminals.
J Neurosci 14:54535460.
Warburton DM, Wesnes K, Edwards J, and Larrad D (1985) Scopolamine and the
sensory conditioning of hallucinations. Neuropsychobiology 14:198202.
Ware JJ, van den Bree MB, and Munaf MR (2011) Association of the CHRNA5-A3-B4
gene cluster with heaviness of smoking: a meta-analysis. Nicotine Tob Res 13:11671175.
Warpman U, Friberg L, Gillespie A, Hellstrm-Lindahl E, Zhang X, and Nordberg A
(1998) Regulation of nicotinic receptor subtypes following chronic nicotinic agonist
exposure in M10 and SH-SY5Y neuroblastoma cells. J Neurochem 70:20282037.
Warren MW, Zheng W, Kobeissy FH, Cheng Liu M, Hayes RL, Gold MS, Larner SF,
and Wang KK (2007) Calpain- and caspase-mediated alphaII-spectrin and tau
proteolysis in rat cerebrocortical neuronal cultures after ecstasy or methamphetamine exposure. Int J Neuropsychopharmacol 10:479489.
Watkins LR, Hutchinson MR, Johnston IN, and Maier SF (2005) Glia: novel counterregulators of opioid analgesia. Trends Neurosci 28:661669.
Watkins LR, Hutchinson MR, Rice KC, and Maier SF (2009) The toll of opioidinduced glial activation: improving the clinical efficacy of opioids by targeting glia.
Trends Pharmacol Sci 30:581591.
Watterson LR, Kufahl PR, Nemirovsky NE, Sewalia K, Hood LE, and Olive MF
(2013) Attenuation of reinstatement of methamphetamine-, sucrose-, and foodseeking behavior in rats by fenobam, a metabotropic glutamate receptor 5 negative
allosteric modulator. Psychopharmacology (Berl) 225:151159.
Watts VJ and Neve KA (2005) Sensitization of adenylate cyclase by Galpha
i/o-coupled receptors. Pharmacol Ther 106:405421.
Wedzony K, Chocyk A, and Mackowiak M(2013) Potential roles of NCAM/PSANCAM proteins in depression and the mechanism of action of antidepressant
drugs. Pharmacol Rep 65:14711478.
Wedzony K, Klimek V, and Goembiowska K (1993) MK-801 elevates the extracellular concentration of dopamine in the rat prefrontal cortex and increases the
density of striatal dopamine D1 receptors. Brain Res 622:325329.
Weeks HR 3rd, Tadler SC, Smith KW, Iacob E, Saccoman M, White AT, Landvatter
JD, Chelune GJ, Suchy Y, and Clark E, et al. (2013) Antidepressant and neurocognitive effects of isoflurane anesthesia versus electroconvulsive therapy in refractory depression. PLoS One 8:e69809.
Wegelius K and Korpi ER (1995) Ethanol inhibits NMDA-induced toxicity and trophism in cultured cerebellar granule cells. Acta Physiol Scand 154:2534.
Wei W, Faria LC, and Mody I (2004) Low ethanol concentrations selectively augment
the tonic inhibition mediated by delta subunit-containing GABAA receptors in
hippocampal neurons. J Neurosci 24:83798382.
Weiland BJ, Thayer RE, Depue BE, Sabbineni A, Bryan AD, and Hutchison KE
(2015) Daily marijuana use is not associated with brain morphometric measures in
adolescents or adults. J Neurosci 35:15051512.
Weissenborn R and Nutt DJ (2012) Popular intoxicants: what lessons can be learned
from the last 40 years of alcohol and cannabis regulation? J Psychopharmacol 26:
213220.
Weisskopf MG, Castillo PE, Zalutsky RA, and Nicoll RA (1994) Mediation of hippocampal mossy fiber long-term potentiation by cyclic AMP. Science 265:18781882.
Welsh BT, Goldstein BE, and Mihic SJ (2009) Single-channel analysis of ethanol
enhancement of glycine receptor function. J Pharmacol Exp Ther 330:198205.
West EA, Saddoris MP, Kerfoot EC, and Carelli RM (2014) Prelimbic and infralimbic
cortical regions differentially encode cocaine-associated stimuli and cocaineseeking before and following abstinence. Eur J Neurosci 39:18911902.
Westra HA and Stewart SH (1998) Cognitive behavioural therapy and pharmacotherapy: complementary or contradictory approaches to the treatment of anxiety?
Clin Psychol Rev 18:307340.
Wexler BE, Gottschalk CH, Fulbright RK, Prohovnik I, Lacadie CM, Rounsaville BJ,
and Gore JC (2001) Functional magnetic resonance imaging of cocaine craving. Am
J Psychiatry 158:8695.
Wheeler D, Boutelle MG, and Fillenz M (1995) The role of N-methyl-D-aspartate
receptors in the regulation of physiologically released dopamine. Neuroscience 65:
767774.
Whistler JL, Chuang HH, Chu P, Jan LY, and von Zastrow M (1999) Functional
dissociation of mu opioid receptor signaling and endocytosis: implications for the
biology of opiate tolerance and addiction. Neuron 23:737746.

Drug-Induced Neuroplasticity
White DA and Holtzman SG (2005) Periadolescent morphine exposure alters subsequent behavioral sensitivity to morphine in adult rats. Eur J Pharmacol 528:
119123.
Whiteaker P, Sharples CG, and Wonnacott S (1998) Agonist-induced up-regulation of
alpha4beta2 nicotinic acetylcholine receptors in M10 cells: pharmacological and
spatial definition. Mol Pharmacol 53:950962.
Whiting PJ, McKernan RM, and Wafford KA (1995) Structure and pharmacology of
vertebrate GABAA receptor subtypes. Int Rev Neurobiol 38:95138.
Wiebelhaus JM, Grim TW, Owens RA, Lazenka MF, Sim-Selley LJ, Abdullah RA,
Niphakis MJ, Vann RE, Cravatt BF, and Wiley JL, et al. (2015) D9-tetrahydrocannabinol and endocannabinoid degradative enzyme inhibitors attenuate intracranial self-stimulation in mice. J Pharmacol Exp Ther 352:195207.
Wieland HA, Lddens H, and Seeburg PH (1992) A single histidine in GABAA
receptors is essential for benzodiazepine agonist binding. J Biol Chem 267:
14261429.
Wiescholleck V and Manahan-Vaughan D (2013) Long-lasting changes in hippocampal synaptic plasticity and cognition in an animal model of NMDA receptor
dysfunction in psychosis. Neuropharmacology 74:4858.
Wiesenfeld-Hallin Z, de Araja Lucas G, Alster P, Xu XJ, and Hkfelt T (1999)
Cholecystokinin/opioid interactions. Brain Res 848:7889.
Wiggins A, Smith RJ, Shen HW, and Kalivas PW (2011) Integrins modulate relapse
to cocaine-seeking. J Neurosci 31:1617716184.
Wilcox MV, Cuzon Carlson VC, Sherazee N, Sprow GM, Bock R, Thiele TE, Lovinger
DM, and Alvarez VA (2014) Repeated binge-like ethanol drinking alters ethanol
drinking patterns and depresses striatal GABAergic transmission. Neuropsychopharmacology 39:579594.
Wilens TE, Adler LA, Adams J, Sgambati S, Rotrosen J, Sawtelle R, Utzinger L,
and Fusillo S (2008) Misuse and diversion of stimulants prescribed for ADHD:
a systematic review of the literature. J Am Acad Child Adolesc Psychiatry 47:
2131.
Wilens TE, Faraone SV, Biederman J, and Gunawardene S (2003) Does stimulant
therapy of attention-deficit/hyperactivity disorder beget later substance abuse? A
meta-analytic review of the literature. Pediatrics 111:179185.
Wiley JL and Compton AD (2004) Progressive ratio performance following challenge
with antipsychotics, amphetamine, or NMDA antagonists in adult rats treated
perinatally with phencyclidine. Psychopharmacology (Berl) 177:170177.
Williams JT, Christie MJ, and Manzoni O (2001) Cellular and synaptic adaptations
mediating opioid dependence. Physiol Rev 81:299343.
Wilson C, Canning P, and Caravati EM (2010) The abuse potential of propofol. Clin
Toxicol (Phila) 48:165170.
Wilson JM, Kalasinsky KS, Levey AI, Bergeron C, Reiber G, Anthony RM, Schmunk
GA, Shannak K, Haycock JW, and Kish SJ (1996) Striatal dopamine nerve terminal markers in human, chronic methamphetamine users. Nat Med 2:699703.
Wilson RI and Nicoll RA (2001) Endogenous cannabinoids mediate retrograde signalling at hippocampal synapses. Nature 410:588592.
Wilson RI and Nicoll RA (2002) Endocannabinoid signaling in the brain. Science 296:
678682.
Wilson SJ and Sayette MA (2015) Neuroimaging craving: urge intensity matters.
Addiction 110:195203.
Winsauer PJ, Quinton MS, Porter JR, Corll CB, Moerschbaecher JM, Delatte MS,
Leonard ST, and Stroble SB (2004) Effects of MDMA administration on
scopolamine-induced disruptions of learning and performance in rats. Pharmacol
Biochem Behav 79:459472.
Winter JC, Fiorella DJ, Timineri DM, Filipink RA, Helsley SE, and Rabin RA (1999)
Serotonergic receptor subtypes and hallucinogen-induced stimulus control. Pharmacol Biochem Behav 64:283293.
Wisden W, Laurie DJ, Monyer H, and Seeburg PH (1992) The distribution of 13
GABAA receptor subunit mRNAs in the rat brain. I. Telencephalon, diencephalon,
mesencephalon. J Neurosci 12:10401062.
Wisden W and Seeburg PH (1992) GABAA receptor channels: from subunits to
functional entities. Curr Opin Neurobiol 2:263269.
Wise RA (1973) Voluntary ethanol intake in rats following exposure to ethanol on
various schedules. Psychopharmacology (Berl) 29:203210.
Wise RA (1996) Addictive drugs and brain stimulation reward. Annu Rev Neurosci
19:319340.
Wise RA (2006) Role of brain dopamine in food reward and reinforcement. Philos
Trans R Soc Lond B Biol Sci 361:11491158.
Wise RA and Morales M (2010) A ventral tegmental CRF-glutamate-dopamine interaction in addiction. Brain Res 1314:3843.
Wise RA and Rompre PP (1989) Brain dopamine and reward. Annu Rev Psychol 40:
191225.
Wislowska-Stanek A, Lehner M, Skorzewska A, Bidzinski A, Turzynska D,
Sobolewska A, Maciejak P, Szyndler J, and Plaznik A(2008) Inhibition of
neophobia-stimulated c-Fos expression in the dorsomedial part of the prefrontal
cortex in rats pretreated with midazolam. Pharmacol Rep 60:811816.
Witkin JM, Statnick MA, Rorick-Kehn LM, Pintar JE, Ansonoff M, Chen Y, Tucker RC,
and Ciccocioppo R (2014) The biology of Nociceptin/Orphanin FQ (N/OFQ) related to
obesity, stress, anxiety, mood, and drug dependence. Pharmacol Ther 141:283299.
Witkin JM, Tzavara ET, Davis RJ, Li X, and Nomikos GG (2005) A therapeutic role
for cannabinoid CB1 receptor antagonists in major depressive disorders. Trends
Pharmacol Sci 26:609617.
Witschi R, Punnakkal P, Paul J, Walczak JS, Cervero F, Fritschy JM, Kuner R, Keist
R, Rudolph U, and Zeilhofer HU (2011) Presynaptic alpha2-GABAA receptors in
primary afferent depolarization and spinal pain control. J Neurosci 31:81348142.
Wohrl R, Eisenach S, Manahan-Vaughan D, Heinemann U, and von Haebler D (2007)
Acute and long-term effects of MK-801 on direct cortical input evoked homosynaptic and heterosynaptic plasticity in the CA1 region of the female rat. Eur J
Neurosci 26:28732883.
Wolf ME (2002) Addiction: making the connection between behavioral changes and
neuronal plasticity in specific pathways. Mol Interv 2:146157.

1003

Wolf ME and Ferrario CR (2010) AMPA receptor plasticity in the nucleus accumbens
after repeated exposure to cocaine. Neurosci Biobehav Rev 35:185211.
Wolf ME, Mangiavacchi S, and Sun X (2003) Mechanisms by which dopamine
receptors may influence synaptic plasticity. Ann N Y Acad Sci 1003:241249.
Wolf ME, Xue CJ, Li Y, and Wavak D (2000) Amphetamine increases glutamate
efflux in the rat ventral tegmental area by a mechanism involving glutamate
transporters and reactive oxygen species. J Neurochem 75:16341644.
Wooltorton JR, Pidoplichko VI, Broide RS, and Dani JA (2003) Differential desensitization and distribution of nicotinic acetylcholine receptor subtypes in midbrain dopamine areas. J Neurosci 23:31763185.
Woolverton WL, Ricaurte GA, Forno LS, and Seiden LS (1989) Long-term effects of
chronic methamphetamine administration in rhesus monkeys. Brain Res 486:
7378.
Workman AD, Charvet CJ, Clancy B, Darlington RB, and Finlay BL (2013) Modeling
transformations of neurodevelopmental sequences across mammalian species.
J Neurosci 33:73687383.
Wright MJ Jr, Angrish D, Aarde SM, Barlow DJ, Buczynski MW, Creehan KM,
Vandewater SA, Parsons LH, Houseknecht KL, and Dickerson TJ, et al. (2012)
Effect of ambient temperature on the thermoregulatory and locomotor stimulant effects
of 4-methylmethcathinone in Wistar and Sprague-Dawley rats. PLoS One 7:e44652.
Wu X and Castrn E (2009) Co-treatment with diazepam prevents the effects of
fluoxetine on the proliferation and survival of hippocampal dentate granule cells.
Biol Psychiatry 66:58.
Wu X and Reddy DS (2012) Integrins as receptor targets for neurological disorders.
Pharmacol Ther 134:6881.
Wu YN, Shen KZ, and Johnson SW (1999) Presynaptic inhibition preferentially
reduces in NMDA receptor-mediated component of transmission in rat midbrain
dopamine neurons. Br J Pharmacol 127:14221430.
Xi ZX, Peng XQ, Li X, Song R, Zhang HY, Liu QR, Yang HJ, Bi GH, Li J, and Gardner
EL (2011) Brain cannabinoid CB receptors modulate cocaines actions in mice. Nat
Neurosci 14:11601166.
Xi ZX, Spiller K, Pak AC, Gilbert J, Dillon C, Li X, Peng XQ, and Gardner EL (2008)
Cannabinoid CB1 receptor antagonists attenuate cocaines rewarding effects:
experiments with self-administration and brain-stimulation reward in rats. Neuropsychopharmacology 33:17351745.
Xia Y, Portugal GS, Fakira AK, Melyan Z, Neve R, Lee HT, Russo SJ, Liu J,
and Morn JA (2011) Hippocampal GluA1-containing AMPA receptors mediate
context-dependent sensitization to morphine. J Neurosci 31:1627916291.
Xiang K, Earl DE, Davis KM, Giovannucci DR, Greenfield LJ Jr, and Tietz EI (2008)
Chronic benzodiazepine administration potentiates high voltage-activated calcium
currents in hippocampal CA1 neurons. J Pharmacol Exp Ther 327:872883.
Xiang K and Tietz EI (2007) Benzodiazepine-induced hippocampal CA1 neuron
alpha-amino-3-hydroxy-5-methylisoxasole-4-propionic acid (AMPA) receptor plasticity linked to severity of withdrawal anxiety: differential role of voltage-gated
calcium channels and N-methyl-D-aspartic acid receptors. Behav Pharmacol 18:
447460.
Xiao C, Nashmi R, McKinney S, Cai H, McIntosh JM, and Lester HA (2009) Chronic
nicotine selectively enhances alpha4beta2* nicotinic acetylcholine receptors in the
nigrostriatal dopamine pathway. J Neurosci 29:1242812439.
Xiao C, Zhou C, Li K, Davies DL, and Ye JH (2008) Purinergic type 2 receptors at
GABAergic synapses on ventral tegmental area dopamine neurons are targets for
ethanol action. J Pharmacol Exp Ther 327:196205.
Xie Z and Miller GM (2009) Trace amine-associated receptor 1 as a monoaminergic
modulator in brain. Biochem Pharmacol 78:10951104.
Xu C, Zhang W, Rondard P, Pin JP, and Liu J (2014) Complex GABAB receptor
complexes: how to generate multiple functionally distinct units from a single receptor. Front Pharmacol 5:12.
Xu F, Gainetdinov RR, Wetsel WC, Jones SR, Bohn LM, Miller GW, Wang YM,
and Caron MG (2000) Mice lacking the norepinephrine transporter are supersensitive to psychostimulants. Nat Neurosci 3:465471.
Xu H, Kotak VC, and Sanes DH (2010) Normal hearing is required for the emergence
of long-lasting inhibitory potentiation in cortex. J Neurosci 30:331341.
Xu SX, Zhou ZQ, Li XM, Ji MH, Zhang GF, and Yang JJ (2013) The activation of
adenosine monophosphate-activated protein kinase in rat hippocampus contributes to the rapid antidepressant effect of ketamine. Behav Brain Res 253:305309.
Xue YX, Xue LF, Liu JF, He J, Deng JH, Sun SC, Han HB, Luo YX, Xu LZ, and Wu P,
et al. (2014) Depletion of perineuronal nets in the amygdala to enhance the erasure
of drug memories. J Neurosci 34:66476658.
Yamada S, Harano M, Annoh N, Nakamura K, and Tanaka M (1999) Involvement of
serotonin 2A receptors in phencyclidine-induced disruption of prepulse inhibition
of the acoustic startle in rats. Biol Psychiatry 46:832838.
Yamamoto BK, Moszczynska A, and Gudelsky GA (2010) Amphetamine toxicities:
classical and emerging mechanisms. Ann N Y Acad Sci 1187:101121.
Yamamoto BK and Raudensky J (2008) The role of oxidative stress, metabolic compromise, and inflammation in neuronal injury produced by amphetamine-related
drugs of abuse. J Neuroimmune Pharmacol 3:203217.
Yamamoto BK and Zhu W (1998) The effects of methamphetamine on the production
of free radicals and oxidative stress. J Pharmacol Exp Ther 287:107114.
Yan QS, Reith ME, Jobe PC, and Dailey JW (1997) Dizocilpine (MK-801) increases
not only dopamine but also serotonin and norepinephrine transmissions in the
nucleus accumbens as measured by microdialysis in freely moving rats. Brain Res
765:149158.
Yanai J, Avraham Y, Levy S, Maslaton J, Pick CG, Rogel-Fuchs Y, and Zahalka EA
(1992) Alterations in septohippocampal cholinergic innervations and related
behaviors after early exposure to heroin and phencyclidine. Brain Res Dev Brain
Res 69:207214.
Yang X, Yang Q, Wang X, Luo C, Wan Y, Li J, Liu K, Zhou M, and Zhang C (2014)
MicroRNA expression profile and functional analysis reveal that miR-206 is
a critical novel gene for the expression of BDNF induced by ketamine. Neuromolecular Med 16:594605.

1004

Korpi et al.

Yap JJ, Covington HE 3rd, Gale MC, Datta R, and Miczek KA (2005) Behavioral
sensitization due to social defeat stress in mice: antagonism at mGluR5 and NMDA
receptors. Psychopharmacology (Berl) 179:230239.
Yardley MM, Neely M, Huynh N, Asatryan L, Louie SG, Alkana RL, and Davies DL
(2014) Multiday administration of ivermectin is effective in reducing alcohol intake
in mice at doses shown to be safe in humans. Neuroreport 25:10181023.
Yardley MM, Wyatt L, Khoja S, Asatryan L, Ramaker MJ, Finn DA, Alkana RL,
Huynh N, Louie SG, and Petasis NA, et al. (2012) Ivermectin reduces alcohol
intake and preference in mice. Neuropharmacology 63:190201.
Yasoshima Y and Yamamoto T (2005) Effects of midazolam on the expression of
conditioned taste aversion in rats. Brain Res 1043:115123.
Yates SL, Bencherif M, Fluhler EN, and Lippiello PM (1995) Up-regulation of nicotinic
acetylcholine receptors following chronic exposure of rats to mainstream cigarette
smoke or alpha 4 beta 2 receptors to nicotine. Biochem Pharmacol 50:20012008.
Ye JH, Tao L, Zhu L, Krnjevic K, and McArdle JJ (2001) Ethanol inhibition of glycineactivated responses in neurons of ventral tegmental area of neonatal rats.
J Neurophysiol 86:24262434.
Ye JH, Wang F, Krnjevic K, Wang W, Xiong ZG, and Zhang J (2004) Presynaptic
glycine receptors on GABAergic terminals facilitate discharge of dopaminergic
neurons in ventral tegmental area. J Neurosci 24:89618974.
Ye Q and Miao QL (2013) Experience-dependent development of perineuronal nets
and chondroitin sulfate proteoglycan receptors in mouse visual cortex. Matrix Biol
32:352363.
Yeckel MF, Kapur A, and Johnston D (1999) Multiple forms of LTP in hippocampal
CA3 neurons use a common postsynaptic mechanism. Nat Neurosci 2:625633.
Yee BK, Hauser J, Dolgov VV, Keist R, Mhler H, Rudolph U, and Feldon J (2004) GABA
receptors containing the alpha5 subunit mediate the trace effect in aversive and appetitive conditioning and extinction of conditioned fear. Eur J Neurosci 20:19281936.
Yee BK, Keist R, von Boehmer L, Studer R, Benke D, Hagenbuch N, Dong Y,
Malenka RC, Fritschy JM, and Bluethmann H, et al. (2005) A schizophreniarelated sensorimotor deficit links alpha 3-containing GABAA receptors to a dopamine
hyperfunction. Proc Natl Acad Sci USA 102:1715417159.
Yeh SY and De Souza EB (1991) Lack of neurochemical evidence for neurotoxic
effects of repeated cocaine administration in rats on brain monoamine neurons.
Drug Alcohol Depend 27:5161.
Yevenes GE, Moraga-Cid G, Peoples RW, Schmalzing G, and Aguayo LG (2008) A
selective G betagamma-linked intracellular mechanism for modulation of a ligandgated ion channel by ethanol. Proc Natl Acad Sci USA 105:2052320528.
Yevenes GE, Peoples RW, Tapia JC, Parodi J, Soto X, Olate J, and Aguayo LG (2003)
Modulation of glycine-activated ion channel function by G-protein betagamma
subunits. Nat Neurosci 6:819824.
Yin HH, Knowlton BJ, and Balleine BW (2004) Lesions of dorsolateral striatum
preserve outcome expectancy but disrupt habit formation in instrumental learning.
Eur J Neurosci 19:181189.
Yin HH, Knowlton BJ, and Balleine BW (2005a) Blockade of NMDA receptors in the
dorsomedial striatum prevents action-outcome learning in instrumental conditioning. Eur J Neurosci 22:505512.
Yin HH, Ostlund SB, Knowlton BJ, and Balleine BW (2005b) The role of the dorsomedial striatum in instrumental conditioning. Eur J Neurosci 22:513523.
Yokel RA and Pickens R (1973) Self-administration of optical isomers of amphetamine and methylamphetamine by rats. J Pharmacol Exp Ther 187:2733.
Yonezaki K, Uchimoto K, Miyazaki T, Asakura A, Kobayashi A, Takase K, and Goto T
(2015) Postanesthetic effects of isoflurane on behavioral phenotypes of adult male
C57BL/6J mice. PLoS One 10:e0122118.
Yoshida T, Uchigashima M, Yamasaki M, Katona I, Yamazaki M, Sakimura K, Kano
M, Yoshioka M, and Watanabe M (2011) Unique inhibitory synapse with particularly rich endocannabinoid signaling machinery on pyramidal neurons in basal
amygdaloid nucleus. Proc Natl Acad Sci USA 108:30593064.
Young C, Jevtovic-Todorovic V, Qin YQ, Tenkova T, Wang H, Labruyere J, and Olney
JW (2005) Potential of ketamine and midazolam, individually or in combination, to
induce apoptotic neurodegeneration in the infant mouse brain. Br J Pharmacol
146:189197.
Young C, Straiko MM, Johnson SA, Creeley C, and Olney JW (2008) Ethanol causes
and lithium prevents neuroapoptosis and suppression of pERK in the infant mouse
brain. Neurobiol Dis 31:355360.
Yu D, Zhang L, Eisel J-L, Bertrand D, Changeux J-P, and Weight FF (1996) Ethanol
inhibition of nicotinic acetylcholine type alpha 7 receptors involves the aminoterminal domain of the receptor. Mol Pharmacol 50:10101016.
Yu X, Wang G, Gilmore A, Yee AX, Li X, Xu T, Smith SJ, Chen L, and Zuo Y (2013)
Accelerated experience-dependent pruning of cortical synapses in ephrin-A2
knockout mice. Neuron 80:6471.
Yuan J, McCann U, and Ricaurte G (1997) Methylphenidate and brain dopamine
neurotoxicity. Brain Res 767:172175.
Yuan T, Mameli M, O Connor EC, Dey PN, Verpelli C, Sala C, Perez-Otano I,
Lscher C, and Bellone C (2013) Expression of cocaine-evoked synaptic plasticity
by GluN3A-containing NMDA receptors. Neuron 80:10251038.
Ycel M, Solowij N, Respondek C, Whittle S, Fornito A, Pantelis C, and Lubman DI
(2008) Regional brain abnormalities associated with long-term heavy cannabis use.
Arch Gen Psychiatry 65:694701.
Zaczek R, Culp S, and De Souza EB (1991) Interactions of [3H]amphetamine with rat
brain synaptosomes. II. Active transport. J Pharmacol Exp Ther 257:830835.
Zajaczkowski W, Hetman M, Nikolaev E, Quack G, Danysz W, and Kaczmarek L
(2000) Behavioural evaluation of long-term neurotoxic effects of NMDA receptor
antagonists. Neurotox Res 1:299310.
Zalesky A, Solowij N, Ycel M, Lubman DI, Takagi M, Harding IH, Lorenzetti V,
Wang R, Searle K, and Pantelis C, et al. (2012) Effect of long-term cannabis use on
axonal fibre connectivity. Brain 135:22452255.
Zammit S, Allebeck P, Andreasson S, Lundberg I, and Lewis G (2002) Self reported
cannabis use as a risk factor for schizophrenia in Swedish conscripts of 1969:
historical cohort study. BMJ 325:11991203.

Zangen A, Solinas M, Ikemoto S, Goldberg SR, and Wise RA (2006) Two brain sites
for cannabinoid reward. J Neurosci 26:49014907.
Zapata A, Minney VL, and Shippenberg TS (2010) Shift from goal-directed to habitual cocaine seeking after prolonged experience in rats. J Neurosci 30:
1545715463.
Zarate CA Jr, Brutsche N, Laje G, Luckenbaugh DA, Venkata SL, Ramamoorthy A,
Moaddel R, and Wainer IW (2012a) Relationship of ketamines plasma metabolites with
response, diagnosis, and side effects in major depression. Biol Psychiatry 72:331338.
Zarate CA Jr, Brutsche NE, Ibrahim L, Franco-Chaves J, Diazgranados N, Cravchik
A, Selter J, Marquardt CA, Liberty V, and Luckenbaugh DA (2012b) Replication of
ketamines antidepressant efficacy in bipolar depression: a randomized controlled
add-on trial. Biol Psychiatry 71:939946.
Zeilhofer HU, Wildner H, and Yvenes GE (2012) Fast synaptic inhibition in spinal
sensory processing and pain control. Physiol Rev 92:193235.
Zemkova H, Tvrdonova V, Bhattacharya A, and Jindrichova M (2014) Allosteric modulation of ligand gated ion channels by ivermectin. Physiol Res 63 (Suppl 1):S215S224.
Zhang HY, Bi GH, Li X, Li J, Qu H, Zhang SJ, Li CY, Onaivi ES, Gardner EL, and Xi ZX,
et al. (2015a) Species differences in cannabinoid receptor 2 and receptor responses to
cocaine self-administration in mice and rats. Neuropsychopharmacology 40:10371051.
Zhang HY, Gao M, Liu QR, Bi GH, Li X, Yang HJ, Gardner EL, Wu J, and Xi ZX (2014)
Cannabinoid CB2 receptors modulate midbrain dopamine neuronal activity and
dopamine-related behavior in mice. Proc Natl Acad Sci USA 111:E5007E5015.
Zhang L, Coffey LL, and Reith ME (1997) Regulation of the functional activity of the
human dopamine transporter by protein kinase C. Biochem Pharmacol 53:677688.
Zhang L, Li J, Liu N, Wang B, Gu J, Zhang M, Zhou Z, Jiang Y, Zhang L, and Zhang L
(2012) Signaling via dopamine D1 and D3 receptors oppositely regulates cocaineinduced structural remodeling of dendrites and spines. Neurosignals 20:1534.
Zhang S, Edelmann L, Liu J, Crandall JE, and Morabito MA (2008) Cdk5 regulates
the phosphorylation of tyrosine 1472 NR2B and the surface expression of NMDA
receptors. J Neurosci 28:415424.
Zhang T, Zhang L, Liang Y, Siapas AG, Zhou FM, and Dani JA (2009) Dopamine
signaling differences in the nucleus accumbens and dorsal striatum exploited by
nicotine. J Neurosci 29:40354043.
Zhang X, Ju G, and Le Gal La Salle G (1991) Fos expression in GHB-induced generalized absence epilepsy in the thalamus of the rat. Neuroreport 2:469472.
Zhang Y, Butelman ER, Schlussman SD, Ho A, and Kreek MJ (2005) Effects of the
plant-derived hallucinogen salvinorin A on basal dopamine levels in the caudate
putamen and in a conditioned place aversion assay in mice: agonist actions at
kappa opioid receptors. Psychopharmacology (Berl) 179:551558.
Zhang Y, Cudmore RH, Lin DT, Linden DJ, and Huganir RL (2015b) Visualization of
NMDA receptor-dependent AMPA receptor synaptic plasticity in vivo. Nat Neurosci 18:402407.
Zhao X, Venkata SL, Moaddel R, Luckenbaugh DA, Brutsche NE, Ibrahim L, Zarate
CA Jr, Mager DE, and Wainer IW (2012) Simultaneous population pharmacokinetic modelling of ketamine and three major metabolites in patients with
treatment-resistant bipolar depression. Br J Clin Pharmacol 74:304314.
Zheng Y, Russell B, Schmierer D, and Laverty R (1997) The effects of aminorex and
related compounds on brain monoamines and metabolites in CBA mice. J Pharm
Pharmacol 49:8996.
Zhou C, Douglas JE, Kumar NN, Shu S, Bayliss DA, and Chen X (2013) Forebrain HCN1
channels contribute to hypnotic actions of ketamine. Anesthesiology 118:785795.
Zhou D, Lebel C, Lepage C, Rasmussen C, Evans A, Wyper K, Pei J, Andrew G,
Massey A, and Massey D, et al. (2011) Developmental cortical thinning in fetal
alcohol spectrum disorders. Neuroimage 58:1625.
Zhou FC, Anthony B, Dunn KW, Lindquist WB, Xu ZC, and Deng P (2007) Chronic
alcohol drinking alters neuronal dendritic spines in the brain reward center nucleus accumbens. Brain Res 1134:148161.
Zhou Q, Verdoorn TA, and Lovinger DM (1998) Alcohols potentiate the function of
5-HT3 receptor-channels on NCB-20 neuroblastoma cells by favouring and stabilizing the open channel state. J Physiol 507:335352.
Zhou W, Dong L, Wang N, Shi JY, Yang JJ, Zuo ZY, and Zhou ZQ (2014) Akt mediates
GSK-3b phosphorylation in the rat prefrontal cortex during the process of ketamine
exerting rapid antidepressant actions. Neuroimmunomodulation 21:183188.
Zhou Y, Oudin MJ, Gajendra S, Sonego M, Falenta K, Williams G, Lalli G,
and Doherty P (2015) Regional effects of endocannabinoid, BDNF and FGF receptor signalling on neuroblast motility and guidance along the rostral migratory
stream. Mol Cell Neurosci 64:3243.
Zhu L and Ye JH (2005) The role of G proteins in the activity and ethanol modulation
of glycine-induced currents in rat neurons freshly isolated from the ventral tegmental area. Brain Res 1033:102108.
Zhu PJ and Lovinger DM (2005) Retrograde endocannabinoid signaling in a postsynaptic neuron/synaptic bouton preparation from basolateral amygdala. J Neurosci 25:61996207.
Zhu Y, Hsu MS, and Pintar JE (1998) Developmental expression of the mu, kappa,
and delta opioid receptor mRNAs in mouse. J Neurosci 18:25382549.
Zimmer A, Zimmer AM, Hohmann AG, Herkenham M, and Bonner TI (1999) Increased mortality, hypoactivity, and hypoalgesia in cannabinoid CB1 receptor
knockout mice. Proc Natl Acad Sci USA 96:57805785.
Zorumski CF, Mennerick S, and Izumi Y (2014) Acute and chronic effects of ethanol
on learning-related synaptic plasticity. Alcohol 48:117.
Zuo W, Chen L, Wang L, and Ye JH (2013) Cocaine facilitates glutamatergic transmission and activates lateral habenular neurons. Neuropharmacology 70:180189.
Zou X, Patterson TA, Divine RL, Sadovova N, Zhang X, Hanig JP, Paule MG, Slikker
W Jr, and Wang C (2009) Prolonged exposure to ketamine increases neurodegeneration in the developing monkey brain. Int J Dev Neurosci 27:727731.
Zuo Y, Kuryatov A, Lindstrom JM, Yeh JZ, and Narahashi T (2002) Alcohol modulation of neuronal nicotinic acetylcholine receptors is alpha subunit dependent.
Alcohol Clin Exp Res 26:779784.
Zweben JE, Cohen JB, Christian D, Galloway GP, Salinardi M, Parent D, and Iguchi M
(2004) Psychiatric symptoms in methamphetamine users. Am J Addict 13:181190.

Potrebbero piacerti anche