Hostname: page-component-8448b6f56d-t5pn6 Total loading time: 0 Render date: 2024-04-16T04:35:40.525Z Has data issue: false hasContentIssue false

Cannabinoid signalling in embryonic and adult neurogenesis: possible implications for psychiatric and neurological disorders

Published online by Cambridge University Press:  16 May 2018

Rúbia W. de Oliveira*
Affiliation:
Department of Pharmacology and Therapeutics, State University of Maringá, Maringá, Paraná, Brazil
Cilene L. Oliveira
Affiliation:
Department of Physiological Sciences, Federal University of Santa Catarina, Florianópolis, Santa Catarina, Brazil
Francisco S. Guimarães
Affiliation:
Department of Pharmacology, School of Medicine of Ribeirão Preto, University of São Paulo, Ribeirão Preto, Brazil Center of Interdisciplinary Research on Applied Neurosciences (NAPNA), University of São Paulo, Ribeirão Preto, Brazil
Alline C. Campos
Affiliation:
Department of Pharmacology, School of Medicine of Ribeirão Preto, University of São Paulo, Ribeirão Preto, Brazil Center of Interdisciplinary Research on Applied Neurosciences (NAPNA), University of São Paulo, Ribeirão Preto, Brazil
*
Author for correspondence: Dr. Rúbia W. de Oliveira, Laboratory of Neuropsychopharmacology, Department of Pharmacology and Therapeutics, State University of Maringá, Av. Colombo 5790, K-68, 104a, Maringá, Paraná 87020-900, Brazil. Tel: +55 44 30115165; Fax: +55 44 30114999; E-mail: rmmwoliveira@uem.br.
Rights & Permissions [Opens in a new window]

Abstract

Cannabinoid signalling modulates several aspects of brain function, including the generation and survival of neurons during embryonic and adult periods. The present review intended to summarise evidence supporting a role for the endocannabinoid system on the control of neurogenesis and neurogenesis-dependent functions. Studies reporting participation of cannabinoids on the regulation of any step of neurogenesis and the effects of cannabinoid compounds on animal models possessing neurogenesis-dependent features were selected from Medline. Qualitative evaluation of the selected studies indicated that activation of cannabinoid receptors may change neurogenesis in embryonic or adult nervous systems alongside rescue of phenotypes in animal models of different psychiatric and neurological disorders. The text offers an overview on the effects of cannabinoids on central nervous system development and the possible links with psychiatric and neurological disorders such as anxiety, depression, schizophrenia, brain ischaemia/stroke and Alzheimer’s disease. An understanding of the mechanisms by which cannabinoid signalling influences developmental and adult neurogenesis will help foster the development of new therapeutic strategies for neurodevelopmental, psychiatric and neurological disorders.

Type
Review Article
Copyright
© Scandinavian College of Neuropsychopharmacology 2018 

Summations

  • Cannabinoid signalling modulates several aspects of brain function, including generation and survival of neurons during embryonic and adult periods.

  • Psychiatric and neurological disorders alter the dynamics of adult hippocampal neurogenesis by either increasing or decreasing neurogenesis.

  • Manipulations of cannabinoid signalling may restore or prevent neurogenic deficits in animal models that mimic some features of psychiatric and neurological conditions.

Considerations

  • Due to methodological limitations in the field of psychiatric and neurological disorders, mechanisms linking cannabinoids, neurogenesis and pathophysiology are still unclear.

  • This review detected the need for studies comparing the effects of acute and long-term treatment with cannabinoid on neurogenesis and associated functions during different life stages (mainly the critical periods of neuroplasticity).

  • This review detected the need for further work to establish the effects of cannabinoids on dysfunctional neurogenesis in animal models and human studies.

  • In future studies, a systematic review of the literature should be performed to increase the value of the evidence.

Introduction

A substantial body of evidence has demonstrated the involvement of cannabinoid signalling in regulating neurogenesis in embryonic or adult central nervous system (CNS) in physiological and/or pathological conditions. This is a narrative review intended to summarise the evidence supporting a role for the endocannabinoid system (ECBS) on the control of neurogenesis and neurogenesis-dependent functions. We selected studies reporting the participation of cannabinoids on the regulation of any step of the neurogenic process and showing effects of cannabinoid compounds on animal models of psychiatric and neurological disorders with neurogenesis-dependent features. From the selected literature, we extracted information regarding how cannabinoid compounds and manipulations of the ECBS affected the above-mentioned processes. We also advocated that the influence of cannabinoids on CNS development may be an opportunity to understand psychiatric and neurological disorders.

Neurogenesis in embryonic and adult CNS

Neurogenesis is the process by which functional neurons are produced in the nervous systems of all animals (Reference Gage, Kempermann, Palmer, Peterson and Ray1,Reference Lindsey and Tropepe2). In mammals, including humans, neurons in the peripheral nervous system and CNS are primarily generated during the embryonic and early postnatal periods (Reference Czaja, Fornaro and Geuna3). From early to adult life, neurogenesis remains active only in few discrete regions of the brain (Reference Altman and Das4,Reference Jordan, Ma, Ming and Song5). Although the functions of neurogenesis in the adult mammalian brain are controversial, its existence seems undisputed (Reference Ming and Song6).

Newborn neurons have been found in adult rats, mice, non-human primates and humans (Reference Lindsey and Tropepe2,Reference Altman and Das4Reference Kornack and Rakic10). The magnitude of the renewing of the neuronal population exhibits variations when compared across species and age of the subjects (Reference Spalding, Bergmann, Alkass, Bernard, Salehpour, Huttner, Bostrom, Westerlund, Vial, Buchholz, Possnert, Mash, Druid and Frisen11Reference Ihunwo, Tembo and Dzamalala13). For example, it has been reported that 0.004% of the dentate gyrus (DG) neurons are added daily in each human hippocampus, while in 2-month-old mice is 0.3–0.6% and for 5–16-year-old macaque is 0.04% per day (Reference Bergami, Masserdotti, Temprana, Motori, Eriksson, Gobel, Yang, Conzelmann, Schinder, Gotz and Berninger14). However, stereological methods have shown that the neuronal turnover in adult human brains is reduced as compared to mice and macaques, with an age-dependent decline of neuroblasts (Reference Eriksson, Perfilieva, Bjork-Eriksson, Alborn, Nordborg, Peterson and Gage9,Reference Spalding, Bergmann, Alkass, Bernard, Salehpour, Huttner, Bostrom, Westerlund, Vial, Buchholz, Possnert, Mash, Druid and Frisen11,Reference Bergami, Masserdotti, Temprana, Motori, Eriksson, Gobel, Yang, Conzelmann, Schinder, Gotz and Berninger14).

In adult or embryonic stages, neurogenesis process encompasses steps organised in time and space shaping the mammalian nervous system (Reference Urban and Guillemot15). The adequate balance between cell birth, survival, death and integration into the circuitries is fundamental for keeping the regular shape of the CNS and, consequently, for keeping its function (Reference Jessberger, Toni, Clemenson, Ray and Gage16Reference Andersen, Urban, Achimastou, Ito, Simic, Ullom, Martynoga, Lebel, Goritz, Frisen, Nakafuku and Guillemot19). For a detailed description of neurogenic processes, we suggest the reading of Paridaen and Huettner (Reference Paridaen and Huttner20) for embryonic neurogenesis and Bond et al. (Reference Bond, Ming and Song21) for adult neurogenesis. For the purposes of the present review, only selected aspects of neurogenesis will be described in the following text.

Newborn cells in the embryonic or adult CNS come from series of divisions of the neural stem cells (NSC). Originated from embryonic totipotent cells, NSC may proliferate or differentiate into new lineages by giving rise to progenitors committed to glial or neuronal phenotypes (Reference Guan, Chang, Rolletschek and Wobus22) (Fig. 1). The NSC, as well as the progenitors, may undergo symmetrical divisions forming two cells identical to themselves (rapid proliferation) or asymmetrical divisions generating a clone of itself and a different cell type (slow proliferation, slow differentiation) or two different cell types (rapid differentiation) (Reference Gotz and Huttner23) (Fig. 1). Glial or neuronal progenitors may differentiate into glioblasts or neuroblasts, respectively (Reference Qian, Shen, Goderie, He, Capela, Davis and Temple24) (Fig. 1). Glioblasts may proliferate and mature in the place of their birth or migrate to other regions maturing far away from their origin (Reference Guan, Chang, Rolletschek and Wobus22) while neuroblasts often migrate, mature and integrate circuits far away from their progenitors (Reference Alvarez-Buylla25). The migration of neuroblasts to their final destinations may be dependent on the scaffold of radial glia (Reference Qian, Shen, Goderie, He, Capela, Davis and Temple24,Reference Cayre, Canoll and Goldman26), or ‘tunnels’ of astrocytes (Reference Lois and Alvarez-Buylla27) or chains of neuroblasts (Reference Lindsey and Tropepe2,Reference Cameron, Woolley, McEwen and Gould28) (Fig. 1).

Fig. 1 Schematic representation of the steps in embryonic or adult neurogenesis in the central nervous system. Neural stem cells, neuronal progenitors and glial progenitors may undergo symmetric or asymmetric divisions. Symmetrical divisions produce two ‘daughters’ that are identical to their precursors and each other. Asymmetrical divisions produce two different ‘daughters’, one that is identical to their precursors and another ‘daughter’ that is different from the ‘sister’ and the precursor. Symmetrical divisions expand the pool of precursors (proliferation step) more rapidly than the asymmetrical divisions. However, asymmetrical divisions give rise to cells with a new phenotype (differentiation step). Therefore, neural stem cells may differentiate into progenitors committed to neuronal or glial phenotypes. Neuronal progenitors may differentiate into neuroblasts, whereas glial progenitors may differentiate into different types of glioblasts. Progenitors also may become quiescent( non-dividing state). Neuroblasts and glioblasts maintain their self-renewing capacity until maturation. Cell death may occur at any step of the process. For a review and more detailed description of neurogenic steps, we suggest the studies by Paridaen and Huettner (Reference Paridaen and Huttner20) (for embryonic neurogenesis) and Bond et al. (Reference Bond, Ming and Song21) (for adult neurogenesis).

In embryonic CNS, neuronal progenitors are localised mainly in the subventricular zone (SVZ) of all ventricles and, strictly controlled, neurogenesis occurs widespread in the nervous system (Reference Qian, Shen, Goderie, He, Capela, Davis and Temple24,Reference Cipriani, Nardelli, Verney, Delezoide, Guimiot, Gressens and Adle-Biassette29). Under physiological conditions, adult neurogenesis seems confined to the SVZ-olfactory bulb system (SVZ-OB) and the DG of the hippocampus. In the SVZ-OB, neuronal progenitors are found throughout the longitudinal extension of the lateral walls of the lateral ventricles differentiating into neuroblasts while moving away of the SVZ through the rostral migratory stream (RMS) (Reference Pencea, Bingaman, Freedman and Luskin30). The RMS is like a tunnel, pavement with astrocytes, where chains of neural progenitors and neuroblasts (in different stages of development) migrate towards the OB (Reference Peretto, Merighi, Fasolo and Bonfanti31,Reference Sawamoto, Wichterle, Gonzalez-Perez, Cholfin, Yamada, Spassky, Murcia, Garcia-Verdugo, Marin, Rubenstein, Tessier-Lavigne, Okano and Alvarez-Buylla32). In the adult hippocampus, the neural progenitors are in the subgranular layer of the DG from where they migrate in chains while differentiating into neuroblasts, towards the granular layer of the DG (Reference Lindsey and Tropepe2,Reference Cameron, Woolley, McEwen and Gould28). In their final destinations, the neuroblasts will find their fate by settling, maturing, integrating the existing circuitry or dying (Reference Gage, Kempermann, Palmer, Peterson and Ray1,Reference Cameron, Woolley, McEwen and Gould28,Reference Alvarez-Buylla and Garcia-Verdugo33).

A plethora of regulatory mechanisms orchestrates neurogenesis in embryos and adults (Reference Kintner34). For example, paracrine factors, neurotransmitters or hormones may favour or disrupt proliferation, differentiation, migration or maturity by interacting with receptors in the progenitors or other cells in different levels of differentiation and commitment (Reference Jagasia, Song, Gage and Lie35,Reference Pathania, Yan and Bordey36). In addition, diffusible and membrane-bound factors from target regions may repel or attract neuroblasts, slowing down or speeding up their maturation and integration in the circuitry at the final destination (Reference Hagg37). The presence of synthetic and degradation enzymes for the endocannabinoids as well as cannabinoid receptors in NSC and progenitor cells suggests that ECBS may play a role in the control of neurogenesis in embryos and adults (Reference Katona, Urban, Wallace, Ledent, Jung, Piomelli, Mackie and Freund38,Reference Harkany, Keimpema, Barabas and Mulder39).

Cannabinoids and the ECBS

For decades, the term cannabinoids have described a class of compounds derived from the plant Cannabis spp. Currently, the term is essentially used to describe three types of substances: phytocannabinoids, synthetic cannabinoids and endocannabinoids (Reference Pertwee40). More than 100 phytocannabinoids have been identified and isolated from the Cannabis sativa, including its two major components: Δ9-tetrahydrocannabinol (THC), responsible for the psychological and subjective effects of the plant, and cannabidiol, the main non-psychotomimetic compound (Reference Mechoulam and Gaoni41,Reference Turner, Williams, Iversen and Whalley42). Search for endogenous sites, explaining the effects of THC on behaviour, led to the discovery of the ECBS. In the late 1980s, Devane et al. (Reference Devane, Dysarz, Johnson, Melvin and Howlett43) identified a specific protein G-coupled receptor activated by THC in the rat CNS, which was later cloned and named CB1 receptor (Reference Matsuda, Lolait, Brownstein, Young and Bonner44). Afterwards, a second cannabinoid receptor was also described and named CB2 (Reference Munro, Thomas and Abu-Shaar45). CB1 and CB2 receptors are Gi/o-coupled protein receptors blocking calcium channels and activating potassium channels reducing cell firing rate and neurotransmitter release (Reference Szabo and Schlicker46) (Fig. 2).

Fig. 2 Classical representation of endocannabinoid signalling in the adult brain. Anandamide (AEA) and 2-arachidonoyl glycerol (2-AG) are produced ‘on demand’ in calcium (Ca2+)-dependent manner (via the previous activation of a metabotropic or ionotropic receptor). After the synthesis of endocannabinoids by specialised enzymes, they act as retrograde massagers by activating CB1 receptors located at pre-synaptic terminals. CB1 is a Gi/o-coupled receptor, and its activation reduces Ca2+ currents and increases K+ currents, leading to the inhibition of neurotransmitter release. The actions of 2-AG and AEA are terminated by enzymatic hydrolysis; fatty acid amide hydrolase (FAAH) and monoacylglycerol lipase (MAGL) degrade AEA and 2-AG, respectively. The CB1 receptor is also expressed in astrocytes and microglia and the CB2 receptor is expressed in activated microglia and putatively expressed in neurons (still under debate). CB1, type 1 cannabinoid receptor; CB2, type 2 cannabinoid receptor; DAGL, diacylglycerol lipase; NAPE-PLD, n-acyl phosphatidylethanolamine-specific phospholipase D.

The initial characterisation of CB1 receptors in the CNS indicated that these receptors are expressed in axons, cell bodies and dendrites (Reference Tsou, Mackie, Sanudo-Pena and Walker47). In 2001, Wilson and Nicoll (Reference Wilson and Nicoll48) found CB1 receptors located in the axon terminals participating in the endocannabinoid mediated-retrograde signalling in the hippocampus controlling the release of gamma-aminobutyric acid (GABA). Following the initial finding, activation of the CB1 receptor was shown to inhibit the release of other neurotransmitters, such as glutamate, serotonin and dopamine (Reference Takahashi and Castillo49,Reference Lau and Schloss50). In adult brains, CB1 activation was also associated with the control of short-term neuronal reactivity in glutamatergic and peptidergic synapses (Reference Wilson and Nicoll48,Reference Yoshida, Hashimoto, Zimmer, Maejima, Araishi and Kano51,Reference Diana and Marty52). CB1 activation also exerts neuroprotective effects by reducing glutamate-induced excitotoxicity (Reference Marsicano, Goodenough, Monory, Hermann, Eder, Cannich, Azad, Cascio, Gutierrez, van der Stelt, Lopez-Rodriguez, Casanova, Schutz, Zieglgansberger, Di Marzo, Behl and Lutz53) and stimulating neuroplasticity (Reference Fogaca, Galve-Roperh, Guimaraes and Campos54). Expression of functional CB2 receptors has been found in specific populations of cells (microglial cells, neurons and NSCs) in the CNS, but at lower levels than CB1 (Reference Onaivi, Ishiguro, Gong, Patel, Perchuk, Meozzi, Myers, Mora, Tagliaferro, Gardner, Brusco, Akinshola, Liu, Hope, Iwasaki, Arinami, Teasenfitz and Uhl55Reference Lisboa, Gomes, Guimaraes and Campos57). The specific functions and cellular consequences of CB2 activation in the CNS are still under investigation but seem also related to the control of the release of neurotransmitters. For example, the CB2 receptor agonist JWH133 decreased the amount of dopamine in the nucleus accumbens of rodents submitted to a cocaine-induced self-administration paradigm (Reference Xi, Peng, Li, Song, Zhang, Liu, Yang, Bi, Li and Gardner58). In microglial cells, activation of CB2 receptors reduced the secretion of cytokines that function as neuromodulators changing neuronal firing and subsequently neurotransmitter release (Reference Lisboa, Gomes, Guimaraes and Campos57).

The first endogenous ligands for CB receptors were the arachidonoyl ethanolamide or anandamide (AEA) and 2-arachidonoyl glycerol (2-AG), derived from the hydrolysis of arachidonic acid (Reference Devane, Hanus, Breuer, Pertwee, Stevenson, Griffin, Gibson, Mandelbaum, Etinger and Mechoulam59,Reference Mechoulam, Ben-Shabat, Hanus, Ligumsky, Kaminski, Schatz, Gopher, Almog, Martin, Compton, Pertwee, Griffin, Bayewitch, Barg and Vogel60). In the CNS, AEA is synthesised mainly by n-acyl phosphatidylethanolamine phospholipase D, whereas 2-AG is produced by the α and β isoforms of diacylglycerol lipase (DAGL). Once produced and released, in a calcium-dependent manner (Reference Saito, Wotjak and Moreira61), AEA and 2-AG may interact with CB receptors located in pre- and post-synaptic membranes or may be hydrolysed by the enzymes fatty acid amide hydrolase (FAAH) and monoacylglycerol lipase (MAGL), respectively (Reference Katona, Urban, Wallace, Ledent, Jung, Piomelli, Mackie and Freund38) (Fig. 2). Endocannabinoids production and release from postsynaptic neuronal compartments occur ‘on demand’, upon cell depolarisation, being reduced by their own action as retrograde messengers on pre-synaptic inhibitory CB1 receptors (Reference Saito, Wotjak and Moreira61) (Fig. 2). Embryonic and adult regions of the brain with neurogenic potential express genes coding for receptors and enzymes of the ECBS system, which may interfere with pre-existing or newly formed networks (Reference Katona, Urban, Wallace, Ledent, Jung, Piomelli, Mackie and Freund38,Reference Campos, Paraíso-Luna, Fogaça, Guimarães and Galve-Roperh62).

Cannabinoids and embryonic neurogenesis

The ECBS seems capable of regulating some features of the neurogenic process in the embryonic hippocampus and cerebral cortex (Reference Palazuelos, Aguado, Egia, Mechoulam, Guzman and Galve-Roperh56,Reference Aguado, Monory, Palazuelos, Stella, Cravatt, Lutz, Marsicano, Kokaia, Guzman and Galve-Roperh63Reference Diaz-Alonso, Aguado, de Salas-Quiroga, Ortega, Guzman and Galve-Roperh66). The increase of the intracellular calcium in embryonic NSC and immature neurons induced the production of endocannabinoids (Reference Maccarrone, Guzman, Mackie, Doherty and Harkany67). Growth factors, such as fibroblast growth factor and nerve growth factor, may increase 2-AG levels via the activation of phospholipase C or tropomyosin receptor kinase A receptor (Reference Keimpema, Tortoriello, Alpar, Capsoni, Arisi, Calvigioni, Hu, Cattaneo, Doherty, Mackie and Harkany68,Reference Maison, Walker, Walsh, Williams and Doherty69). 2-AG, synthetised approximately 1000-fold higher than AEA in embryonic brain, seem to favour neural maturation and cell proliferation (Reference Keimpema, Alpar, Howell, Malenczyk, Hobbs, Hurd, Watanabe, Sakimura, Kano, Doherty and Harkany70Reference Oudin, Hobbs and Doherty72). Indeed, the pharmacological inhibition of DAGL, responsible for the 2-AG synthesis, with RHC-80276 reduced the proliferation of embryonic NSC in cultures (Reference Goncalves, Suetterlin, Yip, Molina-Holgado, Walker, Oudin, Zentar, Pollard, Yanez-Munoz, Williams, Walsh, Pangalos and Doherty73). Besides, an isoform of the enzyme DAGL co-localises with CB1 receptors in developing neurons during the growth of the axonal cones (Reference Oudin, Hobbs and Doherty72). A role for AEA is unclear once the inhibition of enzymes for synthesis (Reference Campos, Ortega, Palazuelos, Fogaca, Aguiar, Diaz-Alonso, Ortega-Gutierrez, Vazquez-Villa, Moreira, Guzman, Galve-Roperh and Guimaraes74) or degradation- (Reference Aguado, Monory, Palazuelos, Stella, Cravatt, Lutz, Marsicano, Kokaia, Guzman and Galve-Roperh63) induced proliferation of embryonic NSC.

Actions of the endocannabinoids on neural development seem to mediate by CB1 and CB2 receptors, which expressions may vary over the course of neurogenesis (Fig. 3). Indeed, the receptor CB2 is more abundant in less committed cells, whereas CB1 receptor is predominantly expressed during neuronal lineage specification (Reference Harkany, Guzman, Galve-Roperh, Berghuis, Devi and Mackie71,Reference Mato, Del Olmo and Pazos75) (Fig. 3). In addition, cannabinoid receptors seem functional during the development of the CNS once that cannabinoid receptor agonist WIN 55,212-2 stimulated the binding of [35S] GTPγS in the tissue of embryonic brain (Reference Berrendero, Mendizabal, Murtra, Kieffer and Maldonado76). In the embryonic cortex, genetic ablation of the CB1 receptor inhibited proliferation of NSC, favoured neuronal fate commitment and neurite growth (Reference Keimpema, Alpar, Howell, Malenczyk, Hobbs, Hurd, Watanabe, Sakimura, Kano, Doherty and Harkany70). Activation of CB1 in cortical neural precursors with the agonist HU-210 promoted the expansion of NSC pool and promoted survival by inducing Pax6 and T-box TF (Tbr2) (Reference Aguado, Palazuelos, Monory, Stella, Cravatt, Lutz, Marsicano, Kokaia, Guzman and Galve-Roperh64). Pax6 is an important transcription factor involved in regulating cortical progenitor proliferation, neurogenesis and the formation of cortical layers, whereas Trb2 promotes the generation and proliferation of intermediate precursors that give rise to pyramidal glutamatergic neurons in the cortex during neurodevelopment (Reference Urban and Guillemot15). Activation of cannabinoid receptors by AEA, 2-AG or WIN55-212,2 may also promote astroglial cell differentiation in vitro (Reference Aguado, Palazuelos, Monory, Stella, Cravatt, Lutz, Marsicano, Kokaia, Guzman and Galve-Roperh64). Despite their viability, fertility and normal brain morphology (Reference Marsicano, Goodenough, Monory, Hermann, Eder, Cannich, Azad, Cascio, Gutierrez, van der Stelt, Lopez-Rodriguez, Casanova, Schutz, Zieglgansberger, Di Marzo, Behl and Lutz53), CB1 knockout mice presented higher mortality, reduced locomotor activity and hypoalgesia when compared with heterozygous littermates (Reference Zimmer, Zimmer, Hohmann, Herkenham and Bonner77).

Fig. 3 Schematic representation of the neurogenesis steps in the central nervous system of embryos (a) and adults (b), along with the putative expression of the endocannabinoid system in different cell populations. 2-AG, 2-arachidonoylglycerol; AEA, anandamide; CB1, type 1 cannabinoid receptor; CB2, type 2 cannabinoid receptor; DAGL, diacylglycerol lipase; FAAH, fatty acid amide hydrolase; MAGL, monoacylglycerol lipase; NAPE-PLD, n-acyl phosphatidylethanolamine-specific phospholipase D.

In humans, the ectopic expression of CB1 and CB2 receptors is associated with defective development of the cortex (Reference Zurolo, Iyer, Spliet, Van Rijen, Troost, Gorter and Aronica78). Endocannabinoid signalling controls the proliferation of pyramidal cell progenitors and the radial migration of immature pyramidal cells in the embryonic cortex (Reference Mulder, Aguado, Keimpema, Barabas, Ballester Rosado, Nguyen, Monory, Marsicano, Di Marzo, Hurd, Guillemot, Mackie, Lutz, Guzman, Lu, Galve-Roperh and Harkany79). The CB1 receptor is expressed in intermediate progenitor cells (Tbr2+) that later differentiate into pyramidal cells (Reference Diaz-Alonso, Aguado, de Salas-Quiroga, Ortega, Guzman and Galve-Roperh66,Reference Mulder, Aguado, Keimpema, Barabas, Ballester Rosado, Nguyen, Monory, Marsicano, Di Marzo, Hurd, Guillemot, Mackie, Lutz, Guzman, Lu, Galve-Roperh and Harkany79,Reference Bisogno, Howell, Williams, Minassi, Cascio, Ligresti, Matias, Schiano-Moriello, Paul, Williams, Gangadharan, Hobbs, Di Marzo and Doherty80). Zurolo et al. (Reference Zurolo, Iyer, Spliet, Van Rijen, Troost, Gorter and Aronica78) observed unexpectedly high expression of CB1 receptors in dysplastic neurons in the early stages of human corticogenesis associated with cortical malformations and intractable epilepsy (focal cortical dysplasia). According to Diaz-Alonso et al. (Reference Diaz-Alonso, Aguado, de Salas-Quiroga, Ortega, Guzman and Galve-Roperh66), the CB1 receptor is also involved in organising the cortical layers. In mice lacking CB1 expression in glutamatergic neurons during cortical development, the expression of the proteins (Ctip2/Satb2) was abnormal and the cortical layer V disorganised producing severe motor deficits in adult animals (Reference Keimpema, Tortoriello, Alpar, Capsoni, Arisi, Calvigioni, Hu, Cattaneo, Doherty, Mackie and Harkany68,Reference Maison, Walker, Walsh, Williams and Doherty69). Moreover, Alpar et al. (Reference Alpar, Tortoriello, Calvigioni, Niphakis, Milenkovic, Bakker, Cameron, Hanics, Morris, Fuzik, Kovacs, Cravatt, Parnavelas, Andrews, Hurd, Keimpema and Harkany81) observed enlarged corpus callosum by excessive 2-AG-mediated signalling suggesting abnormal axonal growth of glutamatergic neurons of layer V caused by CB1 hyperactivity. CB1 signalling seems also important to correct placement and integration of GABAergic interneurons during cortical development (Reference Berghuis, Dobszay, Wang, Spano, Ledda, Sousa, Schulte, Ernfors, Mackie, Paratcha, Hurd and Harkany82,Reference Roland, Ricobaraza, Carrel, Jordan, Rico, Simon, Humbert-Claude, Ferrier, McFadden, Scheuring and Lenkei83). In fact, Morozov et al. (Reference Morozov, Torii and Rakic84) observed CB1 receptors expressed in embryonic GABAergic interneurons migrating through a long-distance pathway to differentiate into CB1/CCK+ or CB1/reelin/calretinin+ GABAergic interneurons. In these cells, CB1 activation by endogenous or synthetic cannabinoids regulates axonal growth and the shape of their dendritic arbours (Reference Goncalves, Suetterlin, Yip, Molina-Holgado, Walker, Oudin, Zentar, Pollard, Yanez-Munoz, Williams, Walsh, Pangalos and Doherty73,Reference Berghuis, Dobszay, Wang, Spano, Ledda, Sousa, Schulte, Ernfors, Mackie, Paratcha, Hurd and Harkany82,Reference Roland, Ricobaraza, Carrel, Jordan, Rico, Simon, Humbert-Claude, Ferrier, McFadden, Scheuring and Lenkei83). 2-AG-mediated may also control the differentiation of NSCs into GABAergic neurons and neurite outgrowth in cholinergic neurons (Reference Keimpema, Tortoriello, Alpar, Capsoni, Arisi, Calvigioni, Hu, Cattaneo, Doherty, Mackie and Harkany68) while AEA induced the formation of CB1/TrkB heterocomplexes, promoting interneuron migration (Reference Berghuis, Dobszay, Wang, Spano, Ledda, Sousa, Schulte, Ernfors, Mackie, Paratcha, Hurd and Harkany82). Roles for CB2 receptors during the different stages of brain development remain unclear: the antagonist SR144528 decreased the basal proliferative capacity of NSCs in vitro (Reference Palazuelos, Ortega, Diaz-Alonso, Guzman and Galve-Roperh85); agonist HU-308 induced cell cycle maintenance and neural differentiation (Reference Molina-Holgado, Rubio-Araiz, Garcia-Ovejero, Williams, Moore, Arevalo-Martin, Gomez-Torres and Molina-Holgado86); 2-AG was shown to induce early oligodendrocyte differentiation via CB2 receptors (Reference Alpar, Tortoriello, Calvigioni, Niphakis, Milenkovic, Bakker, Cameron, Hanics, Morris, Fuzik, Kovacs, Cravatt, Parnavelas, Andrews, Hurd, Keimpema and Harkany81).

Cannabinoids and adult neurogenesis

In the adult brain, the ECBS modulates different steps required for neurogenesis: cell proliferation, differentiation, maturation and survival (Fig. 3) (Reference Prenderville, Kelly and Downer87). Cannabinoid receptors activate different intracellular pathways, such as extracellular signal-regulated kinases (ERKs) 1 and 2 (ERK1/2), c-Jun amino-terminal kinases and PI3K/Akt/mTOR, inducing the production of neurotrophins such as brain-derived neurotrophic factor (BDNF) and other molecules that control the proliferation and survival of newborn cells (Reference Harkany, Keimpema, Barabas and Mulder39). Voluntary exercise, a positive regulator of adult neurogenesis, increases AEA levels and promotes cell proliferation in the hippocampus (Reference Hill, Titterness, Morrish, Carrier, Lee, Gil-Mohapel, Gorzalka, Hillard and Christie88). Pre-treatment with the CB1 receptor antagonist AM251 prevented running-induced adult hippocampal neurogenesis (Reference Hill, Titterness, Morrish, Carrier, Lee, Gil-Mohapel, Gorzalka, Hillard and Christie88) Facilitation of the effects of AEA by pharmacological (URB597) or genetic FAAH inhibition increased hippocampal neurogenesis (Reference Diaz-Alonso, Aguado, de Salas-Quiroga, Ortega, Guzman and Galve-Roperh66) and prevented its decrease after trimethylthiazoline exposure (Reference Aguado, Monory, Palazuelos, Stella, Cravatt, Lutz, Marsicano, Kokaia, Guzman and Galve-Roperh63). Conversely, Gonçalves et al. (Reference Goncalves, Suetterlin, Yip, Molina-Holgado, Walker, Oudin, Zentar, Pollard, Yanez-Munoz, Williams, Walsh, Pangalos and Doherty73) observed suppressed proliferation in the SVZ and cell migration SVZ-OB after treatment with RHC33, an inhibitor of 2-AG synthesis. In addition, genetic ablation of DAGLα/β decreases cell proliferation, survival and the number of cells committed to the neuronal fate in the DG (Reference Gao, Vasilyev, Goncalves, Howell, Hobbs, Reisenberg, Shen, Zhang, Strassle, Lu, Mark, Piesla, Deng, Kouranova, Ring, Whiteside, Bates, Walsh, Williams, Pangalos, Samad and Doherty89,Reference Jenniches, Ternes, Albayram, Otte, Bach, Bindila, Michel, Lutz, Bilkei-Gorzo and Zimmer90).

Phytocannabinoids such as THC and cannabidiol might increase or decrease adult hippocampal neurogenesis (Reference Berghuis, Dobszay, Wang, Spano, Ledda, Sousa, Schulte, Ernfors, Mackie, Paratcha, Hurd and Harkany82,Reference Wolf, Bick-Sander, Fabel, Leal-Galicia, Tauber, Ramirez-Rodriguez, Muller, Melnik, Waltinger, Ullrich and Kempermann91,Reference Campos, Fogaca, Scarante, Joca, Sales, Gomes, Sonego, Rodrigues, Galve-Roperh and Guimaraes92). However, acute or chronic (3 weeks) treatment with THC did not change cell proliferation in the DG of adult animals (Reference Campos, Fogaca, Scarante, Joca, Sales, Gomes, Sonego, Rodrigues, Galve-Roperh and Guimaraes92). In the study by Wolf et al. (Reference Wolf, Bick-Sander, Fabel, Leal-Galicia, Tauber, Ramirez-Rodriguez, Muller, Melnik, Waltinger, Ullrich and Kempermann91), adult mice treated with THC (6 weeks) exhibited decreased proliferation and a simultaneous impairment in spatial memory performance.

Adult CB1 knockout mice showed lower rates of proliferation, astrogliogenesis and neurogenesis in the subgranular zone (SGZ) and SVZ (Reference Aguado, Monory, Palazuelos, Stella, Cravatt, Lutz, Marsicano, Kokaia, Guzman and Galve-Roperh63,Reference Campos, Fogaca, Scarante, Joca, Sales, Gomes, Sonego, Rodrigues, Galve-Roperh and Guimaraes92,Reference Jin, Xie, Kim, Parmentier-Batteur, Sun, Mao, Childs and Greenberg93) and kainic acid-induced hippocampal NSC proliferation (Reference Aguado, Monory, Palazuelos, Stella, Cravatt, Lutz, Marsicano, Kokaia, Guzman and Galve-Roperh63). However, results obtained in studies using the treatment with CB1 antagonists or inverse agonists such as rimonabant, are contradictory. For example, rimonabant decreased doublecortin (DCX) expression in the SGZ of the DG and SVZ (Reference Lee, Kim, Yoon and Ryu94). In other studies, a CB1 receptor antagonist/inverse agonist facilitated the proliferation and survival of hippocampal neural precursor cells (Reference Jenniches, Ternes, Albayram, Otte, Bach, Bindila, Michel, Lutz, Bilkei-Gorzo and Zimmer90,Reference Campos, Fogaca, Scarante, Joca, Sales, Gomes, Sonego, Rodrigues, Galve-Roperh and Guimaraes92,Reference Lee, Kim, Yoon and Ryu94). Rodents treated with repeated doses of WIN 55,212-2, a CB1/CB2 receptor agonist, exhibit higher proliferation rates of neural precursor cells in the SVZ and DG (Reference Aguado, Monory, Palazuelos, Stella, Cravatt, Lutz, Marsicano, Kokaia, Guzman and Galve-Roperh63,Reference Goncalves, Suetterlin, Yip, Molina-Holgado, Walker, Oudin, Zentar, Pollard, Yanez-Munoz, Williams, Walsh, Pangalos and Doherty73). In adult CB2 knockout mice, low rates of cell proliferation under basal conditions or in response to kainate-induced excitotoxicity were also observed in the DG (Reference Berghuis, Dobszay, Wang, Spano, Ledda, Sousa, Schulte, Ernfors, Mackie, Paratcha, Hurd and Harkany82). CB2 inverse agonists, such as JTE 907, AM630 or SR144528, also reduced NSC proliferation in the SVZ and SGZ (Reference Goncalves, Suetterlin, Yip, Molina-Holgado, Walker, Oudin, Zentar, Pollard, Yanez-Munoz, Williams, Walsh, Pangalos and Doherty73,Reference Berghuis, Dobszay, Wang, Spano, Ledda, Sousa, Schulte, Ernfors, Mackie, Paratcha, Hurd and Harkany82). These compounds decrease the basal proliferative capacity of NSCs in culture (Reference Berghuis, Dobszay, Wang, Spano, Ledda, Sousa, Schulte, Ernfors, Mackie, Paratcha, Hurd and Harkany82). Repeated administration of a CB2 receptor agonist, HU-308, increases NSC proliferation in the SGZ via the Akt/mTORC1 pathway (Reference Berghuis, Dobszay, Wang, Spano, Ledda, Sousa, Schulte, Ernfors, Mackie, Paratcha, Hurd and Harkany82).

Despite some contradictions, most of the publications examined here indicated the activation of cannabinoid receptors as the main mechanisms by which ECBS may regulate neurogenesis in embryonic and adult mammalian brains. In the next sections, we will speculate on how cannabinoid receptors modulation may change neurogenesis repercuting in the pathophysiology of anxiety, depression, schizophrenia, brain ischaemia and Alzheimer’s disease.

Cannabinoids, neurogenesis and possible implications for psychiatric and neurological disorders

Mental and neurological disorders comprise a broad range of disabling syndromes with different emotional and behavioural symptoms. Aberrant neural development or disruptive mechanisms related to the adult neurogenic niches are potential aetiological factors that precipitate the initial symptoms or the late-onset of these disorders (Reference Han, Lee and Koh95). For example, changes in the mechanisms associated with the neurogenic process in the embryonic and adult brain have been reported in patients with Alzheimer’s disease (AD) (Reference Mu and Gage96,Reference Martinez-Canabal97), schizophrenia (Reference Iannitelli, Quartini, Tirassa and Bersani98) and mood disorders (Reference Campos, Moreira, Gomes, Del Bel and Guimaraes99). In the other way around, psychiatric and neurological disorders may alter the dynamics of adult hippocampal neurogenesis by either increasing or decreasing cell proliferation (Reference Martinez-Canabal97,Reference Duman, Malberg, Nakagawa and D’Sa100). Increased hippocampal cell proliferation has been observed in animal models of Huntington’s disease (Reference Vivar101), ischaemic brain injury (Reference Kawai, Takagi, Miyake-Takagi, Okuyama, Mochizuki and Takeo102) and temporal lobe epilepsy (Reference Parent103,Reference Liu, Curtis, Gibbons, Mee, Bergin, Teoh, Connor, Dragunow and Faull104). Impairments in hippocampal neurogenesis have been reported in animal models of AD (Reference Hollands, Bartolotti and Lazarov105), Parkinson’s disease (Reference Regensburger, Prots and Winner106) and in the postmortem brains of patients with different psychiatric conditions (Reference Kempermann and Kronenberg107). In addition to the loss of existing neurons, a decrease in neurogenesis in subjects with these conditions may indicate that the endogenous capacity of the adult brain for cell renewal and the putative functions of these neurons are compromised or even lost (Reference David, Wang, Samuels, Rainer, David, Gardier and Hen108).

Despite the extensive pre-clinical evidence suggesting that both exogenous and endogenous cannabinoids may regulate neurogenesis, which may be affected by mental and neurological disorders, the link between cannabinoids, neurogenesis and brain disorders are unclear. The weakness of evidence may come from the lack of postmortem studies in brains from patients with neuropsychiatric disorders (Reference David, Wang, Samuels, Rainer, David, Gardier and Hen108). In the next sections, we present evidence suggesting that manipulations of cannabinoid signalling restore or prevent neurogenic deficits in animal models that mimic some features of psychiatric and neurological conditions.

Cannabinoids, adult neurogenesis, and depressive and anxiety disorders

Impairments in hippocampus-dependent functions (e.g. cognitive deficits, affect lability and dysregulated pattern separation) are common symptoms of psychiatric disorders such as major depression, anxiety, schizophrenia and addiction (Reference David, Wang, Samuels, Rainer, David, Gardier and Hen108Reference Kang, Wen, Song, Christian and Ming110). These symptoms may indicate a disrupted function of the hippocampal DG and dysregulation of adult-generated neurons (Reference Yun, Reynolds, Masiulis and Eisch111). Indeed, decreases in hippocampal volume and hippocampal neurogenesis have been considered cellular substrates of major depression (Reference Duman, Malberg, Nakagawa and D’Sa100,Reference Kempermann and Kronenberg107,Reference Sheline112), posttraumatic stress disorder (Reference Kitayama, Vaccarino, Kutner, Weiss and Bremner113Reference Wang, Neylan, Mueller, Lenoci, Truran, Marmar, Weiner and Schuff115) and schizophrenia (Reference Goldman and Mitchell116). The attenuation of hippocampal neurogenesis also facilitates anxiety- and despair-related behaviours in rodents (Reference Hollands, Bartolotti and Lazarov105,Reference Revest, Dupret, Koehl, Funk-Reiter, Grosjean, Piazza and Abrous117). Moreover, adult hippocampal neurogenesis has been suggested to buffer the stress response (Reference Campos, Ortega, Palazuelos, Fogaca, Aguiar, Diaz-Alonso, Ortega-Gutierrez, Vazquez-Villa, Moreira, Guzman, Galve-Roperh and Guimaraes74,Reference Snyder, Soumier, Brewer, Pickel and Cameron118) and is implicated in the therapeutic effects of antidepressants (Reference Santarelli, Saxe, Gross, Surget, Battaglia, Dulawa, Weisstaub, Lee, Duman, Arancio, Belzung and Hen119,Reference Malberg120). Structural changes in the hippocampus are attenuated or reversed by antidepressants, atypical antipsychotics and physical exercise, which are known to positively impact hippocampal neurogenesis (Reference Kempermann, Fabel, Ehninger, Babu, Leal-Galicia, Garthe and Wolf121,Reference Erickson, Voss, Prakash, Basak, Szabo, Chaddock, Kim, Heo, Alves, White, Wojcicki, Mailey, Vieira, Martin, Pence, Woods, McAuley and Kramer122). Therefore, it is likely that some of the actions of cannabinoids might rescue behavioural and/or functional deficits impaired by adult neurogenesis deficiencies.

Despite the extensive pre-clinical evidence suggesting that both exogenous and endogenous cannabinoids regulate adult hippocampal neurogenesis, the mechanisms that link cannabinoids, alterations in adult neurogenesis and affective disorders are still unclear. This lack of clarity is at least partially because postmortem studies of adult hippocampal neurogenesis in brains from patients with neuropsychiatric disorders are rare, and the findings have been mostly inconclusive (Reference Christian, Song and Ming109). For example, a decrease (Reference Boldrini, Underwood, Hen, Rosoklija, Dwork, John Mann and Arango123) or lack of change (Reference Reif, Fritzen, Finger, Strobel, Lauer, Schmitt and Lesch124) in hippocampal cell proliferation has been observed in the hippocampus of patients with major depression. Moreover, depressed patients treated with tricyclic antidepressants or selective serotonin reuptake inhibitors showed increased (Reference Boldrini, Underwood, Hen, Rosoklija, Dwork, John Mann and Arango123) or unchanged (Reference Reif, Fritzen, Finger, Strobel, Lauer, Schmitt and Lesch124) hippocampal cell proliferation.

In rodents, chronic unpredictable stress (CUS) has been used to mimic some depressive-like behaviours and to investigate the underlying cellular and molecular mechanisms of depression (Reference Willner125). CUS not only induces depressive-like behaviours but also impairs long-term potentiation (LTP) and decreases the number of BrdU-labelled neural progenitor cells and DCX-positive immature neurons in the mouse DG (Reference Li, Chen, Liu, Zhang, Liu and Li126Reference Zhang, Wang, Zhong, Liu, Long, Zhao, Gao, Cravatt and Liu128). Otherwise, blockade of 2-AG degradation by the MAGL inhibitor JZL184 enhanced hippocampal neurogenesis, restored LTP in the DG, and produced antidepressant-like effects on mice that were subjected to the CUS model of depression (Reference Zhang, Wang, Zhong, Liu, Long, Zhao, Gao, Cravatt and Liu128) (Table 1). These effects were attributed to an increase in hippocampal neurogenesis that occurred through the activation of the CB1 receptor. However, so far these effects have not been confirmed by other groups. In other study, repeated cannabidiol administration (30 mg/kg for 14 days) exerted anxiolytic-like effects, reduced anhedonia and increased hippocampal neurogenesis in mice that were subjected to CUS (Reference Campos, Ortega, Palazuelos, Fogaca, Aguiar, Diaz-Alonso, Ortega-Gutierrez, Vazquez-Villa, Moreira, Guzman, Galve-Roperh and Guimaraes74). The genetic ablation of proliferating progenitors in the hippocampus of these stressed mice prevented the anxiolytic-like actions of cannabidiol. The authors concluded that repeated cannabidiol administration prevents the effects of CUS through a neurogenesis-dependent mechanism, favouring adaptations to stress. This assumption was supported by the observation that hippocampal adult neurogenesis was not required for the antidepressant-like effect of chronic cannabidiol administration under basal (non-stressed mice) conditions (Reference Schiavon, Bonato, Milani, Guimaraes and Weffort de Oliveira129).

Table 1 Cannabinoids increase adult neurogenesis in animal models of psychiatric conditions

↓, decreases; ↑, increases; DG, dentate gyrus; GFAP-TK, GFAP-thymidine kinase; i.p., intraperitoneal; LTP, long-term potentiation; mPFC, medial prefrontal cortex.

* All males.

Monoacylglycerol lipase inhibitor.

Cannabinoid agonist.

The behavioural and pro-neurogenic effects of cannabinoids on stressed mice involve the activation of both cannabinoid CB1 and CB2 receptors, secondary to an increase in endocannabinoid tone (Reference Campos, Ortega, Palazuelos, Fogaca, Aguiar, Diaz-Alonso, Ortega-Gutierrez, Vazquez-Villa, Moreira, Guzman, Galve-Roperh and Guimaraes74). Indeed, hippocampal neurogenesis is impaired in CB1 knockout mice (Reference Jin, Xie, Kim, Parmentier-Batteur, Sun, Mao, Childs and Greenberg93). Chronic administration of the full and potent CB1/CB2 receptor agonist HU-210 increased hippocampal cell proliferation and produced antidepressant-like effects on rat behaviours (Reference Hill and Gorzalka130). Accordingly, Lee et al. (Reference Lee, Kim, Yoon and Ryu94) have shown that repeated treatment with rimonabant, a CB1 receptor antagonist, caused loss of antidepressant activity and decreased DCX immunoreactivity in the mouse hippocampus. However, it is important to mention that these results have not been confirmed in other studies.

The CB2 receptor-selective agonist HU-308 also exerted proliferation-enhancing effects on the mouse hippocampus (Reference Palazuelos, Ortega, Diaz-Alonso, Guzman and Galve-Roperh85). Furthermore, transgenic mice that overexpress CB2 receptors and were subjected to CUS presented a decrease in depressive-like behaviours and increased expression of the BDNF gene in the hippocampus, suggesting an increase in neuroplasticity (Reference Garcia-Gutierrez, Perez-Ortiz, Gutierrez-Adan and Manzanares131).

Cannabinoids, neurogenesis and schizophrenia

Schizophrenia is a heterogeneous and multifactorial disease that is believed to result from complex interactions between genetic, physiological and environmental factors (Reference Alzheimer’s132). Based on the considerable evidence, schizophrenia may involve the abnormal neurogenesis of embryonic NSCs, a process that would be particularly vulnerable to a number of genetic and/or environmental disturbances during early brain development (Reference Iannitelli, Quartini, Tirassa and Bersani98,Reference Murray and Lewis133Reference Wu, Jew and Lu136). In humans, the use of Cannabis for recreational or medical reasons during pregnancy has been associated with attention deficits, impaired learning and memory, and behavioural changes related to schizophrenia in the offspring (Reference Wu, Jew and Lu136,Reference Richardson, Hester and McLemore137). However, the extent of this association is still controversial (Reference Fried138Reference El Marroun, Hudziak, Tiemeier, Creemers, Steegers, Jaddoe, Hofman, Verhulst, van den Brink and Huizink140). The effects of THC (Reference Tortoriello, Morris, Alpar, Fuzik, Shirran, Calvigioni, Keimpema, Botting, Reinecke, Herdegen, Courtney, Hurd and Harkany141) or synthetic cannabinoids (Reference Sun, Fang, Ren, Chen, Guo, Yan, Song, Zhang and Liao142) on embryonic development are highly variable, depending on the substance. In rodents, reports supporting and refuting the deleterious consequences of in utero and postnatal exposure to THC have been published (Reference Maccarrone, Guzman, Mackie, Doherty and Harkany67,Reference Alpar, Tortoriello, Calvigioni, Niphakis, Milenkovic, Bakker, Cameron, Hanics, Morris, Fuzik, Kovacs, Cravatt, Parnavelas, Andrews, Hurd, Keimpema and Harkany81,Reference de Salas-Quiroga, Diaz-Alonso, Garcia-Rincon, Remmers, Vega, Gomez-Canas, Lutz, Guzman and Galve-Roperh143). Due to the lack of conclusive data, the American Congress of Obstetricians and Gynaecologists (http://www.acog.org/) discourages the use of marijuana during pregnancy or lactation. Excellent reviews have been published on the topic of Cannabis use and neurodevelopment (Reference Maccarrone, Guzman, Mackie, Doherty and Harkany67,Reference Alpar, Tortoriello, Calvigioni, Niphakis, Milenkovic, Bakker, Cameron, Hanics, Morris, Fuzik, Kovacs, Cravatt, Parnavelas, Andrews, Hurd, Keimpema and Harkany81,Reference Richardson, Hester and McLemore137).

Regarding adult hippocampal neurogenesis, a previous study reported the higher expression of the polysialylated form of the neural cell adhesion molecule (PSA-NCAM), a marker of immature neurons, in the hippocampus of patients with schizophrenia in the absence of changes in total cell number (Reference Barbeau, Liang, Robitalille, Quirion and Srivastava144). Other studies reported a decrease in the number of cells positive for the proliferation marker Ki-67 in the hippocampus of patients with schizophrenia (Reference Reif, Fritzen, Finger, Strobel, Lauer, Schmitt and Lesch124,Reference Allen, Fung and Weickert145). Walton et al. (Reference Walton, Zhou, Kogan, Shin, Webster, Gross, Heusner, Chen, Miyake, Tajinda, Tamura, Miyakawa and Matsumoto146) identified an immature DG (iDG) in patients with schizophrenia. The iDG is characterised by greater hippocampal cell proliferation, an increase in the levels of markers of immature neurons (e.g. calretinin and DCX), and the lack of markers of mature neurons (e.g. calbindin). From a functional point of view, mice with an iDG exhibit several behavioural traits that reflect both positive and negative symptoms commonly observed in patients with schizophrenia, including hyperactivity and deficits in social interaction, nest building, and working memory (Reference Walton, Zhou, Kogan, Shin, Webster, Gross, Heusner, Chen, Miyake, Tajinda, Tamura, Miyakawa and Matsumoto146). Thus, disturbed hippocampal adult neurogenesis is related to cognitive deficits and other symptoms observed in patients with schizophrenia (Reference Reif, Fritzen, Finger, Strobel, Lauer, Schmitt and Lesch124). Susceptibility genes for schizophrenia, such as neuregulin-1, disrupted-in-schizophrenia 1 (DISC1), neuronal PAS domain-containing protein 3 (NPAS3) and fatty acid binding protein 7 (Fabp7), regulate adult hippocampal neurogenesis and are involved in the expression of schizophrenia-like behaviours in rodents (Reference Kang, Wen, Song, Christian and Ming110). For example, Fabp7-deficient mice show impaired hippocampal neurogenesis and a decrease in prepulse inhibition of the acoustic startle reflex (Reference Maekawa, Takashima, Matsumata, Ikegami, Kontani, Hara, Kawashima, Owada, Kiso, Yoshikawa, Inokuchi and Osumi147), indicating abnormalities in sensorimotor gating. SREB2, an orphan G-protein-coupled receptor expressed in the DG of patients with schizophrenia, impairs cognitive function and negatively regulates hippocampal adult neurogenesis in SREB2 Tg mice (Reference Chen, Kogan, Gross, Zhou, Walton, Shin, Heusner, Miyake, Tajinda, Tamura and Matsumoto148). Accordingly, DG-irradiated rats present behavioural abnormalities in social interactions and working memory, which are also often observed in patients with schizophrenia (Reference Iwata, Suzuki, Wakuda, Seki, Thanseem, Matsuzaki, Mamiya, Ueki, Mikawa, Sasaki, Suda, Yamamoto, Tsuchiya, Sugihara, Nakamura, Sato, Takei, Hashimoto and Mori149). Therefore, impaired adult hippocampal neurogenesis might contribute to hippocampal structural abnormalities and be associated with the behavioural and cognitive symptoms of schizophrenia (Reference Reif, Fritzen, Finger, Strobel, Lauer, Schmitt and Lesch124,Reference Nelson, Saykin, Flashman and Riordan150Reference Ganzola, Maziade and Duchesne153).

Although the effects of antipsychotic drugs on adult hippocampal neurogenesis and hippocampus-dependent behaviours are not entirely clear (Reference Newton and Duman154,Reference Balu and Lucki155), the neurogenic actions of atypical antipsychotics have been at least partially correlated with beneficial effects on negative and cognitive symptoms of schizophrenia. Haloperidol, a typical antipsychotic drug that controls positive symptoms of schizophrenia by opposing the excessive stimulation of D2 receptors, fails to alleviate negative symptoms, such as flattened affect and cognitive deficits (Reference Meltzer and Sumiyoshi156), and has no effect or even decreases hippocampal neurogenesis (Reference Eisch, Barrot, Schad, Self and Nestler157Reference Benninghoff, Grunze, Schindler, Genius, Schloesser, van der Ven, Dehning, Wiltfang, Moller and Rujescu159). On the other hand, atypical antipsychotics, such as olanzapine, risperidone (Reference Wakade, Mahadik, Waller and Chiu160), clozapine (Reference Halim, Weickert, McClintock, Weinberger and Lipska161) and ziprasidone (Reference Benninghoff, Grunze, Schindler, Genius, Schloesser, van der Ven, Dehning, Wiltfang, Moller and Rujescu159,Reference Peng, Zhang, Wang, Chen, Xue, Wang, Yang, Chen, Liu, Kuang and Tan162), increase cell proliferation in both neurogenic regions (i.e. the hippocampal SGZ and SVZ). Chronic treatment with olanzapine also increases the number of proliferating cells in the prelimbic cortex of rats (Reference Kodama, Fujioka and Duman163). Increased neurogenesis contributes to neuronal replenishment and might explain the observed amelioration of cognitive and negative symptoms elicited by atypical antipsychotics.

According to animal and human studies, CB1 and CB2 receptor functions, as well as AEA and 2-AG levels, are involved in the pathophysiology of schizophrenia (Reference Fakhoury164). CP-55940, a CB1/CB2 receptor agonist, abolished the oscillatory activity at the θ frequency and impaired the sensory gating function in the limbic circuitry of rats, further supporting the connection between Cannabis abuse and an increased risk of developing schizophrenia (Reference Hajos, Hoffmann and Kocsis165). A cross-sectional survey study published in 2004 suggested that Cannabis abuse during the critical period of neuroplasticity in adolescence is associated with positive and negative manifestations of psychosis (Reference Stefanis, Delespaul, Henquet, Bakoula, Stefanis and Van Os166). As mentioned above, the ECBS regulates fundamental developmental processes such as cell proliferation, migration, differentiation, synaptogenesis and survival during patterning of the CNS (Reference Maccarrone, Guzman, Mackie, Doherty and Harkany67,Reference Keimpema, Alpar, Howell, Malenczyk, Hobbs, Hurd, Watanabe, Sakimura, Kano, Doherty and Harkany70,Reference Zimmer, Zimmer, Hohmann, Herkenham and Bonner77). Accordingly, changes in ECBS-related genes have been reported in the brains of patients with schizophrenia (Reference Ujike, Takaki, Nakata, Tanaka, Takeda, Kodama, Fujiwara, Sakai and Kuroda167Reference Martinez-Gras, Hoenicka, Ponce, Rodriguez-Jimenez, Jimenez-Arriero, Perez-Hernandez, Ampuero, Ramos-Atance, Palomo and Rubio169).

Only a few researchers have explored the link between neurogenesis, schizophrenia and cannabinoids. In the study by Bortolato et al. (Reference Bortolato, Bini, Frau, Devoto, Pardu, Fan and Solbrig170), a 2-week administration of the potent non-selective cannabinoid receptor agonist WIN 55,212-2 (2 mg/kg) to juvenile male Lewis rats increased the survival of new cells, mainly neural glial antigen 2- or glial fibrillary acidic protein-positive cells, in the striatum and prefrontal cortex, two key terminal fields of dopaminergic pathways. The same treatment increased striatal dopamine metabolism and turnover in adulthood. The neurochemical changes were accompanied by behavioural alterations that are potentially related to attention deficits, such as slow reaction time and increased novelty-seeking behaviours (Table 1). The authors concluded that cannabinoid receptor agonism by WIN 55,212-2 might impact behaviours related to high dopaminergic metabolism and alter frontostriatal neurogenesis and gliogenesis.

Cannabinoids, adult neurogenesis and brain ischaemia

Hypoxia or ischaemia during prenatal asphyxia, severe hypotensive shock, atrial fibrillation, cardiac arrest (i.e. global brain ischaemia), or embolic/thrombotic occlusion of one or more cerebral vessels [i.e. focal brain ischaemia or stroke (Reference Martinez-Orgado, Fernandez-Lopez, Lizasoain and Romero171,Reference Ritz, van Buchem and Daemen172)] severely impairs brain blood perfusion. The process of pathological ischaemia begins with the breakdown of ion homoeostasis in the neuronal membrane caused by energy collapse, leading to anoxic depolarisation, massive glutamate release and oxidative stress in adjacent postsynaptic cells. These changes occur within minutes and comprise the acute excitotoxic phase of brain ischaemia, culminating in necrotic cell death in the infarcted region. In the subsequent hours to days (i.e. the reperfusion phase), further neurovascular changes occur when blood and oxygen re-enter the infarcted area, including membrane degradation, mitochondrial damage, neuroinflammation and apoptosis. A series of protective mechanisms, including neurogenesis and angiogenesis, may be activated to counteract these pathological ischaemic events (Reference Dirnagl173Reference Wiltrout, Lang, Yan, Dempsey and Vemuganti176). Increased hippocampal neurogenesis promotes spatial memory recovery after focal (Reference Pu, Jiang, Hu, Xia, Hong, Zhang, Gao, Chen and Shi177) and global (Reference Mori, Meyer, Soares, Milani, Guimaraes and de Oliveira178) brain ischaemia, whereas the inhibition of hippocampal neurogenesis exacerbates ischaemia-induced cognitive impairments (Reference Heurteaux, Widmann, Moha ou Maati, Quintard, Gandin, Borsotto, Veyssiere, Onteniente and Lazdunski175,Reference Pu, Jiang, Hu, Xia, Hong, Zhang, Gao, Chen and Shi177Reference Han, Wu, Han, Yang, Wang and Fang179). Nonetheless, a substantial proportion of newly generated neurons dies after ischaemic insult (Reference Arvidsson, Collin, Kirik, Kokaia and Lindvall174). Therefore, therapeutic agents protecting against ischaemic brain injury should, ideally, be able to exert multiple effects on impeding the ischaemic cascade propagation, as well as stimulating the proliferation and differentiation of new neural cells to repair damaged areas (Reference Heurteaux, Widmann, Moha ou Maati, Quintard, Gandin, Borsotto, Veyssiere, Onteniente and Lazdunski175).

Concerning the mechanisms of neuroprotection, CB1 receptor activation may prevent neuronal death and stimulate neurogenesis after brain ischaemia. In a pioneer study, Nagayama et al. (Reference Nagayama, Sinor, Simon, Chen, Graham, Jin and Greenberg180), have shown that the synthetic cannabinoid agonist WIN 55,212-2 decreased hippocampal neuronal loss after transient global cerebral ischaemia and reduced infarct volume after permanent focal cerebral ischaemia. These effects were blocked by the specific CB1 receptor antagonist SR141716A (Reference Nagayama, Sinor, Simon, Chen, Graham, Jin and Greenberg180). In another study, WIN 55,212-2 (0.1 mg/kg, single doses) enhanced cell proliferation, oligodendrogenesis and neuroblast generation in the striatum and SVZ of newborn rats exposed to acute hypoxia-ischaemia (Reference Fernandez-Lopez, Pradillo, Garcia-Yebenes, Martinez-Orgado, Moro and Lizasoain181).

Using a model of focal brain ischaemia [i.e. middle cerebral artery occlusion (MCAO)], Sun et al. (Reference Sun, Fang, Ren, Chen, Guo, Yan, Song, Zhang and Liao142) reported an increase in the expression of CB1 receptors in the ischaemic penumbra area 2 h after the ischaemic insult. The administration of WIN 55,212-2 (9 mg/kg, i.v.) significantly attenuated brain swelling and reduced the infarct volume (Table 2). WIN 55,212-2 also promoted the proliferation of NG2-positive cells in the ischaemic penumbra area and ipsilateral SVZ following the ischaemic insult. The selective CB1 receptor antagonist rimonabant (1 mg/kg, i.v.) partially blocked the effects of WIN 55,212-2. Moreover, Caltana et al. (Reference Caltana, Saez, Aronne and Brusco182) reported neuroprotective effects of the CB1 receptor agonist arachidonyl-2-chloroethylamide (ACEA) on mice subjected to MCAO. An ACEA treatment counteracted the functional impairments and attenuated the astrocytic reaction and neuronal death in ischaemic mice. ACEA also affected neural plasticity by increasing dendritic thickness and synaptogenesis in the brains of ischaemic mice. In contrast, treatment with the CB1 antagonist AM251 decreased these parameters. Thus, CB1 receptors stimulate adult neurogenesis following brain ischaemia. However, the simultaneous activation of both CB1 and CB2 receptors might be necessary for neuroprotection in response to ischaemic injuries. For example, Fernández-López et al. (Reference Fernandez-Lopez, Martinez-Orgado, Nunez, Romero, Lorenzo, Moro and Lizasoain183) showed that the combined administration of the CB1 antagonist SR141716 and the CB2 antagonist SR144528 reversed the neuroprotective effects of WIN 55,212-2 on brain slices from 7-day-old Wistar rats exposed to oxygen-glucose deprivation.

Table 2 Cannabinoids agonists increase adult neurogenesis in animal models of brain ischaemia and Alzheimer’s disease

↓, decreases; ↑, increases; Aβ, β-amyloid; ACEA, arachidonyl-2-chloroethylamide; APP, amyloid precursor protein; BDNF, brain-derived neurotrophic factor; DG, dentate gyrus; FAAH, fatty acid amide hydrolase; i.c.v., intracerebroventricular; i.p., intraperitoneal; i.v., intra vascular; OB, olfactory bulb; s.c., subcutaneus; SVZ, subventricular zone.

* All males.

Cannabinoid receptor agonist.

CB1 receptor agonist.

** CB1 receptor agonist.

†† CB2 agonist.

Recently, an important role for CB2 receptor in poststroke spontaneous recovery has been reported. Bravo-Ferrer et al. (Reference Bravo-Ferrer, Cuartero, Zarruk, Pradillo, Hurtado, Romera, Diaz-Alonso, Garcia-Segura, Guzman, Lizasoain, Galve-Roperh and Moro184) have demonstrated that subacute pharmacological blockage of the CB2 receptor with SR144528 or after CB2 genetic deletion inhibited stroke-induced neurogenesis by reducing the migration of neuroblasts toward the injured cortex, after permanent middle artery occlusion in mice.

CB1 and CB2 receptors are also associated with postnatal oligodendrogenesis. CB1 receptor activation increases the number of glial precursors in the rat SVZ. In addition, CB2 receptor activation increases PS-NCAM expression, which is required for the migration of oligodendrocyte precursors (Reference Arevalo-Martin, Garcia-Ovejero, Gomez, Rubio-Araiz, Navarro-Galve, Guaza, Molina-Holgado and Molina-Holgado185). Furthermore, modulation of the inflammatory response by CB2 receptors reduces damage and increases neuronal survival during the initial and later phases of ischaemic brain injury (Reference Mori, Meyer, Soares, Milani, Guimaraes and de Oliveira178,Reference Fernandez-Lopez, Martinez-Orgado, Nunez, Romero, Lorenzo, Moro and Lizasoain183). However, further studies are necessary to determine the mechanisms by which CB1 and CB2 receptor signalling contribute to the neuroplastic effects of cannabinoids on brain ischaemia.

Cannabinoids, adult neurogenesis and AD

AD is the most common form of dementia among the elderly (Reference Alzheimer’s132,Reference Gale, Acar and Daffner186). Memory impairments, cognitive and functional deterioration, and olfactory deficits are characteristic symptoms of this disease. Although a small proportion of AD cases (<5%) have a genetic basis (familial AD), the majority of cases are sporadic with an as yet unknown aetiology (Reference Gotz and Ittner187,Reference Dorszewska, Prendecki, Oczkowska, Dezor and Kozubski188). The pathological hallmarks of AD are the presence of amyloid senile plaques composed of extracellular deposits of β-amyloid (Aβ) peptide derived from aberrant processing of the transmembrane amyloid precursor protein (APP) and the hyperphosphorylation of the microtubule-associated protein τ, resulting in formation of the intracellular neurofibrillary tangles that impair inter-neuronal communication (Reference Sperling, Mormino and Johnson189Reference Mi and Johnson191). The brains of patients with AD show signs of neurodegeneration, oxidative damage, neuroinflammation and reduced cholinergic activity in areas related to memory processing (Reference Schliebs and Arendt192). Synapse loss in the hippocampus and neocortex has been considered the primary structural correlate of cognitive decline in patients with AD (Reference Mann193,Reference Raskin, Cummings, Hardy, Schuh and Dean194).

Changes in adult hippocampal neurogenesis have been reported in AD (Reference Martinez-Canabal97,Reference Radad, Moldzio, Al-Shraim, Kranner, Krewenka and Rausch195). A moderate decline in hippocampal neurogenesis (Reference Crews, Adame, Patrick, Delaney, Pham, Rockenstein, Hansen and Masliah196) and a failure in neuronal maturation (Reference Jin, Peel, Mao, Xie, Cottrell, Henshall and Greenberg197) have been observed in postmortem brains of patients with AD. On the other hand, increase in the proliferation of hippocampal progenitor cells was detected during the onset, middle and advanced stages of AD (Reference Jin, Peel, Mao, Xie, Cottrell, Henshall and Greenberg197,Reference Perry, Johnson, Ekonomou, Perry, Ballard and Attems198). One study showed an increase in the levels of several immature neuronal markers, such as DCX, PS-NCAM, neurogenic differentiation factor and TUC-4, in a cohort of patients with the senile AD (Reference Jin, Peel, Mao, Xie, Cottrell, Henshall and Greenberg197). In a younger cohort of presenile patients with a faster and more severe disease course, however, these results were not replicated (Reference Boekhoorn, Joels and Lucassen199). Nevertheless, increased hippocampal neurogenesis in AD patients may represent a compensatory mechanism for endogenous brain repair and to counteract disease-related inflammation (Reference Martinez-Canabal97).

The neuropathological and cognitive features of patients with AD have been successfully mimicked in transgenic models by manipulating genes involved in the familial AD, such as APP, presenisilin-1 and presenilin-2, which lead to the production and deposition of Aβ plaques (Reference Bilkei-Gorzo200). Interestingly, these genes also modulate neurogenesis (Reference Marlatt and Lucassen201). Similar to human patients with AD, transgenic animal models of AD develop severe cognitive deficits and hippocampal degeneration (Reference Bilkei-Gorzo200). However, the results regarding adult neurogenesis are again highly variable, probably because of methodological differences in the age of the animals, transgene expression, Aβ deposition and neurotransmitter levels. Both decreased and increased hippocampal neurogenesis have been reported in transgenic models of AD (Reference Marlatt and Lucassen201).

Several reports point out a possible implication of the ECBS in AD in the modulation of events occurring during the course of AD progression evaluated from early- to late symptomatic AD-likes stages, in postmortem AD brains and genetically modified mice (Reference Kalifa, Polston, Allard and Manaye202,Reference Aso, Palomer, Juves, Maldonado, Munoz and Ferrer203,Reference Marchalant, Baranger, Wenk, Khrestchatisky and Rivera204). In brains of AD patients, the microglial CB1 receptor is increased mostly in plaque-bearing areas (Reference Ramirez, Blazquez, Gomez del Pulgar, Guzman and de Ceballos205), while neuronal CB1 receptor expression is reduced in the hippocampus and prefrontal cortex (Reference Ramirez, Blazquez, Gomez del Pulgar, Guzman and de Ceballos205,Reference Westlake, Howlett, Bonner, Matsuda and Herkenham206). An upregulation on the FAAH levels on plaque-associated astrocytes has been also reported in postmortem AD brains (Reference Benito, Nunez, Tolon, Carrier, Rabano, Hillard and Romero207). However, other authors have demonstrated no changes in CB receptors expression in the hippocampus or cortex of AD patients (Reference Lee, Agacinski, Williams, Wilcock, Esiri, Francis, Wong, Chen and Lai208Reference Ahmad, Goffin, Van den Stock, De Winter, Cleeren, Bormans, Tournoy, Persoons, Van Laere and Vandenbulcke210). Recent studies have also not found any difference in the CB1 protein level in the hippocampus of AD transgenic mice in a pre-symptomatic stage of AD (Reference Bedse, Romano, Cianci, Lavecchia, Lorenzo, Elphick, Laferla, Vendemiale, Grillo, Altieri, Cassano and Gaetani211,Reference Maccarrone, Totaro, Leuti, Giacovazzo, Scipioni, Mango, Coccurello, Nistico and Oddi212). Otherwise, the CB2 expression is increased in the hippocampus and prefrontal cortex in postmortem brains of AD patients (Reference Benito, Nunez, Tolon, Carrier, Rabano, Hillard and Romero207,Reference Solas, Francis, Franco and Ramirez213) and also in a mouse model of Aβ amyloidosis (Reference Savonenko, Melnikova, Wang, Ravert, Gao, Koppel, Lee, Pletnikova, Cho, Sayyida, Hiatt, Troncoso, Davies, Dannals, Pomper and Horti214), suggesting the involvement of CB2 receptors in the pathogenesis of AD.

Nevertheless, strategies targeting adult neurogenesis with cannabinoids have been used as a means to mitigate the symptoms of AD under several experimental conditions (Reference Marchalant, Baranger, Wenk, Khrestchatisky and Rivera204,Reference Aso, Andres-Benito and Ferrer215,Reference Watt and Karl216). The CB1 receptor agonist ACEA at pre-symptomatic or at early stages reduced the cognitive deficits and decreased inflammatory response in the vicinity of Aβ plaques in transgenic animals (Reference Aso, Palomer, Juves, Maldonado, Munoz and Ferrer203). CB2 receptor agonists also reduced inflammation induced by Aβ production and deposition, promoted Aβ clearance and increased cell viability in the presence of Aβ (Reference Aso, Andres-Benito and Ferrer215,Reference Aso, Juves, Maldonado and Ferrer217). Moreover, CB2 selective and CB1-CB2 mixed agonists prevent memory impairments in AD rats and mice after chronic administration (Reference Ramirez, Blazquez, Gomez del Pulgar, Guzman and de Ceballos205,Reference Aso, Juves, Maldonado and Ferrer217,Reference Martin-Moreno, Reigada, Ramirez, Mechoulam, Innamorato, Cuadrado and de Ceballos218). Finally, treatment with cannabidiol reduced Aβ-induced neuroinflammation (Reference Esposito, Scuderi, Valenza, Togna, Latina, De Filippis, Cipriano, Carratu, Iuvone and Steardo219,Reference Chen, Bromley-Brits, He, Cai, Zhang and Song220), rescued spatial memory deficits and promoted microglial migration, a cellular mechanism that may enable the removal of Aβ deposits (Reference Martin-Moreno, Reigada, Ramirez, Mechoulam, Innamorato, Cuadrado and de Ceballos218).

Considering the role of cannabinoids on adult neurogenesis, Esposito et al. (Reference Esposito, Scuderi, Valenza, Togna, Latina, De Filippis, Cipriano, Carratu, Iuvone and Steardo219) have shown that 15 days of cannabidiol (10 mg/kg) counteracts the Aβ-induced DCX depletion and stimulates basal neurogenesis in rats injected with Aβ into the hippocampus. This therapeutic effect was attributed to the selective activation of PPAR-γ receptors by cannabidiol, since previous injections of GW9662, a selective PPAR-γ antagonist, abolished these effects. However, chronic treatment with the synthetic cannabinoid agonist HU-210 failed to produce any beneficial effects on APP23/PS45 double transgenic AD mice. HU-210 treatment did not improve cognitive deficits measured in the water maze and contextual fear conditioning tasks had no effect on Aβ generation or plaque formation in the brains of AD transgenic mice and did not affect adult hippocampal neurogenesis. Chronic treatment with high doses of HU-210 (20 mg/kg) even decreased hippocampal neurogenesis in AD transgenic mice (Reference Chen, Bromley-Brits, He, Cai, Zhang and Song220). Further work is necessary to elucidate the effects of cannabinoids on altered hippocampal neurogenesis observed in experimental AD animal models.

Conclusions and perspectives

Drugs that are currently available to treat psychiatric and neurological disorders are frequently associated with delayed and partial therapeutic responses, as well as substantial side effects (Reference Kang, Wen, Song, Christian and Ming110). Thus, new and more efficient drugs are required. Based on the results presented here regarding neurogenesis and the relevance of the ECBS to CNS functions, pharmacological approaches based on cannabinoids may offer a promising strategy to both treat and prevent several brain disorders.

In the present review, we summarised the main lines of evidence supporting the effects of cannabinoids on CNS development, their impacts on proliferative processes in the adult brain, and the possible implications of ECBS-induced neurogenesis in psychiatric and neurological conditions. The vast majority the studies reviewed here examined the role of cannabinoids in adult hippocampal neurogenesis, probably reflecting the extent of the literature on the relationship between hippocampal function and the behavioural and cognitive symptoms of psychiatric and neurological disorders. However, the effects of these drugs on CNS embryogenesis and their possible associations with the pathogenesis of these disorders require further investigation.

Several questions remain to be answered, including the precise mechanism by which cannabinoids regulate neurogenesis and cell fate, as well the relevance of non-cannabinoid receptor-mediated mechanisms (e.g. TRPV1, GPR55, and PPAR-γ receptors).

Notably, although this topic is beyond of the scope of the present review, studies have reported that disrupted neurogenesis confers susceptibility to addictive behaviours in rodents. Most drugs of abuse suppress neurogenesis, and the recovery of drug-impaired neurogenesis may be an important mechanism to improve neuroplasticity during abstinence and, therefore, recovery (Reference Mandyam and Koob221). Cannabis is the most commonly used illicit drug worldwide, and although researchers have been extensively studied the effects of Cannabis use on neurodevelopment, the effects of THC or marijuana on adult neurogenesis are still under debate (Reference Richardson, Hester and McLemore137,Reference Fried138). Therefore, new studies comparing the acute and long-term effects of cannabinoid signalling on facilitating neurogenesis and brain functions during different life stages (mainly the critical periods of neuroplasticity) are needed.

Supplementary material

To view supplementary material for this article, please visit https://doi.org/10.1017/neu.2018.11

Acknowledgements

The authors would like to thank the members of our research groups for cultivating an inspiring scientific environment. The authors thank Franciele F. Scarante and Marco Aurélio Mori, PhD, for their assistance in designing the figures. The authors would like to apologise to the researchers whose studies were not cited here due to space limitations. R.M.W.O. and C.L.O. are recipients of CNPq and CAPES grants. A.C.C. and F.S.G. are recipients of FAPESP grants (numbers 15/05551-0 and 12/17626-7, respectively).

References

1. Gage, FH, Kempermann, G, Palmer, TD, Peterson, DA Ray, J (1998) Multipotent progenitor cells in the adult dentate gyrus. J Neurobiol 36, 249266.Google Scholar
2. Lindsey, BW Tropepe, V (2006) A comparative framework for understanding the biological principles of adult neurogenesis. Prog Neurobiol 80, 281307.Google Scholar
3. Czaja, K, Fornaro, M Geuna, S (2012) Neurogenesis in the adult peripheral nervous system. Neural Regen Res 7, 10471054.Google Scholar
4. Altman, J Das, GD (1965) Autoradiographic and histological evidence of postnatal hippocampal neurogenesis in rats. J Comp Neurol 124, 319335.Google Scholar
5. Jordan, JD, Ma, DK, Ming, GL Song, H (2007) Cellular niches for endogenous neural stem cells in the adult brain. CNS Neurol Disord Drug Targets 6, 336341.Google Scholar
6. Ming, GL Song, H (2011) Adult neurogenesis in the mammalian brain: significant answers and significant questions. Neuron 70, 687702.Google Scholar
7. Corotto, FS, Henegar, JA Maruniak, JA (1993) Neurogenesis persists in the subependymal layer of the adult mouse brain. Neurosci Lett 149, 111114.Google Scholar
8. Garcia-Verdugo, JM, Doetsch, F, Wichterle, H, Lim, DA Alvarez-Buylla, A (1998) Architecture and cell types of the adult subventricular zone: in search of the stem cells. J Neurobiol 36, 234248.Google Scholar
9. Eriksson, PS, Perfilieva, E, Bjork-Eriksson, T, Alborn, AM, Nordborg, C, Peterson, DA Gage, FH (1998) Neurogenesis in the adult human hippocampus. Nat Med 4, 13131317.Google Scholar
10. Kornack, DR Rakic, P (2001) The generation, migration, and differentiation of olfactory neurons in the adult primate brain. Proc Natl Acad Sci U S A 98, 47524757.Google Scholar
11. Spalding, KL, Bergmann, O, Alkass, K, Bernard, S, Salehpour, M, Huttner, HB, Bostrom, E, Westerlund, I, Vial, C, Buchholz, BA, Possnert, G, Mash, DC, Druid, H Frisen, J (2013) Dynamics of hippocampal neurogenesis in adult humans. Cell 153, 12191227.Google Scholar
12. Snyder, JS Cameron, HA (2012) Could adult hippocampal neurogenesis be relevant for human behavior? Behav Brain Res 227, 384390.Google Scholar
13. Ihunwo, AO, Tembo, LH Dzamalala, C (2016) The dynamics of adult neurogenesis in human hippocampus. Neural Regen Res 11, 18691883.Google Scholar
14. Bergami, M, Masserdotti, G, Temprana, SG, Motori, E, Eriksson, TM, Gobel, J, Yang, SM, Conzelmann, KK, Schinder, AF, Gotz, M Berninger, B (2015) A critical period for experience-dependent remodeling of adult-born neuron connectivity. Neuron 85, 710717.Google Scholar
15. Urban, N Guillemot, F (2014) Neurogenesis in the embryonic and adult brain: same regulators, different roles. Front Cell Neurosci 8, 396.Google Scholar
16. Jessberger, S, Toni, N, Clemenson, GD Jr., Ray, J Gage, FH (2008) Directed differentiation of hippocampal stem/progenitor cells in the adult brain. Nat Neurosci 11, 888893.Google Scholar
17. Carr, VM Farbman, AI (1993) The dynamics of cell death in the olfactory epithelium. Exp Neurol 124, 308314.Google Scholar
18. Almeida, OF, Conde, GL, Crochemore, C, Demeneix, BA, Fischer, D, Hassan, AH, Meyer, M, Holsboer, F Michaelidis, TM (2000) Subtle shifts in the ratio between pro- and antiapoptotic molecules after activation of corticosteroid receptors decide neuronal fate. FASEB J 14, 779790.Google Scholar
19. Andersen, J, Urban, N, Achimastou, A, Ito, A, Simic, M, Ullom, K, Martynoga, B, Lebel, M, Goritz, C, Frisen, J, Nakafuku, M Guillemot, F (2014) A transcriptional mechanism integrating inputs from extracellular signals to activate hippocampal stem cells. Neuron 83, 10851097.Google Scholar
20. Paridaen, JT Huttner, WB (2014) Neurogenesis during development of the vertebrate central nervous system. EMBO Rep 15, 351364.Google Scholar
21. Bond, AM, Ming, GL Song, H (2015) Adult mammalian neural stem cells and neurogenesis: five decades later. Cell Stem Cell 17, 385395.Google Scholar
22. Guan, K, Chang, H, Rolletschek, A Wobus, AM (2001) Embryonic stem cell-derived neurogenesis. Retinoic acid induction and lineage selection of neuronal cells. Cell Tissue Res 305, 171176.Google Scholar
23. Gotz, M Huttner, WB (2005) The cell biology of neurogenesis. Nat Rev Mol Cell Biol 6, 777788.Google Scholar
24. Qian, X, Shen, Q, Goderie, SK, He, W, Capela, A, Davis, AA Temple, S (2000) Timing of CNS cell generation: a programmed sequence of neuron and glial cell production from isolated murine cortical stem cells. Neuron 28, 6980.Google Scholar
25. Alvarez-Buylla, A (1990) Mechanism of neurogenesis in adult avian brain. Experientia 46, 948955.Google Scholar
26. Cayre, M, Canoll, P Goldman, JE (2009) Cell migration in the normal and pathological postnatal mammalian brain. Prog Neurobiol 88, 4163.Google Scholar
27. Lois, C Alvarez-Buylla, A (1994) Long-distance neuronal migration in the adult mammalian brain. Science 264, 11451148.Google Scholar
28. Cameron, HA, Woolley, CS, McEwen, BS Gould, E (1993) Differentiation of newly born neurons and glia in the dentate gyrus of the adult rat. Neuroscience 56, 337344.Google Scholar
29. Cipriani, S, Nardelli, J, Verney, C, Delezoide, AL, Guimiot, F, Gressens, P Adle-Biassette, H (2016) Dynamic expression patterns of progenitor and pyramidal neuron layer markers in the developing human hippocampus. Cereb Cortex 26, 12551271.Google Scholar
30. Pencea, V, Bingaman, KD, Freedman, LJ Luskin, MB (2001) Neurogenesis in the subventricular zone and rostral migratory stream of the neonatal and adult primate forebrain. Exp Neurol 172, 116.Google Scholar
31. Peretto, P, Merighi, A, Fasolo, A Bonfanti, L (1997) Glial tubes in the rostral migratory stream of the adult rat. Brain Res Bull 42, 921.Google Scholar
32. Sawamoto, K, Wichterle, H, Gonzalez-Perez, O, Cholfin, JA, Yamada, M, Spassky, N, Murcia, NS, Garcia-Verdugo, JM, Marin, O, Rubenstein, JL, Tessier-Lavigne, M, Okano, H Alvarez-Buylla, A (2006) New neurons follow the flow of cerebrospinal fluid in the adult brain. Science 311, 629632.Google Scholar
33. Alvarez-Buylla, A Garcia-Verdugo, JM (2002) Neurogenesis in adult subventricular zone. J Neurosci 22, 629634.Google Scholar
34. Kintner, C (2002) Neurogenesis in embryos and in adult neural stem cells. J Neurosci 22, 639643.Google Scholar
35. Jagasia, R, Song, H, Gage, FH Lie, DC (2006) New regulators in adult neurogenesis and their potential role for repair. Trends Mol Med 12, 400405.Google Scholar
36. Pathania, M, Yan, LD Bordey, A (2010) A symphony of signals conducts early and late stages of adult neurogenesis. Neuropharmacology 58, 865876.Google Scholar
37. Hagg, T (2005) Molecular regulation of adult CNS neurogenesis: an integrated view. Trends Neurosci 28, 589595.Google Scholar
38. Katona, I, Urban, GM, Wallace, M, Ledent, C, Jung, KM, Piomelli, D, Mackie, K Freund, TF (2006) Molecular composition of the endocannabinoid system at glutamatergic synapses. J Neurosci 26, 56285637.Google Scholar
39. Harkany, T, Keimpema, E, Barabas, K Mulder, J (2008) Endocannabinoid functions controlling neuronal specification during brain development. Mol Cell Endocrinol 286, S84S90.Google Scholar
40. Pertwee, RG (2005) Pharmacological actions of cannabinoids. Handb Exp Pharmacol 168, 151.Google Scholar
41. Mechoulam, R Gaoni, Y (1965) A total synthesis of dl-delta-1-tetrahydrocannabinol, the active constituent of hashish. J Am Chem Soc 87, 32733275.Google Scholar
42. Turner, SE, Williams, CM, Iversen, L Whalley, BJ (2017) Molecular pharmacology of phytocannabinoids. Prog Chem Org Nat Prod 103, 61101.Google Scholar
43. Devane, WA, Dysarz, FA 3rd, Johnson, MR, Melvin, LS Howlett, AC (1988) Determination and characterization of a cannabinoid receptor in rat brain. Mol Pharmacol 34, 605613.Google Scholar
44. Matsuda, LA, Lolait, SJ, Brownstein, MJ, Young, AC Bonner, TI (1990) Structure of a cannabinoid receptor and functional expression of the cloned cDNA. Nature 346, 561564.Google Scholar
45. Munro, S, Thomas, KL Abu-Shaar, M (1993) Molecular characterization of a peripheral receptor for cannabinoids. Nature 365, 6165.Google Scholar
46. Szabo, B Schlicker, E (2005) Effects of cannabinoids on neurotransmission. Handb Exp Pharmacol 168, 327365.Google Scholar
47. Tsou, K, Mackie, K, Sanudo-Pena, MC Walker, JM (1999) Cannabinoid CB1 receptors are localized primarily on cholecystokinin-containing GABAergic interneurons in the rat hippocampal formation. Neuroscience 93, 969975.Google Scholar
48. Wilson, RI Nicoll, RA (2001) Endogenous cannabinoids mediate retrograde signalling at hippocampal synapses. Nature 410, 588592.Google Scholar
49. Takahashi, KA Castillo, PE (2006) The CB1 cannabinoid receptor mediates glutamatergic synaptic suppression in the hippocampus. Neuroscience 139, 795802.Google Scholar
50. Lau, T Schloss, P (2008) The cannabinoid CB1 receptor is expressed on serotonergic and dopaminergic neurons. Eur J Pharmacol 578, 137141.Google Scholar
51. Yoshida, T, Hashimoto, K, Zimmer, A, Maejima, T, Araishi, K Kano, M (2002) The cannabinoid CB1 receptor mediates retrograde signals for depolarization-induced suppression of inhibition in cerebellar Purkinje cells. J Neurosci 22, 16901697.Google Scholar
52. Diana, MA Marty, A (2004) Endocannabinoid-mediated short-term synaptic plasticity: depolarization-induced suppression of inhibition (DSI) and depolarization-induced suppression of excitation (DSE). Br J Pharmacol 142, 919.Google Scholar
53. Marsicano, G, Goodenough, S, Monory, K, Hermann, H, Eder, M, Cannich, A, Azad, SC, Cascio, MG, Gutierrez, SO, van der Stelt, M, Lopez-Rodriguez, ML, Casanova, E, Schutz, G, Zieglgansberger, W, Di Marzo, V, Behl, C Lutz, B (2003) CB1 cannabinoid receptors and on-demand defense against excitotoxicity. Science 302, 8488.Google Scholar
54. Fogaca, MV, Galve-Roperh, I, Guimaraes, FS Campos, AC (2013) Cannabinoids, neurogenesis and antidepressant drugs: Is there a link? Curr Neuropharmacol 11, 263275.Google Scholar
55. Onaivi, ES, Ishiguro, H, Gong, JP, Patel, S, Perchuk, A, Meozzi, PA, Myers, L, Mora, Z, Tagliaferro, P, Gardner, E, Brusco, A, Akinshola, BE, Liu, QR, Hope, B, Iwasaki, S, Arinami, T, Teasenfitz, L Uhl, GR (2006) Discovery of the presence and functional expression of cannabinoid CB2 receptors in brain. Ann N Y Acad Sci 1074, 514536.Google Scholar
56. Palazuelos, J, Aguado, T, Egia, A, Mechoulam, R, Guzman, M Galve-Roperh, I (2006) Non-psychoactive CB2 cannabinoid agonists stimulate neural progenitor proliferation. FASEB J 20, 24052407.Google Scholar
57. Lisboa, SF, Gomes, FV, Guimaraes, FS Campos, AC (2016) Microglial cells as a link between cannabinoids and the immune hypothesis of psychiatric disorders. Front Neurol 7, 5.Google Scholar
58. Xi, ZX, Peng, XQ, Li, X, Song, R, Zhang, HY, Liu, QR, Yang, HJ, Bi, GH, Li, J Gardner, EL (2011) Brain cannabinoid CB(2) receptors modulate cocaine’s actions in mice. Nat Neurosci 14, 11601166.Google Scholar
59. Devane, WA, Hanus, L, Breuer, A, Pertwee, RG, Stevenson, LA, Griffin, G, Gibson, D, Mandelbaum, A, Etinger, A Mechoulam, R (1992) Isolation and structure of a brain constituent that binds to the cannabinoid receptor. Science 258, 19461949.Google Scholar
60. Mechoulam, R, Ben-Shabat, S, Hanus, L, Ligumsky, M, Kaminski, NE, Schatz, AR, Gopher, A, Almog, S, Martin, BR, Compton, DR, Pertwee, RG, Griffin, G, Bayewitch, M, Barg, J Vogel, Z (1995) Identification of an endogenous 2-monoglyceride, present in canine gut, that binds to cannabinoid receptors. Biochem Pharmacol 50: 83--90.Google Scholar
61. Saito, VM, Wotjak, CT Moreira, FA (2010) Pharmacological exploitation of the endocannabinoid system: new perspectives for the treatment of depression and anxiety disorders? Rev Bras Psiquiatr 32(Suppl. 1):S7S14.Google Scholar
62. Campos, AC, Paraíso-Luna, J, Fogaça, MV, Guimarães, FS Galve-Roperh, I (2017) Cannabinoids as regulators of neural development and adult neurogenesis. Lipidomics of Stem Cells 6, 117136.Google Scholar
63. Aguado, T, Monory, K, Palazuelos, J, Stella, N, Cravatt, B, Lutz, B, Marsicano, G, Kokaia, Z, Guzman, M Galve-Roperh, I (2005) The endocannabinoid system drives neural progenitor proliferation. FASEB J 19, 17041706.Google Scholar
64. Aguado, T, Palazuelos, J, Monory, K, Stella, N, Cravatt, B, Lutz, B, Marsicano, G, Kokaia, Z, Guzman, M Galve-Roperh, I (2006) The endocannabinoid system promotes astroglial differentiation by acting on neural progenitor cells. J Neurosci 26, 15511561.Google Scholar
65. Diaz-Alonso, J, Guzman, M Galve-Roperh, I (2012) Endocannabinoids via CB(1) receptors act as neurogenic niche cues during cortical development. Philos Trans R Soc Lond B Biol Sci 367, 32293241.Google Scholar
66. Diaz-Alonso, J, Aguado, T, de Salas-Quiroga, A, Ortega, Z, Guzman, M Galve-Roperh, I (2015) CB1 Cannabinoid receptor-dependent activation of mTORC1/Pax6 signaling drives Tbr2 expression and basal progenitor expansion in the developing mouse cortex. Cereb Cortex 25, 23952408.Google Scholar
67. Maccarrone, M, Guzman, M, Mackie, K, Doherty, P Harkany, T (2014) Programming of neural cells by (endo)cannabinoids: from physiological rules to emerging therapies. Nat Rev Neurosci 15, 786801.Google Scholar
68. Keimpema, E, Tortoriello, G, Alpar, A, Capsoni, S, Arisi, I, Calvigioni, D, Hu, SS, Cattaneo, A, Doherty, P, Mackie, K Harkany, T (2013) Nerve growth factor scales endocannabinoid signaling by regulating monoacylglycerol lipase turnover in developing cholinergic neurons. Proc Natl Acad Sci U S A 110, 19351940.Google Scholar
69. Maison, P, Walker, DJ, Walsh, FS, Williams, G Doherty, P (2009) BDNF regulates neuronal sensitivity to endocannabinoids. Neurosci Lett 467, 9094.Google Scholar
70. Keimpema, E, Alpar, A, Howell, F, Malenczyk, K, Hobbs, C, Hurd, YL, Watanabe, M, Sakimura, K, Kano, M, Doherty, P Harkany, T (2013) Diacylglycerol lipase alpha manipulation reveals developmental roles for intercellular endocannabinoid signaling. Sci Rep 3, 2093.Google Scholar
71. Harkany, T, Guzman, M, Galve-Roperh, I, Berghuis, P, Devi, LA Mackie, K (2007) The emerging functions of endocannabinoid signaling during CNS development. Trends Pharmacol Sci 28, 8392.Google Scholar
72. Oudin, MJ, Hobbs, C Doherty, P (2011) DAGL-dependent endocannabinoid signalling: roles in axonal pathfinding, synaptic plasticity and adult neurogenesis. Eur J Neurosci 34, 16341646.Google Scholar
73. Goncalves, MB, Suetterlin, P, Yip, P, Molina-Holgado, F, Walker, DJ, Oudin, MJ, Zentar, MP, Pollard, S, Yanez-Munoz, RJ, Williams, G, Walsh, FS, Pangalos, MN Doherty, P (2008) A diacylglycerol lipase-CB2 cannabinoid pathway regulates adult subventricular zone neurogenesis in an age-dependent manner. Mol Cell Neurosci 38, 526536.Google Scholar
74. Campos, AC, Ortega, Z, Palazuelos, J, Fogaca, MV, Aguiar, DC, Diaz-Alonso, J, Ortega-Gutierrez, S, Vazquez-Villa, H, Moreira, FA, Guzman, M, Galve-Roperh, I Guimaraes, FS (2013) The anxiolytic effect of cannabidiol on chronically stressed mice depends on hippocampal neurogenesis: involvement of the endocannabinoid system. Int J Neuropsychopharmacol 16, 14071419.Google Scholar
75. Mato, S, Del Olmo, E Pazos, A (2003) Ontogenetic development of cannabinoid receptor expression and signal transduction functionality in the human brain. Eur J Neurosci 17, 17471754.Google Scholar
76. Berrendero, F, Mendizabal, V, Murtra, P, Kieffer, BL Maldonado, R (2003) Cannabinoid receptor and WIN 55 212-2-stimulated [35S]-GTPgammaS binding in the brain of mu-, delta- and kappa-opioid receptor knockout mice. Eur J Neurosci 18, 21972202.Google Scholar
77. Zimmer, A, Zimmer, AM, Hohmann, AG, Herkenham, M Bonner, TI (1999) Increased mortality, hypoactivity, and hypoalgesia in cannabinoid CB1 receptor knockout mice. Proc Natl Acad Sci U S A 96, 57805785.Google Scholar
78. Zurolo, E, Iyer, AM, Spliet, WG, Van Rijen, PC, Troost, D, Gorter, JA Aronica, E (2010) CB1 and CB2 cannabinoid receptor expression during development and in epileptogenic developmental pathologies. Neuroscience 170, 2841.Google Scholar
79. Mulder, J, Aguado, T, Keimpema, E, Barabas, K, Ballester Rosado, CJ, Nguyen, L, Monory, K, Marsicano, G, Di Marzo, V, Hurd, YL, Guillemot, F, Mackie, K, Lutz, B, Guzman, M, Lu, HC, Galve-Roperh, I Harkany, T (2008) Endocannabinoid signaling controls pyramidal cell specification and long-range axon patterning. Proc Natl Acad Sci U S A 105, 87608765.Google Scholar
80. Bisogno, T, Howell, F, Williams, G, Minassi, A, Cascio, MG, Ligresti, A, Matias, I, Schiano-Moriello, A, Paul, P, Williams, EJ, Gangadharan, U, Hobbs, C, Di Marzo, V Doherty, P (2003) Cloning of the first sn1-DAG lipases points to the spatial and temporal regulation of endocannabinoid signaling in the brain. J Cell Biol 163, 463468.Google Scholar
81. Alpar, A, Tortoriello, G, Calvigioni, D, Niphakis, MJ, Milenkovic, I, Bakker, J, Cameron, GA, Hanics, J, Morris, CV, Fuzik, J, Kovacs, GG, Cravatt, BF, Parnavelas, JG, Andrews, WD, Hurd, YL, Keimpema, E Harkany, T (2014) Endocannabinoids modulate cortical development by configuring Slit2/Robo1 signalling. Nat Commun 5, 4421.Google Scholar
82. Berghuis, P, Dobszay, MB, Wang, X, Spano, S, Ledda, F, Sousa, KM, Schulte, G, Ernfors, P, Mackie, K, Paratcha, G, Hurd, YL Harkany, T (2005) Endocannabinoids regulate interneuron migration and morphogenesis by transactivating the TrkB receptor. Proc Natl Acad Sci U S A 102, 1911519120.Google Scholar
83. Roland, AB, Ricobaraza, A, Carrel, D, Jordan, BM, Rico, F, Simon, A, Humbert-Claude, M, Ferrier, J, McFadden, MH, Scheuring, S Lenkei, Z (2014) Cannabinoid-induced actomyosin contractility shapes neuronal morphology and growth. Elife 3, e03159.Google Scholar
84. Morozov, YM, Torii, M Rakic, P (2009) Origin, early commitment, migratory routes, and destination of cannabinoid type 1 receptor-containing interneurons. Cereb Cortex 19(Suppl. 1):i7889.Google Scholar
85. Palazuelos, J, Ortega, Z, Diaz-Alonso, J, Guzman, M Galve-Roperh, I (2012) CB2 cannabinoid receptors promote neural progenitor cell proliferation via mTORC1 signaling. J Biol Chem 287, 11981209.Google Scholar
86. Molina-Holgado, F, Rubio-Araiz, A, Garcia-Ovejero, D, Williams, RJ, Moore, JD, Arevalo-Martin, A, Gomez-Torres, O Molina-Holgado, E (2007) CB2 cannabinoid receptors promote mouse neural stem cell proliferation. Eur J Neurosci 25, 629634.Google Scholar
87. Prenderville, JA, Kelly, AM Downer, EJ (2015) The role of cannabinoids in adult neurogenesis. Br J Pharmacol 172, 39503963.Google Scholar
88. Hill, MN, Titterness, AK, Morrish, AC, Carrier, EJ, Lee, TT, Gil-Mohapel, J, Gorzalka, BB, Hillard, CJ Christie, BR (2010) Endogenous cannabinoid signaling is required for voluntary exercise-induced enhancement of progenitor cell proliferation in the hippocampus. Hippocampus 20, 513523.Google Scholar
89. Gao, Y, Vasilyev, DV, Goncalves, MB, Howell, FV, Hobbs, C, Reisenberg, M, Shen, R, Zhang, MY, Strassle, BW, Lu, P, Mark, L, Piesla, MJ, Deng, K, Kouranova, EV, Ring, RH, Whiteside, GT, Bates, B, Walsh, FS, Williams, G, Pangalos, MN, Samad, TA Doherty, P (2010) Loss of retrograde endocannabinoid signaling and reduced adult neurogenesis in diacylglycerol lipase knock-out mice. J Neurosci 30, 20172024.Google Scholar
90. Jenniches, I, Ternes, S, Albayram, O, Otte, DM, Bach, K, Bindila, L, Michel, K, Lutz, B, Bilkei-Gorzo, A Zimmer, A (2016) Anxiety, stress, and fear response in mice with reduced endocannabinoid levels. Biol Psychiatry 79, 858868.Google Scholar
91. Wolf, SA, Bick-Sander, A, Fabel, K, Leal-Galicia, P, Tauber, S, Ramirez-Rodriguez, G, Muller, A, Melnik, A, Waltinger, TP, Ullrich, O Kempermann, G (2010) Cannabinoid receptor CB1 mediates baseline and activity-induced survival of new neurons in adult hippocampal neurogenesis. Cell Commun Signal 8, 12.Google Scholar
92. Campos, AC, Fogaca, MV, Scarante, FF, Joca, SRL, Sales, AJ, Gomes, FV, Sonego, AB, Rodrigues, NS, Galve-Roperh, I Guimaraes, FS (2017) Plastic and neuroprotective mechanisms involved in the therapeutic effects of cannabidiol in psychiatric disorders. Front Pharmacol 8, 269.Google Scholar
93. Jin, K, Xie, L, Kim, SH, Parmentier-Batteur, S, Sun, Y, Mao, XO, Childs, J Greenberg, DA (2004) Defective adult neurogenesis in CB1 cannabinoid receptor knockout mice. Mol Pharmacol 66, 204208.Google Scholar
94. Lee, S, Kim, DH, Yoon, SH Ryu, JH (2009) Sub-chronic administration of rimonabant causes loss of antidepressive activity and decreases doublecortin immunoreactivity in the mouse hippocampus. Neurosci Lett 467, 111116.Google Scholar
95. Han, MH, Lee, EH Koh, SH (2016) Current opinion on the role of neurogenesis in the therapeutic strategies for Alzheimer disease, Parkinson disease, and ischemic stroke; considering neuronal voiding function. Int Neurourol J 20, 276287.Google Scholar
96. Mu, Y Gage, FH (2011) Adult hippocampal neurogenesis and its role in Alzheimer’s disease. Mol Neurodegener 6, 85.Google Scholar
97. Martinez-Canabal, A (2014) Reconsidering hippocampal neurogenesis in Alzheimer’s disease. Front Neurosci 8, 147.Google Scholar
98. Iannitelli, A, Quartini, A, Tirassa, P Bersani, G (2017) Schizophrenia and neurogenesis: a stem cell approach. Neurosci Biobehav Rev 80, 414442.Google Scholar
99. Campos, AC, Moreira, FA, Gomes, FV, Del Bel, EA Guimaraes, FS (2012) Multiple mechanisms involved in the large-spectrum therapeutic potential of cannabidiol in psychiatric disorders. Philos Trans R Soc Lond B Biol Sci 367, 33643378.Google Scholar
100. Duman, RS, Malberg, J, Nakagawa, S D’Sa, C (2000) Neuronal plasticity and survival in mood disorders. Biol Psychiatry 48, 732739.Google Scholar
101. Vivar, C (2015) Adult hippocampal neurogenesis, aging and neurodegenerative diseases: possible strategies to prevent cognitive impairment. Curr Top Med Chem 15, 21752192.Google Scholar
102. Kawai, T, Takagi, N, Miyake-Takagi, K, Okuyama, N, Mochizuki, N Takeo, S (2004) Characterization of BrdU-positive neurons induced by transient global ischemia in adult hippocampus. J Cereb Blood Flow Metab 24, 548555.Google Scholar
103. Parent, JM (2003) Injury-induced neurogenesis in the adult mammalian brain. Neuroscientist 9, 261272.Google Scholar
104. Liu, YW, Curtis, MA, Gibbons, HM, Mee, EW, Bergin, PS, Teoh, HH, Connor, B, Dragunow, M Faull, RL (2008) Doublecortin expression in the normal and epileptic adult human brain. Eur J Neurosci 28, 22542265.Google Scholar
105. Hollands, C, Bartolotti, N Lazarov, O (2016) Alzheimer’s disease and hippocampal adult neurogenesis; exploring shared mechanisms. Front Neurosci 10, 178.Google Scholar
106. Regensburger, M, Prots, I Winner, B (2014) Adult hippocampal neurogenesis in Parkinson’s disease: impact on neuronal survival and plasticity. Neural Plast 2014, 454696.Google Scholar
107. Kempermann, G Kronenberg, G (2003) Depressed new neurons--adult hippocampal neurogenesis and a cellular plasticity hypothesis of major depression. Biol Psychiatry 54, 499503.Google Scholar
108. David, DJ, Wang, J, Samuels, BA, Rainer, Q, David, I, Gardier, AM Hen, R (2010) Implications of the functional integration of adult-born hippocampal neurons in anxiety-depression disorders. Neuroscientist 16, 578591.Google Scholar
109. Christian, KM, Song, H Ming, GL (2014) Functions and dysfunctions of adult hippocampal neurogenesis. Annu Rev Neurosci 37, 243262.Google Scholar
110. Kang, E, Wen, Z, Song, H, Christian, KM Ming, GL (2016) Adult neurogenesis and psychiatric disorders. Cold Spring Harb Perspect Biol 8, 9.Google Scholar
111. Yun, S, Reynolds, RP, Masiulis, I Eisch, AJ (2016) Re-evaluating the link between neuropsychiatric disorders and dysregulated adult neurogenesis. Nat Med 22, 12391247.Google Scholar
112. Sheline, YI (1996) Hippocampal atrophy in major depression: a result of depression-induced neurotoxicity? Mol Psychiatry 1, 298299.Google Scholar
113. Kitayama, N, Vaccarino, V, Kutner, M, Weiss, P Bremner, JD (2005) Magnetic resonance imaging (MRI) measurement of hippocampal volume in posttraumatic stress disorder: a meta-analysis. J Affect Disord 88, 7986.Google Scholar
114. Karl, A, Schaefer, M, Malta, LS, Dorfel, D, Rohleder, N Werner, A (2006) A meta-analysis of structural brain abnormalities in PTSD. Neurosci Biobehav Rev 30, 10041031.Google Scholar
115. Wang, Z, Neylan, TC, Mueller, SG, Lenoci, M, Truran, D, Marmar, CR, Weiner, MW Schuff, N (2010) Magnetic resonance imaging of hippocampal subfields in posttraumatic stress disorder. Arch Gen Psychiatry 67, 296303.Google Scholar
116. Goldman, MB Mitchell, CP (2004) What is the functional significance of hippocampal pathology in schizophrenia? Schizophr Bull 30, 367392.Google Scholar
117. Revest, JM, Dupret, D, Koehl, M, Funk-Reiter, C, Grosjean, N, Piazza, PV Abrous, DN (2009) Adult hippocampal neurogenesis is involved in anxiety-related behaviors. Mol Psychiatry 14, 959967.Google Scholar
118. Snyder, JS, Soumier, A, Brewer, M, Pickel, J Cameron, HA (2011) Adult hippocampal neurogenesis buffers stress responses and depressive behaviour. Nature 476, 458461.Google Scholar
119. Santarelli, L, Saxe, M, Gross, C, Surget, A, Battaglia, F, Dulawa, S, Weisstaub, N, Lee, J, Duman, R, Arancio, O, Belzung, C Hen, R (2003) Requirement of hippocampal neurogenesis for the behavioral effects of antidepressants. Science 301, 805809.Google Scholar
120. Malberg, JE (2004) Implications of adult hippocampal neurogenesis in antidepressant action. J Psychiatry Neurosci 29, 196205.Google Scholar
121. Kempermann, G, Fabel, K, Ehninger, D, Babu, H, Leal-Galicia, P, Garthe, A Wolf, SA (2010) Why and how physical activity promotes experience-induced brain plasticity. Front Neurosci 4, 189.Google Scholar
122. Erickson, KI, Voss, MW, Prakash, RS, Basak, C, Szabo, A, Chaddock, L, Kim, JS, Heo, S, Alves, H, White, SM, Wojcicki, TR, Mailey, E, Vieira, VJ, Martin, SA, Pence, BD, Woods, JA, McAuley, E Kramer, AF (2011) Exercise training increases size of hippocampus and improves memory. Proc Natl Acad Sci U S A 108, 30173022.Google Scholar
123. Boldrini, M, Underwood, MD, Hen, R, Rosoklija, GB, Dwork, AJ, John Mann, J Arango, V (2009) Antidepressants increase neural progenitor cells in the human hippocampus. Neuropsychopharmacology 34, 23762389.Google Scholar
124. Reif, A, Fritzen, S, Finger, M, Strobel, A, Lauer, M, Schmitt, A Lesch, KP (2006) Neural stem cell proliferation is decreased in schizophrenia, but not in depression. Mol Psychiatry 11, 514522.Google Scholar
125. Willner, P (2017) The chronic mild stress (CMS) model of depression: history, evaluation and usage. Neurobiol Stress 6, 7893.Google Scholar
126. Li, YF, Chen, HX, Liu, Y, Zhang, YZ, Liu, YQ Li, J (2006) Agmatine increases proliferation of cultured hippocampal progenitor cells and hippocampal neurogenesis in chronically stressed mice. Acta Pharmacol Sin 27, 13951400.Google Scholar
127. Li, B, Yamamori, H, Tatebayashi, Y, Shafit-Zagardo, B, Tanimukai, H, Chen, S, Iqbal, K Grundke-Iqbal, I (2008) Failure of neuronal maturation in Alzheimer disease dentate gyrus. J Neuropathol Exp Neurol 67, 7884.Google Scholar
128. Zhang, Z, Wang, W, Zhong, P, Liu, SJ, Long, JZ, Zhao, L, Gao, HQ, Cravatt, BF Liu, QS (2015) Blockade of 2-arachidonoylglycerol hydrolysis produces antidepressant-like effects and enhances adult hippocampal neurogenesis and synaptic plasticity. Hippocampus 25, 1626.Google Scholar
129. Schiavon, AP, Bonato, JM, Milani, H, Guimaraes, FS Weffort de Oliveira, RM (2016) Influence of single and repeated cannabidiol administration on emotional behavior and markers of cell proliferation and neurogenesis in non-stressed mice. Prog Neuropsychopharmacol Biol Psychiatry 64, 2734.Google Scholar
130. Hill, MN Gorzalka, BB (2005) Is there a role for the endocannabinoid system in the etiology and treatment of melancholic depression? Behav Pharmacol 16, 333352.Google Scholar
131. Garcia-Gutierrez, MS, Perez-Ortiz, JM, Gutierrez-Adan, A Manzanares, J (2010) Depression-resistant endophenotype in mice overexpressing cannabinoid CB(2) receptors. Br J Pharmacol 160, 17731784.Google Scholar
132. Alzheimer’s, A (2016) 2016 Alzheimer’s disease facts and figures. Alzheimers Dement 12, 459509.Google Scholar
133. Murray, RM Lewis, SW (1987) Is schizophrenia a neurodevelopmental disorder? Br Med J (Clin Res Ed) 295, 681682.Google Scholar
134. Weinberger, DR (1987) Implications of normal brain development for the pathogenesis of schizophrenia. Arch Gen Psychiatry 44, 660669.Google Scholar
135. Garey, L (2010) When cortical development goes wrong: schizophrenia as a neurodevelopmental disease of microcircuits. J Anat 217, 324333.Google Scholar
136. Wu, CS, Jew, CP Lu, HC (2011) Lasting impacts of prenatal cannabis exposure and the role of endogenous cannabinoids in the developing brain. Future Neurol 6, 459480.Google Scholar
137. Richardson, KA, Hester, AK McLemore, GL (2016) Prenatal cannabis exposure – the ‘first hit’ to the endocannabinoid system. Neurotoxicol Teratol 58, 514.Google Scholar
138. Fried, PA (2002) Conceptual issues in behavioral teratology and their application in determining long-term sequelae of prenatal marihuana exposure. J Child Psychol Psychiatry 43, 81102.Google Scholar
139. Hall, W Degenhardt, L (2009) Adverse health effects of non-medical cannabis use. Lancet 374, 13831391.Google Scholar
140. El Marroun, H, Hudziak, JJ, Tiemeier, H, Creemers, H, Steegers, EA, Jaddoe, VW, Hofman, A, Verhulst, FC, van den Brink, W Huizink, AC (2011) Intrauterine cannabis exposure leads to more aggressive behavior and attention problems in 18-month-old girls. Drug Alcohol Depend 118, 470474.Google Scholar
141. Tortoriello, G, Morris, CV, Alpar, A, Fuzik, J, Shirran, SL, Calvigioni, D, Keimpema, E, Botting, CH, Reinecke, K, Herdegen, T, Courtney, M, Hurd, YL Harkany, T (2014) Miswiring the brain: Delta9-tetrahydrocannabinol disrupts cortical development by inducing an SCG10/stathmin-2 degradation pathway. EMBO J 33, 668685.Google Scholar
142. Sun, J, Fang, YQ, Ren, H, Chen, T, Guo, JJ, Yan, J, Song, S, Zhang, LY Liao, H (2013) WIN55,212-2 protects oligodendrocyte precursor cells in stroke penumbra following permanent focal cerebral ischemia in rats. Acta Pharmacol Sin 34, 119128.Google Scholar
143. de Salas-Quiroga, A, Diaz-Alonso, J, Garcia-Rincon, D, Remmers, F, Vega, D, Gomez-Canas, M, Lutz, B, Guzman, M Galve-Roperh, I (2015) Prenatal exposure to cannabinoids evokes long-lasting functional alterations by targeting CB1 receptors on developing cortical neurons. Proc Natl Acad Sci U S A 112, 1369313698.Google Scholar
144. Barbeau, D, Liang, JJ, Robitalille, Y, Quirion, R Srivastava, LK (1995) Decreased expression of the embryonic form of the neural cell adhesion molecule in schizophrenic brains. Proc Natl Acad Sci U S A 92, 27852789.Google Scholar
145. Allen, KM, Fung, SJ Weickert, CS (2016) Cell proliferation is reduced in the hippocampus in schizophrenia. Aust N Z J Psychiatry 50, 473480.Google Scholar
146. Walton, NM, Zhou, Y, Kogan, JH, Shin, R, Webster, M, Gross, AK, Heusner, CL, Chen, Q, Miyake, S, Tajinda, K, Tamura, K, Miyakawa, T Matsumoto, M (2012) Detection of an immature dentate gyrus feature in human schizophrenia/bipolar patients. Transl Psychiatry 2, e135.Google Scholar
147. Maekawa, M, Takashima, N, Matsumata, M, Ikegami, S, Kontani, M, Hara, Y, Kawashima, H, Owada, Y, Kiso, Y, Yoshikawa, T, Inokuchi, K Osumi, N (2009) Arachidonic acid drives postnatal neurogenesis and elicits a beneficial effect on prepulse inhibition, a biological trait of psychiatric illnesses. PLoS One 4, e5085.Google Scholar
148. Chen, Q, Kogan, JH, Gross, AK, Zhou, Y, Walton, NM, Shin, R, Heusner, CL, Miyake, S, Tajinda, K, Tamura, K Matsumoto, M (2012) SREB2/GPR85, a schizophrenia risk factor, negatively regulates hippocampal adult neurogenesis and neurogenesis-dependent learning and memory. Eur J Neurosci 36, 25972608.Google Scholar
149. Iwata, Y, Suzuki, K, Wakuda, T, Seki, N, Thanseem, I, Matsuzaki, H, Mamiya, T, Ueki, T, Mikawa, S, Sasaki, T, Suda, S, Yamamoto, S, Tsuchiya, KJ, Sugihara, G, Nakamura, K, Sato, K, Takei, N, Hashimoto, K Mori, N (2008) Irradiation in adulthood as a new model of schizophrenia. PLoS One 3, e2283.Google Scholar
150. Nelson, MD, Saykin, AJ, Flashman, LA Riordan, HJ (1998) Hippocampal volume reduction in schizophrenia as assessed by magnetic resonance imaging: a meta-analytic study. Arch Gen Psychiatry 55, 433440.Google Scholar
151. Dhikav, V Anand, KS (2007) Is hippocampal atrophy a future drug target? Med Hypotheses 68, 13001306.Google Scholar
152. Steen, RG, Mull, C, McClure, R, Hamer, RM Lieberman, JA (2006) Brain volume in first-episode schizophrenia: systematic review and meta-analysis of magnetic resonance imaging studies. Br J Psychiatry 188, 510518.Google Scholar
153. Ganzola, R, Maziade, M Duchesne, S (2014) Hippocampus and amygdala volumes in children and young adults at high-risk of schizophrenia: research synthesis. Schizophr Res 156, 7686.Google Scholar
154. Newton, SS Duman, RS (2007) Neurogenic actions of atypical antipsychotic drugs and therapeutic implications. CNS Drugs 21, 715725.Google Scholar
155. Balu, DT Lucki, I (2009) Adult hippocampal neurogenesis: regulation, functional implications, and contribution to disease pathology. Neurosci Biobehav Rev 33, 232252.Google Scholar
156. Meltzer, HY Sumiyoshi, T (2008) Does stimulation of 5-HT(1A) receptors improve cognition in schizophrenia? Behav Brain Res 195, 98102.Google Scholar
157. Eisch, AJ, Barrot, M, Schad, CA, Self, DW Nestler, EJ (2000) Opiates inhibit neurogenesis in the adult rat hippocampus. Proc Natl Acad Sci U S A 97, 75797584.Google Scholar
158. Malberg, JE, Eisch, AJ, Nestler, EJ Duman, RS (2000) Chronic antidepressant treatment increases neurogenesis in adult rat hippocampus. J Neurosci 20, 91049110.Google Scholar
159. Benninghoff, J, Grunze, H, Schindler, C, Genius, J, Schloesser, RJ, van der Ven, A, Dehning, S, Wiltfang, J, Moller, HJ Rujescu, D (2013) Ziprasidone--not haloperidol--induces more de-novo neurogenesis of adult neural stem cells derived from murine hippocampus. Pharmacopsychiatry 46, 1015.Google Scholar
160. Wakade, CG, Mahadik, SP, Waller, JL Chiu, FC (2002) Atypical neuroleptics stimulate neurogenesis in adult rat brain. J Neurosci Res 69, 7279.Google Scholar
161. Halim, ND, Weickert, CS, McClintock, BW, Weinberger, DR Lipska, BK (2004) Effects of chronic haloperidol and clozapine treatment on neurogenesis in the adult rat hippocampus. Neuropsychopharmacology 29, 10631069.Google Scholar
162. Peng, Z, Zhang, R, Wang, H, Chen, Y, Xue, F, Wang, L, Yang, F, Chen, Y, Liu, L, Kuang, F Tan, Q (2013) Ziprasidone ameliorates anxiety-like behaviors in a rat model of PTSD and up-regulates neurogenesis in the hippocampus and hippocampus-derived neural stem cells. Behav Brain Res 244, 18.Google Scholar
163. Kodama, M, Fujioka, T Duman, RS (2004) Chronic olanzapine or fluoxetine administration increases cell proliferation in hippocampus and prefrontal cortex of adult rat. Biol Psychiatry 56, 570580.Google Scholar
164. Fakhoury, M (2017) Role of the endocannabinoid system in the pathophysiology of schizophrenia. Mol Neurobiol 54, 768778.Google Scholar
165. Hajos, M, Hoffmann, WE Kocsis, B (2008) Activation of cannabinoid-1 receptors disrupts sensory gating and neuronal oscillation: relevance to schizophrenia. Biol Psychiatry 63, 10751083.Google Scholar
166. Stefanis, NC, Delespaul, P, Henquet, C, Bakoula, C, Stefanis, CN Van Os, J (2004) Early adolescent cannabis exposure and positive and negative dimensions of psychosis. Addiction 99, 13331341.Google Scholar
167. Ujike, H, Takaki, M, Nakata, K, Tanaka, Y, Takeda, T, Kodama, M, Fujiwara, Y, Sakai, A Kuroda, S (2002) CNR1, central cannabinoid receptor gene, associated with susceptibility to hebephrenic schizophrenia. Mol Psychiatry 7, 515518.Google Scholar
168. De Marchi, N, De Petrocellis, L, Orlando, P, Daniele, F, Fezza, F Di Marzo, V (2003) Endocannabinoid signalling in the blood of patients with schizophrenia. Lipids Health Dis 2, 5.Google Scholar
169. Martinez-Gras, I, Hoenicka, J, Ponce, G, Rodriguez-Jimenez, R, Jimenez-Arriero, MA, Perez-Hernandez, E, Ampuero, I, Ramos-Atance, JA, Palomo, T Rubio, G (2006) (AAT)n repeat in the cannabinoid receptor gene, CNR1: association with schizophrenia in a Spanish population. Eur Arch Psychiatry Clin Neurosci 256, 437441.Google Scholar
170. Bortolato, M, Bini, V, Frau, R, Devoto, P, Pardu, A, Fan, Y Solbrig, MV (2014) Juvenile cannabinoid treatment induces frontostriatal gliogenesis in Lewis rats. Eur Neuropsychopharmacol 24, 974985.Google Scholar
171. Martinez-Orgado, J, Fernandez-Lopez, D, Lizasoain, I Romero, J (2007) The seek of neuroprotection: introducing cannabinoids. Recent Pat CNS Drug Discov 2, 131139.Google Scholar
172. Ritz, K, van Buchem, MA Daemen, MJ (2013) The heart-brain connection: mechanistic insights and models. Neth Heart J 21, 5557.Google Scholar
173. Dirnagl, U (2012) Pathobiology of injury after stroke: the neurovascular unit and beyond. Ann N Y Acad Sci 1268, 2125.Google Scholar
174. Arvidsson, A, Collin, T, Kirik, D, Kokaia, Z Lindvall, O (2002) Neuronal replacement from endogenous precursors in the adult brain after stroke. Nat Med 8, 963970.Google Scholar
175. Heurteaux, C, Widmann, C, Moha ou Maati, H, Quintard, H, Gandin, C, Borsotto, M, Veyssiere, J, Onteniente, B Lazdunski, M (2013) NeuroAiD: properties for neuroprotection and neurorepair. Cerebrovasc Dis 35(Suppl. 1):17.Google Scholar
176. Wiltrout, C, Lang, B, Yan, Y, Dempsey, RJ Vemuganti, R (2007) Repairing brain after stroke: a review on post-ischemic neurogenesis. Neurochem Int 50, 10281041.Google Scholar
177. Pu, H, Jiang, X, Hu, X, Xia, J, Hong, D, Zhang, W, Gao, Y, Chen, J Shi, Y (2016) Delayed docosahexaenoic acid treatment combined with dietary supplementation of omega-3 fatty acids promotes long-term neurovascular restoration after ischemic stroke. Transl Stroke Res 7, 521534.Google Scholar
178. Mori, MA, Meyer, E, Soares, LM, Milani, H, Guimaraes, FS de Oliveira, RM (2017) Cannabidiol reduces neuroinflammation and promotes neuroplasticity and functional recovery after brain ischemia. Prog Neuropsychopharmacol Biol Psychiatry 75, 94105.Google Scholar
179. Han, H, Wu, LM, Han, MX, Yang, WM, Wang, YX Fang, ZH (2016) Diabetes impairs spatial learning and memory and hippocampal neurogenesis via BDNF in rats with transient global ischemia. Brain Res Bull 124, 269277.Google Scholar
180. Nagayama, T, Sinor, AD, Simon, RP, Chen, J, Graham, SH, Jin, K Greenberg, DA (1999) Cannabinoids and neuroprotection in global and focal cerebral ischemia and in neuronal cultures. J Neurosci 19, 29872995.Google Scholar
181. Fernandez-Lopez, D, Pradillo, JM, Garcia-Yebenes, I, Martinez-Orgado, JA, Moro, MA Lizasoain, I (2010) The cannabinoid WIN55212-2 promotes neural repair after neonatal hypoxia-ischemia. Stroke 41, 29562964.Google Scholar
182. Caltana, L, Saez, TM, Aronne, MP Brusco, A (2015) Cannabinoid receptor type 1 agonist ACEA improves motor recovery and protects neurons in ischemic stroke in mice. J Neurochem 135, 616629.Google Scholar
183. Fernandez-Lopez, D, Martinez-Orgado, J, Nunez, E, Romero, J, Lorenzo, P, Moro, MA Lizasoain, I (2006) Characterization of the neuroprotective effect of the cannabinoid agonist WIN-55212 in an in vitro model of hypoxic-ischemic brain damage in newborn rats. Pediatr Res 60, 169173.Google Scholar
184. Bravo-Ferrer, I, Cuartero, MI, Zarruk, JG, Pradillo, JM, Hurtado, O, Romera, VG, Diaz-Alonso, J, Garcia-Segura, JM, Guzman, M, Lizasoain, I, Galve-Roperh, I Moro, MA (2017) Cannabinoid type-2 receptor drives neurogenesis and improves functional outcome after stroke. Stroke 48, 204212.Google Scholar
185. Arevalo-Martin, A, Garcia-Ovejero, D, Gomez, O, Rubio-Araiz, A, Navarro-Galve, B, Guaza, C, Molina-Holgado, E Molina-Holgado, F (2008) CB2 cannabinoid receptors as an emerging target for demyelinating diseases: from neuroimmune interactions to cell replacement strategies. Br J Pharmacol 153, 216225.Google Scholar
186. Gale, SA, Acar, D Daffner, KR (2018) Dementia. Am J Med. Epub ahead of print.Google Scholar
187. Gotz, J Ittner, LM (2008) Animal models of Alzheimer’s disease and frontotemporal dementia. Nat Rev Neurosci 9, 532544.Google Scholar
188. Dorszewska, J, Prendecki, M, Oczkowska, A, Dezor, M Kozubski, W (2016) Molecular basis of familial and sporadic Alzheimer’s disease. Curr Alzheimer Res 13, 952963.Google Scholar
189. Sperling, R, Mormino, E Johnson, K (2014) The evolution of preclinical Alzheimer’s disease: implications for prevention trials. Neuron 84, 608622.Google Scholar
190. Dubois, B, Hampel, H, Feldman, HH, Scheltens, P, Aisen, P, Andrieu, S, Bakardjian, H, Benali, H, Bertram, L, Blennow, K, Broich, K, Cavedo, E, Crutch, S, Dartigues, JF, Duyckaerts, C, Epelbaum, S, Frisoni, GB, Gauthier, S, Genthon, R, Gouw, AA, Habert, MO, Holtzman, DM, Kivipelto, M, Lista, S, Molinuevo, JL, O’Bryant, SE, Rabinovici, GD, Rowe, C, Salloway, S, Schneider, LS, Sperling, R, Teichmann, M, Carrillo, MC, Cummings, J Jack, CR Jr., Proceedings of the Meeting of the International Working G, the American Alzheimer’s Association on “The Preclinical State of AD”, July, & Washington DC USA (2016) Preclinical Alzheimer’s disease: definition, natural history, and diagnostic criteria. Alzheimers Dement 12, 292323.Google Scholar
191. Mi, K Johnson, GV (2006) The role of tau phosphorylation in the pathogenesis of Alzheimer’s disease. Curr Alzheimer Res 3, 449463.Google Scholar
192. Schliebs, R Arendt, T (2006) The significance of the cholinergic system in the brain during aging and in Alzheimer’s disease. J Neural Transm (Vienna) 113, 16251644.Google Scholar
193. Mann, DM (1996) Pyramidal nerve cell loss in Alzheimer’s disease. Neurodegeneration 5, 423427.Google Scholar
194. Raskin, J, Cummings, J, Hardy, J, Schuh, K Dean, RA (2015) Neurobiology of Alzheimer’s disease: integrated molecular, physiological, anatomical, biomarker, and cognitive dimensions. Curr Alzheimer Res 12, 712722.Google Scholar
195. Radad, K, Moldzio, R, Al-Shraim, M, Kranner, B, Krewenka, C Rausch, WD (2017) Recent advances on the role of neurogenesis in the adult brain: therapeutic potential in Parkinson’s and Alzheimer’s diseases. CNS Neurol Disord Drug Targets 16, 740748.Google Scholar
196. Crews, L, Adame, A, Patrick, C, Delaney, A, Pham, E, Rockenstein, E, Hansen, L Masliah, E (2010) Increased BMP6 levels in the brains of Alzheimer’s disease patients and APP transgenic mice are accompanied by impaired neurogenesis. J Neurosci 30, 1225212262.Google Scholar
197. Jin, K, Peel, AL, Mao, XO, Xie, L, Cottrell, BA, Henshall, DC Greenberg, DA (2004) Increased hippocampal neurogenesis in Alzheimer’s disease. Proc Natl Acad Sci U S A 101, 343347.Google Scholar
198. Perry, EK, Johnson, M, Ekonomou, A, Perry, RH, Ballard, C Attems, J (2012) Neurogenic abnormalities in Alzheimer’s disease differ between stages of neurogenesis and are partly related to cholinergic pathology. Neurobiol Dis 47, 155162.Google Scholar
199. Boekhoorn, K, Joels, M Lucassen, PJ (2006) Increased proliferation reflects glial and vascular-associated changes, but not neurogenesis in the presenile Alzheimer hippocampus. Neurobiol Dis 24, 114.Google Scholar
200. Bilkei-Gorzo, A (2014) Genetic mouse models of brain ageing and Alzheimer’s disease. Pharmacol Ther 142, 244257.Google Scholar
201. Marlatt, MW Lucassen, PJ (2010) Neurogenesis and Alzheimer’s disease: biology and pathophysiology in mice and men. Curr Alzheimer Res 7, 113125.Google Scholar
202. Kalifa, S, Polston, EK, Allard, JS Manaye, KF (2011) Distribution patterns of cannabinoid CB1 receptors in the hippocampus of APPswe/PS1DeltaE9 double transgenic mice. Brain Res 1376, 94100.Google Scholar
203. Aso, E, Palomer, E, Juves, S, Maldonado, R, Munoz, FJ Ferrer, I (2012) CB1 agonist ACEA protects neurons and reduces the cognitive impairment of AbetaPP/PS1 mice. J Alzheimers Dis 30, 439459.Google Scholar
204. Marchalant, Y, Baranger, K, Wenk, GL, Khrestchatisky, M Rivera, S (2012) Can the benefits of cannabinoid receptor stimulation on neuroinflammation, neurogenesis and memory during normal aging be useful in AD prevention? J Neuroinflammation 9, 10.Google Scholar
205. Ramirez, BG, Blazquez, C, Gomez del Pulgar, T, Guzman, M de Ceballos, ML (2005) Prevention of Alzheimer’s disease pathology by cannabinoids: neuroprotection mediated by blockade of microglial activation. J Neurosci 25, 19041913.Google Scholar
206. Westlake, TM, Howlett, AC, Bonner, TI, Matsuda, LA Herkenham, M (1994) Cannabinoid receptor binding and messenger RNA expression in human brain: an in vitro receptor autoradiography and in situ hybridization histochemistry study of normal aged and Alzheimer’s brains. Neuroscience 63, 637652.Google Scholar
207. Benito, C, Nunez, E, Tolon, RM, Carrier, EJ, Rabano, A, Hillard, CJ Romero, J (2003) Cannabinoid CB2 receptors and fatty acid amide hydrolase are selectively overexpressed in neuritic plaque-associated glia in Alzheimer’s disease brains. J Neurosci 23, 1113611141.Google Scholar
208. Lee, JH, Agacinski, G, Williams, JH, Wilcock, GK, Esiri, MM, Francis, PT, Wong, PT, Chen, CP Lai, MK (2010) Intact cannabinoid CB1 receptors in the Alzheimer’s disease cortex. Neurochem Int 57, 985989.Google Scholar
209. Mulder, J, Zilberter, M, Pasquare, SJ, Alpar, A, Schulte, G, Ferreira, SG, Kofalvi, A, Martin-Moreno, AM, Keimpema, E, Tanila, H, Watanabe, M, Mackie, K, Hortobagyi, T, de Ceballos, ML Harkany, T (2011) Molecular reorganization of endocannabinoid signalling in Alzheimer’s disease. Brain 134, 10411060.Google Scholar
210. Ahmad, R, Goffin, K, Van den Stock, J, De Winter, FL, Cleeren, E, Bormans, G, Tournoy, J, Persoons, P, Van Laere, K Vandenbulcke, M (2014) In vivo type 1 cannabinoid receptor availability in Alzheimer’s disease. Eur Neuropsychopharmacol 24, 242250.Google Scholar
211. Bedse, G, Romano, A, Cianci, S, Lavecchia, AM, Lorenzo, P, Elphick, MR, Laferla, FM, Vendemiale, G, Grillo, C, Altieri, F, Cassano, T Gaetani, S (2014) Altered expression of the CB1 cannabinoid receptor in the triple transgenic mouse model of Alzheimer’s disease. J Alzheimers Dis 40, 701712.Google Scholar
212. Maccarrone, M, Totaro, A, Leuti, A, Giacovazzo, G, Scipioni, L, Mango, D, Coccurello, R, Nistico, R Oddi, S (2018) Early alteration of distribution and activity of hippocampal type-1 cannabinoid receptor in Alzheimer’s disease-like mice overexpressing the human mutant amyloid precursor protein. Pharmacol Res. Epub ahead of print.Google Scholar
213. Solas, M, Francis, PT, Franco, R Ramirez, MJ (2013) CB2 receptor and amyloid pathology in frontal cortex of Alzheimer’s disease patients. Neurobiol Aging 34, 805808.Google Scholar
214. Savonenko, AV, Melnikova, T, Wang, Y, Ravert, H, Gao, Y, Koppel, J, Lee, D, Pletnikova, O, Cho, E, Sayyida, N, Hiatt, A, Troncoso, J, Davies, P, Dannals, RF, Pomper, MG Horti, AG (2015) Cannabinoid CB2 receptors in a mouse model of Abeta amyloidosis: immunohistochemical analysis and suitability as a PET biomarker of neuroinflammation. PLoS One 10, e0129618.Google Scholar
215. Aso, E, Andres-Benito, P Ferrer, I (2016) Delineating the efficacy of a Cannabis-based medicine at advanced stages of dementia in a murine model. J Alzheimers Dis 54, 903912.Google Scholar
216. Watt, G Karl, T (2017) In vivo evidence for therapeutic properties of cannabidiol (CBD) for Alzheimer’s disease. Front Pharmacol 8, 20.Google Scholar
217. Aso, E, Juves, S, Maldonado, R Ferrer, I (2013) CB2 cannabinoid receptor agonist ameliorates Alzheimer-like phenotype in AbetaPP/PS1 mice. J Alzheimers Dis 35, 847858.Google Scholar
218. Martin-Moreno, AM, Reigada, D, Ramirez, BG, Mechoulam, R, Innamorato, N, Cuadrado, A de Ceballos, ML (2011) Cannabidiol and other cannabinoids reduce microglial activation in vitro and in vivo: relevance to Alzheimer’s disease. Mol Pharmacol 79, 964973.Google Scholar
219. Esposito, G, Scuderi, C, Valenza, M, Togna, GI, Latina, V, De Filippis, D, Cipriano, M, Carratu, MR, Iuvone, T Steardo, L (2011) Cannabidiol reduces Abeta-induced neuroinflammation and promotes hippocampal neurogenesis through PPARgamma involvement. PLoS One 6, e28668.Google Scholar
220. Chen, B, Bromley-Brits, K, He, G, Cai, F, Zhang, X Song, W (2010) Effect of synthetic cannabinoid HU210 on memory deficits and neuropathology in Alzheimer’s disease mouse model. Curr Alzheimer Res 7, 255261.Google Scholar
221. Mandyam, CD Koob, GF (2012) The addicted brain craves new neurons: putative role for adult-born progenitors in promoting recovery. Trends Neurosci 35, 250260.Google Scholar
Figure 0

Fig. 1 Schematic representation of the steps in embryonic or adult neurogenesis in the central nervous system. Neural stem cells, neuronal progenitors and glial progenitors may undergo symmetric or asymmetric divisions. Symmetrical divisions produce two ‘daughters’ that are identical to their precursors and each other. Asymmetrical divisions produce two different ‘daughters’, one that is identical to their precursors and another ‘daughter’ that is different from the ‘sister’ and the precursor. Symmetrical divisions expand the pool of precursors (proliferation step) more rapidly than the asymmetrical divisions. However, asymmetrical divisions give rise to cells with a new phenotype (differentiation step). Therefore, neural stem cells may differentiate into progenitors committed to neuronal or glial phenotypes. Neuronal progenitors may differentiate into neuroblasts, whereas glial progenitors may differentiate into different types of glioblasts. Progenitors also may become quiescent( non-dividing state). Neuroblasts and glioblasts maintain their self-renewing capacity until maturation. Cell death may occur at any step of the process. For a review and more detailed description of neurogenic steps, we suggest the studies by Paridaen and Huettner (20) (for embryonic neurogenesis) and Bond et al. (21) (for adult neurogenesis).

Figure 1

Fig. 2 Classical representation of endocannabinoid signalling in the adult brain. Anandamide (AEA) and 2-arachidonoyl glycerol (2-AG) are produced ‘on demand’ in calcium (Ca2+)-dependent manner (via the previous activation of a metabotropic or ionotropic receptor). After the synthesis of endocannabinoids by specialised enzymes, they act as retrograde massagers by activating CB1 receptors located at pre-synaptic terminals. CB1 is a Gi/o-coupled receptor, and its activation reduces Ca2+ currents and increases K+ currents, leading to the inhibition of neurotransmitter release. The actions of 2-AG and AEA are terminated by enzymatic hydrolysis; fatty acid amide hydrolase (FAAH) and monoacylglycerol lipase (MAGL) degrade AEA and 2-AG, respectively. The CB1 receptor is also expressed in astrocytes and microglia and the CB2 receptor is expressed in activated microglia and putatively expressed in neurons (still under debate). CB1, type 1 cannabinoid receptor; CB2, type 2 cannabinoid receptor; DAGL, diacylglycerol lipase; NAPE-PLD, n-acyl phosphatidylethanolamine-specific phospholipase D.

Figure 2

Fig. 3 Schematic representation of the neurogenesis steps in the central nervous system of embryos (a) and adults (b), along with the putative expression of the endocannabinoid system in different cell populations. 2-AG, 2-arachidonoylglycerol; AEA, anandamide; CB1, type 1 cannabinoid receptor; CB2, type 2 cannabinoid receptor; DAGL, diacylglycerol lipase; FAAH, fatty acid amide hydrolase; MAGL, monoacylglycerol lipase; NAPE-PLD, n-acyl phosphatidylethanolamine-specific phospholipase D.

Figure 3

Table 1 Cannabinoids increase adult neurogenesis in animal models of psychiatric conditions

Figure 4

Table 2 Cannabinoids agonists increase adult neurogenesis in animal models of brain ischaemia and Alzheimer’s disease

Supplementary material: File

de Oliveira et al. supplementary material

de Oliveira et al. supplementary material 1

Download de Oliveira et al. supplementary material(File)
File 26.3 KB