Canna~Fangled Abstracts

G-Protein-Coupled Receptors in Adult Neurogenesis

By July 5, 2012No Comments
pm8
G-Protein-Coupled Receptors in Adult Neurogenesis
David R. Sibley, ASSOCIATE EDITOR
This article has been cited by other articles in PMC.

Abstract

The importance of adult neurogenesis has only recently been accepted, resulting in a completely new field of investigation within stem cell biology. The regulation and functional significance of adult neurogenesis is currently an area of highly active research. G-protein-coupled receptors (GPCRs) have emerged as potential modulators of adult neurogenesis. GPCRs represent a class of proteins with significant clinical importance, because approximately 30% of all modern therapeutic treatments target these receptors. GPCRs bind to a large class of neurotransmitters and neuromodulators such as norepinephrine, dopamine, and serotonin. Besides their typical role in cellular communication, GPCRs are expressed on adult neural stem cells and their progenitors that relay specific signals to regulate the neurogenic process. This review summarizes the field of adult neurogenesis and its methods and specifies the roles of various GPCRs and their signal transduction pathways that are involved in the regulation of adult neural stem cells and their progenitors. Current evidence supporting adult neurogenesis as a model for self-repair in neuropathologic conditions, adult neural stem cell therapeutic strategies, and potential avenues for GPCR-based therapeutics are also discussed.

I. Introduction

Only a few decades ago, scientists thought that certain cells in the body, such as cardiac myocytes and brain cells, were nonrenewable. We now know that these cells can be regenerated through specific processes involving stem cells that exist throughout life. The first evidence of adult neurogenesis was reported in the 1960s by Joseph Altman, who showed that neurons in adult rats incorporated [3H]thymidine (Altman, 1962). However, it was not until the 1990s that the idea of adult neurogenesis became widely accepted, when it was shown that the subventricular zone (SVZ1) of the lateral ventricles (Reynolds and Weiss, 1992Richards et al., 1992) and the subgranular zone (SGZ) of the hippocampal dentate gyrus (Gage et al., 1995Palmer et al., 1997) contain self-renewing neural stem cells (NSCs) that give rise to new neural cells. The existence of adult neurogenesis in humans was confirmed in 1998 (Eriksson et al., 1998).

G-protein-coupled receptors (GPCRs) are the largest family of membrane receptors in eukaryotes. Although the exact number of GPCRs is unknown, nearly a thousand genes encoding for GPCRs have been identified in the human genome (Takeda et al., 2002), of which approximately half are receptors for endogenous ligands. Also called heptahelical receptors, GPCRs are integral membrane proteins composed of an extracellular N terminus, seven transmembrane α-helices connected by intracellular and extracellular loops, and an intracellular C terminus. When activated, GPCRs transduce signals from outside the cell to intracellular pathways, resulting in cellular responses.

GPCRs affect the transduction of signals through heterotrimeric G-proteins, which exist bound to the inner side of the cytoplasmic membrane. G-proteins consist of three subunits, α, β, and γ, that are altered by activated GPCRs. When a ligand binds the GPCR on the cell’s outside surface, it drives a conformational change, thus activating the receptor. The activated receptor then functions as a guanine-nucleotide exchange factor, exchanging GDP for GTP on the Gα subunit of the G-protein. Subsequently, the Gα-GTP subunit dissociates from the Gβγ dimer and the GPCR. Both the GTP-bound Gα and free Gβγ subunits can induce different intracellular signaling cascades and/or downstream effector proteins (e.g., adenylyl cyclases, phospholipase C, various ion channels). Because the Gα subunit possesses intrinsic enzymatic GTPase activity, it eventually hydrolyzes the GTP back to GDP, allowing Gα to reassemble with the Gβγ subunit and GPCR, returning the GPCR and G-protein to their original states. The activity of the Gα subunit is modulated by other proteins, such as the regulators of G protein signaling proteins, a type of GTPase-activating protein that accelerates GTP hydrolysis, thereby reducing the signaling (Sjögren et al., 2010). In addition, GPCRs can transduce signals without G protein involvement through G protein-independent signaling (noncanonical) pathways (Wei et al., 2003;Shenoy et al., 2006).

GPCRs are essential in the processes of neurotransmission, cell proliferation, and organ-specific function (Luttrell, 2008). Not surprisingly, GPCRs are important drug targets with at least 30% of all modern therapeutics acting at these receptors (Overington et al., 2006Lagerström and Schiöth, 2008). The GPCR neurotransmitter systems involved in adult neurogenesis are discussed in this review. These encompass those primarily considered neuromodulators such as norepinephrine (NE), dopamine, and serotonin. Neuromodulators regulate long-range paracrine or nonsynaptic signaling through neuronal projections into the SVZ and SGZ, the two major neurogenic areas of the adult mammalian brain. Therefore, it is not surprising that the GPCRs are involved in the regulation of NSCs and their progenitors. Furthermore, increasing evidence points to the involvement of other GPCR ligands in adult neurogenesis, such as chemokines, peptide hormones, endogenous opioids, and Wnt proteins, to name a few. In this review, the general features of NSCs, methods for studying adult neurogenesis, and role of the brain vascular niche and choroid plexus in adult neurogenesis are summarized, followed by a comprehensive examination of the GPCR systems involved in modulating adult neurogenesis (Table 1). It will conclude with discussions on adult neurogenesis in pathological conditions, use of NSC therapy in the central nervous system (CNS), and the therapeutic potential of GPCR-based NSC strategies.

TABLE 1

Role of GPCRs in adult neurogenesis

II. General Features of Neural Stem Cells and Progenitors

A. Characteristics of Neural Stem Cells in the Subventricular Zone

NSCs are defined by their ability to self-replicate and differentiate into multiple cell types found in the CNS, including neurons, astrocytes, and oligodendrocytes (Gage, 2000). Neural progenitor cells are cells that do not fully meet all of the attributes of an NSC such as neuronal pluripotency and self-renewal (Potten and Loeffler, 1990). The identity of resident adult NSCs is debated. It has been proposed that NSCs are subependymal cells (Morshead et al., 1994), astrocytes (Doetsch et al., 1999), or multiciliated ependymal cells (Johansson et al., 1999). The astrocyte theory is the most accepted and is illustrated in Fig. 1. NSCs are located in the SVZ and are represented by a subset of slowly dividing radial-like astrocytic cells called type B cells that stain positive for glial fibrillary acidic protein (GFAP). The type B can proliferate to generate rapidly dividing cells called transient amplifying progenitors (TAP or type C cells). TAP cells lose the GFAP marker associated with type B cells, but gain the Dlx2+ epitope. These TAP cells then generate migrating neuroblasts (type A cells), distinguished by the neuronal marker observed in chain-migrating neurons called polysialylated neural cell adhesion molecule (NCAM) (Bonfanti and Theodosis, 1994). This marker is important in chain migration as NCAM-deficient mice have significantly smaller olfactory bulbs (Tomasiewicz et al., 1993Cremer et al., 1994).

Fig. 1.

Adult SVZ neurogenesis. B cells (purple) are astrocyte-like cells that serve as the SVZ stem cell (self-renews) and contact the ventricle lumen. C cells (light green) are rapidly dividing TAP cells derived through division from the B cells. C cells divide 

Molecular markers for both stem cells and their progenitors, however, are not sufficiently specific to positively identify these cell types. Common markers for NSCs and NPCs such as nestin, bromodeoxyuridine (BrdU), and doublecortin (DCX) can also be detected in other cell types found in the brain such as reactive astrocytes or cancer cells (Lendahl et al., 1990Clarke et al., 1994Kaneko et al., 2000Komitova and Eriksson, 2004). Although there is no single marker for stem and/or progenitor cells, the colocalization of several markers may be used to identify (and separate) these cells. The use of several markers simultaneously with state-of-the-art flow cytometry is presently the “gold standard” in the identification and separation of cells. However, flow cytometry cannot isolate large quantities of cells for potential therapeutic uses (Preffer and Dombkowski, 2009Tárnok et al., 2010).

The SVZ is one of only two discrete regions in the adult CNS capable of active neurogenesis. The neurons in the SVZ are continuously generated and migrate as tangentially oriented chains (Doetsch and Alvarez-Buylla, 2006) to the olfactory bulb from all rostrocaudal locations of the SVZ through a region called the rostral migratory stream (Fig. 1). The oligodendrocytes migrate laterally and dorsally into the corpus callosum, the fornix, and the fiber tracks of the striatum (Menn et al., 2006). This chain-like migration occurs in rodents and various primates, but not humans (Sanai et al., 2004). Differences between species could be due to the longer distance that separates the olfactory bulbs from the cerebrum in humans (Sanai et al., 2004) or because rodents generate more olfactory bulb neurons out of survival. Rodents produce thousands of olfactory bulb neurons every day (Alvarez-Buylla, 1997). Once these newly formed neurons reach the olfactory bulb, they migrate radially to the outer cell layers and differentiate into GABAergic interneurons (Alvarez-Buylla, 1997). In rodents, it is estimated that the time for neural progenitors in the SVZ to migrate to the olfactory bulb and differentiate is approximately 15 days (Lois and Alvarez-Buylla, 1994).

B. Characteristics of Neural Stem Cells in the Subgranular Zone

Because the hippocampus is a center of memory and learning, adult neurogenesis in the SGZ has huge potential to affect memory functions. In a distinct but parallel manner, quiescent NSCs in the SGZ are a population of radial astrocytic cells containing GFAP, Sox2, and Nestin markers (type 1 cells) (Suh et al., 2007). Sex-determining region Y-box 2 (Sox2) is a transcription factor that maintains undifferentiated state of NSCs. Nestin is an intermediate filament protein that is expressed in NSCs considered in the early stages of development. These NSCs are located within the SGZ, with their radial processes projecting through the granular cell layer. Shorter tangential processes extend along the border of the hilus and granule cell layer (Ming and Song, 2005) (Fig. 2). Similar to SVZ cells, SGZ NSCs generate actively dividing nonradial transient amplifying progenitors (type 2 cells) that lose the GFAP marker (Sox2+, Nestin+, GFAP−). Type 2 cells generate neuroblasts that are DCX-positive and differentiate and migrate locally into the glutamatergic dentate granule cell layer (Mu et al., 2010) (Fig. 2), a process that takes approximately 4 weeks (Cameron et al., 1993). New granule cells in the dentate gyrus are continuously generated locally in all mammals examined so far, including humans (Ming and Song, 2005), and survive for an extended period of at least 2 years (Eriksson et al., 1998). The adult SVZ generates the largest pool of dividing progenitor cells. In rats, 20% of the adult SVZ cell population is proliferating (Sajad et al., 2011). In the mouse SGZ, approximately 0.01% of the granule cell population is dividing (Kempermann et al., 1997Cameron and McKay, 2001), whereas in monkeys, it is only 0.004% (Kornack and Rakic, 1999). Therefore, adult neurogenesis is considered a rare event. Adult neurogenesis has also been reported in other areas of the brain such as the neocortex and striatum, but at much lower levels (Gould et al., 1999Gritti et al., 1999Rietze et al., 2000Bédard et al., 2006).

Fig. 2.

Adult SGZ neurogenesis. Astrocytes (AS; dark purple) self-renew and also divide to give rise to progenitors (D cells; blue), which differentiate into new granule cells (G cells; orange). These newly born granule cells integrate into the granule cell layer 

III. Adult versus Embryonic Neurogenesis

Both embryonic and adult neural stem cells are defined by their ability to self-renew, to be relatively quiescent, and to differentiate into multiple cell types. However, their cellular properties are different. Embryonic stem cells are isolated from the blastocyst stage of an embryo (Martin, 1981) and have the potential to generate the three germ layers in vitro and in vivo (Keller et al., 1993Kennedy and Keller, 2003). Adult NSCs have a much more restricted potential and differentiate only into neurons and glia.

During development, embryonic NSCs are derived from radial glia and neuroepithelial cells lining the neural tubes. In a temporally defined sequence, subsets of neurons are generated first, followed by astrocytes and then oligodendrocytes. The sequential maturation of the different types of neurons in the developing CNS is only recently becoming uncovered (Okano and Temple, 2009). However, cortical neurogenesis is essentially completed during the embryonic period, and gliogenesis occurs largely in the first month of postnatal life (Okano and Temple, 2009). The timing of this switch from neurogenesis to gliogenesis involves both the down-regulation of neurogenic genes and the activation of proglial genes, such as the nuclear factors 1A and 1B (Deneen et al., 2006), through transcription factors (Gritti et al., 1999), epigenetic modification (Shimozaki et al., 2005), Notch signaling (Sanosaka et al., 2009), and miRNAs (Liu and Zhao, 2009).

Adult NSCs originate from the embryonic neuroepithelial radial glia, with a subset of these cells persisting in the adult in specialized regions in the brain such as the SVZ and SGZ (Doetsch, 2003). An important regulator of the temporal progression of adult neurogenesis has been found to involve miRNA-124 (Cheng et al., 2009), which first appears between the transition from TAP into neuroblasts. The number of different types of neurons generated when adult NSCs are transplanted into embryonic brains is much less compared with embryonic NSCs (Temple, 2001).

The fundamental question of why adult NSCs are more restricted in their neurogenic potential than embryonic NSCs may be due to their microenvironment, termed the stem cell “niche.” Adult SVZ cells lie in a vascular recess, a highly specialized microdomain of extracellular matrix and circulating small molecules that provide spatial and regulatory clues (Shen et al., 2008Tavazoie et al., 2008). In addition, the ability and frequency of adult NSCs to self-renew deteriorates with age (Kuhn et al., 1996;Ahn and Joyner, 2005) and is associated with a decline of miRNA let-7b (Nishino et al., 2008).

IV. Methods for Analyzing Adult Neurogenesis

A. Bromodeoxyuridine Labeling

BrdU labeling in conjunction with the detection of various cell types and cell cycle markers is the main method of studying adult neurogenesis today (Miller and Nowakowski, 1988), replacing [3H]thymidine, which was used previously. The major advantage of BrdU is that it does not require any previous information from the cells to be labeled and can be used across species and cell types. As a thymidine analog, it is incorporated into DNA during DNA synthesis. However, there are limitations when using BrdU to study neurogenesis. One major drawback is that BrdU labels both cell division (S-phase) and DNA synthesis associated with DNA repair, gene duplication without cell division, and apoptosis (Taupin, 2007). In addition, BrdU is toxic and has been found to negatively affect the proliferation of neural progenitors in vitro (Ross et al., 2008). Using lower doses of BrdU (50 mg/kg) minimizes labeling of non–S-phase DNA synthesis and prevents BrdU toxicity (Cooper-Kuhn and Kuhn, 2002). Furthermore, detection of adult neurogenesis requires not only BrdU colocalization with neuronal markers but also proof of the absence of apoptosis. Detailed methods for analyzing neural stem and progenitor cells via BrdU are found in Kelly et al. (2009).

B. Genetic Marking with Viruses

An alternative to BrdU for assigning proliferative properties of neurogenesis is the expression of transgenes from retroviruses. The Maloney murine leukemia virus lacks nuclear import mechanisms, so viral integration and expression of DNA can occur only during mitosis. A recombinant form of the virus containing green fluorescent protein (GFP) is used. Upon viral integration, GFP is expressed, indicating cell division (Lewis and Emerman, 1994). The use of a live reporter makes retroviruses ideal for transducing a mixed culture of stem cells or targeting live cells both in vitro and in vivo, allowing direct visualization and analysis of live, recently divided cells (Stitelman et al., 2010).

Lentiviruses have also been used for the detection of dividing cells. Although retroviruses only infect cells that are actively replicating, lentiviruses will infect all cells. Lentivirus-mediated expression, besides being used to study the properties of stem cells and progenitors, has the potential to deliver therapy-based approaches in the CNS (Kouroupi et al., 2010) and to reprogram adult NSCs to non-neuronal cell lineages (Forsberg et al., 2010). Efficient suppression of transgene expression has been recently achieved using a microRNA-regulated lentiviral reporter system (Sachdeva et al., 2010). Detailed methods for either viral system are found in Peltier and Schaffer (2010).

Adenoassociated viruses have the ability to infect specific cell types based on the virus serotype capsid and act as a vector, inserting genes of interest into target cells. Adenoassociated viruses have been used to enhance or knockout expression of various regulators of neurogenesis such as brain-derived neurotrophic factor (Henry et al., 2007), cyclin-dependent kinase 5 (Lagace et al., 2008), and interleukin-4 (Kiyota et al., 2010). This DNA virus normally expresses genes transiently and is very dependent upon the injection site and serotype. Furthermore, the insert size is limited (Landgren and Curtis, 2011), and the integration frequency may be different between mitotic and postmitotic cells (Han et al., 2008).

C. Neurospheres

Neurosphere cultures have been widely used to study neural stem/progenitor cells (Gottlieb, 2002;Singec and Quinones-Hinojosa, 2008). Neurospheres refer to the sphere-forming assay used to identify stem cells based upon their functional capacity to self-renew and express pluripotency in vitro (Reynolds and Weiss, 1992). Stem cells are cultured under nonadherent conditions to form free-floating aggregates or spheres (Rietze and Reynolds, 2006). Because SGZ tissue results in more adherent cells in culture, this assay is almost exclusively used with SVZ tissue. The tissue is dissected under a microscope, then enzymatically dissociated into single cells and grown in a serum-free medium while in the presence of epidermal growth factor (EGF) (Reynolds and Weiss, 1992). A small population of cells begins to divide, initially adhering to the plate. As the sphere grows, it detaches and becomes free-floating. The majority of the cells at this point express the intermediate filament nestin, which is present in neuroepithelial cells. To demonstrate self-renewal properties, neurospheres are mechanically dissociated and recultured at the single-cell level (to give rise to clones; i.e., clonal) in the presence of EGF, a smaller subset of these cells reforming the neurospheres (referred to as secondary neurospheres). The number of single cells that reform neurospheres is a semiquantitative way to determine the number of stem cells in vivo (Pastrana et al., 2011). To demonstrate pluripotency, each neurosphere is mechanically dissociated and plated on an adherent substrate without EGF or other growth factors, initiating the differentiation process. Immunocytochemical detection of neuronal and glial markers determines whether all three cell types (neurons, astrocytes, oligodendrocytes) are present. Neurospheres can be cultured from areas outside of the SVZ, such as in the spinal cord (Vescovi et al., 1993Weiss et al., 1996), but the addition of fibroblast growth factor (FGF) in addition to EGF is required.

Clonal neurospheres are not homogeneous but are composed of a population of stem cells, progenitors, and differentiated cells (Parker et al., 2005). When generating clonal neurospheres by plating single cells, neurospheres were observed to frequently fuse (Mori et al., 2006Singec et al., 2006). In addition, neurosphere formation is predisposed to cells that are poised for proliferation or are actively dividing, so it is likely that it does not detect quiescent stem cells or their intrinsic properties (Pastrana et al., 2009). Cells other than true stem cells can also give rise to neurospheres (Stingl, 2009). These limitations indicate that the neurosphere assay alone cannot be used to define stem cells in vivo. Detailed methods and video on the neurosphere assay can be found at Azari et al. (2010).

D. Monolayer Cultures

Adherent monolayer culture is another method commonly used to study neural stem/progenitor cells (Gottlieb, 2002Ray, 2008). This method is based on the same underlying principles and uses methods similar to those used in the neurosphere culture system. In both methods, the cells are cultured in defined serum-free medium with EGF and/or FGF. However, instead of plating the cells on a nonadhesive substrate to generate free-floating spheres of cells (Reynolds et al., 1992), the progenitor cells are cultured in the presence of substrates such a polyornithine, laminin, or fibronectin coated on the cultureware. Under these conditions, the cells will adhere to the coated substrates and form monolayers (Ray et al., 1993).

Monolayer cultures sidestep some of the major limitations of neurospheres, namely the poor penetration of substances into tightly packed spheres and the fact that the neurospheres often contain a heterogenous population of cells at different stages of development (Conti and Cattaneo, 2010). By allowing cells to be more isolated and continuously bathed by the medium, NSCs grown in monolayers maintain a higher degree of homogeneity. The monolayer method also has the advantage of allowing direct access to cells for pharmacological testing and electrophysiology recordings. Nonetheless, because the culture system can greatly influence the properties of the NSCs, it is important that no single in vitro assay alone be used to define stem cells in vivo. Detailed methods of the isolation and monolayer cultivation of neural precursor cells can be found at Babu et al. (2011).

V. Regulation of Adult Neurogenesis in the Brain Vascular Niche and Choroid Plexus

The vasculature is emerging as an important site in the maintenance of the stem cell niche by providing spatial cues and regulatory signals. As arterioles traverse deeper into the brain, they lose their smooth muscle layer and essentially become tubes of endothelial cells (Girouard and Iadecola, 2006). The adult SVZ and SGZ contain an extensive vascular plexus in which dividing stem cells and TAPs are located near microvessels (Palmer et al., 2000). Tight junctions between endothelial cells and perivascular astrocytes are integral components of the blood-brain barrier (Abbott et al., 2006). The perivascular end-feet of astrocytes are closely applied to the microvessel wall through the basal lamina. Small-molecule diffusion from these endothelial cells is involved in the neurogenic process (Shen et al., 2004). SVZ blood vessels promote stem cell homeostasis and self-renewal by acting as a scaffold to support the seeding of the SVZ niche (Tavazoie et al., 2008). EphA/ephrin tyrosine kinase signaling plays a crucial role in the maintenance of the brain vascular system. In a recent study, Ephrin-A5-deficient mice display narrower hippocampal capillaries (Hara et al., 2010). These mice also have impaired SGZ-mediated neurogenesis.

Endothelial cells in the SVZ vascular niche may allow key small molecules such as vascular endothelial growth factor and FGF2 to diffuse to the NSCs and TAPs, which are established signals in the neurogenic process (Leventhal et al., 1999Shen et al., 2004). In addition to small molecules, hormones and their antagonists that cross the blood-brain barrier may also have direct access to the SVZ. This provides a potential mechanism for GPCR-based therapeutics to regulate neurogenesis.

The choroid plexus is composed of modified ependymal and epithelial cells located in each of the four ventricles that produce cerebrospinal fluid (CSF). The choroid plexus is also composed of many capillaries and set apart from the ventricles by a layer of epithelial cells. Liquid derived from the plasma filters through these cells and becomes the CSF. The choroid plexus also removes waste products, foreign substances, and excess neurotransmitters from the CSF (Wright and Saito, 1986).

GPCRs expressed in the choroid plexus may have direct or indirect actions on the SVZ, depending upon the innervation and local release of the hormone. The sympathetic nerves innervate the choroid plexus almost exclusively by the superior cervical ganglion (Edvinsson et al., 1974Lindvall et al., 1978b). Cholinergic and peptidergic systems also innervate the choroid plexus (Lindvall et al., 1978aAndo et al., 1986). Serotoninergic fibers are not found in the choroid plexus (Napoleone et al., 1982) but fibers from the dorsal raphe nucleus end at the ependymal surface and release serotonin in the CSF.

GPCRs found in the choroid plexus include the adhesion receptor GPR124, which is up-regulated after brain injury (Pickering et al., 2008). Other GPCRs confirmed or suspected to be expressed in the choroid plexus include the serotonin 5-HT2C (Pazos and Palacios, 1985Conn et al., 1986Labasque et al., 2008), 5-HT6 (Roberts et al., 2002), cannabinoid (Ashton et al., 2004Suárez et al., 2010), histamine H4 (Maślińska et al., 2009), orexin A (Tafuriet al., 2009), sphingosine-1-phosphate1–5 or EDG1–5 (Caballero et al., 2009), P2Y (Johansson et al., 2007), secretin (Siu et al., 2006), melanocortin type 2 (Nimura et al., 2006), corticotropin-releasing factor 2α (Chen et al., 2005), metabotropic glutamate (Gillard et al., 2003), somatostatin (Katz et al., 2002), vasopressin 1b (Zemo and McCabe, 2001), adrenomedullin (Juaneda et al., 2001Kobayashi et al., 2001), luteinizing hormone/chorionic gonadotropin (Lei and Rao, 2001), kinin B1 (Mahabeer et al., 2000), dopamine (Mignini et al., 2000), CXCR3 (Van Der Meer et al., 2001), endothelin B (Nakagomi et al., 2000), and the α1-adrenergic (Papay et al., 2006Gupta et al., 2009).

VI. G-Protein-Coupled Receptor Regulation of Adult Neurogenesis

A. Adrenergics

Adrenergic receptors (ARs) mediate the sympathetic nervous system by binding the endogenous catecholamines epinephrine and NE. Raymond Ahlquist (Ahlquist, 1948) introduced the initial concept of different AR subtypes (α and β). Nine AR subtypes (α1A, α1B, α1D, α2A, α2B, α2C, β1, β2, and β3) are now characterized. These receptors are all activated by the same catecholamines but result in various downstream effects (Bylund, 2005).

The adult hippocampus receives strong noradrenergic innervation from the locus ceruleus (Loy et al., 1980), and it is known that the early maturation of neural progenitor cells involves NE (Kärkkäinen et al., 2009). Pharmacological agents that increase NE levels can enhance hippocampal neurogenesis (Malberg et al., 2000) and the survival and differentiation of granule cells (Rizk et al., 2006) and olfactory bulb neurons (Bauer et al., 2003Veyrac et al., 2005). NE was shown to activate self-renewal and produce multipotent neural precursors in the adult mouse hippocampus (Jhaveri et al., 2010). Intrahippocampal injection of a β3-AR agonist or systemic injection of isoproterenol, a β-AR nonselective agonist, increased proliferation and the number of nestin/GFAP double-positive neural precursors (Jhaveri et al., 2010). Studies using a noradrenergic neurotoxin, N-(2-chloroethyl)-N-ethyl-2-bromo benzylamine hydrochloride (DSP-4), suggest that NE regulates proliferation but not survival or differentiation of precursors in the adult rat hippocampus (Kulkarni et al., 2002). However, this neurotoxin has been shown to increase norepinephrine release (Kask et al., 1997) and regenerative sprouting (Fritschy and Grzanna, 1992) as compensatory mechanisms.

Previous studies indicate that α1-ARs increase proliferation of embryonic neuroepithelial cells, suggesting that this subtype might influence adult progenitor proliferation (Popovik and Haynes, 2000;Kulkarni et al., 2002). α1-ARs stimulate proliferation (Hiramoto et al., 2006) and migration (Hiramoto et al., 2008) and protect against stress-induced death of mouse embryonic brain-derived neural progenitor cells through a caspase 3/7-independent mechanism in vitro (Ohashi et al., 2007). We have shown that α1A-AR subtype stimulation increases neurogenesis and gliogenesis in adult mice in vivo (Gupta et al., 2009). Adult mice expressing wild-type (WT) α1A-ARs tagged with enhanced GFP, mice with constitutively active mutant receptors under the control of a large fragment of the endogenous promoter, or normal WT mice treated with the α1A-AR selective agonist cirazoline, display increased BrdU incorporation in both the SVZ and SGZ, increased migration of α1A-AR positive cells, and increased numbers of Brdu/Nestin double-positive cells. In neurospheres derived from these mice, α1A-AR stimulation increases neuron and oligodendrocyte differentiation through a phosphoinositide 3-kinase (PI3K) survival mechanism that selectively increased the apoptosis of astrocytes.

α2-ARs exist both on target neurons receiving noradrenergic innervation and as α2-autoreceptors that decrease NE release from the presynaptic terminal. In the adult, the α2-AR agonists clonidine and guanabenz decrease proliferation of hippocampal progenitors with no effect on survival or differentiation (Yanpallewar et al., 2010). Adult hippocampal progenitors in vitro express all the α2-AR subtypes that decrease neurosphere frequency and BrdU incorporation, effects that are blocked with the α2-AR antagonist yohimbine. This study also observed no changes in hippocampal progenitor proliferation in any of the α2-AR knockout (KO) mice or with α2-AR antagonists, suggesting that α2-ARs induce a stimulatory, but not a basally, tonic inhibitory effect on adult SGZ progenitors. In contrast, another study (Rizk et al., 2006) reported enhanced progenitor survival after long-duration treatment with the α2-AR antagonist dexefaroxan.

B. Cannabinoids

The cannabinoid system plays a major role in the CNS and is emerging as a key regulator of adult neurogenesis, neuronal cell fate, and neuroprotection (Gowran et al., 2011). The two currently known cannabinoid (CB) receptors, CB1 and CB2, are both GPCRs. The CB1 receptors are expressed primarily in the CNS, whereas CB2 receptors are found mainly in hematopoietic and immune cells. Endocannabinoids are the endogenous ligands of CB receptors. Endocannabinoids promote long-lasting adult plasticity in a few areas of the brain, particularly in the dentate gyrus. In contrast, most drugs of abuse decrease adult hippocampal neurogenesis (Eisch et al., 2000Abrous et al., 2002Nixon and Crews, 2002),

In murine NPCs, anandamide, a nonselective endogenous cannabinoid, affects fate determination in vitro by promoting first glial and then neuronal differentiation without effects on survival or apoptosis (Soltys et al., 2010). These effects were shown to depend upon the phosphorylation of a cAMP-responsive element binding protein. Activation of diacylglycerol lipases involved in the synthesis of 2-arachidonoylglycerol, as well as CB1/CB2 receptor agonism, increases proliferation of cultured NSCs or hippocampal progenitors, whereas CB receptor antagonism and CB receptor-deficient cells result in inhibited proliferation (Aguado et al., 2005Jiang et al., 2005Palazuelos et al., 2006Goncalves et al., 2008). However, the CB2 receptor-mediated effect is more pronounced in the SVZ of older mice, suggesting that CB2 receptor agonists may counteract the natural decline in adult neurogenesis that is associated with aging (Goncalves et al., 2008Marchalant et al., 20082009a,b).

Neurosphere proliferation is regulated by both CB1 and CB2 receptor agonism through PI3K/AKT signaling (Molina-Holgado et al., 2007). Endocannabinoids are also thought to be involved with neuronal survival functions through PI3K/AKT phosphorylation (Galve-Roperh et al., 2008Maison et al., 2009) and through inhibition of glutamate neurotoxicity (Galve-Roperh et al., 2008). In addition, cannabinoids oppose the antineurogenic effects of neuronal nitric oxide (Kim et al., 2006b).

The plant-derived extracts Δ9-tetrahydrocannabinol (THC) and cannabidiol (CBD) produce differential effects on spatial learning and neurogenesis when fed to nestin-GFP reporter mice (Wolf et al., 2010). CBD, working through the CB1 receptor, does not impair learning but increases adult hippocampal neurogenesis, whereas THC reduces learning without affecting neurogenesis. CB1 receptor-deficient mice have reduced adult neurogenesis in both the SVZ and SGZ (Jin et al., 2004b). Because of the lack of psychoactive effects, CBD may represent a more promising candidate for therapeutic applications than THC (Pertwee, 2009). There is one report that the endogenous endocannabinoid anandamide and its analog methanandamide may inhibit the differentiation of cortical neuron progenitors to mature neurons and reduce adult hippocampal neurogenesis through attenuation of the Rap1/B-Raf/ERK pathway (Rueda et al., 2002).

It is possible that exo- and endocannabinoids may have differential effects on hippocampal neurogenesis as a result of their full versus partial agonism at CB receptors (Fride and Mechoulam, 2003). The CB1 receptor and the endocannabinoid-inactivating enzyme fatty acid amide hydrolase are expressed, in vivo and in vitro, in postnatal radial glia, adult nestin/GFAP-positive progenitor SVZ cells (Aguado et al., 2006Arévalo-Martín et al., 2007), and nestin/sox2-positive cells (Aguado et al., 2005;Mulder et al., 2008). In both cell culture and in postnatal CB1 receptor-deficient mice, CB1 receptors seem to regulate both progenitor proliferation and astrogliogenesis (Aguado et al., 2006). CB1 receptor activation increases the number of Oligo2-positive cells, whereas CB2 receptor activation increases polysialylated NCAM expression, both found on oligodendrocytes (Arévalo-Martín et al., 2007), through survival mechanisms involving PI3K/AKT (Molina-Holgado et al., 2002). Therefore, diseases and disorders characterized by glial deficiency potentially could be modulated through treatment with cannabinoids.

C. Chemokines (Stromal Cell-Derived Factor 1/CXC Chemokine Receptor Type 4)

Chemokines are small cytokines or proteins that are categorized into four groups. CXC chemokines (or α-chemokines) promote the migration of neutrophils and lymphocytes. CC chemokines (or β-chemokines) induce the migration of monocytes, natural killer cells, and dendritic cells. C chemokines (or γ-chemokines) attract T-cell precursors to the thymus, and CX3C chemokines (or δ-chemokines) serve as chemoattractants and adhesion molecules (Bonecchi et al., 2009).

CXCL12 [also known as stromal cell-derived factor 1 (SDF-1)] is the only known physiological ligand for the chemokine receptor CXCR4. It is a small chemokine protein of 8 to 13 kDa but is a well known mediator of neural progenitor cell migration during development. Mice that lack either the CXCR4 receptor or SDF-1 show abnormal development of the granule layer of the cerebellum (Ma et al., 1998;Zou et al., 1998) and the dentate gyrus of the hippocampus (Bagri et al., 2002Lu et al., 2002), which suggests that this receptor system is a critical regulator of neurogenesis. During development, SDF-1 is expressed by cells lining migratory paths and by cells located at the end of migratory paths. The migrating cells themselves express the receptor CXCR4 (Ma et al., 1998Zou et al., 1998). Anatomical studies using in situ hybridization and immunohistochemistry to localize chemokine receptors in the brain have demonstrated that they are expressed in the adult SGZ, SVZ, and olfactory bulb (Tran and Miller, 2005). Cells expressing these receptors have also been shown to express nestin and the TLX(tailless) gene—both markers for neural progenitors (Shi et al., 2004). In adult neural progenitor cells, SDF-1 induces migration and differentiation by increasing the expression of matrix metalloproteinases 3 and 9 (Barkho et al., 2008). In the postnatal brain, SDF-1 and CXCR4 are expressed in granule neurons and dividing neural progenitors in the dentate gyrus (Lu et al., 2002Banisadr et al., 2003;Bhattacharyya et al., 2008). SDF-1 plays a novel role as a neurotransmitter in the dentate gyrus and increases the strength of GABAergic inputs to the pool of dividing neural progenitors (Bhattacharyya et al., 2008).

Excitation-neurogenesis coupling is the process by which neuronal electrical activity affects the production of new neurons (Deisseroth et al., 2004). For example, GABAergic inputs and activation of GABAA receptors has been shown to promote neuronal differentiation and development in the adult hippocampus (Tozuka et al., 2005Ge et al., 2006). Excitation-neurogenesis is the tonic activation of CXCR4 by SDF-1 in newly formed granule cells. It is suggested to be essential for neurogenesis-dependent long-term memory in the adult hippocampus (Kolodziej et al., 2008).

Human neural progenitor cell proliferation is regulated through the CXCR4/AKT-1/Forkhead box O3a signaling pathway (Wu et al., 2009) and changes in cell cycle proteins that affect neuronal survival (Khan et al., 2003). AKT-1 promotes proliferation by interacting with 14-3-3 proteins that sequester p21 or by increasing the expression of cyclin D proteins (Muise-Helmericks et al., 1998Zhou et al., 2001). In addition, AKT-1 phosphorylates and inhibits the winged-helix family of transcription factors (Forkhead box O3a), which are negative regulators of cell cycle progression (Brunet et al., 1999Nakamura et al., 2000Brunet et al., 2001). There are contrasting reports as to whether SDF-1 promotes quiescence (Krathwohl and Kaiser, 2004) or proliferation (Gong et al., 2006). One study that used primary cells from rat cortex found that SDF-1 promotes neural progenitor cell proliferation via the ERK1/2 and PI-3 kinase pathways (Gong et al., 2006), whereas the other study observed that SDF-1 inhibits growth and differentiation and induces quiescence of human NPCs in isolated intermediate progenitor cells, neurospheres, and hippocampal slice cultures (Krathwohl and Kaiser, 2004). Species differences or preparation variance are the most likely cause of contrasting results.

D. Dopamine

Dopamine is a catecholamine that is synthesized by neurons most commonly found in the substantia nigra, ventral tegmental area, and hypothalamus. Dopamine neurological function encompasses roles in punishment and reward, mood, sleep, attention, and learning (Lima et al., 2009). Dopamine receptors are classified into two groups, D1-like (D1 and D5) and D2-like (D2, D3, and D4), on the basis of their structure and interactions with different G-proteins. D1-like receptors activate adenylyl cyclase through Gs, whereas D2-like receptors inhibit adenylyl cyclase activity through Gi (Holmes et al., 2004).

Dopamine receptors are known to be expressed in neurogenic regions. Dopamine receptors, particularly D3, are highly expressed in the developing brain in areas especially active in neurogenesis and persist in the adult SVZ (Diaz et al., 1997). In the adult rodent brain, both D1- and D2-like receptors have been identified in SVZ-derived neurospheres (Coronas et al., 2004), specifically in β-III immunopositive neuronal cells. mRNA expression and in vivo studies have demonstrated a particularly strong D3 receptor presence in embryonic and adult SVZ neurogenic regions (Diaz et al., 1997Van Kampen et al., 2004). Fluorescence-activated cell sorting of adult mouse SVZ cells using hGFAP-GFP and DCX-GFP mice showed that TAP cells express D3 receptors, whereas stem cell-like astrocytes and neuroblasts do not (Kim et al., 2010).

Dopamine is known to enhance adult neurogenesis (Winner et al., 2009). Depletion of dopamine in rodents decreases precursor cell proliferation in both the subependymal zone and the SGZ. Proliferation in the SVZ diminishes significantly after dopaminergic deafferentation. However, the number of neural progenitor cells expressing the proneuronal cell fate determinant Pax-6 increases in the SVZ. Survival and quantitative cell fate analysis of newly generated cells revealed that 6-hydroxydopamine (6-OHDA) lesions induce opposing effects in the two different neurogenic regions of the brain: a transient decrease in the granule cell layer causes a sustained increase of newly generated neurons in the glomerular layer (Winner et al., 2006). Destruction of dopamine neurons in the substantia nigra and ventral tegmental areas reduces the number of proliferating neural precursors in the SVZ by 40% (Baker et al., 2004). Inhibition of dopaminergic transmission in adult rats in vivo using the D2-like antagonist haloperidol led to inconsistent results (Wakade et al., 2002Kippin et al., 2005). However, systemic treatment of either normal or dopamine-depleted rats with the D2/3-like agonists ropinirole or 7-hydroxy-N,N-di-n-propyl-2-aminotetralin significantly increases precursor cell proliferation in the SVZ (Höglinger et al., 2004Van Kampen et al., 2004Van Kampen and Robertson, 2005), resulting in increased differentiation of neurons. These studies identify dopamine D2-like receptors as a regulator of adult neurogenesis with implications for the potential use of endogenous neural precursors in cell replacement strategies for Parkinson’s disease (PD).

There are conflicting reports on the ability of dopamine to promote neurogenesis, although these studies are in the minority. One study that explored 7-hydroxy-N,N-di-n-propyl-2-aminotetralin on human and murine NPCs derived from the fetal midbrain found no effects on proliferation, survival, or neurogenesis (Milosevic et al., 2007). Another study showed that repetitive cocaine administration decreases proliferation in the adult rat hippocampus with no effect on differentiation or survival (Yamaguchi et al., 2004Domínguez-Escribà et al., 2006). It is possible that these discrepancies in the literature are due to activation of distinct subtypes of dopamine receptors, particularly in the SGZ, at different developmental stages. A recent study revealed that dopamine is particularly effective in modulating the function of adult newborn neurons in the SGZ but not in mature dentate granule cells (Mu et al., 2011). It is also possible that optimum dopaminergic modulation of neurogenesis requires simultaneous activation of D1- and D2-like receptors, shown to rescue proliferation in the 6-OHDA animal model with significantly greater increases in neuronal differentiation in the olfactory bulb than with levodopa alone (O’Keeffe et al., 2009a).

D2-like receptors are thought to promote adult neurogenesis. In vivo, D3 receptor antagonism reduces the numbers of newborn neurons that reach the olfactory bulb and decreases progenitor cell proliferation but does not change the number of BrdU-retaining (stem) cells. No effect of the D1-like agonist 2,3,4,5-tetrahydro-7,8-dihydroxy-1-phenyl-1H-3-benzazepine (SKF 38393) on cell proliferation is observed (Höglinger et al., 2004). These studies suggest that D2-like receptors (i.e., D3 receptors) are expressed on transit-amplifying progenitor cells but not SVZ stem cell-like astrocytes (Kim et al., 2010).

A variety of factors are known to influence dopamine-induced neurogenesis. It has been proposed that the mechanism of D2 receptor-mediated adult neurogenesis in mice involves ciliary neurotrophic factor (CNTF) (Yang et al., 2008b). Dopaminergic denervation in adult mice reduces CNTF mRNA by approximately 60%, whereas systemic treatment with the D2 receptor agonist quinpirole increases CNTF mRNA in the SVZ and SGZ. Quinpirole acted on postsynaptic receptors in that it reversed the reduced proliferation seen after dopaminergic denervation in wild-type mice. In vitro studies using human NPCs show that treatment with NMDA during proliferation and differentiation increases the amount of tyrosine-hydroxylase-immunopositive cells, which is reversed by memantine (Wegner et al., 2009). These results suggest that N-methyl-d-aspartate glutamate receptors in differentiating human NPCs are important regulators of dopaminergic neurogenesis in vitro. Another factor involved in regulating dopamine-induced proliferation in the SVZ is EGF, but not FGF, in a PKC-dependent manner (O’Keeffe et al., 2009b). In vivo dopamine depletion decreases proliferation in the SVZ concomitant with reduced EGF and is reversed by administration of levodopa. In addition, EGFR-positive cells are depleted in patients with Parkinson’s disease (O’Keeffe et al., 2009b).

E. Glutamate

Glutamate is the major excitatory neurotransmitter of the CNS. The glutamate system mediates its effects through both ionotropic ligand-gated ion channel receptors and metabotropic glutamate receptors (mGluR), which couple to G proteins. The main glutamatergic input to the hippocampal dentate gyrus is from the entorhinal cortex via the perforant path (Collingridge and Lester, 1989). Most of the actions of glutamate on adult neurogenesis have been ascribed to its ionotropic receptors. The effects of the mGluRs on adult neurogenesis are unclear and conflicting. There are eight known subtypes of mGluRs divided into three groups: group 1 (mGluR1 and -5), group 2 (mGluR2 and -3), and group 3 (mGluRs 4, 6, 7, and 8).

The subtypes mGluR3, mGluR4, and mGluR5 are expressed in postnatal NSCs (Canudas et al., 2004Di Giorgi Gerevini et al., 20042005). Subtype-specific effects of mGluRs on adult neurogenesis remain unclear. mGluR1 antagonism, but not mGluR4/5 blockade, reduces cell proliferation in hippocampal organotypic slices, which suggests that mGluR1 receptors may facilitate neurogenesis (Baskys et al., 2005). In cultured mouse NPCs, mGlu2/3R or mGlu5R blockade reduces cell proliferation and survival, whereas mGluR5 activation increases neurogenesis (Di Giorgi-Gerevini et al., 2005). Additional studies have shown that treatment with the mGluR2/3 blocker (1R,2R,3R,5R,6R)-2-amino-3-(3,4-dichlorobenzyloxy)-6-fluorobicyclo[3.1.0]hexane-2,6-dicarboxylic acid (MGS0039) enhances cell proliferation in vivo (Yoshimizu and Chaki, 2004). It was recently shown that mGluR3 activation inhibits astrocytic differentiation of SVZ-derived NSCs through a mechanism involving bone morphogenetic protein receptor signaling (Ciceroni et al., 2010) and that treatment with the novel mGluR2 agonist (2R,4R)-4-aminopyrrolidine-2,4-dicarboxylate inhibits diffuse brain injury-induced neurogenesis (Feng et al., 2011). There is evidence suggesting a link between mGlu4R function in cerebellar granule cell neuroprogenitors and medulloblastomas and mGlu3R-mediated proliferation and survival of forebrain neural stem/progenitor cells and malignant gliomas [clearly, more study is needed to determine the significance of glutamate in regulating adult neurogenesis (for review, see Melchiorri et al., 2007)].

F. Lysophosphatidic Acid and Sphingosine 1-Phosphate

Lysophosphatidic acid (LPA) and sphingosine-1-phosphate (S1P) are extracellular bioactive phospholipids that have recently emerged as important influences on normal nervous system development. These ligands act through at least five specific GPCRs for each system, LPA1–5 (Chun et al., 2002Meyer zu Heringdorf and Jakobs, 2007) and S1P1–5 (Rosen et al., 2009).

During embryonic development, LPA regulates cortical growth, proliferation, differentiation, and cell survival (Kingsbury et al., 2003). In embryonic tissue studies, LPA induces proliferation of murine cortical neuroblasts in vitro (Contos et al., 2000) but not in vivo (Kingsbury et al., 2003). However, it has no proliferative effect on hippocampal neural progenitor cells of the embryonic rat (Harada et al., 2004). These contrasting results could suggest that the effects of LPA are dependent upon the tissue source, species, and developmental stage of the subject (for review, see Pébay et al., 2007).

During development, LPA1 is expressed in neural progenitor cells, suggesting a regulatory function in neurogenesis (Hecht et al., 1996). In the embryonic mouse brain, LPA1 is expressed in bands of cells adjacent to several ventricles and near the lateral ventricle in the ventricular zone of the neocortex, locations that are coincident with neurogenesis (McGiffert et al., 2002).

In human embryonic stem cell-derived neurospheres, all five of the LPA receptors are expressed. LPA specifically inhibits the differentiation of NSCs toward neurons without affecting proliferation, whereas it maintains the differentiation of NSCs toward astrocytes (Dottori et al., 2008). These effects are not blocked by pertussis toxin, which inhibits Gi signaling, and are only partially inhibited by (S)-phosphoric acid mono-{2-octadec-9-enoylamino-3-[4-(pyridin-2-ylmethoxy)-phenyl]-propyl} ester (VPC32183), an antagonist of the LPA1 and LPA3 receptors. In contrast, these effects are fully blocked by a combination of PI3K/AKT and Rho/Rock inhibitors, suggesting the involvement of LPA receptors other than LPA1/3.

LPA is also known to increase adult neurogenesis by promoting the differentiation of cells toward neuronal lineages (Svetlov et al., 2004). In neurospheres obtained from mouse postnatal forebrain, LPA induces proliferation of cells coexpressing Sca-1 and AC133, markers of primitive hematopoietic and NSCs, along with the LPA receptor subtypes 1 to 3 (Svetlov et al., 2004). These effects can be blocked by diacylglycerol-pyrophosphate, an antagonist of LPA1 and LPA3 receptors. In adult mice, LPA1 is thought to promote adult neurogenesis. Although 50% of LPA1-null mice die perinatally, a variant that arises spontaneously after colony expansion, termed maLPA1, exhibit reduced ventricular zone, altered neuronal markers, and increased cortical death (Estivill-Torrús et al., 2008). Subsequent examination of these mice shows defects in proliferation, differentiation, and survival of newly formed neurons in the SGZ (Matas-Rico et al., 2008). In addition, conditions known to induce neurogenesis, including enriched environment and voluntary exercise, are impaired in the maLPA1 mice. Analysis of trophic factors in maLPA1-null mice demonstrate alterations in brain-derived neurotrophic factor and insulin growth factor 1 levels after enrichment and exercise (Matas-Rico et al., 2008). These mice also display increased anxiety and spatial memory defects (Santin et al., 2009).

S1P1 is expressed in bands of cells adjacent to several ventricles and near the lateral ventricle in the embryonic mouse brain (McGiffert et al., 2002). The expression pattern of S1P1 also encompasses the hippocampal primordia, olfactory bulb, and ganglionic eminence (McGiffert et al., 2002). In humans, embryonic SC-derived NSCs express S1P1,3, which induces proliferation and morphological changes of neural progenitor cells (Harada et al., 2004Dottori et al., 2008). LPA1 and S1P5>1–3 are expressed in embryonic and adult rodent OPCs (Dawson et al., 2003Terai et al., 2003Jaillard et al., 2005;Novgorodov et al., 2007), whereas fetal human OPCs express higher levels of S1P1 (Miron et al., 2008). OPC maturation requires first the formation of processes involved in OPC migration, the actual migration of OPCs and is complete upon its myelination of neurons (Baumann and Pham-Dinh, 2001). Both LPA and S1P act on OPCs and inhibit oligodendrocyte maturation (Dawson et al., 2003Jaillard et al., 2005). S1P5 is preferentially expressed in OPCs and is responsible for the inhibition of their migration in rodents (Novgorodov et al., 2007). In human OPCs, S1P5,3 activation with the sphingosine analog fongolimod (FTY720), a systemic immunomodulatory therapy for multiple sclerosis, has a biphasic effect on maturation. Short-term (1 day) FTY720 treatment causes initial process retraction, whereas a 2-day treatment stimulates process formation and increases cell survival in a S1P1-ERK-dependent manner (Coelho et al., 2007Miron et al., 2008). There are also reports that S1P signaling cross-talk with neurotrophin-3 stimulates proliferation and survival of OPCs through a cAMP-response element-binding protein (Saini et al., 20042005).

G. Melatonin

Melatonin is a widespread, naturally occurring compound. Many of its physiological effects are mediated through its activation of the GPCRs, MT1 and MT2. Its secretion into the blood from the pineal gland varies in a daily cycle, thereby allowing the regulation of sleep and circadian rhythm (Dubocovich et al., 2003). Adult hippocampal neurogenesis is affected by circadian rhythms and sleep deprivation (Holmes et al., 2004Guzman-Marin et al., 2005), thus implying a significant role for melatonin. In vitro, melatonin increased the number of new neurons differentiated from adult mouse hippocampal neural precursor cells by promoting cell survival without affecting proliferation (Ramírez-Rodríguez et al., 2009Rennie et al., 2009). Although a previous study revealed increased proliferation in the dentate gyrus of maternally separated rats seems contradictory, this study only measured BrdU incorporation without the use of neurogenetic cell markers (Kim et al., 2004). Melatonin’s effect on neuronal cell survival were recapitulated in vivo by application of exogenous melatonin (8 mg/kg) and demonstrated antidepressant behavior in mice (Ramírez-Rodríguez et al., 2009). Melatonin also increases neuritogenesis and dendritogenesis with agonist treatment, leading to greater complexity of the dendritic tree (Ramirez-Rodriguez et al., 2011). Melatonin has also been shown to ameliorate the irradiation-induced decline in adult hippocampal neurogenesis, suggested to occur through its ability to scavenge free radicals (Manda et al., 2009). This benefit may be useful to combat the side effects of brain radiotherapy.

H. Muscarinic

While cholinergic activity plays a significant role in adult neurogenesis, most of these effects are documented through nicotine and the acetylcholine receptor ion channels (for review, see Campbell et al., 2011). Little is known about the role of cholinergic receptors in the GPCR system and their role in adult neurogenesis. Through immunohistochemistry and reverse transcription-polymerase chain reaction, M1 and M4 receptors were expressed in immature neurons in the adult mouse olfactory bulb or dentate gyrus (Kaneko et al., 2006) or those that colabeled with BrdU in the rat subgranular zone (Mohapel et al., 2005). The M2, M3, and M4 muscarinic receptors are expressed in embryonic rat neuroepithelial cells, and muscarinic agonists increased proliferation and differentiation of neurons (Ma et al., 2000). The nonselective muscarinic agonist, oxotremorine could alleviate the stress-induced decrease in the proliferation, survival, and differentiation of progenitor cells in the hippocampus (Veena et al., 2011). However, in contrast to an earlier study, muscarinic blockage with scopolamine only affected survival of BrdU+ cells without changing neural proliferation or differentiation in the unstressed adult rat hippocampus (Kotani et al., 2006). These results could indicate that muscarinic activity becomes important in adult hippocampal neurogenesis only under stress conditions.

I. Opioids

The classic opioid receptor system modulates many different aspects of physiology but is most known for its control of nociception (Fields, 2004) and reward-related behavior (Nestler, 2004). Four major subtypes of opioid receptors have been cloned: μ-opioid receptors (MORs), δ-opioid receptors (DORs), κ-opioid receptors (KORs), and nociception receptors (Waldhoer et al., 2004). All are GPCRs and are activated by endogenous opioid peptides.

Endogenous opioids and exogenously administered synthetic opiate agonists acting on the MORs, DORs, and KORs have been shown to affect proliferation and differentiation in various embryonic neural cell types. In embryonic NSCs, stimulation of DORs using (+)-4-[(αR)-α-(2S,5R)-4-allyl-2,5-dimethyl-1-piperazinyl)-3-methoxybenzyl]-N,N-diethylbenzamide (SNC80) promotes neuronal differentiation through PI3K/PKC/Ca2+/calmodulin-dependent protein kinase II/mitogen-activated protein kinase kinase, but MOR agonism ([d-Ala2,N-Me-Phe4,Gly5-ol]enkephalin) or KOR stimulation [trans-(±)-3,4-dichloro-N-methyl-N-[2-(1-pyrrolidinyl)cyclohexyl]benzeneacetamide (U50,488H)] have no effect (Narita et al., 2006). In embryonic stem cells derived from a mouse blastocyst, both MOR and KOR promote proliferation and differentiation into neural progenitors through ERK (Kim et al., 2006a). MOR and KOR functionality is found in neural progenitor-derived oligodendrocytes and induces their differentiation via ERK and p38 (Hahn et al., 2010). Activation of opioid receptors blocks the differentiation of both astrocytes and neurons in retinoic acid-induced neural progenitors (Hahn et al., 2010). It is noteworthy that ERK signaling drives the inhibition of astrogenesis, whereas p38 drives the inhibition of neurogenesis.

In early human gestation, both DOR and MOR subtypes are expressed strongly in the SVZ during gestational weeks 11 to 16, but decrease by week 20. DORs and MORs are expressed in multipotential stem cells, newly differentiated neurons, and developing glial cells. However, migrating neurons express negligible levels of either subtype (Tripathi et al., 2008). There is robust expression of KORs in highly enriched (>90% nestin-positive) human fetal brain-derived neural progenitors, and receptor stimulation increases proliferation and migration. These effects are partially blocked by the KOR antagonist, nor-binaltorphimine (Sheng et al., 2007). These many contrasts between these studies on proliferation and differentiation may be due to different mechanisms operating in opioid-mediated neurogenesis versus gliogenesis or may be due to species variations.

In the dentate gyrus of the adult rat hippocampus, all three opioid receptor subtypes have been found both at the mRNA and protein levels (Delfs et al., 1994Mansour et al., 1995a,b). MOR is highly expressed in the brainstem and spinal cord, regulating functions related to these brain regions such as addiction, alertness, and nociception (Arvidsson et al., 1995Uhl et al., 1999). However, MORs are also well known to regulate hippocampal function (Morris and Johnston, 1995Terman et al., 2000Guo et al., 2005). MORs are expressed on interneurons and primary cell types in all hippocampal regions (Meibach and Maayani, 1980Arvidsson et al., 1995Svoboda et al., 1999Drake and Milner, 2002), including the granule cell layer (GCL) of the dentate gyrus.

The role of opioids in proliferation and differentiation in the adult hippocampus has been complicated and conflicting. Opiates decrease adult SGZ proliferation by inhibiting progenitor proliferation, maturation, and survival (Eisch et al., 2000Kahn et al., 2005Arguello et al., 2008) and by altering the progenitor cell cycle (Mandyam et al., 20042007Arguello et al., 2008). Long-term morphine exposure decreases SGZ neurogenesis by inhibiting dividing cells (particularly those in S phase) and inhibiting progenitor cell progression to a more mature neuronal stage (Arguello et al., 2009). MOR-KO mice exhibit normal proliferation and differentiation but enhanced survival of new neurons and more granule cells (Harburg et al., 2007). In contrast, incubation with the opioid receptor antagonist naloxone reduces proliferation in cultured adult rat hippocampal progenitors with a coincident increase in the differentiation of neurons, but a decrease in astrogliogenesis and oligodendrogenesis is observed (Persson et al., 2003). These inconsistencies are potentially due to the involvement of multiple opioid receptor subtypes in the regulation of neurogenesis and gliogenesis and to differential effects of opioid antagonists on hippocampal progenitor proliferation (Persson et al., 2004).

Opioid receptors have also been shown to decrease proliferation in the SVZ, most studies concentrating on early postnatal animals. Opioid agonists consistently decrease SVZ proliferation in an age-dependent manner. In the postnatal rat SVZ, morphine or β-endorphins, ligands known to interact with opioid receptors in nociception (Bartolome et al., 1994), decrease proliferation in early development but show no effect on older animals (Kornblum et al., 1987Lorber et al., 1990). In addition, β-endorphin antagonism by a sulfated fragment of cholecystokinin inhibits DNA synthesis (Bartolome et al., 1994). In addition, DNA synthesis in the P11 forebrain is decreased in response to synthetic enkephalin, an opioid receptor agonist (Vértes et al., 1982).

Endogenous opioids modulate adult neurogenesis and gliogenesis in the SVZ by inhibiting cell proliferation as observed in songbirds (Khurshid et al., 2010). These data support an inhibitory mode of action for opioid ligands on cellular proliferation in the SVZ of postnatal mammals. In contrast, studies in the adult rat find that short-term morphine treatment increases [3H]thymidine uptake in the SVZ (Messing et al., 1979Miller et al., 1982). Cellular proliferation, however, could not be blocked by naltrexone, an opioid receptor antagonist, suggesting that a nonopioid receptor-mediated process may be involved.

J. Peptide Hormones

Vasoactive intestinal polypeptide (VIP) is a peptide neurotransmitter released by GABAergic interneurons in the dentate gyrus. VIP and its receptors (VPAC1 and VPAC2) are expressed in developing and adult dentate gyrus, and VIP blockade causes impairments in associative learning abilities (Gozes et al., 1995).

VIP is known to promote embryonic neurogenesis. It shortens the cell cycle of embryonic neuroepithelial cells (Gressens et al., 1998) and promotes neuronal differentiation of embryonic hippocampal neurons in culture (Blondel et al., 2000). VPAC2 receptor activation shifts the fate of symmetrically dividing neural progenitors toward a nestin-only phenotype, without changing rates of proliferation. In contrast, selective VPAC1 receptor stimulation moves neural progenitor fate toward granule cell neurogenesis (Zaben et al., 2009). Adult Vipr2(−/−) mice exhibit reduced progenitor survival and SGZ neurogenesis, demonstrating a trophic role for VPAC2 receptors in vivo. Type 2 nestin-positive precursors in vivo are also reduced, consistent with a role for VPAC2 in maintaining this cell population (Zaben et al., 2009). This seminal work provides substantial evidence for differential modification of neurogenesis by the VPAC subtypes.

Pituitary adenylate cyclase-activating polypeptide (PACAP) is a 38-amino acid C-terminally α-amidated neuropeptide that was originally isolated from the hypothalamus on the basis of its ability to stimulate adenylyl cyclase activity in rat anterior pituitary cells (Miyata et al., 1989). Alternative processing of the PACAP precursor can generate a 27-amino acid α-amidated peptide (PACAP27) that possesses the same biological activity as the 38-residue peptide (Miyata et al., 1990). PACAP exhibits 68% sequence identity with VIP. The activities of PACAP are mediated through PAC1 receptors (PAC1) that have a much lower affinity for VIP. However, VPAC1 and VPAC2 receptors recognize both PACAP and VIP with high affinity (Vaudry et al., 2000).

During embryonic development, PACAP is expressed in the germinal neuroepithelia and possibly embryonic stem cells. PACAP interacts with sonic hedgehog to reduce proliferation of granule cell precursors (Nicot et al., 2002), increase neurite outgrowth (Gonzalez et al., 1997), and promote survival and differentiation, processes that are required for the development of the cerebellum (Falluel-Morel et al., 2007). PACAP and PAC1 are highly expressed and colocalized in neural progenitors of the mouse cortex at embryonic day 14.5, and induce proliferation via Gq-mediated, but not Gs-mediated, phospholipase C (PLC)/inositol trisphosphate-dependent signaling. These progenitors differentiate mostly into astrocytes (Nishimoto et al., 2007). PACAP also increases DNA synthesis in oligodendrocyte progenitors isolated from neonatal rat brain but delays their maturation (Lee et al., 2001).

PAC1 is expressed in both the SVZ and SGZ of the adult mouse brain and is thought to promote adult neurogenesis. Cultured NSCs/progenitors isolated from the lateral ventricle wall of adult mice express PAC1 and proliferate in vitro in response to two PAC1 agonists, PACAP and maxadilan, but not VIP, at physiologic concentrations. This indicates that PAC1 is a mediator of NSC/progenitor proliferation (Mercer et al., 2004Ohta et al., 2006). In vitro characterization studies showed that PACAP is capable of inducing NSCs to form multipotent neurospheres that can self-renew and differentiate into both neurons and glia (Mercer et al., 2004) through a PKA-dependent pathway (Ohta et al., 2006). In vivo intracerebroventricular infusion of PACAP increases proliferation in both the SVZ and SGZ (Mercer et al., 2004Ohta et al., 2006). Recent studies using adult PACAP(−/−) mice indicate decreased survival of newly divided cells in the SGZ, but no difference in proliferation or differentiation compared with control mice under normal or enriched conditions (Ago et al., 2011). In addition, PACAP effectively attenuates apoptosis in the spinal cord of injured rats (Chen and Tzeng, 2005).

VIP is a member of the secretin/VIP/glucagon peptide family and is involved in the positive regulation of embryonic neurogenesis. Secretin (−/−) mice display a decreased number of BrdU-labeled neurons and a dramatic increase in neural progenitor cell apoptosis in in the SGZ during the early postnatal period (Jukkola et al., 2011).

In the adult, VIP also promotes neurogenesis. Adult secretin(−/−) mice have a reduced dentate gyrus volume, decreased long-term potentiation, and impaired spatial learning ability (Jukkola et al., 2011). Taken together, these results suggest a protective and proliferative role in the adult brain for the secretin/VIP/PACAP peptide family that regulates SGZ neurogenesis.

K. Neuropeptide Y

Neuropeptide Y (NPY) is a 36-amino acid peptide that belongs to a family of peptides that includes pancreatic polypeptide and peptide YY. NPY is secreted by the hypothalamus and is widely distributed in the central and peripheral nervous systems. The peptide exerts its biological effects through five GPCRs (Y1–Y5), four of which are functional in humans (Y1, Y2, Y4, Y5).

NPY is an important neuromodulator that regulates mood, endocrine function, blood pressure, nociception, appetite, body weight regulation, and cognition (Michel et al., 1998). Subtypes Y1 and Y5 have known roles in the stimulation of feeding, whereas Y2 and Y4 seem to have roles in appetite inhibition and satiety (Kamiji and Inui, 2007MacNeil and Douglas, 2007). NPY receptors are also known to be neuroprotective (Xapelli et al., 2006). The Y2 and Y5 receptors were shown to be protective in kainate-induced excitotoxicity in neuronal cultures, as well as in vivo after intrahippocampal kainate injection (Smiałowska et al., 2009). Y2R agonists also show neuroprotective activity in the ischemic middle cerebral artery occlusion model (Smiałowska et al., 2009).

In humans, NPY immunoreactivity is present in all five layers of the cortex of the inferior parietal lobe and the white matter under this cortex (Krivokuća et al., 2010). The dentate gyrus also shows particularly high NPY-like immunoreactivity (Dumont et al., 1992). Particular neuronal cell types that are immunoreactive for NPY include the Cajal-Retzius, known to be involved in embryonic neurogenesis (Bielle et al., 2005) and Alzheimer’s disease (Baloyannis, 2005).

NPY increases cell proliferation in the rat SVZ through an ERK pathway but does not affect the self-renewal of NSCs (Thiriet et al., 2011). NPY is also involved in ATP-induced neuroproliferation in adult mouse olfactory epithelium (Jia and Hegg, 2010). Using KO mice in vivo or cell cultures with specific NPY receptor agonists and antagonists, the neuroproliferative and neuronal differentiating effects of NPY in the SVZ were found to be mediated by the Y1 receptor subtype through ERK-mediated signaling (Hansel et al., 2001Howell et al., 20032005Agasse et al., 2008Decressac et al., 2009). Mice with a targeted deletion of NPY contain half as many dividing olfactory neuronal precursor cells as do controls and develop significantly fewer olfactory neurons by adulthood (Hansel et al., 2001).

NPY promotes SGZ neurogenesis as well, and this effect is also thought to be mediated by the Y1 receptor subtype through ERK signaling (Hansel et al., 2001Howell et al., 20032005Agasse et al., 2008;Decressac et al., 2009). Intracerebroventricular administration of NPY increases SGZ cell proliferation and promotes neuronal differentiation in adult mice through the Y1, but not Y2, receptor (Decressac et al., 2011). Furthermore, NSCs/progenitors from the postnatal hippocampus are sensitive to the combination of NPY and FGF2, which significantly shortens the cell cycle time of nestin-positive NSCs more than either factor alone, enhancing rates of proliferation (Rodrigo et al., 2010).

NPY cells express other neurotransmitters, such as nitric-oxide synthase, which influence progenitor cell proliferation, migration, and neurite outgrowth (Reif et al., 2004Chen et al., 2005). Within the hippocampus, there is an increasing appreciation of the role of GABAergic interneurons as a mechanism for excitation-neurogenesis coupling (Deisseroth and Malenka, 2005). Many in vivo manipulations that affect electrical activity influence neuron production from neural stem/progenitor cells in the dentate gyrus (for review, see Lehmann et al., 2005). GABAergic excitation is also known to promote neuronal differentiation in these cells (Tozuka et al., 2005). NPY is coreleased with GABA by GABAergic interneurons under high-frequency firing conditions and induces proliferation of nestin-positive NSCs cultured from the whole hippocampus (Howell et al., 2003).

Seizures are known to induce NPY expression (Cardoso et al., 2010). NPY then partly mediates seizure-induced NSC proliferation in the SGZ in vivo (Howell et al., 2007). Furthermore, Y1-deficient mice show decreased seizure-induced proliferation in the SGZ, as well as in the subcallosal zone, where seizures induce glial proliferation (Howell et al., 2007).

L. Purinergic Receptors

Two families of purinergic GPCRs exist: adenosine receptors (or P1 receptors) and P2Y receptors (Fredholm et al., 1994). In humans, there are four adenosine receptors (A1, A2A, A2B, A3), with the A1and A2A receptors playing important roles in the brain. Adenosine is the endogenous ligand for the adenosine receptors (Fredholm et al., 2001). P2Y receptors are another family of purinergic GPCRs that is stimulated by nucleotides such as ATP, ADP, UTP, UDP, and UDP-glucose. To date, 12 P2Y receptors have been cloned in humans: P2Y1–2, 4–6, 8–14 (Abbracchio et al., 2006). The gaps in the numbering of the P2Y subtypes are due to previous incorrect classification.

Adenosine modulates neuronal activity presynaptically by inhibiting or facilitating transmitter release and postsynaptically by affecting the action of other neurotransmitters. It acts nonsynaptically by hyperpolarizing or depolarizing neurons (Ribeiro et al., 2002). Adenosine has been shown to play protective roles in the hippocampus, which is highly sensitive to ischemia and hypoxia (Sebastião et al., 2001).

The amount of information on the neurogenic effects of the P1 adenosine receptor is very limited. Cells isolated from the mouse fetal midbrain increase neuron formation when adenosine is applied (Delic and Zimmermann, 2010). However, there are no studies exploring neurogenic potential of P1 adenosine receptors in adult cell or animal models.

ATP-dependent purinergic signaling through the P2Y1 receptor has been associated with developmental neurogenesis (Weissman et al., 2004), as radial glial signaling required for embryonic neurogenesis involves P2Y1 ATP receptors. Embryonic NPCs express P2Y purinergic receptors and release ATP themselves to mobilize intracellular calcium and increase progenitor proliferation. Furthermore, P2Y receptor antagonists suppress proliferation and instead promote differentiation into neurons and glia in vitro, whereas subsequent removal of purinergic inhibition restores progenitor cell expansion (Lin et al., 2007). ATP, UTP, and adenosine-5′-O-(2-thiodiphosphate), in conjunction with EGF, induce converging signals that cause proliferation and then guide adult neural progenitor migration through the formation of stress fibers, actin cytoskeleton, AKT, and focal adhesion kinase in cultured adult murine NSCs (Grimm et al., 2009Grimm et al., 2010).

In human NSCs, ATP and P2Y receptors have been identified as mitogens for v-myc immortalized neural progenitor cells through calcium release and PI3K (Ryu et al., 2003). Activation of P2Y induces DNA synthesis in glial cells (Neary and Zhu, 1994Milenkovic et al., 2003) as well as radial glia (Uckermann et al., 2002). NTPDase, which degrades active ATP, is selectively localized to the adult rat SVZ along with P2Y1 and P2Y4 receptors (Lin et al., 2007). In the adult mouse SVZ, type B cells and residual radial glia cells in the hippocampus selectively express NTPDase2 (Shukla et al., 2005Mishra et al., 2006;Langer et al., 2007). In addition, neurospheres cultured from the adult mouse SVZ express NTDPase2, P2Y1, and P2Y2 receptors. P2Y1 and P2Y2 receptor agonism augments proliferation in the presence of growth factors, whereas neurospheres isolated from P2Y1-deficient mice show reduced proliferation (Mishra et al., 2006). In another study using the selective P2Y1 receptor antagonist 2′-deoxy-N6-methyladenosine 3′,5′-bisphosphate (MRS 2179), primary neurospheres from the adult mouse SVZ demonstrated A2A and P2Y1 receptor-mediated inhibition of neurosphere generation and proliferation (Stafford et al., 2007). However, MRS 2179 may be only ~11-fold selective for the P2Y1 over the P2X1receptor ligand-gated ion channels (Brown et al., 2000).

Previous studies have shown that expression of P2Y receptors in astrocytes is altered when connexin43, a gap junction protein, is down-regulated (Suadicani et al., 2003) and that the growth rate of Cx43-null astrocytes is decreased compared with that of WT cells (Dermietzel et al., 2000). In embryonic neurospheres isolated from connexin 43-deficient mice, the concomitant decrease in P2Y1 receptors was found to regulate the decrease in proliferation and migration as a result of this gap-junction channel loss (Scemes et al., 2003). These results are consistent with an in vitro model of CNS differentiation in which neural progenitor cells were found to be coupled by Cx43 channels, and blockade of gap junctional communication strongly reduced proliferation and differentiation of neural progenitors (Duval et al., 2002).

M. Serotonin

Serotonin, or 5-hydroxytryptamine (5-HT), plays a major role in the CNS and is an important regulator of adult neurogenesis (Banasr et al., 2004Malberg, 2004Warner-Schmidt and Duman, 2006). Reduced 5-HT function resulting in impaired adult neurogenesis has been hypothesized to be involved in major depression (Duman et al., 2000Jacobs et al., 2000Kempermann, 2002). Serotonergic neurons innervate the entire CNS, originating from cells in the raphe nuclei (Lorez and Richards, 1982) and the reticular formation of the brain stem. At least 15 subtypes of serotonin receptors exist, including 5-HT1–5 and 5-HT7. Within each receptor type, there are subcategories of receptors. Except for the 5-HT3 receptor, all of the 5-HT receptors are GPCRs. The roles of the 5-HT1A and 5-HT2A/2C receptor subtypes in adult neurogenesis are the most understood.

Selective serotonin reuptake inhibitors (SSRIs) increase 5-HT levels by blocking its reuptake at the presynaptic serotonergic nerve terminal. Long-term SSRI treatment, which is commonly used to decrease symptoms of depression and other mood disorders, has been shown to increase cell proliferation in both the adult rat hippocampus in vivo and in neural cultures in vitro (Malberg et al., 2000Lee et al., 2001Manev et al., 2001). The SSRI-induced proliferative effect seems to be mediated predominantly through activation of 5-HT1A receptors (Santarelli et al., 2003), in a postsynaptic effect (Huang and Herbert, 2005). Administration of a partial 5-HT1A agonist will also increase adult neurogenesis (Grabiec et al., 2009) through a similar mechanism. In contrast, 5-HT1A receptor blockade reduces cell proliferation in the dentate gyrus (Radley and Jacobs, 2002). However, no change in cell proliferation is seen in 5-HT1A receptor-KO mice (Santarelli et al., 2003).

Lesions to serotonergic inputs result in decreased adult neurogenesis in both the SGZ and SVZ (Brezun and Daszuta, 1999), which can be functionally restored by the transplantation of serotonergic tissue to the lesion site (Brezun and Daszuta, 2000a,b). Serotonin depletion also reduces survival and proliferation in cultured neurospheres derived from adult mouse hippocampus (Benninghoff et al., 2010).

Evidence suggests that 5-HT2 receptors are also involved in regulating neurogenesis, especially in the SVZ. In addition to 5-HT1A receptor activation, 5-HT2 receptor agonists enhance cell proliferation in the SZV (Banasr et al., 2004). Oddly, 5-HT2A/2C antagonists have also been found to increase adult neurogenesis in the SVZ (Wang et al., 2004; Green et al., 2006). A previous study found that fluoxetine enhances the division of early progenitor (TAP) cells but not stem-like cells (Encinas et al., 2006). It has been discovered that serotonin may exert differential effects on neural stem/progenitor cells (Soumier et al., 2010) via opposing actions mediated through different 5-HT receptors during different stages.

In the adult hippocampus, long-term 5-HT1A receptor activation and 5-HT2A blockade increased proliferation, whereas long-term 5-HT2 receptor stimulation reduced proliferation. Long-term 5-HT2receptor stimulation increased survival and enhanced maturation (Klempin et al., 2010). Other studies have shown that long-term 5-HT2A/2C blockade increases the proliferation of progenitor cells in the adult hippocampus (Jha et al., 2008). It is noteworthy that 5-HT4 receptor agonists have been discovered to increase adult neurogenesis in a much shorter time frame than SSRIs, possibly by increasing serotonergic drive and enhancing cAMP-responsive element-binding protein phosphorylation (Lucas et al., 2007). The overall effect of the various 5-HT receptor-mediated actions of serotonin is increased adult neurogenesis. Because of its critical role in depression and other mood disorders, how and to what extent serotonin affects adult neurogenesis is an important question that warrants extensive further investigation.

N. Wnt

Wnt proteins bind to receptors of the atypical GPCR frizzled and low-density lysoprotein-related protein5/6 families on the cell surface. To date, at least 17 human genes have been cloned for Wnt signaling glycoproteins: Wnt1, -2, -2B, -3, -3A, -4, -5A, -5B, -6, -7A, -7B, -8A, -8B, -9A, -9B, -10A, -10B, -11, and -16 (Katoh and Katoh, 2005). Wnt binding to its receptors induces the nuclear translocation of β-catenin that acts as a transcriptional cofactor of lymphoid enhancer binding factor 1 and modifies gene transcription (canonical Wnt pathway). Wnt can also signal in noncanonical pathways independent of β-catenin, leading to small G-protein activation and cytoskeletal changes or Ca2+signaling with activation of heterotrimeric G-proteins, causing various cell responses (Force et al., 2007).

Wnt receptors are implicated in adult hippocampal neurogenesis through the regulation of proliferation, migration, and differentiation (Zhou et al., 2004Lie et al., 2005Yu et al., 2006Wexler et al., 2009). The Wnt 3 receptor is expressed and the β-catenin pathway is active in the hippocampal SGZ. Overexpression of Wnt3 is sufficient to increase adult neurogenesis in the SGZ in vitro and in vivo. In contrast, blockade of Wnt signaling abolishes neurogenesis almost completely in vivo (Lie et al., 2005). Low-density lysoprotein-related protein 6 mutant mice exhibit reduced production and proliferation of dentate granule neurons and abnormalities in the radial glial scaffolding (Zhou et al., 2004). In a series of in vitro studies using adult mouse neurospheres, both Wnt3 and Wnt5a promoted neurogenesis, but only Wnt5a-mediated neurogenesis was blocked by PKC inhibition (Yu et al., 2006).

O. Intracellular Signals (Phospholipase C-β1, Phospholipase A2, Phosphodiesterase-4D)

Direct downstream second messengers and effectors that are normally associated with GPCR signaling may be able to mediate adult neurogenesis independently of any GPCR system. These signals include arachidonic acid, diacyl glycerol, inositol 1,4,5-triphosphate, and cAMP.

Enzymes that metabolize arachidonic acid, such as 5-lipoxygenase, cyclooxygenase-2, and prostaglandin E2, have been reported to increase proliferation of neuronal precursors in the adult SGZ when infused into the hippocampus (Uchida et al., 2002). Cyclooxygenase-2-deficient mice show lower numbers of proliferating cells, which are induced to proliferate in the postischemic SGZ (Sasaki et al., 2004). Several phospholipase A2 isoforms are neurotrophic for cerebellar granule neurons (Arioka et al., 2005) or regulate neurite outgrowth and survival in culture (Smalheiser et al., 1996Forlenza et al., 2007Masuda et al., 2008).

PLC is an enzyme that cleaves phosphatidylinositol 4,5-bisphosphate into diacyl glycerol and inositol 1,4,5-trisphosphate. PLC-β1(−/−) mice show increased adult hippocampal neurogenesis, increased adult granule cell density and survival, and stable proliferation rates (Manning et al., 2012). These mice also display aberrant migration of mature granule neurons within the GCL in adulthood, with excessive adult-generated mature neurons residing in the middle and outer GCL.

Mice deficient in phosphodiesterase-4D (PDE4D), an enzyme that catalyzes the hydrolysis of cAMP, display memory enhancement in radial arm maze, water maze, and object recognition tests, along with increased hippocampal neurogenesis. Microinfusion of lentiviral vectors that contained microRNAs targeting long-form PDE4D isoforms into bilateral dentate gyri of the mouse hippocampus down-regulate PDE4D4 and PDE4D5, enhance memory, and raise hippocampal neurogenesis (Li et al., 2011). Long-term PDE4 inhibition by rolipram treatment also increases adult neurogenesis, as evidenced by enhanced proliferation and survival of BrdU-positive neurons in the hippocampal dentate gyrus (Sasaki et al., 2007Li et al., 2009). This suggests that PDE4D normally acts to repress hippocampal neurogenesis.

VII. Implications of Adult Neurogenesis in Pathological Conditions

Adult neurogenesis is altered in brain injuries and several neurological diseases and disorders, including traumatic brain injury and ischemic stroke; neurodegenerative diseases such as Alzheimer’s, Huntington’s, and Parkinson’s diseases (Abdipranoto et al., 2008); demyelinating diseases such as multiple sclerosis; epilepsy and seizures; and psychiatric disorders such as depression and schizophrenia. A detailed review on neurogenesis in pathological states is found in Kempermann (2011a).

A. Adult Neurogenesis in Brain Injury

Traumatic brain injury (TBI) occurs when sudden trauma causes damage to the brain. It is the most common form of acquired brain injury in children, young adults, and older people. Recovery after TBI depends on the balance between neuronal injury and neuroregeneration. TBI typically causes neuronal loss and neurological deficits, especially in hippocampus-dependent cognitive functions. Accumulating evidence from animal studies of both focal and diffuse TBI suggests that neurogenesis is increased after TBI in both the SGZ of the dentate gyrus (Kernie et al., 2001Yu et al., 2008Zheng et al., 2011) and the SVZ of the lateral ventricles (Bye et al., 2011). However, it remains to be demonstrated conclusively whether neurogenesis contributes to recovery after TBI.

Stroke is the leading cause of disability and death in humans and is the result of ischemia, blockage, or hemorrhage to the brain. Numerous studies indicate that focal or global ischemia potently stimulate neurogenesis in adult rodents (Liu et al., 1998Jin et al., 2001Kee et al., 2001) and monkeys (Tonchev et al., 2005Koketsu et al., 2006). In contrast, subarachnoid hemorrhage has been shown to be associated with a decrease in cell proliferation in the SGZ and SVZ (Mino et al., 2003). Although functional neurogenesis has been demonstrated in the hippocampus (van Praag et al., 2002) and newly born neurons in the SVZ have been shown to migrate to the damaged cortex (Arvidsson et al., 2002), the role of ischemia-induced neurogenesis in recovery after an ischemic insult remains unclear.

B. Adult Neurogenesis in Neurodegenerative Diseases

 

1. Alzheimer’s Disease.

Alzheimer’s disease (AD) is a progressive neurodegenerative disease of the CNS and the most common cause of dementia in older people. AD is characterized by the accumulation of plaques and tangles in the brain, particularly in the hippocampus. The plaques contain insoluble deposits of amyloid-β-peptide, whereas the neurofibrillary tangles are composed of aggregates of the hyperphosphorylated microtubule-associated protein τ. The cognitive deficits observed in AD correlate most closely with the τ pathology (Selkoe, 2003Zhang et al., 2012). The role of neurogenesis in AD is unclear and confusing. Increased hippocampal neurogenesis is observed in some animal models of AD, including the Sorl1 KO mouse (Rohe et al., 2008) and mutated APP23 mouse model (Mirochnic et al., 2009). DCX is expressed in neurons in the SGZ of the dentate gyrus; consistent with this, postmortem samples taken from patients with AD possess more DCX-positive cells, suggesting that adult neurogenesis is increased (Jin et al., 2004a). In contrast, adult neurogenesis is reduced in double transgenic mice for mutated APP and presenilin-1 (Niidome et al., 2008) and triple transgenic mice for APP, presilin-1, and τ (Rodríguez et al., 2008). Furthermore, hippocampal volume is reduced in people with AD (Roh et al., 2011). Surprisingly, regulation of neurogenesis seems to remain intact in these models of AD. Reduced overall cholinergic signaling is a well known hallmark of AD, and loss of cholinergic drive has been shown to reduce adult neurogenesis in both the hippocampus and olfactory bulb (Cooper-Kuhn et al., 2004). Conversely, treatment with cholinergic drugs, which increase cholinergic signaling, has been shown to stimulate neurogenesis (Mohapel et al., 2005Kotani et al., 2006).

 

2. Huntington’s Disease.

Huntington’s disease (HD) is a progressive neurodegenerative genetic disorder of the CNS. Initially, HD affects muscle coordination; eventually, however, it leads to cognitive decline and dementia. It is caused by an autosomal-dominant mutation on the Huntingtin (HTT) gene, resulting in an expansion of a CAG triplet repeat stretch and alteration in the subsequent Htt protein. The altered Htt protein causes gradual pathological changes to the brain, giving rise to a wide spectrum of movement, cognitive, and psychiatric signs and symptoms. In genetic animal models of HD, such as the R6 transgenic mouse, adult neurogenesis is reduced (Lazic et al., 2004Gil et al., 2005). In contrast, increased cell proliferation and neurogenesis is observed in postmortem HD human brain tissue (Curtis et al., 2003).

 

3. Parkinson’s Disease.

Parkinson’s disease is a progressive degenerative disorder of the CNS that affects movement. It results from the death of dopaminergic neurons in the substantial nigra. Signs and symptoms of PD include tremor, bradykinesia, muscle rigidity, impaired posture, speech changes, and eventually dementia. PD is characterized by the accumulation of α-synuclein aggregates in the form of Lewy bodies. There is little evidence supporting active neurogenesis in the adult substantia nigra. Furthermore, there does not seem to be reactive neurogenesis in either the 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) or 6-OHDA animal models of PD. One study reported increased BrdU labeling in the substantia nigra after MPTP treatment as well as data supporting neurogenesis in the normal adult substantial nigra (Zhao et al., 2003), whereas another study found no new neurons in the substantia nigra of rodents injected with 6-OHDA (Frielingsdorf et al., 2004). However, after dopaminergic lesions with 6-OHDA or MPTP, the number of intrinsic tyrosine hydroxylase-positive (dopamine, NE) neurons increases in the striatum, and there is a transient elevation of adult neurogenesis in the hippocampus (Winner et al., 2009Park and Enikolopov, 2010). Recent studies suggest that upon injury, reactive astrocytes promote adult DA neurogenesis through the Wnt1/β-catenin signaling pathway (L’Episcopo et al., 20112012). In contrast, decreased neurogenesis has been shown in both the SGZ and SVZ in PD (Höglinger et al., 2004).

 

4. Amyotrophic Lateral Sclerosis.

Amyotrophic lateral sclerosis (ALS) is a fatal neurodegenerative disease characterized by rapidly progressive muscle weakness, disability, and death. It is caused by the degeneration of lower motor neurons located in the spinal cord and the upper motor/cortical neurons in the brainstem and cortex. In mice carrying a mutation of superoxide dismutase 1, a mutation found only in rare familial patients with ALS, increased progenitor cell proliferation, migration, and neurogenesis has been reported (Chi et al., 2006). Currently, there is no effective treatment for ALS.

Adult neurogenesis and neural plasticity is modulated in several neurodegenerative diseases. It is noteworthy that genes often involved in neurodegenerative diseases such as PD (α-synuclein), AD (presenilin-1, τ), and HD (huntingtin) also seem to play important roles in regulating adult neurogenesis (for review, see Winner et al., 2011).

C. Adult Neurogenesis in Demyelinating Disease

Multiple sclerosis (MS) is an inflammatory autoimmune disease characterized by demyelination and axonal scarring. This leads to less effective communication between nerve cells in the brain and spinal cord. MS occurs most often in young adults and women (Compston and Coles, 2008) and is often associated with periods of worsening (relapsing) and improvement (remitting), as some remyelination takes place. Unfortunately, oligodendrocytes, the cells responsible for the making and maintaining of myelin in the CNS, cannot completely rebuild the sheath, because MS destroys oligodendrocytes. In the adult brain, oligodendrocytes arise from a subset of neuron-glial antigen 2-expressing perinatal OPCs (Menn et al., 2006). Normally, few new oligodendrocytes are generated in the adult brain (McCarthy and Leblond, 1988Ehninger and Kempermann, 2003); however, after demyelination injury, some neuron-glial antigen 2 cells will give rise to oligodendrocytes (Redwine and Armstrong, 1998Nait-Oumesmar et al., 1999Levine et al., 2001). Furthermore, increased neurogenesis is observed in long-term lesions of MS (Chang et al., 2008). As the disease advances, the remyelination becomes less effective (Wolswijk, 1998), and neurologic function progressively deteriorates. Patients with MS typically live 5 to 10 years fewer than those without the disease. A recent report suggests that activation of gliogenesis in the SVZ occurs in MS and that SVZ-derived early glial progenitors may give rise to OPCs (Nait-Oumesmar et al., 2007). These early glial progenitors could be a potential therapeutic target for strategies to promoting myelin repair in MS.

D. Adult Neurogenesis in Epilepsy and Seizures

Epilepsy, characterized by recurrent seizures, is the third most common neurological disorder after stroke and AD. Epileptic seizures result from abnormal synchronous neuronal activity in the brain (Fisher et al., 2005). Temporal lobe epilepsy (TLE) is the most common form of epilepsy in adult humans and is often associated with hippocampal lesions. In animal models of epilepsy, adult neurogenesis is profoundly enhanced in both the SGZ and SVZ (Parent et al., 1998Ferland et al., 2002;Jessberger et al., 2007a). Precursor cell proliferation and adult neurogenesis may increase as much as 10-fold after short-term seizures. The magnitude of the increase depends on the severity of the seizures; mild seizures cause a smaller increase than severe (Yang et al., 2008a). Evidence of increased cell proliferation has also been found in tissue from children with hippocampal sclerosis and seizures (Takei et al., 2007). However, chronic TLE is associated with decreased adult hippocampal neurogenesis (Hattiangady et al., 2004). In addition, the seizure-induced increase in neurogenesis seems to decline with age and can become completely absent in old age (Rao et al., 2008Shetty et al., 2011). In fact, short-term seizures in old age may lead to marked neuronal loss, a greater chance of developing chronic TLE, and diminished cognitive function (Hattiangady et al., 2011).

Seizures, in addition to enhancing neurogenesis, seem to abnormally affect neuronal development, polarity, migration, and integration (Jessberger et al., 2007bKraev et al., 2009). Ectopic neurons are found in the hilus and cornu ammonis 3 regions in rodent model of epilepsy (Parent et al., 1999) as well as in tissue from humans with chronic TLE (Thom et al., 2002). Many of these ectopic neurons seem to have incorrectly integrated into the network, because they can be activated by cornu ammonis 3 pyramidal cell firing but not by granule cells (Scharfman et al., 2000McCloskey et al., 2006). It has been hypothesized that this seizure-induced aberrant neurogenesis may actually contribute to the development of epilepsy (Parent, 2007Hattiangady et al., 2008Zhao and Overstreet, 2008). Clearly, the role of neurogenesis in seizures and epilepsy merits further investigation.

E. Adult Neurogenesis in Psychiatric Disorders

 

1. Major Depression.

Major depressive disorder (MDD) is a psychiatric disorder characterized by generalized low mood, low self-esteem, and a loss of interest or pleasure in activities that are usually enjoyable. It is the most prevalent mental health disorder in the United States. The lifetime risk of MDD is almost 25%, and lifetime risk of suicide among untreated patients is nearly 20%; approximately 10% of Americans suffer depression in a given year (Kessler and Wang, 2009).

The idea that depression is due to impaired adult neurogenesis was first proposed in 2000 (Jacobs et al., 2000). A plethora of experimental data has been found that supports the neurogenesis hypothesis of depression; however, an equally compelling amount of evidence exists against the theory (for review, seeKempermann et al., 2008Samuels and Hen, 2011). The hippocampi of patients with chronic MDD are often smaller (Sheline et al., 1996Sheline, 2000). Stress and cortisol levels, which are often elevated in people with MDD, reduce adult hippocampal neurogenesis (Malberg and Duman, 2003). Antidepressants such as the SSRIs increase cell proliferation in the adult hippocampus (Malberg et al., 2000Santarelli et al., 2003Encinas et al., 2006Pinnock et al., 2009). In addition, the negative effects of stress on adult hippocampal neurogenesis are reversed by the SSRI fluoxetine (Malberg and Duman, 2003). Conversely, the antidepressive action of fluoxetine is removed when adult hippocampal neurogenesis is abolished (Santarelli et al., 2003). More recent studies suggest that these proneurogenic effects are species-dependent (Holick et al., 2008) as well as dependent on the particular antidepressant used (Surget et al., 2008). It is now thought that fluoxetine’s antipressant effect possesses both neurogenesis-dependent and -independent components (David et al., 2009). It is noteworthy that several proneurogenic compounds have been found to produce a positive response in tests of anxiety or depression, including FGF2 (Perez et al., 2009) and Notch1 (Guo et al., 2009). Environmental enrichment and exercise, which increase adult hippocampal neurogenesis, are known to improve mood (Van Praag et al., 1999Veena et al., 2009). Current knowledge suggests that impaired adult neurogenesis plays an important role in MDD, although MDD is not a purely neurogenesis-based disorder (Kempermann, 2011a).

 

2. Schizophrenia.

Schizophrenia is a severe chronic disease affecting higher cognitive functions. It is characterized by progressive disintegration of thought processes and emotional responsiveness and often manifests in some combination of hallucinations, paranoia, and disordered speech, thinking, and behavior. The onset of schizophrenia typically occurs in young adults, with lifetime prevalence of schizophrenia at almost 1%. The ability to function normally progressively deteriorates, resulting in significant social and occupational dysfunction. As a result of increased health problems and a very high suicide rate of nearly 5%, patients with schizophrenia typically live 12 to 15 years fewer than unaffected people (van Os and Kapur, 2009).

Decreased adult neurogenesis has been hypothesized to be a contributory factor in schizophrenia (for review, see Kempermann et al., 2008). Hippocampal cell proliferation is decreased in postmortem tissue of patients with schizophrenia (Reif et al., 2006). Furthermore, whole-brain and hippocampal volume is smaller in people with schizophrenia (Steen et al., 2006).

The cause of schizophrenia is unknown, but several genes linked to schizophrenia have also been associated with neurogenesis (Le Strat et al., 2009). The Disc 1 (disrupted in schizophrenia) gene was found in a Scottish family with a high incidence of schizophrenia as well as major depression (Millar et al., 2000). It is highly expressed in neurogenic areas and seems to play an important role in the development and integration of new neurons (Duan et al., 2007) through the AKT-mammalian target of rapamycin pathway (Kim et al., 2009). Disc 1 has also been shown to regulate neurogenesis through comodulation of Wnt-glycogen synthase kinase 3β/β-catenin signaling (Mao et al., 2009). Mutations of Disc 1 or glycogen synthase kinase 3β impair adult neurogenesis (Eom and Jope, 2009).

VIII. Adult Neural Stem Cell Therapy in the Central Nervous System

Adult stem cells possess two major defining properties, the ability to self-renew and the ability to differentiate into multiple specialized cell types. In the mammalian CNS, adult NSCs are able to generate new functional neurons and glia. Although active adult neurogenesis seems to be mostly restricted to the neurogenic SGZ and SVZ regions in the brain, it is exhibited in other areas of the CNS (for review, seeKempermann, 2011b). Adult neurogenesis plays an active and important role in maintaining normal homeostatic processes and contributes to the plasticity of the CNS (for review, see Lledo et al., 2006;Ming and Song, 2011).

The role of adult neurogenesis in neurological disorders and diseases is still poorly characterized and is currently a major research focus. Adult neurogenesis may help to repair the CNS in response to brain injury or disease. In general, acute brain damage tends to transiently increased cell proliferation, whereas chronic neuropathology eventually results in decreased adult neurogenesis. In rodents, NSCs are involved in replacing neurons dying from direct injury (Magavi et al., 2000) or ischemic stroke (Arvidsson et al., 2002Nakatomi et al., 2002). Furthermore, it seems that cell proliferation and neurogenesis initially increase at sites of damage in many seizure models (Bengzon et al., 1997Parent et al., 1998Jessberger et al., 2005) and neurodegenerative diseases such AD (Jin et al., 2004aVan Kampen and Eckman, 2006López-Toledano and Shelanski, 2007) and HD (Curtis et al., 2003). Increased neuroplasticity, a known function of adult neurogenesis, is also observed in many human neurodegenerative disorders (Kempermann, 2011a). NSC-based therapies hold great promise for treating many of the aforementioned neurological disorders. The two major strategies currently used in NSC-based therapy include transplantation of exogenous neural stem/progenitor cells and stimulation of endogenous neural stem/progenitor cells (Taupin, 2006).

Transplantation of fetal neuronal precursor cells has shown some success in animal models of spinal cord injury (McDonald et al., 1999), TBI (Riess et al., 2002), PD (Bjorklund et al., 2002), and spinal muscular atrophy (Corti et al., 2010). However, similar neurotransplantation therapies in humans with PD or HD have yielded mixed results. Whereas some patients with PD exhibited remarkable improvements (Wenning et al., 1997), others showed no benefits (Olanow et al., 2003), and some suffered additional side effects (Hagell et al., 2002). Clinical trials in patients with HD have also encountered problems, as some neurotransplants displayed HD-like neuronal degeneration (Cicchetti et al., 2009), whereas one patient developed mass lesions (Keene et al., 2009).

Adult neuronal precursor cells are also used for transplantation therapy. Adult spinal cord stem cells transplanted locally into the adult dentate gyrus have been shown to generate new neurons in rats (Shihabuddin et al., 2000). Furthermore, neurospheres derived from adult neural stem/progenitor cells administered either intravenously or intrathecally resulted in remyelination and functional recovery in a mouse model of MS (Pluchino et al., 2003).

There are many challenges and technical issues that must be overcome before neurotransplantation of adult or embryonic derived stem/progenitor cells can proceed much further in humans (for review, seeFeng and Gao, 2012). Challenges associated with transplanting these cells into the CNS include improper delivery and migration, mass effect and cell death, cell manufacturing problems, tumor formation, and adverse immune system interactions (Chiu and Rao, 2011).

Stimulation of endogenous neural stem/progenitor cells affords another potential strategy for treating neurological disorders and diseases. Although adult neurogenesis is mostly restricted to the neurogenic SGZ and SVZ regions in the brain, neural stem/progenitor cells seem to reside throughout the CNS (Kempermann, 2011b). The recruitment and activation of these endogenous cells would represent a strategy to promote regeneration and repair of the damaged or diseased CNS.

There are multiple methods to stimulate endogenous neurogenesis for the replacement of lost cells that have not yet been explored. For example, stimulating NSCs in the SVZ to replicate and migrate to the sites of damage and/or degeneration via the rostral migratory stream and possibly the corpus callosum fiber tract (Osman et al., 2011). In theory, this could be accomplished either by direct or indirect stimulation of the appropriate neural stem/progenitor cells. Direct stimulation could include the application of an agent that directly binds to and stimulates the NSC (perhaps through a GPCR-mediated system). Indirect stimulation could involve an agent that produces its effect by causing and/or increasing the release of some other endogenous agent (possibly by inhibiting an enzyme or a neurotransmitter reuptake mechanism), which in turn directly stimulates the NSC. Proneurogenic endogenous therapeutic strategies are currently being pursued by a number of pharmaceutical corporations and biotech companies.

IX. Therapeutic Potential of G-Protein-Coupled Receptor-Based Neural Stem Cell Strategies

Numerous compounds have been found to be proneurogenic. Antidepressant SSRIs are probably the best known, most studied, and some of the few proneurogenic agents used clinically (serotonin, NE, and dopamine reuptake inhibitors, or a combination, are all probably proneurogenic). SSRIs are indirect proneurogenic compounds that produce their effect by increasing 5-HT levels, which in turn stimulate neurogenesis via the activation of 5-HT receptors. In addition to the 5-HT receptors and other GPCR neurotransmitter systems mentioned in this review (section VI), a variety of other agents can enhance adult neurogenesis, including nitric oxide donors, acetylcholinesterase inhibitors, statins, N-methyl-d-aspartate receptor blockers, anti-inflammatory agents, and sex hormones (for review, see Jang et al., 2008Kempermann, 2011c). The major disadvantage with most of these compunds, including the SSRIs, is a lack of specificity for their proneurogenic effects.

Research efforts have recently been undertaken to discover more specific targets and compounds for proneurogenic drug development. Some of these efforts have focused on identifying GPCRs that selectively stimulate NCS activity and/or proliferation of NPCs and differentiation into particular neuronal cell types. These attributes would prove valuable for both stimulation of in vivo endogenous adult neurogenesis and in vitro culture of cells for exogenous transplantation therapy. To date, most research on GPCR-based NSC therapies has focused on stimulating endogenous adult neurogenesis in vivo, either alone or in combination with another agent. Using a combination of in vitro screening, genomics, and bioassays, followed by in vivo neurogenesis and functional behavioral assays, several small-molecule compounds have recently been discovered that interact with GPCRs to stimulate NSC activity. GPCR-mediated proneurogenic drug strategies currently under investigation and/or clinical development include 5-HT receptor agents, α1A-AR compounds, CB2 receptor agonists, DOR peptides, melatonergic agents, mGlu2/3 receptor antagonists, and S1P or LPA receptor compounds. Long-term α1A-AR stimulation improves learning and memory, reduces anxiety and depression, and increases lifespan in mice (Doze et al., 20092011Perez and Doze, 2011). Early preclinical and phase 1 trials with 5-HT receptor agents and mGlu2/3 antagonists have shown promise. The major neurological disorders targeted by GPCR-based NSC therapies thus far include psychiatric disorders, such as anxiety and depression, and select neurodegenerative diseases, including AD, HD, PD, and ALS.

There is a pressing need for better treatments for many neurological disorders, particularly the currently incurable CNS diseases. Because adult neurogenesis seems to be modifiable, stimulation of endogenous neurogenic processes, and/or the use of transplanted exogenous NSCs, has been postulated as a possible therapeutic strategy for several neurological disorders. Hopes have been raised by the prospect of NSC therapy. However, despite intense research efforts, successful NSC therapy has remained elusive. Developing effective strategies for isolating, enriching, and propagating NSCs for transplantation has proven problematic. Furthermore, identifying the molecules and cellular mechanisms required for the proper integration of transplanted stem cells into the injured brain has been difficult. As the processes that regulate adult neurogenesis are more fully understood, it is expected that the dream of functional NSC therapy will be reached. As demonstrated in this review, considerable information of the role of GPCRs in adult neurogenesis has been obtained. Using this knowledge, GPCR-based NSC therapeutic strategies may soon be realized.

X. Concluding Remarks and Future Directions

Many tissues of the body can repair and self-renew, largely because of stem cells and the various mechanisms that regulate their behavior. The concept that the brain does not regenerate was one of the most recent bastions of science shattered. Thus, the science of NSCs is a relatively new field. Although advances are slowly being made, there is a huge potential to ameliorate neurological diseases and counteract the aging process. The major weakness in the scientific evidence for adult neurogenesis is the lack of human research. Although evidence in rats and mice is encouraging, its application to human diseases is not assured. In particular, clinical trials will need to be performed to determine whether the production of new neurons can play therapeutic roles. Given the inherent difficulties of measuring neurogenesis in humans, the identification of molecular correlates and the signal transduction pathways involved in adult neurogenesis may prove to be the best course for future research.

The addition of new neurons to the existing circuitry represents a prominent form of structural plasticity and may contribute to specific brain functions in the adult brain. Research over the past decade has established a close relationship between adult neurogenesis and neuronal activity. Various environmental cues, exercise, learning and cognition-related stimuli, and emotional activities can influence neurogenic activity. With the addition of pharmaceuticals as suggested through GPCR ligands, we may be able to enhance our cognitive functions as well as treat neurological diseases.

Acknowledgments

This work was supported in part by the National Science Foundation [Grant 0347259]; the National Institutes of Health National Center for Research Resources [Grant P20-RR016471]; the National Heart, Lung and Blood Institute [Grant R01-HL098279]; and a Grant in Aid from the American Heart Association, Great Rivers Affiliate. We thank Katie Collette, Haley Amoth, and Elizabeth Sandquist for assistance in editing the manuscript.

Authorship Contributions

Wrote or contributed to the writing of the manuscript: Doze and Perez.

1Abbreviations:

5-HT
5-hydroxytryptamine
6-OHDA
6-hydroxydopamine
AD
Alzheimer’s disease
ALS
amyotrophic lateral sclerosis
AR
adrenergic receptor
BrdU
bromodeoxyuridine
CB
cannabinoid
CBD
cannabidiol
CNS
central nervous system
CNTF
ciliary neurotrophic factor
CSF
cerebrospinal fluid
DCX
doublecortin
DOR
δ-opioid receptor
DSP-4
N-(2-chloroethyl)-N-ethyl-2-bromo benzylamine hydrochloride
EGF
epidermal growth factor
ERK
extracellular signal-regulated kinase
FGF
fibroblast growth factor
GCL
granule cell layer
GFAP
glial fibrillary acidic protein
GFP
green fluorescent protein
GPCR
G-protein-coupled receptor
h
human
KO
knockout
KOR
κ-opioid receptor
LPA
lysophosphatidic acid
MDD
major depressive disorder
mGluR
metabotropic glutamate receptor
MGS0039
(1R,2R,3R,5R,6R)-2-amino-3-(3,4-dichlorobenzyloxy)-6-fluorobicyclo[3.1.0]hexane-2,6-dicarboxylic acid
miRNA
microRNA
MOR
μ-opioid receptor
MPTP
1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine
MRS 2179
2′-deoxy-N6-methyladenosine 3′,5′-bisphosphate
MS
multiple sclerosis
NCAM
neural cell adhesion molecule
NE
norepinephrine
NPY
neuropeptide Y
NTPDase
nucleoside triphosphate diphosphatase
OPC
oligodendrocyte progenitor cell
PACAP
pituitary adenylate cyclase-activating polypeptide
PD
Parkinson’s disease
PDE
phosphodiesterase
PI3K
phosphoinositide 3-kinase
PKC
protein kinase C
PLC
phospholipase C
S1P
sphingosine-1-phosphate
SDF-1
stromal cell-derived factor 1
SGZ
subgranular zone
SKF 38393
2,3,4,5-tetrahydro-7,8-dihydroxy-1-phenyl-1H-3-benzazepine
SNC80
(+)-4-[(αR)-α-(2S,5R)-4-allyl-2,5-dimethyl-1-piperazinyl)-3-methoxybenzyl]-N,N-diethylbenzamide
Sox2
Sex-determining region Y-box 2
SSRI
selective serotonin reuptake inhibitor
SVZ
subventricular zone
TAP
transient amplifying progenitor
TBI
traumatic brain injury
THC
Δ9-tetrahydrocannabinol
TLE
temporal lobe epilepsy
U50488H
trans-(±)-3,4-dichloro-N-methyl-N-[2-(1-pyrrolidinyl)cyclohexyl]benzeneacetamide
VIP
vasoactive intestinal polypeptide
VPAC
vasoactive intestinal polypeptide receptor
VPC32183
(S)-phosphoric acid mono-{2-octadec-9-enoylamino-3-[4-(pyridin-2-ylmethoxy)-phenyl]-propyl} ester
WT
wild type.

 

References

  • Abbott NJ, Rönnbäck L, Hansson E. (2006) Astrocyte-endothelial interactions at the blood-brain barrier. Nat Rev Neurosci 7:41–53. [PubMed]
  • Abbracchio MP, Burnstock G, Boeynaems JM, Barnard EA, Boyer JL, Kennedy C, Knight GE, Fumagalli M, Gachet C, Jacobson KA, et al. (2006) International Union of Pharmacology. LVIII: Update on the P2Y G protein-coupled nucleotide receptors: from molecular mechanisms and pathophysiology to therapy. Pharmacol Rev 58:281–341. [PMC free article] [PubMed]
  • Abdipranoto A, Wu S, Stayte S, Vissel B. (2008) The role of neurogenesis in neurodegenerative diseases and its implications for therapeutic development. CNS Neurol Disord Drug Targets7:187–210. [PubMed]
  • Abrous DN, Adriani W, Montaron MF, Aurousseau C, Rougon G, Le Moal M, Piazza PV. (2002)Nicotine self-administration impairs hippocampal plasticity. J Neurosci 22:3656–3662.[PubMed]
  • Agasse F, Bernardino L, Kristiansen H, Christiansen SH, Ferreira R, Silva B, Grade S, Woldbye DP, Malva JO. (2008) Neuropeptide Y promotes neurogenesis in murine subventricular zone.Stem Cells 26:1636–1645. [PubMed]
  • Ago Y, Yoneyama M, Ishihama T, Kataoka S, Kawada K, Tanaka T, Ogita K, Shintani N, Hashimoto H, Baba A, et al. (2011) Role of endogenous pituitary adenylate cyclase-activating polypeptide in adult hippocampal neurogenesis. Neuroscience 172:554–561. [PubMed]
  • Aguado T, Monory K, Palazuelos J, Stella N, Cravatt B, Lutz B, Marsicano G, Kokaia Z, Guzmán M, Galve-Roperh I. (2005) The endocannabinoid system drives neural progenitor proliferation.FASEB J 19:1704–1706. [PubMed]
  • Aguado T, Palazuelos J, Monory K, Stella N, Cravatt B, Lutz B, Marsicano G, Kokaia Z, Guzmán M, Galve-Roperh I. (2006) The endocannabinoid system promotes astroglial differentiation by acting on neural progenitor cells. J Neurosci 26:1551–1561. [PubMed]
  • Ahlquist RP. (1948) A study of the adrenotropic receptors. Am J Physiol 1153:586–600.[PubMed]
  • Ahn S, Joyner AL. (2005) In vivo analysis of quiescent adult neural stem cells responding to Sonic hedgehog. Nature 437:894–897. [PubMed]
  • Altman J. (1962) Are new neurons formed in the brains of adult mammals? Science 135:1127–1128. [PubMed]
  • Alvarez-Buylla A. (1997) Mechanism of migration of olfactory bulb interneurons. Semin Cell Dev Biol 8:207–213. [PubMed]
  • Ando K, Tagawa T, Ishikawa K, Takamura H, Yasuzumi F. (1986) A comparative study of the innervation of the choroid plexus in amphibia. Experientia 42:394–398. [PubMed]
  • Arévalo-Martín A, García-Ovejero D, Rubio-Araiz A, Gómez O, Molina-Holgado F, Molina-Holgado E. (2007) Cannabinoids modulate Olig2 and polysialylated neural cell adhesion molecule expression in the subventricular zone of post-natal rats through cannabinoid receptor 1 and cannabinoid receptor 2. Eur J Neurosci 26:1548–1559. [PubMed]
  • Arguello AA, Fischer SJ, Schonborn JR, Markus RW, Brekken RA, Eisch AJ. (2009) Effect of chronic morphine on the dentate gyrus neurogenic microenvironment. Neuroscience 159:1003–1010. [PMC free article] [PubMed]
  • Arguello AA, Harburg GC, Schonborn JR, Mandyam CD, Yamaguchi M, Eisch AJ. (2008) Time course of morphine’s effects on adult hippocampal subgranular zone reveals preferential inhibition of cells in S phase of the cell cycle and a subpopulation of immature neurons.Neuroscience 157:70–79. [PMC free article] [PubMed]
  • Arioka M, Cheon SH, Ikeno Y, Nakashima S, Kitamoto K. (2005) A novel neurotrophic role of secretory phospholipases A2 for cerebellar granule neurons. FEBS Lett 579:2693–2701.[PubMed]
  • Arvidsson A, Collin T, Kirik D, Kokaia Z, Lindvall O. (2002) Neuronal replacement from endogenous precursors in the adult brain after stroke. Nat Med 8:963–970. [PubMed]
  • Arvidsson U, Riedl M, Chakrabarti S, Lee JH, Nakano AH, Dado RJ, Loh HH, Law PY, Wessendorf MW, Elde R. (1995) Distribution and targeting of a mu-opioid receptor (MOR1) in brain and spinal cord. J Neurosci 15:3328–3341. [PubMed]
  • Ashton JC, Appleton I, Darlington CL, Smith PF. (2004) Cannabinoid CB1 receptor protein expression in the rat choroid plexus: a possible involvement of cannabinoids in the regulation of cerebrospinal fluid. Neurosci Lett 364:40–42. [PubMed]
  • Azari H, Rahman M, Sharififar S, Reynolds BA. (2010) Isolation and expansion of the adult mouse neural stem cells using the neurosphere assay. J Vis Exp (45):2393. [PMC free article][PubMed]
  • Babu H, Claasen JH, Kannan S, Rünker AE, Palmer T, Kempermann G. (2011) A protocol for isolation and enriched monolayer cultivation of neural precursor cells from mouse dentate gyrus. Front Neurosci 5:89. [PMC free article] [PubMed]
  • Bagri A, Gurney T, He X, Zou YR, Littman DR, Tessier-Lavigne M, Pleasure SJ. (2002) The chemokine SDF1 regulates migration of dentate granule cells. Development 129:4249–4260.[PubMed]
  • Baker SA, Baker KA, Hagg T. (2004) Dopaminergic nigrostriatal projections regulate neural precursor proliferation in the adult mouse subventricular zone. Eur J Neurosci 20:575–579.[PubMed]
  • Baloyannis SJ. (2005) Morphological and morphometric alterations of Cajal-Retzius cells in early cases of Alzheimer’s disease: a Golgi and electron microscope study. Int J Neurosci115:965–980. [PubMed]
  • Banasr M, Hery M, Printemps R, Daszuta A. (2004) Serotonin-induced increases in adult cell proliferation and neurogenesis are mediated through different and common 5-HT receptor subtypes in the dentate gyrus and the subventricular zone. Neuropsychopharmacology 29:450–460. [PubMed]
  • Banisadr G, Skrzydelski D, Kitabgi P, Rostène W, Parsadaniantz SM. (2003) Highly regionalized distribution of stromal cell-derived factor-1/CXCL12 in adult rat brain: constitutive expression in cholinergic, dopaminergic and vasopressinergic neurons. Eur J Neurosci 18:1593–1606.[PubMed]
  • Barkho BZ, Munoz AE, Li X, Li L, Cunningham LA, Zhao X. (2008) Endogenous matrix metalloproteinase (MMP)-3 and MMP-9 promote the differentiation and migration of adult neural progenitor cells in response to chemokines. Stem Cells 26:3139–3149. [PMC free article][PubMed]
  • Bartolome JV, Lorber BA, Bartolome MB. (1994) Brain cholecystokinin and beta-endorphin systems may antagonistically interact to regulate tissue DNA synthesis in rat pups. Brain Res661:19–24. [PubMed]
  • Baskys A, Bayazitov I, Fang L, Blaabjerg M, Poulsen FR, Zimmer J. (2005) Group I metabotropic glutamate receptors reduce excitotoxic injury and may facilitate neurogenesis.Neuropharmacology 49 (Suppl 1):146–156. [PubMed]
  • Bauer S, Moyse E, Jourdan F, Colpaert F, Martel JC, Marien M. (2003) Effects of the alpha 2-adrenoreceptor antagonist dexefaroxan on neurogenesis in the olfactory bulb of the adult rat in vivo: selective protection against neuronal death. Neuroscience 117:281–291. [PubMed]
  • Baumann N, Pham-Dinh D. (2001) Biology of oligodendrocyte and myelin in the mammalian central nervous system. Physiol Rev 81:871–927. [PubMed]
  • Bédard A, Gravel C, Parent A. (2006) Chemical characterization of newly generated neurons in the striatum of adult primates. Exp Brain Res 170:501–512. [PubMed]
  • Bengzon J, Kokaia Z, Elmér E, Nanobashvili A, Kokaia M, Lindvall O. (1997) Apoptosis and proliferation of dentate gyrus neurons after single and intermittent limbic seizures. Proc Natl Acad Sci USA 94:10432–10437. [PMC free article] [PubMed]
  • Benninghoff J, Gritti A, Rizzi M, Lamorte G, Schloesser RJ, Schmitt A, Robel S, Genius J, Moessner R, Riederer P, et al. (2010) Serotonin depletion hampers survival and proliferation in neurospheres derived from adult neural stem cells. Neuropsychopharmacology 35:893–903.[PMC free article] [PubMed]
  • Bhattacharyya BJ, Banisadr G, Jung H, Ren D, Cronshaw DG, Zou Y, Miller RJ. (2008) The chemokine stromal cell-derived factor-1 regulates GABAergic inputs to neural progenitors in the postnatal dentate gyrus. J Neurosci 28:6720–6730. [PMC free article] [PubMed]
  • Bielle F, Griveau A, Narboux-Nême N, Vigneau S, Sigrist M, Arber S, Wassef M, Pierani A. (2005) Multiple origins of Cajal-Retzius cells at the borders of the developing pallium. Nat Neurosci 8:1002–1012. [PubMed]
  • Bjorklund LM, Sánchez-Pernaute R, Chung S, Andersson T, Chen IY, McNaught KS, Brownell AL, Jenkins BG, Wahlestedt C, Kim KS, et al. (2002) Embryonic stem cells develop into functional dopaminergic neurons after transplantation in a Parkinson rat model. Proc Natl Acad Sci USA 99:2344–2349. [PMC free article] [PubMed]
  • Blondel O, Collin C, McCarran WJ, Zhu S, Zamostiano R, Gozes I, Brenneman DE, McKay RD. (2000) A glia-derived signal regulating neuronal differentiation. J Neurosci 20:8012–8020.[PubMed]
  • Bonecchi R, Galliera E, Borroni EM, Corsi MM, Locati M, Mantovani A. (2009) Chemokines and chemokine receptors: an overview. Front Biosci 14:540–551. [PubMed]
  • Bonfanti L, Theodosis DT. (1994) Expression of polysialylated neural cell adhesion molecule by proliferating cells in the subependymal layer of the adult rat, in its rostral extension and in the olfactory bulb. Neuroscience 62:291–305. [PubMed]
  • Brezun JM, Daszuta A. (1999) Depletion in serotonin decreases neurogenesis in the dentate gyrus and the subventricular zone of adult rats. Neuroscience 89:999–1002. [PubMed]
  • Brezun JM, Daszuta A. (2000a) Serotonergic reinnervation reverses lesion-induced decreases in PSA-NCAM labeling and proliferation of hippocampal cells in adult rats. Hippocampus 10:37–46. [PubMed]
  • Brezun JM, Daszuta A. (2000b) Serotonin may stimulate granule cell proliferation in the adult hippocampus, as observed in rats grafted with foetal raphe neurons. Eur J Neurosci 12:391–396. [PubMed]
  • Brown SG, King BF, Kim YC, Jang SY, Burnstock G, Jacobson KA. (2000) Activity of novel adenine nucleotide derivatives as agonists and antagonists at recombinant rat P2X receptors.Drug Dev Res 49:253–259. [PMC free article] [PubMed]
  • Brunet A, Bonni A, Zigmond MJ, Lin MZ, Juo P, Hu LS, Anderson MJ, Arden KC, Blenis J, Greenberg ME. (1999) AKT promotes cell survival by phosphorylating and inhibiting a Forkhead transcription factor. Cell 96:857–868. [PubMed]
  • Brunet A, Park J, Tran H, Hu LS, Hemmings BA, Greenberg ME. (2001) Protein kinase SGK mediates survival signals by phosphorylating the forkhead transcription factor FKHRL1 (FOXO3a). Mol Cell Biol 21:952–965. [PMC free article] [PubMed]
  • Bye N, Carron S, Han X, Agyapomaa D, Ng SY, Yan E, Rosenfeld JV, Morganti-Kossmann MC. (2011) Neurogenesis and glial proliferation are stimulated following diffuse traumatic brain injury in adult rats. J Neurosci Res 89:986–1000. [PubMed]
  • Bylund DB. (2005) Adrenergic receptors: historical perspectives from the 20th century, in The Adrenergic Receptors in the 21st Century (Perez DM, editor. ed) pp 3–21, Humana Press, New Jersey.
  • Caballero S, Swaney J, Moreno K, Afzal A, Kielczewski J, Stoller G, Cavalli A, Garland W, Hansen G, Sabbadini R, et al. (2009) Anti-sphingosine-1-phosphate monoclonal antibodies inhibit angiogenesis and sub-retinal fibrosis in a murine model of laser-induced choroidal neovascularization. Exp Eye Res 88:367–377. [PMC free article] [PubMed]
  • Cameron HA, McKay RD. (2001) Adult neurogenesis produces a large pool of new granule cells in the dentate gyrus. J Comp Neurol 435:406–417. [PubMed]
  • 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 56:337–344. [PubMed]
  • Campbell NR, Fernandes CC, John D, Lozada AF, Berg DK. (2011) Nicotinic control of adult-born neuron fate. Biochem Pharmacol 82:820–827. [PMC free article] [PubMed]
  • Canudas AM, Di Giorgi-Gerevini V, Iacovelli L, Nano G, D’Onofrio M, Arcella A, Giangaspero F, Busceti C, Ricci-Vitiani L, Battaglia G, et al. (2004) PHCCC, a specific enhancer of type 4 metabotropic glutamate receptors, reduces proliferation and promotes differentiation of cerebellar granule cell neuroprecursors. J Neurosci 24:10343–10352. [PubMed]
  • Cardoso A, Freitas-da-Costa P, Carvalho LS, Lukoyanov NV. (2010) Seizure-induced changes in neuropeptide Y-containing cortical neurons: Potential role for seizure threshold and epileptogenesis. Epilepsy Behav 19:559–567. [PubMed]
  • Chang A, Smith MC, Yin X, Fox RJ, Staugaitis SM, Trapp BD. (2008) Neurogenesis in the chronic lesions of multiple sclerosis. Brain 131:2366–2375. [PMC free article] [PubMed]
  • Chen J, Zacharek A, Zhang C, Jiang H, Li Y, Roberts C, Lu M, Kapke A, Chopp M. (2005)Endothelial nitric oxide synthase regulates brain-derived neurotrophic factor expression and neurogenesis after stroke in mice. J Neurosci 25:2366–2375. [PMC free article] [PubMed]
  • Chen WH, Tzeng SF. (2005) Pituitary adenylate cyclase-activating polypeptide prevents cell death in the spinal cord with traumatic injury. Neurosci Lett 384:117–121. [PubMed]
  • Cheng LC, Pastrana E, Tavazoie M, Doetsch F. (2009) miR-124 regulates adult neurogenesis in the subventricular zone stem cell niche. Nat Neurosci 12:399–408. [PMC free article] [PubMed]
  • Chi L, Ke Y, Luo C, Li B, Gozal D, Kalyanaraman B, Liu R. (2006) Motor neuron degeneration promotes neural progenitor cell proliferation, migration, and neurogenesis in the spinal cords of amyotrophic lateral sclerosis mice. Stem Cells 24:34–43. [PMC free article] [PubMed]
  • Chiu AY, Rao MS. (2011) Cell-based therapy for neural disorders—anticipating challenges.Neurotherapeutics 8:744–752. [PMC free article] [PubMed]
  • Chun J, Goetzl EJ, Hla T, Igarashi Y, Lynch KR, Moolenaar W, Pyne S, Tigyi G. (2002)International Union of Pharmacology. XXXIV. Lysophospholipid receptor nomenclature.Pharmacol Rev 54:265–269. [PubMed]
  • Cicchetti F, Saporta S, Hauser RA, Parent M, Saint-Pierre M, Sanberg PR, Li XJ, Parker JR, Chu Y, Mufson EJ, et al. (2009) Neural transplants in patients with Huntington’s disease undergo disease-like neuronal degeneration. Proc Natl Acad Sci USA 106:12483–12488.[PMC free article] [PubMed]
  • Ciceroni C, Mosillo P, Mastrantoni E, Sale P, Ricci-Vitiani L, Biagioni F, Stocchi F, Nicoletti F, Melchiorri D. (2010) mGlu3 metabotropic glutamate receptors modulate the differentiation of SVZ-derived neural stem cells towards the astrocytic lineage. Glia 58:813–822. [PubMed]
  • Clarke SR, Shetty AK, Bradley JL, Turner DA. (1994) Reactive astrocytes express the embryonic intermediate neurofilament nestin. Neuroreport 5:1885–1888. [PubMed]
  • Coelho RP, Payne SG, Bittman R, Spiegel S, Sato-Bigbee C. (2007) The immunomodulator FTY720 has a direct cytoprotective effect in oligodendrocyte progenitors. J Pharmacol Exp Ther323:626–635. [PubMed]
  • Collingridge GL, Lester RA. (1989) Excitatory amino acid receptors in the vertebrate central nervous system. Pharmacol Rev 41:143–210. [PubMed]
  • Compston A, Coles A. (2008) Multiple sclerosis. Lancet 372:1502–1517. [PubMed]
  • Conn PJ, Sanders-Bush E, Hoffman BJ, Hartig PR. (1986) A unique serotonin receptor in choroid plexus is linked to phosphatidylinositol turnover. Proc Natl Acad Sci USA 83:4086–4088. [PMC free article] [PubMed]
  • Conti L, Cattaneo E. (2010) Neural stem cell systems: physiological players or in vitro entities?Nat Rev Neurosci 11:176–187. [PubMed]
  • Cooper-Kuhn CM, Kuhn HG. (2002) Is it all DNA repair? Methodological considerations for detecting neurogenesis in the adult brain. Brain Res Dev Brain Res 134:13–21. [PubMed]
  • Cooper-Kuhn CM, Winkler J, Kuhn HG. (2004) Decreased neurogenesis after cholinergic forebrain lesion in the adult rat. J Neurosci Res 77:155–165. [PubMed]
  • Contos JJ, Fukushima N, Weiner JA, Kaushal D, Chun J. (2000) Requirement for the lpA1 lysophosphatidic acid receptor gene in normal suckling behavior. Proc Natl Acad Sci USA97:13384–13389. [PMC free article] [PubMed]
  • Coronas V, Bantubungi K, Fombonne J, Krantic S, Schiffmann SN, Roger M. (2004) Dopamine D3 receptor stimulation promotes the proliferation of cells derived from the post-natal subventricular zone. J Neurochem 91:1292–1301. [PubMed]
  • Corti S, Nizzardo M, Nardini M, Donadoni C, Salani S, Ronchi D, Simone C, Falcone M, Papadimitriou D, Locatelli F, et al. (2010) Embryonic stem cell-derived neural stem cells improve spinal muscular atrophy phenotype in mice. Brain 133:465–481. [PubMed]
  • Cremer H, Lange R, Christoph A, Plomann M, Vopper G, Roes J, Brown R, Baldwin S, Kraemer P, Scheff S. (1994) Inactivation of the N-CAM gene in mice results in size reduction of the olfactory bulb and deficits in spatial learning. Nature 367:455–459. [PubMed]
  • Curtis MA, Penney EB, Pearson AG, van Roon-Mom WM, Butterworth NJ, Dragunow M, Connor B, Faull RL. (2003) Increased cell proliferation and neurogenesis in the adult human Huntington’s disease brain. Proc Natl Acad Sci USA 100:9023–9027. [PMC free article][PubMed]
  • David DJ, Samuels BA, Rainer Q, Wang JW, Marsteller D, Mendez I, Drew M, Craig DA, Guiard BP, Guilloux JP, et al. (2009) Neurogenesis-dependent and -independent effects of fluoxetine in an animal model of anxiety/depression. Neuron 62:479–493. [PMC free article] [PubMed]
  • Dawson J, Hotchin N, Lax S, Rumsby M. (2003) Lysophosphatidic acid induces process retraction in CG-4 line oligodendrocytes and oligodendrocyte precursor cells but not in differentiated oligodendrocytes. J Neurochem 87:947–957. [PubMed]
  • Decressac M, Prestoz L, Veran J, Cantereau A, Jaber M, Gaillard A. (2009) Neuropeptide Y stimulates proliferation, migration and differentiation of neural precursors from the subventricular zone in adult mice. Neurobiol Dis 34:441–449. [PubMed]
  • Decressac M, Wright B, David B, Tyers P, Jaber M, Barker RA, Gaillard A. (2011) Exogenous neuropeptide Y promotes in vivo hippocampal neurogenesis. Hippocampus 21:233–238.[PubMed]
  • Deisseroth K, Malenka RC. (2005) GABA excitation in the adult brain: a mechanism for excitation-neurogenesis coupling. Neuron 47:775–777. [PubMed]
  • Deisseroth K, Singla S, Toda H, Monje M, Palmer TD, Malenka RC. (2004) Excitation-neurogenesis coupling in adult neural stem/progenitor cells. Neuron 42:535–552. [PubMed]
  • Delfs JM, Kong H, Mestek A, Chen Y, Yu L, Reisine T, Chesselet MF. (1994) Expression of mu opioid receptor mRNA in rat brain: an in situ hybridization study at the single cell level. J Comp Neurol 345:46–68. [PubMed]
  • Delic J, Zimmermann H. (2010) Nucleotides affect neurogenesis and dopaminergic differentiation of mouse fetal midbrain-derived neural precursor cells. Purinergic Signal 6:417–428. [PMC free article] [PubMed]
  • Deneen B, Ho R, Lukaszewicz A, Hochstim CJ, Gronostajski RM, Anderson DJ. (2006) The transcription factor NFIA controls the onset of gliogenesis in the developing spinal cord. Neuron52:953–968. [PubMed]
  • Dermietzel R, Gao Y, Scemes E, Vieira D, Urban M, Kremer M, Bennett MV, Spray DC. (2000)Connexin43 null mice reveal that astrocytes express multiple connexins. Brain Res Brain Res Rev 32:45–56. [PubMed]
  • Di Giorgi Gerevini VD, Caruso A, Cappuccio I, Ricci Vitiani L, Romeo S, Della Rocca C, Gradini R, Melchiorri D, Nicoletti F. (2004) The mGlu5 metabotropic glutamate receptor is expressed in zones of active neurogenesis of the embryonic and postnatal brain. Brain Res Dev Brain Res150:17–22. [PubMed]
  • Di Giorgi-Gerevini V, Melchiorri D, Battaglia G, Ricci-Vitiani L, Ciceroni C, Busceti CL, Biagioni F, Iacovelli L, Canudas AM, Parati E, et al. (2005) Endogenous activation of metabotropic glutamate receptors supports the proliferation and survival of neural progenitor cells. Cell Death Differ 12:1124–1133. [PubMed]
  • Diaz J, Ridray S, Mignon V, Griffon N, Schwartz JC, Sokoloff P. (1997) Selective expression of dopamine D3 receptor mRNA in proliferative zones during embryonic development of the rat brain. J Neurosci 17:4282–4292. [PubMed]
  • Doetsch F. (2003) A niche for adult neural stem cells. Curr Opin Genet Dev 13:543–550.[PubMed]
  • Doetsch F, Alvarez-Buylla A. (1996) Network of tangential pathways for neuronal migration in adult mammalian brain. Proc Natl Acad Sci USA 93:14895–14900. [PMC free article] [PubMed]
  • Doetsch F, Caillé I, Lim DA, García-Verdugo JM, Alvarez-Buylla A. (1999) Subventricular zone astrocytes are neural stem cells in the adult mammalian brain. Cell 97:703–716. [PubMed]
  • Domínguez-Escribà L, Hernández-Rabaza V, Soriano-Navarro M, Barcia JA, Romero FJ, García-Verdugo JM, Canales JJ. (2006) Chronic cocaine exposure impairs progenitor proliferation but spares survival and maturation of neural precursors in adult rat dentate gyrus.Eur J Neurosci 24:586–594. [PubMed]
  • Dottori M, Leung J, Turnley AM, Pébay A. (2008) Lysophosphatidic acid inhibits neuronal differentiation of neural stem/progenitor cells derived from human embryonic stem cells. Stem Cells 26:1146–1154. [PubMed]
  • Doze VA, Handel EM, Jensen KA, Darsie B, Luger EJ, Haselton JR, Talbot JN, Rorabaugh BR. (2009) α1A- and α1B-adrenergic receptors differentially modulate antidepressant-like behavior in the mouse. Brain Res 1285:148–157. [PMC free article] [PubMed]
  • Doze VA, Papay RS, Goldenstein BL, Gupta MK, Collette KM, Nelson BW, Lyons MJ, Davis BA, Luger EJ, Wood SG, et al. (2011) Long-term α1A-adrenergic receptor stimulation improves synaptic plasticity, cognitive function, mood, and longevity. Mol Pharmacol 80:747–758.[PMC free article] [PubMed]
  • Drake CT, Milner TA. (2002) Mu opioid receptors are in discrete hippocampal interneuron subpopulations. Hippocampus 12:119–136. [PubMed]
  • Duan X, Chang JH, Ge S, Faulkner RL, Kim JY, Kitabatake Y, Liu XB, Yang CH, Jordan JD, Ma DK, et al. (2007) Disrupted-In-Schizophrenia 1 regulates integration of newly generated neurons in the adult brain. Cell 130:1146–1158. [PMC free article] [PubMed]
  • Dubocovich ML, Rivera-Bermudez MA, Gerdin MJ, Masana MI. (2003) Molecular pharmacology, regulation and function of mammalian melatonin receptors. Front Biosci8:d1093–d1108. [PubMed]
  • Duman RS, Malberg J, Nakagawa S, D’Sa C. (2000) Neuronal plasticity and survival in mood disorders. Biol Psychiatry 48:732–739. [PubMed]
  • Dumont Y, Martel JC, Fournier A, St-Pierre S, Quirion R. (1992) Neuropeptide Y and neuropeptide Y receptor subtypes in brain and peripheral tissues. Prog Neurobiol 38:125–167.[PubMed]
  • Duval N, Gomès D, Calaora V, Calabrese A, Meda P, Bruzzone R. (2002) Cell coupling and Cx43 expression in embryonic mouse neural progenitor cells. J Cell Sci 115:3241–3251. [PubMed]
  • Edvinsson L, Nielsen KC, Owman CH, West KA. (1974) Adrenergic innervation of the mammalian choroid plexus. Am J Anat 139:299–307.
  • Ehninger D, Kempermann G. (2003) Regional effects of wheel running and environmental enrichment on cell genesis and microglia proliferation in the adult murine neocortex. Cereb Cortex 13:845–851. [PubMed]
  • Eisch AJ, Barrot M, Schad CA, Self DW, Nestler EJ. (2000) Opiates inhibit neurogenesis in the adult rat hippocampus. Proc Natl Acad Sci USA 97:7579–7584. [PMC free article] [PubMed]
  • Encinas JM, Vaahtokari A, Enikolopov G. (2006) Fluoxetine targets early progenitor cells in the adult brain. Proc Natl Acad Sci USA 103:8233–8238. [PMC free article] [PubMed]
  • Eom TY, Jope RS. (2009) Blocked inhibitory serine-phosphorylation of glycogen synthase kinase-3alpha/beta impairs in vivo neural precursor cell proliferation. Biol Psychiatry 66:494–502. [PMC free article] [PubMed]
  • Eriksson PS, Perfilieva E, Björk-Eriksson T, Alborn AM, Nordborg C, Peterson DA, Gage FH. (1998) Neurogenesis in the adult human hippocampus. Nat Med 4:1313–1317. [PubMed]
  • Estivill-Torrús G, Llebrez-Zayas P, Matas-Rico E, Santín L, Pedraza C, De Diego I, Del Arco I, Fernández-Llebrez P, Chun J, De Fonseca FR. (2008) Absence of LPA1 signaling results in defective cortical development. Cereb Cortex 18:938–950. [PubMed]
  • Falluel-Morel A, Chafai M, Vaudry D, Basille M, Cazillis M, Aubert N, Louiset E, de Jouffrey S, Le Bigot JF, Fournier A, et al. (2007) The neuropeptide pituitary adenylate cyclase-activating polypeptide exerts anti-apoptotic and differentiating effects during neurogenesis: focus on cerebellar granule neurones and embryonic stem cells. J Neuroendocrinol 19:321–327. [PubMed]
  • Feng Z, Gao F. (2012) Stem cell challenges in the treatment of neurodegenerative disease. CNS Neurosci Ther 18:142–148. [PubMed]
  • Feng YB, Yao H, Man X, Chi LY, Chi ZF. (2011) Effects of the group II mGlu receptor agonist 2R,4R-APDC on dentate gyrus cell proliferation in the adult rat brain after diffuse brain injury.Neurol Res 33:381–388. [PubMed]
  • Ferland RJ, Gross RA, Applegate CD. (2002) Increased mitotic activity in the dentate gyrus of the hippocampus of adult C57BL/6J mice exposed to the flurothyl kindling model of epileptogenesis. Neuroscience 115:669–683. [PubMed]
  • Fields H. (2004) State-dependent opioid control of pain. Nat Rev Neurosci 5:565–575.[PubMed]
  • Fisher RS, van Emde Boas W, Blume W, Elger C, Genton P, Lee P, Engel J., Jr (2005) Epileptic seizures and epilepsy: definitions proposed by the International League Against Epilepsy (ILAE) and the International Bureau for Epilepsy (IBE). Epilepsia 46:470–472. [PubMed]
  • Force T, Woulfe K, Koch WJ, Kerkelä R. (2007) Molecular scaffolds regulate bidirectional crosstalk between Wnt and classical seven-transmembrane-domain receptor signaling pathways. Sci STKE 397:pe41. [PubMed]
  • Forlenza OV, Mendes CT, Marie SK, Gattaz WF. (2007) Inhibition of phospholipase A2 reduces neurite outgrowth and neuronal viability. Prostaglandins Leukot Essent Fatty Acids 76:47–55.[PubMed]
  • Forsberg M, Carlén M, Meletis K, Yeung MS, Barnabé-Heider F, Persson MA, Aarum J, Frisén J. (2010) Efficient reprogramming of adult neural stem cells to monocytes by ectopic expression of a single gene. Proc Natl Acad Sci USA 107:14657–14661. [PMC free article] [PubMed]
  • Fredholm BB, Abbracchio MP, Burnstock G, Daly JW, Harden TK, Jacobson KA, Leff P, Williams M. (1994) Nomenclature and classification of purinoceptors. Pharmacol Rev 46:143–156. [PubMed]
  • Fredholm BB, IJzerman AP, Jacobson KA, Klotz KN, Linden J. (2001) International Union of Pharmacology. XXV. Nomenclature and classification of adenosine receptors. Pharmacol Rev53:527–552. [PubMed]
  • Fride E, Mechoulam R. (2003) New advances in the identification and physiological roles of the components of the endogenous cannabinoid system, in Molecular Biology of Drug Addiction(Maldonado R, editor. ed) pp 173–179, Humana Press, Totowa, NJ.
  • Frielingsdorf H, Schwarz K, Brundin P, Mohapel P. (2004) No evidence for new dopaminergic neurons in the adult mammalian substantia nigra. Proc Natl Acad Sci USA 101:10177–10182.[PMC free article] [PubMed]
  • Fritschy JM, Grzanna R. (1992) Restoration of ascending noradrenergic projections by residual locus coeruleus neurons: compensatory response to neurotoxin-induced cell death in the adult rat brain. J Comp Neurol 321:421–441. [PubMed]
  • Gage FH. (2000) Mammalian neural stem cells. Science 287:1433–1438. [PubMed]
  • Gage FH, Coates PW, Palmer TD, Kuhn HG, Fisher LJ, Suhonen JO, Peterson DA, Suhr ST, Ray J. (1995) Survival and differentiation of adult neuronal progenitor cells transplanted to the adult brain. Proc Natl Acad Sci USA 92:11879–11883. [PMC free article] [PubMed]
  • Galve-Roperh I, Aguado T, Palazuelos J, Guzmán M. (2008) Mechanisms of control of neuron survival by the endocannabinoid system. Curr Pharm Des 14:2279–2288. [PubMed]
  • Ge S, Goh EL, Sailor KA, Kitabatake Y, Ming GL, Song H. (2006) GABA regulates synaptic integration of newly generated neurons in the adult brain. Nature 439:589–593.[PMC free article] [PubMed]
  • Gillard SE, Tzaferis J, Tsui HC, Kingston AE. (2003) Expression of metabotropic glutamate receptors in rat meningeal and brain microvasculature and choroid plexus. J Comp Neurol461:317–332. [PubMed]
  • Gil JM, Mohapel P, Araújo IM, Popovic N, Li JY, Brundin P, Petersén A. (2005) Reduced hippocampal neurogenesis in R6/2 transgenic Huntington’s disease mice. Neurobiol Dis20:744–751. [PubMed]
  • Girouard H, Iadecola C. (2006) Neurovascular coupling in the normal brain and in hypertension, stroke, and Alzheimer disease. J Appl Physiol 100:328–335. [PubMed]
  • Goncalves MB, Suetterlin P, Yip P, Molina-Holgado F, Walker DJ, Oudin MJ, Zentar MP, Pollard S, Yáñez-Muñoz RJ, Williams G, et al. (2008) A diacylglycerol lipase-CB2 cannabinoid pathway regulates adult subventricular zone neurogenesis in an age-dependent manner. Mol Cell Neurosci 38:526–536. [PubMed]
  • Gong X, He X, Qi L, Zuo H, Xie Z. (2006) Stromal cell derived factor-1 acutely promotes neural progenitor cell proliferation in vitro by a mechanism involving the ERK1/2 and PI-3K signal pathways. Cell Biol Int 30:466–471. [PubMed]
  • Gonzalez BJ, Basille M, Vaudry D, Fournier A, Vaudry H. (1997) Pituitary adenylate cyclase-activating polypeptide promotes cell survival and neurite outgrowth in rat cerebellar neuroblasts. Neuroscience 78:419–430. [PubMed]
  • Gottlieb DI. (2002) Large-scale sources of neural stem cells. Annu Rev Neurosci 25:381–407.[PubMed]
  • Gould E, Reeves AJ, Graziano MS, Gross CG. (1999) Neurogenesis in the neocortex of adult primates. Science 286:548–552. [PubMed]
  • Gowran A, Noonan J, Campbell VA. (2011) The multiplicity of action of cannabinoids: implications for treating neurodegeneration. CNS Neurosci Ther 17:637–644. [PubMed]
  • Gozes I, Fridkin M, Brenneman DE. (1995) A VIP hybrid antagonist: from developmental neurobiology to clinical applications. Cell Mol Neurobiol 15:675–687. [PubMed]
  • Grabiec M, Turlejski K, Djavadian RL. (2009) The partial 5-HT1A receptor agonist buspirone enhances neurogenesis in the opossum (Monodelphis domestica). Eur Neuropsychopharmacol19:431–439. [PubMed]
  • Green AR. (2006) Neuropharmacology of 5-hydroxytryptamine. Br J Pharmacol 147 (Suppl 1):S145–S152. [PMC free article] [PubMed]
  • Gressens P, Paindaveine B, Hill JM, Evrard P, Brenneman DE. (1998) Vasoactive intestinal peptide shortens both G1 and S phases of neural cell cycle in whole postimplantation cultured mouse embryos. Eur J Neurosci 10:1734–1742. [PubMed]
  • Grimm I, Messemer N, Stanke M, Gachet C, Zimmermann H. (2009) Coordinate pathways for nucleotide and EGF signaling in cultured adult neural progenitor cells. J Cell Sci 122:2524–2533. [PubMed]
  • Grimm I, Ullsperger SN, Zimmermann H. (2010) Nucleotides and epidermal growth factor induce parallel cytoskeletal rearrangements and migration in cultured adult murine neural stem cells. Acta Physiol 199:181–189. [PubMed]
  • Gritti A, Frölichsthal-Schoeller P, Galli R, Parati EA, Cova L, Pagano SF, Bjornson CR, Vescovi AL. (1999) Epidermal and fibroblast growth factors behave as mitogenic regulators for a single multipotent stem cell-like population from the subventricular region of the adult mouse forebrain. J Neurosci 19:3287–3297. [PubMed]
  • Gould E, Reeves AJ, Graziano MS, Gross CG. (1999) Neurogenesis in the neocortex of adult primates. Science 286:548–552. [PubMed]
  • Guo M, Xu NJ, Li YT, Yang JY, Wu CF, Pei G. (2005) Morphine modulates glutamate release in the hippocampal CA1 area in mice. Neurosci Lett 381:12–15. [PubMed]
  • Guo YJ, Zhang ZJ, Wang SH, Sui YX, Sun Y. (2009) Notch1 signaling, hippocampal neurogenesis and behavioral responses to chronic unpredicted mild stress in adult ischemic rats.Prog Neuropsychopharmacol Biol Psychiatry 33:688–694. [PubMed]
  • Gupta MK, Papay RS, Jurgens CW, Gaivin RJ, Shi T, Doze VA, Perez DM. (2009) α1-Adrenergic receptors regulate neurogenesis and gliogenesis. Mol Pharmacol 76:314–326. [PMC free article][PubMed]
  • Guzman-Marin R, Suntsova N, Methippara M, Greiffenstein R, Szymusiak R, McGinty D. (2005) Sleep deprivation suppresses neurogenesis in the adult hippocampus of rats. Eur J Neurosci 22:2111–2116. [PubMed]
  • Hagell P, Piccini P, Björklund A, Brundin P, Rehncrona S, Widner H, Crabb L, Pavese N, Oertel WH, Quinn N, et al. (2002) Dyskinesias following neural transplantation in Parkinson’s disease.Nat Neurosci 5:627–628. [PubMed]
  • Hahn JW, Jagwani S, Kim E, Rendell VR, He J, Ezerskiy LA, Wesselschmidt R, Coscia CJ, Belcheva MM. (2010) Mu and kappa opioids modulate mouse embryonic stem cell-derived neural progenitor differentiation via MAP kinases. J Neurochem 112:1431–1441.[PMC free article] [PubMed]
  • Han Z, Zhong L, Maina N, Hu Z, Li X, Chouthai NS, Bischof D, Weigel-Van Aken KA, Slayton WB, Yoder MC, et al. (2008) Stable integration of recombinant adeno-associated virus vector genomes after transduction of murine hematopoietic stem cells. Hum Gene Ther 19:267–278.[PubMed]
  • Hansel DE, Eipper BA, Ronnett GV. (2001) Neuropeptide Y functions as a neuroproliferative factor. Nature 410:940–944. [PubMed]
  • Hara Y, Nomura T, Yoshizaki K, Frisén J, Osumi N. (2010) Impaired hippocampal neurogenesis and vascular formation in ephrin-A5-deficient mice. Stem Cells 28:974–983. [PubMed]
  • Harada J, Foley M, Moskowitz MA, Waeber C. (2004) Sphingosine-1-phosphate induces proliferation and morphological changes of neural progenitor cells. J Neurochem 88:1026–1039. [PubMed]
  • Harburg GC, Hall FS, Harrist AV, Sora I, Uhl GR, Eisch AJ. (2007) Knockout of the mu opioid receptor enhances the survival of adult-generated hippocampal granule cell neurons.Neuroscience 144:77–87. [PMC free article] [PubMed]
  • Hattiangady B, Kuruba R, Shetty AK. (2011) Acute Seizures in Old Age Leads to a Greater Loss of CA1 Pyramidal Neurons, an Increased Propensity for Developing Chronic TLE and a Severe Cognitive Dysfunction. Aging Dis 2:1–17. [PMC free article] [PubMed]
  • Hattiangady B, Rao MS, Shetty AK. (2004) Chronic temporal lobe epilepsy is associated with severely declined dentate neurogenesis in the adult hippocampus. Neurobiol Dis 17:473–490.[PubMed]
  • Hattiangady B, Rao MS, Shetty AK. (2008) Plasticity of hippocampal stem/progenitor cells to enhance neurogenesis in response to kainate-induced injury is lost by middle age. Aging Cell7:207–224. [PMC free article] [PubMed]
  • Hecht JH, Weiner JA, Post SR, Chun J. (1996) Ventricular zone gene-1 (vzg-1) encodes a lysophosphatidic acid receptor expressed in neurogenic regions of the developing cerebral cortex. J Cell Biol 135:1071–1083. [PMC free article] [PubMed]
  • Henry RA, Hughes SM, Connor B. (2007) AAV-mediated delivery of BNDF augments neurogenesis in the normal and quinolinic acid-lesioned adult rat brain. Eur J Neurosci25:3513–3525. [PubMed]
  • Hiramoto T, Ihara Y, Watanabe Y. (2006) α-1 Adrenergic receptors stimulation induces the proliferation of neural progenitor cells in vitro. Neurosci Lett 408:25–28. [PubMed]
  • Hiramoto T, Satoh Y, Takishima K, Watanabe Y. (2008) Induction of cell migration of neural progenitor cells in vitro by alpha-1 adrenergic receptor and dopamine D1 receptor stimulation.Neuroreport 19:793–797. [PubMed]
  • Holmes A, Lachowicz JE, Sibley DR. (2004) Phenotypic analysis of dopamine receptor knockout mice; recent insights into the functional specificity of dopamine receptor subtypes.Neuropharmacology 47:1117–1134. [PubMed]
  • Holmes MM, Galea LA, Mistlberger RE, Kempermann G. (2004) Adult hippocampal neurogenesis and voluntary running activity: circadian and dose-dependent effects. J Neurosci Res 76:216–222. [PubMed]
  • Höglinger GU, Rizk P, Muriel MP, Duyckaerts C, Oertel WH, Caille I, Hirsch EC. (2004)Dopamine depletion impairs precursor cell proliferation in Parkinson disease. Nat Neurosci7:726–735. [PubMed]
  • Holick KA, Lee DC, Hen R, Dulawa SC. (2008) Behavioral effects of chronic fluoxetine in BALB/cJ mice do not require adult hippocampal neurogenesis or the serotonin 1A receptor.Neuropsychopharmacology 33:406–417. [PubMed]
  • Howell OW, Doyle K, Goodman JH, Scharfman HE, Herzog H, Pringle A, Beck-Sickinger AG, Gray WP. (2005) Neuropeptide Y stimulates neuronal precursor proliferation in the post-natal and adult dentate gyrus. J Neurochem 93:560–570. [PubMed]
  • Howell OW, Scharfman HE, Herzog H, Sundstrom LE, Beck-Sickinger A, Gray WP. (2003)Neuropeptide Y is neuroproliferative for post-natal hippocampal precursor cells. J Neurochem86:646–659. [PubMed]
  • Howell OW, Silva S, Scharfman HE, Sosunov AA, Zaben M, Shatya A, McKhann G, 2nd, Herzog H, Laskowski A, Gray WP. (2007) Neuropeptide Y is important for basal and seizure-induced precursor cell proliferation in the hippocampus. Neurobiol Dis 26:174–188. [PubMed]
  • Huang GJ, Herbert J. (2005) The role of 5-HT1A receptors in the proliferation and survival of progenitor cells in the dentate gyrus of the adult hippocampus and their regulation by corticoids. Neuroscience 135:803–813. [PubMed]
  • Jacobs BL, van Praag H, Gage FH. (2000) Adult brain neurogenesis and psychiatry: a novel theory of depression. Mol Psychiatry 5:262–269. [PubMed]
  • Jaillard C, Harrison S, Stankoff B, Aigrot MS, Calver AR, Duddy G, Walsh FS, Pangalos MN, Arimura N, Kaibuchi K, et al. (2005) Edg8/S1P5: an oligodendroglial receptor with dual function on process retraction and cell survival. J Neurosci 25:1459–1469. [PubMed]
  • Jang MH, Sone H, Ming GL. (2008) Regulation of adult neurogenesis by neurotransmitters, inAdult Neurogenesis (Gage FH, Kempermann G, Song H, editors. eds) pp 397–423, Cold Spring Harbor Laboratory Press, Cold Spring Harbor, New York.
  • Jessberger S, Nakashima K, Clemenson GD, Jr, Mejia E, Mathews E, Ure K, Ogawa S, Sinton CM, Gage FH, Hsieh J. (2007a) Epigenetic modulation of seizure-induced neurogenesis and cognitive decline. J Neurosci 27:5967–5975. [PubMed]
  • Jessberger S, Römer B, Babu H, Kempermann G. (2005) Seizures induce proliferation and dispersion of doublecortin-positive hippocampal progenitor cells. Exp Neurol 196:342–351.[PubMed]
  • Jessberger S, Zhao C, Toni N, Clemenson GD, Jr, Li Y, Gage FH. (2007b) Seizure-associated, aberrant neurogenesis in adult rats characterized with retrovirus-mediated cell labeling. J Neurosci 27:9400–9407. [PubMed]
  • Jha S, Rajendran R, Fernandes KA, Vaidya VA. (2008) 5-HT2A/2C receptor blockade regulates progenitor cell proliferation in the adult rat hippocampus. Neurosci Lett 441:210–214. [PubMed]
  • Jhaveri DJ, Mackay EW, Hamlin AS, Marathe SV, Nandam LS, Vaidya VA, Bartlett PF. (2010)Norepinephrine directly activates adult hippocampal precursors via beta3-adrenergic receptors.J Neurosci 30:2795–2806. [PMC free article] [PubMed]
  • Jia C, Hegg CC. (2010) NPY mediates ATP-induced neuroproliferation in adult mouse olfactory epithelium. Neurobiol Dis 38:405–413. [PMC free article] [PubMed]
  • Jiang W, Zhang Y, Xiao L, Van Cleemput J, Ji SP, Bai G, Zhang X. (2005) Cannabinoids promote embryonic and adult hippocampus neurogenesis and produce anxiolytic- and antidepressant-like effects. J Clin Invest 115:3104–3116. [PMC free article] [PubMed]
  • Jin K, Minami M, Lan JQ, Mao XO, Batteur S, Simon RP, Greenberg DA. (2001) Neurogenesis in dentate subgranular zone and rostral subventricular zone after focal cerebral ischemia in the rat. Proc Natl Acad Sci USA 98:4710–4715. [PMC free article] [PubMed]
  • Jin K, Peel AL, Mao XO, Xie L, Cottrell BA, Henshall DC, Greenberg DA. (2004a) Increased hippocampal neurogenesis in Alzheimer’s disease. Proc Natl Acad Sci USA 101:343–347.[PMC free article] [PubMed]
  • Jin K, Xie L, Kim SH, Parmentier-Batteur S, Sun Y, Mao XO, Childs J, Greenberg DA. (2004b)Defective adult neurogenesis in CB1 cannabinoid receptor knockout mice. Mol Pharmacol66:204–208. [PubMed]
  • Johansson CB, Svensson M, Wallstedt L, Janson AM, Frisén J. (1999) Neural stem cells in the adult human brain. Exp Cell Res 253:733–736. [PubMed]
  • Johansson PA, Burnstock G, Dziegielewska KM, Guida E, McIntyre P, Saunders NR. (2007)Expression and localization of P2 nucleotide receptor subtypes during development of the lateral ventricular choroid plexus of the rat. Eur J Neurosci 25:3319–3331. [PubMed]
  • Juaneda C, Dumont Y, Chabot JG, Quirion R. (2001) Autoradiographic distribution of adrenomedullin receptors in the rat brain. Eur J Pharmacol 421:R1–R2. [PubMed]
  • Jukkola PI, Rogers JT, Kaspar BK, Weeber EJ, Nishijima I. (2011) Secretin deficiency causes impairment in survival of neural progenitor cells in mice. Hum Mol Genet 20:1000–1007.[PubMed]
  • Kahn L, Alonso G, Normand E, Manzoni OJ. (2005) Repeated morphine treatment alters polysialylated neural cell adhesion molecule, glutamate decarboxylase-67 expression and cell proliferation in the adult rat hippocampus. Eur J Neurosci 21:493–500. [PubMed]
  • Kamiji MM, Inui A. (2007) Neuropeptide Y receptor selective ligands in the treatment of obesity.Endocr Rev 28:664–684. [PubMed]
  • Kaneko N, Okano H, Sawamoto K. (2006) Role of the cholinergic system in regulating survival of newborn neurons in the adult mouse dentate gyrus and olfactory bulb. Genes Cells 11:1145–1159. [PubMed]
  • Kaneko Y, Sakakibara S, Imai T, Suzuki A, Nakamura Y, Sawamoto K, Ogawa Y, Toyama Y, Miyata T, Okano H. (2000) Musashi1: an evolutionally conserved marker for CNS progenitor cells including neural stem cells. Dev Neurosci 22:139–153. [PubMed]
  • Kärkkäinen V, Louhivuori V, Castrén ML, Akerman KE. (2009) Neurotransmitter responsiveness during early maturation of neural progenitor cells. Differentiation 77:188–198.[PubMed]
  • Kask A, Harro J, Tuomaine P, Rägo L, Männistö PT. (1997) Overflow of noradrenaline and dopamine in frontal cortex after [N-(2-chloroethyl)-N-ethyl-2-bromobenzylamine] (DSP-4) treatment: in vivo microdialysis study in anaesthetized rats. Naunyn Schmiedebergs Arch Pharmacol 355:267–272. [PubMed]
  • Katoh Y, Katoh M. (2005) Identification and characterization of rat Wnt6 and Wnt10a genes in silico. Int J Mol Med 15:527–531. [PubMed]
  • Katz SE, Klisovic DD, O’Dorisio MS, Lynch R, Lubow M. (2002) Expression of somatostatin receptors 1 and 2 in human choroid plexus and arachnoid granulations: implications for idiopathic intracranial hypertension. Arch Ophthalmol 120:1540–1543. [PubMed]
  • Kee NJ, Preston E, Wojtowicz JM. (2001) Enhanced neurogenesis after transient global ischemia in the dentate gyrus of the rat. Exp Brain Res 136:313–320. [PubMed]
  • Keene CD, Chang RC, Leverenz JB, Kopyov O, Perlman S, Hevner RF, Born DE, Bird TD, Montine TJ. (2009) A patient with Huntington’s disease and long-surviving fetal neural transplants that developed mass lesions. Acta Neuropathol 117:329–338. [PMC free article][PubMed]
  • Keller G, Kennedy M, Papayannopoulou T, Wiles MV. (1993) Hematopoietic commitment during embryonic stem cell differentiation in culture. Mol Cell Biol 13:473–486.[PMC free article] [PubMed]
  • Kelly S, Caldwell M, Keasey MP, Cooke JA, Uney JB. (2009) Identifying neural progenitor cells in the adult brain. Methods Mol Biol 549:217–230. [PubMed]
  • Kempermann G. (2002) Regulation of adult hippocampal neurogenesis – implications for novel theories of major depression. Bipolar Disord 4:17–33. [PubMed]
  • Kempermann G. (2011a) Medicine, in Adult Neurogenesis 2: Stem Cells and Neuronal Development in the Adult Brain, pp 518–587, Oxford University Press, New York.
  • Kempermann G. (2011b) Neurogenic and non-neurogenic regions, in Adult Neurogenesis 2: Stem Cells and Neuronal Development in the Adult Brain, pp 275–326, Oxford University Press, New York.
  • Kempermann G. (2011c) Regulation, in Adult Neurogenesis 2: Stem Cells and Neuronal Development in the Adult Brain, pp 327–438, Oxford University Press, New York.
  • Kempermann G, Krebs J, Fabel K. (2008) The contribution of failing adult hippocampal neurogenesis to psychiatric disorders. Curr Opin Psychiatry 21:290–295. [PubMed]
  • Kempermann G, Kuhn HG, Gage FH. (1997) More hippocampal neurons in adult mice living in an enriched environment. Nature 386:493–495. [PubMed]
  • Kennedy M, Keller GM. (2003) Hematopoietic commitment of ES cells in culture. Methods Enzymol 365:39–59. [PubMed]
  • Kernie SG, Erwin TM, Parada LF. (2001) Brain remodeling due to neuronal and astrocytic proliferation after controlled cortical injury in mice. J Neurosci Res 66:317–326. [PubMed]
  • Kessler RC, Wang PS. (2009) Epidemiology of depression, in Handbook of Depression (Gotlib IH, Hammen CL, editors. eds) 2nd ed, pp 5–22, Gilford Press, New York.
  • Khan MZ, Brandimarti R, Musser BJ, Resue DM, Fatatis A, Meucci O. (2003) The chemokine receptor CXCR4 regulates cell-cycle proteins in neurons. J Neurovirol 9:300–314.[PMC free article] [PubMed]
  • Khurshid N, Hameed LS, Mohanasundaram S, Iyengar S. (2010) Opioid modulation of cell proliferation in the ventricular zone of adult zebra finches (Taenopygia guttata). FASEB J24:3681–3695. [PubMed]
  • Kim E, Clark AL, Kiss A, Hahn JW, Wesselschmidt R, Coscia CJ, Belcheva MM. (2006a) Mu- and kappa-opioids induce the differentiation of embryonic stem cells to neural progenitors. J Biol Chem 281:33749–33760. [PMC free article] [PubMed]
  • Kim JY, Duan X, Liu CY, Jang MH, Guo JU, Pow-anpongkul N, Kang E, Song H, Ming GL. (2009) DISC1 regulates new neuron development in the adult brain via modulation of AKT-mTOR signaling through KIAA1212. Neuron 63:761–773. [PMC free article] [PubMed]
  • Kim MJ, Kim HK, Kim BS, Yim SV. (2004) Melatonin increases cell proliferation in the dentate gyrus of maternally separated rats. J Pineal Res 37:193–197. [PubMed]
  • Kim SH, Won SJ, Mao XO, Ledent C, Jin K, Greenberg DA. (2006b) Role for neuronal nitric-oxide synthase in cannabinoid-induced neurogenesis. J Pharmacol Exp Ther 319:150–154.[PubMed]
  • Kim Y, Wang WZ, Comte I, Pastrana E, Tran PB, Brown J, Miller RJ, Doetsch F, Molnár Z, Szele FG. (2010) Dopamine stimulation of postnatal murine subventricular zone neurogenesis via the D3 receptor. J Neurochem 114:750–760. [PMC free article] [PubMed]
  • Kingsbury MA, Rehen SK, Contos JJ, Higgins CM, Chun J. (2003) Non-proliferative effects of lysophosphatidic acid enhance cortical growth and folding. Nat Neurosci 6:1292–1299.[PubMed]
  • Kippin TE, Kapur S, van der Kooy D. (2005) Dopamine specifically inhibits forebrain neural stem cell proliferation, suggesting a novel effect of antipsychotic drugs. J Neurosci 25:5815–5823. [PubMed]
  • Kiyota T, Okuyama S, Swan RJ, Jacobsen MT, Gendelman HE, Ikezu T. (2010) CNS expression of anti-inflammatory cytokine interleukin-4 attenuates Alzheimer’s disease-like pathogenesis in APP+PS1 bigenic mice. FASEB J 24:3093–3102. [PMC free article] [PubMed]
  • Klempin F, Babu H, De Pietri Tonelli D, Alarcon E, Fabel K, Kempermann G. (2010)Oppositional effects of serotonin receptors 5-HT1a, 2, and 2c in the regulation of adult hippocampal neurogenesis. Front Mol Neurosci 3:14. [PMC free article] [PubMed]
  • Kobayashi H, Shiraishi S, Minami S, Yokoo H, Yanagita T, Saitoh T, Mohri M, Wada A. (2001)Adrenomedullin receptors in rat choroid plexus. Neurosci Lett 297:167–170. [PubMed]
  • Koketsu D, Furuichi Y, Maeda M, Matsuoka N, Miyamoto Y, Hisatsune T. (2006) Increased number of new neurons in the olfactory bulb and hippocampus of adult non-human primates after focal ischemia. Exp Neurol 199:92–102. [PubMed]
  • Kolodziej A, Schulz S, Guyon A, Wu DF, Pfeiffer M, Odemis V, Höllt V, Stumm R. (2008) Tonic activation of CXC chemokine receptor 4 in immature granule cells supports neurogenesis in the adult dentate gyrus. J Neurosci 28:4488–4500. [PubMed]
  • Komitova M, Eriksson PS. (2004) Sox-2 is expressed by neural progenitors and astroglia in the adult rat brain. Neurosci Lett 369:24–27. [PubMed]
  • Kornblum HI, Loughlin SE, Leslie FM. (1987) Effects of morphine on DNA synthesis in neonatal rat brain. Dev Brain Res 31:45–52. [PubMed]
  • Kornack DR, Rakic P. (1999) Continuation of neurogenesis in the hippocampus of the adult macaque monkey. Proc Natl Acad Sci USA 96:5768–5773. [PMC free article] [PubMed]
  • Kotani S, Yamauchi T, Teramoto T, Ogura H. (2006) Pharmacological evidence of cholinergic involvement in adult hippocampal neurogenesis in rats. Neuroscience 142:505–514. [PubMed]
  • Kouroupi G, Lavdas AA, Gaitanou M, Thomaidou D, Stylianopoulou F, Matsas R. (2010)Lentivirus-mediated expression of insulin-like growth factor-I promotes neural stem/precursor cell proliferation and enhances their potential to generate neurons. J Neurochem 115:460–474.[PubMed]
  • Kraev IV, Godukhin OV, Patrushev IV, Davies HA, Popov VI, Stewart MG. (2009) Partial kindling induces neurogenesis, activates astrocytes and alters synaptic morphology in the dentate gyrus of freely moving adult rats. Neuroscience 162:254–267. [PubMed]
  • Krathwohl MD, Kaiser JL. (2004) Chemokines promote quiescence and survival of human neural progenitor cells. Stem Cells 22:109–118. [PubMed]
  • Krivokuća D, Puskas L, Puskas N, Erić M. (2010) Morphometric characteristics of neuropeptide Y immunoreactive neurons in cortex of human inferior parietal lobule. Coll Antropol 34:99–104.[PubMed]
  • Kuhn HG, Dickinson-Anson H, Gage FH. (1996) Neurogenesis in the dentate gyrus of the adult rat: age-related decrease of neuronal progenitor proliferation. J Neurosci 16:2027–2033.[PubMed]
  • Kulkarni VA, Jha S, Vaidya VA. (2002) Depletion of norepinephrine decreases the proliferation, but does not influence the survival and differentiation, of granule cell progenitors in the adult rat hippocampus. Eur J Neurosci 16:2008–2012. [PubMed]
  • L’Episcopo F, Tirolo C, Testa N, Caniglia S, Morale MC, Cossetti C, D’Adamo P, Zardini E, Andreoni L, Ihekwaba AE, et al. (2011) Reactive astrocytes and Wnt/β-catenin signaling link nigrostriatal injury to repair in 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine model of Parkinson’s disease. Neurobiol Dis 41:508–527. [PMC free article] [PubMed]
  • L’Episcopo F, Tirolo C, Testa N, Caniglia S, Morale MC, Deleidi M, Serapide MF, Pluchino S, Marchetti B. (2012) Plasticity of subventricular zone neuroprogenitors in MTPT (1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine) mouse model of Parkinson’s disease involves cross talk between inflammatory and Wnt/β-catenin signaling pathways: functional consequences for neuroprotection and repair. J Neurosci 32:2062–2085. [PMC free article] [PubMed]
  • Le Strat Y, Ramoz N, Gorwood P. (2009) The role of genes involved in neuroplasticity and neurogenesis in the observation of a gene-environment interaction (GxE) in schizophrenia. Curr Mol Med 9:506–518. [PubMed]
  • Labasque M, Reiter E, Becamel C, Bockaert J, Marin P. (2008) Physical interaction of calmodulin with the 5-hydroxytryptamine2C receptor C-terminus is essential for G protein-independent, arrestin-dependent receptor signaling. Mol Biol Cell 19:4640–4650.[PMC free article] [PubMed]
  • Lagace DC, Benavides DR, Kansy JW, Mapelli M, Greengard P, Bibb JA, Eisch AJ. (2008) Cdk5 is essential for adult hippocampal neurogenesis. Proc Natl Acad Sci USA 105:18567–18571.[PMC free article] [PubMed]
  • Lagerström MC, Schiöth HB. (2008) Structural diversity of G protein-coupled receptors and significance for drug discovery. Nat Rev Drug Discov 7:339–357. [PubMed]
  • Landgren H, Curtis MA. (2011) Locating and labeling neural stem cells in the brain. J Cell Physiol 226:1–7. [PubMed]
  • Langer D, Ikehara Y, Takebayashi H, Hawkes R, Zimmermann H. (2007) The ectonucleotidases alkaline phosphatase and nucleoside triphosphate diphosphohydrolase 2 are associated with subsets of progenitor cell populations in the mouse embryonic, postnatal and adult neurogenic zones. Neuroscience 150:863–879. [PubMed]
  • Lazic SE, Grote H, Armstrong RJ, Blakemore C, Hannan AJ, van Dellen A, Barker RA. (2004)Decreased hippocampal cell proliferation in R6/1 Huntington’s mice. Neuroreport 15:811–813.[PubMed]
  • Lee M, Lelievre V, Zhao P, Torres M, Rodriguez W, Byun JY, Doshi S, Ioffe Y, Gupta G, de los Monteros AE, et al. (2001) Pituitary adenylyl cyclase-activating polypeptide stimulates DNA synthesis but delays maturation of oligodendrocyte progenitors. J Neurosci 21:3849–3859.[PubMed]
  • Lehmann K, Butz M, Teuchert-Noodt G. (2005) Offer and demand: proliferation and survival of neurons in the dentate gyrus. Eur J Neurosci 21:3205–3216. [PubMed]
  • Lei ZM, Rao CV. (2001) Neural actions of luteinizing hormone and human chorionic gonadotropin. Semin Reprod Med 19:103–109. [PubMed]
  • Lendahl U, Zimmerman LB, McKay RD. (1990) CNS stem cells express a new class of intermediate filament protein. Cell 60:585–595. [PubMed]
  • Leventhal C, Rafii S, Rafii D, Shahar A, Goldman SA. (1999) Endothelial trophic support of neuronal production and recruitment from the adult mammalian subependyma. Mol Cell Neurosci 13:450–464. [PubMed]
  • Levine JM, Reynolds R, Fawcett JW. (2001) The oligodendrocyte precursor cell in health and disease. Trends Neurosci 24:39–47. [PubMed]
  • Lewis PF, Emerman M. (1994) Passage through mitosis is required for oncoretroviruses but not for the human immunodeficiency virus. J Virol 68:510–516. [PMC free article] [PubMed]
  • Li YF, Cheng YF, Huang Y, Conti M, Wilson SP, O’Donnell JM, Zhang HT. (2011)Phosphodiesterase-4D knock-out and RNA interference-mediated knock-down enhance memory and increase hippocampal neurogenesis via increased cAMP signaling. J Neurosci31:172–183. [PMC free article] [PubMed]
  • Li YF, Huang Y, Amsdell SL, Xiao L, O’Donnell JM, Zhang HT. (2009) Antidepressant- and anxiolytic-like effects of the phosphodiesterase-4 inhibitor rolipram on behavior depend on cyclic AMP response element binding protein-mediated neurogenesis in the hippocampus.Neuropsychopharmacology 34:2404–2419. [PMC free article] [PubMed]
  • Lie DC, Colamarino SA, Song HJ, Désiré L, Mira H, Consiglio A, Lein ES, Jessberger S, Lansford H, Dearie AR, et al. (2005) Wnt signalling regulates adult hippocampal neurogenesis. Nature437:1370–1375. [PubMed]
  • Lledo PM, Alonso M, Grubb MS. (2006) Adult neurogenesis and functional plasticity in neuronal circuits. Nat Rev Neurosci 7:179–193. [PubMed]
  • Lima MM, Reksidler AB, Vital MA. (2009) The neurobiology of the substantia nigra pars compacta: from motor to sleep regulation. J Neural Transm Suppl 73:135–145. [PubMed]
  • Lin JH, Takano T, Arcuino G, Wang X, Hu F, Darzynkiewicz Z, Nunes M, Goldman SA, Nedergaard M. (2007) Purinergic signaling regulates neural progenitor cell expansion and neurogenesis. Dev Biol 302:356–366. [PMC free article] [PubMed]
  • Lindvall M, Alumets J, Edvinsson L, Fahrenkrug J, Håkanson R, Hanko J, Owman C, Schaffalitzky de Muckadell OB, Sundler F. (1978a) Peptidergic (VIP) nerves in the mammalian choroid plexus. Neurosci Lett 9:77–82. [PubMed]
  • Lindvall M, Edvinsson L, Owman C. (1978b) Sympathetic nervous control of cerebrospinal fluid production from the choroid plexus. Science 201:176–178. [PubMed]
  • Liu C, Zhao X. (2009) MicroRNAs in adult and embryonic neurogenesis. Neuromolecular Med11:141–152. [PMC free article] [PubMed]
  • Liu J, Solway K, Messing RO, Sharp FR. (1998) Increased neurogenesis in the dentate gyrus after transient global ischemia in gerbils. J Neurosci 18:7768–7778. [PubMed]
  • Lois C, Alvarez-Buylla A. (1994) Long distance neuronal migration in the adult mammalian brain. Science 264:1145–1148. [PubMed]
  • López-Toledano MA, Shelanski ML. (2007) Increased neurogenesis in young transgenic mice overexpressing human APP(Sw, Ind). J Alzheimers Dis 12:229–240. [PubMed]
  • Lorber BA, Freitag SK, Bartolome JV. (1990) Effects of beta-endorphin on DNA synthesis in brain regions of preweanling rats. Brain Res 531:329–332. [PubMed]
  • Lorez HP, Richards JG. (1982) Supra-ependymal serotoninergic nerves in mammalian brain: morphological, pharmacological and functional studies. Brain Res Bull 9:727–741. [PubMed]
  • Loy R, Koziell DA, Lindsey JD, Moore RY. (1980) Noradrenergic innervation of the adult rat hippocampal formation. J Comp Neurol 189:699–710. [PubMed]
  • Lu M, Grove EA, Miller RJ. (2002) Abnormal development of the hippocampal dentate gyrus in mice lacking the CXCR4 chemokine receptor. Proc Natl Acad Sci USA 99:7090–7095.[PMC free article] [PubMed]
  • Lucas G, Rymar VV, Du J, Mnie-Filali O, Bisgaard C, Manta S, Lambas-Senas L, Wiborg O, Haddjeri N, Piñeyro G, et al. (2007) Serotonin4 (5-HT4) receptor agonists are putative antidepressants with a rapid onset of action. Neuron 55:712–725. [PubMed]
  • Luttrell LM. (2008) Reviews in molecular biology and biotechnology: transmembrane signaling by G protein-coupled receptors. Mol Biotechnol 39:239–264. [PubMed]
  • Ma Q, Jones D, Borghesani PR, Segal RA, Nagasawa T, Kishimoto T, Bronson RT, Springer TA. (1998) Impaired B-lymphopoiesis, myelopoiesis, and derailed cerebellar neuron migration in CXCR4- and SDF-1-deficient mice. Proc Natl Acad Sci USA 95:9448–9453. [PMC free article][PubMed]
  • Ma W, Maric D, Li BS, Hu Q, Andreadis JD, Grant GM, Liu QY, Shaffer KM, Chang YH, Zhang L, et al. (2000) Acetylcholine stimulates cortical precursor cell proliferation in vitro via muscarinic receptor activation and MAP kinase phosphorylation. Eur J Neurosci 12:1227–1240.[PubMed]
  • MacNeil DJ, Douglas J. (2007) NPY Y1 and Y5 receptor selective antagonists as anti-obesity drugs. Curr Top Med Chem 7:1721–1733. [PubMed]
  • Magavi SS, Leavitt BR, Macklis JD. (2000) Induction of neurogenesis in the neocortex of adult mice. Nature 405:951–955. [PubMed]
  • Mahabeer R, Naidoo S, Raidoo DM. (2000) Detection of tissue kallikrein and kinin B1 and B2 receptor mRNAs in human brain by in situ RT-PCR. Metab Brain Dis 15:325–335. [PubMed]
  • Maison P, Walker DJ, Walsh FS, Williams G, Doherty P. (2009) BDNF regulates neuronal sensitivity to endocannabinoids. Neurosci Lett 467:90–94. [PubMed]
  • Malberg JE. (2004) Implications of adult hippocampal neurogenesis in antidepressant action. J Psychiatry Neurosci 29:196–205. [PMC free article] [PubMed]
  • Malberg JE, Duman RS. (2003) Cell proliferation in adult hippocampus is decreased by inescapable stress: reversal by fluoxetine treatment. Neuropsychopharmacology 28:1562–1571.[PubMed]
  • Malberg JE, Eisch AJ, Nestler EJ, Duman RS. (2000) Chronic antidepressant treatment increases neurogenesis in adult rat hippocampus. J Neurosci 20:9104–9110. [PubMed]
  • Manda K, Ueno M, Anzai K. (2009) Cranial irradiation-induced inhibition of neurogenesis in hippocampal dentate gyrus of adult mice: attenuation by melatonin pretreatment. J Pineal Res46:71–78. [PubMed]
  • Mandyam CD, Harburg GC, Eisch AJ. (2007) Determination of key aspects of precursor cell proliferation, cell cycle length and kinetics in the adult mouse subgranular zone. Neuroscience146:108–122. [PMC free article] [PubMed]
  • Mandyam CD, Norris RD, Eisch AJ. (2004) Chronic morphine induces premature mitosis of proliferating cells in the adult mouse subgranular zone. J Neurosci Res 76:783–794. [PubMed]
  • Manev H, Uz T, Smalheiser NR, Manev R. (2001) Antidepressants alter cell proliferation in the adult brain in vivo and in neural cultures in vitro. Eur J Pharmacol 411:67–70. [PubMed]
  • Manning EE, Ransome MI, Burrows EL, Hannan AJ. (2010) Increased adult hippocampal neurogenesis and abnormal migration of adult-born granule neurons is associated with hippocampal-specific cognitive deficits in phospholipase C-β1 knockout mice. Hippocampus22:309–319. [PubMed]
  • Mansour A, Fox CA, Akil H, Watson SJ. (1995a) Opioid-receptor mRNA expression in the rat CNS: anatomical and functional implications. Trends Neurosci 18:22–29. [PubMed]
  • Mansour A, Hoversten MT, Taylor LP, Watson SJ, Akil H. (1995b) The cloned mu, delta and kappa receptors and their endogenous ligands: evidence for two opioid peptide recognition cores. Brain Res 700:89–98. [PubMed]
  • Mao Y, Ge X, Frank CL, Madison JM, Koehler AN, Doud MK, Tassa C, Berry EM, Soda T, Singh KK, et al. (2009) Disrupted in schizophrenia 1 regulates neuronal progenitor proliferation via modulation of GSK3beta/beta-catenin signaling. Cell 136:1017–1031. [PMC free article][PubMed]
  • Marchalant Y, Brothers HM, Norman GJ, Karelina K, DeVries AC, Wenk GL. (2009a)Cannabinoids attenuate the effects of aging upon neuroinflammation and neurogenesis.Neurobiol Dis 34:300–307. [PubMed]
  • Marchalant Y, Brothers HM, Wenk GL. (2009b) Cannabinoid agonist WIN-55,212–2 partially restores neurogenesis in the aged rat brain. Mol Psychiatry 14:1068–1069. [PMC free article][PubMed]
  • Marchalant Y, Cerbai F, Brothers HM, Wenk GL. (2008) Cannabinoid receptor stimulation is anti-inflammatory and improves memory in old rats. Neurobiol Aging 29:1894–1901.[PMC free article] [PubMed]
  • Martin GR. (1981) Isolation of a pluripotent cell line from early mouse embryos cultured in medium conditioned by teratocarcinoma stem cells. Proc Natl Acad Sci USA 78:7634–7638.[PMC free article] [PubMed]
  • Maślińska D, Laure-Kamionowska M, Maśliński KT, Wojtecka-Łukasik E, Szukiewicz D, Maśliński (2009) Morphology and immuno-distribution of the histamine H4 receptor and histamine-releasing factor in choroid plexus of patients with paraneoplastic cerebellar degeneration. Inflamm Res 58 (Suppl 1):45–46. [PubMed]
  • Masuda S, Yamamoto K, Hirabayashi T, Ishikawa Y, Ishii T, Kudo I, Murakami M. (2008)Human group III secreted phospholipase A2 promotes neuronal outgrowth and survival.Biochem J 409:429–438. [PubMed]
  • Matas-Rico E, García-Diaz B, Llebrez-Zayas P, López-Barroso D, Santín L, Pedraza C, Smith-Fernández A, Fernández-Llebrez P, Tellez T, Redondo M, et al. (2008) Deletion of lysophosphatidic acid receptor LPA1 reduces neurogenesis in the mouse dentate gyrus. Mol Cell Neurosci 39:342–355. [PMC free article] [PubMed]
  • McCarthy GF, Leblond CP. (1988) Radioautographic evidence for slow astrocyte turnover and modest oligodendrocyte production in the corpus callosum of adult mice infused with 3H-thymidine. J Comp Neurol 271:589–603. [PubMed]
  • McCloskey DP, Hintz TM, Pierce JP, Scharfman HE. (2006) Stereological methods reveal the robust size and stability of ectopic hilar granule cells after pilocarpine-induced status epilepticus in the adult rat. Eur J Neurosci 24:2203–2210. [PubMed]
  • McDonald JW, Liu XZ, Qu Y, Liu S, Mickey SK, Turetsky D, Gottlieb DI, Choi DW. (1999)Transplanted embryonic stem cells survive, differentiate and promote recovery in injured rat spinal cord. Nat Med 5:1410–1412. [PubMed]
  • McGiffert C, Contos JJ, Friedman B, Chun J. (2002) Embryonic brain expression analysis of lysophospholipid receptor genes suggests roles for s1p(1) in neurogenesis and s1p(1–3) in angiogenesis. FEBS Lett 531:103–108. [PubMed]
  • Meibach RC, Maayani S. (1980) Localization of naloxone-sensitive [3H]dihydromorphine binding sites within the hippocampus of the rat. Eur J Pharmacol 68:175–179. [PubMed]
  • Melchiorri D, Cappuccio I, Ciceroni C, Spinsanti P, Mosillo P, Sarichelou I, Sale P, Nicoletti F. (2007) Metabotropic glutamate receptors in stem/progenitor cells. Neuropharmacology53:473–480. [PubMed]
  • Menn B, Garcia-Verdugo JM, Yaschine C, Gonzalez-Perez O, Rowitch D, Alvarez-Buylla A. (2006) Origin of oligodendrocytes in the subventricular zone of the adult brain. J Neurosci26:7907–7918. [PubMed]
  • Mercer A, Rönnholm H, Holmberg J, Lundh H, Heidrich J, Zachrisson O, Ossoinak A, Frisén J, Patrone C. (2004) PACAP promotes neural stem cell proliferation in adult mouse brain. J Neurosci Res 76:205–215. [PubMed]
  • Messing RB, Dodge C, Waymire JC, Lynch GS, Deadwyler SA. (1979) Morphine induced increases in the incorporation of 3H-thymidine into brain striatal DNA. Brain Res Bull 4:615–619. [PubMed]
  • Meyer zu Heringdorf D, Jakobs KH. (2007) Lysophospholipid receptors: signalling, pharmacology and regulation by lysophospholipid metabolism. Biochim Biophys Acta1768:923–940. [PubMed]
  • Michel MC, Beck-Sickinger A, Cox H, Doods HN, Herzog H, Larhammar D, Quirion R, Schwartz T, Westfall T. (1998) XVI. International Union of Pharmacology recommendations for the nomenclature of neuropeptide Y, peptide YY, and pancreatic polypeptide receptors.Pharmacol Rev 50:143–150. [PubMed]
  • Mignini F, Bronzetti E, Felici L, Ricci A, Sabbatini M, Tayebati SK, Amenta F. (2000) Dopamine receptor immunohistochemistry in the rat choroid plexus. J Auton Pharmacol 20:325–332.[PubMed]
  • Milenkovic I, Weick M, Wiedemann P, Reichenbach A, Bringmann A. (2003) P2Y receptor-mediated stimulation of Müller glial cell DNA synthesis: dependence on EGF and PDGF receptor transactivation. Invest Ophthalmol Vis Sci 44:1211–1220. [PubMed]
  • Millar JK, Wilson-Annan JC, Anderson S, Christie S, Taylor MS, Semple CA, Devon RS, St Clair DM, Muir WJ, Blackwood DH, et al. (2000) Disruption of two novel genes by a translocation co-segregating with schizophrenia. Hum Mol Genet 9:1415–1423. [PubMed]
  • Miller CR, O’Steen WK, Deadwyler SA. (1982) Effect of morphine on 3H-thymidine incorporation in the subependyma of the rat: an autoradiographic study. J Comp Neurol208:209–214. [PubMed]
  • Miller MW, Nowakowski RS. (1988) Use of bromodeoxyuridine-immunohistochemistry to examine the proliferation, migration and time of origin of cells in the central nervous system.Brain Res 457:44–52. [PubMed]
  • Milosevic J, Schwarz SC, Maisel M, Poppe-Wagner M, Dieterlen MT, Storch A, Schwarz J. (2007) Dopamine D2/D3 receptor stimulation fails to promote dopaminergic neurogenesis of murine and human midbrain-derived neural precursor cells in vitro. Stem Cells Dev 16:625–635.[PubMed]
  • Ming GL, Song H. (2005) Adult neurogenesis in the mammalian central nervous system. Annu Rev Neurosci 28:223–250. [PubMed]
  • Ming GL, Song H. (2011) Adult neurogenesis in the mammalian brain: significant answers and significant questions. Neuron 70:687–702. [PMC free article] [PubMed]
  • Mino M, Kamii H, Fujimura M, Kondo T, Takasawa S, Okamoto H, Yoshimoto T. (2003)Temporal changes of neurogenesis in the mouse hippocampus after experimental subarachnoid hemorrhage. Neurol Res 25:839–845. [PubMed]
  • Mirochnic S, Wolf S, Staufenbiel M, Kempermann G. (2009) Age effects on the regulation of adult hippocampal neurogenesis by physical activity and environmental enrichment in the APP23 mouse model of Alzheimer disease. Hippocampus 19:1008–1018. [PubMed]
  • Miron VE, Jung CG, Kim HJ, Kennedy TE, Soliven B, Antel JP. (2008) FTY720 modulates human oligodendrocyte progenitor process extension and survival. Ann Neurol 63:61–71.[PubMed]
  • Mishra SK, Braun N, Shukla V, Füllgrabe M, Schomerus C, Korf HW, Gachet C, Ikehara Y, Sévigny J, Robson SC, et al. (2006) Extracellular nucleotide signaling in adult neural stem cells: synergism with growth factor-mediated cellular proliferation. Development 133:675–684.[PubMed]
  • Miyata A, Arimura A, Dahl RR, Minamino N, Uehara A, Jiang L, Culler MD, Coy DH. (1989)Isolation of a novel 38 residue-hypothalamic polypeptide which stimulates adenylate cyclase in pituitary cells. Biochem Biophys Res Commun 164:567–574. [PubMed]
  • Miyata A, Jiang L, Dahl RD, Kitada C, Kubo K, Fujino M, Minamino N, Arimura A. (1990)Isolation of a neuropeptide corresponding to the N-terminal 27 residues of the pituitary adenylate cyclase activating polypeptide with 38 residues (PACAP38). Biochem Biophys Res Commun 170:643–648. [PubMed]
  • Mohapel P, Leanza G, Kokaia M, Lindvall O. (2005) Forebrain acetylcholine regulates adult hippocampal neurogenesis and learning. Neurobiol Aging 26:939–946. [PubMed]
  • Molina-Holgado E, Vela JM, Arévalo-Martín A, Almazán G, Molina-Holgado F, Borrell J, Guaza C. (2002) Cannabinoids promote oligodendrocyte progenitor survival: involvement of cannabinoid receptors and phosphatidylinositol-3 kinase/AKT signaling. J Neurosci 22:9742–9753. [PubMed]
  • Molina-Holgado F, Rubio-Araiz A, García-Ovejero D, Williams RJ, Moore JD, Arévalo-Martín A, Gómez-Torres O, Molina-Holgado E. (2007) CB2 cannabinoid receptors promote mouse neural stem cell proliferation. Eur J Neurosci 25:629–634. [PubMed]
  • Mori H, Ninomiya K, Kino-oka M, Shofuda T, Islam MO, Yamasaki M, Okano H, Taya M, Kanemura Y. (2006) Effect of neurosphere size on the growth rate of human neural stem/progenitor cells. J Neurosci Res 84:1682–1691. [PubMed]
  • Morris BJ, Johnston HM. (1995) A role for hippocampal opioids in long-term functional plasticity. Trends Neurosci 18:350–355. [PubMed]
  • Morshead CM, Reynolds BA, Craig CG, McBurney MW, Staines WA, Morassutti D, Weiss S, van der Kooy D. (1994) Neural stem cells in the adult mammalian forebrain: a relatively quiescent subpopulation of subependymal cells. Neuron 13:1071–1082. [PubMed]
  • Mu Y, Lee SW, Gage FH. (2010) Signaling in adult neurogenesis. Curr Opin Neurobiol 20:416–423. [PMC free article] [PubMed]
  • Mu Y, Zhao C, Gage FH. (2011) Dopaminergic modulation of cortical inputs during maturation of adult-born dentate granule cells. J Neurosci 31:4113–4123. [PMC free article] [PubMed]
  • Muise-Helmericks RC, Grimes HL, Bellacosa A, Malstrom SE, Tsichlis PN, Rosen N. (1998)Cyclin D expression is controlled post-transcriptionally via a phosphatidylinositol 3-kinase/AKT-dependent pathway. J Biol Chem 273:29864–29872. [PubMed]
  • Mulder J, Aguado T, Keimpema E, Barabás K, Ballester Rosado CJ, Nguyen L, Monory K, Marsicano G, Di Marzo V, Hurd YL, et al. (2008) Endocannabinoid signaling controls pyramidal cell specification and long-range axon patterning. Proc Natl Acad Sci USA105:8760–8765. [PMC free article] [PubMed]
  • Nait-Oumesmar B, Decker L, Lachapelle F, Avellana-Adalid V, Bachelin C, Baron-Van Evercooren A. (1999) Progenitor cells of the adult mouse subventricular zone proliferate, migrate and differentiate into oligodendrocytes after demyelination. Eur J Neurosci 11:4357–4366. [PubMed]
  • Nait-Oumesmar B, Picard-Riera N, Kerninon C, Decker L, Seilhean D, Höglinger GU, Hirsch EC, Reynolds R, Baron-Van Evercooren A. (2007) Activation of the subventricular zone in multiple sclerosis: evidence for early glial progenitors. Proc Natl Acad Sci USA 104:4694–4699.[PMC free article] [PubMed]
  • Nakagomi S, Kiryu-Seo S, Kiyama H. (2000) Endothelin-converting enzymes and endothelin receptor B messenger RNAs are expressed in different neural cell species and these messenger RNAs are coordinately induced in neurons and astrocytes respectively following nerve injury.Neuroscience 101:441–449. [PubMed]
  • Nakamura N, Ramaswamy S, Vazquez F, Signoretti S, Loda M, Sellers WR. (2000) Forkhead transcription factors are critical effectors of cell death and cell cycle arrest downstream of PTEN. Mol Cell Biol 20:8969–8982. [PMC free article] [PubMed]
  • Nakatomi H, Kuriu T, Okabe S, Yamamoto S, Hatano O, Kawahara N, Tamura A, Kirino T, Nakafuku M. (2002) Regeneration of hippocampal pyramidal neurons after ischemic brain injury by recruitment of endogenous neural progenitors. Cell 110:429–441. [PubMed]
  • Napoleone P, Sancesario G, Amenta F. (1982) Indoleaminergic innervation of rat choroid plexus: a fluorescence histochemical study. Neurosci Lett 34:143–147. [PubMed]
  • Narita M, Kuzumaki N, Miyatake M, Sato F, Wachi H, Seyama Y, Suzuki T. (2006) Role of delta-opioid receptor function in neurogenesis and neuroprotection. J Neurochem 97:1494–1505. [PubMed]
  • Neary JT, Zhu Q. (1994) Signaling by ATP receptors in astrocytes. NeuroReport 5:1617–1620.[PubMed]
  • Nestler EJ. (2004) Historical review: Molecular and cellular mechanisms of opiate and cocaine addiction. Trends Pharmacol Sci 25:210–218. [PubMed]
  • Nicot A, Lelièvre V, Tam J, Waschek JA, DiCicco-Bloom E. (2002) Pituitary adenylate cyclase-activating polypeptide and sonic hedgehog interact to control cerebellar granule precursor cell proliferation. J Neurosci 22:9244–9254. [PubMed]
  • Niidome T, Taniuchi N, Akaike A, Kihara T, Sugimoto H. (2008) Differential regulation of neurogenesis in two neurogenic regions of APPswe/PS1dE9 transgenic mice. Neuroreport19:1361–1364. [PubMed]
  • Nimura M, Udagawa J, Hatta T, Hashimoto R, Otani H. (2006) Spatial and temporal patterns of expression of melanocortin type 2 and 5 receptors in the fetal mouse tissues and organs. Anat Embryol 211:109–117. [PubMed]
  • Nishimoto M, Furuta A, Aoki S, Kudo Y, Miyakawa H, Wada K. (2007) PACAP/PAC1 autocrine system promotes proliferation and astrogenesis in neural progenitor cells. Glia 55:317–327.[PubMed]
  • Nishino J, Kim I, Chada K, Morrison SJ. (2008) Hmga2 promotes neural stem cell self-renewal in young but not old mice by reducing p16Ink4a and p19Arf Expression. Cell 135:227–239.[PMC free article] [PubMed]
  • Nixon K, Crews FT. (2002) Binge ethanol exposure decreases neurogenesis in adult rat hippocampus. J Neurochem 83:1087–1093. [PubMed]
  • Novgorodov AS, El-Alwani M, Bielawski J, Obeid LM, Gudz TI. (2007) Activation of sphingosine- 1-phosphate receptor S1P5 inhibits oligodendrocyte progenitor migration. FASEB J21:1503–1514. [PubMed]
  • Ohashi H, Nishikawa K, Ayukawa K, Hara Y, Nishimoto M, Kudo Y, Abe T, Aoki S, Wada K. (2007) Alpha 1-adrenoceptor agonists protect against stress-induced death of neural progenitor cells. Eur J Pharmacol 573:20–28. [PubMed]
  • Ohta S, Gregg C, Weiss S. (2006) Pituitary adenylate cyclase-activating polypeptide regulates forebrain neural stem cells and neurogenesis in vitro and in vivo. J Neurosci Res 84:1177–1186.[PubMed]
  • Okano H, Temple S. (2009) Cell types to order: temporal specification of CNS stem cells. Curr Opin Neurobiol 19:112–119. [PubMed]
  • O’Keeffe GC, Barker RA, Caldwell MA. (2009a) Dopaminergic modulation of neurogenesis in the subventricular zone of the adult brain. Cell Cycle 8:2888–2894. [PubMed]
  • O’Keeffe GC, Tyers P, Aarsland D, Dalley JW, Barker RA, Caldwell MA. (2009b) Dopamine-induced proliferation of adult neural precursor cells in the mammalian subventricular zone is mediated through EGF. Proc Natl Acad Sci USA 106:8754–8759. [PMC free article] [PubMed]
  • Olanow CW, Goetz CG, Kordower JH, Stoessl AJ, Sossi V, Brin MF, Shannon KM, Nauert GM, Perl DP, Godbold J, et al. (2003) A double-blind controlled trial of bilateral fetal nigral transplantation in Parkinson’s disease. Ann Neurol 54:403–414. [PubMed]
  • Osman AM, Porritt MJ, Nilsson M, Kuhn HG. (2011) Long-term stimulation of neural progenitor cell migration after cortical ischemia in mice. Stroke 42:3559–3565. [PubMed]
  • Overington JP, Al-Lazikani B, Hopkins AL. (2006) How many drug targets are there? Nat Rev Drug Discov 5:993–996. [PubMed]
  • Palazuelos J, Aguado T, Egia A, Mechoulam R, Guzmán M, Galve-Roperh I. (2006) Non-psychoactive CB2 cannabinoid agonists stimulate neural progenitor proliferation. FASEB J20:2405–2407. [PubMed]
  • Palmer TD, Takahashi J, Gage FH. (1997) The adult rat hippocampus contains primordial neural stem cells. Mol Cell Neurosci 8:389–404. [PubMed]
  • Palmer TD, Willhoite AR, Gage FH. (2000) Vascular niche for adult hippocampal neurogenesis.J Comp Neurol 425:479–494. [PubMed]
  • Papay R, Gaivin R, Jha A, McCune DF, McGrath JC, Rodrigo MC, Simpson PC, Doze VA, Perez DM. (2006) Localization of the mouse α1A-adrenergic receptor (AR) in the brain: α1AAR is expressed in neurons, GABAergic interneurons, and NG2 oligodendrocyte progenitors. J Comp Neurol 497:209–222. [PubMed]
  • Parent JM. (2007) Adult neurogenesis in the intact and epileptic dentate gyrus. Prog Brain Res163:529–540. [PubMed]
  • Parent JM, Janumpalli S, McNamara JO, Lowenstein DH. (1998) Increased dentate granule cell neurogenesis following amygdala kindling in the adult rat. Neurosci Lett 247:9–12. [PubMed]
  • Parent JM, Tada E, Fike JR, Lowenstein DH. (1999) Inhibition of dentate granule cell neurogenesis with brain irradiation does not prevent seizure-induced mossy fiber synaptic reorganization in the rat. J Neurosci 19:4508–4519. [PubMed]
  • Park JH, Enikolopov G. (2010) Transient elevation of adult hippocampal neurogenesis after dopamine depletion. Exp Neurol 222:267–276. [PMC free article] [PubMed]
  • Parker MA, Anderson JK, Corliss DA, Abraria VE, Sidman RL, Park KI, Teng YD, Cotanche DA, Snyder EY. (2005) Expression profile of an operationally-defined neural stem cell clone. Exp Neurol 194:320–332. [PubMed]
  • Pastrana E, Cheng LC, Doetsch F. (2009) Simultaneous prospective purification of adult subventricular zone neural stem cells and their progeny. Proc Natl Acad Sci USA 106:6387–6392. [PMC free article] [PubMed]
  • Pastrana E, Silva-Vargas V, Doetsch F. (2011) Eyes wide open: a critical review of sphere-formation as an assay for stem cells. Cell Stem Cell 8:486–498. [PMC free article] [PubMed]
  • Pazos A, Palacios JM. (1985) Quantitative autoradiographic mapping of serotonin receptors in the rat brain. I. Serotonin-1 receptors. Brain Res 346:205–230. [PubMed]
  • Pébay A, Bonder CS, Pitson SM. (2007) Stem cell regulation by lysophospholipids.Prostaglandins Other Lipid Mediat 84:83–97. [PubMed]
  • Peltier J, Schaffer DV. (2010) Viral packaging and transduction of adult hippocampal neural progenitors. Methods Mol Biol 621:103–116. [PubMed]
  • Perez JA, Clinton SM, Turner CA, Watson SJ, Akil H. (2009) A new role for FGF2 as an endogenous inhibitor of anxiety. J Neurosci 29:6379–6387. [PMC free article] [PubMed]
  • Perez DM, Doze VA. (2011) Cardiac and neuroprotection regulated by α1-adrenergic receptor subtypes. J Recept Signal Transduct Res 31:98–110. [PMC free article] [PubMed]
  • Persson AI, Naylor AS, Jonsdottir IH, Nyberg F, Eriksson PS, Thorlin T. (2004) Differential regulation of hippocampal progenitor proliferation by opioid receptor antagonists in running and non-running spontaneously hypertensive rats. Eur J Neurosci 19:1847–1855. [PubMed]
  • Persson AI, Thorlin T, Bull C, Zarnegar P, Ekman R, Terenius L, Eriksson PS. (2003) Mu- and delta-opioid receptor antagonists decrease proliferation and increase neurogenesis in cultures of rat adult hippocampal progenitors. Eur J Neurosci 17:1159–1172. [PubMed]
  • Pertwee RG. (2009) Emerging strategies for exploiting cannabinoid receptor agonists as medicines. Br J Pharmacol 156:397–411. [PMC free article] [PubMed]
  • Pickering C, Hägglund M, Szmydynger-Chodobska J, Marques F, Palha JA, Waller L, Chodobski A, Fredriksson R, Lagerström MC, Schiöth HB. (2008) The Adhesion GPCR GPR125 is specifically expressed in the choroid plexus and is upregulated following brain injury. BMC Neurosci 9:97. [PMC free article] [PubMed]
  • Pinnock SB, Lazic SE, Wong HT, Wong IH, Herbert J. (2009) Synergistic effects of dehydroepiandrosterone and fluoxetine on proliferation of progenitor cells in the dentate gyrus of the adult male rat. Neuroscience 158:1644–1651. [PMC free article] [PubMed]
  • Pluchino S, Quattrini A, Brambilla E, Gritti A, Salani G, Dina G, Galli R, Del Carro U, Amadio S, Bergami A, et al. (2003) Injection of adult neurospheres induces recovery in a chronic model of multiple sclerosis. Nature 422:688–694. [PubMed]
  • Popovik E, Haynes LW. (2000) Survival and mitogenesis of neuroepithelial cells are influenced by noradrenergic but not cholinergic innervation in cultured embryonic rat neopallium. Brain Res 853:227–235. [PubMed]
  • Potten CS, Loeffler M. (1990) Stem cells: attributes, cycles, spirals, pitfalls and uncertainties. Lessons for and from the crypt. Development 110:1001–1020. [PubMed]
  • Preffer F, Dombkowski D. (2009) Advances in complex multiparameter flow cytometry technology: Applications in stem cell research. Cytometry B Clin Cytom 76:295–314.[PMC free article] [PubMed]
  • Radley JJ, Jacobs BL. (2002) 5-HT1A receptor antagonist administration decreases cell proliferation in the dentate gyrus. Brain Res 955:264–267. [PubMed]
  • Ramírez-Rodríguez G, Klempin F, Babu H, Benítez-King G, Kempermann G. (2009) Melatonin modulates cell survival of new neurons in the hippocampus of adult mice.Neuropsychopharmacology 34:2180–2191. [PubMed]
  • Ramirez-Rodriguez G, Ortíz-López L, Domínguez-Alonso A, Benítez-King GA, Kempermann G. (2011) Chronic treatment with melatonin stimulates dendrite maturation and complexity in adult hippocampal neurogenesis of mice. J Pineal Res 50:29–37. [PubMed]
  • Rao MS, Hattiangady B, Shetty AK. (2008) Status epilepticus during old age is not associated with enhanced hippocampal neurogenesis. Hippocampus 18:931–944. [PMC free article][PubMed]
  • Ray J. (2008) Monolayer cultures of neural stem/progenitor cells, in Adult Neurogenesis (Gage FH, Kempermann G, Song H, editors. eds), 135–157, Cold Spring Harbor Laboratory Press, Cold Spring Harbor, New York.
  • Ray J, Peterson DA, Schinstine M, Gage FH. (1993) Prolieration, differentiation, and long-term culture of primary hippocampal neurons. Proc Natl Acad Sci 90:3602–3606. [PMC free article][PubMed]
  • Redwine JM, Armstrong RC. (1998) In vivo proliferation of oligodendrocyte progenitors expressing PDGFalphaR during early remyelination. J Neurobiol 37:413–428. [PubMed]
  • 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 11:514–522.[PubMed]
  • Reif A, Schmitt A, Fritzen S, Chourbaji S, Bartsch C, Urani A, Wycislo M, Mössner R, Sommer C, Gass P, et al. (2004) Differential effect of endothelial nitric oxide synthase (NOS-III) on the regulation of adult neurogenesis and behaviour. Eur J Neurosci 20:885–895. [PubMed]
  • Rennie K, De Butte M, Pappas BA. (2009) Melatonin promotes neurogenesis in dentate gyrus in the pinealectomized rat. J Pineal Res 47:313–317. [PubMed]
  • Reynolds BA, Tetzlaff W, Weiss S. (1992) A multipotent EGF-responsive striatal embryonic progenitor cell produces neurons and astrocytes. J Neurosci 12:4565–4574. [PubMed]
  • Reynolds BA, Weiss S. (1992) Generation of neurons and astrocytes from isolated cells of the adult mammalian central nervous system. Science 255:1707–1710. [PubMed]
  • Ribeiro JA, Sebastião AM, de Mendonça A. (2002) Adenosine receptors in the nervous system: pathophysiological implications. Prog Neurobiol 68:377–392. [PubMed]
  • Richards LJ, Kilpatrick TJ, Bartlett PF. (1992) De novo generation of neuronal cells from the adult mouse brain. Proc Natl Acad Sci USA 89:8591–8595. [PMC free article] [PubMed]
  • Riess P, Zhang C, Saatman KE, Laurer HL, Longhi LG, Raghupathi R, Lenzlinger PM, Lifshitz J, Boockvar J, Neugebauer E, et al. (2002) Transplanted neural stem cells survive, differentiate, and improve neurological motor function after experimental traumatic brain injury.Neurosurgery 51:1043–1052; discussion 1052—1054. [PubMed]
  • Rietze R, Poulin P, Weiss S. (2000) Mitotically active cells that generate neurons and astrocytes are present in multiple regions of the adult mouse hippocampus. J Comp Neurol 424:397–408.[PubMed]
  • Rietze RL, Reynolds BA. (2006) Neural stem cell isolation and characterization. Methods Enzymol 419:3–23. [PubMed]
  • Rizk P, Salazar J, Raisman-Vozari R, Marien M, Ruberg M, Colpaert F, Debeir T. (2006) The alpha2-adrenoceptor antagonist dexefaroxan enhances hippocampal neurogenesis by increasing the survival and differentiation of new granule cells. Neuropsychopharmacology31:1146–1157. [PubMed]
  • Roberts JC, Reavill C, East SZ, Harrison PJ, Patel S, Routledge C, Leslie RA. (2002) The distribution of 5-HT6 receptors in rat brain: an autoradiographic binding study using the radiolabelled 5-HT6 receptor antagonist [125I]SB-258585. Brain Res 934:49–57. [PubMed]
  • Rodrigo C, Zaben M, Lawrence T, Laskowski A, Howell OW, Gray WP. (2010) NPY augments the proliferative effect of FGF2 and increases the expression of FGFR1 on nestin positive postnatal hippocampal precursor cells, via the Y1 receptor. J Neurochem 113:615–627.[PubMed]
  • Rodríguez JJ, Jones VC, Tabuchi M, Allan SM, Knight EM, LaFerla FM, Oddo S, Verkhratsky A. (2008) A Impaired adult neurogenesis in the dentate gyrus of a triple transgenic mouse model of Alzheimer’s disease. PLoS One 3:e2935. [PMC free article] [PubMed]
  • Roh JH, Qiu A, Seo SW, Soon HW, Kim JH, Kim GH, Kim MJ, Lee JM, Na DL. (2011) Volume reduction in subcortical regions according to severity of Alzheimer’s disease. J Neurol 258:1013–1020. [PubMed]
  • Rohe M, Carlo AS, Breyhan H, Sporbert A, Militz D, Schmidt V, Wozny C, Harmeier A, Erdmann B, Bales KR, et al. (2008) Sortilin-related receptor with A-type repeats (SORLA) affects the amyloid precursor protein-dependent stimulation of ERK signaling and adult neurogenesis. J Biol Chem 283:14826–14834. [PubMed]
  • Rosen H, Gonzalez-Cabrera PJ, Sanna MG, Brown S. (2009) Sphingosine 1-phosphate receptor signaling. Annu Rev Biochem 78:743–768. [PubMed]
  • Ross HH, Levkoff LH, Marshall GP, 2nd, Caldeira M, Steindler DA, Reynolds BA, Laywell ED. (2008) Bromodeoxyuridine induces senescence in neural stem and progenitor cells. Stem Cells26:3218–3227. [PubMed]
  • Rueda D, Navarro B, Martinez-Serrano A, Guzman M, Galve-Roperh I. (2002) The endocannabinoid anandamide inhibits neuronal progenitor cell differentiation through attenuation of the Rap1/B-Raf/ERK pathway. J Biol Chem 277:46645–46650. [PubMed]
  • Ryu JK, Choi HB, Hatori K, Heisel RL, Pelech SL, McLarnon JG, Kim SU. (2003) Adenosine triphosphate induces proliferation of human neural stem cells: Role of calcium and p70 ribosomal protein S6 kinase. J Neurosci Res 72:352–362. [PubMed]
  • Sachdeva R, Jönsson ME, Nelander J, Kirkeby A, Guibentif C, Gentner B, Naldini L, Björklund A, Parmar M, Jakobsson J. (2010) Tracking differentiating neural progenitors in pluripotent cultures using microRNA-regulated lentiviral vectors. Proc Natl Acad Sci USA 107:11602–11607.[PMC free article] [PubMed]
  • Saini HS, Coelho RP, Goparaju SK, Jolly PS, Maceyka M, Spiegel S, Sato-Bigbee C. (2005) Novel role of sphingosine kinase 1 as a mediator of neurotrophin-3 action in oligodendrocyte progenitors. J Neurochem 95:1298–1310. [PubMed]
  • Saini HS, Gorse KM, Boxer LM, Sato-Bigbee C. (2004) Neurotrophin-3 and a CREB-mediated signaling pathway regulate Bcl-2 expression in oligodendrocyte progenitor cells. J Neurochem89:951–961. [PubMed]
  • Sajad M, Chawla R, Zargan J, Umar S, Sadaqat M, Khan HA. (2011) Cytokinetics of adult rat SVZ after EAE. Brain Res 1371:140–149. [PubMed]
  • Samuels BA, Hen R. (2011) Neurogenesis and affective disorders. Eur J Neurosci 33:1152–1159.[PubMed]
  • Sanai N, Tramontin AD, Quiñones-Hinojosa A, Barbaro NM, Gupta N, Kunwar S, Lawton MT, McDermott MW, Parsa AT, Manuel-García Verdugo J, et al. (2004) Unique astrocyte ribbon in adult human brain contains neural stem cells but lacks chain migration. Nature 427:740–744.[PubMed]
  • Sanosaka T, Namihira M, Nakashima K. (2009) Epigenetic mechanisms in sequential differentiation of neural stem cells. Epigenetics 4:89–92. [PubMed]
  • Santarelli L, Saxe M, Gross C, Surget A, Battaglia F, Dulawa S, Weisstaub N, Lee J, Duman R, Arancio O, et al. (2003) Requirement of hippocampal neurogenesis for the behavioral effects of antidepressants. Science 301:805–809. [PubMed]
  • Santin LJ, Bilbao A, Pedraza C, Matas-Rico E, López-Barroso D, Castilla-Ortega E, Sánchez-López J, Riquelme R, Varela-Nieto I, de la Villa P, et al. (2009) Behavioral phenotype of maLPA1-null mice: increased anxiety-like behavior and spatial memory deficits. Genes Brain Behav 8:772–784. [PubMed]
  • Sasaki T, Kitagawa K, Omura-Matsuoka E, Todo K, Terasaki Y, Sugiura S, Hatazawa J, Yagita Y, Hori M. (2007) The phosphodiesterase inhibitor rolipram promotes survival of newborn hippocampal neurons after ischemia. Stroke 38:1597–1605. [PubMed]
  • Sasaki T, Kitagawa K, Yamagata K, Takemiya T, Tanaka S, Omura-Matsuoka E, Sugiura S, Matsumoto M, Hori M. (2004) Amelioration of hippocampal neuronal damage after transient forebrain ischemia in cyclooxygenase-2-deficient mice. J Cereb Blood Flow Metab 24:107–113.[PubMed]
  • Scemes E, Duval N, Meda P. (2003) Reduced expression of P2Y1 receptors in connexin43-null mice alters calcium signaling and migration of neural progenitor cells. J Neurosci 23:11444–11452. [PMC free article] [PubMed]
  • Scharfman HE, Goodman JH, Sollas AL. (2000) Granule-like neurons at the hilar/CA3 border after status epilepticus and their synchrony with area CA3 pyramidal cells: functional implications of seizure-induced neurogenesis. J Neurosci 20:6144–6158. [PubMed]
  • Sebastião AM, de Mendonça A, Ribeiro JA. (2001) Neuroprotection during hypoxic insults: role of adenosine. Drug Dev Res 52:291–295.
  • Selkoe DJ. (2003) Aging, amyloid, and Alzheimer’s disease: a perspective in honor of Carl Cotman. Neurochem Res 28:1705–1713. [PubMed]
  • Sheline YI. (2000) 3D MRI studies of neuroanatomic changes in unipolar major depression: the role of stress and medical comorbidity. Biol Psychiatry 48:791–800. [PubMed]
  • Sheline YI, Wang PW, Gado MH, Csernansky JG, Vannier MW. (1996) Hippocampal atrophy in recurrent major depression. Proc Natl Acad Sci USA 93:3908–3913. [PMC free article][PubMed]
  • Shen Q, Goderie SK, Jin L, Karanth N, Sun Y, Abramova N, Vincent P, Pumiglia K, Temple S. (2004) Endothelial cells stimulate self-renewal and expand neurogenesis of neural stem cells.Science 304:1338–1340. [PubMed]
  • Shen Q, Wang Y, Kokovay E, Lin G, Chuang SM, Goderie SK, Roysam B, Temple S. (2008)Adult SVZ stem cells lie in a vascular niche: a quantitative analysis of niche cell-cell interactions.Cell Stem Cell 3:289–300. [PMC free article] [PubMed]
  • Sheng WS, Hu S, Herr G, Ni HT, Rock RB, Gekker G, Lokensgard JR, Peterson PK. (2007)Human neural precursor cells express functional kappa-opioid receptors. J Pharmacol Exp Ther 322:957–963. [PubMed]
  • Shenoy SK, Drake MT, Nelson CD, Houtz DA, Xiao K, Madabushi S, Reiter E, Premont RT, Lichtarge O, Lefkowitz RJ. (2006) β-arrestin-dependent, G protein-independent ERK1/2 activation by the β2 adrenergic receptor. J Biol Chem 281:1261–1273. [PubMed]
  • Shetty AK, Hattiangady B, Rao MS, Shuai B. (2011) Deafferentation enhances neurogenesis in the young and middle aged hippocampus but not in the aged hippocampus. Hippocampus21:631–646. [PMC free article] [PubMed]
  • Shi Y, Chichung Lie D, Taupin P, Nakashima K, Ray J, Yu RT, Gage FH, Evans RM. (2004)Expression and function of orphan nuclear receptor TLX in adult neural stem cells. Nature427:78–83. [PubMed]
  • Shimozaki K, Namihira M, Nakashima K, Taga T. (2005) Stage- and site-specific DNA demethylation during neural cell development from embryonic stem cells. J Neurochem93:432–439. [PubMed]
  • Shihabuddin LS, Horner PJ, Ray J, Gage FH. (2000) Adult spinal cord stem cells generate neurons after transplantation in the adult dentate gyrus. J Neurosci 20:8727–8735. [PubMed]
  • Shukla V, Zimmermann H, Wang L, Kettenmann H, Raab S, Hammer K, Sévigny J, Robson SC, Braun N. (2005) Functional expression of the ecto-ATPase NTPDase2 and of nucleotide receptors by neuronal progenitor cells in the adult murine hippocampus. J Neurosci Res80:600–610. [PubMed]
  • Singec I, Knoth R, Meyer RP, Maciaczyk J, Volk B, Nikkhah G, Frotscher M, Snyder EY. (2006)Defining the actual sensitivity and specificity of the neurosphere assay in stem cell biology. Nat Methods 3:801–806. [PubMed]
  • Singec I, Quinones-Hinojosa A. (2008) Neurospheres, in Adult Neurogenesis (Gage FH, Kempermann G, Song H, editors. eds) pp 119–134, Cold Spring Harbor Laboratory Press, Cold Spring Harbor, New York.
  • Siu FK, Sham MH, Chow BK. (2006) The prenatal expression of secretin receptor. Ann NY Acad Sci 1070:561–565. [PubMed]
  • Sjögren B, Blazer LL, Neubig RR. (2010) Regulators of G protein signaling proteins as targets for drug discovery. Prog Mol Biol Transl Sci 91:81–119. [PubMed]
  • Smalheiser NR, Dissanayake S, Kapil A. (1996) Rapid regulation of neurite outgrowth and retraction by phospholipase A2-derived arachidonic acid and its metabolites. Brain Res 721:39–48. [PubMed]
  • Smiałowska M, Domin H, Zieba B, Koźniewska E, Michalik R, Piotrowski P, Kajta M. (2009)Neuroprotective effects of neuropeptide Y-Y2 and Y5 receptor agonists in vitro and in vivo.Neuropeptides 43:235–249. [PubMed]
  • Soltys J, Yushak M, Mao-Draayer Y. (2010) Regulation of neural progenitor cell fate by anandamide. Biochem Biophys Res Commun 400:21–26. [PubMed]
  • Soumier A, Banasr M, Goff LK, Daszuta A. (2010) Region- and phase-dependent effects of 5-HT1A and 5-HT2C receptor activation on adult neurogenesis. Eur Neuropsychopharmacol20:336–345. [PubMed]
  • Stafford MR, Bartlett PF, Adams DJ. (2007) Purinergic receptor activation inhibits mitogen-stimulated proliferation in primary neurospheres from the adult mouse subventricular zone.Mol Cell Neurosci 35:535–548. [PubMed]
  • 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 188:510–518. [PubMed]
  • Stingl J. (2009) Detection and analysis of mammary gland stem cells. J Pathol 217:229–241.[PubMed]
  • Stitelman DH, Endo M, Bora A, Muvarak N, Zoltick PW, Flake AW, Brazelton TR. (2010)Robust in vivo transduction of nervous system and neural stem cells by early gestational intra amniotic gene transfer using lentiviral vector. Mol Ther 18:1615–1623. [PMC free article][PubMed]
  • Suadicani SO, De Pina-Benabou MH, Urban-Maldonado M, Spray DC, Scemes E. (2003) Acute downregulation of Cx43 alters P2Y receptor expression levels in mouse spinal cord astrocytes.Glia 42:160–171. [PMC free article] [PubMed]
  • Suárez J, Romero-Zerbo SY, Rivera P, Bermúdez-Silva FJ, Pérez J, De Fonseca FR, Fernández-Llebrez P. (2010) Endocannabinoid system in the adult rat circumventricular areas: an immunohistochemical study. J Comp Neurol 518:3065–3085. [PubMed]
  • Suh H, Consiglio A, Ray J, Sawai T, D’Amour KA, Gage FH. (2007) In vivo fate analysis reveals the multipotent and self-renewal capacities of Sox2+ neural stem cells in the adult hippocampus. Cell Stem Cell 1:515–528. [PMC free article] [PubMed]
  • Surget A, Saxe M, Leman S, Ibarguen-Vargas Y, Chalon S, Griebel G, Hen R, Belzung C. (2008)Drug-dependent requirement of hippocampal neurogenesis in a model of depression and of antidepressant reversal. Biol Psychiatry 64:293–301. [PubMed]
  • Svetlov SI, Ignatova TN, Wang KK, Hayes RL, English D, Kukekov VG. (2004) Lysophosphatidic acid induces clonal generation of mouse neurospheres via proliferation of Sca-1- and AC133-positive neural progenitors. Stem Cells Dev 13:685–693. [PubMed]
  • Svoboda KR, Adams CE, Lupica CR. (1999) Opioid receptor subtype expression defines morphologically distinct classes of hippocampal interneurons. J Neurosci 19:85–95. [PubMed]
  • Tafuri S, Pavone LM, Mastellone V, Spina A, Avallone L, Vittoria A, Staiano N, Scala G. (2009)Expression of orexin A and its receptor 1 in the choroid plexuses from buffalo brain.Neuropeptides 43:73–80. [PubMed]
  • Takeda S, Kadowaki S, Haga T, Takaesu H, Mitaku S. (2002) Identification of G protein-coupled receptor gens from the human genome sequence. FEBS Lett 520:97–101. [PubMed]
  • Takei H, Wilfong A, Yoshor D, Armstrong DL, Bhattacharjee MB. (2007) Evidence of increased cell proliferation in the hippocampus in children with Ammon’s horn sclerosis. Pathol Int57:76–81. [PubMed]
  • Tárnok A, Ulrich H, Bocsi J. (2010) Phenotypes of stem cells from diverse origin. Cytometry A77:6–10. [PubMed]
  • Taupin P. (2006) Therapeutic potential of adult neural stem cells. Recent Pat CNS Drug Discov1:299–303. [PubMed]
  • Taupin P. (2007) BrdU immunohistochemistry for studying adult neurogenesis: paradigms, pitfalls, limitations, and validation. Brain Res Rev 53:198–214. [PubMed]
  • Tavazoie M, Van der Veken L, Silva-Vargas V, Louissaint M, Colonna L, Zaidi B, Garcia-Verdugo JM, Doetsch F. (2008) A specialized vascular niche for adult neural stem cells. Cell Stem Cell 3:279–288. [PubMed]
  • Temple S. (2001) The development of neural stem cells. Nature 414:112–117. [PubMed]
  • Terai K, Soga T, Takahashi M, Kamohara M, Ohno K, Yatsugi S, Okada M, Yamaguchi T. (2003) Edg-8 receptors are preferentially expressed in oligodendrocyte lineage cells of the rat CNS. Neuroscience 116:1053–1062. [PubMed]
  • Terman GW, Drake CT, Simmons ML, Milner TA, Chavkin C. (2000) Opioid modulation of recurrent excitation in the hippocampal dentate gyrus. J Neurosci 20:4379–4388. [PubMed]
  • Thiriet N, Agasse F, Nicoleau C, Guégan C, Vallette F, Cadet JL, Jaber M, Malva JO, Coronas V. (2011) NPY promotes chemokinesis and neurogenesis in the rat subventricular zone. J Neurochem 116:1018–1027. [PubMed]
  • Thom M, Sisodiya SM, Beckett A, Martinian L, Lin WR, Harkness W, Mitchell TN, Craig J, Duncan J, Scaravilli F. (2002) Cytoarchitectural abnormalities in hippocampal sclerosis. J Neuropathol Exp Neurol 61:510–519. [PubMed]
  • Tomasiewicz H, Ono K, Yee D, Thompson C, Goridis C, Rutishauser U, Magnuson T. (1993)Genetic deletion of a neural cell adhesion molecule variant (N-CAM-180) produces distinct defects in the central nervous system. Neuron 11:1163–1174. [PubMed]
  • Tonchev AB, Yamashima T, Sawamoto K, Okano H. (2005) Enhanced proliferation of progenitor cells in the subventricular zone and limited neuronal production in the striatum and neocortex of adult macaque monkeys after global cerebral ischemia. J Neurosci Res 81:776–788. [PubMed]
  • Tozuka Y, Fukuda S, Namba T, Seki T, Hisatsune T. (2005) GABAergic excitation promotes neuronal differentiation in adult hippocampal progenitor cells. Neuron 47:803–815. [PubMed]
  • Tran PB, Miller RJ. (2005) HIV-1, chemokines and neurogenesis. Neurotox Res 8:149–158.[PMC free article] [PubMed]
  • Tripathi A, Khurshid N, Kumar P, Iyengar S. (2008) Expression of delta- and mu-opioid receptors in the ventricular and subventricular zones of the developing human neocortex.Neurosci Res 61:257–270. [PubMed]
  • Uchida K, Kumihashi K, Kurosawa S, Kobayashi T, Itoi K, Machida T. (2002) Stimulatory effects of prostaglandin E2 on neurogenesis in the dentate gyrus of the adult rat. Zoolog Sci 19:1211–1216. [PubMed]
  • Uckermann O, Grosche J, Reichenbach A, Bringmann A. (2002) ATP-evoked calcium responses of radial glial (Müller) cells in the postnatal rabbit retina. J Neurosci Res 70:209–218. [PubMed]
  • Uhl GR, Sora I, Wang Z. (1999) The mu opiate receptor as a candidate gene for pain: polymorphisms, variations in expression, nociception, and opiate responses. Proc Natl Acad Sci USA 96:7752–7755. [PMC free article] [PubMed]
  • Van Der Meer P, Goldberg SH, Fung KM, Sharer LR, González-Scarano F, Lavi E. (2001)Expression pattern of CXCR3, CXCR4, and CCR3 chemokine receptors in the developing human brain. J Neuropathol Exp Neurol 60:25–32. [PubMed]
  • Van Kampen JM, Eckman CB. (2006) Dopamine D3 receptor agonist delivery to a model of Parkinson’s disease restores the nigrostriatal pathway and improves locomotor behavior. J Neurosci 26:7272–7280. [PubMed]
  • Van Kampen JM, Hagg T, Robertson HA. (2004) Induction of neurogenesis in the adult rat subventricular zone and neostriatum following dopamine D3 receptor stimulation. Eur J Neurosci 19:2377–2387. [PubMed]
  • Van Kampen JM, Robertson HA. (2005) A possible role for dopamine D3 receptor stimulation in the induction of neurogenesis in the adult rat substantia nigra. Neuroscience 136:381–386.[PubMed]
  • van Os J, Kapur S. (2009) Schizophrenia. Lancet 374:635–645. [PubMed]
  • van Praag H, Christie BR, Sejnowski TJ, Gage FH. (1999) Running enhances neurogenesis, learning, and long-term potentiation in mice. Proc Natl Acad Sci USA 96:13427–13431.[PMC free article] [PubMed]
  • van Praag H, Schinder AF, Christie BR, Toni N, Palmer TD, Gage FH. (2002) Functional neurogenesis in the adult hippocampus. Nature 415:1030–1034. [PubMed]
  • Vaudry D, Gonzalez BJ, Basille M, Yon L, Fournier A, Vaudry H. (2000) Pituitary adenylate cyclase-activating polypeptide and its receptors: from structure to functions. Pharmacol Rev52:269–324. [PubMed]
  • Veena J, Srikumar BN, Mahati K, Raju TR, Shankaranarayana Rao BS. (2011) Oxotremorine treatment restores hippocampal neurogenesis and ameliorates depression-like behaviour in chronically stressed rats. Psychopharmacology (Berl) 217:239–253. [PubMed]
  • Veena J, Srikumar BN, Raju TR, Shankaranarayana Rao BS. (2009) Exposure to enriched environment restores the survival and differentiation of new born cells in the hippocampus and ameliorates depressive symptoms in chronically stressed rats. Neurosci Lett 455:178–182.[PubMed]
  • Vértes Z, Melegh G, Vértes M, Kovács S. (1982) Effect of naloxone and D-met2-pro5-enkephalinamide treatment on the DNA synthesis in the developing rat brain. Life Sci 31:119–126. [PubMed]
  • Vescovi AL, Reynolds BA, Fraser DD, Weiss S. (1993) bFGF regulates the proliferative fate of unipotent (neuronal) and bipotent (neuronal/astroglial) EGF-generated CNS progenitor cells.Neuron 11:951–966. [PubMed]
  • Veyrac A, Didier A, Colpaert F, Jourdan F, Marien M. (2005) Activation of noradrenergic transmission by alpha2-adrenoceptor antagonists counteracts deafferentation-induced neuronal death and cell proliferation in the adult mouse olfactory bulb. Exp Neurol 194:444–456. [PubMed]
  • Wakade CG, Mahadik SP, Waller JL, Chiu FC. (2002) Atypical neuroleptics stimulate neurogenesis in adult rat brain. J Neurosci Res 69:72–79. [PubMed]
  • Waldhoer M, Bartlett SE, Whistler JL. (2004) Opioid receptors. Annu Rev Biochem 73:953–990. [PubMed]
  • Wang HD, Dunnavant FD, Jarman T, Deutch AY. (2004) Effects of antipsychotic drugs on neurogenesis in the forebrain of the adult rat. Neuropsychopharmacology 29:1230–1238.[PubMed]
  • Warner-Schmidt JL, Duman RS. (2006) Hippocampal neurogenesis: opposing effects of stress and antidepressant treatment. Hippocampus 16:239–249. [PubMed]
  • Wegner F, Kraft R, Busse K, Schaarschmidt G, Härtig W, Schwarz SC, Schwarz J. (2009)Glutamate receptor properties of human mesencephalic neural progenitor cells: NMDA enhances dopaminergic neurogenesis in vitro. J Neurochem 111:204–216. [PubMed]
  • Wei H, Ahn S, Shenoy SK, Karnik SS, Hunyady L, Luttrell LM, Lefkowitz RJ. (2003)Independent β-arrestin 2 and G protein-mediated pathways for angiotensin II activation of extracellular signal-regulated kinases 1 and 2. Proc Natl Acad Sci USA 100:10782–10787.[PMC free article] [PubMed]
  • Weiss S, Dunne C, Hewson J, Wohl C, Wheatley M, Peterson AC, Reynolds BA. (1996)Multipotent CNS stem cells are present in the adult mammalian spinal cord and ventricular neuroaxis. J Neurosci 16:7599–7609. [PubMed]
  • Weissman TA, Riquelme PA, Ivic L, Flint AC, Kriegstein AR. (2004) Calcium waves propagate through radial glial cells and modulate proliferation in the developing neocortex. Neuron43:647–661. [PubMed]
  • Wenning GK, Odin P, Morrish P, Rehncrona S, Widner H, Brundin P, Rothwell JC, Brown R, Gustavii B, Hagell P, et al. (1997) Short- and long-term survival and function of unilateral intrastriatal dopaminergic grafts in Parkinson’s disease. Ann Neurol 42:95–107. [PubMed]
  • Wexler EM, Paucer A, Kornblum HI, Palmer TD, Plamer TD, Geschwind DH. (2009)Endogenous Wnt signaling maintains neural progenitor cell potency. Stem Cells 27:1130–1141.[PMC free article] [PubMed]
  • Winner B, Desplats P, Hagl C, Klucken J, Aigner R, Ploetz S, Laemke J, Karl A, Aigner L, Masliah E, et al. (2009) Dopamine receptor activation promotes adult neurogenesis in an acute Parkinson model. Exp Neurol 219:543–552. [PubMed]
  • Winner B, Geyer M, Couillard-Despres S, Aigner R, Bogdahn U, Aigner L, Kuhn G, Winkler J. (2006) Striatal deafferentation increases dopaminergic neurogenesis in the adult olfactory bulb.Exp Neurol 197:113–121. [PubMed]
  • Winner B, Kohl Z, Gage FH. (2011) Neurodegenerative disease and adult neurogenesis. Eur J Neurosci 33:1139–1151. [PubMed]
  • Wolf SA, Bick-Sander A, Fabel K, Leal-Galicia P, Tauber S, Ramirez-Rodriguez G, Müller A, Melnik A, Waltinger TP, Ullrich O, et al. (2010) Cannabinoid receptor CB1 mediates baseline and activity-induced survival of new neurons in adult hippocampal neurogenesis. Cell Commun Signal 8:12–28. [PMC free article] [PubMed]
  • Wolswijk G. (1998) Chronic stage multiple sclerosis lesions contain a relatively quiescent population of oligodendrocyte precursor cells. J Neurosci 18:601–609. [PubMed]
  • Wright EM, Saito Y. (1986) The choroid plexus as a route from blood to brain. Ann NY Acad Sci481:214–220. [PubMed]
  • Wu Y, Peng H, Cui M, Whitney NP, Huang Y, Zheng JC. (2009) CXCL12 increases human neural progenitor cell proliferation through AKT-1/FOXO3a signaling pathway. J Neurochem109:1157–1167. [PMC free article] [PubMed]
  • Xapelli S, Agasse F, Ferreira R, Silva AP, Malva JO. (2006) Neuropeptide Y as an endogenous antiepileptic, neuroprotective and pro-neurogenic peptide. Recent Pat CNS Drug Discov 1:315–324. [PubMed]
  • Yamaguchi M, Suzuki T, Seki T, Namba T, Juan R, Arai H, Hori T, Asada T. (2004) Repetitive cocaine administration decreases neurogenesis in adult rat hippocampus. Ann NY Acad Sci1025:351–362. [PubMed]
  • Yang F, Wang JC, Han JL, Zhao G, Jiang W. (2008a) Different effects of mild and severe seizures on hippocampal neurogenesis in adult rats. Hippocampus 18:460–468. [PubMed]
  • Yang P, Arnold SA, Habas A, Hetman M, Hagg T. (2008b) Ciliary neurotrophic factor mediates dopamine D2 receptor-induced CNS neurogenesis in adult mice. J Neurosci 28:2231–2241.[PubMed]
  • Yanpallewar SU, Fernandes K, Marathe SV, Vadodaria KC, Jhaveri D, Rommelfanger K, Ladiwala U, Jha S, Muthig V, Hein L, et al. (2010) Alpha2-adrenoceptor blockade accelerates the neurogenic, neurotrophic, and behavioral effects of chronic antidepressant treatment. J Neurosci 30:1096–1109. [PMC free article] [PubMed]
  • Yoshimizu T, Chaki S. (2004) Increased cell proliferation in the adult mouse hippocampus following chronic administration of group II metabotropic glutamate receptor antagonist, MGS0039. Biochem Biophys Res Commun 315:493–496. [PubMed]
  • Yu JM, Kim JH, Song GS, Jung JS. (2006) Increase in proliferation and differentiation of neural progenitor cells isolated from postnatal and adult mice brain by Wnt-3a and Wnt-5a.Mol Cell Biochem 288:17–28. [PubMed]
  • Yu TS, Zhang G, Liebl DJ, Kernie SG. (2008) Traumatic brain injury-induced hippocampal neurogenesis requires activation of early nestin-expressing progenitors. J Neurosci 28:12901–12912. [PMC free article] [PubMed]
  • Zaben M, Sheward WJ, Shtaya A, Abbosh C, Harmar AJ, Pringle AK, Gray WP. (2009) The neurotransmitter VIP expands the pool of symmetrically dividing postnatal dentate gyrus precursors via VPAC2 receptors or directs them toward a neuronal fate via VPAC1 receptors.Stem Cells 27:2539–2551. [PubMed]
  • Zemo DA, McCabe JT. (2001) Salt-loading increases vasopressin and vasopressin 1b receptor mRNA in the hypothalamus and choroid plexus. Neuropeptides 35:181–188. [PubMed]
  • Zhang C, Wang ZJ, Lok KH, Yin M. (2012) β-Amyloid42 induces desensitization of CXC chemokine receptor-4 via formyl peptide receptor in neural stem/progenitor cells. Biol Pharm Bull 35:131–138. [PubMed]
  • Zhao M, Momma S, Delfani K, Carlen M, Cassidy RM, Johansson CB, Brismar H, Shupliakov O, Frisen J, Janson AM. (2003) Evidence for neurogenesis in the adult mammalian substantia nigra. Proc Natl Acad Sci USA 100:7925–7930. [PMC free article] [PubMed]
  • Zhao CS, Overstreet-Wadiche L. (2008) Integration of adult generated neurons during epileptogenesis. Epilepsia 49 (Suppl 5):3–12. [PubMed]
  • Zheng W, Zhuge Q, Zhong M, Chen G, Shao B, Wang H, Mao X, Xie L, Jin K. (2011)Neurogenesis in adult human brain after traumatic brain injury. J Neurotraumahttp://dx.doi.org/10.1089/neu.2010.1579.
  • Zhou BP, Liao Y, Xia W, Spohn B, Lee MH, Hung MC. (2001) Cytoplasmic localization of p21Cip1/WAF1 by AKT-induced phosphorylation in HER-2/neu-overexpressing cells. Nat Cell Biol 3:245–252. [PubMed]
  • Zhou CJ, Zhao C, Pleasure SJ. (2004) Wnt signaling mutants have decreased dentate granule cell production and radial glial scaffolding abnormalities. J Neurosci 24:121–126. [PubMed]
  • Zou YR, Kottmann AH, Kuroda M, Taniuchi I, Littman DR. (1998) Function of the chemokine receptor CXCR4 in haematopoiesis and in cerebellar development. Nature 393:595–599.[PubMed]

Articles from Pharmacological Reviews are provided here courtesy of American Society for Pharmacology and Experimental Therapeutics
pot logo blog2