Review Neural Stem Cells and Stroke - SAGE Journals

0 downloads 0 Views 234KB Size Report
Recently, neural stem cells (NSCs) as therapy for stroke have been developed in preclinical studies. NSCs are harbored in the subventricular zone (SVZ) as well ...
Cell Transplantation, Vol. 22, pp. 619–630, 2013 Printed in the USA. All rights reserved. Copyright  2013 Cognizant Comm. Corp.

0963-6897/13 $90.00 + .00 DOI: http://dx.doi.org/10.3727/096368912X655091 E-ISSN 1555-3892 www.cognizantcommunication.com

Review Neural Stem Cells and Stroke Dah-Ching Ding,*1 Chen-Huan Lin,†1 Woei-Cherng Shyu,†‡ and Shinn-Zong Lin†‡§ *Department of Obstetrics and Gynecology, Buddhist Tzu Chi General Hospital, Tzu Chi University, Hualien, Taiwan, ROC †Center for Neuropsychiatry, China Medical University Hospital, Taichung, Taiwan, ROC ‡Graduate Institute of Immunology, China Medical University, Taichung, Taiwan, ROC §Department of Neurosurgery, China Medical University Beigang Hospital, Yunlin, Taiwan, ROC

Acute ischemic stroke causes a disturbance of neuronal circuitry and disruption of the blood–brain barrier that can lead to functional disabilities. At present, thrombolytic therapy inducing recanalization of the occluded vessels in the cerebral infarcted area is a commonly used therapeutic strategy. However, only a minority of patients have timely access to this kind of therapy. Recently, neural stem cells (NSCs) as therapy for stroke have been developed in preclinical studies. NSCs are harbored in the subventricular zone (SVZ) as well as the subgranular zone of the brain. The microenvironment in the SVZ, including intercellular interactions, extracellular matrix proteins, and soluble factors, can promote NSC proliferation, self-renewal, and multipotency. Endogenous neurogenesis responds to insults of ischemic stroke supporting the existence of remarkable plasticity in the mammalian brain. Homing and integration of NSCs to the sites of damaged brain tissue are complex morphological and physiological processes. This review provides an update on current preclinical cell therapies for stroke, focusing on neurogenesis in the SVZ and dentate gyrus and on recruitment cues that promote NSC homing and integration to the site of the damaged brain. Keywords: Stroke; Homing; Stem cells; Subventricular zone (SVZ); Dentate gyrus; Regeneration

INTRODUCTION Stroke is one of the leading causes of death worldwide, but approximately two thirds of stroke victims survive (17,83). Neural cell dysfunction in the stroke brain region will cause functional disability. Although stroke is a major cause of death, the greatest health and social burden of stroke is disability (41). Thrombolytic treatment using recombinant tissue plasminogen activator has significantly improved the survival. However, only a small proportion of patients can be rescued by thrombolytic treatments due to a limited treatment window (139). It is known that natural neuronal regeneration can occur in response to brain damage, but in general, neuronal proliferation usually fails to repair the damaged brain completely. Therefore, development of new strategies for treating stroke is essential, and stem cell therapy has shown promising preclinical results (44).

Stem cells have, uniquely, the capacity for endless selfrenewal, the potential to differentiate, and depending on their origin, the ability to repopulate various organs with multiple cell types (28). Neural stem cells (NSCs) show tropism to sites of pathology in ischemic stroke (102,124). These remarkable abilities make NSCs an optimal tool for the treatment of stroke that requires a regenerative approach, either by integration into damaged tissue and replacement of missing cells or as delivery vehicles for drugs, growth factors, or functional proteins. NSC therapy for ischemic stroke can be categorized into endogenous and exogenous types. The mechanisms of NSC migration and homing to the injury site are basic to the application of this therapeutic strategy in neuro­degenerative disease. Here, we discuss the features of the normal neurogenic niche in the adult mammalian sub­ventricular zone (SVZ) as well as homing and integration of the NSCs.

Online prepub date: November 1, 2012. 1 These authors provided equal contribution to this work. Address correspondence to Shinn-Zong Lin, M.D., Ph.D., Center for Neuropsychiatry, China Medical University and Hospital, Taichung, Taiwan, ROC. Tel: +886-4-22052121 ext. 6034; Fax: +886-4-220806666; E-mail: [email protected] or Woei-Cherng Shyu, M.D., Ph.D., Center for Neuropsychiatry, China Medical University Hospital, Taichung, Taiwan, ROC. Tel: +886-4-22052121 ext. 7811; Fax: +886-4-22052121 ext. 7810; E-mail: [email protected]

619

620

Presence of NSCs in the Adult Brain It was once thought that neurogenesis occurred only during embryonic development and that the adult central nervous system (CNS) could not generate new neurons (44). However, extensive studies in recent years have shown that this is not true (130). In fact, neurogenesis can occur in two regions of the adult CNS: the SVZ and the subgranular zone of the dentate gyrus (31,37). The progenitor cells of the SVZ region and the subgranular zone of the dentate gyrus can proliferate, migrate, and differentiate into neurons, astrocytes, and oligodendrocytes. The NSCs derived from these zones show the long-term capacity for self-renewal and multipotency. Additionally, NSCs can also be found in other regions of the brain and peripheral nervous system such as the striatum, spinal cord, neocortex, and the adult human olfactory bulb (95,96,138). Subventricular Zone and Neural Stem Cells NSCs have the ability to self-renew, proliferate, and generate multiple cellular lineages both in vitro and in vivo (65). NSCs are present in the CNS throughout life in the SVZ of the adult human brain (108,109). In the SVZ, cells that express glial fibrillary astrocytic protein (GFAP) constitute the genuine NSCs (30). The niche comprises cellular structures, extracellular matrix proteins, and soluble factors (5). The SVZ is one of the main neural stem cell niches. Subventricular Zone (SVZ) The SVZ is located on the borders of the lateral ventricles of the adult mammalian brain (3). A layer of ependymal cell lines the ventricular surface and consists of E1 cells, with multiple basal bodies and multiple long cilia, and E2 cells, with two cilia and complex basal bodies that may act as mechanical or chemical sensors monitoring the composition and flow of cerebrospinal fluid (CSF) (88). The human SVZ is quite different from that of other species because of the presence of a hypocellular gap and a prominent “astrocytic ribbon” in the lateral wall of the lateral ventricles (110). Cell–cell interactions occur in the wall of the lateral ventricle, in the fashion of a “pinwheel architecture,” which refers to the organization of the ependymal lining where processes of astrocytes, in contact with ventricles, are surrounded by ependymal cells (88). However, this organization is not present in the ventricular wall of nonneurogenic regions, suggesting an important role for this intercellular interaction and for contact with the CSF in neurogenesis. NSCs in the adult rodent SVZ are present in a vascular niche, where they interact closely with blood vessels and with other cells (88,119). Endothelial cells comprise vascular structures that contribute to the SVZ neurogenic niche

Ding ET AL.

by releasing soluble factors such as pigment epitheliumderived factor (PEDF), brain-derived neurotrophic factor (BDNF), erythropoietin (EPO), and vascular endothelial growth factor (VEGF) (Fig. 1) (63,70,107,112,120). Extracellular matrix proteins also play an important role in maintaining the SVZ neurogenic niche. Tenascin C, laminin, and collagen type 1 are expressed in the extracellular matrix of the SVZ (Fig. 1) (40,84). These extracellular matrix structures may concentrate growth factors and may be important in the regulation of the microenvironment in the SVZ (84). Soluble factors participate in the maintenance of the SVZ neurogenic niche. Ependymal cells produce noggin, which regulates NSC differentiation (73). Growth factors are important players in the regulation of NSC proliferation and self-renewal. Epidermal growth factor (EGF), basic fibroblast growth factor (bFGF), and hepatocyte growth factor (HGF) contribute to the maintenance of self-renewal and proliferation of NSCs in the SVZ (Fig. 1) (42,91,113). Dentate Gyrus and NSCs The presence of NSCs has also been confirmed in the subgranular zone (SGZ) at the dentate gyrus–hilus interface (61,62,100) and is linked with mood and hippo­ campal function (32,86). Neurogenesis persists in the hippocampus of rats up to 11 months of age (16). However, hippocampal neurogenesis is drastically reduced in aged rats. The neuronal markers NeuN and calbindin-D28k are expressed in NSCs migrating into the SGZ. Stem cells can be isolated from hippocampus-adjacent regions of the subependyma, but the adult dentate gyrus does not contain a population of resident neural stem cells. Indeed, although the lateral ventricle and other ventricular subependymal regions directly adjacent to the hippocampus contain neural stem cells that exhibit long-term self-renewal and multipotentiality, separate neuronal and glial progenitors with limited self-renewal capacity are present in the adult dentate gyrus (118). However, other researchers found considerable capacity for selfrenewal and multipotential progenitors in the dentate gyrus (39,97,98). Adult neurogenesis is dynamic and is thought to consist of proliferation of nestin-expressing NSCs, transiently amplifying progenitors (TAPs), maturation of double­ cortin (DCX)-expressing neuroblasts, and survival of adult-born neurons, which integrate into hippocampal circuitry (33,55). Each stage is discretely regulated by a variety of intrinsic and extrinsic factors, and modulation of adult SGZ neurogenesis by myriad stimuli is the focus of intense research (34,144). However, interaction between NSCs, TAPs, and the neurogenic microenvironment or neurogenic niche needs to be clarified. Recently, it has been reported that Notch 1 is required for maintenance of the reservoir of adult hippocampal stem cells (Fig. 1) (2).

NEURAL STEM CELLS and STROKE

621

Figure 1. Schematic of subventricular zone (SVZ) and dentate gyrus neural progenitor cell (NPC) neurogenesis in the adult brain. (1) Growth factors such as epidermal growth factor (EGF), basic fibroblast growth factor (bFGF), and hepatocyte growth factor (HGF) contribute to the maintenance of self-renewal and proliferation of neural stem cells (NSCs) in the SVZ. (2) Contribution to the SVZ neurogenic niche by endothelial cells composing vascular structures are soluble factors such as pigment epithelium-derived factor (PEDF), brain-derived neurotrophic factor (BDNF), erythropoietin (EPO), and vascular endothelial growth factor (VEGF). (3) Extracellular matrix proteins also play an important role in maintaining the SVZ neurogenic niche. Tenascin C, laminin, and collagen type 1 are expressed in the extracellular matrix of the SVZ. (4) Notch 1 is required for maintenance of the reservoir of adult hippocampal stem cells. (5) CNS injury where factors such as stromal cell-derived factor-1 (SDF-1), leukemic inhibitory factor (LIF), and interleukin (IL)-6 are overexpressed. These factors are important for NSC recruitment and homing.

Source of Neural Stem Cells The availability of NSCs is the first consideration in treatment of neurological diseases. Stem cell therapy for ischemic stroke can be categorized into endogenous and exogenous types. Endogenous Neural Stem Cells Multipotent NSCs can be isolated from the adult rodent CNS and cultivated in vitro (113,136) as well as in adult human brain tissue. Autologous transplantation of NSCs is a potential therapy for stroke. In a rat model of spinal cord injury, autologous transplantation of NSCs derived from SVZ was found to be feasible (103). However, there is very little experimental data regarding adult NSC transplantation in models of stroke. One study investigated the effects of adult-derived exogenous NSC

transplantation in the cortex of ischemic stroke in adult rats (143). The researchers used histology and magnetic resonance imaging (MRI) to analyze the transplanted NSC survival and migration, and they observed some functional recovery. In another study of embolic stroke, neural progenitor cells isolated from the SVZ of the adult rat were labeled by superparamagnetic particles and intracisternally transplanted into the adult rat 48 h after stroke (51). It was found that MRI could identify the cerebral tissue destined to undergo angiogenesis after cell therapy. Studies show that new neurons are produced in the SVZ after a stroke (7,60). New cells migrate and integrate into the damaged tissue, expressing neuronal markers and acquiring a morphological phenotype that resembles neurons after experimental occlusion of the middle cerebral artery (7). Other studies

622

have shown an increase in neurogenesis after ischemic stroke (99,137,142). These studies provide evidence of the regenerative capacity present in the injured brain, but this potential does not seem to be sufficient and different manipulations may be necessary to potentiate endogenous regeneration. Manipulation of endogenous neurogenesis and “encouragement” of migration to sites of injury are important in the therapeutic use of endogenous NSCs. Chemotactic and humoral factors produced during disease serve as homing cues that cause NSC migration to the site of injury (35). Further studies are needed to exploit the endogenous regenerative ability of the mammalian brain. Exogenous Neural Stem Cells Exogenous NSC therapy for stroke is used primarily to reconstruct destroyed neuronal circuits. These cell grafts may exert a neuroprotective effect by secreting survivalpromoting neurotrophic factors either intrinsically or after the introduction of therapeutic transgenes (44). NSCs can be obtained from multiple tissues—skin (64,71,133), embryonic stem cells (36), embryonic NSCs (140), bone marrow and adipose-derived mesenchymal stem cells (27,38,123), fetal stem cells (26), peripheral blood stem cells (122), olfactory ensheathing cells (125), induced pluripotent stem cells (6,19), and fetal and adult nervous systems (4,116,135). Genetic manipulation of isolated NSCs may be necessary to obtain sufficient numbers of stem cells and achieve feasibility in NSC-based therapy (94). Telomerase reverse transcriptase (TERT) and myelocytomatosis viral oncogene homolog (Myc) were used to establish NSC lines (56,115). However, Myc could cause genetic instability of an established cell line. It is essential to thoroughly characterize a cell line before starting human treatment, since isolated NSCs can become tumorigenic after serial passaging and transplantation (121,126). With further research, it may become possible to develop reliable and thoroughly characterized NSC lines for use in the treatment of neurodegenerative diseases. Neural Stem Cell Homing and Recruitment How NSCs home to injury sites is important for cellbased regenerative therapy. NSCs are highly migratory and seem to be attracted to injured brain areas such as ischemic regions (20). While stem cell homing to bone marrow has been widely studied (89), the molecular basis of stem cell pathotrophism is not well understood. Further identification of the mechanisms involved would pave the way for the development of treatments to enhance endogenous mobilization of stem cells in disease states, perhaps

Ding ET AL.

using small molecules. NSCs express a wide variety of receptors that may enable them to respond to many chemotactic signals that emanate from brain pathologies. Chemokine and cytokine production is a common feature of many brain lesions, including stroke, which suggests that these factors could be important in mediating the responses of stem cells to injuries. Inflammation is a key player in the homing and recruitment of NSCs to sites of CNS injury where factors such as stromal cell-derived factor-1 (SDF-1), leukemic inhibitory factor (LIF), and interleukin (IL)-6 are overexpressed (14,46,124,128) (Fig. 1). SDF-1 and CXC Chemokine Receptor-4 (CXCR4). Mutual, reciprocal SDF-1/CXCR4 interaction in NSCs and the ischemic stroke region has been reported (46,85,114,131). During steady-state homeostasis, CXCR4 is expressed by hematopoietic cells and also by stromal cells, which are the main source for SDF-1 in the bone marrow (BM). Stress will increase SDF-1 and CXCR4 levels, which stimulates recruitment of immature and maturing leukocytes from the BM reservoir to damaged organs as part of host defense and repair mechanisms (1,57). Astrocytes and endothelial cells of an ischemic area in the brain increase the expression of SDF-1 and trigger constitutive expression and activation of CXCR4, the receptor for SDF-1, resulting in NSC migration toward ischemic brain explants (49). Exposure to SDF-1 enhances migration and proliferation of Lewis ´ [stage-specific embryonic antigen 1 (SSEA-1); LeX]-positive NSCs, which can differentiate into cells with neuronal phenotypes after transplantation (22). SDF-1 is upregulated in traumatic brain injury (50), suggesting a recruiting mechanism for NSCs to traumatic lesions in the brain. Leukemic Inhibitory Factor and Interleukin-6. LIF is a member of the IL-6 cytokine family, which also includes ciliary neurotrophic factor (CNTF) and IL-6, among others (13,45). In the nervous system, endogenous LIF expression is very low under normal physiological conditions, but it is systematically induced following ischemia (23,128). Relevance of this rise of LIF has been emphasized in the regenerating adult olfactory epithelium where it is necessary for the lesion-induced proliferation of neuronal progenitors (15). Exogenous LIF will promote NSC self-renewal in vivo (14), suggesting that the very rapid and transient overexpression of endogenous LIF observed after lesion could constitute a signal that recruits and amplifies NSCs, making their regenerative potential available for brain repair. Similar results have been obtained with IL-6, another inflammatory cytokine, which decreases the proliferation of neuroblasts and enhances NSC self-renewal,

NEURAL STEM CELLS and STROKE

as demonstrated in a transgenic mouse model of IL-6 overexpression (134). IL-6 may also be involved in the differentiation of NSCs into neurons (12,13). Integrins Integrins are heterodimeric transmembrane molecules consisting of α and β subunits that mediate cell adhesion and migration (21,48). Homing of progenitor cells such as NSCs to ischemic tissue may follow the paradigm of mature leukocytes migrating to inflammatory tissue (29,82,106,119). We have shown that peripheral blood stem cell (PBSC) intracerebral transplantation can significantly improve neurological function following chronic cerebral ischemia in rats, accompanied by increased local cortical cerebral blood flow and β1-integrin expression in the ischemic hemisphere (122). The neurological improvement in this study was blocked by β1-integrin inhibitor (synthetic RDG peptide) (26,122). These results reveal that β1-integrin is necessary for neuroplasticity after intracerebral stem cell transplantation, possibly through the enhanced angiogenesis by homing of stem cells to ischemic sites (26,72,79,122). Whether β1-integrins play the same cellular role or whether different cell types use distinct mechanisms for homing remains to be determined. Further studies are needed to elucidate whether there is a synergism between other adhesion molecules and their counter­ligands in the multistep recruitment of NSCs to ischemic tissue (72,79). Modulation of these integrins may provide novel opportunities for treating cerebral ischemic disease (Fig. 1). Neural Stem Cell Therapy for Stroke Extensive ischemic injury is a neurological disorder caused by multiple factors such as hypoxia, which is a common cause of neurological disability in adults and children. This disorder is characterized by extensive damage of cerebral parenchyma, resulting in formation of a cystic cavity and consequently loss of neural cells and their connections. It leads to the death of multiple neuron types, as well as oligodendrocytes, astrocytes, and endothelial cells (75). Neuronal plasticity and reorganization of neural circuitries contribute to varying degrees of spontaneous recovery, but the majority of patients exhibit persistent motor, sensory or cognitive impairments. At present, there are no effective clinical therapies. Many preclinical studies have used NSCs to treat ische­ mic stroke animal models. Most animal models of ische­ mic stroke entail occlusion of arterial blood supply to the brain by surgical techniques (26,122). Researchers have used these models to study the response of endogenous neurogenesis to stroke and found that hypoxic injury to the brain leads to increased neurogenesis in the SVZ

623

(7,52,99). Moreover, studies of the cues that increase neurogenesis and migration of these NSCs to the site of injury showed that SDF-1 and angiopoietin-1 contribute to the homing process of newly formed neural progenitors to the site of injury (92). Stroke-induced neuro­genesis is maintained in the aged rat brain (24). There is also evidence in humans for enhanced SVZ cell proliferation and neuroblast formation after stroke (54,81,87). In mice, ependymal cells lining the lateral ventricle participate in the neurogenic response to stroke by producing new neuro­blasts, but their survival is poor (18). Different stem cells and their derivatives of rodent and human origin can survive, differentiate into neurons, and restore function after transplantation in the strokedamaged rodent brain (9,78). Intracerebral implantation of NSCs after ischemia is a feasible method of administration of regenerative therapy (53,141). Genetically modified [overexpression of BDNF, nerve growth factor (NGF)] NSCs also have been shown to improve function in a mouse stroke model (68,127). In addition, overexpression of either VEGF or the antiapoptotic factor v-akt murine thymoma viral oncogene homolog 1 (Akt1) in human NSCs promotes angiogenesis and increases neuronal survival, respectively, enhancing the functional improvements in stroke-damaged mice (66,67). Intravenously administered NSCs, previously sorted for the expression of a surface antigen CD49d, home to ischemic lesions in the brain, and thus improve functional recovery after ischemia (43). Intravenously administered human mesenchymal stem cells (MSCs) have been found to reduce stroke-induced deficits in rats, most likely by inducing angiogenesis and improving cerebral blood flow (93). Human embryonic stem (ES) cell-derived MSCs injected intravenously in rats have been shown to migrate to the infarct area, express neuronal and endothelial cell markers, provide neuroprotection, and improve recovery (76). Mouse NSCs delivered intravenously 3 days after stroke in mice have been reported to suppress inflammation and glial scar formation and give rise to delayed neuro­protection and improved functional recovery, starting 18 days after the insult (10). These findings suggest that there is an extended time window for neuroprotection using NSCs. To study the different therapeutic effects of the administration of embryonic versus adult NSCs, rats were subjected to middle cerebral artery occlusion treated with either type of stem cell. It was found that ESCs survive longer and in larger proportions than adult NSCs and that the administration of either cell line decreased the infarct size (129). Transplanted human NSCs can survive and differentiate into neurons that express different neuronal markers (25). Taken together, these data

624

suggest that NSC transplantation is a potential regenerative therapy for stroke. Clinical Trials of Stem Cell Transplantation for Stroke Currently, stem cell therapy in stroke patients is in its infancy. Several small clinical trials with delivery of stem cells in stroke have been completed (78). Slight improvements were detected in some patients after cells from an immortalized human teratocarcinoma cell line were implanted into ischemic or hemorrhagic infarcts in the basal ganglia and cerebral cortex (58,90). However, no substantial clinical improvements were detected after intravenous injection of autologous MSCs in ischemic patients (11). Several clinical trials using intravenous or intra-arterial infusion of autologous bone marrowderived and other stem cells in stroke patients are ongoing or planned (www.clinicaltrials.gov). One trial will use conditionally immortalized NSCs isolated from human fetal cortex to treat stroke patients. Conditionally immortalized NSCs have been shown to decrease motor impairments in a rat stroke model (105), possibly by promoting angiogenesis and improving cerebral blood flow. In summary, these small initial human studies cannot be comparable due to differences in target population, type of cells, timing of injection and mode of delivery,

Ding ET AL.

but they indicate that stem cell therapy may be technically feasible in stroke patients. Side effects were observed in these clinical trials (58,111,117); therefore, safety is still a principle concern. Epilepsy, risk of bleeding or thrombosis at the site of injection and risk of malignant transformation are the major adverse effects that have been observed. No tumorigenic potential of the cells has so far been observed in two clinical trials (11,59). SUMMARY The process of neurogenesis in the mammalian brain and different interactions in the SVZ and dentate gyrus have been revealed in animal studies. This article illustrated the recruitment and integration of NSCs in the stroke mammalian brain. Cell–cell interaction and cell– extracellular matrix interactions are critical to this complex process, including proliferation, differentiation, and migration of NSCs. Growth factors and cytokines contribute to the proliferation, homing, and recruitment of NSCs in the diseased brain. Future Directions of Neural Stem Cell Therapy for Stroke Stem cell research could lead to the development of radical new therapies for neurodegenerative diseases that currently lack effective treatments. Over the past few

Figure 2. Stem cell therapy for stroke. Ischemic stroke leads to the death of multiple neuronal types, as well as astrocytes, oligodendrocytes, and endothelial cells in the cortex and subcortical regions. Stem cell therapy could be used to recovery damaged neural circuitry by transplanting stem cell-derived neuron precursors. Additionally, compounds could be infused that would promote neurogenesis from endogenous SVZ neural stem cells or stem cells could be injected systemically for modulation of inflammation and neuroprotection.

NEURAL STEM CELLS and STROKE

years, there has been continuous progress in developing different kinds of NSCs derived from different origins. Patient-specific cells that may be useful for transplantation can now be produced from induced pluripotent stem (iPS) cells (101). Also, NSCs in the adult brain generate new neurons in response to neurodegeneration (74,142). It has become clear that the characteristics of the pathological environment, such as inflammation, play an important role in the survival, differentiation, and function of both grafted and endogenous cells (47,77,132). Besides cell replacement, stem cells are known to lead to improvements that could also be of clinical value through immunomodulation, trophic actions, neuroprotection and angiogenesis (66,67,69,93,104). Therapeutic approaches using stem cells mainly for neuroprotection by supplying neurotrophic molecules or modulating inflammation will most likely be applied soon in stroke models. The spectrum of stem cell therapy for stroke is illustrated in Figure 2. However, the use of clinical pearls of stem cells to treat stroke will require more basic research so that the mechanisms regulating the proliferation, migration, differentiation, survival, and function of stem cells and their derivatives are better understood and can be effectively controlled. Once researchers have identified a reliable and consistent source of NSCs, it will be essential to determine whether transplanted cells can integrate into the normal anatomy of the damaged area. Research has shown that NSCs transplanted into rodent brains are capable of integration into neuronal circuitry and have the phenotypical and electrical properties of neurons (8,80). Before clinical application of these cells, preclinical evidence of efficacy and safety of various sources of NSCs is needed. Most stroke patients have few or no therapeutic options and are prepared to receive any new treatments. Harnessing the full potential of NSCs requires much more insight into the mechanisms behind regeneration. Clinical translation of stem cell research into treatment applicable to stroke patients will require the collaboration of scientists and clinicians. ACKNOWLEDGMENTS: This work was supported in part by grants from the Chen-Han Foundation for Education, Taiwan Department of Health Clinical Trial and Research Center of Excellence (DOH99-TD-B-111-004), the Topnotch Stroke Research Center (China Medical University, CMU98-CT-24), the National Science Council (NSC97-2314-B-039-036-MY3, Taiwan), China Medical University (DMR-98-120), and Buddhist Tzu Chi General Hospital (TCRD 99-12). The authors declare no conflict of interest.

References   1. Abkowitz, J. L.; Robinson, A. E.; Kale, S.; Long, M. W.; Chen, J. Mobilization of hematopoietic stem cells during homeostasis and after cytokine exposure. Blood 102(4):1249–1253; 2003.

625

  2. Ables, J. L.; Decarolis, N. A.; Johnson, M. A.; Rivera, P. D.; Gao, Z.; Cooper, D. C.; Radtke, F.; Hsieh, J.; Eisch, A. J. Notch1 is required for maintenance of the reservoir of adult hippocampal stem cells. J. Neurosci. 30(31):10484– 10492; 2010.   3. Agasse, F.; Nicoleau, C.; Petit, J.; Jaber, M.; Roger, M.; Benzakour, O.; Coronas, V. Evidence for a major role of endogenous fibroblast growth factor-2 in apoptotic cortexinduced subventricular zone cell proliferation. Eur. J. Neurosci. 26(11):3036–3042; 2007.   4. Alvarez-Buylla, A.; Garcia-Verdugo, J. M. Neurogenesis in adult subventricular zone. J. Neurosci. 22(3):629–634; 2002.   5. Alvarez-Buylla, A.; Lim, D. A. For the long run: Maintaining germinal niches in the adult brain. Neuron 41(5):683– 686; 2004.   6. Amabile, G.; Meissner, A. Induced pluripotent stem cells: Current progress and potential for regenerative medicine. Trends Mol. Med. 15(2):59–68; 2009.   7. Arvidsson, A.; Collin, T.; Kirik, D.; Kokaia, Z.; Lindvall, O. Neuronal replacement from endogenous precursors in the adult brain after stroke. Nat. Med. 8(9):963–970; 2002.   8. Auerbach, J. M.; Eiden, M. V.; McKay, R. D. Transplanted CNS stem cells form functional synapses in vivo. Eur. J. Neurosci. 12(5):1696–1704; 2000.   9. Bacigaluppi, M.; Pluchino, S.; Martino, G.; Kilic, E.; Hermann, D. M. Neural stem/precursor cells for the treatment of ischemic stroke. J. Neurol. Sci. 265(1–2):73–77; 2008. 10. Bacigaluppi, M.; Pluchino, S.; Peruzzotti-Jametti, L.; Kilic, E.; Kilic, U.; Salani, G.; Brambilla, E.; West, M. J.; Comi, G.; Martino, G.; Hermann, D. M. Delayed post-ischemic neuroprotection following systemic neural stem cell transplantation involves multiple mechanisms. Brain 132(Pt 8): 2239–2251; 2009. 11. Bang, O. Y.; Lee, J. S.; Lee, P. H.; Lee, G. Autologous mesenchymal stem cell transplantation in stroke patients. Ann. Neurol. 57(6):874–882; 2005. 12. Barkho, B. Z.; Song, H.; Aimone, J. B.; Smrt, R. D.; Kuwabara, T.; Nakashima, K.; Gage, F. H.; Zhao, X. Identification of astrocyte-expressed factors that modulate neural stem/progenitor cell differentiation. Stem Cells Dev. 15(3):407–421; 2006. 13. Bauer, S.; Kerr, B. J.; Patterson, P. H. The neuropoietic cytokine family in development, plasticity, disease and injury. Nat. Rev. Neurosci. 8(3):221–232; 2007. 14. Bauer, S.; Patterson, P. H. Leukemia inhibitory factor promotes neural stem cell self-renewal in the adult brain. J. Neurosci. 26(46):12089–12099; 2006. 15. Bauer, S.; Rasika, S.; Han, J.; Mauduit, C.; Raccurt, M.; Morel, G.; Jourdan, F.; Benahmed, M.; Moyse, E.; Patterson, P. H. Leukemia inhibitory factor is a key signal for injuryinduced neurogenesis in the adult mouse olfactory epithelium. J. Neurosci. 23(5):1792–1803; 2003. 16. Bottai, D.; Fiocco, R.; Gelain, F.; Defilippis, L.; Galli, R.; Gritti, A.; Vescovi, L. A. Neural stem cells in the adult nervous system. J. Hematother. Stem Cell Res. 12(6):655–670; 2003. 17. Broderick, J.; Brott, T.; Kothari, R.; Miller, R.; Khoury, J.; Pancioli, A.; Gebel, J.; Mills, D.; Minneci, L.; Shukla, R. The Greater Cincinnati/Northern Kentucky Stroke Study: Preliminary first-ever and total incidence rates of stroke among blacks. Stroke 29(2):415–421; 1998.

626

18. Carlen, M.; Meletis, K.; Goritz, C.; Darsalia, V.; Evergren, E.; Tanigaki, K.; Amendola, M.; Barnabe-Heider, F.; Yeung, M. S.; Naldini, L.; Honjo, T.; Kokaia, Z.; Shupliakov, O.; Cassidy, R. M.; Lindvall, O.; Frisén, J.. Forebrain ependymal cells are Notch-dependent and generate neuroblasts and astrocytes after stroke. Nat. Neurosci. 12(3):259–267; 2009. 19. Chambers, S. M.; Fasano, C. A.; Papapetrou, E. P.; Tomishima, M.; Sadelain, M.; Studer, L. Highly efficient neural conversion of human ES and iPS cells by dual inhibition of SMAD signaling. Nat. Biotechnol. 27(3):275–280; 2009. 20. Chen, J.; Zhang, C.; Jiang, H.; Li, Y.; Zhang, L.; Robin, A.; Katakowski, M.; Lu, M.; Chopp, M. Atorvastatin induction of VEGF and BDNF promotes brain plasticity after stroke in mice. J. Cereb. Blood Flow Metab. 25(2):281–290; 2005. 21. Chute, J. P. Stem cell homing. Curr. Opin. Hematol. 13(6):399–406; 2006. 22. Corti, S.; Locatelli, F.; Papadimitriou, D.; Donadoni, C.; Del Bo, R.; Fortunato, F.; Strazzer, S.; Salani, S.; Bresolin, N.; Comi, G. P. Multipotentiality, homing properties, and pyramidal neurogenesis of CNS–derived LeX(ssea-1)+/ CXCR4+ stem cells. FASEB J. 19(13):1860–1862; 2005. 23. Covey, M. V.; Levison, S. W. Leukemia inhibitory factor participates in the expansion of neural stem/progenitors after perinatal hypoxia/ischemia. Neuroscience 148(2):501–509; 2007. 24. Darsalia, V.; Heldmann, U.; Lindvall, O.; Kokaia, Z. Stroke-induced neurogenesis in aged brain. Stroke 36(8):1790–1795; 2005. 25. Darsalia, V.; Kallur, T.; Kokaia, Z. Survival, migration and neuronal differentiation of human fetal striatal and cortical neural stem cells grafted in stroke-damaged rat striatum. Eur. J. Neurosci. 26(3):605–614; 2007. 26. Ding, D. C.; Shyu, W. C.; Chiang, M. F.; Lin, S. Z.; Chang, Y. C.; Wang, H. J.; Su, C. Y.; Li, H. Enhancement of neuroplasticity through upregulation of b1-integrin in human umbilical cord-derived stromal cell implanted stroke model. Neurobiol. Dis. 27(3):339–353; 2007. 27. Ding, D. C.; Shyu, W. C.; Lin, S. Z. Mesenchymal stem cells. Cell Transplant. 20(1):5–14; 2011. 28. Ding, D. C.; Shyu, W. C.; Lin, S. Z.; Li, H. Current concepts in adult stem cell therapy for stroke. Curr. Med. Chem. 13(29):3565–3574; 2006. 29. Doetsch, F. A niche for adult neural stem cells. Curr. Opin. Genet. Dev. 13(5):543–550; 2003. 30. Doetsch, F.; Caille, I.; Lim, D. A.; Garcia-Verdugo, J. M.; Alvarez-Buylla, A. Subventricular zone astrocytes are neural stem cells in the adult mammalian brain. Cell 97(6):703– 716; 1999. 31. Doetsch, F.; Garcia-Verdugo, J. M.; Alvarez-Buylla, A. Cellular composition and three-dimensional organization of the subventricular germinal zone in the adult mammalian brain. J. Neurosci. 17(13):5046–5061; 1997. 32. Doetsch, F.; Hen, R. Young and excitable: The function of new neurons in the adult mammalian brain. Curr. Opin. Neurobiol. 15(1):121–128; 2005. 33. Duan, X.; Kang, E.; Liu, C. Y.; Ming, G. L.; Song, H. Development of neural stem cell in the adult brain. Curr. Opin. Neurobiol. 18(1):108–115; 2008. 34. Eisch, A. J.; Cameron, H. A.; Encinas, J. M.; Meltzer, L. A.; Ming, G. L.; Overstreet-Wadiche, L. S. Adult neurogenesis, mental health, and mental illness: Hope or hype? J. Neurosci. 28(46):11785–11791; 2008.

Ding ET AL.

35. Ekdahl, C. T.; Kokaia, Z.; Lindvall, O. Brain inflammation and adult neurogenesis: The dual role of microglia. Neuroscience 158(3):1021–1029; 2009. 36. Erceg, S.; Ronaghi, M.; Stojkovic, M. Human embryonic stem cell differentiation toward regional specific neural precursors. Stem Cells 27(1):78–87; 2009. 37. Eriksson, P. S.; Perfilieva, E.; Bjork-Eriksson, T.; Alborn, A. M.; Nordborg, C.; Peterson, D. A.; Gage, F. H. Neurogenesis in the adult human hippocampus. Nat. Med. 4(11):1313–1317; 1998. 38. Fu, L.; Zhu, L.; Huang, Y.; Lee, T. D.; Forman, S. J.; Shih, C. C. Derivation of neural stem cells from mesenchymal stem cells: Evidence for a bipotential stem cell population. Stem Cells Dev. 17(6):1109–1121; 2008. 39. Gage, F. H.; Coates, P. W.; Palmer, T. D.; Kuhn, H. G.; Fisher, L. J.; Suhonen, J. O.; Peterson, D. A.; Suhr, S. T.; Ray, J. Survival and differentiation of adult neuronal progenitor cells transplanted to the adult brain. Proc. Natl. Acad. Sci. USA 92(25):11879–11883; 1995. 40. Garcion, E.; Faissner, A.; ffrench-Constant, C. Knockout mice reveal a contribution of the extracellular matrix molecule tenascin-C to neural precursor proliferation and migration. Development 128(13):2485–2496; 2001. 41. Gilman, S. Time course and outcome of recovery from stroke: Relevance to stem cell treatment. Exp. Neurol. 199(1):37–41; 2006. 42. Gritti, A.; Frolichsthal-Schoeller, P.; Galli, R.; Parati, E. A.; Cova, L.; Pagano, S. F.; Bjornson, C. R.; Vescovi, A. L. 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(9):3287– 3297; 1999. 43. Guzman, R.; De Los Angeles, A.; Cheshier, S.; Choi, R.; Hoang, S.; Liauw, J.; Schaar, B.; Steinberg, G. Intracarotid injection of fluorescence activated cell-sorted CD49dpositive neural stem cells improves targeted cell delivery and behavior after stroke in a mouse stroke model. Stroke 39(4):1300–1306; 2008. 44. Haas, S.; Weidner, N.; Winkler, J. Adult stem cell therapy in stroke. Curr. Opin. Neurol. 18(1):59–64; 2005. 45. Heinrich, P. C.; Behrmann, I.; Haan, S.; Hermanns, H. M.; Muller-Newen, G.; Schaper, F. Principles of interleukin (IL)-6-type cytokine signalling and its regulation. Biochem. J. 374(Pt 1):1–20; 2003. 46. Hill, W. D.; Hess, D. C.; Martin-Studdard, A.; Carothers, J. J.; Zheng, J.; Hale, D.; Maeda, M.; Fagan, S. C.; Carroll, J. E.; Conway, S. J. SDF-1 (CXCL12) is upregulated in the ischemic penumbra following stroke: Association with bone marrow cell homing to injury. J. Neuropathol. Exp. Neurol. 63(1):84–96; 2004. 47. Hoehn, B. D.; Palmer, T. D.; Steinberg, G. K. Neurogenesis in rats after focal cerebral ischemia is enhanced by indomethacin. Stroke 36(12):2718–2724; 2005. 48. Hynes, R. O. Integrins: Bidirectional, allosteric signaling machines. Cell 110(6):673–687; 2002. 49. Imitola, J.; Raddassi, K.; Park, K. I.; Mueller, F. J.; Nieto, M.; Teng, Y. D.; Frenkel, D.; Li, J.; Sidman, R. L.; Walsh, C. A.; Snyder, E. Y.; Khoury, S. J. Directed migration of neural stem cells to sites of CNS injury by the stromal cell-derived factor 1a/CXC chemokine receptor-4 pathway. Proc. Natl. Acad. Sci. USA 101(52):18117–18122; 2004. 50. Itoh, T.; Satou, T.; Ishida, H.; Nishida, S.; Tsubaki, M.; Hashimoto, S.; Ito, H. The relationship between SDF-1a/

NEURAL STEM CELLS and STROKE

CXCR4 and neural stem cells appearing in damaged area after traumatic brain injury in rats. Neurol. Res. 31(1):90– 102; 2009. 51. Jiang, Q.; Zhang, Z. G.; Ding, G. L.; Zhang, L.; Ewing, J. R.; Wang, L.; Zhang, R.; Li, L.; Lu, M.; Meng, H.; Arbab, A. S.; Hu, J.; Li, Q. J.; Pourabdollah Nejad, D. S.; Athiraman, H.; Chopp, M. Investigation of neural progenitor cell induced angiogenesis after embolic stroke in rat using MRI. Neuroimage 28(3):698–707; 2005. 52. Jin, K.; Minami, M.; Lan, J. Q.; Mao, X. O.; Batteur, S.; Simon, R. P.; Greenberg, D. A. Neurogenesis in dentate subgranular zone and rostral subventricular zone after focal cerebral ischemia in the rat. Proc. Natl. Acad. Sci. USA 98(8):4710–4715; 2001. 53. Jin, K.; Sun, Y.; Xie, L.; Mao, X. O.; Childs, J.; Peel, A.; Logvinova, A.; Banwait, S.; Greenberg, D. A. Comparison of ischemia-directed migration of neural precursor cells after intrastriatal, intraventricular, or intravenous transplantation in the rat. Neurobiol. Dis. 18(2):366–374; 2005. 54. Jin, K.; Wang, X.; Xie, L.; Mao, X. O.; Zhu, W.; Wang, Y.; Shen, J.; Mao, Y.; Banwait, S.; Greenberg, D. A. Evidence for stroke-induced neurogenesis in the human brain. Proc. Natl. Acad. Sci. USA 103(35):13198–13202; 2006. 55. Kempermann, G.; Jessberger, S.; Steiner, B.; Kronenberg, G. Milestones of neuronal development in the adult hippocampus. Trends Neurosci. 27(8):447–452; 2004. 56. Kerosuo, L.; Piltti, K.; Fox, H.; Angers-Loustau, A.; Hayry, V.; Eilers, M.; Sariola, H.; Wartiovaara, K. Myc increases self-renewal in neural progenitor cells through Miz-1. J. Cell Sci. 121(Pt 23):3941–3950; 2008. 57. Kollet, O.; Shivtiel, S.; Chen, Y. Q.; Suriawinata, J.; Thung, S. N.; Dabeva, M. D.; Kahn, J.; Spiegel, A.; Dar, A.; Samira, S.; Goichberg, P.; Kalinkovich, A.; ArenzanaSeisdedos, F.; Nagler, A.; Hardan, I.; Revel, M.; Shafritz, D. A.; Lapidot, T. HGF, SDF-1, and MMP-9 are involved in stress-induced human CD34+ stem cell recruitment to the liver. J. Clin. Invest. 112(2):160–169; 2003. 58. Kondziolka, D.; Steinberg, G. K.; Wechsler, L.; Meltzer, C. C.; Elder, E.; Gebel, J.; Decesare, S.; Jovin, T.; Zafonte, R.; Lebowitz, J.; Flickinger, J. C.; Tong, D.; Marks, M. P.; Jamieson, C.;. Luu, D.; Bell-Stephens, T.;. Teraoka, J. Neurotransplantation for patients with subcortical motor stroke: A phase 2 randomized trial. J. Neurosurg. 103(1): 38–45; 2005. 59. Kondziolka, D.; Wechsler, L.; Goldstein, S.; Meltzer, C.; Thulborn, K. R.; Gebel, J.; Jannetta, P.; DeCesare, S.; Elder, E. M.; McGrogan, M. Reitman, M. A.; Bynum, L. Transplantation of cultured human neuronal cells for patients with stroke. Neurology 55(4):565–569; 2000. 60. Kuge, A.; Takemura, S.; Kokubo, Y.; Sato, S.; Goto, K.; Kayama, T. Temporal profile of neurogenesis in the subventricular zone, dentate gyrus and cerebral cortex following transient focal cerebral ischemia. Neurol. Res. 31(9):969–976; 2009. 61. Kuhn, H. G.; Dickinson-Anson, H.; Gage, F. H. Neurogenesis in the dentate gyrus of the adult rat: Agerelated decrease of neuronal progenitor proliferation. J. Neurosci. 16(6):2027–2033; 1996. 62. Lagace, D. C.; Whitman, M. C.; Noonan, M. A.; Ables, J. L.; DeCarolis, N. A.; Arguello, A. A.; Donovan, M. H.; Fischer, S. J.; Farnbauch, L. A.; Beech, R. D.; DiLeone, R. J.; Greer, C. A.; Mandyam, C. D.; Eisch, A. J. Dynamic

627

contribution of nestin-expressing stem cells to adult neurogenesis. J. Neurosci. 27(46):12623–12629; 2007. 63. Laguna Goya, R.; Tyers, P.; Barker, R. A. The search for a curative cell therapy in Parkinson’s disease. J. Neurol. Sci. 265(1-2):32–42; 2008. 64. Lako, M.; Armstrong, L.; Cairns, P. M.; Harris, S.; Hole, N.; Jahoda, C. A. Hair follicle dermal cells repopulate the mouse haematopoietic system. J. Cell Sci. 115(Pt 20): 3967–3974; 2002. 65. Laywell, E. D.; Steindler, D. A.; Silver, D. J. Astrocytic stem cells in the adult brain. Neurosurg. Clin. N. Am. 18(1):21-30; 2007. 66. Lee, H. J.; Kim, K. S.; Park, I. H.; Kim, S. U. Human neural stem cells over-expressing VEGF provide neuroprotection, angiogenesis and functional recovery in mouse stroke model. PLoS One 2(1):e156; 2007. 67. Lee, H. J.; Kim, M. K.; Kim, H. J.; Kim, S. U. Human neural stem cells genetically modified to overexpress Akt1 provide neuroprotection and functional improvement in mouse stroke model. PLoS One 4(5):e5586; 2009. 68. Lee, H. J.; Lim, I. J.; Lee, M. C.; Kim, S. U. Human neural stem cells genetically modified to overexpress brainderived neurotrophic factor promote functional recovery and neuroprotection in a mouse stroke model. J. Neurosci. Res. 88(15):3282–3294; 2010. 69. Lee, S. T.; Chu, K.; Jung, K. H.; Kim, S. J.; Kim, D. H.; Kang, K. M.; Hong, N. H.; Kim, J. H.; Ban, J. J.; Park, H. K.; Kim, S. U.; Park, C. G.; Lee, S. K.; Kim, M.; Roh, J. K. Anti-inflammatory mechanism of intravascular neural stem cell transplantation in haemorrhagic stroke. Brain 131(Pt 3): 616–629; 2008. 70. Leventhal, C.; Rafii, S.; Rafii, D.; Shahar, A.; Goldman, S. A. Endothelial trophic support of neuronal production and recruitment from the adult mammalian subependyma. Mol. Cell. Neurosci. 13(6):450–464; 1999. 71. Liang, L.; Bickenbach, J. R. Somatic epidermal stem cells can produce multiple cell lineages during development. Stem Cells 20(1):21–31; 2002. 72. Liao, H.; Huang, W.; Schachner, M.; Guan, Y.; Guo, J.; Yan, J.; Qin, J.; Bai, X.; Zhang, L. b 1 integrin-mediated effects of tenascin-R domains EGFL and FN6-8 on neural stem/progenitor cell proliferation and differentiation in vitro. J. Biol. Chem. 283(41):27927–27936; 2008. 73. Lim, D. A.; Tramontin, A. D.; Trevejo, J. M.; Herrera, D. G.; Garcia-Verdugo, J. M.; Alvarez-Buylla, A. Noggin antagonizes BMP signaling to create a niche for adult neurogenesis. Neuron 28(3):713–726; 2000. 74. Lindvall, O.; Kokaia, Z. Neurogenesis following stroke affecting the adult brain. In: Gage, F.; Kempermann, G.; Song, H., eds. Adult neurogenesis. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press; 2008:549–570. 75. Lindvall, O.; Kokaia, Z. Stem cells in human neuro­ degenerative disorders—time for clinical translation? J. Clin. Invest. 120(1):29–40; 2010. 76. Liu, Y. P.; Seckin, H.; Izci, Y.; Du, Z. W.; Yan, Y. P.; Baskaya, M. K. Neuroprotective effects of mesenchymal stem cells derived from human embryonic stem cells in transient focal cerebral ischemia in rats. J. Cereb. Blood Flow Metab. 29(4):780–791; 2009. 77. Liu, Z.; Fan, Y.; Won, S. J.; Neumann, M.; Hu, D.; Zhou, L.; Weinstein, P. R.; Liu, J. Chronic treatment with minocycline preserves adult new neurons and reduces

628

functional impairment after focal cerebral ischemia. Stroke 38(1):146–152; 2007. 78. Locatelli, F.; Bersano, A.; Ballabio, E.; Lanfranconi, S.; Papadimitriou, D.; Strazzer, S.; Bresolin, N.; Comi, G. P.; Corti, S. Stem cell therapy in stroke. Cell. Mol. Life Sci. 66(5):757–772; 2009. 79. Loulier, K.; Lathia, J. D.; Marthiens, V.; Relucio, J.; Mughal, M. R.; Tang, S. C.; Coksaygan, T.; Hall, P. E.; Chigurupati, S.; Patton, B.; Colognato, H.; Rao, M. S.; Mattson, M. P.; Haydar, T. F.; Ffrench-Constant, C. b1 integrin maintains integrity of the embryonic neocortical stem cell niche. PLoS Biol. 7(8):e1000176; 2009. 80. Lundberg, C.; Englund, U.; Trono, D.; Bjorklund, A.; Wictorin, K. Differentiation of the RN33B cell line into forebrain projection neurons after transplantation into the neonatal rat brain. Exp. Neurol. 175(2):370–387; 2002. 81. Macas, J.; Nern, C.; Plate, K. H.; Momma, S. Increased generation of neuronal progenitors after ischemic injury in the aged adult human forebrain. J. Neurosci. 26(50):13114–13119; 2006. 82. Mazo, I. B.; Gutierrez-Ramos, J. C.; Frenette, P. S.; Hynes, R. O.; Wagner, D. D.; von Andrian, U. H. Hematopoietic progenitor cell rolling in bone marrow microvessels: Parallel contributions by endothelial selectins and vascular cell adhesion molecule 1. J. Exp. Med. 188(3):465–474; 1998. 83. Meairs, S.; Wahlgren, N.; Dirnagl, U.; Lindvall, O.; Rothwell, P.; Baron, J. C.; Hossmann, K.; Engelhardt, B.; Ferro, J.; McCulloch, J.; Kaste, M.; Endres, M.; Koistinaho, J.; Planas, A.; Vivien, D.; Dijkhuizen, R.; Czlonkowska, A.; Hagen, A.; Evans, A.;. De Libero, G.; Nagy, Z.; Rastenyte, D.; Reess, J.; Davalos, A.; Lenzi, G. L.; Amarenco, P.; Hennerici, M. Stroke research priorities for the next decade--A representative view of the European scientific community. Cerebrovasc. Dis. 22(2– 3):75–82; 2006. 84. Mercier, F.; Kitasako, J. T.; Hatton, G. I. Anatomy of the brain neurogenic zones revisited: Fractones and the fibroblast/macrophage network. J. Comp. Neurol. 451(2):170– 188; 2002. 85. Miller, J. T.; Bartley, J. H.; Wimborne, H. J.; Walker, A. L.; Hess, D. C.; Hill, W. D.; Carroll, J. E. The neuroblast and angioblast chemotaxic factor SDF-1 (CXCL12) expression is briefly up regulated by reactive astrocytes in brain following neonatal hypoxic-ischemic injury. BMC Neurosci. 6:63; 2005. 86. Ming, G. L.; Song, H. Adult neurogenesis in the mammalian central nervous system. Annu. Rev. Neurosci. 28:223–250; 2005. 87. Minger, S. L.; Ekonomou, A.; Carta, E. M.; Chinoy, A.; Perry, R. H.; Ballard, C. G. Endogenous neurogenesis in the human brain following cerebral infarction. Regen. Med. 2(1):69–74; 2007. 88. Mirzadeh, Z.; Merkle, F. T.; Soriano-Navarro, M.; GarciaVerdugo, J. M.; Alvarez-Buylla, A. Neural stem cells confer unique pinwheel architecture to the ventricular surface in neurogenic regions of the adult brain. Cell Stem Cell 3(3):265–278; 2008. 89. Nagasawa, T.; Hirota, S.; Tachibana, K.; Takakura, N.; Nishikawa, S.; Kitamura, Y.; Yoshida, N.; Kikutani, H.; Kishimoto, T. Defects of B-cell lymphopoiesis and bonemarrow myelopoiesis in mice lacking the CXC chemokine PBSF/SDF-1. Nature 382(6592):635–638; 1996.

Ding ET AL.

90. Nelson, P. T.; Kondziolka, D.; Wechsler, L.; Goldstein, S.; Gebel, J.; DeCesare, S.; Elder, E. M.; Zhang, P. J.; Jacobs, A.; McGrogan, M.; Lee, V. M.; Trojanowski, J. Q. Clonal human (hNT) neuron grafts for stroke therapy: Neuropathology in a patient 27 months after implantation. Am. J. Pathol. 160(4):1201-1206; 2002. 91. Nicoleau, C.; Benzakour, O.; Agasse, F.; Thiriet, N.; Petit, J.; Prestoz, L.; Roger, M.; Jaber, M.; Coronas, V. Endogenous hepatocyte growth factor is a niche signal for subventricular zone neural stem cell amplification and self-renewal. Stem Cells 27(2):408–419; 2009. 92. Ohab, J. J.; Fleming, S.; Blesch, A.; Carmichael, S. T. A neurovascular niche for neurogenesis after stroke. J. Neurosci. 26(50):13007–13016; 2006. 93. Onda, T.; Honmou, O.; Harada, K.; Houkin, K.; Hamada, H.; Kocsis, J. D. Therapeutic benefits by human mesenchymal stem cells (hMSCs) and Ang-1 gene-modified hMSCs after cerebral ischemia. J. Cereb. Blood Flow Metab. 28(2):329–340; 2008. 94. Ostenfeld, T.; Caldwell, M. A.; Prowse, K. R.; Linskens, M. H.; Jauniaux, E.; Svendsen, C. N. Human neural precursor cells express low levels of telomerase in vitro and show diminishing cell proliferation with extensive axonal outgrowth following transplantation. Exp. Neurol. 164(1):215–226; 2000. 95. Pagano, S. F.; Impagnatiello, F.; Girelli, M.; Cova, L.; Grioni, E.; Onofri, M.; Cavallaro, M.; Etteri, S.; Vitello, F.; Giombini, S.; Solero, C. L.;Parati, E. A. Isolation and characterization of neural stem cells from the adult human olfactory bulb. Stem Cells 18(4):295–300; 2000. 96. Palmer, T. D.; Markakis, E. A.; Willhoite, A. R.; Safar, F.; Gage, F. H. Fibroblast growth factor-2 activates a latent neurogenic program in neural stem cells from diverse regions of the adult CNS. J. Neurosci. 19(19):8487–8497; 1999. 97. Palmer, T. D.; Ray, J.; Gage, F. H. FGF-2-responsive neuronal progenitors reside in proliferative and quiescent regions of the adult rodent brain. Mol. Cell. Neurosci. 6(5):474–486; 1995. 98. Palmer, T. D.; Takahashi, J.; Gage, F. H. The adult rat hippocampus contains primordial neural stem cells. Mol. Cell. Neurosci. 8(6):389–404; 1997. 99. Parent, J. M.; Vexler, Z. S.; Gong, C.; Derugin, N.; Ferriero, D. M. Rat forebrain neurogenesis and striatal neuron replacement after focal stroke. Ann. Neurol. 52(6):802– 813; 2002. 100. Parent, J. M.; Yu, T. W.; Leibowitz, R. T.; Geschwind, D. H.; Sloviter, R. S.; Lowenstein, D. H. Dentate granule cell neurogenesis is increased by seizures and contributes to aberrant network reorganization in the adult rat hippocampus. J. Neurosci. 17(10):3727–3738; 1997. 101. Park, I. H.; Arora, N.; Huo, H.; Maherali, N.; Ahfeldt, T.; Shimamura, A.; Lensch, M. W.; Cowan, C.; Hochedlinger, K.; Daley, G. Q. Disease-specific induced pluripotent stem cells. Cell 134(5):877–886; 2008. 102. Pastori, C.; Librizzi, L.; Breschi, G. L.; Regondi, C.; Frassoni, C.; Panzica, F.; Frigerio, S.; Gelati, M.; Parati, E.; De Simoni, M. G.; de Curtis, M.. Arterially perfused neurosphere-derived cells distribute outside the ischemic core in a model of transient focal ischemia and reperfusion in vitro. PLoS One 3(7):e2754; 2008. 103. Pfeifer, K.; Vroemen, M.; Blesch, A.; Weidner, N. Adult neural progenitor cells provide a permissive guiding sub-

NEURAL STEM CELLS and STROKE

104.

105.

106.

107. 108. 109.

110.

111.

112.

113.

114.

115.

116.

strate for corticospinal axon growth following spinal cord injury. Eur. J. Neurosci. 20(7):1695–1704; 2004. Pluchino, S.; Zanotti, L.; Rossi, B.; Brambilla, E.; Ottoboni, L.; Salani, G.; Martinello, M.; Cattalini, A.; Bergami, A.; Furlan, R.; Comi, G.; Constantin, G.; Martino, G. Neurosphere-derived multipotent precursors promote neuroprotection by an immunomodulatory mechanism. Nature 436(7048):266–271; 2005. Pollock, K.; Stroemer, P.; Patel, S.; Stevanato, L.; Hope, A.; Miljan, E.; Dong, Z.; Hodges, H.; Price, J.; Sinden, J. D. A conditionally immortal clonal stem cell line from human cortical neuroepithelium for the treatment of ischemic stroke. Exp. Neurol. 199(1):143–155; 2006. Prestoz, L.; Relvas, J. B.; Hopkins, K.; Patel, S.; Sowinski, P.; Price, J.; ffrench-Constant, C. Association between integrin-dependent migration capacity of neural stem cells in vitro and anatomical repair following transplantation. Mol. Cell. Neurosci. 18(5):473–484; 2001. Pumiglia, K.; Temple, S. PEDF: Bridging neurovascular interactions in the stem cell niche. Nat. Neurosci. 9(3):299– 300; 2006. Quinones-Hinojosa, A.; Chaichana, K. The human subventricular zone: A source of new cells and a potential source of brain tumors. Exp. Neurol. 205(2):313–324; 2007. Quinones-Hinojosa, A.; Sanai, N.; Gonzalez-Perez, O.; Garcia-Verdugo, J. M. The human brain subventricular zone: Stem cells in this niche and its organization. Neurosurg. Clin. N. Am. 18(1):15–20; 2007. Quinones-Hinojosa, A.; Sanai, N.; Soriano-Navarro, M.; Gonzalez-Perez, O.; Mirzadeh, Z.; Gil-Perotin, S.; Romero-Rodriguez, R.; Berger, M. S.; Garcia-Verdugo, J. M.; Alvarez-Buylla, A. Cellular composition and cytoarchitecture of the adult human subventricular zone: A niche of neural stem cells. J. Comp. Neurol. 494(3):415– 434; 2006. Rabinovich, S. S.; Seledtsov, V. I.; Banul, N. V.; Poveshchenko, O. V.; Senyukov, V. V.; Astrakov, S. V.; Samarin, D. M.; Taraban, V. Y. Cell therapy of brain stroke. Bull. Exp. Biol. Med. 139(1):126–128; 2005. Ramirez-Castillejo, C.; Sanchez-Sanchez, F.; AndreuAgullo, C.; Ferron, S. R.; Aroca-Aguilar, J. D.; Sanchez, P.; Mira, H.; Escribano, J.; Farinas, I. Pigment epithelium-derived factor is a niche signal for neural stem cell renewal. Nat. Neurosci. 9(3):331–339; 2006. Reynolds, B. A.; Weiss, S. Generation of neurons and astrocytes from isolated cells of the adult mammalian central nervous system. Science 255(5052):1707–1710; 1992. Robin, A. M.; Zhang, Z. G.; Wang, L.; Zhang, R. L.; Katakowski, M.; Zhang, L.; Wang, Y.; Zhang, C.; Chopp, M. Stromal cell-derived factor 1a mediates neural progenitor cell motility after focal cerebral ischemia. J. Cereb. Blood Flow Metab. 26(1):125–134; 2006. Roy, N. S.; Chandler-Militello, D.; Lu, G.; Wang, S.; Goldman, S. A. Retrovirally mediated telomerase immortalization of neural progenitor cells. Nat. Protoc. 2(11):2815–2825; 2007. Sanai, N.; Tramontin, A. D.; Quinones-Hinojosa, A.; Barbaro, N. M.; Gupta, N.; Kunwar, S.; Lawton, M. T.; McDermott, M. W.; Parsa, A. T.; Manuel-Garcia Verdugo, J.; Berger, M. S.; Alvarez-Buylla, A. Unique astrocyte ribbon in adult human brain contains neural stem cells but lacks chain migration. Nature 427(6976):740–744; 2004.

629

117. Savitz, S. I.; Dinsmore, J.; Wu, J.; Henderson, G. V.; Stieg, P.; Caplan, L. R. Neurotransplantation of fetal porcine cells in patients with basal ganglia infarcts: A preliminary safety and feasibility study. Cerebrovasc. Dis. 20(2):101–107; 2005. 118. Seaberg, R. M.; van der Kooy, D. Adult rodent neurogenic regions: The ventricular subependyma contains neural stem cells, but the dentate gyrus contains restricted progenitors. J. Neurosci. 22(5):1784–1793; 2002. 119. Shen, Q.; Wang, Y.; Kokovay, E.; Lin, G.; Chuang, S. M.; Goderie, S. K.; Roysam, B.; Temple, S. Adult SVZ stem cells lie in a vascular niche: A quantitative analysis of niche cell-cell interactions. Cell Stem Cell 3(3):289–300; 2008. 120. Shingo, T.; Sorokan, S. T.; Shimazaki, T.; Weiss, S. Erythropoietin regulates the in vitro and in vivo production of neuronal progenitors by mammalian forebrain neural stem cells. J. Neurosci. 21(24):9733–9743; 2001. 121. Shiras, A.; Chettiar, S. T.; Shepal, V.; Rajendran, G.; Prasad, G. R.; Shastry, P. Spontaneous transformation of human adult nontumorigenic stem cells to cancer stem cells is driven by genomic instability in a human model of glioblastoma. Stem Cells 25(6):1478–1489; 2007. 122. Shyu, W. C.; Lin, S. Z.; Chiang, M. F.; Su, C. Y.; Li, H. Intracerebral peripheral blood stem cell (CD34+) implantation induces neuroplasticity by enhancing b1 integrinmediated angiogenesis in chronic stroke rats. J. Neurosci. 26(13):3444–3453; 2006. 123. Shyu, W. C.; Lin, S. Z.; Yang, H. I.; Tzeng, Y. S.; Pang, C. Y.; Yen, P. S.; Li, H. Functional recovery of stroke rats induced by granulocyte colony-stimulating factor-stimulated stem cells. Circulation 110(13):1847–1854; 2004. 124. Shyu, W. C.; Lin, S. Z.; Yen, P. S.; Su, C. Y.; Chen, D. C.; Wang, H. J.; Li, H. Stromal cell-derived factor-1 a promotes neuroprotection, angiogenesis, and mobilization/ homing of bone marrow-derived cells in stroke rats. J. Pharmacol. Exp. Ther. 324(2):834–849; 2008. 125. Shyu, W. C.; Liu, D. D.; Lin, S. Z.; Li, W. W.; Su, C. Y.; Chang, Y. C.; Wang, H. J.; Wang, H. W.; Tsai, C. H.; Li, H. Implantation of olfactory ensheathing cells promotes neuroplasticity in murine models of stroke. J. Clin. Invest. 118(7):2482–2495; 2008. 126. Siebzehnrubl, F. A.; Jeske, I.; Muller, D.; Buslei, R.; Coras, R.; Hahnen, E.; Huttner, H. B.; Corbeil, D.; Kaesbauer, J.; Appl, T.; von Hörsten, S.; Blümcke, I. Spontaneous in vitro transformation of adult neural precursors into stemlike cancer cells. Brain Pathol. 19(3):399–408; 2009. 127. Sun, C.; Zhang, H.; Li, J.; Huang, H.; Cheng, H.; Wang, Y.; Li, P.; An, Y. Modulation of the major histocompatibility complex by neural stem cell-derived neurotrophic factors used for regenerative therapy in a rat model of stroke. J. Transl. Med. 8:77; 2010. 128. Suzuki, S.; Tanaka, K.; Nogawa, S.; Ito, D.; Dembo, T.; Kosakai, A.; Fukuuchi, Y. Immunohistochemical detection of leukemia inhibitory factor after focal cerebral ischemia in rats. J. Cereb. Blood Flow Metab. 20(4):661–668; 2000. 129. Takahashi, K.; Yasuhara, T.; Shingo, T.; Muraoka, K.; Kameda, M.; Takeuchi, A.; Yano, A.; Kurozumi, K.; Agari, T.; Miyoshi, Y.; Kinugasa, K.; Date, I. Embryonic neural stem cells transplanted in middle cerebral artery occlusion model of rats demonstrated potent therapeutic effects, compared to adult neural stem cells. Brain Res. 1234:172–182; 2008.

630

130. Temple, S. The development of neural stem cells. Nature 414(6859):112–117; 2001. 131. Thored, P.; Arvidsson, A.; Cacci, E.; Ahlenius, H.; Kallur, T.; Darsalia, V.; Ekdahl, C. T.; Kokaia, Z.; Lindvall, O. Persistent production of neurons from adult brain stem cells during recovery after stroke. Stem Cells 24(3):739–747; 2006. 132. Thored, P.; Heldmann, U.; Gomes-Leal, W.; Gisler, R.; Darsalia, V.; Taneera, J.; Nygren, J. M.; Jacobsen, S. E.; Ekdahl, C. T.; Kokaia, Z.; Lindvall, O. Long-term accumulation of microglia with proneurogenic phenotype concomitant with persistent neurogenesis in adult subventricular zone after stroke. Glia 57(8):835–849; 2009. 133. Toma, J. G.; Akhavan, M.; Fernandes, K. J.; BarnabeHeider, F.; Sadikot, A.; Kaplan, D. R.; Miller, F. D. Isolation of multipotent adult stem cells from the dermis of mammalian skin. Nat. Cell Biol. 3(9):778–784; 2001. 134. Vallieres, L.; Campbell, I. L.; Gage, F. H.; Sawchenko, P. E. Reduced hippocampal neurogenesis in adult transgenic mice with chronic astrocytic production of interleukin-6. J. Neurosci. 22(2):486–492; 2002. 135. Vescovi, A. L.; Gritti, A.; Galli, R.; Parati, E. A. Isolation and intracerebral grafting of nontransformed multipotential embryonic human CNS stem cells. J. Neurotrauma 16(8):689–693; 1999. 136. Wachs, F. P.; Couillard-Despres, S.; Engelhardt, M.; Wilhelm, D.; Ploetz, S.; Vroemen, M.; Kaesbauer, J.; Uyanik, G.; Klucken, J.; Karl, C.; Tebbing, J.; Svendsen, C.; Weidner, N.; Kuhn, H. G.; Winkler, J.; Aigner, L. High efficacy of clonal growth and expansion of adult neural stem cells. Lab. Invest. 83(7):949–962; 2003. 137. Yagita, Y.; Kitagawa, K.; Ohtsuki, T.; Takasawa, K.; Miyata, T.; Okano, H.; Hori, M.; Matsumoto, M.

Ding ET AL.

138.

139.

140.

141.

142. 143.

144.

Neurogenesis by progenitor cells in the ischemic adult rat hippocampus. Stroke 32(8):1890–1896; 2001. Yamamoto, S.; Nagao, M.; Sugimori, M.; Kosako, H.; Nakatomi, H.; Yamamoto, N.; Takebayashi, H.; Nabeshima, Y.; Kitamura, T.; Weinmaster, G.; Nakamura, K.; Nakafuku, M. Transcription factor expression and Notch-dependent regulation of neural progenitors in the adult rat spinal cord. J. Neurosci. 21(24):9814–9823; 2001. Zaleska, M. M.; Mercado, M. L.; Chavez, J.; Feuerstein, G. Z.; Pangalos, M. N.; Wood, A. The development of stroke therapeutics: Promising mechanisms and translational challenges. Neuropharmacology 56(2):329–341; 2009. Zhang, P.; Li, J.; Liu, Y.; Chen, X.; Kang, Q. Transplanted human embryonic neural stem cells survive, migrate, differentiate and increase endogenous nestin expression in adult rat cortical peri-infarction zone. Neuropathology 29(4):410–421; 2009. Zhang, P.; Li, J.; Liu, Y.; Chen, X.; Lu, H.; Kang, Q.; Li, W.; Gao, M. Human embryonic neural stem cell transplantation increases subventricular zone cell proliferation and promotes peri-infarct angiogenesis after focal cerebral ischemia. Neuropathology 31(4):384–391; 2011. Zhang, Z. G.; Chopp, M. Neurorestorative therapies for stroke: Underlying mechanisms and translation to the clinic. Lancet Neurol. 8(5):491–500; 2009. Zhang, Z. G.; Jiang, Q.; Zhang, R.; Zhang, L.; Wang, L.; Arniego, P.; Ho, K. L.; Chopp, M. Magnetic resonance imaging and neurosphere therapy of stroke in rat. Ann. Neurol. 53(2):259–263; 2003. Zhao, C.; Deng, W.; Gage, F. H. Mechanisms and functional implications of adult neurogenesis. Cell 132(4):645– 660; 2008.