Next Article in Journal
Cytomegalovirus (CMV) Pneumonitis: Cell Tropism, Inflammation, and Immunity
Next Article in Special Issue
High and Low Molecular Weight Hyaluronic Acid Differentially Influences Oxylipins Synthesis in Course of Neuroinflammation
Previous Article in Journal
Prospective Application of Activity-Based Proteomic Profiling in Vision Research-Potential Unique Insights into Ocular Protease Biology and Pathology
Previous Article in Special Issue
Perspectives for Ezrin and Radixin in Astrocytes: Kinases, Functions and Pathology
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Modeling Neurological Disorders with Human Pluripotent Stem Cell-Derived Astrocytes

1
iPSC-based Drug Discovery and Development Team, RIKEN BioResource Research Center (BRC), Kyoto 619-0237, Japan
2
Center for iPS Cell Research and Application (CiRA), Kyoto University, Kyoto 606-8507, Japan
3
Medical-risk Avoidance based on iPS Cells Team, RIKEN Center for Advanced Intelligence Project (AIP), Kyoto 606-8507, Japan
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2019, 20(16), 3862; https://doi.org/10.3390/ijms20163862
Submission received: 12 June 2019 / Revised: 30 July 2019 / Accepted: 31 July 2019 / Published: 8 August 2019

Abstract

:
Astrocytes play vital roles in neurological disorders. The use of human induced pluripotent stem cell (iPSC)-derived astrocytes provides a chance to explore the contributions of astrocytes in human diseases. Here we review human iPSC-based models for neurological disorders associated with human astrocytes and discuss the points of each model.

1. Introduction

Astrocytes are the most abundant glial cell type in the central nerve system (CNS) and have multiple roles on neuronal development and function. Astrocytes regulate synaptogenesis, modulate synaptic plasticity, provide metabolic support to neurons, secrete or absorb neurotransmitters from synapses, regulate extracellular ion concentrations, support the brain–blood barrier (BBB), and promote myelination in the white matter. Conversely, their dysfunction has been implicated in several neurological disorders such as neurodegenerative diseases, neurodevelopmental diseases, epilepsy and astroglioma [1,2,3].

2. Diversity of Astrocyte Functions and Phenotypes

Astrocytes are extremely specialized and heterogeneous in terms of morphology and function throughout the CNS. According to classical taxonomy, astrocytes are divided into two major classes, protoplasmic astrocytes in the gray matter and fibrous astrocytes in the white matter [4]. Protoplasmic astrocytes have highly branched and complex processes that contact blood vessels and neurons, whereas fibrous astrocytes have elongated, more simple processes that contact blood vessels, oligodendrocytes, and axons at the nodes of Ranvier [5]. Astrocytes are coupled by gap junctions and form large networks [6]. Astrocytes also make contact and form gap-junction coupling with oligodendrocytes in different brain regions, termed panglial networks [7,8]. Recent studies found that astrocyte morphology and function diversify with age and location in the brain [2,9,10,11,12,13,14]. For example, mouse neocortical protoplasmic astrocytes within layer II–IV exhibit distinct morphologies, structural interactions with synapses, and molecular expressions [15]. How astrocytes acquire these different phenotypes during brain development and maturation is largely unknown.
Astrocytes acquire different reactive phenotypes when they respond to different pathological stimuli, such as infection, ischemia, neurodegenerative disorders, and aging. Reactive astrocytes have modulated gene expression profiles that regulate both structural and biochemical changes [9,11,16]. Two major different types of reactive astrocytes, A1 and A2, were shown to be induced by injury and disease in the adult CNS [17]. A1 astrocytes, which are induced by injury, neuroinflammation, and neurodegenerative disease, produce proinflammatory molecules and secrete molecules that are toxic to neural cells. Several cytokines including IL-1α, tumor necrosis factor (TNF) and C1q released by activated microglia can promote the formation of reactive A1 astrocytes such that the A1 astrocytes no longer promote neuronal survival, outgrowth, or synapse formation, nor phagocytize synapses or debris [6]. On the other hand, A2 astrocytes, which are induced by ischemia, secrete neuroprotective cytokines including thrombospondins (TSPs) to provide neurotrophic support and modulate inflammatory responses that promote neuronal survival and tissue repair [17]. Non-reactive astrocytes become reactive astrocytes with a spectrum of potential molecular, cellular, and functional changes, including neurotoxic or neuroprotective properties, in response to a wide range of extracellular signals and stress. Reactive changes in astrocyte properties occur in a context-dependent manner, and under each circumstance that would induce a continuum heterogeneous population of reactive astrocytes [18,19]. It is hypothesized that reactive astrocytes play a role not only in aging and age-related diseases including neurodegenerative disease, but also in neurodevelopmental diseases [9,16]. A1 astrocytes have been identified in the post-mortem brain tissues of individuals with neurological diseases, including Alzheimer’s disease (AD), Parkinson’s disease (PD), multiple sclerosis (MS), amyotrophic lateral sclerosis (ALS), and Huntington’s disease (HD). Astrocytes display different phenotypes depending on the pathogenesis [3]. Moreover, astrocyte reactivity is induced by normal aging [9]. It was shown that astrocytes in aged mouse brains express genes associated with reactivity, particularly in the hippocampus and striatum, which are areas linked to neurodegenerative diseases [13,16]. Thus, astrocyte phenotypic changes have been implicated in the pathogenesis of a wide range of neurological disorders.

3. Human PSC-Derived Astrocytes as In Vitro Models

Human and rodent astrocytes show significant differences [5,11,20], with human cortical astrocytes being larger, more complex and diverse [20]. For example, human cortical protoplasmic astrocytes are nearly 2.6-fold larger in diameter and extend 10-fold more glial fibrillary acidic protein (GFAP)-expressing primary processes than their rodent counterparts [20]. Furthermore, human cortical protoplasmic astrocytes propagate Ca2+ waves approximately 4-fold faster than rodent cortical protoplasmic astrocytes [20]. The two types also show distinct gene expression profiles [11]. These differences must be considered when studying animal models. However, there is limited access to astrocytes from patient brain samples. Human astrocytoma cell lines [21] and primary human adult and fetal astrocytes [11] are available as in vitro cellular models, but fail to fully recapitulate human diseases.
Human pluripotent stem cells (PSCs), which include human embryonic stem cells (ESCs) [22] and human induced pluripotent stem cells (iPSCs) [23], can be used to model neurological diseases. Several protocols exist for the directed differentiation of human PSCs toward astrocytes. Furthermore, human ESCs and iPSCs with gene modification mimicking the mutation of the inherited disease can be generated, as too can iPSCs from the patient without gene modification. Accordingly, several studies have used human PSCs to model astrocyte function or dysfunction in neurological diseases (Table 1).

4. Methods for Generating Human PSC-Derived Astrocytes

Since the derivation of neural precursors from human ESCs was reported in 2001 [41,42], various differentiation protocols for neural cells including astrocytes from human PSCs have been described [24,29,43,44,45,46,47,48,49,50]. Most of them were developed by recapturing embryonic developmental events, although astrocyte development is not fully understood. Protocols for generating astrocytes from human PSCs generally consist of four main steps: (1) Conversion of the undifferentiated human PSCs into rosette-forming neuroepithelial cells; (2) Induction of the neural stem cells with or without morphogens for regional patterning; (3) Glia lineage specification by long-term expansion of the induced neural stem cells either in adhesion or suspension culture with growth factors; and (4) Astrocyte terminal differentiation and maturation [43,51].
To obtain neural stem cells, the serum-free culture of embryoid body (SFEB) method [47] and the SFEB with quick aggregation (SFEBq) method [48] have been used as well as the dual inhibition of SMAD signaling in adherent human PSC cultures [52]. Alternatively, long-term self-renewing human iPSC-derived neuroepithelial stem cells (ltNES) [53] can be used for astrocyte differentiation as well as neuronal differentiation [49].
The combination of morphogens and growth factors, such as FGF8, RA, or SHH [54,55,56], were used for rostral-caudal and dorsal-ventral patterning during the induction and regional specification of neural stem cells [57]. The induced cells expressed distinct homeodomain transcription factors and displayed phenotypic differences [57]. Using these techniques, researchers separately generated region-specific astrocyte subtypes including spinal cord astrocytes, midbrain astrocytes, and neocortical astrocytes [43,57,58,59].
Astrocyte lineage specification and terminal differentiation are conducted with growth factor-rich undefined medium such as fetal bovine serum-containing medium, growth factor-defined medium or growth factor-reduced medium. The efficiency of the terminal differentiation is independent of the growth factors used. On the other hand, serum accelerates terminal differentiation. Although astrocytes can be expanded easily in vitro, they tend to undergo senescence with the exhaustion of growth factors [60,61]. Serum components also may change the astrocyte phenotype. Indeed, acutely purified primary astrocytes exhibit extensive process-bearing morphologies under serum-free media, but exhibit polygonal fibroblast-like morphologies under serum-containing media [11]. Furthermore, a serum-containing medium induces a phenotypic change and reactivity to human iPSC-derived astrocytes [35]. Thus, in vitro culture conditions should be controlled very carefully. Two groups [43,57] have described methods to generate functional astrocytes from human PSCs under serum-free conditions. The maturation of astrocytes is also important to recapitulate the physiological function or disease phenotype. Long-term cultivation as a three-dimensional (3D) structure can accelerate the maturation of iPSC-derived astrocytes. Sloan et al. conducted up to 20 months cultivation of iPSC-derived astrocytes in 3D cortical neurospheres. Astrocytes after long-time culture closely resembled primary human fetal astrocytes and showed similar global gene-expression patterns and functions to that of primary human astrocytes [62]. It is likely that the astrocyte function requires integration of the cells into a system. Yet, it is unknown if astrocytes in the entire brain form just a single (syncytium) or multiple networks (syncytia) through gap junctional coupling [63]. iPSC-derived astrocytes partially recapitulate functional networks by simply assessing calcium influx spread [43,49]. However, these systems lack a robust astrocytic network which can modulate neuronal activity, signal transmission, or vessel contraction. Additionally, it is the future target to define the “healthy or normal” character in Ca2+ oscillation or spread for disease modeling, because of the wide-variety of Ca2+ spreading mechanisms [64]. Therefore, it is the incoming step to construct a culture system, consisting of multiple cell types to mimic human brain and model disorders in the future.
Alternatively, the direct conversion of skin fibroblasts to astrocytes was reported [44]. The defined transcription factors nuclear factor I A (NFIA), nuclear factor I B (NFIB), and SRY-box transcription factor 9 (SOX9) induce an astrocytic phenotype in human fibroblasts [44]. In addition, a recent study demonstrated that human PSCs acquire glial fate by the transient forced activation of NFIA and SOX9 tandemly or NFIA alone, and further differentiated into functional astrocytes [46,65]. Understanding the molecular mechanisms in astrocyte specification will lead to improved molecular activation to enable PSC-derived astrocytes with better operability and better yield in the future.

5. Modeling Neurological Disorders Using Human PSC-Derived Astrocytes

Below we highlighted the modeling of several neurological disorders with human PSC-derived astrocytes.

5.1. Alzheimer’s Disease

Alzheimer’s disease (AD) is the most common dementia worldwide. In AD, astrocyte pathology is expressed even before neuronal death. Astrocytes in rodent AD models and AD patients become reactive, have aberrant calcium signaling, and change their gene expression profile and metabolism [3]. AD astrocytes display elevated levels of GFAP expression and GABA production and release [66]. Moreover, amyloid-β(Aβ) accumulates in AD astrocytes. Most cases of AD are late-onset sporadic AD, 1–2% are early-onset familial AD with underlying gene mutations in amyloid precursor protein (APP) or presenilin-1 and -2 (PSEN1/2) [67].
We generated iPSCs from one sporadic AD patient and two familial AD patients with APP E693Δ mutation and APP V717L mutation. The AD iPSCs were differentiated into neurons and astrocytes, in which Aβ oligomers were found to have accumulated [24]. We further used the AD iPSC model for drug screening and discovered that docosahexaenoic acid (DHA) may be effective for a subset of AD patients to reduce or prevent the symptoms [24].
Oksanen et al. generated iPSCs from three AD patients with PSEN1 exon 9 deletion and analyzed the AD iPSC-derived astrocytes [25]. The AD astrocytes exhibited increased Aβ production, altered cytokine release, and dysregulated Ca2+ homeostasis [25]. Furthermore, due to the altered metabolism, the AD astrocytes showed increased oxidative stress and reduced lactate secretion, as well as compromised neuronal supportive function, as evidenced by altering Ca2+ transients in the healthy neurons.

5.2. Parkinson’s Disease

Parkinson’s disease (PD) is the second most common neurodegenerative disease and is characterized by a significant loss of ventral midbrain dopaminergic neurons in the substantia nigra pars compacta. A pathological hallmark of PD is the accumulation of Lewy bodies (LBs), which mainly consist of α-synuclein, and α-synuclein aggregates spread from neuron to neuron in PD lesions [68]. Most PD cases are sporadic, but monogenic mutations in 17 genes have been identified and implicated in the PD pathogenesis [69]. Mutations in the leucine-rich repeat kinase 2 (LRRK2) gene are the most common cause of inherited PD, particularly the G2019S mutation in LRRK2 [69].
Di Domenico et al. generated iPSC-derived astrocytes and dopaminergic neurons from familial PD patients carrying the G2019S mutation and compared the cells with iPSC-derived astrocytes and dopaminergic neurons from healthy controls [26]. They demonstrated that α-synuclein, secreted from the PD iPSC-derived astrocytes, exerted neurotoxic function on surrounding dopaminergic neurons, leading to a non-cell-autonomous neuronal dysfunction. In addition, they identified dysfunctional chaperone-mediated autophagy (CMA), impaired macroautophagy, and progressive α-synuclein accumulation in the PD iPSC-derived astrocytes. Furthermore, the reactivation of CMA protected the PD iPSC-derived astrocytes and dopaminergic neurons via the clearance of α-synuclein accumulation [26]. These findings described a non-cell-autonomous contribution of astrocytes during PD pathogenesis.

5.3. Huntington’s Disease

Huntington’s disease (HD) is a progressive and fatal neurodegenerative disorder caused by the trinucleotide repeat expansion (CAG) in exon 1 of the huntingtin (HTT) gene. The CAG repeat expansion causes an expanded polyglutamine tract at the amino terminus of the HTT protein. The mutant HTT protein accumulates and aggregates not only in neurons, including medium spiny neurons of the striatum, but also in other neural cell types such as astrocytes in the brain of HD patients [70].
Juopperi et al. generated iPSCs from a father with adult-onset HD and 50 CAG repeats and his daughter with juvenile HD and 109 CAG repeats [27]. The astrocytes derived from the HD-specific iPSCs exhibited a large number of cytoplasmic, electron clear vacuoles a phenomenon seen in HD patients. The severity of this phenotype was dependent on the number of CAG repeats.

5.4. Tauopathy

Mutations in the gene encoding the microtubule-associated protein TAU (MAPT) are a common cause of frontotemporal dementia (FTD), a group of neurodegenerative diseases characterized by progressive nerve cell loss in the frontal lobes and the temporal lobes in the brain.
Hallmann et al. generated iPSCs from one FTD patient carrying the N279K MAPT mutation to establish a cell model of the astrocyte pathology in FTD. In their study, FTD iPSC-derived neurons and astrocytes were analyzed. The FTD astrocytes showed disease-associated changes in the TAU expression, increased vulnerability to oxidative stress and increased protein ubiquitination. Co-culture of healthy control neurons with FTD iPSC-derived astrocytes demonstrated that the astrocytes induced alterations in stress-response and gene-expression profiles in the neurons [28].

5.5. TDP-43 Proteinopathy

Transactive response DNA-binding protein (TDP-43) forms ubiquitinated inclusions in the cytoplasm and nuclei of neurons and astrocytes in patients with amyotrophic lateral sclerosis (ALS) or frontotemporal lobar dementia (FTLD). Consistently, mutations in the TDP-43 gene are a common cause of familial ALS and FTLD.
Serio et al. generated functional astrocytes from human iPSCs carrying a TDP-43 mutation and showed that the mutant TDP43 astrocytes exhibited increased levels of TDP-43, impaired subcellular localization of TDP-43, and decreased cell survival [29]. However, the mutant TDP-43 astrocytes did not adversely affect the survival of cocultured neurons, suggesting a glial cell-autonomous pathological phenotype is associated with the TDP-43 proteinopathy.

5.6. Amyotrophic Lateral Sclerosis

ALS is a group of rare neurological diseases characterized by the progressive degeneration of motor neurons, resulting in progressive muscle weakness and respiratory failure. More than 90% of ALS cases are considered sporadic, with the remaining considered familial. Among familial ALS, 20% are linked to various point mutations in the Cu/Zn superoxide dismutase 1 (SOD1) gene.
Wada et al. generated human ESCs overexpressing SOD1 mutant (G93A) by gene editing and then differentiated them to spinal motor neurons and astrocytes [30]. The astrocytes were found to secrete factors toxic to the spinal motor neurons [30]. Tyzack et al. showed that ephrin type-B receptor 1 (EphB1) is upregulated in injured motor neurons and that EphB1 induces astrocytic STAT3 signaling followed by a protective and anti-inflammatory signature in astrocytes [31]. They generated astrocytes from ALS patient-iPSCs carrying SOD1 mutation (D90A) and demonstrated that EphB1 and the downstream pathway is disrupted [31]. These ALS studies identified the increase of astrocytic toxicity and dysfunction of the neuroprotective astrocytic response.

5.7. Spinal Muscular Atrophy

Spinal muscular atrophy (SMA) is a group of childhood neurodegenerative disorders caused by loss or mutations in the survival motor neuron 1 (SMN1). SMA is characterized by a loss of motor neurons or anterior horn cells, resulting in progressive muscle weakness and muscle wasting in proximal muscles, respiratory difficulty, paralysis, and death.
McGivern et al. established iPSC-derived astrocytes from patients with SMA, and reported that the astrocytes had larger cell bodies and shorter processes, reduced GDNF expression, increased basal calcium levels and decreased response to ATP stimulation [32]. The SMA iPSC-derived astrocyte models suggested that reactive astrocytes are involved in motor neuron loss in SMA.

5.8. Alexander Disease

Alexander disease (AxD) is a rare inherited neurodegenerative disorder that primarily affects astrocytes. A variety of GFAP gene mutations causes astrocyte dysfunction in AxD [71,72]. The common neuropathological features of AxD are the degeneration of white matter (leukodystrophy) and the formation of cytoplasmic inclusions, termed Rosenthal fibers, within the astrocytes [71,72,73].
We generated iPSCs from three AxD patients and differentiated them into astrocytes, in which Rosenthal fiber-like structures and GFAP-positive aggregates were observed [33]. The AxD astrocytes produced and released more inflammatory cytokines than healthy control iPSC-derived astrocytes, suggesting that the AxD astrocyte models recapitulate the leukodystrophy and a variety of other AxD features [33]. Li et al. also established AxD astrocyte models from three AxD patient-iPSCs [34]. These iPS-derived astrocytes exhibited Rosenthal fiber-like structures and GFAP aggregation. Using a co-culture system of AxD iPS-derived astrocytes and oligodendrocytes, Li et al. found that the astrocytes secreted more molecules that inhibited oligodendrocyte progenitor cell function and impair myelination [34].

5.9. Multiple Sclerosis

Multiple sclerosis (MS) is an auto-inflammatory disease of the CNS. It is characterized by demyelination followed by axonal loss and neurodegeneration [74]. In MS, reactive astrocytes are present around the lesions [74,75], however, the pathogenesis and cause of MS is largely unknown.
Perriot et al. generated iPSCs from four MS patients [35]. They found IL-1b, TNF-a, and IL-6, all cytokines important for the neuroinflammation seen in MS, in a serum-free culture system triggered a specific reactivity in the iPSC-derived astrocytes.

5.10. Autism Spectrum Disorder

Autism spectrum disorder (ASD) is a group of neurodevelopmental disorders. A small percentage of ASD is syndromic ASD with copy number variants (CNVs) or mutations in ASD-risk genes, but most are non-syndromic ASD with unknown genetic etiology [76].
Russo et al. generated iPSCs from a clinically well-characterized cohort of three individuals with non-syndromic ASD sharing common behaviors [36]. Analyzing the synaptogenesis and neuronal activity of a mixed neuron culture derived from ASD iPSCs, they found that the neurons had a significant decrease in synaptic gene expression and protein levels, glutamate neurotransmitter release, and, consequently, reduced spontaneous firing rate. Moreover, co-culture experiments of the ASD iPSC-derived astrocytes and neurons revealed that cytokines including interleukin-6 (IL-6) secreted from the astrocytes, interfered with proper neuronal development.

5.11. Rett Syndrome

Rett syndrome (RTT) is a rare neurodevelopmental disorder caused by a variety of mutations in the Methyl-CpG-binding protein2 (MECP2) gene on the X chromosome. RTT patients show abnormalities in cognitive, sensory, emotional, motor and autonomic functions, but the disease mechanism is not well understood. Mouse models have shown that MeCP2 deficiency in astrocytes causes abnormal BDNF regulation, cytokine production, and neuronal dendritic induction [1,77].
Williams et al. generated iPSC lines from RTT patients carrying the V247X (valine 247 to stop codon, nonsense), R294X (arginine 294 to stop codon, nonsense) or R306C (arginine 306 to cysteine, missense) mutations [78] and differentiated them into astrocytes [37].
Neurons co-cultured with astrocytes derived from RTT iPSCs showed shorter neurite length and fewer terminal ends than if co-cultured with control astrocytes [37]. These phenotypes are consistent with RTT pathologies. Moreover, the study demonstrated that both insulin-like growth factor 1 (IGF1) and GPE (a peptide containing the first three amino acids of IGF-1) partially rescued the neuronal deficits caused by the RTT astrocytes. Andoh-Noda et al. generated two iPSC lines from RTT twins with a de novo frame-shift mutation in MECP2 (G269AfsX288) [38,79]. They demonstrated that the RTT iPSC lines did not express detectable MeCP2 protein during any stage of differentiation into neural cells and that MeCP2 deficiency triggers a change in the astrocytic gene expression, yielding accelerated astrocyte formation from RTT iPSC-derived neural stem cells [38]. Finally, Delépine et al. identified altered microtubule dynamics and impaired vesicular transport in RTT iPSC-derived astrocytes carrying the MECP2 gene mutation (R294X) [39]. These studies have shed light on astrocytic abnormalities in RTT and suggested that astrocytes might be a target of RTT therapy.

5.12. Epilepsy

Epilepsy is one of the common neurological disorder in which hypersynchronous neuronal firing occurs as a consequence of intensive burst activity from groups of neurons. Epilepsy can be genetic or acquired, although the cause is unknown in most cases. Childhood epileptic encephalopathies (CEEs) such as Dravet syndrome, Angelman syndrome, and Rett syndrome (RTT) are increasingly linked to specific genetic mutations. iPSC-derived neural cells, including GABAergic or glutamatergic neurons, have come to be a common tool to model epileptic disorders [80]. Astrocyte regulates neurotransmission by encasing thousands of synapses to form tripartite synapses between pre- and postsynaptic structures of two neurons and the surrounding astrocytic process [81]. Astrocytes also contact blood vessels and other glial cells, affecting neurotransmission. Therefore, astrocytes have a key role in the seizure activity, in addition to the well-established neurogenic mechanisms [3,82].
A total of 70–80% of patients with RTT develop seizures. As mentioned in earlier sections, astrocytes derived from iPSC of RTT patients do not have enough ability to support neurite outgrowth or synapse formation [37]. In the case of drug-induced epilepsy, single-cell transcriptome analysis of cocultured iPSC-derived neurons and astrocytes clarified that the iPSC-derived astrocyte upregulate the activities of AMPA and NMDA receptors, which is directly related to epileptiform discharge by gabazine or kaliotoxin [40].

6. Closing Remarks

Astrocytes play a key role in many neurological disorders, but details are still lacking. Astrocytes derived from human PSCs with disease-specific gene mutations can be a powerful tool to understanding the cellular phenotypes and disease mechanisms in vitro. These models are expected to clarify how astrocytes contribute to neurological disorders including the underlying mechanisms.

Funding

This research was funded in part by a grant for Core Center for iPS Cell Research of Research Center Network for Realization of Regenerative Medicine from AMED to H.I. and from KAKENHI (18K18452, 18H02717) to T.K, and H.I., KAKENHI (17K16121) to T.K, and KAKENHI (19K07986) to M.S.

Acknowledgments

We would like to express our sincere gratitude to all our coworkers and collaborators: Ran Shibukawa, Yukako Sagara, Yasue Okanishi, Kazuma Kamata, Aya Washida, Kayoko Tsukita, Takako Enami, Ayako Nagahashi for technical assistance; Peter Karagiannis for critical reading and editing of the manuscript; and Makiko Yasui, Mikie Iijima, Nozomi Kawabata, Ayumi Suzuki, Kiyomi Hatae, and Miwa Fujita for their valuable administrative support.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Molofsky, A.V.; Krenick, R.; Ullian, E.; Tsai, H.-H.; Deneen, B.; Richardson, W.D.; Barres, B.A.; Rowitch, D.H. Astrocytes and disease: A neurodevelopmental perspective. Genes Dev. 2012, 26, 891–907. [Google Scholar] [CrossRef] [PubMed]
  2. Lin, C.-C.J.; Yu, K.; Hatcher, A.; Huang, T.-W.; Lee, H.K.; Carlson, J.; Weston, M.C.; Chen, F.; Zhang, Y.; Zhu, W.; et al. Identification of diverse astrocyte populations and their malignant analogs. Nat. Neurosci. 2017, 20, 396–405. [Google Scholar] [Green Version]
  3. Dossi, E.; Vasile, F.; Rouach, N. Human astrocytes in the diseased brain. Brain Res. Bull. 2018, 136, 139–156. [Google Scholar] [CrossRef] [PubMed]
  4. Miller, R.; Raff, M. Fibrous and protoplasmic astrocytes are biochemically and developmentally distinct. J. Neurosci. 1984, 4, 585–592. [Google Scholar] [CrossRef] [PubMed]
  5. Lundgaard, I.; Osorio, M.J.; Kress, B.T.; Sanggaard, S.; Nedergaard, M. White matter astrocytes in health and disease. Neuroscience 2014, 276, 161–173. [Google Scholar] [CrossRef] [PubMed]
  6. Gutnick, M.J.; Connors, B.; Ransom, B.R. Dye-coupling between glial cells in the guinea pig neocortical slice. Brain Res. 1981, 213, 486–492. [Google Scholar] [CrossRef]
  7. Griemsmann, S.; Hoft, S.P.; Bedner, P.; Zhang, J.; von Staden, E.; Beinhauer, A.; Degen, J.; Dublin, P.; Cope, D.W.; Richter, N.; et al. Characterization of panglial gap junction networks in the thalamus, neocortex, and hippocampus reveals a unique population of glial cells. Cereb. Cortex 2015, 25, 3420–3433. [Google Scholar] [CrossRef] [PubMed]
  8. Beiersdorfer, A.; Scheller, A.; Kirchhoff, F.; Lohr, C. Panglial gap junctions between astrocytes and olfactory ensheathing cells mediate transmission of ca2+ transients and neurovascular coupling. Glia 2019, 67, 1385–1400. [Google Scholar] [CrossRef] [PubMed]
  9. Liddelow, S.A.; Guttenplan, K.A.; Clarke, L.E.; Bennett, F.C.; Bohlen, C.J.; Schirmer, L.; Bennett, M.L.; Münch, A.E.; Chung, W.-S.; Peterson, T.C.; et al. Neurotoxic reactive astrocytes are induced by activated microglia. Nature 2017, 541, 481–487. [Google Scholar] [CrossRef]
  10. Zhang, Y.; Chen, K.; Sloan, S.A.; Bennett, M.L.; Scholze, A.R.; O’Keeffe, S.; Phatnani, H.P.; Guarnieri, P.; Caneda, C.; Ruderisch, N.; et al. An RNA-Sequencing Transcriptome and Splicing Database of Glia, Neurons, and Vascular Cells of the Cerebral Cortex. J. Neurosci. 2014, 34, 11929–11947. [Google Scholar] [CrossRef]
  11. Zhang, Y.; Sloan, S.A.; Clarke, L.E.; Caneda, C.; Plaza, C.A.; Blumenthal, P.D.; Vogel, H.; Steinberg, G.K.; Edwards, M.S.; Li, G.; et al. Purification and characterization of progenitor and mature human astrocytes reveals transcriptional and functional differences with mouse. Neuron 2016, 89, 37–53. [Google Scholar] [CrossRef] [PubMed]
  12. Rao, V.T.; Fuh, S.-C.; Sawaya, R.; Moore, C.S.; Ho, M.-K.; Bedell, B.J.; Antel, J.P.; Ludwin, S.K.; Sarnat, H.B.; Bar-Or, A. MicroRNA Expression Patterns in Human Astrocytes in Relation to Anatomical Location and Age. J. Neuropathol. Exp. Neurol. 2016, 75, 156–166. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Boisvert, M.M.; Erikson, G.A.; Shokhirev, M.N.; Allen, N.J. The Aging Astrocyte Transcriptome from Multiple Regions of the Mouse Brain. Cell Rep. 2018, 22, 269–285. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Emsley, J.G.; Macklis, J.D. Astroglial heterogeneity closely reflects the neuronal-defined anatomy of the adult murine CNS. Neuron Glia Boil. 2006, 2, 175–186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Lanjakornsiripan, D.; Pior, B.-J.; Kawaguchi, D.; Furutachi, S.; Tahara, T.; Katsuyama, Y.; Suzuki, Y.; Fukazawa, Y.; Gotoh, Y. Layer-specific morphological and molecular differences in neocortical astrocytes and their dependence on neuronal layers. Nat. Commun. 2018, 9, 1623. [Google Scholar] [CrossRef] [PubMed]
  16. Clarke, L.E.; Liddelow, S.A.; Chakraborty, C.; Münch, A.E.; Heiman, M.; Barres, B.A. Normal aging induces a1-like astrocyte reactivity. Proc. Natl. Acad. Sci. USA 2018, 115, E1896–E1905. [Google Scholar] [CrossRef] [PubMed]
  17. Zamanian, J.L.; Xu, L.; Foo, L.C.; Nouri, N.; Zhou, L.; Giffard, R.G.; Barres, B.A. Genomic Analysis of Reactive Astrogliosis. J. Neurosci. 2012, 32, 6391–6410. [Google Scholar] [CrossRef] [Green Version]
  18. Sofroniew, M.V. Astrogliosis. Cold Spring Harb. Perspect. Biol. 2014, 7, a020420. [Google Scholar] [CrossRef]
  19. Liddelow, S.A.; Barres, B.A. Reactive Astrocytes: Production, Function, and Therapeutic Potential. Immunity 2017, 46, 957–967. [Google Scholar] [CrossRef] [Green Version]
  20. Oberheim, N.A.; Takano, T.; Han, X.; He, W.; Lin, J.H.C.; Wang, F.; Xu, Q.; Wyatt, J.D.; Pilcher, W.; Ojemann, J.G.; et al. Uniquely hominid features of adult human astrocytes. J. Neurosci. 2009, 29, 3276–3287. [Google Scholar] [CrossRef]
  21. Malik, N.; Wang, X.; Shah, S.; Efthymiou, A.G.; Yan, B.; Heman-Ackah, S.; Zhan, M.; Rao, M. Comparison of the Gene Expression Profiles of Human Fetal Cortical Astrocytes with Pluripotent Stem Cell Derived Neural Stem Cells Identifies Human Astrocyte Markers and Signaling Pathways and Transcription Factors Active in Human Astrocytes. PLoS ONE 2014, 9, e96139. [Google Scholar] [CrossRef] [PubMed]
  22. Thomson, J.A. Embryonic Stem Cell Lines Derived from Human Blastocysts. Science 1998, 282, 1145–1147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Takahashi, K.; Tanabe, K.; Ohnuki, M.; Narita, M.; Ichisaka, T.; Tomoda, K.; Yamanaka, S. Induction of Pluripotent Stem Cells from Adult Human Fibroblasts by Defined Factors. Cell 2007, 131, 861–872. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Kondo, T.; Asai, M.; Tsukita, K.; Kutoku, Y.; Ohsawa, Y.; Sunada, Y.; Imamura, K.; Egawa, N.; Yahata, N.; Okita, K.; et al. Modeling alzheimer’s disease with ipscs reveals stress phenotypes associated with intracellular abeta and differential drug responsiveness. Cell Stem Cell 2013, 12, 487–496. [Google Scholar] [CrossRef] [PubMed]
  25. Oksanen, M.; Petersen, A.J.; Naumenko, N.; Puttonen, K.; Lehtonen, S.; Gubert Olive, M.; Shakirzyanova, A.; Leskela, S.; Sarajarvi, T.; Viitanen, M.; et al. Psen1 mutant ipsc-derived model reveals severe astrocyte pathology in alzheimer’s disease. Stem Cell Rep. 2017, 9, 1885–1897. [Google Scholar] [CrossRef] [PubMed]
  26. Di Domenico, A.; Carola, G.; Calatayud, C.; Pons-Espinal, M.; Munoz, J.P.; Richaud-Patin, Y.; Fernandez-Carasa, I.; Gut, M.; Faella, A.; Parameswaran, J.; et al. Patient-specific ipsc-derived astrocytes contribute to non-cell-autonomous neurodegeneration in parkinson’s disease. Stem Cell Rep. 2019, 12, 213–229. [Google Scholar] [CrossRef] [PubMed]
  27. Juopperi, T.A.; Kim, W.R.; Chiang, C.-H.; Yu, H.; Margolis, R.L.; Ross, C.A.; Ming, G.-L.; Song, H. Astrocytes generated from patient induced pluripotent stem cells recapitulate features of Huntington’s disease patient cells. Mol. Brain 2012, 5, 17. [Google Scholar] [CrossRef]
  28. Hallmann, A.-L.; Araúzo-Bravo, M.J.; Mavrommatis, L.; Ehrlich, M.; Röpke, A.; Brockhaus, J.; Missler, M.; Sterneckert, J.; Schöler, H.R.; Kuhlmann, T.; et al. Astrocyte pathology in a human neural stem cell model of frontotemporal dementia caused by mutant TAU protein. Sci. Rep. 2017, 7, 42991. [Google Scholar] [CrossRef]
  29. Serio, A.; Bilican, B.; Barmada, S.J.; Ando, D.M.; Zhao, C.; Siller, R.; Burr, K.; Haghi, G.; Story, D.; Nishimura, A.L.; et al. Astrocyte pathology and the absence of non-cell autonomy in an induced pluripotent stem cell model of TDP-43 proteinopathy. Proc. Natl. Acad. Sci. USA 2013, 110, 4697–4702. [Google Scholar] [CrossRef] [Green Version]
  30. Wada, T.; Goparaju, S.K.; Tooi, N.; Inoue, H.; Takahashi, R.; Nakatsuji, N.; Aiba, K. Amyotrophic Lateral Sclerosis Model Derived from Human Embryonic Stem Cells Overexpressing Mutant Superoxide Dismutase 1. Stem Cells Transl. Med. 2012, 1, 396–402. [Google Scholar] [CrossRef]
  31. Tyzack, G.E.; Hall, C.E.; Sibley, C.R.; Cymes, T.; Forostyak, S.; Carlino, G.; Meyer, I.F.; Schiavo, G.; Zhang, S.-C.; Gibbons, G.M.; et al. A neuroprotective astrocyte state is induced by neuronal signal EphB1 but fails in ALS models. Nat. Commun. 2017, 8, 1164. [Google Scholar] [CrossRef] [PubMed]
  32. McGivern, J.V.; Patitucci, T.N.; Nord, J.A.; Barabas, M.-E.A.; Stucky, C.L.; Ebert, A.D. Spinal muscular atrophy astrocytes exhibit abnormal calcium regulation and reduced growth factor production. Glia 2013, 61, 1418–1428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Kondo, T.; Funayama, M.; Miyake, M.; Tsukita, K.; Era, T.; Osaka, H.; Ayaki, T.; Takahashi, R.; Inoue, H. Modeling Alexander disease with patient iPSCs reveals cellular and molecular pathology of astrocytes. Acta Neuropathol. Commun. 2016, 4, 607. [Google Scholar] [CrossRef] [PubMed]
  34. Li, L.; Tian, E.; Chen, X.; Chao, J.; Klein, J.; Qu, Q.; Sun, G.; Sun, G.; Huang, Y.; Warden, C.D.; et al. GFAP Mutations in Astrocytes Impair Oligodendrocyte Progenitor Proliferation and Myelination in an hiPSC Model of Alexander Disease. Cell Stem Cell 2018, 23, 239–251. [Google Scholar] [CrossRef] [PubMed]
  35. Perriot, S.; Mathias, A.; Perriard, G.; Canales, M.; Jonkmans, N.; Merienne, N.; Meunier, C.; El Kassar, L.; Perrier, A.L.; Laplaud, D.-A.; et al. Human Induced Pluripotent Stem Cell-Derived Astrocytes Are Differentially Activated by Multiple Sclerosis-Associated Cytokines. Stem Cell Rep. 2018, 11, 1199–1210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Russo, F.B.; Freitas, B.C.; Pignatari, G.C.; Fernandes, I.R.; Sebat, J.; Muotri, A.R.; Beltrão-Braga, P.C.B. Modeling the Interplay Between Neurons and Astrocytes in Autism Using Human Induced Pluripotent Stem Cells. Boil. Psychiatry 2018, 83, 569–578. [Google Scholar] [CrossRef]
  37. Williams, E.C.; Zhong, X.; Mohamed, A.; Li, R.; Liu, Y.; Dong, Q.; Ananiev, G.E.; Mok, J.C.C.; Lin, B.R.; Lu, J.; et al. Mutant astrocytes differentiated from Rett syndrome patients-specific iPSCs have adverse effects on wild-type neurons. Hum. Mol. Genet. 2014, 23, 2968–2980. [Google Scholar] [CrossRef] [PubMed]
  38. Andoh-Noda, T.; Akamatsu, W.; Miyake, K.; Matsumoto, T.; Yamaguchi, R.; Sanosaka, T.; Okada, Y.; Kobayashi, T.; Ohyama, M.; Nakashima, K.; et al. Differentiation of multipotent neural stem cells derived from Rett syndrome patients is biased toward the astrocytic lineage. Mol. Brain 2015, 8, 136. [Google Scholar] [CrossRef] [PubMed]
  39. Delepine, C.; Meziane, H.; Nectoux, J.; Opitz, M.; Smith, A.B.; Ballatore, C.; Saillour, Y.; Bennaceur-Griscelli, A.; Chang, Q.; Williams, E.C.; et al. Altered microtubule dynamics and vesicular transport in mouse and human mecp2-deficient astrocytes. Hum. Mol. Genet. 2016, 25, 146–157. [Google Scholar] [CrossRef]
  40. Ishii, M.N.; Yamamoto, K.; Shoji, M.; Asami, A.; Kawamata, Y. Human induced pluripotent stem cell (hipsc)-derived neurons respond to convulsant drugs when co-cultured with hipsc-derived astrocytes. Toxicology 2017, 389, 130–138. [Google Scholar] [CrossRef]
  41. Zhang, S.-C.; Wernig, M.; Duncan, I.D.; Brüstle, O.; Thomson, J.A. In vitro differentiation of transplantable neural precursors from human embryonic stem cells. Nat. Biotechnol. 2001, 19, 1129–1133. [Google Scholar] [CrossRef] [PubMed]
  42. Reubinoff, B.E.; Itsykson, P.; Turetsky, T.; Pera, M.F.; Reinhartz, E.; Itzik, A.; Ben-Hur, T. Neural progenitors from human embryonic stem cells. Nat. Biotechnol. 2001, 19, 1134–1140. [Google Scholar] [CrossRef] [PubMed]
  43. Krencik, R.; Zhang, S.-C. Directed Differentiation of Functional Astroglial Subtypes from Human Pluripotent Stem Cells. Nat. Protoc. 2011, 6, 1710–1717. [Google Scholar] [CrossRef] [PubMed]
  44. Caiazzo, M.; Giannelli, S.; Valente, P.; Lignani, G.; Carissimo, A.; Sessa, A.; Colasante, G.; Bartolomeo, R.; Massimino, L.; Ferroni, S.; et al. Direct conversion of fibroblasts into functional astrocytes by defined transcription factors. Stem Cell Rep. 2015, 4, 25–36. [Google Scholar] [CrossRef] [PubMed]
  45. Shaltouki, A.; Peng, J.; Liu, Q.; Rao, M.S.; Zeng, X. Efficient Generation of Astrocytes from Human Pluripotent Stem Cells in Defined Conditions. Stem Cells 2013, 31, 941–952. [Google Scholar] [CrossRef]
  46. Li, X.; Tao, Y.; Bradley, R.; Du, Z.; Tao, Y.; Kong, L.; Dong, Y.; Jones, J.; Yan, Y.; Harder, C.R.; et al. Fast Generation of Functional Subtype Astrocytes from Human Pluripotent Stem Cells. Stem Cell Rep. 2018, 11, 998–1008. [Google Scholar] [CrossRef]
  47. Watanabe, K.; Kamiya, D.; Nishiyama, A.; Katayama, T.; Nozaki, S.; Kawasaki, H.; Watanabe, Y.; Mizuseki, K.; Sasai, Y. Directed differentiation of telencephalic precursors from embryonic stem cells. Nat. Neurosci. 2005, 8, 288–296. [Google Scholar] [CrossRef] [Green Version]
  48. Eiraku, M.; Watanabe, K.; Matsuo-Takasaki, M.; Kawada, M.; Yonemura, S.; Matsumura, M.; Wataya, T.; Nishiyama, A.; Muguruma, K.; Sasai, Y. Self-Organized Formation of Polarized Cortical Tissues from ESCs and Its Active Manipulation by Extrinsic Signals. Cell Stem Cell 2008, 3, 519–532. [Google Scholar] [CrossRef] [Green Version]
  49. Lundin, A.; Delsing, L.; Clausen, M.; Ricchiuto, P.; Sanchez, J.; Sabirsh, A.; Ding, M.; Synnergren, J.; Zetterberg, H.; Brolén, G.; et al. Human iPS-Derived Astroglia from a Stable Neural Precursor State Show Improved Functionality Compared with Conventional Astrocytic Models. Stem Cell Rep. 2018, 10, 1030–1045. [Google Scholar] [CrossRef] [Green Version]
  50. Kondo, T.; Funayama, M.; Tsukita, K.; Hotta, A.; Yasuda, A.; Nori, S.; Kaneko, S.; Nakamura, M.; Takahashi, R.; Okano, H.; et al. Focal Transplantation of Human iPSC-Derived Glial-Rich Neural Progenitors Improves Lifespan of ALS Mice. Stem Cell Rep. 2014, 3, 242–249. [Google Scholar] [CrossRef] [Green Version]
  51. Tyzack, G.; Lakatos, A.; Patani, R. Human Stem Cell-Derived Astrocytes: Specification and Relevance for Neurological Disorders. Curr. Stem Cell Rep. 2016, 2, 236–247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. 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. 2009, 27, 275–280. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Falk, A.; Koch, P.; Kesavan, J.; Takashima, Y.; Ladewig, J.; Alexander, M.; Wiskow, O.; Tailor, J.; Trotter, M.; Pollard, S.; et al. Capture of neuroepithelial-like stem cells from pluripotent stem cells provides a versatile system for in vitro production of human neurons. PLoS ONE 2012, 7, e29597. [Google Scholar] [CrossRef] [PubMed]
  54. Lippmann, E.S.; Williams, C.E.; Ruhl, D.A.; Estevez-Silva, M.C.; Chapman, E.R.; Coon, J.J.; Ashton, R.S. Deterministic HOX Patterning in Human Pluripotent Stem Cell-Derived Neuroectoderm. Stem Cell Rep. 2015, 4, 632–644. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Osakada, F.; Takahashi, M. Neural induction and patterning in Mammalian pluripotent stem cells. CNS Neurol. Disord.-Drug Targets 2011, 10, 419–432. [Google Scholar] [CrossRef] [PubMed]
  56. Tao, Y.; Zhang, S.-C. Neural Subtype Specification from Human Pluripotent Stem Cells. Cell Stem Cell 2016, 19, 573–586. [Google Scholar] [CrossRef] [Green Version]
  57. Krencik, R.; Weick, J.P.; Liu, Y.; Zhang, Z.-J.; Zhang, S.-C. Specification of Transplantable Astroglial Subtypes from Human Pluripotent Stem Cells. Nat. Biotechnol. 2011, 29, 528–534. [Google Scholar] [CrossRef] [PubMed]
  58. Roybon, L.; Lamas, N.J.; Garcia-Diaz, A.; Yang, E.J.; Sattler, R.; Jackson-Lewis, V.; Kim, Y.A.; Kachel, C.A.; Rothstein, J.D.; Przedborski, S.; et al. Human Stem Cell-Derived Spinal Cord Astrocytes with Defined Mature or Reactive Phenotypes. Cell Rep. 2013, 4, 1035–1048. [Google Scholar] [CrossRef] [Green Version]
  59. Holmqvist, S.; Brouwer, M.; Djelloul, M.; Diaz, A.G.; Devine, M.J.; Hammarberg, A.; Fog, K.; Kunath, T.; Roybon, L. Generation of human pluripotent stem cell reporter lines for the isolation of and reporting on astrocytes generated from ventral midbrain and ventral spinal cord neural progenitors. Stem Cell Res. 2015, 15, 203–220. [Google Scholar] [CrossRef] [Green Version]
  60. Mi, H.; Barres, B.A. Purification and Characterization of Astrocyte Precursor Cells in the Developing Rat Optic Nerve. J. Neurosci. 1999, 19, 1049–1061. [Google Scholar] [CrossRef]
  61. Morrison, R.S.; De Vellis, J. Growth of purified astrocytes in a chemically defined medium. Proc. Natl. Acad. Sci. USA 1981, 78, 7205–7209. [Google Scholar] [CrossRef] [PubMed]
  62. Sloan, S.A.; Darmanis, S.; Huber, N.; Khan, T.A.; Birey, F.; Caneda, C.; Reimer, R.; Quake, S.R.; Barres, B.A.; Paşca, S.P. Human Astrocyte Maturation Captured in 3D Cerebral Cortical Spheroids Derived from Pluripotent Stem Cells. Neuron 2017, 95, 779–790. [Google Scholar] [CrossRef] [PubMed]
  63. Kiyoshi, C.M.; Du, Y.; Zhong, S.; Wang, W.; Taylor, A.T.; Xiong, B.; Ma, B.; Terman, D.; Zhou, M. Syncytial isopotentiality: A system-wide electrical feature of astrocytic networks in the brain. Glia 2018, 66, 2756–2769. [Google Scholar] [CrossRef] [PubMed]
  64. Khakh, B.S.; Deneen, B. The Emerging Nature of Astrocyte Diversity. Annu. Rev. Neurosci. 2019, 42, 187–207. [Google Scholar] [CrossRef] [PubMed]
  65. Tchieu, J.; Calder, E.L.; Guttikonda, S.R.; Gutzwiller, E.M.; Aromolaran, K.A.; Steinbeck, J.A.; Goldstein, P.A.; Studer, L. NFIA is a gliogenic switch enabling rapid derivation of functional human astrocytes from pluripotent stem cells. Nat. Biotechnol. 2019, 37, 267–275. [Google Scholar] [CrossRef] [PubMed]
  66. Kobayashi, E.; Nakano, M.; Kubota, K.; Himuro, N.; Mizoguchi, S.; Chikenji, T.; Otani, M.; Mizue, Y.; Nagaishi, K.; Fujimiya, M. Activated forms of astrocytes with higher GLT-1 expression are associated with cognitive normal subjects with Alzheimer pathology in human brain. Sci. Rep. 2018, 8, 1712. [Google Scholar] [CrossRef] [PubMed]
  67. Bertram, L.; Tanzi, R.E. Genome-wide association studies in alzheimer’s disease. Hum. Mol. Genet. 2009, 18, R137–R145. [Google Scholar] [CrossRef] [PubMed]
  68. Lee, S.-J.; Desplats, P.; Lee, H.-J.; Spencer, B.; Masliah, E. Cell-to-Cell Transmission of α-Synuclein Aggregates. In Amyloid Proteins: Methods and Protocols; Sigurdsson, E.M., Calero, M., Gasset, M., Eds.; Humana Press: Totowa, NJ, USA, 2012; pp. 347–359. [Google Scholar]
  69. Hernandez, D.G.; Reed, X.; Singleton, A.B. Genetics in parkinson disease: Mendelian versus non-mendelian inheritance. J. Neurochem. 2016, 139 (Suppl. 1), 59–74. [Google Scholar] [CrossRef]
  70. Shin, J.-Y.; Fang, Z.-H.; Yu, Z.-X.; Wang, C.-E.; Li, S.-H.; Li, X.-J. Expression of mutant huntingtin in glial cells contributes to neuronal excitotoxicity. J. Cell Boil. 2005, 171, 1001–1012. [Google Scholar] [CrossRef]
  71. Yoshida, T.; Nakagawa, M. Clinical aspects and pathology of alexander disease, and morphological and functional alteration of astrocytes induced by gfap mutation. Neuropathol. Off. J. Jpn. Soc. Neuropathol. 2012, 32, 440–446. [Google Scholar] [CrossRef]
  72. Messing, A.; Brenner, M.; Feany, M.B.; Nedergaard, M.; Goldman, J.E. Alexander disease. J. Neurosci. 2012, 32, 5017–5023. [Google Scholar] [CrossRef] [PubMed]
  73. Tomokane, N.; Iwaki, T.; Tateishi, J.; Iwaki, A.; Goldman, J.E. Rosenthal fibers share epitopes with alpha B-crystallin, glial fibrillary acidic protein, and ubiquitin, but not with vimentin. Immunoelectron microscopy with colloidal gold. Am. J. Pathol. 1991, 138, 875–885. [Google Scholar] [PubMed]
  74. Brosnan, C.F.; Raine, C.S. The astrocyte in multiple sclerosis revisited. Glia 2013, 61, 453–465. [Google Scholar] [CrossRef] [PubMed]
  75. Perriard, G.; Mathias, A.; Enz, L.; Canales, M.; Schluep, M.; Gentner, M.; Schaeren-Wiemers, N.; Du Pasquier, R.A. Interleukin-22 is increased in multiple sclerosis patients and targets astrocytes. J. Neuroinflamm. 2015, 12, 119. [Google Scholar] [CrossRef] [PubMed]
  76. Sztainberg, Y.; Zoghbi, H.Y. Lessons learned from studying syndromic autism spectrum disorders. Nat. Neurosci. 2016, 19, 1408–1417. [Google Scholar] [CrossRef] [PubMed]
  77. Maezawa, I.; Swanberg, S.; Harvey, D.; LaSalle, J.M.; Jin, L.-W. Rett syndrome astrocytes are abnormal and spread MeCP2 deficiency through gap junctions. J. Neurosci. 2009, 29, 5051–5061. [Google Scholar] [CrossRef]
  78. Ananiev, G.; Williams, E.C.; Li, H.; Chang, Q. Isogenic Pairs of Wild Type and Mutant Induced Pluripotent Stem Cell (iPSC) Lines from Rett Syndrome Patients as In Vitro Disease Model. PLoS ONE 2011, 6, e25255. [Google Scholar] [CrossRef]
  79. Miyake, K.; Yang, C.; Minakuchi, Y.; Ohori, K.; Soutome, M.; Hirasawa, T.; Kazuki, Y.; Adachi, N.; Suzuki, S.; Itoh, M.; et al. Comparison of genomic and epigenomic expression in monozygotic twins discordant for rett syndrome. PLoS ONE 2013, 8, e66729. [Google Scholar]
  80. Zhou, R.; Jiang, G.; Tian, X.; Wang, X. Progress in the molecular mechanisms of genetic epilepsies using patient-induced pluripotent stem cells. Epilepsia Open 2018, 3, 331–339. [Google Scholar] [CrossRef]
  81. Perea, G.; Navarrete, M.; Araque, A. Tripartite synapses: Astrocytes process and control synaptic information. Trends Neurosci. 2009, 32, 421–431. [Google Scholar] [CrossRef]
  82. Tian, G.F.; Azmi, H.; Takano, T.; Xu, Q.; Peng, W.; Lin, J.; Oberheim, N.; Lou, N.; Wang, X.; Zielke, H.R.; et al. An astrocytic basis of epilepsy. Nat. Med. 2005, 11, 973–981. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Table 1. Modeling astrocyte contributions in neurological disorders with human pluripotent stem cells.
Table 1. Modeling astrocyte contributions in neurological disorders with human pluripotent stem cells.
DiseaseSource Gene Mutation/ModificationReference
Alzheimer’s disease (AD)Patient derived iPSCsAPP E693Δ mutation
APP V717L mutation
Sporadic AD
Kondo et al., 2013 [24]
Alzheimer’s disease (AD)Patient derived iPSCsPSEN1 exon 9 deletionOksanen et al., 2017 [25]
Parkinson’s disease (PD)Patient derived iPSCsLRRK2 G2019S mutationdi Domenico et al., 2019 [26]
Huntington’s disease (HD)Patient derived iPSCsHTT CAG repeat expansionJuopperi et al. 2012 [27]
TauopathyPatient derived iPSCsMAPT N279K mutationHallmann et al. 2017 [28]
TDP-43 proteinopathyPatient derived iPSCsTDP-43 M337V mutationSerio et al., 2013 [29]
Amyotrophic lateral sclerosis (ALS)Human ESCsSOD1 G93A overexpressingWada et al., 2012 [30]
Amyotrophic lateral sclerosis (ALS)Patient derived iPSCsSOD1 D90A mutationTyzack et al., 2017 [31]
Spinal muscular atrophy (SMA)Patient derived iPSCsSMN1 exons 7 and 8 deletionMcGivern et al., 2013 [32]
Alexander’s disease (AxD)Patient derived iPSCsGFAP R239C mutation
GFAP E69K mutation
GFAP L264P mutation
Kondo et al., 2016 [33]
Alexander’s disease (AxD)Patient derived iPSCsGFAP R239C mutation
GFAP R79C mutation
GFAP M73K mutation
Li et al., 2018 [34]
Multiple sclerosis (MS)Patient derived iPSCs Perriot et al., 2018 [35]
Autism spectrum disorder (ASD)Patient derived iPSCs Russo et al., 2018 [36]
Rett syndrome (RTT)Patient derived iPSCsMECP2 V247X mutation
MECP2 R294X mutation
MECP2 R306C mutation
Williams et al., 2014 [37]
Rett syndrome (RTT)Patient derived iPSCsMECP2 G269AfsX288 mutationAndoh-Noda et al., 2015 [38]
Rett syndrome (RTT)Patient derived iPSCsMECP2 R294X mutationDelépine et al., 2016 [39]
Drug-induced epilepsyHuman iPSCs Ishii et al., 2017 [40]

Share and Cite

MDPI and ACS Style

Suga, M.; Kondo, T.; Inoue, H. Modeling Neurological Disorders with Human Pluripotent Stem Cell-Derived Astrocytes. Int. J. Mol. Sci. 2019, 20, 3862. https://doi.org/10.3390/ijms20163862

AMA Style

Suga M, Kondo T, Inoue H. Modeling Neurological Disorders with Human Pluripotent Stem Cell-Derived Astrocytes. International Journal of Molecular Sciences. 2019; 20(16):3862. https://doi.org/10.3390/ijms20163862

Chicago/Turabian Style

Suga, Mika, Takayuki Kondo, and Haruhisa Inoue. 2019. "Modeling Neurological Disorders with Human Pluripotent Stem Cell-Derived Astrocytes" International Journal of Molecular Sciences 20, no. 16: 3862. https://doi.org/10.3390/ijms20163862

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop