welcome to stewart anderson lab | stewart anderson lab ......cell stem cell article directed...

31
Cell Stem Cell Article Directed Differentiation and Functional Maturation of Cortical Interneurons from Human Embryonic Stem Cells Asif M. Maroof, 1,2,8 Sotirios Keros, 4,5 Jennifer A. Tyson, 3 Shui-Wang Ying, 4 Yosif M. Ganat, 1,2 Florian T. Merkle, 8 Becky Liu, 1,2 Adam Goulburn, 3 Edouard G. Stanley, 6 Andrew G. Elefanty, 6 Hans Ruedi Widmer, 7 Kevin Eggan, 8 Peter A. Goldstein, 4 Stewart A. Anderson, 3,9, * and Lorenz Studer 1,2, * 1 Center for Stem Cell Biology 2 Developmental Biology Program Sloan-Kettering Institute for Cancer Research, 1275 York Ave, New York, NY 10065, USA 3 Department of Psychiatry 4 Department of Anesthesiology 5 Division of Pediatric Neurology, Department of Pediatrics Weill Cornell Medical College, 1300 York Ave, New York, NY 10021, USA 6 Monash Immunology and Stem Cell Laboratories, Monash University, Building 75, STRIP1, West Ring Road, Clayton, Victoria 3800, Australia 7 Department of Neurosurgery, University of Bern, Inselspital, CH-3010 Bern, Switzerland 8 Department of Stem Cell and Regenerative Biology, Harvard University, Cambridge, MA 02138, USA 9 Department of Psychiatry, Children’s Hospital of Philadelphia and University of Pennsylvania Medical School, Philadelphia, PA 19104, USA *Correspondence: [email protected] (S.A.A.), [email protected] (L.S.) http://dx.doi.org/10.1016/j.stem.2013.04.008 SUMMARY Human pluripotent stem cells are a powerful tool for modeling brain development and disease. The human cortex is composed of two major neuronal populations: projection neurons and local interneu- rons. Cortical interneurons comprise a diverse class of cell types expressing the neurotransmitter GABA. Dysfunction of cortical interneurons has been impli- cated in neuropsychiatric diseases, including schizo- phrenia, autism, and epilepsy. Here, we demonstrate the highly efficient derivation of human cortical inter- neurons in an NKX2.1::GFP human embryonic stem cell reporter line. Manipulating the timing of SHH activation yields three distinct GFP+ populations with specific transcriptional profiles, neurotrans- mitter phenotypes, and migratory behaviors. Further differentiation in a murine cortical environment yields parvalbumin- and somatostatin-expressing neurons that exhibit synaptic inputs and electrophysiological properties of cortical interneurons. Our study defines the signals sufficient for modeling human ventral forebrain development in vitro and lays the founda- tion for studying cortical interneuron involvement in human disease pathology. INTRODUCTION Human pluripotent stem cells (hPSCs) are a powerful tool for the study of human development and disease and for applica- tions in regenerative medicine. The use of hPSCs differentiated toward CNS lineages has been of particular interest given the lack of effective therapies for many neurodegenerative and neuropsychiatric disorders and the availability of protocols for efficiently directing neuronal specification in vitro (Chambers et al., 2009). Early studies using hPSCs have been primarily geared toward neurodegenerative disorders, which are known to affect specific neuron types such as midbrain dopamine neurons in Parkinson’s disease (PD; Kriks et al., 2011) or motor neurons in amyotrophic lateral sclerosis (ALS; Dimos et al., 2008) and spinal muscular atrophy (SMA; Ebert et al., 2009). More recent studies suggest the possibility of tackling complex neuronal disorders such as schizophrenia (Brennand et al., 2011) or autism-related syndromes (Marchetto et al., 2010; Pas xca et al., 2011). Unlike in PD, ALS, and SMA, the neuron types critical for modeling schizophrenia or autism are less well defined, and no attempts have been made to direct neuron subtype identity in those studies. There has been considerable progress in establishing protocols for the deriva- tion of human embryonic stem cell (ESC)-derived cortical projection neurons (Espuny-Camacho et al., 2013; Shi et al., 2012). However, inhibitory neurons such as cortical inter- neurons may be particularly important in schizophrenia or autism (Insel, 2010; Lewis et al., 2005). We have previously demonstrated the derivation of cortical interneurons using an Lhx6::GFP reporter mouse ESC line (Maroof et al., 2010). How- ever, the efficiency of cortical interneuron generation was low, and it was uncertain whether those conditions would apply for generating human cortical interneurons from PSCs. Modeling the development of human cortical interneurons in vitro is of particular interest, because their developmental origin is controversial: studies suggest considerable differences across mammalian species (Letinic et al., 2002; Yu and Zecevic, 2011). Furthermore, the protracted in vivo development of several cortical interneuron types (Anderson et al., 1995) repre- sents an additional challenge for modeling their differentiation using human cells in vitro. Cell Stem Cell 12, 559–572, May 2, 2013 ª2013 Elsevier Inc. 559 STEM 1331

Upload: others

Post on 04-Feb-2021

0 views

Category:

Documents


0 download

TRANSCRIPT

  • Cell Stem Cell

    Article

    Directed Differentiation and FunctionalMaturation of Cortical Interneuronsfrom Human Embryonic Stem CellsAsif M. Maroof,1,2,8 Sotirios Keros,4,5 Jennifer A. Tyson,3 Shui-Wang Ying,4 Yosif M. Ganat,1,2 Florian T. Merkle,8

    Becky Liu,1,2 Adam Goulburn,3 Edouard G. Stanley,6 Andrew G. Elefanty,6 Hans Ruedi Widmer,7 Kevin Eggan,8

    Peter A. Goldstein,4 Stewart A. Anderson,3,9,* and Lorenz Studer1,2,*1Center for Stem Cell Biology2Developmental Biology Program

    Sloan-Kettering Institute for Cancer Research, 1275 York Ave, New York, NY 10065, USA3Department of Psychiatry4Department of Anesthesiology5Division of Pediatric Neurology, Department of Pediatrics

    Weill Cornell Medical College, 1300 York Ave, New York, NY 10021, USA6Monash Immunology and StemCell Laboratories, MonashUniversity, Building 75, STRIP1,West Ring Road, Clayton, Victoria 3800, Australia7Department of Neurosurgery, University of Bern, Inselspital, CH-3010 Bern, Switzerland8Department of Stem Cell and Regenerative Biology, Harvard University, Cambridge, MA 02138, USA9Department of Psychiatry, Children’s Hospital of Philadelphia and University of Pennsylvania Medical School, Philadelphia, PA 19104, USA

    *Correspondence: [email protected] (S.A.A.), [email protected] (L.S.)

    http://dx.doi.org/10.1016/j.stem.2013.04.008

    SUMMARY

    Human pluripotent stem cells are a powerful toolfor modeling brain development and disease. Thehuman cortex is composed of two major neuronalpopulations: projection neurons and local interneu-rons. Cortical interneurons comprise a diverse classof cell types expressing the neurotransmitter GABA.Dysfunction of cortical interneurons has been impli-cated in neuropsychiatric diseases, including schizo-phrenia, autism, and epilepsy. Here, we demonstratethe highly efficient derivation of human cortical inter-neurons in an NKX2.1::GFP human embryonic stemcell reporter line. Manipulating the timing of SHHactivation yields three distinct GFP+ populationswith specific transcriptional profiles, neurotrans-mitter phenotypes, and migratory behaviors. Furtherdifferentiation in amurine cortical environment yieldsparvalbumin- and somatostatin-expressing neuronsthat exhibit synaptic inputs and electrophysiologicalproperties of cortical interneurons. Our study definesthe signals sufficient for modeling human ventralforebrain development in vitro and lays the founda-tion for studying cortical interneuron involvement inhuman disease pathology.

    INTRODUCTION

    Human pluripotent stem cells (hPSCs) are a powerful tool for

    the study of human development and disease and for applica-

    tions in regenerative medicine. The use of hPSCs differentiated

    toward CNS lineages has been of particular interest given the

    lack of effective therapies for many neurodegenerative and

    neuropsychiatric disorders and the availability of protocols for

    efficiently directing neuronal specification in vitro (Chambers

    et al., 2009). Early studies using hPSCs have been primarily

    geared toward neurodegenerative disorders, which are known

    to affect specific neuron types such as midbrain dopamine

    neurons in Parkinson’s disease (PD; Kriks et al., 2011) or motor

    neurons in amyotrophic lateral sclerosis (ALS; Dimos et al.,

    2008) and spinal muscular atrophy (SMA; Ebert et al., 2009).

    More recent studies suggest the possibility of tackling complex

    neuronal disorders such as schizophrenia (Brennand et al.,

    2011) or autism-related syndromes (Marchetto et al., 2010;

    Pasxca et al., 2011). Unlike in PD, ALS, and SMA, the neurontypes critical for modeling schizophrenia or autism are less

    well defined, and no attempts have been made to direct

    neuron subtype identity in those studies. There has been

    considerable progress in establishing protocols for the deriva-

    tion of human embryonic stem cell (ESC)-derived cortical

    projection neurons (Espuny-Camacho et al., 2013; Shi et al.,

    2012). However, inhibitory neurons such as cortical inter-

    neurons may be particularly important in schizophrenia or

    autism (Insel, 2010; Lewis et al., 2005). We have previously

    demonstrated the derivation of cortical interneurons using an

    Lhx6::GFP reporter mouse ESC line (Maroof et al., 2010). How-

    ever, the efficiency of cortical interneuron generation was low,

    and it was uncertain whether those conditions would apply for

    generating human cortical interneurons from PSCs. Modeling

    the development of human cortical interneurons in vitro is

    of particular interest, because their developmental origin is

    controversial: studies suggest considerable differences across

    mammalian species (Letinic et al., 2002; Yu and Zecevic,

    2011). Furthermore, the protracted in vivo development of

    several cortical interneuron types (Anderson et al., 1995) repre-

    sents an additional challenge for modeling their differentiation

    using human cells in vitro.

    Cell Stem Cell 12, 559–572, May 2, 2013 ª2013 Elsevier Inc. 559

    STEM 1331

    mailto:[email protected]:[email protected]://dx.doi.org/10.1016/j.stem.2013.04.008

  • Here, we present a small-molecule-based strategy for the

    efficient induction of human cortical interneurons. Exposure to

    the tankyrase inhibitor XAV939 enhances forebrain differentia-

    tion. Using an NKX2.1::GFP hESC reporter line, we demonstrate

    the selective derivation of three distinct ventral forebrain precur-

    sor populations by combining XAV939 treatment with the timed

    activation of sonic hedgehog (SHH) signaling. Importantly,

    results in both hPSCs and human embryos reveal differences

    between mouse and human forebrain development, such as

    the human-specific, yet transient, FOXA2 expression within the

    ventral forebrain. Finally, mature functional properties and

    the expression of late cortical interneuron markers, including

    parvalbumin (PV) and somatostatin (SST), demonstrate the

    feasibility of studying human cortical interneuron differentiation

    from hPSCs in vitro despite their protracted development in vivo.

    RESULTS

    Combined Inhibition of Wnt and Activation of SHHSignaling Triggers Efficient Induction of NKX2.1+ NeuralPrecursorsThe first step in modeling cortical interneuron development is

    the robust induction of forebrain precursors. Specification of

    anterior and forebrain fate is considered a default program

    during neural differentiation of hPSCs (Chambers et al., 2009;

    Eiraku et al., 2008; Espuny-Camacho et al., 2013; Gaspard

    et al., 2008). However, not all cell lines adopt anterior neural fates

    at equal efficiencies (Kim et al., 2011; Wu et al., 2007). Patterning

    factors secreted within differentiating cultures such as fibroblast

    growth factors (FGFs), Wnts, or retinoids can suppress forebrain

    induction and trigger the induction of caudal cell fates. We

    recently reported that early, high-dose SHH treatment also

    suppresses forebrain markers such as FOXG1 via inhibition of

    DKK1 induction (Fasano et al., 2010). This is a concern for

    deriving cortical interneurons that are thought to require strong

    SHH activation at the NKX2.1+ progenitor stage (Xu et al.,

    2010). Here, we tested whether pharmacological inhibition of

    WNT signaling can improve FOXG1 induction and subsequently

    enable the controlled, SHH-mediated ventralization toward

    NKX2.1+ forebrain progenitor fates (Figure 1A).

    Neural differentiation of hESCs via the dual SMAD-inhibition

    protocol (Chambers et al., 2009) using Noggin + SB431542

    (NSB) robustly induced FOXG1+/PAX6+ precursors. Replace-

    ment of Noggin with the ALK2/3 inhibitor LDN-193189 (LSB)

    induced PAX6 expression equally well but showed a trend to-

    ward lower percentages of FOXG1+ cells (Figures 1B and 1C).

    Adding recombinant DKK1 or the tankyrase inhibitor XAV939

    (Huang et al., 2009), inhibitors of canonical Wnt signaling,

    enhanced FOXG1 expression in LSB-treated cultures (Figures

    1B–1D). Importantly, the effect of XAV939 on FOXG1 induction

    was consistent (Figure 1E) across multiple independent hESC

    (HES-3 and WA-09) and human induced PSC (iPSC) lines (C72

    line, Papapetrou et al., 2009; SeV6, Kriks et al., 2011). Therefore,

    the use of three small molecules (XAV939, LSB, and SB431542;

    termed XLSB) enables rapid and robust induction of forebrain

    fates across human ESCs and iPSC lines.

    We next wanted to test our ability to induce ventral fates using

    XLSB in the presence of SHH activators (Figure 1A). The tran-

    scription factor NKX2.1 is a marker of ventral prosencephalic

    progenitor populations (Sussel et al., 1999; Xu et al., 2004). Using

    a previously established NKX2.1::GFP knockin reporter hESC

    line (Goulburn et al., 2011), we observed GFP induction by day

    10 following activation of the SHH pathway by recombinant

    SHH (R&D Systems, C-25II) and the smoothened activator

    purmorphamine (Figure 1F). We found maximal induction of

    NKX2.1::GFP at day 18 upon treatment with 1 mM purmorph-

    amine and 5nM SHH (Figure 1G), a condition referred to as

    ‘‘SHH’’ for the remainder of the manuscript. Our past studies

    on the derivation of hESC-derived floor plate cells demonstrated

    that early treatment with SHH (day 1 of differentiation) is critical

    for the efficient induction of FOXA2 (Fasano et al., 2010). In

    contrast, we observed here that cells exposed to SHH at late

    differentiation stages (day 10 of the XLSB protocol) remain

    competent for inducing NKX2.1::GFP as measured by fluores-

    cence-activated cell sorting (FACS) (Figure 1H). Therefore,

    XLSB+SHH treatment represents a strategy for generating

    NKX2.1+ progenitors at high efficiencies.

    Timing of SHH Exposure Induces Distinct VentralProgenitor PopulationsThe efficient induction of NKX2.1 largely independent of the

    timing of SHH exposure raised the question of whether timing

    impacts the subtype of NKX2.1+ neural progenitors generated.

    Analogous to rodents, NKX2.1 is expressed during early neural

    human development in both the FOXG1+ telencephalon and

    the FOXG1-negative ventral diencephalon (Figure 2A; Kerwin

    et al., 2010; Rakic and Zecevic, 2003). During embryonic

    mouse development, NKX2.1 expression in the telencephalon

    (FOXG1+) is restricted to the ventral (subcortical) domain. In

    contrast, reports in human fetal tissue suggested that NKX2.1

    may not be restricted to the ventral forebrain, instead extending

    into the dorsal forebrain and cortical anlage (Rakic and Zecevic,

    2003; Yu and Zecevic, 2011). However, those studies were

    largely based on the analysis of second-trimester human

    fetuses, making it difficult to distinguish whether NKX2.1+ cells

    were induced in the dorsal forebrain or migrated dorsally after

    induction in the ventral domain (Fertuzinhos et al., 2009). By

    performing immunocytochemical analyses in early-stage human

    embryos (Carnegie stage 15 [CS15],�38 days post conception),we observed restriction of NKX2.1 expression to the ventral

    forebrain (Figures 2B–2E) comparable to the developing mouse

    CNS. The NKX2.1 domain in the human embryonic ventral

    forebrain was further subdivided into an OLIG2+ ganglionic

    eminence and an OLIG2-negative preoptic-area anlage (Figures

    2C and 2D).

    To address whether in vitro timing of SHH exposure affects the

    regional identity of the resulting hPSC-derived NKX2.1::GFP+

    cells, we compared the identities of sorted cells at day 18 of

    differentiation (Figure 2F). In agreement with our previous results

    on floor-plate induction (Fasano et al., 2010), we observed that

    early SHH exposure (days 2–18) suppressed FOXG1 induction

    (Figure 2G, left panel) despite robust induction of NKX2.1::GFP

    (see Figure 1H). In contrast, both day 6–18 and day 10–18

    treatment groups showed expression of FOXG1 but differed

    in OLIG2 expression (Figure 2G, middle and right panels).

    We hypothesize that differential expression of FOXG1 and

    OLIG2 reflect the distinct NKX2.1+ precursor domains observed

    during early human development (Figures 2A–2E) and mimic the

    STEM 1331

    Cell Stem Cell

    hESC-Derived Cortical Interneurons

    560 Cell Stem Cell 12, 559–572, May 2, 2013 ª2013 Elsevier Inc.

  • Olig2+ domains during mouse forebrain development (Flames

    et al., 2007). More than 90% of the NKX2.1::GFP+ cells in the

    6–18 and 10–18 protocols coexpressed FOXG1, and nearly

    80% of the GFP+ cells in the 10–18 protocol coexpressed

    OLIG2 (Figure 2H).

    The impact of the timing of SHH exposure on the derivation of

    specific ventral precursor populations was further examined by

    global transcriptome analysis (Figures 2I–2K; Table S1 available

    online; all raw data are available in theGene ExpressionOmnibus

    [GEO], http://www.ncbi.nlm.nih.gov/geo/). A direct comparison

    of day 10–18 versus untreated cultures (no SHH) confirmed

    the robust induction of ventral specification markers such as

    NKX2.1, NKX2.2, ASCL1, SIX6, OLIG2, and NKX6.2, which

    were induced at the expense of dorsal forebrain markers such

    Figure 1. Wnt Inhibition and Activation of SHH Signaling Yields Highly Efficient Derivation of Forebrain Fates and NKX2.1 Induction

    (A) Schematic of the differentiation protocol in the dual-SMAD inhibition paradigm for generating anterior neural progenitors.

    (B–E) When either DKK1 or XAV939, both Wnt-signaling antagonists, was added to the dual-SMAD inhibition protocol (DLSB or XLSB), there was a significant

    increase in the percentage of FOXG1+ cells (B) without a loss of PAX6 expression (C): **p < 0.01; ***p < 0.001; n.s., not significant; using ANOVA followed by

    Scheffe test. (D) Representative immunofluorescent image for FOXG1 (red) and PAX6 (green) expression at day 10 following XLSB treatment. Single-channel

    fluorescent images of the marked region are shown in the three right panels. (E) Robust telencephalic specification using XLSBwas also observed at comparable

    efficiencies in hiPSCs (lines SeV6 and C72; n = 4).

    (F–H) Addition of SHH signaling to the XLSB protocol significantly enhanced the production of NKX2.1::GFP-expressing progenitors. (F) Added from day 4, 5 nM

    SHH (C24II) and 1 mm purmorphamine showed synergistic effects in inducing NKX2.1::GFP expression at day 10 (***p < 0.001, compared to SHH). A range

    of concentrations of SHH and purmorphamine are compared at day 18 in (G), and again, cotreatment was greatly superior to quite high concentrations of

    either SHH or purmorphamine alone (***p < 0.001, compared to no SHH, using ANOVA followed by Scheffe test). (H) Delaying the timing of SHH exposure

    between 2 and 10 days of differentiation did not dramatically affect the efficiency of NKX2.1::GFP induction measured at day 18 (***p < 0.001 compared to 0–18,

    using ANOVA followed by Scheffe test). P, purmorphamine; S, sonic hedgehog.

    Data are from hESC line HES-3 (NKX2.1::GFP) in (B), (C), and (F)–(H); from hESC line WA-09/H9 in (D); and from hESC line WA-09/H9 and hiPSC lines SeV6

    and C72 in (E). The scale bar in (D) represents 125 mm. Data in (B), (C), and (E)–(H) are represented as the mean ± SEM.

    STEM 1331

    Cell Stem Cell

    hESC-Derived Cortical Interneurons

    Cell Stem Cell 12, 559–572, May 2, 2013 ª2013 Elsevier Inc. 561

    http://www.ncbi.nlm.nih.gov/geo/

  • Figure 2. Timing of SHH Exposure Determines the Regional Identity of NKX2.1::GFP-Expressing Progenitors

    (A) Model of human prosencephalon (sagittal view at CS14) with expression of forebrain patterning markers based on published data (Kerwin et al., 2010).

    (B–E) Coronal (oblique) hemisection of the human prosencephalon at CS15 demonstrates expression of NKX2.1, OLIG2, and PAX6. NKX2.1 and OLIG2 are

    expressed in various regions throughout the ventral prosencephalon, whereas PAX6 is restricted to the dorsal prosencephalon and the eye. The expression of

    these proteins is nonoverlapping, except in the ganglionic eminence (C), where OLIG2 and NKX2.1 are coexpressed. The scale bar in (B) represents 200 mm.

    (F) Schematic illustration of the distinct time periods of SHH and purmorphamine treatment used in combination with the XLSB protocol.

    (legend continued on next page)

    STEM 1331

    Cell Stem Cell

    hESC-Derived Cortical Interneurons

    562 Cell Stem Cell 12, 559–572, May 2, 2013 ª2013 Elsevier Inc.

  • as EMX2 and PAX6 (Figure 2I). There were no significant differ-

    ences in the expression of general anterior markers such as

    FOXG1 and OTX2 (Figure 2I), supporting the notion that both

    populations are of a forebrain identity. Comparison of day 10–

    18 and day 2–18 treated cultures (Figure 2J) illustrated differ-

    ences in the expression of anterior markers such as FOXG1

    versus diencephalic markers such as RAX. In contrast, day 10–

    18 versus day 6–18 treated cultures showed less pronounced

    differences in forebrain marker expression (Figure 2K). Overall,

    our data demonstrate that the window of SHH treatment signifi-

    cantly impacts the identity of hPSC-derived NKX2.1 precursors.

    One surprising finding from the gene-expression data was the

    induction of FOXA2 in all SHH-treated cultures (Figure 2I).

    Expression of FOXA2, a marker thought to be largely restricted

    to the floor plate within the developing CNS, was confirmed at

    day 18 of differentiation in NKX2.1::GFP+ cells of all three SHH

    treatment paradigms (Figures S1A and S1B). To address

    whether FOXA2 expression represents an in vitro artifact

    following strong extrinsic activation of SHH signaling or whether

    FOXA2 expression occurs in the telencephalon during in vivo

    development, we performed histological analyses in the devel-

    oping mouse and human forebrain. In mouse embryos (embry-

    onic day 11.5 [E11.5]), FOXA2 expression within the mouse

    prosencephalon was segregated from the FOXG1 and NKX2.1

    domains (Figure S1C). However, immunohistochemical analysis

    for FOXA2 in primary human forebrain tissue (CS15 embryo)

    confirmed expression in the telencephalic NKX2.1+ ganglionic

    eminence (Figures S1D–S1F). These in vivo data are in agree-

    ment with the in vitro hPSC differentiation data demonstrating

    (transient) FOXA2 expression during human ventral forebrain

    development.

    Neuronal Differentiation and Migratory Properties ofhESC-Derived NKX2.1+ PrecursorsCortical interneurons are important for balancing excitation and

    inhibition within cortical circuitry and play critical roles in control-

    ling developmental plasticity. Cortical interneuron dysfunction

    has been implicated in various pathological conditions such as

    epilepsy, autism, and schizophrenia (Baraban et al., 2009; Insel,

    2010; Lewis et al., 2005; Peñagarikano et al., 2011). Our data

    demonstrate that day 10–18 treated cells express cortical inter-

    neuron progenitor markers such asOLIG2,MASH1, andNKX6.2.

    At day 18 of differentiation, NKX2.1+ cells exhibit progenitor cell

    properties based on Nestin and Ki67 expression. Prolonged cul-

    ture in the absence of extrinsic SHH activation led to a gradual

    decrease in Ki67 expression (Figures 3A–3C, upper panels)

    and increased percentages of cells expressing the neuronal pre-

    cursor markers DCX and TUJ1 (Figures 3A–3C, lower panels). In

    addition, many GFP+ cells in the 10–18 and 2–18 conditions ex-

    press g-aminobutyric acid (GABA) and subsequently calbindin

    (Figures 3A–3C, lower panels), a marker of tangentially migrating

    cortical interneuron precursors (Anderson et al., 1997). Impor-

    tantly, western blot analysis at day 32 showed that LHX6 protein,

    a marker of postmitotic, migratory cortical interneuron precur-

    sors in the mouse, was selectively enriched in the 10–18 condi-

    tion. In contrast, the hypothalamic marker RAX was selectively

    enriched in the 2–18 condition (Figure 3D). Additional immunocy-

    tochemical data showed expression of DLX2 and ASCL1 in most

    neuronal progeny derived from the day 10–18 NKX2.1::GFP+

    progenitors (Figures 3E–3G). These data indicate that following

    withdrawal of extrinsic SHH activators, NKX2.1::GFP+ progeni-

    tors give rise to region-specific postmitotic neurons including

    immature cortical interneurons.

    We next asked whether hESC-derived putative cortical inter-

    neuron precursors (day 10–18 group) exhibit the characteristic

    migratory potential observed for primary cortical interneurons

    in the mouse (Anderson et al., 1997) and human (Letinic et al.,

    2002) brain. In both species, major subclasses of cortical inter-

    neurons undergo tangential migration on their way from the

    ventral telencephalon into the cortex (Anderson et al., 1997; Fer-

    tuzinhos et al., 2009; Letinic et al., 2002). NKX2.1::GFP+ cells at

    day 32 of differentiation were collected via FACS and injected

    into forebrain slices isolated from E13.5 mouse embryos. The

    cell injection was carefully targeted to the medial ganglionic

    eminence under microscopic visual guidance (Figure 3H). The

    migratory potential of NKX2.1::GFP+ cells and their ability to

    reach the cortex (see zone 2 in Figure 3H) were assessed for

    each treatment group (2–18, 6–18, and 10–18 of SHH treatment;

    Figures 3I–3L). At day 2 (Figure 3M), and more pronounced at

    day 6 after injection (Figure 3N), GFP+ cells were observed

    migrating from the injection site toward the cortex (zone 2).

    Remarkably, human cells from the 10–18 treatment, but not

    the two other treatment groups, showed a robust propensity

    formigration into the neocortical portions of themurine slice (Fig-

    ures 3I–3N). These data suggest that day 10–18 treated cells

    exhibit migratory properties previously described for interneuron

    precursors derived from the primary mouse medial ganglionic

    eminence (MGE) (Sussel et al., 1999; Wichterle et al., 1999).

    To further probe the migratory capacity of the day 2–18

    versus 10–18 treated cells, we transplanted FACS-isolated

    NKX2.1::GFP+ cells into the neonatal neocortex of genetically

    immunocompromised mice. Four weeks after transplantation

    we observed extensive migration in both the radial and tangen-

    tial planes within the mouse cortex, but only for the day 10–18

    group (Figures 3O and 3P). Similar results were obtained upon

    transplantation into the adult cortex (Figure S2), though overall

    migration of NKX2.1::GFP+ cells by day 30 in vivo was less

    extensive compared to the neonatal grafts.

    (G andH) Immunofluorescence for OLIG2 and FOXG1 in theNKX2.1::GFP line at day 18 of differentiation. Treatmentwith SHH after day 6 (6–18 and 10–18 groups)

    significantly increases the percentage of NKX2.1::GFP+ cells that coexpressed FOXG1, as quantified in (H). The scale bar in (G) represents 50 mm. Treatment

    with SHH after day 10 (10–18 group) enhanced the derivation of NKX2.1::GFP+ cells coexpressing OLIG2 (data are the mean ± SEM; *p < 0.05, ***p < 0.001,

    compared to 2–18 using ANOVA followed by Scheffe test). Expression of FOXG1, NKX2.1, and OLIG2 indicates a pattern characteristic of the ganglionic

    eminence (Tekki-Kessaris et al., 2001).

    (I–K) Microarray data from cells sorted for NKX2.1::GFP expression at day 18 of differentiation, comparing gene-expression levels between the SHH day 10–18

    protocol versus no SHH (I), the day 10–18 versus the day 2–18 protocol (J), and the day 10–18 versus day 6–18 protocol (K). Red bars indicate genes expressed at

    higher levels in the SHH 10–18 protocol; blue bars indicate genes expressed at lower levels in the day 10–18 protocol. All changes are significant at p < 0.001.

    See also Figure S1 and Table S1.

    STEM 1331

    Cell Stem Cell

    hESC-Derived Cortical Interneurons

    Cell Stem Cell 12, 559–572, May 2, 2013 ª2013 Elsevier Inc. 563

  • To determine the extent of interneuron precursor maturation

    in vivo, we assessed morphological appearance and the expres-

    sion of interneuron-related markers in the grafted cells from the

    day 10–18 group (Figure S3). At 1–2 months after grafting into

    the neonatal cortex most human cells retained undifferentiated

    bipolar or unipolar morphologies and NKX2.1 expression.

    Figure 3. Conversion of Cycling Neural Progenitors to Neuronal Precursors and Assessment of Their Migratory Potential

    (A–C) At day 18, cells from the three indicated protocols were subjected to FACS forNKX2.1::GFP expression, then replated and evaluated for colabeling with the

    markers indicated. For all three protocols there was a decline in colabeling with markers of progenitors (upper panels: red line, nestin; blue line, Ki67) and an

    increase in markers of neuronal differentiation (lower panels: green, GABA; yellow, TUJ1; purple, doublecortin [DCX]; pink, calbindin). Data are the mean ± SEM.

    (D) Western blotting showed an increase in the hypothalamic-enriched protein RAX in the 2 to 18 condition and an increase in the MGE-enriched protein LHX6 in

    the 10 to 18 condition. Cells were sorted for NKX2.1::GFP prior to analysis.

    (E–G) At day 32, many of the NKX2.1::GFP+ cells from the 10 to 18 condition also expressed DLX2 and ASCL1.

    (H–N) Grafting of day 32-sortedNKX2.1::GFP+ cells into theMGE of E13.5 coronal mouse slice cultures. (H) Schematic of coronal hemisection demonstrating the

    site of transplantation and the zones for quantification of migration. (I and J) In both the 2 to 18 and the 6 to 18 conditions, very few cells migrated into zone 1, and

    fewer into zone 2. (K and L) Only the 10 to 18 condition demonstrated significant and robust migration into the cortical and striatal regions, with many GFP+ cells

    exhibiting bipolar morphologies consistent with a migratory cell (L). The regions where GFP+ cells were detected were quantified 2 (M) and 6 (N) days post

    transplantation (DPT). Data are the mean ± SEM; *p < 0.05, **p < 0.01 using ANOVA followed by Scheffe test.

    (O and P) Transplantation of day 32-sorted NKX2.1::GFP+ cells (day 10 to 18 protocol) into the neocortex of neonatal mice followed by their evaluation in fixed

    sections at postnatal day 30. In marked contrast to the MGE-like cells from the SHH 10–18 protocol (P), neither the SHH 2–18 protocol (O) nor the SHH 6–18

    protocol (data not shown) resulted in extensive migration from the graft site. The scale bar in (P) represents 200 mm.

    See also Figures S2 and S3.

    STEM 1331

    Cell Stem Cell

    hESC-Derived Cortical Interneurons

    564 Cell Stem Cell 12, 559–572, May 2, 2013 ª2013 Elsevier Inc.

  • Furthermore, essentially all grafted cells expressed the neuronal

    precursor marker DCX, including those with multipolar morphol-

    ogies (Figure S3). As in rodents, the large majority of human

    interneuron precursors expressed GABA (46/51 GFP+ cells

    located at the section surface; Figure S3). In contrast, there

    was no colabeling for PV or SST, two proteins marking the

    major subclasses of Nkx2.1-lineage interneurons. These data

    suggest that even by 6 or 8 weeks posttransplant, the Nkx2.1-

    GFP+ cells had not yet differentiated into mature interneurons

    (Figure S3).

    One explanation for the apparent lack of differentiation of

    the grafted GFP+ cells in the mouse host cortex could be that

    NKX2.1 is downregulated following maturation, as occurs in

    mouse cortical interneurons (Marin et al., 2000), resulting in the

    loss of GFP expression. For testing this possibility, sections

    were examined for expression of SC121, a human-specific pan-

    neuronal marker (Kelly et al., 2004). As expected, all GFP+ cells

    also expressed SC121 (Figure S3). At 6 weeks after transplan-

    tation, less than 10% of human cells were single positive for

    SC121 and lacked GFP expression (25 out of 372 SC121+ cells).

    In sum, perhaps due to the protracted rate of cortical inter-

    neuron maturation in vivo, transplantation studies did not result

    in mature cortical interneurons within the first 2 months of

    transplant.

    Phenotypic and Synaptic Maturationof NKX2.1+ Neurons In VitroWe have previously studied cortical interneuron development

    using a coculture system in which mouse embryonic MGE-

    derived progenitors are plated over a ‘‘feeder culture’’ com-

    posed mainly of mouse cortical pyramidal neurons and glia (Xu

    et al., 2004). To determine whether aspects of human inter-

    neuron maturation would be accelerated in this system relative

    to the xenografts, NKX2.1::GFP+ cells were collected via FACS

    at day 32 and replated onto cultures of dissociated embryonic

    mouse cortex (Figure 4A). The cortical cells were isolated

    from E13.5 mouse embryos, a stage prior to the immigration of

    the ventrally derived interneurons into the dorsal neocortex

    (Anderson et al., 1997). Thus, this coculture system mimics

    aspects of normal development with human NKX2.1+ cells

    developing in the presence of the glutamatergic cortical pyrami-

    dal neurons. Studies were performed in parallel using GFP+ cells

    derived from each of the three SHH treatment regimens (Figures

    4B–4E). Roughly 80% of the GFP+ cells from the 10–18 treated

    ‘‘MGE-like’’ cultures, versus 40% from the 2–18 ‘‘hypothalam-

    ic-like’’ culture and 15% from the 6–18 culture, expressed

    GABA. Likewise, the 6–18 culture that is enriched for telence-

    phalic (FOXG1+) cells but lacks expression of the interneuron

    precursor marker LHX6 (Figure 3D) was enriched for choline

    acetyltransferase (ChAT) neurons.

    We next determined whether GFP+ cells from day 10–18

    treated cultures (enriched for GABA neurons) undergo

    GABAergic synaptic transmission. Whole-cell patch-clamp

    electrophysiological studies (Figures 4F–4K) of identified GFP+

    cells showed spontaneous firing in both day 10–18 and day 6–

    18 treated cultures (Figures 4H and 4I). However, only the day

    10–18 group showed modulation of the firing rate of NKX2.1::

    GFP+ cells in response to the GABA type A (GABA-A) receptor

    antagonist bicuculline (Figure 4J). Because the cortical feeders

    contained very few murine interneurons (in both day 10–18

    and day 6–18 coculture conditions), these results indicate that

    hPSC-derived GABAergic neurons mediate the inhibitory synap-

    tic output. Accordingly, in the day 6–18 cultures in which

    GABAergic neurons are more rare (Figure 4C), the firing rates

    following bicuculline exposure remained unchanged (Figure 4K).

    To further analyze synaptic inputs onto the NKX2.1::GFP+

    cells from the day 10–18 protocol, we examined spontaneous

    postsynaptic currents and localization of inhibitory or excitatory

    synaptic markers. GFP+ cells cultured on the mouse cortical

    feeder for 30 days expressed high levels of vesicular GABA

    transporter (VGAT), present within the presynaptic terminal

    of GABAergic synapses. The subcellular localization of VGAT

    within GFP+ cells closely matched expression of the GABAergic

    postsynaptic marker gephyrin (Figures 5A–5E). Whole-cell

    patch-clamp analyses demonstrated that NKX2.1::GFP+ cells

    receive spontaneous inhibitory postsynaptic currents (sIPSCs;

    Figure 5F), which are reversibly blocked by the addition of the

    GABA-A receptor antagonist bicuculline. In addition, NKX2.1::

    GFP+ cells also receive excitatory inputs, as demonstrated

    by the presence of vesicular glutamate transporter 1 (VGLUT1)

    expression adjacent to GFP+ putative dendrites and colabeling

    with the postsynaptic excitatory synapse marker PSD-95

    (Figure 5G–5K). Consistent with the presence of glutamatergic

    synaptic inputs, spontaneous excitatory postsynaptic currents

    (sEPSCs; Figure 5L) were also readily detected in the NKX2.1::

    GFP+ neurons. Additional analysis of spontaneous synaptic

    activity in NKX2.1::GFP+ cells is presented in Figure S4 and

    Table S2.

    Cortical interneurons include a diverse set of neuron types

    with distinct roles in cortical development and function (Ba-

    tista-Brito and Fishell, 2009). The relatively rapid pace of matura-

    tion in our mouse cortical feeder system allowed us to assess

    coexpression of more mature interneuron markers and to

    define neurotransmitter phenotypes beyond GABA and ChAT.

    In the day 2–18 group we observed a large percentage of

    GFP+ cells expressing tyrosine hydroxylase (TH) (Figure 6A),

    consistent with the dopaminergic nature of NKX2.1-lineage

    hypothalamic neurons (Yee et al., 2009). Immunohistochemistry

    showed high percentages of neuronal nitric oxide synthase

    (nNOS) cells in GFP+ cells from the day 2–18 and Calbindin in

    those from the day 10–18 group (Figures 6B and 6C). We also

    observed expression of markers characteristic of differentiated

    cortical interneurons, including SST and PV (Figures 6D and

    6E), again mainly in the day 10–18 group. Representative

    images for each subgroup marker, coexpressed together with

    NKX2.1::GFP, are presented in Figures 6F–6J. No vasoactive

    intestinal polypeptide (VIP)- or calretinin-expressing neurons

    were observed, suggesting a lack of cells exhibiting features of

    the caudal ganglionic eminence (Xu et al., 2004). The derivation

    of NKX2.1+/PV+ neurons was particularly remarkable given the

    late developmental expression of this marker in cortical inter-

    neurons in vivo (Zecevic et al., 2011). Only a subset of the cells

    counted as GFP+/PV+ by the automated image analyses

    (Operetta high-content scanner, see Experimental Procedures)

    showed high levels of PV expression and exhibited inter-

    neuron-like morphologies (about 5% of total GFP+ cells, Fig-

    ure 6J). Our results point to the need for further optimization

    of the derivation of selective cortical interneuron subtypes.

    STEM 1331

    Cell Stem Cell

    hESC-Derived Cortical Interneurons

    Cell Stem Cell 12, 559–572, May 2, 2013 ª2013 Elsevier Inc. 565

  • However, the fact that PV+ hPSC-derived neurons could be

    generated at all suggests that the mouse cortical coculture

    strategy facilitates expression of cortical interneuron subtype

    markers.

    Given the protracted maturation of Nkx2.1-GFP+, putative

    interneuron precursors in vivo, we were surprised to obtain

    morphological, marker-based, and physiological evidence of

    interneuron-like differentiation in the coculture system. We next

    tested whether similar results can be obtained by growing the

    same Nkx2.1::GFP+ cells on cultures enriched for human

    cortical projection neurons. The human cortical feeders were

    obtained upon differentiation of hESCs (modified XLSB condi-

    tions) followed by FACS for CD44�/CD184+/CD24+ neurons(Yuan et al., 2011) (Figures S5A–S5D). The cortical identity of

    the hESC-derived neurons was confirmed by the expression

    of VGLUT1, MAP2, CTIP2, and SATB2. Only very few GFAP+

    or Nestin+ cells were present under those conditions (Figures

    S5E–S5T).

    Figure 4. Maturation of NKX2.1+ Cells into Physiologically Active Neurons

    (A) Preparation of cortical excitatory neuron cultures from E13.5 mice, onto which human NKX2.1::GFP+ cells (after FACS at day 32) are plated. The coculture

    system was critical for promoting neuronal maturation given the protracted in vivo maturation rates of NKX2.1+ cells.

    (B–E) After 30 DIV, cultures from the SHH 10–18 conditions were enriched for NKX2.1::GFP+ cells that coexpress GABA (B), quantified in (C) (mean ± SEM;

    *p < 0.05, **p < 0.01 using ANOVA followed by Scheffe test). In contrast, only the SHH 6–18 condition was enriched for NKX2.1::GFP colabeling with ChAT (D),

    quantified in (E) (mean ± SEM; *p < 0.05 using ANOVA followed by Scheffe test).

    (F and G) Spiking patterns of SHH day 10–18 (F) and SHH day 6–18 (G) neurons recorded at 28 DIV. Action potentials were initiated by protocols shown at

    the bottom.

    (H–K) Spontaneous spiking was recorded from cultures enriched for GABAergic (SHH day 10 to 18) and cholinergic (6 to 18) neurons in the absence (H and I) and

    the presence (J and K) of the GABA-A receptor antagonist bicuculline. Bicuculline had little effect on the spontaneous firing activity in the 6 to 18 condition,

    consistent with the lack of GABAergic cells from either the mouse feeder or the human NKX2.1::GFP+ cells generated by this protocol.

    The scale bars in (B) and (D) represent 50 mm.

    STEM 1331

    Cell Stem Cell

    hESC-Derived Cortical Interneurons

    566 Cell Stem Cell 12, 559–572, May 2, 2013 ª2013 Elsevier Inc.

  • Interestingly, we observed a consistent difference in electro-

    physiological maturation of NKX2.1+ cells plated on mouse

    versus human cortical feeders as shown in Table S2. Recordings

    were obtained at either 14–16 days in vitro (DIV) or 20–30 DIV

    for the mouse feeder condition and 20–30 DIV for the human

    cortical feeder condition. Analysis of the basic electrophysio-

    logical properties of both the 10– 18 and 6–18 groups plated

    on mouse feeders showed more mature characteristics at

    later time points of differentiation. There was significant hyper-

    polarization of the resting membrane potential (RMP) with time,

    as well as a decrease in the input resistance (Ri). In addition,

    the action potentials became faster and narrower, as evidenced

    by the decrease in the rise time and half-width. In contrast, GFP+

    neurons from the day 10–18 condition that were recorded

    20–30 days after plating on the human feeder layer retained

    relatively immature characteristics, comparable to the mea-

    surements seen in the younger 14–16 DIV neurons plated onto

    a mouse feeder layer. The RMP, action potential rise time, action

    potential half-width, and Ri were significantly different in the

    cortical feeder condition compared to the 10–18 neurons on

    mouse feeders recorded at the same DIV range. Sample traces

    depicting qualitative firing properties under the various condi-

    tions are shown in Figure S6.

    These preliminary findings suggest that the pace of functional

    maturation is influenced by the local environment, though the

    corresponding mechanisms remain to be determined. Although

    it is possible that species-specific maturation rates trigger

    the timing of cortical interneuron maturation (e.g., by regulating

    activity), it is also possible that there are simply more astrocytes

    in mouse versus human feeders (Johnson et al., 2007). Future

    studies should include coculture with primary cortical neurons

    of human embryonic origin for further corroboration of our

    Figure 5. NKX2.1::GFP+ GABAergic Interneurons Receive Both Excitatory and Inhibitory Synaptic Inputs

    (A–E) Collapsed z stack confocal image showing NKX2.1::GFP+, the vesicular GABA transporter (VGAT; red in A), and the postsynaptic GABAergic marker

    gephyrin (blue in A). The dendrites of this GFP+ cell that colabel with gephyrin are receiving VGAT-expressing presynaptic terminals (arrows). In addition, a GFP+

    axonal process formed a VGAT+ presynaptic terminal adjacent to a GFP-negative, gephyrin-expressing postsynaptic process (asterisk).

    (F) Whole-cell patch clamp reveals sIPSCs recorded from an NKX2.1::GFP+ neuron (SHH 10 to 18 protocol), which are reversibly blocked by the addition of the

    GABA-A receptor antagonist bicuculline.

    (G–K) Collapsed z stack confocal image showingNKX2.1::GFP, VGLUT1 (red in G), and the postsynaptic marker PSD-95 (blue in C). This GFP+ cell has dendrites

    that colabel with PSD-95 and are adjacent to VGLUT1-expressing presynaptic terminals. Note the presence of a GFP-negative cell expressing VGLUT1 (red in G,

    arrowheads), confirming the presence of excitatory glutamatergic neurons in the culture. (L) Consistent with the apparent presence of glutamatergic synaptic

    inputs, sEPSCs were detected in theNKX2.1::GFP+ neurons (10 to 18). All cells were plated on mouse cortical feeder following FACS forNKX2.1::GFP at day 32.

    See also Figure S4.

    STEM 1331

    Cell Stem Cell

    hESC-Derived Cortical Interneurons

    Cell Stem Cell 12, 559–572, May 2, 2013 ª2013 Elsevier Inc. 567

  • findings. Independent of the mechanism, our data demonstrate

    that the mouse coculture conditions enhance maturation of

    hESC-derived cortical interneurons and enable functional

    in vitro studies.

    DISCUSSION

    The use of WNT-inhibitory molecules such as DKK1 has previ-

    ously been proposed for the derivation of telencephalic and optic

    progenitors during mouse and human ESC differentiation

    (Lamba et al., 2006; Watanabe et al., 2005). We demonstrate

    that the tankyrase inhibitor XAV939 (Huang et al., 2009) can

    replace recombinant DKK1 for enhancing forebrain induction

    using the dual-SMAD inhibition protocol. The use of only three

    small molecules (XLSB protocol) renders our induction condi-

    tions well defined, cost effective, and robust across multiple

    cell lines. XLSB thus represents a general platform for generating

    forebrain lineages including cortical interneurons.

    A recent paper demonstrated the importance of SHH dosage

    during in vitro ventral forebrain specification toward GSX2+,

    NKX2.1-negative progenitors capable of generating medium

    spiny striatal neurons (Ma et al., 2012). In addition to the SHH

    dose, timing can also affect the efficiency of ventral cell-fate

    specification (Fasano et al., 2010). A remarkable feature of the

    current study is that the difference in timing of SHH activation

    is sufficient for triggering the generation of distinct ventral

    progenitors of divergent anterior-posterior identity. We have

    previously reported the derivation of hESC-derived progenitors

    expressing markers of the hypothalamic anlage (Kriks et al.,

    2011) following early activation of SHH signaling in the presence

    of FGF8. Our day 2–18 data suggest that addition of FGF8 is not

    critical for directing hypothalamic progenitor fate. Cholinergic

    neuron derivation from NKX2.1::GFP+ progenitors of the 6–18

    group should be of great value for modeling and treating disor-

    ders associated with loss of cognitive function. For example, in

    Alzheimer’s disease basal forebrain cholinergic neurons are

    among the neurons most vulnerable during early stages of the

    disease (Whitehouse et al., 1982). However, the main focus of

    the current study is the derivation of human cortical interneuron

    lineages triggered in the day 10–18 protocol.

    Our study points to several surprising differences in human

    versus mouse forebrain development, such as human-specific

    expression of FOXA2 in hESC-derived ventral forebrain progen-

    itors and in CS15 human forebrain tissue. We speculate that

    transient FOXA2 expression reflects a prolonged requirement

    for SHH signaling during human development given the obvious

    differences in brain size and extended proliferation of SHH-

    dependent progenitors. To further address this hypothesis, it

    will be important to establish a time course of FOXA2 expression

    during human ventral forebrain development and to determine

    whether FOXA2 is linked to SHH expression as commonly

    observed in other FOXA2+ CNS domains such as the floor plate

    (Placzek and Briscoe, 2005).

    Past studies have reported species differences in cortical

    interneuron development, based on the presence of ASCL1+

    cells within the second-trimester human fetal cortex (Letinic

    et al., 2002). Those data led to the hypothesis that many, and

    perhaps most, human cortical interneurons are born within the

    cortex itself (Letinic et al., 2002). In contrast, studies in mice

    have demonstrated a subcortical origin of most, if not all, cortical

    Figure 6. Neurochemical Profiling of NKX2.1::GFP+ Cells Grown on Mouse Cortical Feeders for 30 DIV

    Cells were labeled by immunofluorescence for the markers indicated, and the results were quantified in graphs (A–E: mean ± SEM; *p < 0.05, **p < 0.01, ***p <

    0.001 using ANOVA followed by Scheffe test: (*) p < 0.05 when compared directly to the 6–18 group; the p value did not reach significance in the standard Scheffe

    test: p = 0.08). Representative images of cellular labeling are shown in (F–J). In the SHH 2 to 18 condition, most of the cells colabeled with TH (A and F) and nNOS

    (B and G). In the 10 to 18 condition, many of the GFP+ cells colabeled with calbindin (Calb; C and H), somatostatin (SST; (D and I), and parvalbumin (PV; E and J),

    each of which is present in subpopulations of mature cortical interneurons in humans. All cells were plated on mouse cortical feeder following FACS for

    NKX2.1::GFP at day 32. The scale bars in (F)–(J) represent 50 mm. See also Figures S5 and S6 and Table S2.

    STEM 1331

    Cell Stem Cell

    hESC-Derived Cortical Interneurons

    568 Cell Stem Cell 12, 559–572, May 2, 2013 ª2013 Elsevier Inc.

  • interneurons (Fogarty et al., 2007; Xu et al., 2004). Although

    our study was not specifically geared toward addressing this

    issue, we found restricted NKX2.1 expression within the ventral

    forebrain in CS15 human embryo and a complete lack of

    expression in the dorsal forebrain (see also the Allen Brain Atlas,

    http://human.brain-map.org/). Furthermore, the ability of hESC-

    derived cortical interneurons to migrate in our slice culture

    assay and in the developing murine cortex in vivo suggests

    that human cortical interneuron precursors have a capacity for

    tangential migration similar to that of their mouse counterparts.

    However, a distinctive feature in hESC-derived cortical inter-

    neurons was the persistent expression of NKX2.1. Nkx2.1 is

    extinguished in mouse cortical interneuron precursors by the

    time they leave the MGE prior to entering the cortex (Marin

    et al., 2000). Our in vitro data are consistent with findings in the

    human neocortex in vivo showing the presence of postmitotic

    NKX2.1+ neurons (Fertuzinhos et al., 2009). However, our data

    do not rule out the possibility that a subset of our human

    ESC-derived cells correspond to striatal interneurons, given

    that those share markers of cortical interneurons in the mouse

    while retaining NKX2.1 expression (Marin et al., 2000).

    Current paradigms of modeling human psychiatric disease

    such as schizophrenia feature the use of patient-specific iPSC-

    derived neurons (Brennand et al., 2011; Marchetto et al., 2010;

    Pasxca et al., 2011). However, those published studies were per-formed in mixed neural cultures of unclear neuronal subtype

    identity and with limited characterization of subtype-specific

    synaptic and functional properties. There is a convergence of

    postmortem findings that attempt to link genetic defects to psy-

    chiatric disorders, such as the interneuron-associated Erbb4

    receptor in schizophrenia (Fazzari et al., 2010). Therefore, it is

    essential to have access to purified populations of mature

    cortical interneurons for modeling human disease. Our results

    demonstrate that the highly efficient derivation of cortical

    interneurons is possible following timed exposure to develop-

    mental cues.

    We report that putative hESC-derived GABAergic inter-

    neurons receive synaptic inputs fromboth other human interneu-

    rons and from excitatory mouse projection neurons. In addition,

    cells with neurochemical properties of cortical interneurons

    adopt fairly mature physiological properties within 30 days of

    coculture. Future studies will be required to address the mecha-

    nisms of accelerated in vitro maturation of the NKX2.1::GFP+

    neurons on mouse cortical cultures and to determine whether

    species-specific timing factors are involved, as suggested

    from our preliminary studies using hESC-derived cortical

    feeders. However, our data demonstrate that the proposed

    culture system can yield synaptically active cortical interneurons

    in vitro, a key prerequisite for modeling cortical interneuron

    pathologies in psychiatric disorders such as schizophrenia

    and autism. The generation of hESC-derived PV-expressing

    neurons and the presence of relatively rapid spiking, nonaccom-

    modating neurons in these cultures is of particular interest given

    the implications of PV interneuron dysfunction in schizophrenia

    (Lewis et al., 2005). Fast-spiking PV+ cortical interneurons are

    observed late during primate prenatal development and

    continue their maturation into early adulthood (Anderson et al.,

    1995; Insel, 2010). Given the important role of PV+ neurons

    under various pathological conditions our data suggest that

    disease modeling of such states may be feasible using our

    differentiation strategies. Another key for the future will be the

    development of protocols that allow the enriched generation of

    specific cortical interneuron subgroups, such as sSST+ versus

    PV+ cells. Our previous results in MGE progenitors indicate

    that this decision may yet again be under the control of SHH

    signaling, with higher levels promoting SST+ and lower levels

    promoting PV+ neurons (Xu et al., 2010).

    Finally, future studies will be required to address the full in vivo

    potential of human in vitro-derived cortical interneurons for ap-

    plications in regenerative medicine. Our current study illustrates

    the remarkable migratory potential of the hESC-derived cortical

    interneurons upon transplantation into the neonatal mouse

    cortex. Transplantation studies into several adult CNS models

    of disease will be of particular interest given the potential use

    of cortical interneuron grafts in modulating pathological seizure

    activity (Baraban et al., 2009), treating aspects of Parkinson’s

    disease (Martı́nez-Cerdeño et al., 2010), and inducing learning

    and plasticity within the postnatal brain (Southwell et al., 2010).

    One specific challenge for such transplantation studies is the

    protracted maturation of grafted hPSC-derived cortical inter-

    neuron precursors in vivo. Preliminary longer-term transplanta-

    tion studies into the adult mouse cortex confirm their continued

    slow maturation rate, with NKX2.1+ putative interneuron pre-

    cursors retaining immature growth cones and showing limited

    integration at 3 months post grafting (data not shown). Those

    data are in contrast to our in vitro coculture work, wherein rapid

    functional integration and phenotypic maturation was readily

    achieved. In sum, this study provides a framework for the gener-

    ation of distinct ventral prosencephalic neuron types that can

    be used to study various aspects of development and serves

    as a powerful in vitro platform for studying the dysfunction of

    specific neuron types implicated in a myriad of neuropsychiatric

    diseases.

    EXPERIMENTAL PROCEDURES

    hESC Culture and Neural Differentiation

    hESCs (WA-09, HES-3 (NKX2.1::GFP), and iPSCs (C72 and SeV6) were

    maintained on mouse embryonic fibroblasts and dissociated with accutase

    (Innovative Cell Technologies) for differentiation or dispase for passaging

    (Chambers et al., 2009). Differentiation media were described previously

    (Kriks et al., 2011): knockout serum replacer (KSR) and N2 medium for neural

    induction, and neurobasal medium + B27 (GIBCO) and N2 supplements

    (Invitrogen) for neuronal differentiation. Small-molecule compounds were as

    follows: XAV939 (2 mM; Stemgent;), LDN193189 (100 nM; Stemgent),

    SB431542 (10 mM; Tocris Bioscience), purmorphamine (1–2 mM; Calbiochem),

    ascorbic acid (200 mM), and dibutyryl-cyclic AMP (200 mM; both from

    Sigma-Aldrich). Recombinant growth factors were as follows: SHH (C25II;

    50–500 ng/ml), Noggin (125 ng/ml), DKK1 (250 ng/ml), BDNF (10 ng/ml), all

    from R&D Systems, and FGF2 (5–10 ng/ml; Promega).

    E13.5 Slice Migration Analysis

    Coronal telencephalic slices (250 mm) were prepared as described previously

    (Anderson et al., 2001) andmaintained in neurobasal medium/B27 with 5 ng/ml

    FGF2 (Promega). hESC-derived NKX2.1::GFP+ cells were isolated by FACS

    and carefully injected (�5,000 cells) into the periventricular region of theMGE using an oocyte microinjector (Nanoject II; Drummond Scientific). After

    2–6 days, cultures were fixed and immunostained. For cell counting, zone 1

    was defined as the region from the ventricular zone of the lateral ganglionic

    eminence extending to the pallial-subpallial border, and zone 2 was defined

    as the region from the lateral cortex through the neocortex to the cortical hem.

    STEM 1331

    Cell Stem Cell

    hESC-Derived Cortical Interneurons

    Cell Stem Cell 12, 559–572, May 2, 2013 ª2013 Elsevier Inc. 569

    http://human.brain-map.org/

  • Mouse and Human Cortical Culture Preparation

    The cortical feeder cultures were prepared by dissociating dorsal cortices

    from 250mm coronal slices of E13.5 mouse brain and cultured as described

    previously (Xu et al., 2010). For human cortical feeder preparation, cells were

    subjected to XLSB-mediated neural induction in the presence of cylcopamine

    and purified for CD24+/CD44�/CD184+ at day 32 (see Supplemental Experi-mental Procedures for details).

    Gene-Expression Profiling

    Total RNA was isolated at day 18 of differentiation from FACS-sorted

    NKX2.1::GFP+ cells from three varying differentiation conditions and a

    ‘‘No SHH’’ condition that did not contain any GFP-expressing cells (Trizol,

    Sigma-Aldrich). Samples in triplicate for each group were processed for

    Illumina bead arrays (Illumina HT-12) by the Memorial Sloan-Kettering

    Cancer Center genomics core facility according to the specifications of the

    manufacturer.

    Immunocytochemical Analyses

    Cultures were fixed in 4% paraformaldehyde and processed for immunocyto-

    chemistry. Secondary antibodies were species-specific Alexa-dye conjugates

    (Invitrogen). The use of the primary antibodies is summarized in Table S3.

    For the automated quantification (Operetta, PerkinElmer), see Supplemental

    Experimental Procedures.

    Transplantation Studies

    Transplantation into the neonatal cortex of NOD-SCID IL2RG�/� mice wasconducted as described previously (Wonders et al., 2008). In brief, �50 3103 cells were injected at following coordinates from bregma (2.0mm A,

    2.5mm L, 1.0mm D), targeting cortical layers 3–6. All animal experiments

    were carried out in accordance with institutional and National Institutes of

    Health guidelines.

    Human Tissue Analysis

    Tissue derived from aborted human fetuses was obtained from the Depart-

    ment of Gynecology, University Bern, Switzerland. The study was approved

    by the ethics committee of the Medical Faculty of the University of Bern,

    Switzerland, and the ethics committee of the state of Bern, Switzerland

    (no. 52/91, 71/94, and 188/95). Written consent was given by the women

    seeking abortion.

    Statistical Analysis

    For all experiments, analysis was derived from at least three independent ex-

    periments. Statistical analysis was performed using ANOVA followed by a

    Scheffe test to for significance among individual groups (SYSTAT v.13).

    FACS

    All cells were dissociated using accutase (Stem Cell Technologies) for 30 min,

    then resuspended in neurobasal medium/B27 with Y27632 (10 mM; Tocris

    Bioscience) and sorted on a BD FACSAria II (gating for side scatter [SSC],

    forward scatter [FSC], and DAPI exclusion). For postsort analysis, data was

    processed using FlowJo (Tree Star) software.

    Electrophysiology

    Whole-cell voltage and current clamp recordings were performed as des-

    cribed previously (Ying et al., 2007). Details are provided in Supplemental

    Experimental Procedures.

    ACCESSION NUMBERS

    The raw data reported in this paper have been deposited in the GEO under

    accession number GSE46098.

    SUPPLEMENTAL INFORMATION

    Supplemental Information includes six figures, three tables, and Supplemental

    Experimental Procedures and can be found with this article online at http://dx.

    doi.org/10.1016/j.stem.2013.04.008.

    ACKNOWLEDGMENTS

    Wewould like to thank J. Hendrikx (Sloan-Kettering Institute [SKI] FlowCytom-

    etry Core), A. Viale (SKI Genomics Core laboratory), M. Leversha (Molecular

    Cytogenetics Core), and E. Tu and M. Tomishima (SKI Stem Cell Facility) for

    excellent technical support. We further thank the Department of Gynecology

    (directors: M. Mueller and D. Surbek), University of Bern, Switzerland for

    providing us with the human fetal tissue. Studies using the NKX2.1::GFP

    hESC line were supported by grants from NYSTEM (S.A.), the European Com-

    mission project NeuroStemcell (L.S.), and the C.V. Starr Foundation (S.A. and

    L.S.). Experiments using hESC line WA-09 or iPSC lines were supported by

    NIMH grants 2RO1 MH066912 (S.A.) and 1RC1MH089690 (S.A. and L.S.)

    and by the Swiss National Science Foundation grant no. 31003A_135565

    (H.R.W.).

    Received: May 8, 2012

    Revised: March 2, 2013

    Accepted: April 12, 2013

    Published: May 2, 2013

    REFERENCES

    Anderson, S.A., Classey, J.D., Condé, F., Lund, J.S., and Lewis, D.A. (1995).

    Synchronous development of pyramidal neuron dendritic spines and parval-

    bumin-immunoreactive chandelier neuron axon terminals in layer III of monkey

    prefrontal cortex. Neuroscience 67, 7–22.

    Anderson, S.A., Eisenstat, D.D., Shi, L., and Rubenstein, J.L. (1997).

    Interneuron migration from basal forebrain to neocortex: dependence on Dlx

    genes. Science 278, 474–476.

    Anderson, S.A., Marı́n, O., Horn, C., Jennings, K., and Rubenstein, J.L. (2001).

    Distinct cortical migrations from the medial and lateral ganglionic eminences.

    Development 128, 353–363.

    Baraban, S.C., Southwell, D.G., Estrada, R.C., Jones, D.L., Sebe, J.Y., Alfaro-

    Cervello, C., Garcı́a-Verdugo, J.M., Rubenstein, J.L., and Alvarez-Buylla, A.

    (2009). Reduction of seizures by transplantation of cortical GABAergic inter-

    neuron precursors into Kv1.1 mutant mice. Proc. Natl. Acad. Sci. USA 106,

    15472–15477.

    Batista-Brito, R., and Fishell, G. (2009). The developmental integration of

    cortical interneurons into a functional network. Curr. Top. Dev. Biol. 87,

    81–118.

    Brennand, K.J., Simone, A., Jou, J., Gelboin-Burkhart, C., Tran, N., Sangar, S.,

    Li, Y., Mu, Y., Chen, G., Yu, D., et al. (2011). Modelling schizophrenia using

    human induced pluripotent stem cells. Nature 473, 221–225.

    Chambers, S.M., Fasano, C.A., Papapetrou, E.P., Tomishima, M., Sadelain,

    M., and Studer, L. (2009). Highly efficient neural conversion of human ES

    and iPS cells by dual inhibition of SMAD signaling. Nat. Biotechnol. 27,

    275–280.

    Dimos, J.T., Rodolfa, K.T., Niakan, K.K., Weisenthal, L.M., Mitsumoto, H.,

    Chung, W., Croft, G.F., Saphier, G., Leibel, R., Goland, R., et al. (2008).

    Induced pluripotent stem cells generated from patients with ALS can be

    differentiated into motor neurons. Science 321, 1218–1221.

    Ebert, A.D., Yu, J., Rose, F.F., Jr., Mattis, V.B., Lorson, C.L., Thomson, J.A.,

    and Svendsen, C.N. (2009). Induced pluripotent stem cells from a spinal

    muscular atrophy patient. Nature 457, 277–280.

    Eiraku, M., Watanabe, K., Matsuo-Takasaki, M., Kawada, M., Yonemura, S.,

    Matsumura, M., Wataya, T., Nishiyama, A., Muguruma, K., and Sasai, Y.

    (2008). Self-organized formation of polarized cortical tissues from ESCs and

    its active manipulation by extrinsic signals. Cell Stem Cell 3, 519–532.

    Espuny-Camacho, I., Michelsen, K.A., Gall, D., Linaro, D., Hasche, A.,

    Bonnefont, J., Bali, C., Orduz, D., Bilheu, A., Herpoel, A., et al. (2013).

    Pyramidal neurons derived from human pluripotent stem cells integrate

    efficiently into mouse brain circuits in vivo. Neuron 77, 440–456.

    Fasano, C.A., Chambers, S.M., Lee, G., Tomishima, M.J., and Studer, L.

    (2010). Efficient derivation of functional floor plate tissue from human embry-

    onic stem cells. Cell Stem Cell 6, 336–347.

    STEM 1331

    Cell Stem Cell

    hESC-Derived Cortical Interneurons

    570 Cell Stem Cell 12, 559–572, May 2, 2013 ª2013 Elsevier Inc.

    http://dx.doi.org/10.1016/j.stem.2013.04.008http://dx.doi.org/10.1016/j.stem.2013.04.008

  • Fazzari, P., Paternain, A.V., Valiente, M., Pla, R., Luján, R., Lloyd, K., Lerma, J.,

    Marı́n, O., and Rico, B. (2010). Control of cortical GABA circuitry development

    by Nrg1 and ErbB4 signalling. Nature 464, 1376–1380.

    Fertuzinhos, S., Krsnik, Z., Kawasawa, Y.I., Rasin, M.R., Kwan, K.Y., Chen,

    J.G., Judas, M., Hayashi, M., and Sestan, N. (2009). Selective depletion of

    molecularly defined cortical interneurons in human holoprosencephaly with

    severe striatal hypoplasia. Cereb. Cortex 19, 2196–2207.

    Flames, N., Pla, R., Gelman, D.M., Rubenstein, J.L., Puelles, L., and Marı́n, O.

    (2007). Delineation of multiple subpallial progenitor domains by the combina-

    torial expression of transcriptional codes. J. Neurosci. 27, 9682–9695.

    Fogarty, M., Grist, M., Gelman, D., Marı́n, O., Pachnis, V., and Kessaris, N.

    (2007). Spatial genetic patterning of the embryonic neuroepithelium generates

    GABAergic interneuron diversity in the adult cortex. J. Neurosci. 27, 10935–

    10946.

    Gaspard, N., Bouschet, T., Hourez, R., Dimidschstein, J., Naeije, G., van den

    Ameele, J., Espuny-Camacho, I., Herpoel, A., Passante, L., Schiffmann, S.N.,

    et al. (2008). An intrinsic mechanism of corticogenesis from embryonic stem

    cells. Nature 455, 351–357.

    Goulburn, A.L., Alden, D., Davis, R.P., Micallef, S.J., Ng, E.S., Yu, Q.C., Lim,

    S.M., Soh, C.L., Elliott, D.A., Hatzistavrou, T., et al. (2011). A targeted

    NKX2.1 human embryonic stem cell reporter line enables identification of

    human basal forebrain derivatives. Stem Cells 29, 462–473.

    Huang, S.M., Mishina, Y.M., Liu, S., Cheung, A., Stegmeier, F., Michaud,

    G.A., Charlat, O., Wiellette, E., Zhang, Y., Wiessner, S., et al. (2009).

    Tankyrase inhibition stabilizes axin and antagonizes Wnt signalling. Nature

    461, 614–620.

    Insel, T.R. (2010). Rethinking schizophrenia. Nature 468, 187–193.

    Johnson, M.A., Weick, J.P., Pearce, R.A., and Zhang, S.C. (2007). Functional

    neural development from human embryonic stem cells: accelerated synaptic

    activity via astrocyte coculture. J. Neurosci. 27, 3069–3077.

    Kelly, S., Bliss, T.M., Shah, A.K., Sun, G.H., Ma, M., Foo, W.C., Masel, J.,

    Yenari, M.A., Weissman, I.L., Uchida, N., et al. (2004). Transplanted human

    fetal neural stem cells survive, migrate, and differentiate in ischemic rat

    cerebral cortex. Proc. Natl. Acad. Sci. USA 101, 11839–11844.

    Kerwin, J., Yang, Y., Merchan, P., Sarma, S., Thompson, J., Wang, X.,

    Sandoval, J., Puelles, L., Baldock, R., and Lindsay, S. (2010). The HUDSEN

    Atlas: a three-dimensional (3D) spatial framework for studying gene expression

    in the developing human brain. J. Anat. 217, 289–299.

    Kim, H., Lee, G., Ganat, Y., Papapetrou, E.P., Lipchina, I., Socci, N.D.,

    Sadelain, M., and Studer, L. (2011). miR-371-3 expression predicts neural

    differentiation propensity in human pluripotent stem cells. Cell Stem Cell 8,

    695–706.

    Kriks, S., Shim, J.W., Piao, J., Ganat, Y.M., Wakeman, D.R., Xie, Z., Carrillo-

    Reid, L., Auyeung, G., Antonacci, C., Buch, A., et al. (2011). Dopamine neurons

    derived from human ES cells efficiently engraft in animal models of Parkinson’s

    disease. Nature 480, 547–551.

    Lamba, D.A., Karl, M.O., Ware, C.B., and Reh, T.A. (2006). Efficient generation

    of retinal progenitor cells from human embryonic stem cells. Proc. Natl. Acad.

    Sci. USA 103, 12769–12774.

    Letinic, K., Zoncu, R., and Rakic, P. (2002). Origin of GABAergic neurons in the

    human neocortex. Nature 417, 645–649.

    Lewis, D.A., Hashimoto, T., and Volk, D.W. (2005). Cortical inhibitory neurons

    and schizophrenia. Nat. Rev. Neurosci. 6, 312–324.

    Ma, L., Hu, B., Liu, Y., Vermilyea, S.C., Liu, H., Gao, L., Sun, Y., Zhang, X., and

    Zhang, S.C. (2012). Human embryonic stem cell-derived GABA neurons

    correct locomotion deficits in quinolinic acid-lesioned mice. Cell Stem Cell

    10, 455–464.

    Marchetto, M.C., Carromeu, C., Acab, A., Yu, D., Yeo, G.W., Mu, Y., Chen, G.,

    Gage, F.H., and Muotri, A.R. (2010). A model for neural development and

    treatment of Rett syndrome using human induced pluripotent stem cells.

    Cell 143, 527–539.

    Marin, O., Anderson, S.A., and Rubenstein, J.L. (2000). Origin and molecular

    specification of striatal interneurons. J. Neurosci. 20, 6063–6076.

    Maroof, A.M., Brown, K., Shi, S.H., Studer, L., and Anderson, S.A. (2010).

    Prospective isolation of cortical interneuron precursors from mouse embry-

    onic stem cells. J. Neurosci. 30, 4667–4675.

    Martı́nez-Cerdeño, V., Noctor, S.C., Espinosa, A., Ariza, J., Parker, P., Orasji, S.,

    Daadi, M.M., Bankiewicz, K., Alvarez-Buylla, A., and Kriegstein, A.R. (2010).

    Embryonic MGE precursor cells grafted into adult rat striatum integrate and

    amelioratemotor symptoms in6-OHDA-lesioned rats.Cell StemCell6, 238–250.

    Papapetrou, E.P., Tomishima, M.J., Chambers, S.M., Mica, Y., Reed, E.,

    Menon, J., Tabar, V., Mo, Q., Studer, L., and Sadelain, M. (2009).

    Stoichiometric and temporal requirements of Oct4, Sox2, Klf4, and c-Myc

    expression for efficient human iPSC induction and differentiation. Proc. Natl.

    Acad. Sci. USA 106, 12759–12764.

    Pasxca, S.P., Portmann, T., Voineagu, I., Yazawa, M., Shcheglovitov, A., Pasxca,A.M., Cord, B., Palmer, T.D., Chikahisa, S., Nishino, S., et al. (2011). Using

    iPSC-derived neurons to uncover cellular phenotypes associated with

    Timothy syndrome. Nat. Med. 17, 1657–1662.

    Peñagarikano, O., Abrahams, B.S., Herman, E.I., Winden, K.D., Gdalyahu, A.,

    Dong, H., Sonnenblick, L.I., Gruver, R., Almajano, J., Bragin, A., et al. (2011).

    Absence of CNTNAP2 leads to epilepsy, neuronal migration abnormalities,

    and core autism-related deficits. Cell 147, 235–246.

    Placzek, M., and Briscoe, J. (2005). The floor plate: multiple cells, multiple

    signals. Nat. Rev. Neurosci. 6, 230–240.

    Rakic, S., and Zecevic, N. (2003). Emerging complexity of layer I in human

    cerebral cortex. Cereb. Cortex 13, 1072–1083.

    Shi, Y., Kirwan, P., Smith, J., Robinson, H.P., and Livesey, F.J. (2012). Human

    cerebral cortex development from pluripotent stem cells to functional excit-

    atory synapses. Nat. Neurosci. 15, 477–486, S1.

    Southwell, D.G., Froemke, R.C., Alvarez-Buylla, A., Stryker, M.P., and Gandhi,

    S.P. (2010). Cortical plasticity induced by inhibitory neuron transplantation.

    Science 327, 1145–1148.

    Sussel, L., Marin, O., Kimura, S., and Rubenstein, J.L. (1999). Loss of Nkx2.1

    homeobox gene function results in a ventral to dorsal molecular respecification

    within the basal telencephalon: evidence for a transformation of the pallidum

    into the striatum. Development 126, 3359–3370.

    Tekki-Kessaris, N., Woodruff, R., Hall, A.C., Gaffield, W., Kimura, S., Stiles,

    C.D., Rowitch, D.H., and Richardson, W.D. (2001). Hedgehog-dependent

    oligodendrocyte lineage specification in the telencephalon. Development

    128, 2545–2554.

    Watanabe, K., Kamiya, D., Nishiyama, A., Katayama, T., Nozaki, S., Kawasaki,

    H., Watanabe, Y., Mizuseki, K., and Sasai, Y. (2005). Directed differentiation of

    telencephalic precursors fromembryonic stemcells. Nat.Neurosci.8, 288–296.

    Whitehouse, P.J., Price, D.L., Struble, R.G., Clark, A.W., Coyle, J.T., and

    Delon, M.R. (1982). Alzheimer’s disease and senile dementia: loss of neurons

    in the basal forebrain. Science 215, 1237–1239.

    Wichterle, H., Garcia-Verdugo, J.M., Herrera, D.G., and Alvarez-Buylla, A.

    (1999). Young neurons from medial ganglionic eminence disperse in adult

    and embryonic brain. Nat. Neurosci. 2, 461–466.

    Wonders, C.P., Taylor, L., Welagen, J., Mbata, I.C., Xiang, J.Z., and Anderson,

    S.A. (2008). A spatial bias for the origins of interneuron subgroups within the

    medial ganglionic eminence. Dev. Biol. 314, 127–136.

    Wu, H., Xu, J., Pang, Z.P., Ge, W., Kim, K.J., Blanchi, B., Chen, C., Südhof,

    T.C., and Sun, Y.E. (2007). Integrative genomic and functional analyses reveal

    neuronal subtype differentiation bias in human embryonic stem cell lines. Proc.

    Natl. Acad. Sci. USA 104, 13821–13826.

    Xu, Q., Cobos, I., De La Cruz, E., Rubenstein, J.L., and Anderson, S.A. (2004).

    Origins of cortical interneuron subtypes. J. Neurosci. 24, 2612–2622.

    Xu, Q., Guo, L., Moore, H., Waclaw, R.R., Campbell, K., and Anderson, S.A.

    (2010). Sonic hedgehog signaling confers ventral telencephalic progenitors

    with distinct cortical interneuron fates. Neuron 65, 328–340.

    Yee, C.L., Wang, Y., Anderson, S., Ekker, M., and Rubenstein, J.L. (2009).

    Arcuate nucleus expression of NKX2.1 and DLX and lineages expressing

    these transcription factors in neuropeptide Y(+), proopiomelanocortin(+), and

    tyrosine hydroxylase(+) neurons in neonatal and adult mice. J. Comp.

    Neurol. 517, 37–50.

    STEM 1331

    Cell Stem Cell

    hESC-Derived Cortical Interneurons

    Cell Stem Cell 12, 559–572, May 2, 2013 ª2013 Elsevier Inc. 571

  • Ying, S.W., Jia, F., Abbas, S.Y., Hofmann, F., Ludwig, A., and Goldstein, P.A.

    (2007). Dendritic HCN2 channels constrain glutamate-driven excitability in

    reticular thalamic neurons. J. Neurosci. 27, 8719–8732.

    Yu, X., and Zecevic, N. (2011). Dorsal radial glial cells have the potential to

    generate cortical interneurons in human but not in mouse brain. J. Neurosci.

    31, 2413–2420.

    Yuan, S.H., Martin, J., Elia, J., Flippin, J., Paramban, R.I., Hefferan, M.P., Vidal,

    J.G., Mu, Y., Killian, R.L., Israel, M.A., et al. (2011). Cell-surface marker signa-

    tures for the isolation of neural stem cells, glia and neurons derived from hu-

    man pluripotent stem cells. PLoS ONE 6, e17540.

    Zecevic, N., Hu, F., and Jakovcevski, I. (2011). Interneurons in the developing

    human neocortex. Dev. Neurobiol. 71, 18–33.

    STEM 1331

    Cell Stem Cell

    hESC-Derived Cortical Interneurons

    572 Cell Stem Cell 12, 559–572, May 2, 2013 ª2013 Elsevier Inc.

  • Cell Stem Cell, Volume 12

    Supplemental Information

    Directed Differentiation and Functional

    Maturation of Cortical Interneurons

    from Human Embryonic Stem Cells

    Asif M. Maroof, Sotirios Keros, Jennifer A. Tyson, Shui-Wang Ying, Yosif M. Ganat, Florian T. Merkle,

    Becky Liu, Adam Goulburn, Edouard G. Stanley, Andrew G. Elefanty, Hans Ruedi Widmer Kevin Eggan,

    Peter A. Goldstein, Stewart A. Anderson, and Lorenz Studer

  • Figure S1. FOXA2 Is Transiently Expressed in the hPSC-Derived NKX2.1+ Lineages and in the Human Ganglionic Eminence, Related to Figure 2 A) Time course of FOXA2 expression after sorting NKX2.1::GFP+ cells at day 18, followed by culture in the absence of SHH for two additional weeks. Data represent mean ± SEM. B) Representative immunocytochemistry image of day 10-18 treated cultures stained for FOXG1, FOXA2 and GFP (day 18). C) Embryonic day 11.5 coronal section of a developing mouse brain showing the expression of FOXG1 (red), NKX2.1 (green), and FOXA2 (D'Amato et al.). There is no expression of FOXG1 in the diencephalon (Di) and no FOXA2 expression in the telencephalon (Tel). NKX2.1 is expressed in the medial ganglionic eminence (MGE) and the

  • hypothalamus (HYPO), but is not expressed in the floor plate (FP) and does not co-label with FOXA2. D) Carnegie stage 15 (CS15; approximately 5.5 gestational weeks) human embryo sectioned for immunohistochemical analysis. E,F) Adjacent coronal (oblique) hemisections of the human CS15 prosencephalon demonstrating the expression of FOXA2 and NKX2.1. While FOXA2 appears to be specifically expressed in the ganglionic eminence (E) NKX2.1 is expressed in various regions of the ventral prosencephalon (F).

    This figure demonstrates differences in FOXA2 expression between human and mouse, suggestive of an alternative mechanism to generate NKX2.1 progenitors in the ventral forebrain. However, FOXA2 expression in these NKX2.1 cells may be transient as shown in vitro (A) and as suggested by in situ hybridization analyses of later stage human embryos (Allen Brain Atlas, http://human.brain-map.org/).

    http://human.brain-map.org/

  • Figure S2. NKX2.1::GFP+ Neuronal Precursors Distribute within the Adult Mouse Cortex, Related to Figure 3 NKX2.1::GFP expressing cells were sorted at day 32 of differentiation and grafted into adult IL2RG immunocompromised mice. While the 2 to 18 (A) and the 6 to 18 (B) condition produced regions of GFP expression in the cortex after 30 days, the GFP cell bodies did not leave the graft core. Only in the 10 to 18 condition (C) did GFP expressing cells leave the graft core and distribute throughout the cortical parenchyma. In addition the GFP expressing cells from the 10 to 18 condition distribute throughout the cortical layers and display neuronal morphologies (D).

  • Figure S3. Slow Pace of Differentiation by Nkx2.1::GFP+ Cells Following Transplantation into Neonatal Mouse Neocortex, Related to Figure 3 Shown are examples of transplants of the Nkx2.1::GFP line, differentiated to day 32 with the “MGE-like protocol (SHH 10-18), then subjected to FACS for GFP, transplanted into the neonatal neocortex of genetically immune-compromised mice, and evaluated in fixed sections after 6 weeks (see methods). A) Low-magnification view of GFP immunofluorescence labeling. Scale bar=100µm. The boxed area shows three cells at higher magnification in panel (A1). Note that even after 6 weeks of differentiation in vivo most of these cells continue to have relatively undifferentiated appearance with bipolar or unipolar processes, often tipped by growth cones

  • (arrow). Scale bar=20µm. B) shows immunofluorescence for the human specific cell marker SC121 (B,B1,B3) and GFP (B,B1,B2). All GFP+ cells express SC121, and most but not quite all SC121 cells at this age express GFP (arrows; ). As we see similarly low percentages of SC121+/GFP(-) cells two weeks after transplantation (not shown), it is not clear whether these cells represent downregulation of Nkx2.1 and GFP or represent GFP(-) cells that came through the FACS. B) Scale bar=35µm, B1-3 scale bar=40µm. C, C1) shows the same view of a section with immunofluorescence for GABA (red) and GFP. Two strongly co-labeled cells are indicated by the arrows. As GABA immunodetection penetrates poorly into sections, co-labeling was quantified by selectively counting only those GFP+ cells with cell bodies located within 5 microns of the cut edges of the section. By this approach, 46 of 51 cells (90.2%) from two 6-week transplants were strongly co-labeled. Scale bar=40µm. D) shows immunofluorescence for GFP, Nkx2.1 (red) and the marker of neuronal precursors DCX (blue; pseudocolored Cy5 signal). All the GFP+ cells have Nkx2.1-expressing nuclei (arrows), and also co-label for DCX. Scale bar=40µm.

  • Figure S4. Electrophysiological Recordings from NKX2.1::GFP+ Neurons of the Day 10–18 Group, Related to Figure 5 A) Examples of spontaneous and miniature IPSCs recorded from a day 10-18 group neuron. B) Overlay of the averaged sIPSCs (gray; n=150) and mIPSCs (black; n=56). Neurons were recorded on day 26-30 on mouse cortical feeders. C) Cumulative probability histogram of IPSCs (gray) and mIPSCs (black). D) Grouped data (mean ± SEM) comparing the frequency of sIPSPs (gray; n=\ markers (MAP2), astrocytic markers (GFAP), GABA, and cortical markers (SATB2, TBR1) in human ESC derived cortical precursors at day 30 of differentiation. All cells were plated on mouse or human ESC-derived cortical feeder following FACS for NKX2.1::GFP at day 32.

  • Figure S5

    Figure S5. Phenotypic Characterization of Human ESC-Derived versus Primary Mouse Cortical Feeder Cells, Related to Figure 6

  • A-D) FACS-mediated isolation of cortical neuron precursors based on expression of CD24 and lack of expression of CD44 and CXCR4 (CD184). E-L) Immunocytochemical analysis for expression of neuronal markers (MAP2), astrocytic markers (GFAP), GABA, and cortical markers (SATB2, TBR1) in hESC-derived cortical precursors at day 30 of differentiation. M-T) Immunocytochemical analysis for expression of neuronal markers (MAP2), astrocytic markers (GFAP), GABA, and cortical markers (SATB2, TBR1) in mouse E13.5 cortical cultures. Overall, we observed comparable marker expression between primary mouse and hPSC derived cortical feeders.

  • Figure S6. Representative Current-Clamp Recordings from NKX2.1+ Cortical Interneurons, Related to Figure 6 A-C) Representative tracings from GFP-positive neurons from the day 10-18 treated group recorded on mouse feeders (A,B) or human ES-derived neurons enriched for putative cerebral cortical projection neurons (C). Depolarized voltage responses were initiated using protocols shown below each set of traces. The holding potential was –60 mV for all recordings. All cells were plated on mouse or human ESC-derived cortical feeder following FACS for NKX2.1::GFP at day 32. These data are related to Figure 6.

  • Table S1. Fold Changes and Significance of Differentially Expressed Transcripts from Figure 2I, Related to Figure 2

    day 10-18 versus no SHH

    day 10-18 versus day 2-18

    day 10-18 versus day 6-18

    Probeset F-Change p-value

    Probeset

    F-Change p-value

    Probeset

    F-Change p-value

    SIX6 7.88 6.35E-09

    ZIC2 9.25 1.56E-09

    SOX21 3.14 3.37E-07

    NKX2-1 6.92 3.40E-10

    FOXG1 6.70 6.70E-05

    SP8 2.81 1.40E-08

    LIPA 6.47 8.53E-09

    PCSK5 3.39 3.35E-05

    PCDH19 2.65 2.29E-06

    IFIT1 5.73 5.21E-09

    ZIC3 3.19 2.25E-04

    FOXG1 2.39 6.69E-05

    SNORD3A 5.37 4.69E-07

    NDP 3.12 2.76E-04

    OTX2 2.38 5.20E-08

    OLIG1 4.70 2.51E-07

    WNT5B 2.66 2.90E-04

    C14ORF23 2.37 5.84E-06

    NKX2-2 4.62 8.69E-09

    PLCH1 2.56 1.52E-05

    SHISA2 2.36 6.38E-05

    PTCH1 3.96 2.07E-06

    KCNK12 2.54 2.74E-04

    PCSK5 2.19 1.67E-05

    OLIG2 3.90 1.15E-05

    CDH11 2.44 4.13E-05

    MYCN 2.19 1.14E-05

    NKX6-2 3.74 2.12E-07

    MYCN 2.41 2.18E-06

    FRZB 2.12 6.43E-05

    HES1 3.46 2.42E-09

    REC8 2.08 1.04E-04

    DOCK11 2.11 2.29E-04

    ASCL1 3.00 1.42E-06

    HCRT 2.08 3.20E-04

    FOXO1 2.08 1.74E-04

    FOXA2 2.63 6.86E-07

    LMO2 2.06 2.05E-05

    ALCAM 2.08 3.61E-04

    SEMA3A -2.00 4.75E-04

    KLF6 -2.21 4.29E-06

    DCX -2.11 2.73E-04

    SEMA3C -2.14 4.49E-06

    EFNB2 -2.24 2.35E-04

    CHRNA3 -2.69 2.69E-06

    MEIS1 -2.14 1.05E-04

    APOE -2.41 8.27E-05

    GAP43 -2.96 5.08E-06

    DLX5 -2.22 2.04E-05

    RAX -2.57 2.70E-05

    CTGF -3.06 1.94E-05

    NR2F2 -2.64 2.63E-06

    LAMB1 -2.64 1.10E-05

    ANKRD1 -3.60 1.35E-06

    EVI1 -2.68 3.30E-06

    GAP43 -2.76 2.45E-04

    POU3F1 -3.70 3.79E-07

    PDGFC -2.80 4.38E-06

    CTGF -4.43 9.83E-07

    SYT4 -4.35 4.97E-08

    DCT -3.52 9.43E-11

    ANKRD1 -4.46 1.28E-07

    GRP -4.59 4.02E-08

    PAX6 -4.64 2.32E-08

    HEY1 -4.48 2.55E-04

    EMP1 -4.68 5.10E-06

    GAS1 -5.43 9.25E-08

    LEFTY2 -4.67 2.27E-06

    NKX2-2 -4.74 7.66E-09

    CCL2 -5.85 7.78E-08

    EPCAM -4.93 6.11E-05

    STMN2 -5.50 2.16E-09

    EMX2 -6.08 1.00E-10

    COL1A1 -4.98 4.33E-08

    VGF -5.86 1.97E-09

    ALDH1A1 -8.99 9.36E-11

    EMP1 -6.14 8.08E-06

    MMP7 -6.62 4.37E-09

  • Table S2. Electrophysiological Properties of hESC-Derived NKX2.1::GFP+ Neurons on Mouse versus Human Cortical Feeders and Averaged Properties of GABAergic Synaptic Currents in GFP+ GABAergic Interneurons, Related to Figure 6 A) Electrophysiological properties of hESC-derived NKX2.1::GFP+ neurons (day 10-18 and day 6-18 group) on mouse versus human cortical feeders (related to Figure 6).

    DIV: days in vitro; RI: Input resistance; AP: action potential; AHP: after-hyperpolarization. *: P < 0.01, unpaired t-test, compared to corresponding 14-16 DIV group **: P < 0.001, unpaired t-test, compared to corresponding 14-16 DIV group †: P < 0.001, unpaired t-test, compared to 10 to 18 mouse (20-30 DIV) ††: P < 0.0001, unpaired t-test, compared to 10 to 18 mouse (20-30 DIV) Action potential properties were measured at threshold as elicited by current injection.

    10 to 18 on mouse feeder 6 to 18 on mouse feeder 10 to 18 on human feeder

    DIV 14-16 20-30 14-16 20-30 20-30

    RMP (mV) -38.6 ± 2.4 -65.6 ± 1.6** -36.1 ± 2.4 -62.6 ± 3.6** -44.8 ± 6.6†

    AP Amp (mV) 57.4 ± 8.3 66.2 ± 4.1 56.2 ± 7.3 64.2 ± 5.1 62.5 ± 5.1

    AP Rise time

    (ms)

    1.472 ± 0.07 1.08 ± 0.03** 1.52 ± 0.08 1.11 ± 0.09* 1.9 ± 0.38†

    AP ½ Width

    (ms)

    3.05 ± 0.06 2.42 ± 0.12* 3.02 ± 0.05 2.5 ± 0.11* 4.8 ± 0.32††

    AHP peak

    (mV)

    10.1 ± 3.5 15.6 ± 2.4 11.2 ± 3.5 16.3 ± 2.6 8.8 ± 3.72

    RI (GΩ) 1.5 ± 0.1 0.72 ± 0.05** 0.94 ± 0.05 0.68 ± 0.12 1.66 ± 0.2††

    No. of cells 8 15 8 12 5

  • B) Averaged properties of GABAergic synaptic currents in GFP-positive GABAergic interneurons (related to Figure 6).

    sIPSC mIPSC

    Amplitude (pA) 62.6 ± 5.2 35.1 ± 2.3*

    Rise time (ms) 2.4 ± 0.3 2.3 ± 0.4

    Decay time (ms) 66.2 ± 5.4 51.2 ± 3.6 *

    ½ width (ms) 54.5 ± 5.3 41.3 ± 3.2*

    No. of cells 8 4

    Recordings were made at 26-30 DIV. *:P < 0.05, unpaired t-test

  • Table S3. List of Primary Antibodies and Their Sources and Concentrations Used in the Current Study, Related to Figures 1–6

    Antibody Species Dilution Source Cat. No.

    ASCL1 Mouse 1/250 BD Pharmingen 556604 Calbindin Rabbit 1/3000 Swant CB38a Calretinin Mouse 1/2000 Swant 6B3

    ChAT Goat 1/100 Millipore AB144P DLX2 Rabbit 1/200 gift from J. Rubenstein n/a

    Doublecortin Goat 1/100 SCBT sc-8067 FOXA2 Goat 1/200 SCBT sc-6554 FOXG1 Rabbit 1/500 Neuracell NC-FAB FOXG1 Rabbit 1/1000 gift from Eseng Lai n/a GABA Rabbit 1/4000 Sigma Aldrich A2052

    GAPDH Rabbit 1/10000 Thermo Scientific PA1-16780 Gephyrin Mouse 1/2000 Synaptic Systems 147011

    GFAP Goat 1/300 SCBT sc-6170 GFP Chicken 1/1000 Abcam ab13970 KI67 Rabbit 1/1000 Thermo Scientific RM-9106 KI67 Mouse 1/200 DAKO MIB-1 LHX6 rabbit 1/500 Abcam ab22885 MAP2 Chicken 1/10000 Abcam AB5392 Nestin Mouse 1/500 Neuromics MO15012

    Nkx2.1 (TTF1) Rabbit 1/1000 Epitomics 2044-1 Nkx2.1 (TTF1) Goat 1/100 SCBT sc-8762

    NOS Rabbit 1/1000 Immunostar 24287 OLIG2 Rabbit 1/1000 Millipore AB9610

    Parvalbumin Mouse 1/2000 Millipore MAB1572 Parvalbumin Rabbit 1/2000 Swant PV25

    PAX6 Mouse 1/250 DSHB PAX6 PAX6 Rabbit 1/1000 Covance PRB-278P

    PSD-95 Rabbit 1/2000 Zymed 51-6900 RAX Mouse 1/250 Abnova H00008575-B01

    SATB2 Mouse 1/100 Abcam ab51502 SC121 Mouse 1/1000 Stemcells AB-121-U-050

    Somatostatin Rat 1/250 Millipore MAB354 TBR1 Rabbit 1/1000 Abcam ab31940

    TH Mouse 1/1000 Immunostar 22941 TuJ1 Mouse 1/500 Covance MMS-435P VGAT Rabbit 1/2000 Synaptic Systems 131003

    VGLUT1 guinea pig 1/5000 Millipore AB5905 VIP Rabbit 1/1000 Immunostar 20077

  • SUPPLEMENTAL EXPERIMENTAL PROCEDURES Western blot Cells were lysed with Ripa Buffer (50mM Tris- HCl pH 8.0, 120mM NaCl, 5mM EDTA, 0.5% NP-40, and a tablet of protease inhibitor). Protein concentration was measured by the Bradford Assay (Bio-Rad). For immunoblot, whole cell extract (4 x 106 cells per sample) were boiled for 5 minutes in Laemmli sample buffer before separation by electrophoresis on 4-12% Bis- Tris gel in MOPS SDS running buffer at 100V for 2 hours using a iBlot electroblotter (Invitrogen). Protein was transferred to nitrocellulose membrane using the iBlot gel transfer device (Invitrogen). Non- specific protein binding was prevented by blocking the membrane with 3% BSA and PBS-T (0.1% Tween- 20). Membrane was incubated at 4 degrees overnight in 3% BSA and PBS-T with primary antibodies (LHX6 (1:500; Abcam), RAX (1:250; Abnova), GAPDH (1:10,000; Thermo Scientific). After five washes with PBS-T (0.1% Tween-20), the blot was incubated with respective secondary antibodies (Rabbit (1:10,000), Mouse (1:20,000) at room temperature. ECL kit, Western Lightning Plus- ECL, was used for detection according to manufacturer's instruction (PerkinElmer). Secondary antibodies were all purchased from Jackson Immunoresearch. ECL kit, Western Lightning Plus- ECL, was used for detection according to manufacturer's instruction (PerkinElmer). Electrophysiology Whole-cell voltage and current clamp recordings were performed as described previously (Ying et al., 2007), and the methods were slightly modified for this study. Briefly, neurons were visualized using a Nikon microscope (ECLIPSE FN1) equipped with a 4x objective and a 40x water immersion objective and GFP-positive neurons were identified using epifluorescence optics. Neurons were recorded at 23 – 24 °C. Input resistance was measured from voltage response elicited by intracellular injection of a small current pulse (-2 or 5 pA, 500 ms). Electrical signals were obtained using a Multiclamp 700B amplifier connected to an interface using Clampex 10.2 software (Molecular Devices, Foster City, CA). Liquid junction potentials were calculated and corrected off-line (Ying and Golds