+ All Categories
Home > Documents > Nestin transgenic ES cells and in vitro neurogenesis

Nestin transgenic ES cells and in vitro neurogenesis

Date post: 03-Feb-2022
Category:
Upload: others
View: 4 times
Download: 0 times
Share this document with a friend
15
Introduction Neurogenesis is considered to be the most complex event of organogenesis during embryonic development involving a precise signalling and cellular interaction cascade in order to generate a functional neural network. This comprises various subtypes of neurons, astroglia and oligodendroglia. Unlike the earlier belief that mammalian nervous system stem cells are confined to the embryo and that adult brain lacks the capacity to regenerate, recent studies demonstrate the presence of neural stem cells in both fetal and adult hypothalamus, olfactory bulb, subventricular zone and the dentate gyrus of hippocampus (Reynolds and Weiss, 1992; Gage, 2000; Bjornson et al., 1999; Doetsch et al., 1999; van der Kooy and Weiss, 2000). Since there is generally a recurrence of embryonic phenotypes in adults after injury (Laywell et al., 1992; Lendahl, 1997; Rossi et al., 1997), whether as a means of repair or merely a default state, it is mandatory to investigate the early embryonic events in order to understand the significance of this phenomenon. Hence, pluripotent embryonic stem (ES) cells recapitulating the in vivo events in a relatively precise manner may serve as an ideal model system for the investigation of early embryonic developmental processes (Okabe et al., 1996). In particular, tissue-specific promoter-mediated targeting is an efficient approach in the identification, isolation and functional characterization of lineage-specific populations in the heterogeneous cell mass of the ES-cell system (Kolossov et al., 1998). In the present study the ES cells have been targeted with neural-specific enhancer driving live reporter expression in order to understand early neurulation events. The intermediate filament protein, nestin, a well known marker for neural stem cells, is expressed in the majority of mitotically active CNS and PNS progenitors (Lendahl et al., 1990; Lendahl, 1997; Cattaneo and McKay, 1990; Mujtaba et al., 1998); it is downregulated upon differentiation (Zimmerman et al., 1994; Lothian and Lendahl, 1997) and reported to reappear upon injury (Lendahl, 1997; Krum and Rosenstein, 1999; Namiki and Tator, 1999; Pekny et al., 1999). Thus, the cells expressing nestin show all the characteristic features of stem cells such as multipotency, self-renewal and regeneration. Hence, nestin could serve as an efficient candidate marker gene in order to unravel early neurogenic proceedings from ES cells in vitro. Unlike Tα1 tubulin, the unipotent neuronal progenitor marker whose expression is confined only to the pre- and post-mitotic neurons (Wang et al., 1998; Roy et al., 2000a; Roy et al., 2000b), nestin represents a more broad spectral, multipotent neural lineage marker expressing not only in neurons but in glia as well (Hockfield and McKay, 1985; Messam et al., 2000). Thus, the 1471 To gain insight into early events of neurogenesis, transgenic embryonic stem (ES) cells were generated using the enhanced green fluorescence protein (EGFP) reporter gene under the regulatory control of the neural stem cell marker, nestin. The expression of EGFP in undifferentiated ES cells suggested that the onset of endogenous nestin occurred before neurulation. Upon differentiation of ES cells, the EGFP expression became confined to the neural lineage and asynchrony in ES-cell-derived neural differentiation was evident. The EGFP intensity was prominent in the proliferative progenitors and unipolar neurons, whereas downregulation occurred in differentiating bi- and multipolar neurons. This was corroborated quantitatively using flow cytometry where maximal generation of neural progenitors was observed 4-12 days post-plating. The proliferative potential of neural progenitors as well as glia, in contrast to post-mitotic neurons, was also evident by time-lapse microscopy. The functional characterization of progenitors revealed an absence of voltage-activated inward currents, whereas the Na + current (I Na ) was detected prior to Ca 2+ currents (I Ca ) in differentiating neurons. Additionally, inhibitory receptor-operated channels could be detected at these early stages of development in bi- and multipolar neurons suggesting that the pre-committed progenitors had retained their intrinsic ability to give rise to functional neurons. This study sheds new light on early events of neurogenesis defining the quantitative and qualitative aspects and demarcating the functional neural cell types from ES cells in vitro. Movies available on-line Key words: ES cells, Nestin, EGFP, Neural progenitor, Neurogenesis Summary Quantitation and functional characterization of neural cells derived from ES cells using nestin enhancer- mediated targeting in vitro Nibedita Lenka* ,‡ , Zhong J. Lu, Philipp Sasse, Jürgen Hescheler and Bernd K. Fleischmann Institute of Neurophysiology, University of Cologne, Cologne, Germany *Present address: National Centre For Cell Science, Pune University Campus, Ganeshkhind, Pune 411007, Maharashtra, India Author for correspondence (e-mail: [email protected]) Accepted 13 January 2002 Journal of Cell Science 115, 1471-1485 (2002) © The Company of Biologists Ltd Research Article
Transcript
Page 1: Nestin transgenic ES cells and in vitro neurogenesis

IntroductionNeurogenesis is considered to be the most complex event oforganogenesis during embryonic development involving aprecise signalling and cellular interaction cascade in order togenerate a functional neural network. This comprises varioussubtypes of neurons, astroglia and oligodendroglia. Unlike theearlier belief that mammalian nervous system stem cells areconfined to the embryo and that adult brain lacks the capacityto regenerate, recent studies demonstrate the presence of neuralstem cells in both fetal and adult hypothalamus, olfactory bulb,subventricular zone and the dentate gyrus of hippocampus(Reynolds and Weiss, 1992; Gage, 2000; Bjornson et al., 1999;Doetsch et al., 1999; van der Kooy and Weiss, 2000). Sincethere is generally a recurrence of embryonic phenotypes inadults after injury (Laywell et al., 1992; Lendahl, 1997; Rossiet al., 1997), whether as a means of repair or merely a defaultstate, it is mandatory to investigate the early embryonic eventsin order to understand the significance of this phenomenon.Hence, pluripotent embryonic stem (ES) cells recapitulatingthe in vivo events in a relatively precise manner may serve asan ideal model system for the investigation of early embryonicdevelopmental processes (Okabe et al., 1996). In particular,tissue-specific promoter-mediated targeting is an efficientapproach in the identification, isolation and functional

characterization of lineage-specific populations in theheterogeneous cell mass of the ES-cell system (Kolossov et al.,1998). In the present study the ES cells have been targeted withneural-specific enhancer driving live reporter expression inorder to understand early neurulation events.

The intermediate filament protein, nestin, a well knownmarker for neural stem cells, is expressed in the majority ofmitotically active CNS and PNS progenitors (Lendahl et al.,1990; Lendahl, 1997; Cattaneo and McKay, 1990; Mujtabaet al., 1998); it is downregulated upon differentiation(Zimmerman et al., 1994; Lothian and Lendahl, 1997) andreported to reappear upon injury (Lendahl, 1997; Krum andRosenstein, 1999; Namiki and Tator, 1999; Pekny et al., 1999).Thus, the cells expressing nestin show all the characteristicfeatures of stem cells such as multipotency, self-renewaland regeneration. Hence, nestin could serve as an efficientcandidate marker gene in order to unravel early neurogenicproceedings from ES cells in vitro. Unlike Tα1 tubulin, theunipotent neuronal progenitor marker whose expression isconfined only to the pre- and post-mitotic neurons (Wang etal., 1998; Roy et al., 2000a; Roy et al., 2000b), nestinrepresents a more broad spectral, multipotent neural lineagemarker expressing not only in neurons but in glia as well(Hockfield and McKay, 1985; Messam et al., 2000). Thus, the

1471

To gain insight into early events of neurogenesis, transgenicembryonic stem (ES) cells were generated using theenhanced green fluorescence protein (EGFP) reporter geneunder the regulatory control of the neural stem cell marker,nestin. The expression of EGFP in undifferentiated ES cellssuggested that the onset of endogenous nestin occurredbefore neurulation. Upon differentiation of ES cells, theEGFP expression became confined to the neural lineageand asynchrony in ES-cell-derived neural differentiationwas evident. The EGFP intensity was prominent in theproliferative progenitors and unipolar neurons, whereasdownregulation occurred in differentiating bi- andmultipolar neurons. This was corroborated quantitativelyusing flow cytometry where maximal generation of neuralprogenitors was observed 4-12 days post-plating. Theproliferative potential of neural progenitors as well as glia,in contrast to post-mitotic neurons, was also evident by

time-lapse microscopy. The functional characterizationof progenitors revealed an absence of voltage-activatedinward currents, whereas the Na+ current (I Na) wasdetected prior to Ca2+ currents (ICa) in differentiatingneurons. Additionally, inhibitory receptor-operatedchannels could be detected at these early stages ofdevelopment in bi- and multipolar neurons suggesting thatthe pre-committed progenitors had retained their intrinsicability to give rise to functional neurons. This study shedsnew light on early events of neurogenesis defining thequantitative and qualitative aspects and demarcating thefunctional neural cell types from ES cells in vitro.

Movies available on-line

Key words: ES cells, Nestin, EGFP, Neural progenitor, Neurogenesis

Summary

Quantitation and functional characterization of neuralcells derived from ES cells using nestin enhancer-mediated targeting in vitroNibedita Lenka* ,‡, Zhong J. Lu, Philipp Sasse, Jürgen Hescheler and Bernd K. FleischmannInstitute of Neurophysiology, University of Cologne, Cologne, Germany*Present address: National Centre For Cell Science, Pune University Campus, Ganeshkhind, Pune 411007, Maharashtra, India‡Author for correspondence (e-mail: [email protected])

Accepted 13 January 2002Journal of Cell Science 115, 1471-1485 (2002) © The Company of Biologists Ltd

Research Article

Page 2: Nestin transgenic ES cells and in vitro neurogenesis

1472

generation and demarcation of both neurons and glia wouldpave the way in exploring the underlying mechanism(s) ofneurogenesis as well as gliogenesis in the ES-cell modelsystem.

Previous studies on the nestin gene using transgenic micedemonstrated that the second intron has the necessary cis-acting enhancer motifs for driving the reporter gene expressionin a neuron-specific manner in the developing CNS(Zimmerman et al., 1994; Lothian and Lendahl, 1997;Josephson et al., 1998; Lothian et al., 1999; Yaworsky andKappen, 1999). Accordingly, the present investigation wascarried out to identify, quantify and functionally characterizethe ES-cell-derived, development-dependent, proliferatingneuronal progenitors as well as the differentiating neuronsbased on nestin intron-II-driven EGFP expression. This studyprovides clues to the multifaceted and dynamic process ofneurogenesis using the powerful model of stem celldifferentiation in vitro that would otherwise have been acomplicated task in vivo.

Materials and MethodsVectorsThe Nes 1852 tk/lacZ, a human nestin promoter vector construct wasa kind gift of Urban Lendahl, Sweden (Lothian and Lendahl, 1997).The 2 Kb nestin intron II enhancer-tk promoter fragment was excisedfrom the aforementioned vector construct using HindIII and NotIrestriction enzymes. Subsequently, this fragment was subcloned intothe pEGFP1 vector (Clontech, Germany) at the SmaI site by blunt endligation and termed h-Nestin-EGFP, where the EGFP reporter is underthe regulation of the nestin-tk promoter. The insertion and orientationof the promoter to pEGFP vector was verified by both restrictiondigestion and sequencing using an automated sequencer (ABI). Toobtain the tkEGFP vector construct, the nestin intron II was releasedfrom h-Nestin-EGFP by SalI restriction digestion and by subsequentrecircularization of the vector.

Cell culture and transfectionThe ES cells of the line D3 were maintained as described before(Wobus et al., 1991). Briefly, ES cells were grown on inactivatedfeeder cells in DMEM supplemented with 15% fetal bovine serum(FBS), non-essential amino acids, 2 mM glutamine, 50 µg/ml Penn-Strep (all from Gibco BRL, Germany), 0.1 mM β-mercaptoethanol(Sigma, Germany), and 1000 U/ml recombinant murine leukemiainhibitory factor (LIF, ESGRO, Gibco). In separate experiments,following linearization using the HindIII restriction enzyme ~5×106

cells were transfected independently by electroporation using ~30 µgof the h-Nestin-EGFP or tkEGFP reporter constructs, respectively,following the standard protocol. The NeoR clones were picked after10-12 days of G418 selection and propagated using the same medium.G418 selection was maintained throughout the propagation period.

Neuronal differentiationDifferentiation of ES cells was initiated by cell aggregation followingthe hanging drop method (Wobus et al., 1991) in DMEM supplementedwith 10% FBS. All-trans-retinoic acid (RA) was used as the inducingagent for neuronal differentiation as described (Strubing et al., 1995)with slight modifications. The neural differentiation was alsomonitored using the conditioned medium (Okabe et al., 1996). Sincethe time window (~7-10 days post-plating) for optimal neuralprogenitor generation was similar irrespective of the medium used, thedata from RA-induced progenitors are presented here for convenience.In brief, following trypsinization, the ES cells (500cells/20µl drop)

were exposed to RA (10–7 M) for 3 days during hanging drop followedby suspension culture of embryoid bodies (EBs) for 4 days and platingwithout RA. Alternatively, RA was added to the medium while platingthe EBs and the medium was replenished with fresh medium withoutRA on the fourth day post-plating (i.e. d7+4).

RT-PCRTotal RNA was isolated from the undifferentiated EGFP-positivetransgenic nestin ES-cell clones and the wild-type ES-cell line D3,grown with or without feeders, as well as from murine blastocystsusing high pure RNA isolation kit (Roche Molecular Biochemicals,Germany). To ascertain the presence of nestin transcripts in thesecells, the total RNA was subjected to reverse-transcription-based PCRfollowing the manufacturer’s instructions (Gibco-BRL). The nestin-specific primers (5′ GGATACAGCTTTATTCAAGG 3′ and 5′ CAG-CCGCTGAAGTTCACTCT 3′; GenBank, accession no. C78523)were designed from the retrieved mouse cDNA sequence, which alsocorresponded to the C-terminus of the latest reported full lengthmouse nestin gene (GenBank accession no. AF076623) at positions5959-5940 and 5481-5500, respectively. The house-keeping HPRT(hypoxanthine phosphoribosyltransferase) (Johansson and Wiles,1995) and β-actin primers were used as internal positive controls forPCR and designed accordingly to decipher the genuineness ofamplified products, based on their size from cDNA, but not fromcontaminating genomic DNA. The RNA samples without the reversetranscriptase served as the negative control, and the PCR product fromeach sample was resolved on the agarose gel and observed under aUV-transluminator.

FACS analysis The EGFP expression of ES-cell-derived clones was analyzed atvarious developmental stages using a FACSCaliburTM flow cytometer(Becton Dickinson) equipped with a 488 nm argon-ion-laser (15 mW)as described (Kolossov et al., 1998). In brief, about 10,000-50,000viable cells were analyzed per sample after isolation usingtrypsinization (0.1% trypsin and 0.01% EDTA) for 2-5 minutes at37°C. Subsequently, the cells were resuspended by gentle triturationusing PBS containing 1 mM Ca2+ and 0.5 mM Mg2+. UntransfectedD3 line ES cells were used as negative controls. The emittedfluorescence of EGFP was measured in log scale at 530 nm (FITCband pass filter) and analyses were performed using CellQuest

software (Becton Dickinson).

ImmunocytochemistryThe specificity of the nestin expression was verified byimmunocytochemistry using antibodies against neurons and gliafollowing the standard protocol. In brief, the ES cells and EBs at variousstages of development grown on glass coverslips were washed with 0.1M PBS, pH 7.4 and fixed with 4% paraformaldehyde for 20 minutes.After washing again with PBS the cells were permeabilized withsolution containing 0.25% Triton X-100 and 0.5 M ammonium chloridein 0.25 M TBS, pH 7.4 for 10 minutes and blocked with 4% goat serumand 0.8% BSA for 1 hour. Subsequently, the cells were exposed to eitherof the primary antibodies: anti-nestin (Rat-401, DSHB, University ofIowa), anti-MAP2, anti-synaptophysin or anti-GFAP (all from Sigmachemicals) for 3 hours at room temperature. After washing (0.25 MTBS, three times for 10 minutes each), the cells were treated with Cy3-conjugated secondary antibody for 1 hour at 37°C for fluorescentlabelling. Finally the cells were washed three times with TBS anddehydrated with ethanol gradients, followed by xylene treatment andmounting with entellan on glass slides. In each case the negative controlwas performed with the substitution of respective primary antibodieswith goat pre-immune serum. The slides were observed under afluorescent microscope to detect EGFP as well as Cy3-labeling.

Journal of Cell Science 115 (7)

Page 3: Nestin transgenic ES cells and in vitro neurogenesis

1473Nestin transgenic ES cells and in vitro neurogenesis

Dissociation of EBs and preparation of single cellsFor isolation, 12-16 EBs were dissociated using collagenase B (RocheMolecular Biochemicals, Germany) and re-plated on gelatin (0.1%)-coated glass coverslips as described (Maltsev et al., 1994). In brief,EBs were dissected and rinsed with PBS followed by collagenase Btreatment (0.1% in PBS) for 30 minutes at 37°C. Subsequently,collagenase was replaced with a medium containing: 85 mM KCl; 30mM K2HPO4; 5 mM MgSO4-7H2O; 1 mM EDTA; 5 mM Na2ATP; 5mM Na-Pyruvate; 5 mM Creatin; 20 mM taurin and 20 mM glucose;pH 7.2). The cells were stirred slowly for 30 minutes, suspended bygentle trituration in DMEM medium and plated onto gelatin-coatedglass coverslips. In initial experiments the central and peripheralregions of the EBs were separated and dissociated individually. Sincethe pattern of differentiation was similar in these two preparations,whole EBs were used for dissociation. Single isolated cells were usedfor immunocytochemical characterizations after 2-5 days of re-platingeither with parallel cultures or with the same, subsequent toelectrophysiological investigations.

Estimation of EGFP intensity and electrophysiologyThe semiquantitative estimation of EGFP intensity was performedas described (Kolossov et al., 1998). For the analysis, the EGFPfluorescence intensity comprising the whole area of the cell wasintegrated and average fluorescence intensities determined in counts.

For patch clamp recordings, isolated neurons of differentdevelopmental stages were investigated, using the whole-cell patchclamp technique (Hamill et al., 1981). The cells were voltage-clampedusing an EPC-9 amplifier (Heka, Germany), held at –80 mV anddepolarizing pulses or ramps were applied (for detail, see figurelegends). For the registration of inward currents ramp depolarizationswere performed and INa was inhibited by tetrodotoxin (TTX, 0.1 µM).For estimation of voltage-dependent Ca2+ currents, the extracellularsolution was exchanged to a Na+-free solution containing Ba2+ ascharge carrier. The expression of the various subtypes of voltage-dependent Ca2+ channels was tested using selective antagonists[Isradipine, ω-Conotoxin GVIA (ω-CgTX), ω-agatoxin IVA (ω-AgaTX), Almone Labs, Israel], which were bath added. For therecording of receptor-operated currents, substances were applied (15milliseconds) through a puffer pipette connected to a pressure ejectionsystem (General Valves, USA) placed into the vicinity of the cell ofinterest (holding potential (HP) –80 mV). The receptor-operatedcurrents were characterized by applying competitive antagonists viathe gravitational perfusion system. The ionic nature of these currentswas analyzed using subtracted voltage ramps. Data were acquired ata sampling rate of 10 kHz, filtered at 1 kHz, stored on hard disk andanalyzed off-line using the Pulse-Fit (Heka) software package.Averaged data are expressed as means±s.e.m. Membrane capacity wasdetermined on-line using the Pulse acquisition software program(Heka). Statistical analysis was performed using paired or unpairedStudent’s t-test, and a P value of <0.05 was considered significant.

The glass coverslips containing the cells were placed into atemperature-controlled (37°C) recording chamber and perfusedcontinuously with extracellular solution by gravity at a rate of 1ml/minute. Substances were applied by exchanging the solution in thechamber, a 90% volume exchange was achieved within approximately20 seconds. Patch pipettes (2-4 MΩ resistance) from borosilicate glassfrom Clark (Electromedical Instruments, UK) were pulled using aZeitz puller (DMZ, Germany). The solutions used had the followingcomposition. Standard external solution, 140 mM NaCl, 5.4 mM KCl,1.8 mM CaCl2, 1 mM MgCl2, 10 mM glucose and 10 mM Hepes (pH7.4, adjusted with NaOH). External solution for measurement of IBa:120 mM D(–)-N-methylglucamine, 10.8 mM BaCl2, 5.4 mM CsCl, 1mM MgCl2, 10 mM glucose and 10 mM Hepes (pH 7.4, adjusted withHCl). Normal intracellular solution: 50 mM KCl, 80 mM K-Asparate,1 mM MgCl2, 3 mM Mg2ATP, 10 mM EGTA and 10 mM Hepes (pH7.2, adjusted with NaOH). Internal solution for recording of inward

currents: 120 mM CsCl, 1 mM MgCl2, 3 mM Mg2ATP, 10 mM Hepesand 10 mM EGTA (pH 7.2, adjusted with CsOH). All these chemicalswere purchased from Sigma (Germany). The toxins were aliquoted,diluted in normal external solution and frozen at –22°C prior to use.

Online supplemental material (time-lapse microscopy) For live monitoring of neurogenesis, one day after isolation (d7+7)the cells were continuously observed for 3 days. An invertedmicroscope (Axiovert 100, Zeiss) equipped with a movable computercontrolled stage (Nikon, Germany), an objective 20× (Plan-Neofluar,Zeiss) and a temperature/CO2-controlled chamber (Nikon) were used.Cells were monitored at 30 minute intervals using alternatetransmission and fluorescent excitation light with a conventionalhalogen lamp and a FITC filter set. Images were acquired using acolour video camera (Sony, DXC 950P, AVT Horn, Germany)controlled by the Lucia software (Nikon) and digitized on-line via avideo frame grabber card (Matrox Corona, Nikon). For fluorescencepictures, the integration mode of the camera (10 single pictures) anda video-based buffer device (Sony MPU-F100P, AVT Horn) wereemployed. Movies 1 and 2 depict Fig. 5A and B, respectively, showingthe neuro- and gliogenesis from neural progenitors. The cellssubsequent to the monitoring were subjected to immonocytochemicalverification and neural specification using neuronal and glial specificantibodies (see http://jcs.biologists.org/supplemental).

ResultsEstablishment of stable transgenic nestin ES-cell clonesand monitoring of EGFP expressionA number of stable G418 resistant transgenic ES-cell clones ofthe line D3 were generated by transfection of the h-Nestin-EGFP construct. Several of these selected clones weredifferentiated into various neural phenotypes following the invitro differentiation protocol as described before using retinoicacid as the inducer and simultaneously monitored for EGFPexpression. 96% of the G418-resistant nestin ES-cell clones(n=100) were observed to be EGFP positive even at theundifferentiated state (Fig. 1A-C). Typically, the clonesexhibited a heterogeneous expression pattern of EGFP duringpropagation on feeder (Fig. 1B,C) or feeder-free culturesindirectly implying the nestin transcription to be active even inundifferentiated ES cells prior to any lineage commitment.Since almost all of our selected, undifferentiated ES-cell cloneshappened to be EGFP positive, the possibility of non-specificity due to positional effect could be ruled out.Additionally, the non-specific EGFP expression pattern due tothe constitutively active heterologous thymidine-kinase (tk)basal promoter was also excluded by performing another set oftransfections using the tkEGFP promoter construct without thenestin intron II segment. In less than 15% of these G418-resistant clones only weak EGFP expression was detected (Fig.1D,E). This is in line with the observation by Lothian andLendahl who demonstrated the absence of any ectopic lacZexpression driven by the same tk basal promoter in transgenicmice (Lothian and Lendahl, 1997).

Validation of early onset of endogenous nestin andconcomitant EGFP expressionTo correlate nestin-driven EGFP expression in undifferentiatedES cells with endogenous nestin expression, both, RT-PCR andimmunostaining were performed. As shown in Fig. 2, by RT-

Page 4: Nestin transgenic ES cells and in vitro neurogenesis

1474

PCR with total RNA isolated from undifferentiated ES cellsfrom one of the EGFP-positive transgenic clones (clone-1) aswell as from untransfected D3 cells grown in presence orabsence of feeders using sequence specific nestin primers, aPCR product of the expected 478 bp size was amplified.Interestingly, the nestin product was detectable in the murineblastocyst RNA (Fig. 2, lower panel, lane 14), but with acomparatively low intensity. This is probably caused by eithera low transcript level or low initial RNA content in the RTreaction containing the blastocyst sample because, by usingtwo housekeeping primers (HPRT and β-actin), the respectiveapparent product intensities with the blastocyst sample werelow compared with that of other samples. Hence, we tookdouble the quantity of reverse transcribed first strand from theblastocyst sample in order (1.5 times loading sample volumeon gel) to scale up the visible intensity of the product duringPCR (Fig. 2, lower panel, lane 8). We further corroboratedthis finding by immunostaining using a monoclonal antibody

against nestin. All the EGFP-positive cells were nestin positivein the feeder-free cultured transgene ES cell clone-1 (Fig.1G,H). Immunostaining with a monoclonal antibody againstnestin in the feeder-free cultured transgenic ES cell clone-1indicated that all EGFP-positive cells were nestin positive (Fig.1G,H). In addition, preliminary studies indicate that the innercell mass (ICM) of murine blastocysts is nestin positive (datanot shown), confirming the observation in ES cells. Unlikeearlier reports on nestin expression (Okabe et al., 1996; Lee etal., 2000) our findings imply that its onset occurs much earlierthan neural plate induction at E7.5.

Nestin intron-II-driven EGFP expression profile duringneural differentiation and neurogenic progressionThe transition from uncommitted ES cells to a neural lineage-defined state was brought about by RA exposure of EBs. Torule out clonal diversity several (n=6) of the clones were

Journal of Cell Science 115 (7)

Fig. 1.Human nestin intron-II-drivenEGFP expression in undifferentiatedES cells and correspondingimmunocytochemical expressionpattern. (A) A single NeoR clone afterelectroporation of h-Nestin-EGFPconstruct and G418 selection for 12days shows heterogeneous EGFPexpression. (B,C) The EGFPexpression in ES cells (arrow) remainspatchy during propagation onmitotically inactive feeders (arrowhead) in clone-1. (D,E) Weak EGFPexpression in one of the few (<15%)EGFP-positive ES-cell clonestransfected with tkEGFP.(B,D) Combined transmission andfluorescent light; (C,E) fluorescencelight alone. (G,H) A complete overlapbetween EGFP expression (G) and (H)nestin-immunoreactivity was observedin undifferentiated ES cells duringfeeder-free propagation. Bars, 100 µM(A), 75µM (D,E), 50 µM (B,C,G,H).

Page 5: Nestin transgenic ES cells and in vitro neurogenesis

1475Nestin transgenic ES cells and in vitro neurogenesis

analyzed to examine the nestin intron-II-driven EGFPexpression profile and specificity in ES-cell-derived neuronsduring differentiation. During cell aggregation andcommencement of differentiation, localized expression ofEGFP in EBs marked the transition. Moreover, under no timepoints studied was there ever complete absence of EGFP-expressing cells in the EBs. After 1 or 2 days post-plating(d7+1/2), the EBs showed a few intense bright shining areas(Fig. 3A,B). By d7+2/3, the cells at the periphery exhibited anepithelial phenotype implicating the differentiation intoectodermal lineage. Starting from d7+4, outgrowths from thecentral part of the EB with areas of intense EGFP expressionwere detected (Fig. 3C,D). This further signified the presenceof neuroepithelial cells as confirmed by immunostaining withthe nestin antibody (Fig. 3J,K).

By d4-6 of plating, few EGFP-positive uni-/bi-/multipolarneurons were observed at the periphery ofthe differentiating EBs along with the glialcells, most probably the radial glia (Rat401 positive) (Hockfield and McKay,1985). However, the number of neuronswith distinct neurite outgrowthssignificantly increased with time afterplating, associated with weakening ofEGFP expression (Fig. 3D). By d8-15,bundles of neurons forming extensivenetworks were observed (Fig. 3E-H).As would be expected, progressiveneurogenesis was accompanied by anincrease in the number and the length ofneurite outgrowths. (Fig. 3, compare Ewith F-H). Most of the terminallydifferentiated neurons lost EGFP, thusmimicking the endogenous nestinexpression pattern as well as reporter geneexpression in vivo (Zimmerman et al.,1994; Lothian and Lendahl, 1997).Nevertheless, in many differentiatingneurons with distinct neurite outgrowthsweak EGFP expression was preserved;this was probably due to accumulationbecause of its slow metabolic rate (Fig.3F-H). No significant differences werenoticed in the neural differentiationpattern, irrespective of the application ofRA, during hanging drop preparation orplating in contradiction to an earlier report(Rohwedel et al., 1999).

Neural specificity of EGFPexpression during differentiationImmunocytochemical studies on wholeEBs (Fig. 3J-M) confirmed neuralspecific confinement of nestin enhancer-driven EGFP expression at differentdevelopmental stages (d7+4 to d7+12).A similar neural differentiation patternwas observed in isolated cells (Fig. 4A-K). All EGFP-positive cells stained withthe antibody directed against nestin,

proving the neural-specific expression of EGFP. TheseEGFP+/nestin+ cells (Fig. 3J,K; Fig. 4C-D) were thought tobe neural progenitors because of their proliferation (BrDU+

or Ki67+; data not shown) (Fig. 5; time-lapse observation)and differentiation potential, giving rise to neurons with uni-, bi- and multipolar morphologies as well as glia; as discernedby immunostaining with both neuron (synaptophysin, MAP2)and glial (GFAP) specific antibodies, respectively (Fig. 3L,M;Fig. 4E-K). The differentiated bi- and multipolar neuronswith longer processes were either weak or negative for EGFP(Fig. 4A,B), but stained positively with MAP2, a marker forpost-mitotic neurons (Fig. 3L,M; Fig. 4J,K). These cellsmaintained synaptic connections as evident fromsynaptophysin staining (Fig. 4G,H) and appeared to formnetworks with adjacent neurons and glia. More oftendifferentiating neurons grew on top of or adjacent to a

Fig. 2.RT-PCR revealed the presence of nestin transcripts in undifferentiated ES cells as wellas murine blastocysts. The upper panel shows nestin transcripts from D3 undifferentiated EScells (lane 2), transgenic h-Nestin-EGFP ES cells (n-ES cells; lane 5) grown on feeders andfrom D3 undifferentiated ES cells grown without feeders (lane 8). The lower panel showsnestin transcripts in the transgenic ES cells without feeders (lane 2) and in the murineblastocysts (lanes 8,14). The length of the nestin PCR product corresponds to the expectedsize of 478 bp. The corresponding products for housekeeping HPRT (upper panel, lanes4,7,10; lower panel, lanes 5,10) and β-actin (lower panel, lanes 6,12) are shown as controls.Lanes 3,6,9 (upper panel) and 3,4,7,9,11,13 (lower panel) represent the respective negativecontrols where the reverse transcriptase was omitted from the reaction.

Page 6: Nestin transgenic ES cells and in vitro neurogenesis

1476

monolayer of flat glial cells establishing extensiveconnections with each other (Fig. 4J,K). These observationsfurther confirmed that the EGFP/nestin positive neuralprogenitors retained the ability of multipotentiality since theywere able to generate MAP2-positive neurons and GFAP-positive glial cells upon differentiation. Notably, at everydevelopmental stage investigated there was a representativepopulation from proliferating (BrDU+; Ki67+; nestin+) anddifferentiating cells (MAP2+; GFAP+), indicating anasynchronous profile in neural differentiation.

Monitoring of neural differentiation with time-lapsemicroscopyA clearer picture of neurogenic progression emerged in time-lapse experiments. As seen in Fig. 5A, a single EGFP-positiveprogenitor was observed to divide into two and then into threecells. The triplet underwent positional changes between 6 and10 hours after the inception of monitoring (see movie;http://jcs.biologists.org/supplemental). Subsequently, one ofthese cells underwent morphogenic change and showed neuriteoutgrowth to become uni- and then bipolar. At subsequent time

Journal of Cell Science 115 (7)

Fig. 3.EGFP expression profileduring differentiation and neuralspecification. The expression wasprominent in progenitors anddownregulated in differentiatingneurons. Localized EGFPexpression in the EB (A,B; d7+2)and in outgrowths (arrow) of theEB (C; d7+4) post-plating.(C) Arrowhead, central part ofEB. (D) EBs (d7+7) showdifferentiating neurons (arrow) atthe periphery with EGFPexpression (see inset); arrowheadshows the central part of the EB.(E) d7+14; three clusters ofneurons (1,2,3) with short andlong processes. (F) (d7+20); withfurther plating time cluster 2shows an extension of neuriteoutgrowths from the central cellmass. (G,H) Magnified view ofthe same cluster, where someneurites are still EGFP positive.The arrowhead indicates thepresence of EGFP-positive neuralprogenitors in the vicinity ofmature neurons, and the arrowindicates the lengthening ofneurite processes with longerplating time from d7+14 to d7+20(E-H). (J,K) Almost all of theEGFP-expressing cells (J) in EBs(d7+7) are nestin positive (K); theyellow colour indicates thesuperimposed pattern betweenEGFP and Cy3. (L,M) In thedifferentiated neurons labelledwith MAP2 (M, double filter), aweak or no EGFP signal (L,d7+7) was observed. Combinedtransmission and fluorescentlight: A,C,D-F,G. Fluorescentlight alone: B,H,J,L,inset D. Bars,50 µM (A,B,inset D); 100 µM(C,D,G,H,J-M); 250 µM (E,F).

Page 7: Nestin transgenic ES cells and in vitro neurogenesis

1477Nestin transgenic ES cells and in vitro neurogenesis

periods the extension of a neurite in association with migrationand interaction with other cells became evident from thisobservation (Fig. 5A). One of the other two cells exhibitedfurther divisions at 23 hours (data not shown) and 36.5 hours,indicating the probable occurrence of self-renewingasymmetric division and retention of stem/progenitor potential.Many of the bipolar neurons did undergo a further

morphogenic change into multipolarones, depending on their interactionwith neighbouring neurons or glia (datanot shown), whereas some remainedbipolar until the end of monitoring (Fig.5A). These changes indicate the co-existence of a diverse phenotypicpopulation of mature neurons andneural progenitors at any given timeduring neuronal differentiation from EScells in vitro. Similarly, the generationof cells with glial phenotype withsubsequent division (Fig. 5B; 55 hoursand 58.5 hours), migration andinteraction with each other could beclearly depicted from EGFP-positiveprogenitors (Fig. 5B; 55-66.5 hours).The noteworthy feature was bothsymmetric and asymmetric division ofprogenitor and, unlike the neuronalcells, the glial cells underwentfurther division after undergoingtransformation from a flat (Fig. 5B; 54hours) to a round phenotype (Fig. 5B;54.5 hours) like the progenitor.However, in both cases weakening ofEGFP remained associated with

progressive morphogenesis into neuron and glia.

Flow-cytometric quantification of neurogenesis from EScells in vitroThe qualitative aspect of neurogenesis in vitro wascomplemented by a quantitative assessment performing flow-

Fig. 4. Isolated cells mimic neurogenesisobserved in whole EBs (d7+7).(A,B) Differentiating neurons with bipolar-(arrow) and multipolar (arrowhead)morphology were seen after plating ofisolated cells A: combined transmissionand fluorescence; B: fluorescence. (C,D) Ahigh overlap (D, yellow) between EGFPexpression (C) and nestin labelling wasobserved. Besides the generation ofneurons, EGFP-positive and -negative glialcells (E) were identified based on GFAPstaining (F, superimposed). Flat glial cellswere weak/negative for EGFP (arrowhead),while elongated fibre-like glial cells wereEGFP positive (arrow). (G,H) Theformation of synaptic connections betweenisolated neurons was confirmed bysynaptophysin staining (H). As expectedfor differentiated neurons, low or no EGFPsignal (G) was detected in synaptophysin-positive cells. Little overlap (K,superimposed) between MAP-2 staining(K) and EGFP expression (J) was detected.The flat glial cells (green) anddifferentiating neurons (red) were found toclosely associate. Bars, 25 µM.

Page 8: Nestin transgenic ES cells and in vitro neurogenesis

1478

cytometry at different stages of development. In several of theEGFP-positive clones analyzed (n=4), the undifferentiated EScells exhibited about 40-50% of bright and weak EGFPfluorescence each when compared with D3 wild-type ES cells(Fig. 6A). Upon differentiation by cell aggregation, RAtreatment and prior to plating (Fig. 6A; left panel, 7+0), therewas a clear leftward shift in the EGFP peak (28% weak: 20-100 intensity range; and 2% bright: 100-1000 range) with themajority (70%) of cells being EGFP negative (1-20 range). Inthe post-plating EBs, EGFP fluorescence further intensified inline with the microscopic observation pattern reported above.About 1-6% of the total bright shining population of cells fromthe RA-treated group (clone-1) was very intense after plating,with EGFP intensity levels beyond the 1000 range (Fig. 6A).

This population was observed even upto 24 days after plating. As seen inFig. 6A, there was a rightward shift inthe EGFP-positive peak in the 4 daypost-plated EBs (70% EGFP+: 20-10,000 range; 30% EGFP–: 1-20range) indicating a biphasicdistribution pattern that remainedalmost the same in the 10-12 day post-plated EBs (52-57% EGFP+ and 43-48% EGFP–) and gradually decreasedwith the course of time from 12 to 24days post-plating (20% EGFP+ and80% EGFP–). This clearly compliedwith an increase in the number of post-mitotic differentiating neurons duringthe course of differentiation asdirectly monitored with time-lapsemicroscopy. The time course of theneurogenic development indicatedthat the percentage of the brightestEGFP-positive (EGFP+++: 1000-10,000 intensity range) progenitor

population in the plated EBs remained stable between d7+4(7.2±2.0%) and d7+10/12 (6.0±1.3%) of plating and declinedto 2.5±1.0% after 24 days of plating (Fig. 6B). Similarly, theproportion of cells belonging to the bright EGFP-positive(EGFP++: 100-1000 range) group showed gradual decline inEGFP intensity (from 27.9±3.3% to 10.9±5.0%), implying theloss of residual EGFP in these differentiating neuralpopulations during a longer plating period. By contrast, theproportion of weak EGFP-positive cells (EGFP+: 20-100range) remained unchanged (24±11.3% to 28.7±2.3%) and analmost exponential rise in the EGFP-negative cells from36.1±2.0% to 62.6±17.3% could be observed between day 4and day 24 after plating (Fig. 6B). Taken together, these datasuggested an asynchronous neural differentiation profile and

Journal of Cell Science 115 (7)

Fig. 5.Time-lapse experiments showedlive the neural progenitor proliferationand differentiation into neurons as well asglia. (A) Transmitted light andfluorescence monitoring of isolated EB-derived cells. Starting from a singleprogenitor (arrow), characterized by itstypical morphology and prominent EGFPexpression, cell division (2.5 h, 6 h, 36.5h) and differentiation into neuronal(unipolar, 11.5 h; bipolar, 12 h) cells(arrowhead) was observed in arepresentative experiment. Uponmigration and further differentiation theEGFP fluorescence declined. (B) In linewith the retained capacity of proliferationof glia (arrow), prior to division rounding-up of the glial cell was noticed (54.5 h).After division, migration, cellularinteraction as well as flattening of thecells occurred. The cell labelled witharrowhead is likely of other origin. Bars,30 µM (A); 50 µM (B).

Page 9: Nestin transgenic ES cells and in vitro neurogenesis

1479Nestin transgenic ES cells and in vitro neurogenesis

that the maximum number of neural progenitors was generatedfrom ES cells in vitro between 4 and 12 days after EB plating.A similar profile was observed in EBs treated with RA duringplating (data not shown).

In contrast to the RA-treated groups, cells without RAexposure displayed intensities hardly beyond the 103 log andthe majority (71-90%) of post-plated EBs were EGFP negative(Fig. 6A, right panel). Additionally, the percentage of EGFP-positive cells increased from 9% (d7+4) to 28% (d7+24) aftertwo to three weeks of plating. This clearly indicated the effectof RA on early onset neural specification and neurogenesis thatwas otherwise delayed in the untreated ones. It was furthercorroborated by immunocytochemical findings where theRA-untreated EBs displayed low percentage of MAP2immunoreactivity after 2-3 weeks of plating (data not shown).

Functional features of transgenic nestin ES-cell-derivedneuronsIt is well known that ion channels determine development aswell as function of neurons. Therefore, we have focused onthe development-dependent expression of ion channels, inparticular voltage-activated inward currents. In fact, the livelabelling of ES-cell-derived neurons enabled for the first timethe investigation of ion channel expression in isolated neuronsat very early stages of neurogenesis. Since the differentiationof ES-cell-derived neurons was found to be asynchronous,cells undergoing functional characterization were classifiedaccording to the following criteria: (1) time after plating;and (2) morphological characteristics and semi-quantitativeassessment of EGFP expression. Isolated neurons from d3-19after plating were investigated using the classic whole cellpatch-clamp technique combined with single cell fluorescenceimaging to estimate EGFP intensity. In line with our previousobservations, neural progenitors (1804±223, n=25) andunipolar neurons (1585±457, n=25) were characterized bysignificantly higher EGFP intensities than more differentiatedneurons (bipolar neurons, 545±86, n=66; multipolar neurons,482±77, n=56). Many of the multipolar neurons, probably ofa more advanced state of differentiation, were withoutdetectable fluorescence (n=8).

When the expression pattern of ion channels wascharacterized in the four population of neurons, no voltage-dependent ion currents were detected in neuralprogenitors/apolar cells during the entire developmental stage(Fig. 7A, n=25). Indeed, at this stage ramp depolarizations aswell as single voltage steps (data not shown) did not yieldmacroscopic currents. EGFP-positive neural progenitors werecarefully selected (morphological criteria) in order to avoidcontamination by other cell types. The unipolar EGFP-positiveneurons (Fig. 7B) expressed voltage-dependent outwardrectifying currents (n=7), whereas no voltage-activated inwardcurrents, neither INa nor IBa, could be detected in the largemajority of cells (22 out of 24). The outward rectifying currentsrecorded in the bipolar as well as multipolar neurons wereidentified as K+ currents based on their sensitivity towards theK+ channel blocker, 4-aminopyridine (4-AP). 4-AP (2 mM)inhibited 73.69±8.2% of the aggregate outward current (HP –80mV, step potential +50 mV) (n=6, data not shown). Interestingly,at the early developmental stage (EDS, d7+3/4), INa but not IBa(Fig. 7E) was detected in 86% of bipolar neurons (n=7), whereas

almost all (80%) the multipolar neurons (n=5) expressed bothINa and IBa. At subsequent stages (LDS, d7-19) all the bipolarand multipolar neurons functionally expressed both currents. INawas further characterized using the Na+ channel selectiveantagonist TTX which, as expected for neuronal Na+ channels,blocked INacompletely at a concentration of 0.1 µΜ (Fig. 7C,D).INa densities in bipolar neurons increased significantly withplating time (Fig. 7F) from 48.1±8.3 to 106.4±10.1 pA/pF(n=60) and in multipolar neurons (data not shown) from40.4±10.9 to 80.7±7.0 pA/pF (n=50), respectively. INa density inthe bipolar neurons was higher than that in the multipolarneurons (Fig. 7G). To further envisage whether there werequalitative and/or quantitative differences in the expressionpattern of IBasubtypes during development, we investigated bothbipolar and multipolar neurons by applying selectiveantagonists. In four typical experiments of bipolar neurons wefound that the L-type Ca2+ channel antagonist Isradipine (3 µΜ)suppressed 22.7±3.6% of the aggregate IBa, the N-type specificCa2+ channel blocker ω-CgTX (3 µM) suppressed 14.2±3.7%and the P/Q type channel blocker ω-AgaTX (0.1 µM) suppressed8.5±1.6%, respectively (Fig. 8B,C). The remaining Cd2+-sensitive (50 µM CdCl2) component amounted to 54.6±8.2% ofcontrol IBa. Similarly, in four typical multipolar neurons thefraction of different Ca2+ channel subtypes amounted to: L-type,27.3±1.9%; N-type, 9.1±3.2%; P/Q type, 16.2±3.7%; and other-types, 47.4±5.6%, respectively (Fig. 8D).

Since receptor-operated ion channels are known to play anessential role not only in carrying out fast synaptic transmissionbut also in the mediation of trophic signals, and migration andsynaptic arrangement during neuronal development, we testedtheir functional expression at different time points duringneurogenesis. The neural progenitors (n=3) and unipolarneurons (n=4) did not respond to γ-aminobutyric (GABA),glycine, kainate and NMDA. Clear agonist responses to GABA(1 mM) and glycine (1 mM) were detected (Fig. 9A,B) inbipolar (n=8, current density 52.1±22.0 pA/pF) and multipolarneurons (n=8, current density 119.1±51.1 pA/pF). Moreover, asexpected, the receptor-operated currents were blocked by thecompetitive antagonists bicuculline (100 nM, n=8) andstrychnine (30 µM, n=8), respectively (HP –80 mV; Fig. 9A,B).To further confirm the ionic nature of the GABA and glycine-evoked currents, voltage ramps in the presence of the agonistwere subtracted from control ramps. As can be seen in the insetof Fig. 9A, the subtracted voltage ramps for GABA yielded areversal potential of –28.3±1.1 mV (n=5), a value close to thecalculated Cl– equilibrium potential of –28.3 mV at 35°C. Wecould also observe kainate responses in bipolar (n=8) as well asmultipolar cells (n=7), however current amplitude was low withslow activation kinetics (data not shown).

DiscussionNestin as the name spun from neuro-epithelial stem cell hasbeen considered to be an efficient marker for proliferating cellswhose expression is downregulated upon differentiation.Hence, examining the nestin intron-II-driven EGFP expressionin our ES-cell model helped not only in demarcating theprogenitor population and neural subtypes but also inunravelling the quantitative, qualitative and functionalneurogenic developmental pattern both prior and subsequent tothe onset of neurogenesis.

Page 10: Nestin transgenic ES cells and in vitro neurogenesis

1480 Journal of Cell Science 115 (7)

Page 11: Nestin transgenic ES cells and in vitro neurogenesis

1481Nestin transgenic ES cells and in vitro neurogenesis

Neural ontogeny and nestin expressionNestin expression in transgenic mice was reported to occur asearly as stage E7.5 (Zimmerman et al., 1994; Lothian andLendahl, 1997; Josephson et al., 1998; Lothian et al., 1999;

Yaworsky and Kappen, 1999), although it remains unclearwhether or not it was investigated at earlier stages. In thetransgenic ES-cell model and in murine blastocysts wedemonstrated that nestin was expressed before lineagespecification. The reported cDNA sequence in the GenBankdatabase from 3.5 dpc murine blastocyst (accession no.C78523) having sequence similarity to mouse, rat, and humannestin genes corroborated our findings on the presence ofnestin transcript in vivo. Similar observations on earlyexpression of other tissue-specific cytoskeletal proteins havealso been described (Bain et al., 1995). Recently, Streit et al.reported that neural induction in chick embryo occurred beforegastrulation, indirectly implicating that nestin, being a neuralstem cell marker, is probably evolutionarily conserved andmight express prior to gastrulation (Streit et al., 2000).However, the role of nestin, a cytoskeletal protein belonging tothe intermediate filament family, prior to the onset ofneurulation remains to be determined. It is possible that nestin,being one of the earliest expressing cytoskeletal proteins, isrequired for laying a strong cytoarchitectural foundation duringearly embryonic development.

Promoter/enhancer targeted neural specificationPrevious studies from our group had already demonstrated theefficacy of promoter-mediated targeting in the ES-cell system

for a better understanding of cardiomyogenesis (Kolossov etal., 1998). The current investigation allowed us to exploredetails of in vitro neurogenesis. Previously, StevenGoldman’s group used a promoter-targeted and EGFP-reporter-based transient expression system in primarycultures for neural cell type selection and isolation (Wang etal., 1998; Roy et al., 2000a; Roy et al., 2000b). However,here we took advantage of the versatility of transgenic nestinES-cell clones combined with stable EGFP expression. Theimmunocytochemical, electrophysiological and time-lapseobservations provided direct proof of the neural lineageconfinement of nestin intron-II-driven EGFP expression in

Fig. 7.The functional characterization of EGFP-positive cellsdemonstrated a development-dependent pattern of ion channelexpression. Ramp depolarizations (150 ms from –100 mV to 50mV, Hp –80 mV) showed that at the apolar stage (A) no voltage-dependent ion channels were detected, whereas unipolar stages (B)expressed voltage-activated outward currents. Starting from thebipolar stage, TTX hypersensitive INa was detected (C,D). In theearly developmental stage (EDS), bipolar neurons INa weredetected first, whereas IBa appeared at later stages (E). While INadensity in bipolar neurons increased (P=0.01) during furtherdevelopment (F), the mean current density in multipolar neuronswas less (P=0.05) than that of bipolar neurons (G).

Fig. 6.The quantitative evaluation of in vitro neurogenesis usingflow-cytometry revealed a development-dependent distributionpattern. A clear delay in the generation of strong EGFP-positiveprogenitors was observed in the retinoic acid untreated (right panels)versus the treated (left panels) EBs. The maximal generation ofprogenitors occurred between d4 and d10 after plating in the RA-treated cells. Notice the decline in the number of EGFP-positive cellsaccompanied by a compensatory increase in EGFP-negative cells,suggesting a transition to the differentiated neuronal population.While the population of undifferentiated D3 ES cells (A) was usedfor calibration of the system (estimate of the range of auto-fluorescence), a relatively homogenous EGFP expression was noticedin the undifferentiated h-nestin-EGFP cells (B). The EGFPfluorescence intensity on the horizontal axis is displayed as log scale;the vertical axis represents the percentage of gated cells. Numbers inthe top-left corner indicate the differentiation stage. (B) Aquantitative evaluation of the in vitro neurogenesis was obtained byaveraging two to three experiments per time point of differentiation.This analysis clearly showed that the decline in the strong EGFP-positive (EGFP+++ and EGFP++) and weak EGFP-positive (EGFP+)cell population was accompanied by a parallel increase in EGFP-negative (EGFP–) cells.

Page 12: Nestin transgenic ES cells and in vitro neurogenesis

1482

our pluripotent transgenic ES-cell clones. The absence ofEGFP expression in the spontaneously beating clusters (datanot shown) generated from the differentiating EBs followingthe cardiac differentiation protocol (Kolossov et al., 1998) infact supported this. Additionally, the flow-cytometry-assistedquantification of stably expressed nestin-driven EGFPprovided insight to demarcating the time and development-dependent neural progenitor population induction.

Neural differentiation and neurogenic quantificationBased on the qualitative microscopic observation, time-lapsemonitoring and concurrent quantitative FACS, two distincttemporal nestin induction patterns (lineage– and lineage+), asdiscerned by EGFP expression, were revealed. The biphasicEGFP expression pattern indeed indicated the decrease inuncommitted ES cells upon differentiation by LIF withdrawaland cell aggregation and concomitant increase in lineage-

specified neural progenitors. This indicates that a critical timewindow exists during this one week regimen for the neural cellfate decision. Hence, selecting these population of nestin-driven, EGFP-expressing ES cells and EB-derived cellsthrough FACS, and conditioning them to a selective lineagesuch as neurons, astrocytes and oligodendrocytes would furtherprovide us with useful information regarding the guiding cuesprior to lineage commitment and specification. Indeed, thetime-lapse monitoring of whole EBs (data not shown) aswell as of isolated cells unequivocally demonstrated theproliferation, migration and differentiation of EGFP-positiveneural progenitors into neurons and glia. Thus, further study inthis regard could answer the critical question whether neuronsand glia are generated from a common progenitor and theexisting crosstalk between these two cell types, as proposed ina number of studies (Tamada et al., 1998; Vernadakis, 1996).

The time course in neurogenic progression revealed thatthe maximum number of neural progenitors was generated

Journal of Cell Science 115 (7)

Fig. 8.Functional expressionof IBa during early stages ofneuronal development. (A) TheIBa current-voltage relationship(voltage steps lasting for 150milliseconds, from –60 to +50mV in 10 mV increments, HP–80 mV) in a representativebipolar LDS neuronal cell. Thethreshold of activation wasbetween –60 and –50 mV andpeak currents were measured at–10 mV. (B) The fraction ofdifferent IBa subtypescomposing the whole cellcurrent was evaluated usingselective antagonists.(C,D) The percentage of thedifferent current componentsdid not significantly differbetween bipolar- andmultipolar neurons.

Page 13: Nestin transgenic ES cells and in vitro neurogenesis

1483Nestin transgenic ES cells and in vitro neurogenesis

between 4 and 12 days after EB plating. Accordingly, we couldsubdivide the transgenic EB-derived cells broadly into threegroups. The brightest shining, EGFP-positive cells on platedEBs were categorized into group I, which included the sub-population of mitotically active neural progenitors as revealedby nestin immunostaining and semi-quantitative fluorescencedetection in isolated cells. Group II, with medium and lowEGFP, included the population of weak shining bi- andmultipolar neurons as well as glia. EGFP-negative cells werecategorized under group III, which included the moredifferentiated neurons and glia along with other non-neural celltypes. Hence, the overall heterogeneity in the full length humannestin intron-II-enhancer-driven EGFP expression in ES cellsupon differentiation reflects clearly the endogenous nestinexpression as reported (Yaworsky and Kappen, 1999) and the

asynchrony in neural progenitor generation. Although thesignificance of this heterogeneity is not well understood, thetemporal- and region-dependent differential expression ofspecific transcription factors leading to neural stem cellheterogeneity might be the causal basis (Yaworsky andKappen, 1999).

Ion channel expression during neural developmentVoltage-dependent ion channels were acclaimed to be cellularfingerprints for the properties and patterns of neuronal cellsduring differentiation (Takahashi and Okamura, 1998). Inearlier studies on the whole EB (Strubing et al., 1995; Arnholdet al., 2000), only cells with distinct neuronal morphologylocated at the periphery could be characterized. We have takenadvantage of single cell isolation and EGFP labelling for theidentification and functional characterization; this allowed easydetection of neural progenitors and unipolar neurons that wereotherwise almost impossible to discern from non-neural celltypes. Based on morphology and EGFP intensity we coulddistinguish four populations of neuronal cells with differentelectrophysiological characteristics. The undifferentiatedneural progenitors did not display voltage-dependent ionchannels, whereas unipolar neurons were found to expressvoltage-dependent 4-AP sensitive K+ channels. By contrast,differentiated neurons such as bipolar and multipolar neuronsexpressed voltage-dependent K+ and Ca2+, as well as TTX-sensitive neuronal Na+ channels. Earlier reports (Barish, 1991;Grantyn et al., 1989; Gottmann et al., 1989) in culturedneuronal precursors suggested the expression of low-voltage-activated Ca2+ currents prior to the appearance of voltage-dependent Na+- and Ca2+ currents. However, in line with ourfindings, Bain et al. reported that, within the first four days ofplating, there was a small number of ES-cell-derived neuron-like cells that lacked voltage-activated inward currents,although their differentiation state was not defined (Bain et al.,1995).

The cell-type-specific ion channel expression might servespecific functions during neuronal maturation. It is possible thatneurite outgrowth in conjunction with inter-synapticconnections determine the expression of voltage-activatedinward currents or vice versa. Neuronal Na+ channels weredetected prior to the onset of voltage-dependent Ca2+ channelsin differentiating neurons indirectly suggesting the possibleexistence of two diverse (lagging and leading in terms of ionchannel expression), ES-cell-derived bipolar neuronal subtypes.Similar differences in the inception of ICa expression betweencultured chick sensory and autonomic neuronal precursors havebeen reported earlier (Gottmann et al., 1988). Since the Ca2+

entry into neuronal cells through voltage-gated Ca2+ channelsinfluences neuronal excitability as well as synaptic transmission(Augustine et al., 1987; Spitzer et al., 1994), the absence of ICain unipolar as well as some bipolar neurons at d7+3/4 suggestedthat these differentiating, relatively young, neurons had not yetestablished connection with their neighbouring counterpartsand hence lacked the functional expression of these ionchannels. Pharmacological evaluation of subtypes of voltage-dependent Ca2+ currents in differentiated neurons revealed thatthese already expressed N- and P/Q-type Ca2+ currents thatwere expected to be found in neuronal and neuroendocrine cells(Scherubl et al., 1993). In addition, receptor-operated channels

Fig. 9.Functional expression of receptor-operated channels duringearly neuronal development. GABA (A) as well as glycine (B)evoked fast-activating and inactivating inward currents (HP –80 mV)in bipolar strong EGFP-positive cells. Co-application of the selectiveantagonists Bicuculline (A) and Strychnine (B) almost completelyinhibited the activation of GABA- and glycine-evoked inwardcurrents, respectively. The ionic nature was determined byperforming ramp depolarizations (from –100 mV to 50 mV in 150milliseconds) in the presence and absence of the agonist yielding areversal potential of –27 mV, as expected for a Cl– current (see insetin A).

Page 14: Nestin transgenic ES cells and in vitro neurogenesis

1484

prevalently of the inhibitory type were detected in bipolar andmultipolar neurons. Taken together, the expression of functionalion channels in the developing neurons from ES cells in vitroseems to be related not to the stage (day after plating) alone, asproposed earlier (Bain et al., 1995; Strubing et al., 1995), butprimarily to the morphology. This demonstrates that ion channelexpression parallels closely the cellular functional demandsduring neurogenesis.

Thus, the use of live reporter EGFP under the regulatorycontrol of the neural-specific enhancer nestin has given us themeans to investigate not only its expression, but alsoquantitative- and functional characteristics of neurogenesisduring early stages of development. This system would help toincrease the diversity of self-renewing and multipotent neuralprogenitors by using fluorescence activated cell sorting(FACS) and, consequently, allow their use in experimentaltransplantation purposes.

We thank U. Lendahl (Karolinska Institute, Sweden) for kindlyproviding the nestin promoter construct, A. Wobus for D3 ES cells,H. Bohlen and O. Manzke (University of Cologne, Germany) forhelping with flow-cytometry analysis, N. Smyth (University ofCologne, Germany) for providing the murine blastocysts, and S.Ullrich, E. Kolossov, C. Andressen, S. Arnhold and W. Bloch(University of Cologne) for critically reading the manuscript. Thisstudy was supported by a grant from BMBF (0311474/6) to J.H.

ReferencesArnhold, S., Andressen, C., Angelov, D. N., Vajna, R., Volsen, S. G.,

Hescheler, J. and Addicks, K. (2000). Embryonic stem-cell derivedneurones express a maturation dependent pattern of voltage-gatedcalcium channels and calcium-binding proteins. Int. J. Dev. Neurosci.18,201-212.

Augustine, G. J., Charlton, M. P. and Smith, S. J. (1987). Calcium actionin synaptic transmitter release. Annu. Rev. Neurosci. 10, 633-693.

Bain, G., Kitchens, D., Yao, M., Huettner, J. E. and Gottlieb, D. I. (1995).Embryonic stem cells express neuronal properties in vitro. Dev. Biol. 168,342-357.

Barish, M. E. (1991). Voltage-gated calcium currents in cultured embryonicXenopus spinal neurones. J. Physiol. 444, 523-543.

Bjornson, C. R., Rietze, R. L., Reynolds, B. A., Magli, M. C. and Vescovi,A. L. (1999). Turning brain into blood: a hematopoietic fate adopted byadult neural stem cells in vivo. Science 283, 534-537.

Cattaneo, E. and McKay, R. (1990). Proliferation and differentiation ofneuronal stem cells regulated by nerve growth factor. Nature347, 762-765.

Doetsch, F., Caille, I., Lim, D. A., Garcia-Verdugo, J. M. and Alvarez-Buylla, A. (1999). Subventricular zone astrocytes are neural stem cells inthe adult mammalian brain. Cell 97, 703-716.

Gage, F. H. (2000). Mammalian neural stem cells. Science287, 1433-1438.Gottmann, K., Dietzel, I. D., Lux, H. D., Huck, S. and Rohrer, H. (1988).

Development of inward currents in chick sensory and autonomic neuronalprecursor cells in culture. J. Neurosci. 8, 3722-3732.

Gottmann, K., Dietzel, I. D., Lux, H. D. and Ruedel, C. (1989). Proton-induced Na+ current develops prior to voltage-dependent Na+ and Ca2+

currents in neuronal precursor cells from chick dorsal root ganglion.Neurosci. Lett. 99, 90-94.

Grantyn, R., Perouansky, M., Rodriguez-Tebar, A. and Lux, H. D. (1989).Expression of depolarizing vol. Brain Res. Dev. Brain Res. 49, 150-155.

Hamill, O. P., Marty, A., Neher, E., Sakmann, B. and Sigworth, F. J. (1981).Improved patch-clamp techniques for high-resolution current recordingfrom cells and cell-free membrane patches. Pflugers Arch. 391, 85-100.

Hockfield, S. and McKay, R. D. (1985). Identification of major cell classesin the developing mammalian nervous system. J. Neurosci. 5, 3310-3328.

Johansson, B. M. and Wiles, M. V. (1995). Evidence for involvement ofactivin A and bone morphogenetic protein 4 in mammalian mesoderm andhematopoietic development. Mol. Cell. Biol. 15, 141-151.

Josephson, R., Muller, T., Pickel, J., Okabe, S., Reynolds, K., Turner, P.A., Zimmer, A. and McKay, R. D. (1998). POU transcription factors

control expression of CNS stem cell-specific genes. Development125, 3087-3100.

Kolossov, E., Fleischmann, B. K., Liu, Q., Bloch, W., Viatchenko-Karpinski, S., Manzke, O., Ji, G. J., Bohlen, H., Addicks, K. andHescheler, J. (1998). Functional characteristics of ES cell-derived cardiacprecursor cells identified by tissue-specific expression of the greenfluorescent protein. J. Cell Biol. 143, 2045-2056.

Krum, J. M. and Rosenstein, J. M. (1999). Transient coexpression of nestin,GFAP, and vascular endothelial growth factor in mature reactive astrogliafollowing neural grafting or brain wounds. Exp. Neurol. 160, 348-360.

Laywell, E. D., Dorries, U., Bartsch, U., Faissner, A., Schachner, M. andSteindler, D. A. (1992). Enhanced expression of the developmentallyregulated extracellular matrix molecule tenascin following adult braininjury. Proc. Natl. Acad. Sci. USA 89, 2634-2638.

Lee, S. H., Lumelsky, N., Studer, L., Auerbach, J. M. and McKay, R. D.(2000). Efficient generation of midbrain and hindbrain neurons from mouseembryonic stem cells. Nat. Biotech. 18, 675-679.

Lendahl, U. (1997). Transgenic analysis of central nervous system developmentand regeneration. Acta Anaesthesiol. Scand. Suppl. 110, 116-118.

Lendahl, U., Zimmerman, L. B. and McKay, R. D. (1990). CNS stem cellsexpress a new class of intermediate filament protein. Cell 60, 585-595.

Lothian, C. and Lendahl, U. (1997). An evolutionarily conserved region inthe second intron of the human nestin gene directs gene expression to CNSprogenitor cells and to early neural crest cells. Eur. J. Neurosci. 9, 452-462.

Lothian, C., Prakash, N., Lendahl, U. and Wahlstrom, G. M. (1999).Identification of both general and region-specific embryonic CNS enhancerelements in the nestin promoter. Exp. Cell Res. 248, 509-519.

Maltsev, V. A., Wobus, A. M., Rohwedel, J., Bader, M. and Hescheler, J.(1994). Cardiomyocytes differentiated in vitro from embryonic stem cellsdevelopmentally express cardiac-specific genes and ionic currents. Circ.Res. 75, 233-244.

Messam, C. A., Hou, J. and Major, E. O. (2000). Coexpression of nestin inneural and glial cells in the developing human CNS defined by a human-specific anti-nestin antibody. Exp. Neurol. 161, 585-596.

Mujtaba, T., Mayer-Proschel, M. and Rao, M. S. (1998). A common neuralprogenitor for the CNS and PNS. Dev. Biol. 200, 1-15.

Namiki, J. and Tator, C. H. (1999). Cell proliferation and nestin expressionin the ependyma of the adult rat spinal cord after injury. J. Neuropathol.Exp. Neurol. 58, 489-498.

Okabe, S., Forsberg-Nilsson, K., Spiro, A. C., Segal, M. and McKay, R. D.(1996). Development of neuronal precursor cells and functional postmitoticneurons from embryonic stem cells in vitro. Mech. Dev. 59, 89-102.

Pekny, M., Johansson, C. B., Eliasson, C., Stakeberg, J., Wallen, A.,Perlmann, T., Lendahl, U., Betsholtz, C., Berthold, C. H. and Frisen, J.(1999). Abnormal reaction to central nervous system injury in mice lackingglial fibrillary acidic protein and vimentin. J. Cell Biol. 145, 503-514.

Reynolds, B. A. and Weiss, S. (1992). Generation of neurons and astrocytesfrom isolated cells of the adult mammalian central nervous system. Science255, 1707-1710.

Rohwedel, J., Guan, K. and Wobus, A. M. (1999). Induction of cellulardifferentiation by retinoic acid in vitro. Cells Tissues Organs 165, 190-202.

Rossi, F., Bravin, M., Buffo, A., Fronte, M., Savio, T. and Strata, P. (1997).Intrinsic properties and environmental factors in the regeneration of adultcerebellar axons. Prog. Brain Res. 114, 283-296.

Roy, N. S., Benraiss, A., Wang, S., Fraser, R. A., Goodman, R., Couldwell,W. T., Nedergaard, M., Kawaguchi, A., Okano, H. and Goldman, S. A.(2000a) Promoter-targeted selection and isolation of neural progenitor cellsfrom the adult human ventricular zone. J. Neurosci. Res. 59, 321-331.

Roy, N. S., Wang, S., Jiang, L., Kang, J., Benraiss, A., Harrison-Restelli,C., Fraser, R. A., Couldwell, W. T., Kawaguchi, A., Okano, H. et al.(2000b) In vitro neurogenesis by progenitor cells isolated from the adulthuman hippocampus. Nat. Med. 6, 271-277.

Scherubl, H., Kleppisch, T., Zink, A., Raue, F., Krautwurst, D. andHescheler, J. (1993). Major role of dihydropyridine-sensitive Ca2+ channelsin Ca2+-induced calcitonin secretion. Am. J. Physiol. 264, E354-E360.

Spitzer, N. C., Gu, X. and Olson, E. (1994). Action potentials, calciumtransients and the control of differentiation of excitable cells. Curr. Opin.Neurobiol. 4, 70-77.

Streit, A., Berliner, A. J., Papanayotou, C., Sirulnik, A. and Stern, C. D.(2000). Initiation of neural induction by FGF signalling before gastrulation.Nature 406, 74-78.

Strubing, C., Ahnert-Hilger, G., Shan, J., Wiedenmann, B., Hescheler, J.and Wobus, A. M. (1995). Differentiation of pluripotent embryonic stem

Journal of Cell Science 115 (7)

Page 15: Nestin transgenic ES cells and in vitro neurogenesis

1485Nestin transgenic ES cells and in vitro neurogenesis

cells into the neuronal lineage in vitro gives rise to mature inhibitory andexcitatory neurons. Mech. Dev. 53, 275-287.

Takahashi, K. and Okamura, Y. (1998). Ion channels and early developmentof neural cells. Physiol. Rev. 78, 307-337.

Tamada, Y., Tanaka, M., Munekawa, K., Hayashi, S., Okamura, H., Kubo,T., Hisa, Y. and Ibata, Y. (1998). Neuron-glia interaction in thesuprachiasmatic nucleus: a double labeling light and electron microscopicimmunocytochemical study in the rat. Brain Res. Bull. 45, 281-287.

van der Kooy, D. and Weiss, S. (2000). Why stem cells? Science287, 1439-1441.

Vernadakis, A. (1996). Glia-neuron intercommunications and synapticplasticity. Prog. Neurobiol. 49, 185-214.

Wang, S., Wu, H., Jiang, J., Delohery, T. M., Isdell, F. and Goldman, S. A.

(1998). Isolation of neuronal precursors by sorting embryonic forebraintransfected with GFP regulated by the T alpha 1 tubulin promoter. Nat.Biotechnol. 16, 196-201.

Wobus, A. M., Wallukat, G. and Hescheler, J. (1991). Pluripotent mouseembryonic stem cells are able to differentiate into cardiomyocytesexpressing chronotropic responses to adrenergic and cholinergic agents andCa2+ channel blockers. Differentiation48, 173-182.

Yaworsky, P. J. and Kappen, C. (1999). Heterogeneity of neural progenitorcells revealed by enhancers in the nestin gene. Dev. Biol. 205, 309-321.

Zimmerman, L., Parr, B., Lendahl, U., Cunningham, M., McKay, R.,Gavin, B., Mann, J., Vassileva, G. and McMahon, A. (1994). Independentregulatory elements in the nestin gene direct transgene expression to neuralstem cells or muscle precursors. Neuron 12, 11-24.


Recommended