Download Nestin Is Required for the Proper SelfRenewal of Neural Stem Cells

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Signal transduction wikipedia , lookup

Cell growth wikipedia , lookup

Cell cycle wikipedia , lookup

Apoptosis wikipedia , lookup

Extracellular matrix wikipedia , lookup

Cytokinesis wikipedia , lookup

Tissue engineering wikipedia , lookup

Mitosis wikipedia , lookup

SULF1 wikipedia , lookup

Cell encapsulation wikipedia , lookup

Cell culture wikipedia , lookup

Programmed cell death wikipedia , lookup

Organ-on-a-chip wikipedia , lookup

Cellular differentiation wikipedia , lookup

List of types of proteins wikipedia , lookup

Amitosis wikipedia , lookup

Transcript
TISSUE-SPECIFIC STEM CELLS
Nestin Is Required for the Proper Self-Renewal of Neural Stem Cells
DONGHYUN PARK,a,c ANDY PENG XIANG,b FRANK FUXIANG MAO,a,b LI ZHANG,a CHUN-GUANG DI,b XIAO-MEI LIU,b
YUAN SHAO,b BAO-FENG MA,b JAE-HYUN LEE,a KWON-SOO HA,c NOAH WALTON,a BRUCE T. LAHNa,b
a
Department of Human Genetics, University of Chicago, Howard Hughes Medical Institute, Chicago, Illinois,
USA; bCenter for Stem Cell Biology and Tissue Engineering, Sun Yat-sen University, The Key Laboratory for
Stem Cells and Tissue Engineering, Ministry of Education, Guangzhou, People’s Republic of China; cDepartment
of Molecular and Cellular Biochemistry and Vascular System Research Center, Kangwon National University
School of Medicine, Kangwon-do, Republic of Korea
Key Words. Nestin • Intermediate filament • Apoptosis • Neural stem cells
ABSTRACT
The intermediate filament protein, nestin, is a widely
employed marker of multipotent neural stem cells (NSCs).
Recent in vitro studies have implicated nestin in a number
of cellular processes, but there is no data yet on its in vivo
function. Here, we report the construction and functional
characterization of Nestin knockout mice. We found that
these mice show embryonic lethality, with neuroepithelium
of the developing neural tube exhibiting significantly fewer
NSCs and much higher levels of apoptosis. Consistent with
this in vivo observation, NSC cultures derived from
knockout embryos show dramatically reduced self-renewal
ability that is associated with elevated apoptosis but no
overt defects in cell proliferation or differentiation. Unexpectedly, nestin deficiency has no detectable effect on the
integrity of the cytoskeleton. Furthermore, the knockout
of Vimentin, which abolishes nestin’s ability to polymerize
into intermediate filaments in NSCs, does not lead to any
apoptotic phenotype. These data demonstrate that nestin is
important for the proper survival and self-renewal of
NSCs, and that this function is surprisingly uncoupled
from nestin’s structural involvement in the cytoskeleton.
STEM CELLS 2010;28:2162–2171
Disclosure of potential conflicts of interest is found at the end of this article.
INTRODUCTION
The cytoskeleton of eukaryotic cells is composed of three
major types of filamentous structures: actin microfilaments,
microtubules, and intermediate filaments (IFs). As its name
implies, IF is intermediate in diameter and can be assembled
from a large family of IF proteins [1]. One member of the IF
protein family is nestin, which is found only in vertebrates
thus far. Since its identification, nestin has been unequivocally
accepted as a marker of NSCs both during embryonic development and in the adult brain [2, 3]. Numerous in vivo and in
vitro studies now rely on nestin expression to track the proliferation, migration, and differentiation of NSCs.
During embryogenesis, nestin expression can also be
found in tissues outside of the central nervous system (CNS),
especially the developing muscle [4]. Interestingly, most nestin-positive cells in early development are stem/progenitor
populations engaged in active proliferation [5]. Once these
cells become differentiated and cease to divide, nestin expres-
sion is downregulated, often with the concomitant upregulation of other tissue-specific IF proteins such as glial fibrillary
acidic protein (GFAP) in astrocytes, a-internexin, and neurofilaments in neurons and desmin in muscle [5].
In the adult, nestin-expressing cells are found frequently
(though not necessarily exclusively) in areas of regeneration,
where they might function as a reservoir of stem/progenitor
cells capable of proliferation and differentiation. The correlation of nestin expression with cell proliferation is also
observed in neoplastic transformation. For example, abundant
nestin expression was found in several cancers such as neuroblastoma, glioma, and melanoma and higher levels of expression seem to correlate with greater malignancy [6, 7].
The mammalian IF protein family comprises 60–70 members, which are classified into several types based on expression
patterns and biochemical properties [1]. Nestin, a type VI IF
protein, is known to interact preferentially with several type III
and type IV IF proteins such as vimentin and a-internexin [8].
Evidence indicates that unlike most other IF proteins, nestin is
not able to polymerize by itself, but instead incorporates into
Author contributions: D.P. and A.P.X.: conception and design, provision of study material, collection and assembly of data, data
analysis and interpretation, manuscript writing, final approval of manuscript; F.F.M., L.Z., X.-M.L., Y.S., B.-F.M., C.-G.D., and J.H.L.:
provision of study material, collection and assembly of data, data analysis and interpretation; K.-S.H.: provision of study material; N.W.:
manuscript writing; B.T.L.: conception and design, provision of study material, data analysis and interpretation, manuscript writing, final
approval of manuscript. D.P. and A.P.X. contributed equally to this article.
Correspondence: Donghyun Park, Ph.D., e-mail: [email protected]; or Andy Peng Xiang, Ph.D., e-mail: [email protected].
cn; or Bruce T. Lahn. e-mail: [email protected] Received November 5, 2009; accepted for publication August 27, 2010; first
published online in Stem Cells Express October 20, 2010; available online without subscription through the open access option.
C AlphaMed Press 1066-5099/2010/$30.00/0 doi: 10.1002/stem.541
V
STEM CELLS 2010;28:2162–2171 www.StemCells.com
Park, Xiang, Mao et al.
2163
the IF network by copolymerizing with other type III or IV IF
proteins [8, 9]. The inability of nestin to polymerize by itself is
presumably due to a very short N-terminal ‘‘head’’ domain that,
in other IF proteins, is essential for filament assembly. Nestin is
depolymerized during mitosis and is reincorporated into the IF
network in G1 phase. The depolymerization of nestin has been
suggested to be regulated by CDC2-mediated phosphorylation
[10]. During mitosis, nestin was shown to promote the phosphorylation-dependent disassembly of vimentin [11]. Thus, nestin
has been suggested to take part in coordinating changes in the
IF network within actively dividing cells, though the biological
significance of these regulations is not clear.
The in vivo physiological function of nestin is essentially
unknown. Evidence from in vitro experiments suggests that
nestin plays a role in promoting cell survival and proliferation. Knockdown of nestin reduced cell growth in cultured
neuroblastoma and astrocytoma cells [12, 13] and sensitized
the immortalized neuronal precursor cell line ST15A to H2O2induced apoptosis [14]. On the other hand, ectopic overexpression of nestin provided cytoprotective effect on ST15A
against H2O2 treatment [14].
To examine the in vivo function of nestin, we generated
nestin-deficient mice. In contrast to the overtly normal phenotypes of mice lacking type III or IV IF proteins [15–24], nestin deficiency causes embryonic lethality. We further show
that nestin is important for the survival and self-renewal of
NSCs during embryonic development. Surprisingly, we found
that this function of nestin does not require its structural
incorporation into the cytoskeleton.
MATERIALS
AND
METHODS
Animals and Cell Cultures
Nes genomic region encompassing coding region of exon 1
was deleted by homologous recombination The targeting vector has a drug selection cassette containing the puromycin resistance gene fused to the DTK gene [25], which is driven by
mouse Pgk promoter and followed by the bovine growth hormone polyadenylation signal. This cassette is flanked in the
50 by a 1,033-bp fragment corresponding to genomic positions 1033 to 1 of the mouse Nes locus, and in the 30 by
a 6,054-bp fragment corresponding to genomic positions
1,003–7,056 of Nes (position 1 denotes the first base of the
coding region of Nes in C57BL/6J reference sequence and
position 1 denotes the base immediately upstream of the
coding region). The targeting vector was electroporated into
the mouse 129 ES cell line ES-E14TG2a (American Type
Culture Collection, Manassas, VA, http://www.atcc.org) followed by puromycin selection. Resistant colonies were
picked and examined by polymerase chain reaction (PCR)
using forward primer corresponding to the mouse genomic
sequence just upstream of the 50 homology region and
reverse primer corresponding to sequence in the selection
cassette of the targeting vector. Only homologous integration
events can yield PCR products of the right size. PCR-positive
colonies were further tested by Southern blotting. Targeted
ES cell clones were injected into C57BL/6J blastocysts to
produce chimeric animals; the Nes knockout allele was bred
into C57BL/6J background; and animals were maintained and
crossed using standard procedures (Cyagen Biosciences,
Guangzhou, China, http://www.cyagen.com/en).
Cell Cultures
NSC cultures were derived and maintained as described previously with minor modification [26]. Briefly, mouse embryos
www.StemCells.com
were removed at embryonic day (E) 11.5–12, and the neural
tube was mechanically dissected from surrounding tissue. Isolated neural tube was incubated in 0.05% trypsin for 10
minutes with gentle trituration to dissociate cells, followed by
plating on dishes coated with poly-L-lyine and laminin. Cells
were maintained in NEP basal medium supplemented with
bFGF (20 ng/ml) and EGF (20 ng/ml). The NEP basal medium consisted of Dulbecco’s modified Eagle’s medium-F12
(Invitrogen, Carlsbad, CA, http://www.invitrogen.com/site/us/
en/home) supplemented with additives as described [27]. For
inhibition of caspase activity, the pan-caspase inhibitor zVAD (Promega, Madison, WI, http://www.promega.com) was
added to the media at 20 lM. For the disruption of microtubules, NSCs were treated with 0.2 or 1 lM of colchicine for
30 minutes before microtubule staining. For the disruption of
actin microfilaments, cells were treated with 20 or 100 nM
cytochalasin D for 30 minutes before microfilament staining.
For cyclin-dependent kinase (CDK) inhibition, cells were
treated with various concentrations of roscovitine (a CDK1/2/
5 inhibitor) or 3-amino-1H pyrazolo[3,4-b]quinoxaline (a
CDK1/5 inhibitor) for 24 hours, which were the maximal levels of treatment without significantly increasing caspase-3/7
activity in wild-type NSCs. For clonogenic neurosphere-forming assay, dissociated cells from passage 2 neurospheres were
plated in 96-well plates at an average density of one cell per
well. All wells were visually inspected under the microscope
to identify wells that contain only one cell each. Clonogenic
ability was calculated as the percentage of the one-celled
wells that produced neurospheres in 7 days. The experiment
was carried out three times, each with two 96-well plates, to
produce mean 6 SD.
Fluorescence Staining
Embryos were collected and fixed with 4% paraformaldehyde
(PFA) and used to obtain 16-lm sections for staining. Cell
cultures were also fixed with 4% PFA for staining. Tissue
sections or cell cultures were incubated with the following
primary antibodies: mouse mAb against nestin (Chemicon,
Billerica, MA, http://www.millipore.com, 1:200 dilution), rabbit pAb against vimentin (Abcam, Cambridge, MA, http://
www.abcam.com, prediluted), rabbit pAb against neurofilament light chain (NF-L; Abcam, 1:500 dilution), mouse mAB
against Hu (Molecular probes, Carlsbad, CA, http://www.
invitrogen.com/site/us/en/home, 1:20 dilution), rabbit pAb
against active caspase-3 (aCASP3; Promega, 1:200 dilution),
mouse mAb against cytochrome c (BD Biosciences, Franklin
Lakes, NJ, http://www.bdbiosciences.com, 1:100 dilution),
rabbit pAb against Ki67 (Abcam, 1:200 dilution), mouse mAb
against SOX2 (R&D Systems, Minneapolis, MN, http://
www.rndsystems.com, 1:125 dilution), rabbit pAb against
GFAP (Dako, Carpinteria, CA, http://www.dako.com, 1:500
dilution), and rabbit pAb against phosphorylated histone H3
(Upstate, Billerica, MA, http://www.millipore.com, 1:50 dilution). This was followed by incubation with FITC, rhodamine,
or Cy5-conjugated secondary antibodies (Jackson ImmunoResearch, West Grove, PA, http://www.jacksonimmuno.com/
home.asp). For detecting caspase activity in situ, 10 lM
FITC-VAD-FMK (Promega) was added to culture media, followed by fixation with 4% PFA in PBS. For detecting phosphatidylserine, FITC-conjugated annexin V (BD Biosciences,
Franklin Lakes, NJ, http://www.bdbiosciences.com, 1:20 dilution) was added to culture media without fixation. Nuclei
were counterstained with DAPI (Sigma, Saint Louis, MO,
http://www.sigmaaldrich.com) or Hoechst 33342 (Sigma). For
visualizing actin microfilaments, fixed cells were stained with
rhodamine-conjugated phalloidin (Invitrogen, 1:250 dilution).
Fluorescence TUNEL labeling was performed with DeadEnd
2164
Fluoreometric TUNEL System (Promega) according to vendor’s instructions. To obtain apoptotic index for each embryo,
a set of 15–20 sections of 16-lm thickness at 200 lm
intervals were used for fluorescence TUNEL staining. Signal
intensity in each pixel over twice the background level was
integrated in neural tube areas of the sections. This was
divided by total area of the region of interest. Three embryos
for each genotype were analyzed and averaged, and the value
was multiplied by an arbitrary scaling factor to set the mean
of the wild type to unit. Mitotic index was obtained from
phospho-H3 fluorescence staining in the same way as apoptotic index. To perform quantitation of SOX2 and Ki67 staining
in sections, SOX2- or Ki67-positive cells were counted and
scaled to the number of total cells based on DAPI (blue)
staining of nuclei. Additionally, SOX2 and Ki67 double-positive cells were counted and scaled to the number of SOX2positive cells. More than five frozen sections were analyzed
per embryo, with two embryos analyzed and averaged for
each genotype. Confocal imaging was performed. Quantitation
of aCASP3 and SOX2 staining in sections was performed in
the same manner.
BrdU Labeling
For in vivo BrdU labeling, pregnant mice were i.p. injected
with 100 mg BrdU (Sigma) per kg body weight, and the
embryos were harvested 6 hours after injection. For in vitro
BrdU labeling, cultured cells were treated with 10 lM BrdU,
and cells were fixed and examined 2 hours after treatment.
BrdU detection was performed with anti-BrdU antibody
(Roche Applied Science, http://www.roche-applied-science.
com/index.jsp, 6 lg/ml final concentration) according to vendor’s instructions. BrdU-positive cells were counted against
nuclei counterstained with hematoxylin. To perform quantitation, BrdU-positive cells were counted and scaled to the number of total cells based on hematoxylin counterstaining of
nuclei. For the in vivo experiment, more than five frozen sections were analyzed for each embryo, with two embryos analyzed and averaged for each genotype. For the in vitro experiment, 10 microscopic areas were examined for each genotype
and averaged.
Quantitation of Caspase-3/7 Enzymatic
Activity Assay
A luminogenic caspase-3/7 substrate (Promega) containing the
tetrapeptide sequence DEVD was used to measure caspase-3/
7 activity. The assay was performed in 96-well format according to vendor’s instructions. Luminescence was measured on
a Safire 2 instrument (Tecan, Durham, NC, http://
www.tecan.com).
Cell Migration Assay
QCM Chemotaxis (8-lm pore size) Cell Migration Assay
(Chemicon) was used for the cell migration assay. The assay
was performed in 24-well format according to manufacturer’s instruction. The insert membrane was coated with
poly-L-lysine and laminin. NSCs were plated in the inserts
(on the top side of the membrane) at a density of 105 viable
cells per cm2. Migrated cells (cells below the membrane)
were quantified by colorimetric measurement 12 hours after
plating.
Expression Analysis
RNA was prepared from NSC cultures obtained from E11.5
embryos as described earlier. Samples were prepared and analyzed using GeneChip Mouse Genome 430 2.0 Arrays (Affymetrix, Santa Clara, CA, http://www.affymetrix.com) following vendor’s protocol. For each genotype, data were obtained
Nestin Function in Neural Stem Cells
from two independently derived cell lines, and each cell line
was subjected to microarray analysis in two separate replicate
experiments. The dChip software was employed to analyze
microarray data using default parameters.
RESULTS
Targeted Disruption of Nestin Causes
Embryonic Lethality
We used homologous recombination to generate mice lacking
exon 1 of Nes, the gene encoding nestin (Fig. 1A, 1B). In
homozygous embryos, the absence of normal Nes mRNA was
demonstrated by Northern blot using a probe against exon 4
(Fig. 1C) and RT-PCR analysis (data not shown). The absence of nestin protein was further confirmed by Western blot
using an antibody against a nestin epitope encoded in exon 4
(Fig. 1D) and immunofluorescence staining (Fig. 1E).
Intercross between heterozygotes of C57BL/6J background yielded only three homozygous pups of a total of 95,
along with 27 wild types and 65 heterozygotes. All three newborn Nes/ animals were significantly smaller than their littermates, and two of them died at around 3 weeks after birth.
One Nes/ male survived for several months but was sterile. This data indicates that Nes knockout causes embryonic
lethality with around 90% penetrance. To determine the embryonic stage at which lethality occurs, embryos from heterozygous intercrosses were removed at various stages and genotyped. Homozygotes were observed at a frequency of 25%
up to E8.5 but were slightly smaller than their littermates by
E8.5. From thereon, the percentage of observed number of
Nes/ embryos over the Mendelian expectation declined
steadily (Fig. 1F). Furthermore, the Nes/ embryos found
after E8.5 were notably smaller than their wild-type littermates, indicating developmental delay. Additionally, a fraction of Nes/ embryos after E8.5 did not display heartbeat
at the time of dissection, suggesting that they were already
dead and were in the process of being resorbed. Nesþ/ mice
were viable, fertile, and indistinguishable from their wildtype littermates at all developmental stages examined, even
though Nes mRNA expression in these animals decreased to
about half of the wild-type level (Fig. 1C). We performed
extensive histological examinations of embryos that were
alive at the time of dissection. We identified dramatic neural
tube defects (see below). We failed to notice any overt abnormalities in other organ systems except smaller size. However,
we cannot rule out the existence of histologically subtle
defects outside the CNS and their potential contribution to
lethality.
Neural Tube in Nes2/2 Embryos Shows
Extensive Cell Death
Given the abundant expression of nestin in NSCs of the
developing neural tube [2], we examined Nes/ embryos
for abnormalities in the neuroepithelium. To visualize NSCs,
we utilized transgenic animals expressing green fluorescent
protein (GFP) under the control of mouse Nes promoter/
enhancer (Nes-GFP transgene). The GFP expression pattern
in these transgenic animals closely recapitulates the endogenous Nes expression, and is therefore an excellent marker
for NSCs [28]. By crossing Nes-GFP transgenic mice with
Nesþ/ animals followed by intercross, we generated NesGFP embryos with all possible genotypes in regards to Nes.
We examined E9.5, E10.5, and E11.5 embryos and found that
in all cases, GFP signal in Nes/ individuals was
Park, Xiang, Mao et al.
2165
Figure 1. Targeted disruption of the mouse Nes gene. (A): Targeting strategy. Genomic structure of wild-type Nes locus is shown at the top.
Open boxes are UTR and closed ones are coding regions. The coding region of exon 1 plus 152 nucleotides of intron 1 were replaced in the targeting vector (middle) by a drug selection cassette. Homology arms are delimited by dashed lines connecting the wild-type locus with the corresponding regions of the targeting vector. The correctly targeted locus is shown at the bottom. The 1.1-kb BamHI-BglI fragment used as Southern
probe is indicated. B: BglI; H: HindIII. On a Southern blot prepared by BglI and HindIII double digestion, this probe should detect the extra HindIII site introduced in the selection cassette. (B): Southern blot detection of Nes genotypes. DNA was digested with BglI and HindIII and hybridized with the probe as depicted. The 9-kb band is from the wild-type allele and the 7.9-kb band is from the knockout allele. (C): Northern blot
analysis of total RNA isolated from Nesþ/þ, Nesþ/, and Nes/ embryos at E11.5. The probe corresponds to a part of Nes exon 4. The top
panel is the Northern image, with Nes/ showing no detectable Nes mRNA, whereas Nesþ/ embryos showing reduced expression. The bottom panel is ethidium bromide staining of the total RNA. (D): Western analysis of NSCs from Nesþ/þ, Nesþ/, and Nes/ genotypes. The
blot was probed with antibodies against nestin (top panel) and vimentin (bottom panel). The Nes/ sample has no nestin immunoreactivity but
no apparent change in vimentin immunoreactivity. (E): Immunofluorescence staining of forebrain sections from Nesþ/þ and Nes/ E11.5
embryos. Nestin is in green and DAPI in blue. There are no nestin-immunoreactive cells in the Nes/ section. Scale bar ¼ 100 lm. (F):
Lethality of Nes/ animals at various stages of development. Nesþ/ male and female mice were crossed, and their progeny recovered at the
developmental stages are indicated. The percent of observed over expected Nes/ individuals drops continuously as development proceeds
(expected number is calculated as one-third of the total number of Nesþ/þ and Nesþ/ animals). The numbers above each data point indicate
the number of Nes/ animals and, in parentheses, the total number of animals. Abbreviation: DAPI, 4’,6-diamidino-2-phenylindole.
dramatically reduced as compared with their littermates (Fig.
2A). The reduction of signal was largely due to much fewer
GFP-positive cells in the neural tube of Nes/ embryos.
However, there is no defect in neural tube closure.
One possible explanation for this observation is that nestin
deficiency in NSCs triggers widespread differentiation of
www.StemCells.com
these cells, resulting in a smaller pool of proliferating NSCs
in Nes/ embryos. To test this possibility, we immunostained E11.5 embryo sections for the early neuronal marker,
Hu, and the mature neuronal maker, NF-L. In both cases, positive cell populations in the neural tube were very comparable
between genotypes (Fig. 2B), indicating that nestin deficiency
2166
Nestin Function in Neural Stem Cells
Figure 2. Increased apoptosis in the neuroepithelium of Nes/ embryos. (A): Fewer neural stem cells in Nes/ embryos. The green fluorescence of E10.5 Nesþ/þ and Nes/ embryos, both carrying the Nes-GFP transgene, was imaged for the whole embryos. The Nes/ embryo
has significantly reduced signal. (B): Proper neuronal differentiation in Nes/ embryos. The dorsal spinal cord sections from E11.5 Nesþ/þ
and Nes/ embryos were immunostained for NF-L (red) and Hu (green) to identify neurons and counterstained with DAPI (blue). Staining patterns of the Nes/ embryo is similar to the Nesþ/þ embryo. Scale bar ¼ 100 lm. (C) Similar fraction of mitotic cells in the neuroepithelium
of Nesþ/þ and Nes/ embryos based on phospho-H3 immunostaining of E11.5 hindbrain sections (upper panels). (D): Comparable mitotic
index between Nesþ/þ and Nes/ E11.5 embryos (n ¼ 3 for each genotype; p > .1 by Student’s t test; see ‘‘Materials and Methods: section
for calculation of mitotic index). (E): Increased apoptosis in Nes/ neuroepithelium based on TUNEL staining of E11.5 hindbrain sections
(upper panels). In both (C) and (E), dashed lines in upper panels indicating boundaries of the neural tube based on DAPI staining in lower panels. Scale bar ¼ 500 lm. (F): Elevated apoptotic index in Nes/ E11.5 embryos (n ¼ 3 for each genotype; p < .0005 by Student’s t test; see
‘‘Materials and Methods’’ section for calculation of apoptotic index). Error bars in both (D) and (F) represent SE. Abbreviations: DAPI, 4’,6-diamidino-2-phenylindole; NF-L, neurofilament light chain.
does not overtly affect the neuronal differentiation of NSCs.
To examine glial differentiation, we performed GFAP staining
in E11.5 neural tubes and found that there is virtually no
GFAPþ cells in either genotype, consistent with the fact that
glial differentiation should not have started at this early developmental stage (Supporting Information Fig. S1).
Another possibility is that the lack of nestin causes NSCs
in Nes/ embryos to proliferate less. To test this, we performed BrdU labeling at E11.5. The proportion of BrdU-positive cells in the neural tube was comparable between genotypes (Supporting Information Fig. S2A, S2B). We further
measured the number of mitotic cells by immunostaining for
phosphorylated histone H3 (phospho-H3), a mitosis marker.
Consistent with BrdU labeling, the mitotic index based on the
fraction of cells positive for phospho-H3 did not differ significantly between genotypes (Fig. 2C, 2D). One issue here is
that the phospho-H3 staining does not differentiate between
NSCs and other cell types. To examine cell proliferation specifically in NSCs, we costained neural tube sections with antibodies against SOX2, a marker for NSCs, and Ki67, a marker
for proliferating cells. We found that among SOX2þ cells,
which are presumably NSCs, the fraction of Ki67þ cells is
comparable between wild-type and Nes/ animals (Supporting Information Fig. S3). Thus, Nes/ does not appear to
have a notable effect on NSC proliferation per second.
The third possibility is increased apoptosis in the neural
tube of Nes/ embryos. We tested this by performing
TUNEL assay in E11.5 embryos. Apoptotic cells as labeled by
TUNEL staining were far more prevalent in the neural tube of
Nes/ embryos as compared with their littermates (Fig. 2E).
The apoptotic index based on the fraction of TUNEL-positive
cells was elevated in Nes/ neural tube by about one order
of magnitude (Fig. 2F). Time course examination of this phenomenon indicates that increased apoptosis can be observed in
Nes/ animals as early as E9. Given that apoptosis is associated with the activation of caspases, we also performed immunostaining for the active form of caspase-3 (aCASP3), a marker
for apoptosis, and confirmed the TUNEL results. In the forebrain, for example, Nes/ embryos had more aCASP3-positive cells relative to their littermates, with the effect becoming
very dramatic as the embryos got older (Supporting Information Fig. S4). To examine apoptosis specifically in NSCs, we
Park, Xiang, Mao et al.
2167
Figure 3. Apoptotic phenotype of cultured Nes/ neural stem cells (NSCs). (A): Increased apoptosis in Nes/ NSCs. Cells were stained with
FITC-VAD-FMK (green) or immunostained for aCASP3 (red) to detect apoptotic cells, and counterstained with DAPI (blue). Scale bar ¼ 50 lm. (B):
Compromised mitochondrial integrity in apoptotic Nes/ NSCs. Cells were immunostained for cytochrome c (red) or stained with FITC-VAD-FMK
(green), to examine mitochondrial integrity or pan-caspase activity, respectively. Although Nes/ cells with little caspase activity maintained mitochondrial integrity (asterisks), Nes/ cells with elevated caspase activity lost mitochondrial integrity as indicated by the diffuse pattern of cytochrome c staining (arrow). Scale bar ¼ 10 lm. Abbreviations: DAPI, 4’,6-diamidino-2-phenylindole; FITC-VAD-FMK, FITC-Val-Ala-Asp-fluoro-methyl-Rerone.
costained neural tube sections for SOX2 and aCASP3. We
found that among SOX2þ cells, which are presumably NSCs,
the fraction of aCASP3þ cells is much higher in Nes/ animals than in wild-type controls (Supporting Information Fig.
S5). The increased apoptosis in NSCs in vivo is further confirmed by in vitro data (see below).
Cultured Nes2/2 NSCs Show Reduced
Self-Renewal Ability and Increased Apoptosis
To further examine the effect of nestin deficiency, we isolated
NSCs from a total of 79 embryos including nine Nes/
embryos on E11.5 and cultured them under standard monolayer
condition. We found that all Nes/ NSC cultures expanded
slowly. Consistent with in vivo data, the rate of cell division
based on BrdU labeling was not significantly different between
Nes/ cells and wild-type cells (Supporting Information Fig.
S2C, S2D), whereas several assays revealed dramatically
increased apoptosis in Nes/ cultures. We stained cells with
antibody against aCASP3 or FITC-VAD-FMK (a FITC-conjugated peptide derivative that stains activated caspases) to visualize apoptotic cells. These two markers colocalized to the
same cells and showed much greater numbers of positive cells
in Nes/ culture (6.4% 6 1.8%) as compared with the wildtype culture (1%; Fig. 3A). This was further corroborated by
staining with fluorescently labeled annexin V, a protein with
high affinity for the membrane phospholipid phosphatidylserine
translocated from the inner to the outer leaflet of the plasma
membrane in apoptotic cells (data not shown). To produce a
www.StemCells.com
more quantitative indicator of apoptosis, we measured caspase3/7 enzymatic activity in the cells. On average, the activity in
Nes/ cells is more than an order of magnitude greater than
in wild-type cells (Fig. 4D).
We tried to block cell death in Nes/ NSCs by the pancaspase inhibitor z-VAD but did not see any effect (data not
shown). Given that cell death pathways involving mitochondrial outer membrane permeabilization cannot be reversed by
the inhibition of caspases [29], this result implies that the mitochondria pathway is involved in cell death triggered by nestin deficiency. Consistent with this possibility, we found that
essentially all apoptotic cells marked by aCASP3 and FITCVAD-FMK in Nes/ NSC culture display reduced integrity
of mitochondrial membranes as visualized by an antibody
against cytochrome c (Fig. 3B).
To examine whether cell death in Nes/ NSCs is secondary to defective cell attachment or movement, we placed
cells in a Boyden chamber with 3-lm membrane pores. We
coated the bottom of the membrane with polylysine/laminin
and seeded cells on the top of the membrane. Twelve hours
after plating, we measured the number of cells that had
migrated through the pores and attached to the substrate on
the bottom side. Nestin deficiency failed to impact cellular
attachment and migration under these conditions (data not
shown), suggesting that cell death is not due to inefficient cell
attachment and/or movement.
To confirm that increased apoptosis indeed affects the
self-renewal ability of NSCs, we performed clonogenic
2168
Nestin Function in Neural Stem Cells
lethality, it remains possible that the lethality phenotype is
due to other defects in nestin knockout animals.
Nestin Deficiency Has Negligible Effect on Gene
Expression in NSCs
Figure 4. Nestin’s requirement for the survival of neural stem cells
(NSCs) is uncoupled from its incorporation into the intermediate filaments (IFs) network. (A–C): NSCs isolated from wild-type (A),
Vim/ (B), and Nes/ (C) embryos were immunostained for nestin (green) and vimentin (red) and counterstained with DAPI (blue).
Nestin becomes depolymerized in Vim/ cells (B), whereas vimentin IF appears normal in Nes/ cells (C). Scale bar ¼ 10 lm. (D):
Elevated caspase-3/7 activity in Nes/ NSCs and Nes/;Vim/
double knockout NSCs but not in Vim/ NSCs. Caspase-3/7 activity was measured by enzymatic assay and the activity in wild-type
cells was set as the unit of measurement. Three independent experiments were performed for each genotype. Error bars represent SE.
Values significantly different from the wild-type value (p < .05 by
Student’s t test) were marked by asterisks.
neurosphere-forming assay, which is a standard test of NSC
self-renewal ability [30, 31]. Fully dissociated NSCs were
plated in 96-well plates with an average seeding density of
one cell per well. Through visual inspection wells that were
seeded with one and only one cell each were identified.
Among these, the neurosphere-forming rate was zero for
Nes/ cells, while being around 21% 6 6% for wild-type
cells.
The above-mentioned in vivo and in vitro results show
that Nes/ NSCs have dramatically elevated levels of apoptosis and reduced self-renewal ability. Nes/ cells otherwise appear normal in proliferation, attachment, and movement. It can thus be concluded that nestin is required for the
proper survival and self-renewal of NSCs. One point worthy
of note is that although increased apoptosis in the neuroepithelium of Nes/ animals could be the cause of embryonic
The apoptotic phenotype of Nes/ NSCs could be due to
nestin deficiency having a direct impact on cellular physiology or it could be secondary to changes in the expression of
other genes caused by the lack of nestin. There are some reasons to suspect that nestin is involved in the regulation of
gene expression. First, mutations in type III and IV IF genes
often result in changes in expression levels of other IF or
extracellular matrix genes [18, 32, 33]. Second, nestin was
reported to interact with nuclear DNA in N-myc-amplified Ntype neuroblastoma cell lines [12], suggesting that nestin may
affect the expression of genes downstream of N-myc.
To examine how nestin deficiency affects gene expression,
we performed Affymatrix microarray analysis on Nes/ and
wild-type NSCs. Somewhat surprisingly, the effect of nestin
deficiency on the transcriptome was minimal. Indeed, the difference in gene expression profile between Nes/ samples
and wild-type samples was smaller than the variation within
either set of samples. Only 16 genes exhibited more than twofold changes between Nes/ and wild-type cells (Supporting
Information Table S1). One of these was Nes, the expression
level of which is about 40-fold less in Nes/ cells than in
wild-type cells according to the array data (note that analysis
of the array data always produces a finite fold difference even
when a gene is completely absent in one of the two samples
compared). The other 15 genes only showed twofold to threefold changes. None of these genes are related to structural
components of the cell such as cytoskeleton or extracellular
matrix and none are related to apoptosis.
Although the above-mentioned data are not informative
about nestin’s molecular function, two messages can be drawn
from the data. First, nestin is unlikely to play a major role in
the regulation of gene expression, including the expression of
cytoskeleton genes and apoptosis genes. Rather, nestin likely
participates directly in some aspect of cellular physiology that
is important for the survival of NSCs. Second, the apoptosis
program in Nes/ NSCs is initiated and executed in a manner that does not require significant changes in gene expression patterns, including the expression of apoptosis genes.
This is not entirely surprising because it is possible that
cells can be poised to undergo apoptosis, and the triggering
event can set off a rapid cascade of downstream events leading to apoptosis without the need for major changes in gene
expression [34].
Nestin’s Function in NSC Survival Is Uncoupled
from its Incorporation into the Cytoskeleton
Nestin is incapable of self-assembly due to its very short Nterminal head domain. Rather, it is thought to incorporate into
the IF network by copolymerizing with other type III or IV IF
proteins [8, 9, 35]. In astrocytes with knockout of Vim (the
gene encoding vimentin), nestin fails to polymerize, indicating
that nestin polymerization requires vimentin in this particular
cell type [9]. To investigate which IF protein(s) copolymerize
with nestin in NSCs, we first examined microarray data for
the expression of IF genes in NSCs. We found that Vim
(besides Nes) was highly expressed in NSCs, whereas the
other members of type III and IV IF genes were absent. Immunostaining results also indicated abundant expression of
vimentin (Fig. 4A) and the absence of GFAP, desmin, and
NF-L (data not shown). Furthermore, the immunostaining signals of vimentin and nestin are extensively colocalized in
Park, Xiang, Mao et al.
NSCs (Fig. 4A). These results suggest that, like in astrocytes,
vimentin is the required partner for nestin polymerization in
NSCs.
To test this, we isolated NSCs from Vim/ embryos and
performed immunostaining of vimentin and nestin. Vimentin
signal was absent from these cells as expected, whereas nestin
appears to be completely depolymerized (Fig. 4B), indicating
that vimentin is indeed an essential partner for nestin polymerization in NSCs.
As an IF protein, nestin has generally been studied in the
context of its involvement in the cytoskeleton. It is therefore
reasonable to assume that the function of nestin is dependent
on its polymerization into the IF network and that nestin’s
function should be abolished in Vim/ cells where nestin
fails to polymerize. We therefore tested apoptotic activities in
Vim/ NSCs, suspecting that these cells would phenocopy
Nes/ NSCs in that they would have elevated levels of apoptosis. To our surprise, we found that caspase-3/7 enzymatic
activity of Vim/ NSCs was not different from that of wildtype cells (Fig. 4D). In addition, no difference in cell expansion rate was found between Vim/ and wild-type NSCs.
The lack of any detectable defect in Vim/ NSCs is consistent with the fact that Vim/ mice do not display any overt
phenotype [15, 33, 36]. Thus, the function of nestin in NSC
survival and self-renewal is independent of its incorporation
into the IF network.
A previous study suggested that nestin is required for the
disassembly of the vimentin IF network during mitosis [11].
This raises a possible explanation for the apoptotic phenotype
of Nes/ NSCs, that is, vimentin IF network is defective in
Nes/ NSCs and this leads primarily to apoptosis that
would otherwise not occur even in Vim/ cells. If this
is the case, then Nes/ and Vim/ double knockout
should rescue the apoptotic phenotype seen in Nes/
cells. To address this possibility, we isolated NSCs from
Nes/;Vim/ double knockout embryos and assayed apoptosis in them. Several measures, including caspase-3/7 activity, showed that Nes/;Vim/ double knockout NSCs displayed the same apoptotic phenotype as Nes/ cells (Fig.
4D). Furthermore, vimentin staining was comparable between
Nes/ and wild-type cells throughout the cell cycle, suggesting that vimentin assembly and disassembly are not
affected by nestin deficiency (Fig. 4C). Thus, elevated apoptosis in Nes/ cells is not due to defects in the vimentin IF
network.
As nestin has been suggested to interact with not only
other IF proteins but also microtubules and actin microfilaments [35], we also examined the integrity of microtubules
and microfilaments by staining with anti-tubulin antibody and
phalloidin (which binds polymerized actin), respectively. We
did not detect any difference between Nes/ NSCs and
wild-type cells (Supporting Information Fig. S6, S7; left panels). We then treated cells with various concentrations of colchicine, which disrupts microtubule polymerization. The disturbance of microtubules in Nes/ NSCs showed a similar
dosage dependency to that of wild-type cells (Supporting Information Fig. S6). Furthermore, when microtubules were disrupted by a high concentration of colchicine (1 lM), actin polymerization was similarly increased in both genotypes
(Supporting Information Fig. S6). We also treated cells with
cytochalasin D, which disrupts actin polymerization. The disturbance of actin microfilaments did not differ between genotypes (Supporting Information Fig. S7). Furthermore, vimentin
appears to be properly polymerized to form IF network in
Nes/ NSCs (Fig. 4C). There is hence no evidence that nestin deficiency causes any defect in microfilaments, microtubules, or vimentin-based IFs of NSCs. Taken together, our
www.StemCells.com
2169
data indicate that the function of nestin in NSC survival is
uncoupled from its structural involvement in the cytoskeleton.
One recent report showed that nestin has a cytoprotective
effect against H2O2-induced apoptosis in the immortalized
neural progenitor cell line ST15A, although the knockdown of
nestin itself did not induce apoptosis in the absence of H2O2
treatment [14]. This study suggested that nestin serves as a
scaffold for CDK5 and inhibits the proapoptotic action of
CDK5 by sequestering it. We tested whether CDK5 signaling
also plays a role in the apoptotic phenotype seen in Nes/
NSCs. We tried to block CDK function in these cells with
various concentrations of roscovitine or 3-amino-1H-pyrazolo[3,4-b]quinoxaline. At 2 lM roscovitine or 100 lM 3amino-1H-pyrazolo[3,4-b]quinoxaline, we observed modest
suppression of apoptosis in Nes/ NSCs based on caspase3/7 activity (Supporting Information Fig. S8). Higher concentrations of either drug did not produce interpretable results
because they significantly raised caspase-3/7 activity in both
wild-type and Nes/ cells, presumably due to drug toxicity.
These data do not lead to a definitive interpretation, but they
do suggest some role of CDK in mediating the apoptotic phenotype of Nes/ cells.
CONCLUSION
Previous studies have generally linked nestin function to its
involvement in the cytoskeleton. Our in vivo and in vitro data
demonstrate that nestin has an important function in the survival and self-renewal of NSCs. We note that this phenotype
is unlikely due to technical artifacts unrelated to the knockout
of the nestin gene. First of all, the same knockout cassette has
been used to knock out several other genes in our lab, and we
have never observed phenotypes similar to nestin knockout,
indicating that the cassette itself is not responsible for the
nestin/ phenotype. Second, knockdown of nestin in several
nestin-expressing tumor cell lines such as glioma and melanoma results in apoptotic phenotype similar to that seen in
nestin/ NSCs (manuscript in preparation).
Importantly, nestin’s function in NSC survival and selfrenewal does not require its incorporation into the cytoskeleton in general or interaction with vimentin in particular. Furthermore, microtubules, actin microfilaments, and vimentin
network appear normal in Nes/ NSCs. There is also no
evidence that nestin plays a significant role in cell attachment
or migration. These data, although not ruling out a possible
role of nestin in the cytoskeleton, do argue strongly that the
requirement of nestin in the proper survival of NSCs in vivo
and in vitro is uncoupled from its structural involvement in
the cytoskeleton. This is rather surprising given the previous
emphasis on nestin’s cytoskeletal function.
The requirement of nestin in the proper survival of NSCs
is consistent with previous reports showing that knockdown
of Nes reduced cell expansion in cultured neuroblastoma,
astrocytoma, and mesangial cells [12, 13, 37]. One of these
reports further showed that knockdown of Nes did not affect
the migration of mesangial cells [37]. However, it is not clear
from these studies whether the slower expansion of Nes
knockdown cells is due to increased cell death or some other
reason. Our data from Nes/ NSCs indicate that nestin promotes cell expansion by supporting proper NSC survival. This
suggests that the negative effect of Nes knockdown on cell
expansion in these other cell types examined by previous
studies may also be due to the fact that nestin supports the
survival of these cell types.
Nestin Function in Neural Stem Cells
2170
There is evidence that the expression of nestin is regulated
by growth factor signaling pathways essential for cell proliferation and/or survival. PDGF, which promotes cell proliferation in mesangial cells, was shown to upregulate nestin
expression [37]. Thrombin was also reported to increase nestin expression and concomitantly stimulate the growth of radial glial cells in vitro [38]. Removal of EGF and bFGF
results in the suppression of nestin expression and also results
in the apoptosis and differentiation of NSCs [39, 40]. These
findings are consistent with a model whereby growth factors
positively regulate nestin expression and nestin in turn executes a part of the downstream effects of growth factor signaling on cell proliferation and survival.
An interesting question is whether nestin plays a role in
the survival of cell types other than embryonic NSCs, such as
adult NSCs and other stem cells in which nestin is expressed.
Given suggestive evidence on the role of nestin in the survival of a few other cell types [12, 13, 37] and given the fact
that cells that strongly express nestin are generally proliferating cells, we suspect that nestin may contribute to the survival
and self-renewal of a number of other rapidly proliferating
stem cell types besides embryonic NSCs.
How does one reconcile nestin’s well-known role as a
major cytoskeletal IF protein with our data showing that nestin’s function in NSC survival does not require its incorporation into the cytoskeleton? We propose two possibilities. In
the first possibility, nestin’s sole function is to support the
survival of NSCs. Nestin’s incorporation into the cytoskeleton
enhances this function but is not absolutely required for it. As
such, vim knockout, which abolishes nestin’s incorporation
into the cytoskeleton, does not have any overt phenotype
because nestin function is still sufficiently preserved. In the
second possibility, nestin has a separate structural function as
a cytoskeletal IF protein that is distinct from its function in
supporting NSC survival. However, this separate structural
function is minor and does not lead to overt phenotype when
abolished in vim knockout. In either case, our data argue
against nestin being a crucial structural protein of the cell.
Some parallels can perhaps be found in a few other IF proteins that appear to possess functions that extend far beyond
their structural roles. For example, K10 interacts directly with
AKT, a multifunctional kinase important for the intracellular
REFERENCES
1
2
3
4
5
6
7
8
9
Chang L, Goldman RD. Intermediate filaments mediate cytoskeletal
crosstalk. Nat Rev Mol Cell Biol 2004;5:601–613.
Lendahl U, Zimmerman LB, McKay RD. CNS stem cells express a
new class of intermediate filament protein. Cell 1990;60:585–595.
Gilyarov AV. Nestin in central nervous system cells. Neurosci Behav
Physiol 2008;38:165–169.
Kachinsky AM, Dominov JA, Miller JB. Myogenesis and the intermediate filament protein, nestin. Dev Biol 1994;165:216–228.
Wiese C, Rolletschek A, Kania G et al. Nestin expression—A property of multi-lineage progenitor cells? Cell Mol Life Sci 2004;61:
2510–2522.
Dahlstrand J, Collins VP, Lendahl U. Expression of the class VI intermediate filament nestin in human central nervous system tumors. Cancer Res 1992;52:5334–5341.
Yang XH, Wu QL, Yu XB et al. Nestin expression in different
tumours and its relevance to malignant grade. J Clin Pathol 2008;61:
467–473.
Steinert PM, Chou YH, Prahlad V et al. A high molecular weight intermediate filament-associated protein in BHK-21 cells is nestin, a
type VI intermediate filament protein. Limited co-assembly in vitro to
form heteropolymers with type III vimentin and type IV alpha-internexin. J Biol Chem 1999;274:9881–9890.
Eliasson C, Sahlgren C, Berthold CH et al. Intermediate filament protein partnership in astrocytes. J Biol Chem 1999;274:23996–24006.
relay of signals that leads to cell growth or death [41]. Ectopic expression of K10 in vitro and in vivo significantly
reduces cell proliferation in a dose-dependent manner [41].
Mouse skin keratinocytes lacking K17 show depressed protein
translation and decreased AKT/mTOR signaling activity [42].
K8 and K18 are known to offer cytoprotection during liver
damage [43]. Although the cytoprotection effect of K8/K18
proteins is not yet clearly understood, their interaction with
stress proteins such as Hsp70, Mrj Hsp27, c-Jun, and PKC-e
appear to be involved ([44] and references therein). Lamins
have been proposed to play important roles in chromatin organization, DNA replication, transcription, and DNA repair
[45]. Thus, recent studies are pointing increasingly to the possibility that many IF proteins do much more than just contributing to the structural integrity of cells. Rather, they may also
have important regulatory functions in cell growth and death
through dynamic interactions with nonstructural proteins. To
this end, nestin is perhaps the most notable example in that it
possesses an important function in the survival and selfrenewal of NSCs, which appears entirely uncoupled from its
participation in the cytoskeleton.
ACKNOWLEDGMENTS
We thank Drs. Charles Babinet and David Markovitz for Vim
knockout mice and Dr. Masahiro Yamaguchi for Nes-GFP transgenic mice. This work was supported in part by the Human Frontier Science Program (D.P.), National Natural Science
Foundation of China (30671023; A.P.X.), National Basic
Research Program of China (2009CB522100 and
2010CB945400; A.P.X.), and Key Scientific and Technological
Projects of Guangdong Province (2007A032100003; A.P.X.).
DISCLOSURE OF POTENTIAL
CONFLICTS OF INTEREST
The authors indicate no potential conflicts of interest.
10 Sahlgren CM, Mikhailov A, Hellman J et al. Mitotic reorganization of
the intermediate filament protein nestin involves phosphorylation by
cdc2 kinase. J Biol Chem 2001;276:16456–16463.
11 Chou YH, Khuon S, Herrmann H et al. Nestin promotes the phosphorylation-dependent disassembly of vimentin intermediate filaments
during mitosis. Mol Biol Cell 2003;14:1468–1478.
12 Thomas SK, Messam CA, Spengler BA et al. Nestin is a potential mediator of malignancy in human neuroblastoma cells. J Biol Chem
2004;279:27994–27999.
13 Wei LC, Shi M, Cao R et al. Nestin small interfering RNA (siRNA)
reduces cell growth in cultured astrocytoma cells. Brain Res 2008;
1196:103–112.
14 Sahlgren CM, Pallari HM, He T et al. A nestin scaffold links Cdk5/
p35 signaling to oxidant-induced cell death. EMBO J 2006;25:
4808.
15 Colucci-Guyon E, Portier MM, Dunia I et al. Mice lacking vimentin
develop and reproduce without an obvious phenotype. Cell 1994;79:
679–694.
16 Elder GA, Friedrich VL, Jr., Kang C et al. Requirement of heavy neurofilament subunit in the development of axons with large calibers. J
Cell Biol 1998;143:195–205.
17 Milner DJ, Weitzer G, Tran D et al. Disruption of muscle architecture
and myocardial degeneration in mice lacking desmin. J Cell Biol
1996;134:1255–1270.
18 Elder GA, Friedrich VL, Jr., Bosco P et al. Absence of the mid-sized
neurofilament subunit decreases axonal calibers, levels of light neurofilament (NF-L), and neurofilament content. J Cell Biol 1998;141:
727–739.
Park, Xiang, Mao et al.
2171
19 Zhu Q, Couillard-Despres S, Julien JP. Delayed maturation of regenerating myelinated axons in mice lacking neurofilaments. Exp Neurol
1997;148:299–316.
20 Zhu Q, Lindenbaum M, Levavasseur F et al. Disruption of the NF-H
gene increases axonal microtubule content and velocity of neurofilament transport: Relief of axonopathy resulting from the toxin beta,
beta0 -iminodipropionitrile. J Cell Biol 1998;143:183–193.
21 Li Z, Colucci-Guyon E, Pincon-Raymond M et al. Cardiovascular
lesions and skeletal myopathy in mice lacking desmin. Dev Biol
1996;175:362–366.
22 Levavasseur F, Zhu Q, Julien JP. No requirement of alpha-internexin
for nervous system development and for radial growth of axons. Brain
Res Mol Brain Res 1999;69:104–112.
23 Pekny M, Leveen P, Pekna M et al. Mice lacking glial fibrillary acidic
protein display astrocytes devoid of intermediate filaments but develop
and reproduce normally. EMBO J 1995;14:1590–1598.
24 McCall MA, Gregg RG, Behringer RR et al. Targeted deletion in
astrocyte intermediate filament (Gfap) alters neuronal physiology.
Proc Natl Acad Sci USA 1996;93:6361–6366.
25 Chen YT, Bradley A. A new positive/negative selectable marker,
puDeltatk, for use in embryonic stem cells. Genesis 2000;28:31–35.
26 Mujtaba T, Piper DR, Kalyani A et al. Lineage-restricted neural precursors can be isolated from both the mouse neural tube and cultured
ES cells. Dev Biol 1999;214:113–127.
27 Bottenstein JE, Sato GH. Growth of a rat neuroblastoma cell line in serumfree supplemented medium. Proc Natl Acad Sci USA 1979;76:514–517.
28 Yamaguchi M, Saito H, Suzuki M et al. Visualization of neurogenesis
in the central nervous system using nestin promoter-GFP transgenic
mice. Neuroreport 2000;11:1991–1996.
29 Kroemer G, Martin SJ. Caspase-independent cell death. Nat Med
2005;11:725–730.
30 Singec I, Knoth R, Meyer RP et al. Defining the actual sensitivity and
specificity of the neurosphere assay in stem cell biology. Nat Methods
2006;3:801–806.
31 Louis SA, Rietze RL, Deleyrolle L et al. Enumeration of neural stem
and progenitor cells in the neural colony-forming cell assay. Stem
Cells 2008;26:988–996.
32 Triolo D, Dina G, Lorenzetti I et al. Loss of glial fibrillary acidic protein (GFAP) impairs Schwann cell proliferation and delays nerve
regeneration after damage. J Cell Sci 2006;119:3981–3993.
33 Menet V, Gimenez y Ribotta M, Chauvet N et al. Inactivation of the
glial fibrillary acidic protein gene, but not that of vimentin, improves
neuronal survival and neurite growth by modifying adhesion molecule
expression. J Neurosci 2001;21:6147–6158.
34 Elmore S. Apoptosis: A review of programmed cell death. Toxicol
Pathol 2007;35:495–516.
35 Herrmann H, Aebi U. Intermediate filaments and their associates:
Multi-talented structural elements specifying cytoarchitecture and
cytodynamics. Curr Opin Cell Biol 2000;12:79–90.
36 Eckes B, Dogic D, Colucci-Guyon E et al. Impaired mechanical stability, migration and contractile capacity in vimentin-deficient fibroblasts. J Cell Sci 1998;111(part 13):1897–1907.
37 Daniel C, Albrecht H, Ludke A et al. Nestin expression in repopulating mesangial cells promotes their proliferation. Lab Invest 2008;88:
387–397.
38 Wautier F, Wislet-Gendebien S, Chanas G et al. Regulation of nestin
expression by thrombin and cell density in cultures of bone mesenchymal stem cells and radial glial cells. BMC Neurosci 2007;8:104.
39 Pollard SM, Conti L, Sun Y et al. Adherent neural stem (NS) cells
from fetal and adult forebrain. Cereb Cortex 2006;16(suppl 1):
i112–i120.
40 Sun Y, Pollard S, Conti L et al. Long-term tripotent differentiation
capacity of human neural stem (NS) cells in adherent culture. Mol
Cell Neurosci 2008;38:245–258.
41 Paramio JM, Casanova ML, Segrelles C et al. Modulation of cell proliferation by cytokeratins K10 and K16. Mol Cell Biol 1999;19:
3086–3094.
42 Kim S, Wong P, Coulombe PA. A keratin cytoskeletal protein regulates protein synthesis and epithelial cell growth. Nature 2006;441:
362–365.
43 Zatloukal K, Stumptner C, Lehner M et al. Cytokeratin 8 protects
from hepatotoxicity, and its ratio to cytokeratin 18 determines the
ability of hepatocytes to form Mallory bodies. Am J Pathol 2000;156:
1263–1274.
44 Coulombe PA, Wong P. Cytoplasmic intermediate filaments revealed
as dynamic and multipurpose scaffolds. Nat Cell Biol 2004;6:699–
706.
45 Dechat T, Pfleghaar K, Sengupta K et al. Nuclear lamins: Major factors in the structural organization and function of the nucleus and
chromatin. Genes Dev 2008;22:832–853.
See www.StemCells.com for supporting information available online.