23.06.2016 Views

Stem Cells

28UrpzM

28UrpzM

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Neuron<br />

Oligodendrocyte<br />

Neural<br />

progenitor<br />

cell<br />

Astrocyte<br />

FOCUS<br />

<strong>Stem</strong> <strong>Cells</strong>


focus<br />

<strong>Stem</strong> <strong>Cells</strong><br />

EDITORS<br />

Barbara Pauly<br />

Senior Editor<br />

pauly@embo.org | T +49 6221 8891 109<br />

Barbara joined EMBO Reports in September 2008. She completed her PhD at the University<br />

of Munich, focusing on signal transduction in the fresh water polyp Hydra. She worked at the<br />

University of California at Berkeley as a post-doctoral researcher, studying the role of the actin<br />

cytoskeleton in endocytosis in mammalian cells.<br />

Daniel Klimmeck<br />

Editor<br />

d.klimmeck@embojournal.org | T +49 6221 8891 407<br />

Daniel received his PhD in 2008 with work on ion channel signaling in sensory neurons in the<br />

laboratory of Stephan Frings at the University of Heidelberg. As a postdoc, he focused on the<br />

molecular characterization of cancer and hematopoietic stem cells with Andreas Trumpp at<br />

the DKFZ and Jeroen Krijgsveld at EMBL. Daniel joined The EMBO Journal in 2015.<br />

Celine Carret<br />

Editor<br />

c.carret@embomolmed.org | T +49 6221 8891 411<br />

EMBO<br />

Molecular<br />

Medicine<br />

Céline Carret completed her PhD at the University of Montpellier, France, characterising host<br />

immunodominant antigens to fight babesiosis, a parasitic disease caused by a unicellular<br />

Apicomplexan parasite closely related to the malaria agent Plasmodium. She further<br />

developed her post-doctoral career on malaria working at the Wellcome Trust Sanger Institute<br />

in Cambridge, UK and Instituto de Medicina Molecular in Lisbon, Portugal. Céline joined EMBO<br />

Molecular Medicine as a Scientific Editor in March 2011.<br />

Maria Polychronidou<br />

Editor<br />

msb@embo.org | T +49 6221 8891 410<br />

Maria received her PhD from the University of Heidelberg, where she studied the role of<br />

nuclear membrane proteins in development and aging. During her post-doctoral work, she<br />

focused on the analysis of tissue-specific regulatory functions of Hox transcription factors<br />

using a combination of computational and genome-wide methods.<br />

emboj.embopress.org | embor.embopress.org | embomolmed.embopress.org | msb.embopress.org


UPCOMING MEETINGS<br />

5th May<br />

22nd May<br />

3rd June<br />

EMBO workshop: Neural control of metabolism and eating<br />

behaviour, Portugal<br />

THE STEM CELL NICHE - DEVELOPMENT & DISEASE,<br />

Copenhagen, Denmark<br />

EMBO | EMBL Symposium: Hematopoietic <strong>Stem</strong> <strong>Cells</strong>: from the<br />

embryo to the aging organism, Heidelberg, Germany<br />

Céline Carret (EMBO Molecular Medicine)<br />

Barbara Pauly (EMBO Reports)<br />

Barbara Pauly (EMBO Reports) and Daniel<br />

Klimmeck (The EMBO Journal)<br />

6th June EMBL partnership: Translational medicine, Germany Céline Carret (EMBO Molecular Medicine)<br />

12th June<br />

GRC conference: Membrane Transporters: Translating<br />

Molecules to Medicine, Lucca, Italy<br />

Barbara Pauly (EMBO Reports)<br />

22nd June ISSCR 2016 Annual Meeting, USA Daniel Klimmeck (The EMBO Journal)<br />

26th June EMBO/EMBL symposia: Innate Immunity, Germany Céline Carret (EMBO Molecular Medicine)<br />

18th July Synthetic Biology: Engineering, Evolution & Design, USA Maria Polychronidou (Molecular Systems Biology)<br />

7th August<br />

16th August<br />

GRC: Tissue Niches & Resident <strong>Stem</strong> <strong>Cells</strong> in Adult Epithelia,<br />

Hong Kong, PRC<br />

EMBO Workshop: Molecular Mechanisms of Ageing and<br />

Regeneration, GR<br />

Daniel Klimmeck (The EMBO Journal)<br />

Daniel Klimmeck (The EMBO Journal)<br />

27th August EMBL Conference: Transcription and Chromatin, Germany Maria Polychronidou (Molecular Systems Biology)<br />

12th September Annual German <strong>Stem</strong> Cell Network Conference, GER Daniel Klimmeck (The EMBO Journal)<br />

14th September<br />

EMBL Conference: Proteomics in Cell Biology and Disease<br />

Mechanisms, Germany<br />

Maria Polychronidou (Molecular Systems Biology)<br />

18th September Behr Symposium: <strong>Stem</strong> <strong>Cells</strong> and Cancer, Heidelberg Barbara Pauly (EMBO Reports) and Daniel<br />

Klimmeck (The EMBO Journal)<br />

13th October 9th Berling Summer Meeting, Germany Maria Polychronidou (Molecular Systems Biology)<br />

26th October Keystone: Translational vaccinology for global health, UK Céline Carret (EMBO Molecular Medicine)<br />

11th December Cell Symposium, Hallmarks of Cancer, BE Daniel Klimmeck (The EMBO Journa<br />

or other meetings attended by editors, please see http://t.co/Oxv86K96Ct


CONTENTS<br />

Full Articles<br />

EMBO Reports<br />

Yap1 is dispensable for self-renewal but required for proper<br />

differentiation of mouse embryonic stem (ES) cells<br />

HaeWon Chung, Bum-Kyu Lee, Nadima Uprety, Wenwen Shen, Jiwoon Lee, Jonghwan Kim<br />

EMBO Reports Published online 25.02.2016<br />

DOI:10.15252/embr.201540933<br />

Resistance of glia-like central and peripheral neural stem<br />

cells to genetically induced mitochondrial dysfunction —<br />

differential effects on neurogenesis<br />

Blanca Díaz-Castro, Ricardo Pardal, Paula García-Flores, Verónica Sobrino, Rocío Durán,<br />

José I Piruat, José López-Barneo<br />

EMBO Reports Published online 21.09.2015<br />

DOI:10.15252/embr.201540982<br />

UTX inhibits EMT-induced breast CSC properties by<br />

epigenetic repression of EMT genes in cooperation with<br />

LSD1 and HDAC1<br />

Hee-Joo Choi, Ji-Hye Park, Mikyung Park, Hee-Young Won, Hyeong-seok Joo, Chang Hoon<br />

Lee, Jeong-Yeon Lee, Gu Kong<br />

EMBO Reports Published online 24.08.2015<br />

DOI:10.15252/embr.201540244<br />

Chromatin remodeling and bivalent histone modifications<br />

in embryonic stem cells<br />

Arigela Harikumar, Eran Meshorer<br />

EMBO Reports Published online 09.11.2015<br />

DOI:10.15252/embr.201541011<br />

continued overleaf


CONTENTS<br />

One-page highlights<br />

EMBO Reports<br />

Integrative genomics positions MKRN1 as a novel<br />

ribonucleoprotein within the embryonic stem cell gene<br />

regulatory network<br />

Paul A Cassar, Richard L Carpenedo, Payman Samavarchi-Tehrani, Jonathan B Olsen,<br />

Chang Jun Park, Wing Y Chang, Zhaoyi Chen, Chandarong Choey, Sean Delaney, Huishan<br />

Guo, Hongbo Guo, R Matthew Tanner, Theodore J Perkins, Scott A Tenenbaum, Andrew<br />

Emili, Jeffrey L Wrana, Derrick Gibbings, William L Stanford<br />

EMBO Reports Published online 11.08.2015<br />

DOI:10.15252/embr.201540974;<br />

Selective influence of Sox2 on POU transcription factor<br />

binding in embryonic and neural stem cells<br />

Tapan Kumar Mistri, Arun George Devasia, Lee Thean Chu, Wei Ping Ng, Florian Halbritter,<br />

Douglas Colby, Ben Martynoga, Simon R Tomlinson, Ian Chambers, Paul Robson, Thorsten<br />

Wohland<br />

EMBO Reports Published online 11.08.2015<br />

DOI:10.15252/embr.201540467;<br />

Mitochondrial metabolism in hematopoietic stem cells<br />

requires functional FOXO3<br />

Pauline Rimmelé, Raymond Liang, Carolina L Bigarella, Fatih Kocabas, Jingjing Xie,<br />

Madhavika N Serasinghe, Jerry Chipuk, Hesham Sadek, Cheng Cheng Zhang, Saghi<br />

Ghaffari<br />

EMBO Reports Published online 24.07.2015<br />

DOI:10.15252/embr.201439704;<br />

DAZL regulates Tet1 translation in murine embryonic<br />

stem cells<br />

Maaike Welling, Hsu-Hsin Chen, Javier Muñoz, Michael U Musheev, Lennart Kester,<br />

Jan Philipp Junker, Nikolai Mischerikow, Mandana Arbab, Ewart Kuijk, Lev Silberstein,<br />

Peter V Kharchenko, Mieke Geens, Christof Niehrs, Hilde van de Velde, Alexander van<br />

Oudenaarden, Albert JR Heck, Niels Geijsen<br />

EMBO Reports Published online 15.06.2015<br />

DOI:10.15252/embr.201540538<br />

Distinct germline progenitor subsets defined through<br />

Tsc2–mTORC1 signaling<br />

Robin M Hobbs, Hue M La, Juho-Antti Mäkelä, Toshiyuki Kobayashi, Tetsuo Noda, Pier<br />

Paolo Pandolfi<br />

EMBO Reports Published online 19.02.2015<br />

DOI:10.15252/embr.201439379<br />

continued overleaf


CONTENTS<br />

One-page highlights continued<br />

EMBO Reports<br />

Epigenetic predisposition to reprogramming fates in<br />

somatic cells<br />

Maayan Pour, Inbar Pilzer, Roni Rosner, Zachary D Smith, Alexander Meissner, Iftach<br />

Nachman<br />

EMBO Reports Published online 19.01.2015<br />

DOI:10.15252/embr.201439264;<br />

Harnessing the apoptotic programs in cancer stem-like cells<br />

Ying-Hua Wang, David T Scadden<br />

EMBO Reports Published online 07.08.2015<br />

DOI:10.15252/embr.201439675<br />

Effects of inflammation on stem cells: together they strive?<br />

Caghan Kizil, Nikos Kyritsis, Michael Brand<br />

EMBO Reports Published online 04.03.2015<br />

DOI:10.15252/embr.201439702;


Scientific Report<br />

Yap1 is dispensable for self-renewal but required<br />

for proper differentiation of mouse embryonic stem<br />

(ES) cells<br />

HaeWon Chung, Bum-Kyu Lee, Nadima Uprety, Wenwen Shen, Jiwoon Lee & Jonghwan Kim *<br />

Abstract<br />

Yap1 is a transcriptional co-activator of the Hippo pathway. The<br />

importance of Yap1 in early cell fate decision during embryogenesis<br />

has been well established, though its role in embryonic stem (ES)<br />

cells remains elusive. Here, we report that Yap1 plays crucial roles in<br />

normal differentiation rather than self-renewal of ES cells. Yap1-<br />

depleted ES cells maintain undifferentiated state with a typical<br />

colony morphology as well as robust alkaline phosphatase activity.<br />

These cells also retain comparable levels of the core pluripotent<br />

factors, such as Pou5f1 and Sox2, to the levels in wild-type ES cells<br />

without significant alteration of lineage-specific marker genes.<br />

Conversely, overexpression of Yap1 in ES cells promotes nuclear<br />

translocation of Yap1, resulting in disruption of self-renewal and<br />

triggering differentiation by up-regulating lineage-specific genes.<br />

Moreover, Yap1-deficient ES cells show impaired induction of<br />

lineage markers during differentiation. Collectively, our data demonstrate<br />

that Yap1 is a required factor for proper differentiation of<br />

mouse ES cells, while remaining dispensable for self-renewal.<br />

Keywords embryonic stem cells; Yap1; differentiation; Hippo pathway;<br />

self-renewal<br />

Subject Categories Signal Transduction; <strong>Stem</strong> <strong>Cells</strong><br />

DOI 10.15252/embr.201540933 | Received 24 June 2015 | Revised 8 January<br />

2016 | Accepted 22 January 2016 | Published online 25 February 2016<br />

EMBO Reports (2016) 17: 519–529<br />

Introduction<br />

The Hippo signaling pathway, modulated by cell density and cell–<br />

cell contact, is implicated in diverse cellular processes including cell<br />

proliferation [1–5], apoptosis [5,6], and organ size control [7,8].<br />

Yap1 is a transcriptional co-activator of the Hippo pathway and is<br />

known to play a crucial role in the segregation of inner cell mass<br />

(ICM) and trophectoderm (TE) during early embryogenesis [9–13].<br />

While Yap1 resides in the nucleus of trophectodermal cells and<br />

functions as a critical co-activator for TE development, it is sequestered<br />

mainly in the cytoplasm of ICM as a phosphorylated inactive<br />

form due to active Hippo signaling [9]. However, the role of Yap1 in<br />

ICM is still elusive [9,12]. In addition to Yap1, Taz and Tead family<br />

members are also crucial players in the Hippo pathway. A<br />

homologue of Yap1, Taz, shares redundant functions such as<br />

controlling cell proliferation and sensing mechanical stress [14,15].<br />

Furthermore, Tead proteins—important in TE differentiation during<br />

early embryogenesis, form a complex with Yap1 and are known to<br />

activate their downstream target genes [16,17].<br />

Two different observations on the roles of Yap1 in embryonic<br />

stem (ES) cells are of note. Recent studies suggested that Yap1<br />

plays an important role in the maintenance of mouse ES cells as<br />

an active factor in the nucleus [18,19]. These works showed<br />

knockdown (KD) of Yap1 promotes differentiation of ES cells while<br />

overexpression (OE) of Yap1 not only enhances self-renewal but<br />

also inhibits differentiation of ES cells, even under neuronal differentiation<br />

conditions [18]. However, the study showing nuclear<br />

localization of Yap1 in ES cells is somewhat contradictory to the<br />

function of the Hippo signaling, since mouse ES cells grow as<br />

tightly packed colonies. It has been suggested that high cell density<br />

or cell–cell contact activates the Hippo signaling and subsequent<br />

sequestration of Yap1 in the cytoplasm of various cell lines such<br />

as HaCaT and NIH-3T3 [1,3,7,20]. Accordingly, another recent<br />

study has claimed that both Yap1 and Taz are dispensable for<br />

self-renewal of ES cells in 2i (Gsk3b and Mek inhibitors) culture<br />

condition [21]. In this case, Yap1- and Taz-depleted ES cells maintained<br />

undifferentiated state under differentiation-promoting<br />

culture conditions [21]. Consistent with this observation, studies<br />

of neuronal differentiation from ES cells have shown that high cell<br />

density, which activates the Hippo signaling and sequesters Yap1<br />

in the cytoplasm, blocks differentiation of ES cells [22,23], suggesting<br />

that nuclear localization of Yap1 might be important in normal<br />

differentiation of ES cells.<br />

In the current study, we show that Yap1 is dispensable for the<br />

maintenance of ES cells but critical in their differentiation. Additional<br />

testing of Yap1-associated factors including Tead family<br />

proteins and Taz also supports the dispensability of Yap1 for the<br />

self-renewal of ES cells. In line with gradual up-regulation of Yap1<br />

level upon differentiation of ES cells, OE of Yap1 in ES cells<br />

enhances nuclear abundance of Yap1 accompanied by induction of<br />

various lineage-specific marker genes. On the contrary, Yap1-<br />

depleted ES cells showed impaired differentiation. Taken together,<br />

Department of Molecular Biosciences, Institute for Cellular and Molecular Biology, Center for Systems and Synthetic Biology, The University of Texas at Austin, Austin, TX, USA<br />

*Corresponding author. Tel: +1 512 232 8046; E-mail: jonghwankim@mail.utexas.edu<br />

ª 2016 The Authors EMBO reports Vol 17 | No 4 | 2016 519


EMBO reports Yap1 is required for differentiation of ES cells HaeWon Chung et al<br />

our data demonstrate a critical role of Yap1 in normal differentiation<br />

rather than self-renewal of ES cells.<br />

Results and Discussion<br />

Yap1 is dispensable for self-renewal of mouse ES cells<br />

Previous studies reported that Yap1 is required in the maintenance<br />

of mouse ES cells by showing that KD of Yap1 promotes<br />

differentiation of ES cells [18,19]. Conversely, another study<br />

claimed that double KD of Yap1 and Taz in 2i media does not<br />

disrupt self-renewal of ES cells [21]. To decipher the roles of Yap1<br />

in self-renewal and pluripotency of ES cells, we first performed KD<br />

of Yap1 using lentivirus-delivered shRNAs in J1 mouse ES cells<br />

(Fig EV1A and Dataset EV1). In contrast to the previous reports<br />

[18,19], we found that even with > 85% of Yap1 KD, ES cells maintain<br />

normal colony morphology with high alkaline phosphatase<br />

(AP) activity, whereas ES cells with KD of Pou5f1 undergo differentiation<br />

accompanied by loss of AP activity, as expected (Fig 1A).<br />

We additionally found that these Yap1-depleted ES cells show<br />

comparable proliferation rate to that of control ES cells (Fig EV1B).<br />

In agreement with these observations, overall expression levels of<br />

ES cell core pluripotency factors, such as Pou5f1 and Nanog, as well<br />

as several lineage-specific regulators were not significantly altered<br />

upon KD of Yap1 (Figs 1B–D and EV1C). We validated our<br />

observation by testing two additional ES cell lines (E14 and CJ7)<br />

and confirmed that KD of Yap1 does not significantly affect features<br />

of normal self-renewing ES cells (Fig EV2A–F).<br />

To rule out the possibility of off-target effects and incomplete<br />

depletion due to shRNA-mediated KD strategies, we additionally<br />

established Yap1 knockout (KO) ES cell lines harboring premature<br />

stop codons on both alleles by CRISPR-Cas9-based genome editing<br />

strategies (Dataset EV1) [24,25]. Consistent with the KD results,<br />

Yap1 KO ES cells sustained self-renewing status and showed normal<br />

ES colony morphology and high AP activity, and levels of<br />

pluripotency-related genes were comparable to those of wild-type<br />

ES cells (Figs 1E–G and EV2G–L). Yap1 KO ES cells were able<br />

to maintain self-renewal for more than 1 month in culture<br />

(Fig EV3A–C). Taken together, these results indicate that Yap1 is<br />

dispensable for self-renewal of mouse ES cells.<br />

To further validate the dispensability of Yap1 in self-renewal of<br />

ES cells, we sought to monitor the global gene expression profiles of<br />

Yap1 KD and Yap1 KO ES cells using RNA-seq approaches. As<br />

expected, comparison of expression profiles between ES and differentiating<br />

ES cells revealed many differentially expressed genes<br />

(DEGs) (Fig 1H and Dataset EV2). However, expression levels of<br />

these genes were not altered significantly upon KD or KO of Yap1<br />

(Fig 1H) which was further confirmed by RT–qPCR (Fig 1I). Overall,<br />

these results indicate that the depletion of Yap1 does not trigger<br />

differentiation of ES cells.<br />

Unlike differentiating ES cells, a hierarchical clustering of<br />

global expression data revealed that Yap1 KD and Yap1 KO ES<br />

cells were clustered together with normal and control ES cells,<br />

indicating that Yap1-deficient ES cells have similar expression profiles<br />

to those of normal ES cells (Fig 1J). We also investigated the<br />

activity of previously defined functional modules in ES cells (Core<br />

and PRC) [26]. Module activity is defined as an averaged<br />

expression of all genes in each module. Briefly, the Core module<br />

includes core pluripotency factors such as Pou5f1, Nanog, and<br />

Sox2, most of which are highly expressed in self-renewing<br />

ES cells. On the other hand, the PRC module includes many<br />

lineage-specific regulators, such as Fgf5, Bmp4, and Hand1, most<br />

of which are repressed in ES cells. Since differentiation of ES cells<br />

decreases the activity of Core module but increases the activity of<br />

PRC module [26], we sought to examine module activities upon<br />

KD or KO of Yap1 to test whether cells maintain self-renewal. As<br />

shown in Fig 1K, ES cells with depleted Yap1 did not show<br />

down-regulation of Core module activity or up-regulation of PRC<br />

module activity, suggesting that Yap1-depleted ES cells largely<br />

maintain self-renewing and undifferentiated states. Further correlation<br />

analyses verified that the global expression patterns of<br />

Yap1-depleted ES cells showed higher correlation with those of<br />

control ES cells (R 2 = 0.978 for Yap1 KD, R 2 = 0.977 for Yap1<br />

KO1, and R 2 = 0.979 for Yap1 KO2) than differentiating ES cells<br />

(R 2 = 0.781) (Fig 1L). Collectively, these data provide strong<br />

evidence that the depletion of Yap1 does not significantly alter<br />

the self-renewal of ES cells.<br />

Figure 1. Yap1 is dispensable for self-renewal of J1 mouse ES cells (see also Figs EV1–EV3).<br />

A Colony morphology and alkaline phosphatase (AP) activity of ES cells upon KD of Yap1 and Pou5f1. KD1 and KD2 indicate two different shRNA sequences tested. All<br />

the following cell morphology and AP staining pictures were taken two passages (4 days) after lentivirus infection unless otherwise stated.<br />

B, C mRNA expression levels of Pou5f1, Nanog, Sox2, Esrrb (B), and Yap1 (C) upon KD of Yap1. All the following mRNA samples were harvested 4 days after lentivirus<br />

infection while passaged every 2 days unless otherwise stated. Data are represented as mean SD.<br />

D Protein levels of Yap1, Pou5f1, and Nanog upon KD of Yap1. All the following protein samples were harvested 4 days after lentivirus infection while passaged every<br />

2 days unless otherwise stated.<br />

E Colony morphology and AP activity of mouse embryonic stem cells (ESC) and three Yap1 KO clones (KO1-KO3).<br />

F mRNA levels of Pou5f1, Nanog, Sox2, and Esrrb upon KO of Yap1. Data are represented as mean SD.<br />

G Protein levels of Yap1, Pou5f1, and Nanog in Yap1 KO clones.<br />

H A heatmap showing relative mRNA expression levels of 3,605 genes differentially expressed (> twofold) between ES cells and differentiating ES cells (dESC). Genes<br />

were sorted by the fold changes of gene expression between dESC and ES cells. Corresponding gene expression profiles obtained from Yap1 KO1, Yap1 KO2, and<br />

Yap1 KD cells are also shown.<br />

I mRNA expression levels of lineage-specific marker genes upon KD of Yap1. dESC were used as control cells.<br />

J A heatmap showing Pearson’s correlation coefficients of gene expression profiles obtained from ESC, control virus-infected ES cells (Control), dESC, Yap1 KD cells,<br />

and Yap1 KO cells.<br />

K Relative average module activities (Core and PRC) in Yap1 KD1 cells, KO cells, and dESC. Module activities were normalized by the data obtained in ES cells. Data<br />

are represented as mean SEM.<br />

L Scatter plots showing log 10 (FPKM) values of genes in Yap1 KD1 cells and Control (upper left panel), dESC and ESC (bottom left panel), and Yap1 KO cells and ES<br />

cells (right two panels). Pearson’s correlation coefficients (R 2 ) are indicated. “FPKM” indicates fragments per kilobase of transcript per million fragments mapped.<br />

▸<br />

520<br />

EMBO reports Vol 17 | No 4 | 2016 ª 2016 The Authors


HaeWon Chung et al Yap1 is required for differentiation of ES cells EMBO reports<br />

A Control Yap1 KD1 Yap1 KD2 Pou5f1KD<br />

AP staining Bright field<br />

AP staining Bright field<br />

200µm 200µm 200µm<br />

200µm<br />

200µm 200µm 200µm<br />

200µm<br />

E<br />

ESC Yap1 KO1 Yap1 KO2 Yap1 KO3<br />

200µm 200µm 200µm 200µm<br />

200µm 200µm 200µm 200µm<br />

Relative gene expression<br />

2<br />

1<br />

0<br />

B<br />

F<br />

Relative gene expression<br />

Pou5f1<br />

Nanog<br />

Sox2<br />

Esrrb<br />

Control<br />

1.5<br />

1<br />

0.5<br />

Yap1 KD1<br />

Yap1 KD2<br />

ESC<br />

Yap1 KO1<br />

dESC<br />

C<br />

Relative Yap1 expression<br />

Yap1 KO2<br />

Yap1 KO3<br />

0<br />

Pou5f1Nanog Sox2 Esrrb<br />

1<br />

0<br />

Control<br />

Yap1 KD1<br />

Yap1 KD2<br />

D<br />

Yap1<br />

Pou5f1<br />

G<br />

Yap1<br />

Pou5f1<br />

Nanog<br />

Gapdh<br />

Nanog<br />

Gapdh<br />

Control<br />

ESC<br />

Yap1 KO1<br />

Yap1 KD1<br />

Yap1 KD2<br />

Yap1 KO2<br />

Yap1 KO3<br />

H<br />

J<br />

1,826 genes<br />

1,779 genes<br />

ESC<br />

Yap1 KO1<br />

Yap1 KO2<br />

Control<br />

Yap1 KD1<br />

dESC<br />

dESC<br />

Yap1 KO1<br />

Yap1 KO2<br />

Yap1 KD1<br />

ESC<br />

Yap1 KO1<br />

Yap1 KO2<br />

Control<br />

Yap1 KD 1<br />

dESC<br />

log2<br />

+3.0<br />

- 3.0<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

I<br />

Relative gene expression<br />

K<br />

Module activity<br />

1000<br />

100<br />

10<br />

1<br />

0<br />

1.4<br />

0.7<br />

0<br />

-0.7<br />

Yap1 KO1<br />

Control<br />

Yap1 KD1<br />

Yap1 KD2<br />

dESC<br />

Gbx2 Nes Gata2 T Sox17 Nodal Cdx2 Gata3<br />

Core<br />

PRC<br />

Yap1 KO2<br />

Yap1 KD1<br />

dESC<br />

L<br />

Yap1 KD1<br />

dESC<br />

-2<br />

-2<br />

4<br />

2<br />

0<br />

-2<br />

4<br />

2<br />

0<br />

-2<br />

0<br />

0<br />

R² = 0.978<br />

Control<br />

2 4<br />

R² = 0.781<br />

2 4<br />

ESC<br />

Yap1 KO1<br />

Yap1 KO2<br />

-2<br />

-2<br />

4<br />

2<br />

0<br />

0<br />

-2<br />

4<br />

2<br />

-2<br />

0<br />

0<br />

R² = 0.977<br />

2 4<br />

ESC<br />

R² = 0.979<br />

2 4<br />

ESC<br />

Figure 1.<br />

ª 2016 The Authors EMBO reports Vol 17 | No 4 | 2016<br />

521


EMBO reports Yap1 is required for differentiation of ES cells HaeWon Chung et al<br />

Taz and Tead family proteins are not required for self-renewal of<br />

ES cells and do not compensate Yap1 functions in Yap1-depleted<br />

ES cells<br />

Taz is homologous to Yap1 and has similar functions to Yap1, such<br />

as regulation of proliferation and activation of TE lineage markers<br />

[14,27]. To rule out the possibility of compensation by Taz in Yap1<br />

KD ES cells, we performed both single and double KD of Yap1 and<br />

Taz using shRNAs under drug selections (blasticidin and puromycin,<br />

respectively). J1 ES cells with depletion of both Yap1 and<br />

Taz (> 85% of KD for each) maintained typical colony morphology<br />

as well as high AP activity (Fig 2A and B). Additionally, the levels<br />

of pluripotency markers, such as Pou5f1 and Nanog, as well as various<br />

lineage markers were not significantly affected by either single<br />

or double KD of Yap1 and Taz (Fig 2C and D), indicating that the<br />

dispensability of Yap1 in self-renewing ES cells is not due to the<br />

compensatory effect of Taz.<br />

Since Yap1 is known to require Tead family proteins to activate<br />

its downstream target genes in NIH-3T3 and MCF10A cell lines<br />

[17,28,29], we investigated whether Tead proteins are also<br />

dispensable for the maintenance of ES cells. To do so, we first<br />

performed KD of Tead2 in ES cells. In contrast to the previous<br />

report [19], we did not observe any significant alteration of cell<br />

morphology or reduced AP activity upon down-regulation of Tead2<br />

(at least > 90% of KD in mRNA levels) (Fig EV4A and B). In<br />

accordance with the colony morphology, Tead2 KD ES cells<br />

expressed similar levels of pluripotency genes compared to wildtype<br />

ES cells (Fig EV4C and D). These results were confirmed by<br />

generation of three independent Tead2 KO ES cell clones using<br />

CRISPR-Cas9 strategies (Dataset EV1). These Tead2 KO clones also<br />

maintained self-renewal without differentiation (Fig EV4E–G). We<br />

additionally conducted triple KD of Tead1/3/4 with triple drug<br />

selection (at least > 80% of KD for each), and did not observe any<br />

significant alteration of cell morphology or AP activity which is in<br />

contrast to the previous report [18] (Fig 2E and F). Similar to the<br />

results obtained from the KD of Yap1, triple Tead KD ES cells<br />

expressed comparable levels of pluripotency genes shown in wildtype<br />

ES cells without any significant activation of lineage-specific<br />

regulators (Fig 2G–I). The results were further validated by double<br />

KD of Tead1/3 in Tead4 KO cells. ES cells with Tead4 KO and<br />

Tead1/3 KD also maintained self-renewal (Fig EV4H–J). Collectively,<br />

our data suggest that both Yap1 and Yap1-associated<br />

proteins such as Taz and Tead are not required for self-renewal of<br />

ES cells.<br />

Yap1 is induced and translocated into the nucleus upon<br />

differentiation of ES cells<br />

In order to investigate the roles of Yap1 in differentiation of ES cells,<br />

we examined the expression level of Yap1 in self-renewing mouse<br />

ES cells as well as upon differentiation of ES cells. Analysis of<br />

published mRNA expression data obtained upon time-course differentiation<br />

of embryoid body (EB) [30] revealed that Yap1 is<br />

moderately expressed in ES cells while its expression gradually<br />

increases upon differentiation (Fig 3A). We differentiated mouse J1<br />

ES cells by the withdrawal of leukemia inhibitory factor (LIF) in the<br />

culture media and examined the level of Yap1. Consistent with the<br />

results from the EB differentiation, both mRNA and protein levels of<br />

Yap1 were moderately increased upon spontaneous differentiation<br />

(Fig 3B and C).<br />

Since active Hippo signaling leads to phosphorylation and<br />

cytoplasmic sequestration of Yap1, blocking Yap1’s function as a<br />

transcriptional coactivator [7,9,12], we examined the levels of phospho-Yap1<br />

and its subsequent localization in both self-renewing and<br />

differentiating ES cells. Western blot analysis showed that Yap1 is<br />

highly phosphorylated in self-renewing ES cells, but the level of<br />

phospho-Yap1 is reduced in differentiating ES cells (Fig 3D). Given<br />

the fact that phospho-Yap1 is sequestered in the cytoplasm [7,9,12],<br />

we examined Yap1 localization by immunofluorescence (IF).<br />

Consistent with hyper-phosphorylation of Yap1 in ES cells, IF results<br />

revealed that Yap1 resides primarily in the cytoplasm of selfrenewing<br />

ES cells (Fig 3E–G). However, upon differentiation of<br />

multiple mouse ES cell lines we tested (J1, CJ7, and E14), Yap1 was<br />

translocated into the nucleus (Figs 3E–G and EV5A–D). Cytoplasmic<br />

Yap1 in ES cells could be attributed to compact ES cell colonies with<br />

active Hippo signaling [7], while lower cell density of differentiating<br />

ES cells growing in a monolayer leads to inactive Hippo signaling,<br />

resulting in the nuclear localization of Yap1.<br />

We further investigated the activity of nuclear Yap1 using a<br />

synthetic Yap1-responsive luciferase (8xGTIIC) construct as previously<br />

designed for the measurement of Yap1 transcriptional activity<br />

in mechanical stress condition (Fig 3H) [15,19,31,32]. The luciferase<br />

construct contains repeated Yap1-Tead binding motifs (eight times)<br />

in front of the minimal cTNT promoter followed by a luciferase<br />

reporter gene [15,32,33]. As shown in Fig 3I and J, we observed a<br />

significant increase in luciferase activity in both ES cells with transient<br />

OE of Yap1 and in differentiating ES cells compared to the<br />

reporter activity in self-renewing ES cells, thereby indicating the<br />

increased level of nuclear Yap1 either by OE of Yap1 or by ES cell<br />

differentiation promotes transcription of the reporter gene. An<br />

induced level and nuclear localization of Yap1 were also confirmed,<br />

along with the increased Yap1 activity in Pou5f1 KD ES cells<br />

undergoing TE differentiation (Fig 3K–M).<br />

Yap1 is required for normal differentiation of ES cells<br />

As we observed increased expression levels and nuclear localization<br />

of Yap1 in differentiating ES cells (Fig 3), we hypothesized that<br />

Yap1 may have critical roles in differentiation of ES cells. To address<br />

this, we tested differentiation potential of Yap1 KD ES cells. Upon<br />

3 days of differentiation, completely differentiated and monolayered<br />

cellular morphology with reduced AP activity were observed<br />

in control ES cells. However, Yap1 KD cells maintained typical<br />

colony morphology with high AP activity comparable to that of selfrenewing<br />

ES cells even after 2–3 days of differentiation (Fig 4A).<br />

We further found that expression levels of some pluripotency factors<br />

such as Sox2 and Esrrb were relatively highly maintained in Yap1-<br />

depleted cells upon differentiation, although the expression of other<br />

core factors, Pou5f1 and Nanog, was decreased similarly to their<br />

levels in control cells upon differentiation (Fig EV6A). Moreover,<br />

up-regulation of various lineage-specific markers, such as Nes, T,<br />

Gsc, Gata6, Cdx2, and Gata3, was significantly impaired during<br />

differentiation of Yap1-depleted cells (Fig EV6B), suggesting that the<br />

depletion of Yap1 affects differentiation potential of ES cells.<br />

In order to get further insight into the roles of Yap1 in global<br />

transcriptional regulation during differentiation, we analyzed gene<br />

522<br />

EMBO reports Vol 17 | No 4 | 2016 ª 2016 The Authors


HaeWon Chung et al Yap1 is required for differentiation of ES cells EMBO reports<br />

A B C<br />

Control Taz KD Yap1 KD Taz/Yap1 KD<br />

Taz Yap1<br />

1.5<br />

D<br />

Relative gene expression<br />

Bright field<br />

AP staining<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

Control<br />

Taz KD<br />

200µm 200µm 200µm 200µm<br />

200µm 200µm 200µm 200µm<br />

Yap1 KD<br />

Taz/Yap1 KD<br />

Relative gene expression<br />

1<br />

0.5<br />

0<br />

Control<br />

Taz KD<br />

Yap1 KD1<br />

Taz/Yap1 KD<br />

Gbx2 Nes Gata2 T Nodal Gata3 Gsc Amot Mixl1 Fzd6 Edn1 Cited1 Msx2 Krt18<br />

Relative gene expression<br />

2<br />

1.5<br />

1<br />

0.5<br />

Control<br />

Taz KD<br />

Yap1 KD<br />

Taz/Yap1 KD<br />

0<br />

Pou5f1 Nanog Sox2 Esrrb<br />

E F G<br />

Control Tead1/3/4 KD<br />

Tead1<br />

Tead3<br />

Tead4<br />

1<br />

I<br />

Bright field<br />

AP staining<br />

2<br />

Control<br />

200µm 200µm<br />

200µm 200µm<br />

Tead 1/3/4 KD<br />

Relative gene expresison<br />

0.5<br />

0<br />

Control<br />

Tead1/3/4 KD<br />

Relative gene expression<br />

1<br />

0.5<br />

0<br />

Control<br />

Tead1/3/4 KD<br />

Pou5f1<br />

Nanog<br />

Sox2<br />

Esrrb<br />

H<br />

Tead1<br />

Tead3<br />

Tead4<br />

Pou5f1<br />

Gapdh<br />

Control<br />

Tead1/3/4 KD<br />

N.S.<br />

Relative gene expression<br />

1.5<br />

1<br />

0.5<br />

0<br />

Gbx2<br />

Gata2<br />

T<br />

Sox17<br />

Nodal<br />

Cdx2<br />

Gata3<br />

Isl1<br />

Gsc<br />

Gata4<br />

Gata6<br />

Pitx2<br />

Eomes<br />

Sox13<br />

Tgfb2<br />

Amot<br />

Mixl1<br />

Fzd6<br />

Edn1<br />

Cited1<br />

Msx2<br />

Krt18<br />

Figure 2. Taz and Tead family proteins are not required for the self-renewal of J1 mouse ES cells (see also Fig EV4).<br />

A Colony morphology and AP activity of ES cells upon KD of Yap1 and Taz.<br />

B–D mRNA expression levels of Yap1 and Taz (B), Pou5f1, Nanog, Sox2, and Esrrb (C), and lineage-specific marker genes (D) upon KD of Yap1 and Taz. Data are<br />

represented as mean SD.<br />

E Colony morphology and AP activity of ES cells upon KD of Tead1/Tead3/Tead4.<br />

F, G mRNA expression levels of Tead1, Tead3, and Tead4 (F) and Pou5f1, Nanog, Sox2, and Esrrb (G) upon KD of Tead 1/3/4. Data are represented as mean SD.<br />

H Protein levels of Tead1, Tead3, Tead4, and Pou5f1 upon KD of Tead1/Tead3/Tead4. N.S., non-specific.<br />

I mRNA expression levels of lineage-specific marker genes upon KD of Tead1/Tead3/Tead4. Data are represented as mean SD.<br />

ª 2016 The Authors EMBO reports Vol 17 | No 4 | 2016<br />

523


EMBO reports Yap1 is required for differentiation of ES cells HaeWon Chung et al<br />

A<br />

Relative gene expression<br />

2 Yap1<br />

Pou5f1<br />

1.5 Gata6<br />

1<br />

0.5<br />

0<br />

0<br />

0.25<br />

0.5<br />

0.75<br />

1<br />

1.5<br />

2<br />

4<br />

7<br />

B<br />

Relative Yap1 expression<br />

2<br />

1<br />

0<br />

ESC<br />

dESC<br />

C<br />

Yap1<br />

Pou5f1<br />

Actb<br />

D<br />

p-Yap1<br />

Yap1<br />

Time (days)<br />

0 2 4 6<br />

0 2 4<br />

Time (Days)<br />

E<br />

ESC<br />

dESC<br />

F<br />

ESC<br />

dESC<br />

G<br />

ESC<br />

dESC<br />

Yap1<br />

20um<br />

20um<br />

20um<br />

20um<br />

20um<br />

20um<br />

Merge DAPI<br />

20um<br />

20um<br />

Merge DAPI<br />

20um<br />

20um<br />

Merge DAPI<br />

20um<br />

20um<br />

20um<br />

20um<br />

20um<br />

20um<br />

20um<br />

20um<br />

Merge<br />

Merge<br />

Merge<br />

Yap1<br />

Yap1<br />

5um 5um 5um 5um 5um 5um<br />

H<br />

Yap1<br />

Tead<br />

8xGTIIC promoter luciferase<br />

Relative luc activity<br />

I<br />

12<br />

9<br />

6<br />

3<br />

0<br />

Control<br />

**<br />

Yap1 OE<br />

J<br />

Relative luc activity<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

ESC<br />

K<br />

** **<br />

6<br />

dESC<br />

Relative Yap1 expression<br />

4<br />

2<br />

0<br />

Control<br />

Pou5f1 KD<br />

L<br />

Control<br />

Pou5f1 KD<br />

Yap1 DAPI Merge<br />

20um 20um 20um<br />

20um 20um 20um<br />

M<br />

Relative luc activity<br />

3<br />

2<br />

1<br />

0<br />

Control<br />

**<br />

Pou5f1 KD<br />

Figure 3.<br />

expression profiles obtained from RNA-seq of normal ES cells<br />

and Yap1-depleted ES cells before and after differentiation. As<br />

shown in Fig 4B, gene expression patterns of DEGs (Yap1 KD ES<br />

cells/wild-type ES cells) upon differentiation showed an inverse<br />

correlation with the expression patterns of wild-type differentiating<br />

cells over self-renewing ES cells (Dataset EV3). The heatmap results<br />

524<br />

EMBO reports Vol 17 | No 4 | 2016 ª 2016 The Authors


HaeWon Chung et al Yap1 is required for differentiation of ES cells EMBO reports<br />

◀<br />

Figure 3. Yap1 is up-regulated and translocated into nucleus during ES cell differentiation (see also Fig EV5).<br />

A Relative mRNA levels of Yap1, Pou5f1, and Gata6 during time-course embryoid body (EB) differentiation. Gene expression data were obtained from GSE3749. Pou5f1<br />

and Gata6 serve as representative ES cell marker and lineage-specific marker, respectively.<br />

B Relative Yap1 mRNA levels in ES cells (ESC) and differentiating ES cells (dESC) (LIF withdrawal for 4 days) and data are represented as mean SD. To differentiate<br />

ES cells, cells were incubated in LIF-withdrawn medium for 4 days. Both ESC and dESC were passaged every 2 days.<br />

C Protein levels of Yap1 and Pou5f1 during time-course differentiation upon LIF withdrawal.<br />

D Phospho-Yap1 levels during time-course differentiation. Samples were normalized by total Yap1 level.<br />

E–G Immunofluorescence (IF) images depicting localization of Yap1 in J1 (E), CJ7 (F), and E14 (G) mouse ESC (top) and dESC (bottom). The white arrow indicates<br />

nucleolus. Bottom panels represent higher magnification of the above panels. Dashed circle indicates nucleus border.<br />

H A schematic diagram depicting a Yap1-responsive luciferase reporter (8xGTIIC) construct.<br />

I Luciferase reporter assay using Yap1-responsive luciferase reporter (8xGTIIC) upon transient overexpression (OE) Yap1 in ES cells. P-values were calculated using<br />

Student’s t-test. Data are represented as mean SD. **P < 0.01. “Control” indicates ES cells infected with control virus not expressing any specific shRNA<br />

sequence.<br />

J Relative activity of Yap1-responsive luciferase reporter gene in ESC and dESC. P-values were calculated using Student’s t-test. Data are represented as mean SD.<br />

**P < 0.01.<br />

K Relative Yap1 mRNA levels in Control and Pou5f1 KD ES cells. Data are represented as mean SD. **P < 0.01. “Control” indicates ES cells infected with control<br />

virus not expressing any specific shRNA sequence.<br />

L IF images depicting localization of Yap1 in Control and Pou5f1 KD ES cells.<br />

M Relative activity of Yap1-responsive luciferase reporter gene upon Pou5f1 KD in ES cells. P-values were calculated using Student’s t-test. Data are represented as<br />

mean SD. **P < 0.01.<br />

clearly revealed that Yap1-depleted ES cells are not properly differentiated.<br />

Additional analyses of the Core and PRC module activity<br />

consistently indicated that Yap1 depletion causes stronger Core<br />

module activity with weaker PRC module activity during differentiation,<br />

indicating that KD of Yap1 delayed or impaired proper differentiation<br />

of ES cells (Fig 4C). Gene ontology (GO) term analysis using<br />

the genes that are not properly induced in Yap1-depleted cells<br />

compared to wild-type ES cells upon differentiation also revealed<br />

that these genes are implicated in various development-related<br />

processes, such as blood vessel development, chordate embryonic<br />

development, and in utero embryonic development (Fig EV6C).<br />

These collectively demonstrate that adequate levels of Yap1 are<br />

critical in normal differentiation of ES cells.<br />

Ectopic expression of Yap1 in ES cells is sufficient to induce<br />

up-regulation of lineage marker genes<br />

To further test roles of Yap1 in ES cell differentiation, we performed<br />

OE of Yap1 in ES cells. Yap1 mainly resides in the cytoplasm of selfrenewing<br />

ES cells. While OE of Yap1 increases both nuclear and<br />

cytoplasmic Yap1 levels, we detected more nuclear Yap1 in Yap1 OE<br />

cells, indicating that exogenous Yap1 can translocate into the<br />

nucleus and act on its target genes (Fig EV6D and E). Yap1 OE cells<br />

also showed flattened morphology similar to that of differentiating<br />

ES cells with reduced AP activity even in the presence of LIF<br />

(Fig 4D). We further examined global gene expression profiles of<br />

Yap1 OE cells, and a clustering analysis showed that the DEGs upon<br />

OE of Yap1 (Yap1 OE/ES cells) are highly similar to the DEGs of<br />

differentiating ES cells over self-renewing ES cells (Fig 4E and<br />

Dataset EV4). Consistently, the activity of the Core module was<br />

significantly decreased upon OE of Yap1, while the PRC module<br />

activity was dramatically increased (Figs 4F and EV6F). These<br />

results suggest that OE of Yap1 is sufficient to trigger ES cell<br />

differentiation. Additional GO term analysis revealed that genes<br />

up-regulated upon OE of Yap1 are significantly enriched in developmental<br />

processes, such as chordate embryonic development,<br />

skeletal system development, and embryonic organ development<br />

(Fig 4G), further demonstrating that the ectopic expression of Yap1<br />

promotes differentiation of ES cells.<br />

Unlike the well-established functions of the Hippo signaling<br />

pathway in the first cell fate decision, the roles of Yap1 in ES cells,<br />

ICM, and during differentiation of ES cells or ICM are still not well<br />

understood. Here, we reveal that Yap1, a transcriptional effector of<br />

Hippo pathway, is a crucial factor implicated in differentiation<br />

rather than self-renewal of ES cells. In contrast to the previous<br />

reports [18,19], a depletion of Yap1 does not show any significant<br />

effect on the maintenance of multiple ES cells we tested. This is<br />

consistent with the nonessential roles of Yap1 in ES cells grown<br />

Figure 4. Yap1 is required for differentiation of ES cells (see also Fig EV6).<br />

A Colony morphology and AP activity of Control and Yap1 KD ES cells upon differentiation. Morphology and AP staining pictures were taken 2 days after differentiation.<br />

B A heatmap showing relative mRNA expression levels of 1,995 genes differentially expressed (> twofold) between Yap1 KD ES cells and Control upon 4 days of<br />

differentiation. Genes were sorted by the fold changes of gene expression between Yap1 KD ES cells and Control (first column). Corresponding gene expression<br />

changes between ES cells (ESC) and differentiating ES cells (dESC) are shown in the second column.<br />

C Relative average module activities (Core and PRC modules) between Yap1 KD ES cells and Control cells upon differentiation. Data are represented as mean SEM.<br />

D Colony morphology and AP activity in Yap1 OE cells. Two different Yap1 OE clones (OE1 and OE2) and pool of Yap1 OE (OE pool) were used. Cell morphology and AP<br />

staining pictures were taken 3 weeks after electroporation.<br />

E A heatmap showing relative mRNA expression levels of 2,137 genes differentially expressed (> twofold) between Yap1 OE ES cells and control ES cells. Genes were<br />

sorted by the fold changes of gene expression between Yap1 OE ES cells and control ES cells (first column) and corresponding gene expression profiles obtained from<br />

dESC are shown.<br />

F Relative average module activities (Core and PRC modules) between Yap1 OE cells and control cells are shown. Data are represented as mean SEM.<br />

G Genes up-regulated in Yap1 OE cells were tested using David 6.7. Significantly enriched gene ontology (GO) terms (biological functions) are shown. Developmental<br />

process-related GO terms are highlighted in red.<br />

▸<br />

ª 2016 The Authors EMBO reports Vol 17 | No 4 | 2016<br />

525


=<br />

EMBO reports Yap1 is required for differentiation of ES cells HaeWon Chung et al<br />

A B C<br />

Bright field AP staining<br />

Yap1 KD1/Control<br />

(Differentiation)<br />

dESC/ESC<br />

Control<br />

Yap1 KD<br />

Control<br />

Yap1 OE1<br />

Yap1 OE2<br />

Yap1 OE<br />

pool<br />

200µm<br />

200µm<br />

D<br />

Bright field<br />

200µm<br />

200µm<br />

200µm<br />

200µm<br />

200µm<br />

200µm<br />

AP staining<br />

200µm<br />

200µm<br />

200µm<br />

200µm<br />

1,269 genes 868 genes<br />

1,450 genes 545 genes<br />

E<br />

Yap1 OE1<br />

dESC<br />

log2<br />

+3.0<br />

-3.0 -0.4<br />

log2<br />

+3.0<br />

-3.0<br />

Module activity<br />

Module activity<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

F<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

Core<br />

Core<br />

PRC<br />

PRC<br />

G<br />

chordate embryonic development<br />

skeletal system development<br />

regulation of cell proliferation<br />

embryonic organ development<br />

positive regulation of<br />

developmental process<br />

embryonic morphogenesis<br />

heart development<br />

vasculature development<br />

in utero embryonic development<br />

tube development<br />

Biological process<br />

0 2 4 6 8<br />

-log10(P-value)<br />

Figure 4.<br />

526<br />

EMBO reports Vol 17 | No 4 | 2016 ª 2016 The Authors


HaeWon Chung et al Yap1 is required for differentiation of ES cells EMBO reports<br />

under the 2i condition [21]. In agreement with the dispensability<br />

of Yap1 in ES cells, other key effectors of the Hippo pathway<br />

(Tead family proteins and Taz) do not compensate Yap1, and are<br />

also not necessary for the maintenance of ES cells, implying that<br />

the transcriptional effectors of the Hippo pathway are at least<br />

dispensable for self-renewal of mouse ES cells. In addition, we<br />

show that Yap1 is mainly sequestered in the cytoplasm of selfrenewing<br />

ES cells, while it localizes in the nucleus upon<br />

differentiation. The predominant nuclear localization of Yap1 in<br />

differentiating ES cells may be due to inactive Hippo signaling in<br />

the cells growing with lower density. Consistently, OE of Yap1 in<br />

ES cells triggers nuclear localization of Yap1 and induces differentiation<br />

along with activation of diverse lineage markers. Our global<br />

expression analyses further suggest that Yap1 may promote<br />

differentiation by activating differentiation-related genes rather<br />

than repressing pluripotency-related genes, which is consistent<br />

with the observation of Yap1 KO embryo, which dies around E10<br />

due to the defects in yolk sac vasculogenesis, chorioallantoic<br />

fusion, and embryonic axis elongation [34]. Notably, a recent<br />

study done in human ES cells suggested that YAP1 represses<br />

mesendoderm differentiation [35], possibly due to the differences<br />

in signaling pathways between human and mouse ES cells [36,37].<br />

In-depth investigation of the impaired regulation of three lineagespecific<br />

genes as well as TE-related genes in Yap1 KD ES cells<br />

upon differentiation may provide further insights into the functions<br />

of Yap1 in early embryogenesis. Together, our data establish that<br />

Yap1 is a critical regulator for proper differentiation but<br />

dispensable for self-renewal of ES cells.<br />

Materials and Methods<br />

Cell culture<br />

J1, E14, and CJ7 mouse ES cells were maintained in Dulbecco’s<br />

modified Eagle’s medium (DMEM, Gibco Ref. 11965) supplemented<br />

with 18% of fetal bovine serum (FBS), penicillin/streptomycin/<br />

L-glutamine (Gibco Ref. 10378), MEM nonessential amino acid<br />

(Gibco Ref. 11140), nucleosides (Millipore Cat. ES-008-D), 100 lM<br />

b-mercaptoethanol (Sigma M3148), and 1,000 U/ml recombinant<br />

leukemia inhibitory factor (LIF, Millipore Cat. ESG1107). ES cells<br />

were cultured on 0.1% gelatin-coated plates at 37°C and 5% CO 2<br />

and passaged every 2 days. HEK 293T cells were maintained in<br />

DMEM supplemented with 10% of FBS and penicillin/streptomycin/<br />

L-glutamine. To differentiate ES cells, cells were washed three times<br />

with the media without LIF and then incubated for 4 days while<br />

passaging every 2 days.<br />

shRNA lentiviral production and infection<br />

12-well plate with virus containing media supplemented with<br />

10 lg/ml polybrene (Millipore Cat. TR-1003-G). After 1 day of infection,<br />

cells are selected with appropriate antibiotics and passaged<br />

every 2 days. Cell morphology, AP staining, protein and mRNA<br />

levels were examined two passages after the infection.<br />

Luciferase reporter gene assay<br />

For the luciferase reporter gene assay, 2.5 × 10 5 J1 ES cells in each<br />

24 well were co-transfected with 100 ng of the GTIIC vector, 5 ng of<br />

PGL4.75 vector containing a Renilla reporter gene as an internal<br />

control reporter using Lipofectamine 3000 (Life Technologies, Cat.<br />

L3000008) and then cultivated for 24 h. To measure luciferase<br />

reporter gene activity, cells were washed two times with PBS, lysed,<br />

and the luciferase activities were measured using the Dual<br />

Luciferase assay kit (Promega, E1910).<br />

Immunofluorescence<br />

~3 × 10 5 /ml ES cells were plated on 0.1% gelatin pre-coated<br />

l-Slide VI 0.4 (Ibidi Cat. 80606). Slides were fixed with 3.7%<br />

paraformaldehyde for 15 min at room temperature and permeabilized<br />

with 0.5% Triton X-100 for 10 min. Slides were then incubated<br />

with blocking solution (3% BSA and 1% normal horse serum in<br />

PBS) for 1 h at room temperature, Yap1 primary antibody solution<br />

(1:200 dilution, Santa Cruz sc-101199) overnight at 4°C, and<br />

secondary antibody solution (1:1,000 dilution) conjugated to Alexa<br />

Fluor 488 for 1 h at room temperature. Lastly, slides were mounted<br />

with ProLong Gold antifade reagent with DAPI (Molecular Probes<br />

P36935) and imaged on a Zeiss 710 laser scanning confocal and<br />

structured illumination microscope.<br />

RNA sequencing and data processing<br />

One lg of RNAs was used to generate Illumina-compatible sequencing<br />

libraries using mRNA isolation kit (NEB, E7490L) and RNA<br />

library prep kit (NEB, E7530S) according to the manufacturer’s<br />

protocol. Adapter ligation was done with sample-specific barcodes.<br />

RNA-seq libraries were sequenced using an Illumina NextSeq 500<br />

machine. Single-end reads from RNA-seq were mapped onto the<br />

mouse genome assembly (mm9) using default setting of Tophat2.<br />

Transcript-level expression analysis was performed using Cuffdiff to<br />

calculate FPKM (fragments per kilobase of transcript per million<br />

mapped reads) [38].<br />

Data deposition<br />

The RNA-seq data used in this study were deposited at the Gene<br />

Expression Omnibus (GEO) under the accession number GSE69669.<br />

HEK 293T cells were plated at ~6 × 10 6 cells per 100 mm 2 and then<br />

transfected with 6 lg of pLKO.1 shRNA vector (Sigma) (Dataset<br />

EV1), 4 lg of pCMV-D8.9, and 2 lg of VSVG plasmids using 30 ll of<br />

Fugene 6 (Promega Ref. 2692), according to the manufacturer’s<br />

protocol. After 24 h, HEK 293T medium was replaced with ES<br />

medium. Two days after transfection, supernatant containing viral<br />

particles was collected and filtered through 0.45-lm Supor <br />

membrane (PALL Ref. 4654). ~2 × 10 5 ES cells were plated on<br />

Expanded View for this article is available online.<br />

Acknowledgements<br />

We thank all the members of Kim laboratory for their help and support and<br />

the ICMB Microscopy and Imaging Facility as well as the Genome Sequencing<br />

and Analysis Facility (GSAF) at UT-Austin. The project is supported by<br />

R01GM112722 from NIH/NIGMS and R1106 from the Cancer Prevention<br />

Research Institute of Texas (CPRIT) to J.K. J.K. is a CPRIT scholar.<br />

ª 2016 The Authors EMBO reports Vol 17 | No 4 | 2016<br />

527


EMBO reports Yap1 is required for differentiation of ES cells HaeWon Chung et al<br />

Author contributions<br />

HWC, NU, and WS performed the experiments. HWC and B-KL analyzed<br />

RNA-seq data. HWC, B-KL, JL, and JK conceived work and wrote the<br />

manuscript.<br />

Conflict of interest<br />

The authors declare that they have no conflict of interest.<br />

References<br />

1. Kim N, Koh E (2011) E-cadherin mediates contact inhibition of proliferation<br />

through Hippo signaling-pathway components. Proc Natl Acad Sci<br />

USA 2011: 11930 – 11935<br />

2. Schlegelmilch K, Mohseni M, Kirak O, Pruszak J, Rodriguez JR, Zhou D,<br />

Kreger BT, Vasioukhin V, Avruch J, Brummelkamp TR et al (2011) Yap1<br />

acts downstream of a-catenin to control epidermal proliferation. Cell<br />

144: 782 – 795<br />

3. Mori M, Triboulet R, Mohseni M, Schlegelmilch K, Shrestha K,<br />

Camargo FD, Gregory RI (2014) Hippo signaling regulates microprocessor<br />

and links cell-density-dependent miRNA biogenesis to cancer. Cell<br />

156: 893 – 906<br />

4. Silvis MR, Kreger BT, Lien W-H, Klezovitch O, Rudakova GM, Camargo<br />

FD, Lantz DM, Seykora JT, Vasioukhin V (2011) a-catenin is a tumor<br />

suppressor that controls cell accumulation by regulating the localization<br />

and activity of the transcriptional coactivator Yap1. Sci Signal 4: ra33<br />

5. Huang J, Wu S, Barrera J, Matthews K, Pan D (2005) The Hippo signaling<br />

pathway coordinately regulates cell proliferation and apoptosis by inactivating<br />

Yorkie, the Drosophila homolog of YAP. Cell 122: 421 – 434<br />

6. Lee K-K, Yonehara S (2012) Identification of mechanism that couples<br />

multisite phosphorylation of Yes-associated protein (YAP) with transcriptional<br />

coactivation and regulation of apoptosis. J Biol Chem 287:<br />

9568 – 9578<br />

7. Zhao B, Wei X, Li W, Udan RS, Yang Q, Kim J, Xie J, Ikenoue T, Yu J, Li L<br />

et al (2007) Inactivation of YAP oncoprotein by the Hippo pathway is<br />

involved in cell contact inhibition and tissue growth control. Genes Dev<br />

21: 2747 – 2761<br />

8. Camargo FD, Gokhale S, Johnnidis JB, Fu D, Bell GW, Jaenisch R, Brummelkamp<br />

TR (2007) YAP1 increases organ size and expands undifferentiated<br />

progenitor cells. Curr Biol 17: 2054 – 2060<br />

9. Nishioka N, Inoue K, Adachi K, Kiyonari H, Ota M, Ralston A, Yabuta N,<br />

Hirahara S, Stephenson RO, Ogonuki N et al (2009) The Hippo signaling<br />

pathway components Lats and Yap pattern Tead4 activity to distinguish<br />

mouse trophectoderm from inner cell mass. Dev Cell 16: 398 – 410<br />

10. Hirate Y, Hirahara S, Inoue K-I, Suzuki A, Alarcon VB, Akimoto K, Hirai T,<br />

Hara T, Adachi M, Chida K et al (2013) Polarity-dependent distribution<br />

of angiomotin localizes Hippo signaling in preimplantation embryos.<br />

Curr Biol 23: 1181 – 1194<br />

11. Rayon T, Menchero S, Nieto A, Xenopoulos P, Crespo M, Cockburn K,<br />

Cañon S, Sasaki H, Hadjantonakis A-K, de la Pompa JL et al (2014)<br />

Notch and hippo converge on Cdx2 to specify the trophectoderm<br />

lineage in the mouse blastocyst. Dev Cell 30: 410 – 422<br />

12. Cockburn K, Biechele S, Garner J, Rossant J (2013) The Hippo pathway<br />

member Nf2 is required for inner cell mass specification. Curr Biol 23:<br />

1195 – 1201<br />

13. Wicklow E, Blij S, Frum T, Hirate Y, Lang RA, Sasaki H, Ralston A (2014)<br />

HIPPO pathway members restrict Sox2 to the inner cell mass where it<br />

promotes ICM fates in the mouse blastocyst. PLoS Genet 10: e1004618<br />

14. Imajo M, Ebisuya M, Nishida E (2014) Dual role of YAP and TAZ in<br />

renewal of the intestinal epithelium. Nat Cell Biol 17: 7 – 19<br />

15. Dupont S, Morsut L, Aragona M, Enzo E, Giulitti S, Cordenonsi M, Zanconato<br />

F, Le Digabel J, Forcato M, Bicciato S et al (2011) Role of YAP/TAZ<br />

in mechanotransduction. Nature 474: 179 – 183<br />

16. Ota M, Sasaki H (2008) Mammalian Tead proteins regulate cell proliferation<br />

and contact inhibition as transcriptional mediators of Hippo<br />

signaling. Development 135: 4059 – 4069<br />

17. Li Z, Zhao B, Wang P, Chen F, Dong Z, Yang H, Guan KL, Xu Y (2010)<br />

Structural insights into the YAP and TEAD complex. Genes Dev 24:<br />

235 – 240<br />

18. Lian I, Kim J, Okazawa H, Zhao J, Zhao B, Yu J, Chinnaiyan A, Israel MA,<br />

Goldstein LSB, Abujarour R et al (2010) The role of YAP transcription<br />

coactivator in regulating stem cell self-renewal and differentiation.<br />

Genes Dev 24: 1106 – 1118<br />

19. Tamm C, Böwer N, Annerén C (2011) Regulation of mouse embryonic<br />

stem cell self-renewal by a Yes-YAP-TEAD2 signaling pathway downstream<br />

of LIF. J Cell Sci 124: 1136 – 1144<br />

20. Varelas X, Samavarchi-Tehrani P, Narimatsu M, Weiss A, Cockburn K,<br />

Larsen BG, Rossant J, Wrana JL (2010) The Crumbs complex couples cell<br />

density sensing to Hippo-dependent control of the TGF-b-SMAD pathway.<br />

Dev Cell 19: 831 – 844<br />

21. Azzolin L, Panciera T, Soligo S, Enzo E, Bicciato S, Dupont S, Bresolin S,<br />

Frasson C, Basso G, Guzzardo V et al (2014) YAP/TAZ incorporation in<br />

the b-catenin destruction complex orchestrates the Wnt response. Cell<br />

158: 157 – 170<br />

22. Ying Q-L, Stavridis M, Griffiths D, Li M, Smith A (2003) Conversion of<br />

embryonic stem cells into neuroectodermal precursors in adherent<br />

monoculture. Nat Biotechnol 21: 183 – 186<br />

23. Tropepe V, Hitoshi S, Sirard C, Mak TW, Rossant J, van der Kooy D (2001)<br />

Direct Neural Fate Specification from Embryonic <strong>Stem</strong> <strong>Cells</strong>. Neuron 30:<br />

65 – 78<br />

24. Cong L, Ran FA, Cox D, Lin S, Barretto R, Habib N, Hsu PD, Wu X, Jiang<br />

W, Marraffini LA et al (2013) Multiplex genome engineering using<br />

CRISPR/Cas systems. Science 339: 819 – 823<br />

25. Mali P, Yang L, Esvelt KM, Aach J, Guell M, DiCarlo JE, Norville JE, Church<br />

GM (2013) RNA-guided human genome engineering via Cas9. Science<br />

339: 823 – 826<br />

26. Kim J, Woo AJ, Chu J, Snow JW, Fujiwara Y, Kim CG, Cantor AB, Orkin SH<br />

(2010) A Myc network accounts for similarities between embryonic stem<br />

and cancer cell transcription programs. Cell 143: 313 – 324<br />

27. Home P, Saha B, Ray S, Dutta D, Gunewardena S, Yoo B, Pal A, Vivian JL,<br />

Larson M, Petroff M et al (2012) Altered subcellular localization of transcription<br />

factor TEAD4 regulates first mammalian cell lineage commitment.<br />

Proc Natl Acad Sci USA 109: 7362 – 7367<br />

28. Zhao B, Ye X, Yu J, Li L, Li W, Li S, Yu J, Lin JD, Wang C-Y, Chinnaiyan<br />

AM et al (2008) TEAD mediates YAP-dependent gene induction and<br />

growth control. Genes Dev 22: 1962 – 1971<br />

29. Zhao B, Kim J, Ye X, Lai ZC, Guan KL (2009) Both TEAD-binding and<br />

WW domains are required for the growth stimulation and oncogenic<br />

transformation activity of yes-associated protein. Cancer Res 69:<br />

1089 – 1098<br />

30. Hailesellasse SK, Porter CJ, Palidwor G, Perez-Iratxeta C, Muro EM,<br />

Campbell PA, Rudnicki MA, Andrade-Navarro MA (2007) Gene function<br />

in early mouse embryonic stem cell differentiation. BMC Genom 8: 85<br />

31. Xiao JH, Davidson I, Matthes H, Garnier J-M, Chambon P (1991) Cloning,<br />

expression, and transcriptional properties of the human enhancer factor<br />

TEF-1. Cell 65: 551 – 568<br />

528<br />

EMBO reports Vol 17 | No 4 | 2016 ª 2016 The Authors


HaeWon Chung et al Yap1 is required for differentiation of ES cells EMBO reports<br />

32. Mahoney W, Hong J, Yaffe M, Farrance I (2005) The transcriptional coactivator<br />

TAZ interacts differentially with transcriptional enhancer<br />

factor-1 (TEF-1) family members. Biochem J 225: 217 – 225<br />

33. Farrance I, Mar J, Ordahl C (1992) M-CAT binding factor is related to the<br />

SV40 enhancer binding factor, TEF-1. J Biol Chem 267: 17234 – 17240<br />

34. Morin-Kensicki EM, Boone BN, Howell M, Stonebraker JR, Teed J, Alb JG,<br />

Magnuson TR, O’Neal W, Milgram SL (2006) Defects in yolk sac vasculogenesis,<br />

chorioallantoic fusion, and embryonic axis elongation in mice<br />

with targeted disruption of Yap65. Mol Cell Biol 26: 77 – 87<br />

35. Beyer TA, Weiss A, Khomchuk Y, Huang K, Ogunjimi AA, Varelas X, Wrana<br />

JL (2013) Switch enhancers interpret TGF-b and hippo signaling to<br />

control cell fate in human embryonic stem cells. Cell Rep 5: 1611 – 1624<br />

36. Amit M, Carpenter MK, Inokuma MS, Chiu C-P, Harris CP, Waknitz MA,<br />

Itskovitz-Eldor J, Thomson JA (2000) Clonally derived human embryonic<br />

stem cell lines maintain pluripotency and proliferative potential for<br />

prolonged periods of culture. Dev Biol 227: 271 – 278<br />

37. Nichols J, Chambers I, Taga T, Smith A (2001) Physiological rationale for<br />

responsiveness of mouse embryonic stem cells to gp130 cytokines.<br />

Development 128: 2333 – 2339<br />

38. Trapnell C, Williams BA, Pertea G, Mortazavi A, Kwan G, van Baren<br />

MJ, Salzberg SL, Wold BJ, Pachter L (2010) Transcript assembly and<br />

quantification by RNA-Seq reveals unannotated transcripts and<br />

isoform switching during cell differentiation. Nat Biotechnol 28:<br />

511 – 515<br />

ª 2016 The Authors EMBO reports Vol 17 | No 4 | 2016<br />

529


Scientific Report<br />

Resistance of glia-like central and peripheral neural<br />

stem cells to genetically induced mitochondrial<br />

dysfunction—differential effects on neurogenesis<br />

Blanca Díaz-Castro 1,† , Ricardo Pardal 1,2,3,† , Paula García-Flores 1,3 , Verónica Sobrino 1 , Rocío Durán 1 ,<br />

José I Piruat 1,** & José López-Barneo 1,2,3,*<br />

Abstract<br />

Mitochondria play a central role in stem cell homeostasis. Reversible<br />

switching between aerobic and anaerobic metabolism is critical<br />

for stem cell quiescence, multipotency, and differentiation, as<br />

well as for cell reprogramming. However, the effect of mitochondrial<br />

dysfunction on neural stem cell (NSC) function is unstudied.<br />

We have generated an animal model with homozygous deletion of<br />

the succinate dehydrogenase subunit D gene restricted to cells of<br />

glial fibrillary acidic protein lineage (hGFAP-SDHD mouse). Genetic<br />

mitochondrial damage did not alter the generation, maintenance,<br />

or multipotency of glia-like central NSCs. However, differentiation<br />

to neurons and oligodendrocytes (but not to astrocytes) was<br />

impaired and, hence, hGFAP-SDHD mice showed extensive brain<br />

atrophy. Peripheral neuronal populations were normal in hGFAP-<br />

SDHD mice, thus highlighting their non-glial (non hGFAP + ) lineage.<br />

An exception to this was the carotid body, an arterial chemoreceptor<br />

organ atrophied in hGFAP-SDHD mice. The carotid body<br />

contains glia-like adult stem cells, which, as for brain NSCs, are<br />

resistant to genetic mitochondrial damage.<br />

Keywords carotid body stem cells; mitochondrial dysfunction; neural stem<br />

cells; peripheral versus central neurogenesis<br />

Subject Categories <strong>Stem</strong> <strong>Cells</strong>; Neuroscience<br />

DOI 10.15252/embr.201540982 | Received 7 July 2015 | Revised 20 August<br />

2015 | Accepted 21 August 2015 | Published online 21 September 2015<br />

EMBO Reports (2015) 16: 1511–1519<br />

Introduction<br />

Intermediary metabolism plays a critical role in stem cell maintenance<br />

and differentiation. Somatic stem cells in their niches are<br />

normally in a quiescent state and maintained under a predominantly<br />

anaerobic condition, which helps to preserve them from<br />

excessive production of reactive oxygen species and other stressors<br />

[1,2]. Indeed, hypoxia is believed to be an essential environmental<br />

factor for maintenance of the stemness in embryonic and adult<br />

stem cells [3,4]. Upregulation of hypoxia-inducible glycolytic genes<br />

is a hallmark of somatic stem cells [5,6] as well as embryonic and<br />

induced pluripotent stem cells (iPSCs) [7]. In contrast, oxidative<br />

phosphorylation appears to be necessary for hematopoietic stem<br />

cell (HSC) differentiation to mature cells [8]. An inverse metabolic<br />

switch (i.e., involving a change from aerobic to anaerobic metabolism)<br />

has also been reported to occur upon reprogramming of<br />

adult cells to iPSCs [7,9]. Although the role of mitochondria in<br />

pluripotency and differentiation is a topic at the forefront of stem<br />

cell research [2,7], the actual effect of in vivo mitochondrial<br />

dysfunction on stem cell survival and differentiation is poorly<br />

understood. This is particularly evident in the case of neural stem<br />

cells (NSCs), despite the fact that oxidative phosphorylation is<br />

critical for neuronal homeostasis, possibly more than for any other<br />

cell type.<br />

Adult multipotent NSCs exist both in the central (CNS) and the<br />

peripheral (PNS) nervous system. During CNS development, radial<br />

glia, a cell type originated from the primordial neuroepithelium,<br />

serve as stem cells from which many neurons in cortical and subcortical<br />

areas as well as astrocytes and oligodendrocytes are derived<br />

[10,11]. A population of NSCs of astroglial lineage remains in the<br />

subventricular zone (SVZ) and the hippocampus of the adult<br />

mammalian brain, thereby supporting neurogenesis throughout<br />

life [11]. In contrast with the CNS, the PNS derives from a nonhomogeneous<br />

population of neural crest progenitors that migrate<br />

throughout the embryo and give rise to neurons and peripheral glia<br />

during fetal and early postnatal life [12,13]. Multipotent neural crest<br />

progenitor cells in the PNS also persist in adult life, particularly in<br />

the enteric nervous system (ENS) [14,15] and in the carotid body<br />

(CB), a neural crest-derived peripheral chemoreceptor organ located<br />

at the carotid artery bifurcation that grows in response to hypoxia<br />

[16,17]. Whether CB stem cells and other peripheral adult NSCs<br />

1 Instituto de Biomedicina de Sevilla (IBiS), Hospital Universitario Virgen del Rocío, CSIC, Universidad de Sevilla, Seville, Spain<br />

2 Departamento de Fisiología Médica y Biofísica, Universidad de Sevilla, Sevilla, Spain<br />

3 Centro de Investigación Biomédica en Red sobre Enfermedades Neurodegenerativas (CIBERNED), Madrid, Spain<br />

*Corresponding author. Tel: +34 955 923007; E-mail: lbarneo@us.es<br />

**Corresponding author. Tel: +34 955 923088; E-mail: jpiruat-ibis@us.es<br />

† These authors contributed equally to this work<br />

ª 2015 The Authors EMBO reports Vol 16 | No 11 | 2015 1511


EMBO reports Mitochondrial dysfunction in neural stem cells Blanca Díaz-Castro et al<br />

share a similar glial phenotype and metabolic properties with<br />

central neurogenic progenitors is not known.<br />

We have generated a mouse model with conditional deletion of<br />

the gene encoding the membrane anchoring subunit D of succinate<br />

dehydrogenase restricted to cells expressing Cre recombinase under<br />

control of the human glial fibrillary acidic protein (hGFAP)<br />

promoter (hGFAP-SDHD mouse), which is active in mouse radial<br />

glia (see Materials and Methods). It is known that ablation of this<br />

mitochondrial complex II (MCII) gene in catecholaminergic cells<br />

compromises ATP synthesis and results in oxidative stress and<br />

neuronal loss in the brain and PNS [17,18]. This new animal model<br />

has permitted us to examine experimentally the resistance of neural<br />

stem cells to genetically induced mitochondrial dysfunction, which<br />

has remained untested so far. The experiments have also provided<br />

valuable information on the differential origin and properties of<br />

neural stem cells in the CNS and PNS, which might be relevant to<br />

their ability to support neurogenesis in adult life.<br />

Results and Discussion<br />

Genetic modification of mitochondrial function in embryonic<br />

radial glia alters neuronal maturation during<br />

brain development<br />

Mutant (hGFAP-SDHD) mice were viable and did not exhibit any<br />

obvious alterations at birth. At postnatal day (P) 0, brains of<br />

mutant animals had similar appearance as those of control littermates,<br />

although dilation of ventricles was observed in some cases<br />

(Fig EV1). However, during the second week of life, hGFAP-SDHD<br />

mice exhibited a marked phenotype characterized by a lack of<br />

motor coordination, ataxia and decreased body size (Fig EV2A).<br />

Animals rapidly deteriorated and died between P16 and P18. At<br />

P15, the brains of GFAP-SDHD mice displayed notable malformations<br />

and were approximately 50% smaller than those of controls<br />

(Fig 1A). Anatomical and histological differences between brains<br />

of mutant mice and controls are illustrated in Fig 1B–M. Coronal<br />

sections stained with the neuronal marker NeuN at different<br />

rostro-caudal levels indicated marked atrophy of the cerebral<br />

cortex, particularly the dorso-lateral region, and a virtual absence<br />

of the hippocampus and cerebellum in hGFAP-SDHD mice with<br />

respect to controls, which at this age had already developed a<br />

normal adult brain. Detailed cytoarchitectonic or neuronal identification<br />

analyses were outside the scope of this work; however, the<br />

disappearance of cortical layers in the fronto-parietal region and<br />

atrophy of the corpus callosum were clearly evident histological<br />

hallmarks of hGFAP-SDHD mice (Fig 1B–E). In these animals, the<br />

regions normally occupied by the hippocampus and cerebellum<br />

contained only embryonic primordial rudiments of these structures<br />

(Fig 1F–M). The temporo-cortical areas, basal nuclei, and ventrocaudal<br />

brain (diencephalon and brain stem) appeared less affected<br />

(Fig 1B, C, F and G; see sagittal sections and details of structures<br />

at higher magnification in Fig EV2B–K). Staining of P0 and P15<br />

brains of wild-type and mutant mice with antibodies against GFAP<br />

showed in both cases the presence of numerous GFAP + cells<br />

(Fig EV3). Interestingly, glial cells in hGFAP-SDHD mice had a<br />

radial glia-like phenotype, with long processes, which were not<br />

present in normal GFAP + cells (Fig EV3H, I, K and L). Histological<br />

analyses at an intermediate (P5) stage of development are shown<br />

in Fig EV4. We established that the amount of SdhD functional<br />

allele was clearly diminished in P15 hGFAP-SDHD mice in comparison<br />

with homo- or heterozygous normal littermates (Fig 1N).<br />

MCII activity in brain mitochondria of hGFAP-SDHD mice was also<br />

decreased to less than 30% of that in homozygous controls<br />

(Fig 1O), thus confirming the deletion of the SdhD gene in a significant<br />

proportion of brain cells. In accord with the morphological<br />

data, these biochemical differences were not observed when only<br />

the ventral parts of the brains were used for analysis (data not<br />

shown). SdhD mRNA levels and MCII activity were already<br />

decreased in newborn (P0) hGFAP-SDHD brains, indicating that<br />

ablation of the SdhD gene had taken place during embryonic life,<br />

as soon as the hGFAP promoter became active (see Fig EV1M and<br />

N). To ensure that the Cre-mediated loxP recombination had actually<br />

resulted in deletion of the SdhD alleles in GFAP + cells, we<br />

generated control and hGFAP-SDHD mouse lines that expressed<br />

the enhanced green fluorescent protein (EGFP) driven by the<br />

hGFAP promoter (see Materials and Methods). After dissociation,<br />

GFAP + cells were sorted from two different brain areas (cortex<br />

and striatum) (Fig 1P and Q). In both cases, the levels of SdhD<br />

mRNA were reduced to < 20% of the value in control mice (Fig 1R<br />

and S), thus demonstrating that the SdhD gene was ablated in<br />

GFAP + cells from hGFAP-SDHD mice. These findings suggest that<br />

brain NSCs, along with their progeny of neuroblasts and astrocytes<br />

generated before birth, appear to be little affected by mitochondrial<br />

dysfunction. However, neuronal maturation, a process that occurs<br />

in the perinatal period, seems to be highly dependent on proper<br />

mitochondrial oxidative metabolic function.<br />

The extensive brain atrophy in hGFAP-SDHD mice is in fair<br />

agreement with previous lineage-specific tracing experiments showing<br />

that, during development, radial glia are stem cells from which<br />

many central neurons in the cortical and subcortical areas are<br />

derived [10]. Although these data, using Cre/loxP recombination of<br />

reporter genes, could be susceptible to misinterpretation due to the<br />

failure of reporter activity, our experiments, based on recombination<br />

of an intrinsic gene (SdhD) essential for cell homeostasis, fully confirm<br />

these results. Indeed, regional brain atrophy in the hGFAP-<br />

SDHD mice is almost superimposable on the distribution of labeled<br />

(X-gal + ) cells described in lineage tracing experiments using the<br />

same human GFAP promoter to drive Cre recombinase expression<br />

[10]. Relative preservation of ventral brain areas in hGFAP-SDHD<br />

animals can therefore be explained by the differential distribution of<br />

Cre-mediated recombination, which in hGFAP-Cre mice spares<br />

neural cells in this region.<br />

Adult SVZ stem cells are preserved in SDHD-deficient mice, but<br />

survival of newly generated neurons and oligodendrocytes<br />

is impaired<br />

Adult NSCs in the SVZ, which are considered to derive from radial<br />

glia, also have a GFAP + astrocyte-like phenotype [11]. We therefore<br />

investigated whether mitochondrial dysfunction influenced the<br />

maintenance, proliferation, and/or multipotency of these stem cells.<br />

Neurosphere assays of dispersed SVZ cells obtained from control<br />

and hGFAP-SDHD animals showed that the number of clonal colonies<br />

generated was the same for the two mouse strains (Fig 2A and B).<br />

Self-renewal capacity, as evidenced by the number of secondary<br />

1512<br />

EMBO reports Vol 16 | No 11 | 2015 ª 2015 The Authors


Blanca Díaz-Castro et al Mitochondrial dysfunction in neural stem cells EMBO reports<br />

A B C D E<br />

N<br />

F G H<br />

I<br />

O<br />

J<br />

K<br />

L<br />

M<br />

P<br />

R<br />

S<br />

Q<br />

Figure 1.<br />

ª 2015 The Authors EMBO reports Vol 16 | No 11 | 2015<br />

1513


EMBO reports Mitochondrial dysfunction in neural stem cells Blanca Díaz-Castro et al<br />

◀<br />

Figure 1. Brain structures and mitochondrial complex II activity in wild-type and hGFAP-SDHD mice.<br />

A Photographs of control (top) and hGFAP-SDHD (bottom) brains at P15. Scale bar: 5 mm.<br />

B–M Brain sections of postnatal (P15) control (B, D, F, H, J, L) and hGFAP-SDHD (C, E, G, I, K, M) mice immunostained with NeuN antibody. Scale bars: 1 mm (B, C, F, G,<br />

J, K), 500 lm (D, E, H, I), and 200 lm (L, M). Ctx: cortex; cc: corpus callosum; CPu: caudate putamen; hc: hippocampus; Dg: dentate gyrus.<br />

N Relative levels of functional SdhD allele in the entire brain (P15) as determined by qPCR of genomic DNA. r.u.: relative units (n = 4).<br />

O Succinate–ubiquinone oxidoreductase (SQR) activity of mitochondria isolated from the entire brain (P15) (n = 6).<br />

P, Q Representative profiles of sorted astrocytes from cortex and striatum of mice carrying the hGFAP-EGFP transgene (blue lines). Animals without the construct (black<br />

line) were used as blanks to determine the selection gates.<br />

R, S Relative SdhD mRNA levels in sorted GFAP-expressing astrocytes from the cortex and striatum of control and hGFAP-SDHD mice.<br />

Data information: Data are presented as mean SEM (n = 4–8). Statistical significance: *P ≤ 0.05; **P ≤ 0.01; ***P ≤ 0.001. The ANOVA test with the appropriate<br />

post hoc analysis was applied. See also Figs EV1–EV4.<br />

neurospheres, was also unaffected by the SdhD deletion (Fig EV5).<br />

These data suggest that SVZ stem cells are well preserved in<br />

hGFAP-SDHD mice with defective mitochondria. Nonetheless, the<br />

size of the primary (Fig 2A and C) and secondary (Fig EV5C)<br />

neurospheres was smaller in hGFAP-SDHD mice compared to<br />

controls (see below). Quantitative PCR analysis of SdhD mRNA<br />

from neurospheres confirmed the ablation of the SdhD gene in SVZ<br />

cells of genetically modified mice (Fig 2D). Furthermore, SVZ stem<br />

A B C D<br />

E<br />

F<br />

G<br />

H<br />

I<br />

Figure 2. Effects of mitochondrial dysfunction on subventricular zone neural stem cells.<br />

A Bright-field images of neurospheres obtained from SVZ neural stem cells of P15 control and hGFAP-SDHD mouse brains. Scale bars: 500 lm.<br />

B, C Neurosphere (NS) forming efficiency (B) and core diameter (C) in cultures grown from SVZ of P15 control and hGFAP-SDHD mice (n = 6 cultures/mice for each genotype).<br />

D Quantitative RT–PCR detection of SdhD expression levels in SVZ neurospheres of wild-type (flox/+) and mutant (flox/ , and flox/ cre) mice (n = 3–4 mice on each group).<br />

E Immunofluorescence detection of the neuronal marker Tuj1 in SVZ neural stem cell-derived adherent cultures from P15 control or hGFAP-SDHD brains. Nuclei were<br />

counterstained with DAPI. Scale bar: 25 lm.<br />

F Number of Tuj1 + neurons generated in SVZ neural stem cell adherent cultures in vitro for several days (n = 8 cultures/mice for each genotype).<br />

G Number of GFAP + astrocytes present in adherent cultures of SVZ neurospheres illustrating the resistance of glial cells to the loss of mitochondrial function (n = 5<br />

independent cultures).<br />

H Immunofluorescence detection of the oligodendrocyte marker O4 in SVZ neural stem cell-derived adherent cultures from P15 control or hGFAP-SDHD brains. Nuclei<br />

were counterstained with DAPI. Scale bar: 25 lm.<br />

I Number of O4 + neurons generated in SVZ neural stem cell adherent cultures in vitro for several days (n = 5–7 cultures/mice for each genotype). See also Fig EV5.<br />

Data information: Data are presented as mean SEM. *P ≤ 0.05. The two-tailed Student’s t-test was applied.<br />

1514<br />

EMBO reports Vol 16 | No 11 | 2015 ª 2015 The Authors


Blanca Díaz-Castro et al Mitochondrial dysfunction in neural stem cells EMBO reports<br />

cells from hGFAP-SDHD animals were able to differentiate into<br />

Tuj1 + neurons in vitro; however, survival of these newly generated<br />

neurons was compromised (Fig 2E and F). In contrast, the differentiation<br />

and survival of astrocytes were practically unaltered in SdhDdefective<br />

neurospheres (Fig 2E and G). Survival of differentiated<br />

oligodendrocytes was also decreased in preparations from hGFAP-<br />

SDHD mice (Fig 2H and I). These observations provide direct experimental<br />

support for the notion that, as other progenitor cell types<br />

[5,6], central NSCs rely predominantly on anaerobic metabolism and<br />

thus can survive and maintain their normal function even after<br />

severe mitochondrial damage. However, maintenance of mature<br />

neurons and oligodendrocytes, but not GFAP + astrocytes, is<br />

absolutely dependent on a correctly functioning mitochondrial<br />

metabolism.<br />

hGFAP-SDHD mice have normal PNS development but impaired<br />

carotid body neurogenesis<br />

Despite the gross brain developmental alterations exhibited by the<br />

hGFAP-SDHD mice, they showed an apparently normal PNS. The<br />

morphology and number of neurons in the adrenal medulla, superior<br />

cervical ganglion (SCG), ENS, and dorsal root ganglia, all of<br />

which are derived from neural crest progenitor cells [12], were similar<br />

in P15 hGFAP-SDHD mice compared with controls (Fig 3A–H).<br />

We checked that, as in the CNS glia, GFAP + Schwann cells in<br />

peripheral nerves were also lacking the SdhD alleles, thereby<br />

demonstrating recombination of the floxed SdhD alleles in PNS<br />

structures (Fig 3I and J). Interestingly, ablation of the SdhD gene<br />

did not seem to compromise survival of these peripheral glial cells<br />

(Fig 3K), although extensive loss of TH + adrenal chromaffin cells<br />

and SCG neurons has been observed in previous studies with TH + -<br />

specific SdhD-deficient (TH-SDHD) mice [18]. Quantitative PCR and<br />

histological analyses showed no differences in the SCG of hGFAP-<br />

SDHD and control mice (Fig EV6). A small, and non-significant,<br />

decrease in SdhD mRNA observed in the SCG of hGFAP-SDHD mice<br />

with respect to controls probably reflected SdhD deletion in the<br />

small population of GFAP + glial cells existing in this structure<br />

(Fig EV6A and B). Lack of hGFAP-Cre-dependent activity in peripheral<br />

neurons was further demonstrated in vivo using a LacZ reporter<br />

mice [16], which showed the absence of b-gal staining in TH +<br />

neurons from SCG (Fig EV6C and D). Taken together, these data<br />

suggest that NSCs giving rise to PNS neurons do not have the<br />

GFAP + glial origin that is characteristic of brain NSCs. An exception<br />

to this general rule appears to be the CB, which in newborn (P0)<br />

hGFAP-SDHD mice had similar size to that of control animals<br />

(Fig 4A–C). However, the normal postnatal increase in TH + glomus<br />

cell number and the volume of TH + CB parenchyma observed in<br />

controls were abolished in hGFAP-SDHD mice (Fig 4A–C). Despite<br />

the decrease in CB neuron-like glomus cell number, the population<br />

of CB stem cells in the hGFAP-SDHD mice remained unaltered as<br />

judged by the ability of dispersed CB stem cells to form growing<br />

neurospheres (Fig 4D–F). These observations indicate that unlike<br />

other structures in the PNS, postnatal development of the CB<br />

depends on glia-like GFAP + stem cells that are also resistant to<br />

mitochondrial dysfunction. A population of these glia-like cells<br />

remains dormant in the adult CB and sustains its adaptive growth<br />

upon exposure to chronic hypoxia [16,17]. Other stem cells in the<br />

adult PNS do not seem to contribute to the physiological homeostasis<br />

of the specific tissues where they are located [14,15]. Therefore,<br />

NSCs of glial lineage that reside in the adult CB constitute a<br />

neurogenic niche of similar characteristics to those existing in the<br />

adult brain and confer upon the CB the anatomical and functional<br />

plasticity required for acclimatization to hypoxia [17]. The expression<br />

of GFAP in brain and CB neural stem cells could in fact be the<br />

manifestation of a quiescent state in these cells that is necessary to<br />

maintain their capacity to generate neurons in adulthood.<br />

Mitochondria and neural stem cell proliferation<br />

and differentiation<br />

We have shown that low reliance on mitochondrial oxidative<br />

phosphorylation seems to be a common generic feature of NSCs<br />

during development and in the adult niches. Nonetheless, the way<br />

different stem cell classes achieve a homeostatic anaerobic metabolism<br />

may be variable. Several groups have reported a smaller<br />

number of mitochondria and modifications of their shape in multipotent<br />

HSCs in comparison to more-committed bone marrow<br />

progenitors [2]. Moreover, the high glycolytic flux of quiescent<br />

HSCs residing in hypoxic niches seems to rely on the upregulation<br />

of hypoxia-inducible transcription factors (particularly HIF-1a) and<br />

the subsequent induction of glycolytic enzymes and pyruvate<br />

dehydrogenase kinase to blunt pyruvate-dependent oxidative phosphorylation<br />

[5,6,19]. However, it has been reported that although<br />

stabilization of HIF-1a and HIF-2a is necessary to initiate the metabolic<br />

switch in the early stages of the reprogramming of human<br />

cells to iPSCs, the stabilization of HIF-2a during latter stages<br />

represses reprogramming [9]. HIF-1a-deficient NSCs have normal<br />

mitochondria although they are resistant to hypoxia and have high<br />

glycolytic activity [20]. Recently, we showed that maintenance and<br />

clonal growth of CB stem cells in vitro is unaffected by a broad<br />

range of O 2 tensions [17]. This insensitivity to O 2 tension is also<br />

observed in neurosphere cultures of adult SVZ stem cells [21] as<br />

well as in embryonic stem cells [3]. Although we have observed<br />

that survival of central (SVZ) and peripheral (CB) neural stem cells<br />

Figure 3. Normal development and survival of peripheral neurons in hGFAP-SDHD mice.<br />

A–H Immunofluorescence detection of neuronal markers TH (A, C) and HuC/D (E, G) in the adrenal medulla (A), superior cervical ganglion (C), enteric ganglia at the level of the<br />

distal small intestine (E), and dorsal root ganglion (G) of P15 control and hGFAP-SDHD mice. Scale bars: 200 lm (A, C), 100 lm (E, G). Cell number in adrenal medulla (B),<br />

superior cervical ganglion (D), enteric ganglia (F), and dorsal root ganglion (H) in the same animal models. Data are presented as mean SEM (n = 4–7 per group).<br />

I Immunofluorescence detection of GFAP expression in cross sections of peripheral sciatic nerve illustrating the normal shape of Schwann cells forming myelin<br />

sheaths in P15 control and hGFAP-SDHD mice. Scale bar: 50 lm.<br />

J Results of quantitative RT–PCR to detect SdhD expression in the peripheral nerves of wild-type (flox/+) and mutant (flox/ , and flox/ cre) mice. Data are presented<br />

as mean SEM (n = 3 flox/+, n = 6 flox/ , and n = 8 flox/ cre mice). Statistical significance: *P ≤ 0.05; **P ≤ 0.01. The two-tailed Student’s t-test was applied.<br />

K Number of myelin sheaths per unit area in sciatic nerves of P15 control and mutant mice. Data are presented as mean SEM (n = 4 mice for each genotype). See<br />

also Fig EV6.<br />

▸<br />

ª 2015 The Authors EMBO reports Vol 16 | No 11 | 2015<br />

1515


EMBO reports Mitochondrial dysfunction in neural stem cells Blanca Díaz-Castro et al<br />

A<br />

B<br />

C<br />

D<br />

E<br />

F<br />

G<br />

H<br />

I J K<br />

Figure 3.<br />

1516<br />

EMBO reports Vol 16 | No 11 | 2015 ª 2015 The Authors


Blanca Díaz-Castro et al Mitochondrial dysfunction in neural stem cells EMBO reports<br />

A<br />

B<br />

D<br />

C<br />

E<br />

characteristic properties of stem cells regardless of O 2 tension<br />

levels. Therefore, besides the hypoxic stabilization of HIF, other<br />

intrinsic mechanisms and/or niche factors might contribute to the<br />

anaerobic metabolic state associated with the maintenance of NSC<br />

multipotency.<br />

Although mitochondrial dysfunction was not seen to alter NSC<br />

maintenance and multipotency in the present study, it exerted a<br />

marked effect on the survival of their central and peripheral neural<br />

progeny. Defective mitochondrial metabolism seemed to impair<br />

radial glial cell differentiation, as in P15 animals they appeared in a<br />

state typical of the prenatal brain. However, these mutated radial<br />

glia cells were able to differentiate into neurons, oligodendrocytes,<br />

and astrocytes. Brains from neonatal wild-type and hGFAP-SDHD<br />

mice showed similar sizes, structures, and NeuN + cell densities,<br />

suggesting that neuronal maturation occurring during the first postnatal<br />

weeks rather than neuroblast differentiation is the process that<br />

was actually compromised in mitochondria-defective animals. On<br />

the other hand, we observed the appearance of neurons and<br />

oligodendrocytes in differentiation assays of SVZ stem cells from<br />

hGFAP-SDHD mice, although these cells did not survive longer than<br />

a few days. Understanding the role of mitochondria on stemness<br />

and the switching from quiescence to activity in stem cells located<br />

in the different niches may help in the development of new therapeutic<br />

strategies against cancer stem cells.<br />

Materials and Methods<br />

Figure 4. Impairment of carotid body postnatal maturation with<br />

maintenance of adult stem cells in hGFAP-SDHD mice.<br />

A Immunofluorescence detection of TH expression in newborn (P0) and P15<br />

wild-type (control) and hGFAP-SDHD mouse carotid body (CB). Boundaries of<br />

the CB parenchyma are indicated by the dotted lines. Scale bar: 50 lm.<br />

B, C Number of TH + cells (B) and size (C) per CB at different ages in control<br />

and hGFAP-SDHD mice. Data are presented as mean SEM (n = 3–5<br />

mice in each group). *P ≤ 0.05; **P ≤ 0.01. The two-tailed Student’s<br />

t-test was applied.<br />

D Bright-field images of CB neurospheres obtained from control and<br />

mutant mice at P15. Scale bar: 200 lm.<br />

E, F Neurosphere forming efficiency (E) and diameter (F) of floating cultures<br />

of dispersed CB cells from P15 control and hGFAP-SDHD mice. Data are<br />

presented as mean SEM (n = 6 cultures/mice for each genotype).<br />

is not affected by mitochondrial dysfunction, proliferation of SVZ<br />

progenitors (as estimated by the size of neurosphere cores) was<br />

reduced in preparations from hGFAP-SDHD mice. This could<br />

indicate that the high rate of in vitro proliferation of central<br />

progenitors (in comparison with progenitors in CB neurospheres)<br />

makes them partially dependent on energy obtained by oxidative<br />

phosphorylation. Taken together, these data suggest that a high<br />

glycolytic flux and resistance to mitochondrial dysfunction are<br />

F<br />

Animals<br />

The hGFAP-SDHD strain with SdhD flox/ hGFAP-CRE genotype was<br />

obtained by breeding the previously reported SdhD-flox mouse<br />

strain [18] with a mouse line expressing Cre under control of the<br />

human GFAP promoter, which is active in mouse radial glia<br />

[10,22]. Littermates with SdhD flox/+ and SdhD flox/ genotypes lacking<br />

CRE recombinase are referred to as flox/+ and flox/ , respectively.<br />

Where indicated, results from both genotypes were pooled<br />

and assigned to a control group since no differences between them<br />

were found for the phenotypes tested. Routine genotyping was<br />

performed for the SdhD alleles and the CRE gene by PCR as previously<br />

reported [10,22]. hGFAP-SDHD mice carrying the enhanced<br />

green fluorescence protein (EGFP) under control of the human<br />

GFAP promoter were obtained by breeding with the hGFAP-EGFP<br />

strain [23]. Mice were housed under temperature-controlled conditions<br />

(22°C) on a 12-h light/dark cycle, and provided with food<br />

and water ad libitum. Mice were housed and treated according to<br />

the animal care guidelines of the European Community Council<br />

(86/609/EEC), as well as institutional guidelines approved by the<br />

ethics committee of the Hospital Universitario Virgen del Rocio<br />

and the University of Seville (see Appendix Supplementary<br />

Materials and Methods).<br />

Histochemistry and immunocytochemistry<br />

For detection of NeuN, GFAP, TH, O4, HuC/D, Tuj1, and b-gal in<br />

tissue sections, neurospheres, and dissociated cells, we used standard<br />

staining procedures. Specific details are given in the Appendix<br />

Supplementary Materials and Methods.<br />

ª 2015 The Authors EMBO reports Vol 16 | No 11 | 2015<br />

1517


EMBO reports Mitochondrial dysfunction in neural stem cells Blanca Díaz-Castro et al<br />

Tissue dissociation, cell sorting, and neurosphere assays<br />

Dissociated cells were obtained from brain and CB tissue following<br />

standard procedures. Cell sorting and the generation of SVZ and CB<br />

neurospheres were carried out as indicated previously [16]. See<br />

Appendix Supplementary Materials and Methods.<br />

SdhD DNA and mRNA analyses<br />

DNA and RNA analyses were performed as described previously<br />

[18]. See Appendix Supplementary Materials and Methods.<br />

Mitochondria isolation and complex II activity<br />

Mitochondrial complex II activity was determined according to<br />

Piruat et al [24] with slight modifications. See Appendix Supplementary<br />

Materials and Methods.<br />

Statistics<br />

Data are presented as mean standard error (SEM). Statistical<br />

significance was assessed by ANOVA with appropriate post hoc<br />

analysis. For paired groups, either a two-tailed Student’s t-test with<br />

a Levene test for homogeneity of variances in the case of normal<br />

distribution, or the nonparametric Mann–Whitney U-test in the case<br />

of non-normal distributions, was applied. Normal distribution was<br />

assessed by Shapiro–Wilk test. PASW18 software was used for<br />

statistical analysis.<br />

Expanded View for this article is available online:<br />

http://embor.embopress.org<br />

Acknowledgements<br />

This research is supported by grants from the Botín Foundation (JL-B), the<br />

Spanish Ministry of Science and Innovation SAF program (JL-B, RP and JIP), the<br />

European Research Council (RP and JL-B), and the Junta de Andalucía (JIP). We<br />

thank Alberto Castejón, Valentina Annese, and María José Castro for technical<br />

assistance.<br />

Author contributions<br />

BD-C, RP, PG-F, VS, RD, and JIP performed experiments. BD-C, RP, JIP, and JL-B<br />

designed experiments and contributed to generate a draft of the manuscript.<br />

JL-B coordinated the project.<br />

Conflict of interest<br />

The authors declare that they have no conflict of interest.<br />

References<br />

1. Suda T, Takubo K, Semenza GL (2011) Metabolic regulation of hematopoietic<br />

stem cells in the hypoxic niche. Cell <strong>Stem</strong> Cell 9: 298 – 310<br />

2. Xu X, Duan S, Yi F, Ocampo A, Liu GH, Izpisua Belmonte JC (2013)<br />

Mitochondrial regulation in pluripotent stem cells. Cell Metab 18:<br />

325 – 332<br />

3. Ezashi T, Das P, Roberts RM (2005) Low O 2 tensions and the prevention<br />

of differentiation of hES cells. Proc Natl Acad Sci USA 102:<br />

4783 – 4788<br />

4. Mohyeldin A, Garzon-Muvdi T, Quinones-Hinojosa A (2010) Oxygen in<br />

stem cell biology: a critical component of the stem cell niche. Cell <strong>Stem</strong><br />

Cell 7: 150 – 161<br />

5. Simsek T, Kocabas F, Zheng J, Deberardinis RJ, Mahmoud AI, Olson EN,<br />

Schneider JW, Zhang CC, Sadek HA (2010) The distinct metabolic profile<br />

of hematopoietic stem cells reflects their location in a hypoxic niche.<br />

Cell <strong>Stem</strong> Cell 7: 380 – 390<br />

6. Takubo K, Nagamatsu G, Kobayashi CI, Nakamura-Ishizu A, Kobayashi<br />

H, Ikeda E, Goda N, Rahimi Y, Johnson RS, Soga T et al (2013)<br />

Regulation of glycolysis by Pdk functions as a metabolic checkpoint<br />

for cell cycle quiescence in hematopoietic stem cells. Cell <strong>Stem</strong> Cell<br />

12: 49 – 61<br />

7. Folmes CD, Nelson TJ, Dzeja PP, Terzic A (2012) Energy metabolism<br />

plasticity enables stemness programs. Ann N Y Acad Sci 1254: 82 – 89<br />

8. Inoue S, Noda S, Kashima K, Nakada K, Hayashi J, Miyoshi H (2010)<br />

Mitochondrial respiration defects modulate differentiation but not<br />

proliferation of hematopoietic stem and progenitor cells. FEBS Lett 584:<br />

3402 – 3409<br />

9. Mathieu J, Zhou W, Xing Y, Sperber H, Ferreccio A, Agoston Z,<br />

Kuppusamy KT, Moon RT, Ruohola-Baker H (2014) Hypoxia-inducible<br />

factors have distinct and stage-specific roles during reprogramming of<br />

human cells to pluripotency. Cell <strong>Stem</strong> Cell 14: 592 – 605<br />

10. Malatesta P, Hack MA, Hartfuss E, Kettenmann H, Klinkert W, Kirchhoff<br />

F, Gotz M (2003) Neuronal or glial progeny: regional differences in radial<br />

glia fate. Neuron 37: 751 – 764<br />

11. Kriegstein A, Alvarez-Buylla A (2009) The glial nature of embryonic and<br />

adult neural stem cells. Annu Rev Neurosci 32: 149 – 184<br />

12. Le Douarin N, Dulac C, Dupin E, Cameron-Curry P (1991) Glial cell<br />

lineages in the neural crest. Glia 4: 175 – 184<br />

13. Morrison SJ, White PM, Zock C, Anderson DJ (1999) Prospective<br />

identification, isolation by flow cytometry, and in vivo self-renewal<br />

of multipotent mammalian neural crest stem cells. Cell 96: 737 – 749<br />

14. Joseph NM, He S, Quintana E, Kim YG, Nunez G, Morrison SJ (2011)<br />

Enteric glia are multipotent in culture but primarily form glia in the<br />

adult rodent gut. J Clin Invest 121: 3398 – 3411<br />

15. Laranjeira C, Sandgren K, Kessaris N, Richardson W, Potocnik A, Vanden<br />

Berghe P, Pachnis V (2011) Glial cells in the mouse enteric nervous<br />

system can undergo neurogenesis in response to injury. J Clin Invest<br />

121: 3412 – 3424<br />

16. Pardal R, Ortega-Saenz P, Duran R, Lopez-Barneo J (2007) Glia-like stem<br />

cells sustain physiologic neurogenesis in the adult mammalian carotid<br />

body. Cell 131: 364 – 377<br />

17. Platero-Luengo A, Gonzalez-Granero S, Duran R, Diaz-Castro B, Piruat JI,<br />

Garcia-Verdugo JM, Pardal R, Lopez-Barneo J (2014) AnO 2 -sensitive<br />

glomus cell-stem cell synapse induces carotid body growth in chronic<br />

hypoxia. Cell 156: 291 – 303<br />

18. Diaz-Castro B, Pintado CO, Garcia-Flores P, Lopez-Barneo J, Piruat JI<br />

(2012) Differential impairment of catecholaminergic cell maturation and<br />

survival by genetic mitochondrial complex II dysfunction. Mol Cell Biol<br />

32: 3347 – 3357<br />

19. Romero-Moya D, Bueno C, Montes R, Navarro-Montero O, Iborra FJ, Lopez<br />

LC, Martin M, Menendez P (2013) Cord blood-derived CD34 + hematopoietic<br />

cells with low mitochondrial mass are enriched in hematopoietic<br />

repopulating stem cell function. Haematologica 98: 1022 – 1029<br />

20. Candelario KM, Shuttleworth CW, Cunningham LA (2013) Neural stem/<br />

progenitor cells display a low requirement for oxidative metabolism<br />

independent of hypoxia inducible factor-1alpha expression. J Neurochem<br />

125: 420 – 429<br />

1518<br />

EMBO reports Vol 16 | No 11 | 2015 ª 2015 The Authors


Blanca Díaz-Castro et al Mitochondrial dysfunction in neural stem cells EMBO reports<br />

21. d’Anglemont de Tassigny X., Sirerol-Piquer M.S., Gómez-Pinedo U., Pardal<br />

R., Bonilla S., Capilla-Gonzalez V., López-López I., De la Torre-Laviana F.J.,<br />

García-Verdugo J.M., López-Barneo J. (2015). Resistance of subventricular<br />

neural stem cells to chronic hypoxemia despite structural<br />

disorganization of the germinal center and impairment of neuronal and<br />

oligodendrocyte survival. Hypoxia 3: 15 – 33.<br />

22. Zhuo L, Theis M, Alvarez-Maya I, Brenner M, Willecke K, Messing A<br />

(2001) hGFAP-cre transgenic mice for manipulation of glial and<br />

neuronal function in vivo. Genesis 31: 85 – 94<br />

23. Nolte C, Matyash M, Pivneva T, Schipke CG, Ohlemeyer C, Hanisch UK,<br />

Kirchhoff F, Kettenmann H (2001) GFAP promoter-controlled EGFPexpressing<br />

transgenic mice: a tool to visualize astrocytes and astrogliosis<br />

in living brain tissue. Glia 33: 72 – 86<br />

24. Piruat JI, Pintado CO, Ortega-Saenz P, Roche M, Lopez-Barneo J (2004)<br />

The mitochondrial SDHD gene is required for early embryogenesis, and<br />

its partial deficiency results in persistent carotid body glomus cell<br />

activation with full responsiveness to hypoxia. Mol Cell Biol 24:<br />

10933 – 10940<br />

ª 2015 The Authors EMBO reports Vol 16 | No 11 | 2015<br />

1519


Scientific Report<br />

UTX inhibits EMT-induced breast CSC properties by<br />

epigenetic repression of EMT genes in cooperation<br />

with LSD1 and HDAC1<br />

Hee-Joo Choi 1,† , Ji-Hye Park 2,† , Mikyung Park 3 , Hee-Young Won 1 , Hyeong-seok Joo 1 , Chang Hoon Lee 3 ,<br />

Jeong-Yeon Lee 2,* & Gu Kong 1,2,**<br />

Abstract<br />

The histone H3K27 demethylase, UTX, is a known component of<br />

the H3K4 methyltransferase MLL complex, but its functional association<br />

with H3K4 methylation in human cancers remains largely<br />

unknown. Here we demonstrate that UTX loss induces epithelial–<br />

mesenchymal transition (EMT)-mediated breast cancer stem cell<br />

(CSC) properties by increasing the expression of the SNAIL, ZEB1<br />

and ZEB2 EMT transcription factors (EMT-TFs) and of the transcriptional<br />

repressor CDH1. UTX facilitates the epigenetic silencing of<br />

EMT-TFs by inducing competition between MLL4 and the H3K4<br />

demethylase LSD1. EMT-TF promoters are occupied by c-Myc and<br />

MLL4, and UTX recognizes these proteins, interrupting their transcriptional<br />

activation function. UTX decreases H3K4me2 and H3<br />

acetylation at these promoters by forming a transcriptional repressive<br />

complex with LSD1, HDAC1 and DNMT1. Taken together, our<br />

findings indicate that UTX is a prominent tumour suppressor that<br />

functions as a negative regulator of EMT-induced CSC-like properties<br />

by epigenetically repressing EMT-TFs.<br />

Keywords breast CSC; EMT; UTX<br />

Subject Categories Chromatin, Epigenetics, Genomics & Functional<br />

Genomics; Cancer<br />

DOI 10.15252/embr.201540244 | Received 13 February 2015 | Revised 10 July<br />

2015 | Accepted 14 July 2015 | Published online 24 August 2015<br />

EMBO Reports (2015) 16: 1288–1298<br />

Introduction<br />

A ubiquitously transcribed tetratricopeptide repeat X chromosome<br />

(UTX) functions as a histone demethylase towards di- and tri-methylated<br />

histone H3 on lysine 27 (H3K27me2/H3K27me3) [1]. This<br />

demethylase is also a known component of the mixed-lineage<br />

leukaemia (MLL) 2/3 H3K4 methyltransferase complex, although its<br />

catalytic effect on H3K4me remains unclear [2–4]. Polycomb group<br />

(PcG)-dependent H3K27 methylation has redundant functions in<br />

many biological processes, including stem cell regulation and<br />

tumour development [5–7], although less evidence supports<br />

the functional role of UTX in these cellular processes. Several<br />

studies suggest that UTX regulates somatic cell reprogramming and<br />

embryonic development [8,9], as well as tumour suppression in<br />

leukaemia [10,11]. Furthermore, genomewide analyses have identified<br />

various somatic mutations and deletions of UTX in several<br />

human cancers, including breast, bladder, and renal cancers and<br />

leukaemia [12–14]. However, the demethylation-dependent or<br />

demethylation-independent molecular function of UTX in human<br />

cancers has not been clearly explained, and it remains controversial<br />

whether UTX has tumour-suppressive activity [15–17].<br />

Accumulating evidence shows that many histone methyltransferases<br />

and demethylases are involved in the regulation of<br />

cancer stem cells (CSCs), a small population of stem-like cancer<br />

cells possessing tumorigenic and metastatic capacities [18,19].<br />

Moreover, recent evidence suggests that regulators of the epithelial–<br />

mesenchymal transition (EMT), which repress E-cadherin (encoded<br />

by CDH1) transcriptionally, such as SNAIL, SLUG, TWIST, ZEB1<br />

and ZEB2, play a key role in inducing CSC self-renewal [20–23].<br />

Notably, these EMT transcription factors (EMT-TFs) cooperate with<br />

various chromatin-modifying enzymes (including the histone<br />

methylases G9a and Set8 and the H3K4/K9 demethylase LSD1) as<br />

well as histone deacetylases and DNMTs to form a repressive<br />

complex for CDH1 [24–26]. UTX is also known to interact physically<br />

with transcriptional and epigenetic regulators, such as MLL 2/3<br />

H3K4 methyltransferase [2–4], histone acetyltransferase and the<br />

SWI/SNF complex [27,28], during the assembly of multi-protein<br />

complexes. However, the molecular relationship between UTX and<br />

EMT-TFs in the epigenetic regulation of EMT-associated CSC properties<br />

remains unknown.<br />

Herein, we found that UTX is critical for inhibiting EMT-induced<br />

breast CSC properties by suppressing the c-Myc-dependent<br />

1 Department of Pathology, College of Medicine, Hanyang University, Seoul, Korea<br />

2 Institute for Bioengineering and Biopharmaceutical Research (IBBR), Hanyang University, Seoul, Korea<br />

3 College of Pharmacy, Dongguk University, Seoul, Korea<br />

*Corresponding author. Tel: +82 2 2220 0634; Fax: +82 2 2295 1091; E-mail: jy2jy2@hanyang.ac.kr<br />

**Corresponding author. Tel: +82 2 2290 8251; Fax: +82 2 2295 1091; E-mail: gkong@hanyang.ac.kr<br />

† These authors contributed equally to this work<br />

1288<br />

EMBO reports Vol 16 | No 10 | 2015 ª 2015 The Authors


Hee-Joo Choi et al Repression of EMT-induced CSC by UTX EMBO reports<br />

transcription of SNAIL, ZEB1 and ZEB2 in cooperation with H3K4<br />

demethylase LSD1, HDAC1 and DNMT1. These findings suggest a<br />

novel epigenetic mechanism for UTX during its inhibition of EMT-<br />

TFs as a tumour suppressor in human breast cancer.<br />

Results and Discussion<br />

UTX loss enhances CSC-like properties and the EMT in human<br />

breast cancer<br />

Given the evidence that Drosophila and murine Utx antagonize<br />

Notch signalling and activate tumour suppressor Rb proteins, previous<br />

studies have suggested a putative tumour-suppressive function<br />

for UTX in human cancer [15,16]. Consistently, in human breast<br />

cancer, UTX somatic mutations have been discovered that might be<br />

associated with the loss of UTX function [12]. In contrast, a recent<br />

study suggested that UTX has oncogenic potential in human breast<br />

cancer cell lines [17]. To clearly determine the role of UTX in<br />

human cancer, we evaluated the expression of UTX in several cell<br />

lines derived from several stages of breast tumour development.<br />

UTX expression was downregulated in the cancer cell lines<br />

compared with normal or immortalized cells (Fig 1A). Notably,<br />

UTX mRNA was expressed at lower levels in CD44 + /CD24 /ESA +<br />

cells with stem cell-like characteristics (Fig 1B). Thus, we investigated<br />

whether UTX affects stem-like phenotypes in normal and<br />

malignant breast epithelial cells. The percentage of CD44 + /CD24 /<br />

ESA + cells in the cell population was increased by UTX knockdown<br />

in the immortalized MCF10A cells and in breast cancer cell lines<br />

highly expressing UTX; however, this population declined in<br />

response to UTX overexpression in MDA-MB-231 cells having a low<br />

UTX expression (Figs 1C and EV1A and B). UTX also negatively<br />

regulated mammosphere formation and anchorage-independent<br />

growth (Fig 1D and E). Consistently, in vivo breast tumour xenograft<br />

assays indicated that mice bearing UTX-overexpressing MDA-<br />

MB-231 tumours exhibited retarded tumour initiation and growth<br />

(Table EV1, Fig EV1C), whereas mice injected with UTX-deficient<br />

SK-BR-3 cells presented more rapid tumour formation and growth<br />

than control mice (Table EV1, Fig EV1D). Moreover, UTX knockdown<br />

induced the neoplastic transformation of the non-tumorigenic<br />

MCF10A cell lines in vivo (Fig EV1E). Considering the ability of<br />

CD44 + /CD24 /ESA + cells to transform mammary epithelial cells<br />

[22,23], these data suggest that UTX loss might promote breast<br />

tumorigenesis by expanding stem-like cells and enhancing their selfrenewal<br />

and tumour-initiating capacities.<br />

Recent studies have shown that EMT is essential for generating<br />

CD44 + /CD24 low/ stem-like cells that can convert breast epithelial<br />

cells into tumorigenic cells and can confer breast cancer aggressiveness<br />

[20–23]. Consistent with these findings, UTX knockdown in<br />

MCF10A cells resulted in a mesenchymal-like sporadic long spindle<br />

phenotype (Fig 2A). We also observed a reduced expression of<br />

epithelial markers including E-cadherin but an increased expression<br />

of mesenchymal markers in these cells (Fig 2B–D). Furthermore,<br />

UTX-overexpressing MDA-MB-231 cells exhibited epithelial-like<br />

morphology (Fig 2A), increased epithelial marker expression and<br />

decreased mesenchymal marker expression (Fig 2B–D). Contrary to<br />

a recent study showing decreased invasion after UTX knockdown in<br />

breast cancer [17], EMT promotion resulting from UTX loss<br />

conferred invasion and migration abilities on non-invasive MCF10A<br />

cells, whereas UTX overexpression inhibited these effects in invasive<br />

MDA-MB-231 cells (Fig 2E and F). Consistently, a meta-analysis<br />

of breast cancer patient survival performed using the Kaplan–Meier<br />

Plotter [29] showed that high UTX expression is associated with<br />

favourable overall survival (P < 0.001, log-rank test; n = 741),<br />

relapse-free survival (P


EMBO reports Repression of EMT-induced CSC by UTX Hee-Joo Choi et al<br />

A<br />

B<br />

C<br />

D<br />

E<br />

Figure 1. Loss of UTX enhances CSC-like properties in human breast cancer.<br />

A Screening of protein expression levels by immunoblotting in normal and cancerous breast cell lines.<br />

B UTX mRNA expression in CD44 + /CD24 /ESA + cells (CSC) and other cells (non-CSC) was sorted using flow cytometry from the indicated cell lines and analysed using<br />

qRT-PCR. ***P < 0.001 versus Non-CSC; two-tailed unpaired t-test.<br />

C Flow cytometry analysis was used to measure CD44 + /CD24 /ESA + cell populations in the indicated cell lines. The lower right quadrant indicates the CD44 + /CD24 /<br />

ESA + cells (x-axis, APC-conjugated CD44; y-axis, PE-conjugated CD24). CON si, non-targeting siRNA; UTX si, UTX siRNA; CON, pLVX-puro empty vector; UTX, pLVXpuro-UTX.<br />

*P < 0.05, **P < 0.01, ***P < 0.001 versus controls; two-tailed unpaired t-test.<br />

D, E Mammosphere formation (D) and soft agar colony formation (E) in cells that underexpress or overexpress UTX. After 21 days (E) or as indicated (D), the numbers of<br />

spheres (D) or colonies (E) (> 100 lm diameter) were counted. shCON, control shRNA; shUTX, UTX shRNA. *P < 0.05, **P < 0.01, ***P < 0.001 versus controls;<br />

two-tailed unpaired t-test.<br />

Data information: The data shown in (B–E) represent the means SD of n = 3 independent experiments.<br />

Source data are available online for this figure.<br />

1290<br />

EMBO reports Vol 16 | No 10 | 2015 ª 2015 The Authors


Hee-Joo Choi et al Repression of EMT-induced CSC by UTX EMBO reports<br />

A<br />

B<br />

C<br />

D<br />

E<br />

F<br />

Figure 2.<br />

Loss of UTX induces the EMT and invasion of breast cancer cells.<br />

A Representative bright-field images of cells that underexpress or overexpress UTX. Scale bars: 100 lm.<br />

B, C The expression of epithelial and mesenchymal markers in the indicated cells was measured using immunoblotting (B) and qRT-PCR (C).<br />

D The immunofluorescence staining of E-cadherin (red) and vimentin (green) was detected using confocal microscopy and quantified. DAPI (blue) was used for<br />

nuclear staining. Scale bars: 10 lm.<br />

E, F Chamber transwell assays of cellular invasion or migration by the indicated cells. Invasion and migration of test cells are expressed relative to values for control<br />

cells. Scale bars: 100 lm.<br />

Data information: Error bars in (C–F) indicate the SD (n = 3 independent experiments). *P < 0.05, **P < 0.01, ***P < 0.001 versus controls; two-tailed unpaired t-test.<br />

Source data are available online for this figure.<br />

ª 2015 The Authors EMBO reports Vol 16 | No 10 | 2015<br />

1291


EMBO reports Repression of EMT-induced CSC by UTX Hee-Joo Choi et al<br />

A<br />

B<br />

C<br />

D<br />

E<br />

F<br />

G<br />

Figure 3.<br />

1292<br />

EMBO reports Vol 16 | No 10 | 2015 ª 2015 The Authors


Hee-Joo Choi et al Repression of EMT-induced CSC by UTX EMBO reports<br />

◀<br />

Figure 3. UTX depletion inhibits E-cadherin transcription by recruiting EMT-related transcription factors.<br />

A To measure CDH1 promoter activity, a luciferase (luc) assay was performed in UTX-knockdown MCF10A and in UTX-overexpressing MDA-MB-231 cells that were cotransfected<br />

with either CDH1 wild-type (WT) or E-box mutant (Mut)-luc and b-galactosidase constructs for 24 h.<br />

B, C The expression of EMT-TFs was measured in the indicated cells using immunoblotting (B) and qRT-PCR (C).<br />

D Immunoblotting with cytoplasmic and nuclear fractions to analyse EMT-TF expression. SP1 and a-tubulin were used for loading controls for nuclear and<br />

cytoplasmic proteins, respectively.<br />

E Immunofluorescence analysis showing the expression level and cellular localization of SNAIL, ZEB1 and ZEB2 in UTX-knockdown MCF10A cells. Nuclei were stained<br />

with DAPI. Scale bar: 10 lm.<br />

F ChIP analysis showing the indicated histone methylation status and gene recruitment to the CDH1 promoter in UTX-deficient MCF10A cells. IgG, control IgG for<br />

negative control.<br />

G Flow cytometry analysis of the CD44 + /CD24 /ESA + population in MCF10A cells that were transfected with non-targeting siRNA (CON si) or with UTX-targeting<br />

siRNA (UTX si), together with the indicated EMT-TF siRNAs for 48 h.<br />

Data information: The data shown in (A, C, E, F and G) represent the means SD of n = 3 independent experiments. *P < 0.05, **P < 0.01, ***P < 0.001 versus controls;<br />

†† P < 0.01 versus UTX si; ns, no significance (two-tailed unpaired t-test for A, C, E and F; one-way ANOVA and Scheffe’s post hoc test for G).<br />

Source data are available online for this figure.<br />

these conditions (Fig EV2E and F). Although H3K27me3 was<br />

decreased by EZH2 knockdown in the CDH1 promoter, recruitment<br />

of EMT-TFs to this region was maintained continuously in the UTXknockdown<br />

cells (Fig EV2G). Therefore, these data imply that UTX<br />

regulates EMT and stemness in an EZH2 activity-independent<br />

manner; this finding supports UTX as a crucial inhibitor of EMTassociated<br />

breast tumorigenesis that might overcome the function of<br />

EZH2 in breast cancer.<br />

UTX transcriptionally represses EMT-TF expression by altering<br />

c-Myc-dependent and c-Myc-independent<br />

epigenetic modifications<br />

Little is known about the epigenetic and transcriptional mechanisms<br />

that occur in the genomic regions of EMT-TFs; in contrast,<br />

the CDH1 regulatory mechanism has been well established. Therefore,<br />

we next investigated the molecular mechanism by which UTX<br />

transcriptionally regulates EMT-TFs to inhibit the properties of<br />

EMT-induced CSCs. SNAIL, ZEB1 and ZEB2 were found to have<br />

E-box motifs that bind c-Myc in their proximal promoters; therefore,<br />

we examined the effect of UTX on c-Myc activity towards EMT-<br />

TFs. Although the expression of c-Myc was not altered by UTX<br />

(Fig EV3A), the recruitment of c-Myc to these regions was<br />

enhanced by UTX knockdown in MCF10A cells and was inhibited<br />

by UTX overexpression in MDA-MB-231 cells, as assessed using a<br />

ChIP assay (Figs 4A and EV3B–D). In addition, UTX bound directly<br />

to these regions. Previous studies have shown that many UTXoccupied<br />

genes not only exhibit altered H3K27 methylation states<br />

but also altered H3K4 methylation states [2,3,31], although the<br />

mechanism by which H3K4 methylation is regulated via UTX<br />

remains unclear. To identify whether UTX-associated histone modifications<br />

participate in the transcriptional regulation of EMT-TFs,<br />

we examined the modification status of histone proteins including<br />

H3K27me3 and H3K4me2 at these promoters. Although the level of<br />

H3K27me3 remained unchanged, the H3K4me2 and H3 acetylation<br />

(ac) active histone markers were enriched in the SNAIL, ZEB1 and<br />

ZEB2 promoters in UTX-deficient MCF10A cells. Moreover, the<br />

recruitment of p300 histone acetyltransferase (HAT) and the dissociation<br />

of HDAC1 and DNMT1 transcriptional repressive markers<br />

were observed in these regions. Consistent results were found in<br />

UTX-overexpressing MDA-MB-231 cells. Furthermore, co-immunoprecipitation<br />

(co-IP) of exogenous and endogenous proteins<br />

showed that UTX physically interacts with c-Myc, HDAC1 and<br />

DNMT1 (Fig 4B and C). To determine whether c-Myc acts as a<br />

bridge for the recognition of UTX by the target promoter and consequent<br />

epigenetic changes, we analysed epigenetic changes at the<br />

EMT-TF promoters following siRNA-mediated c-Myc depletion.<br />

Although c-Myc was required for changes in H3ac by UTX, it did<br />

not affect the enrichment of UTX and H3K4me2 in these promoters<br />

(Fig EV4A and B). However, flow cytometry following c-Myc<br />

siRNA treatment showed that c-Myc depletion reverses UTX-knockdown-induced<br />

stem cell expansion (Fig 4D). Taken together, these<br />

results show that UTX represses the transcription of EMT-TFs by<br />

inhibiting c-Myc-dependent histone acetylation and inducing<br />

c-Myc-independent H3K4 demethylation. Together with the results<br />

of a previous study showing the regulation of SNAIL by c-Myc [32],<br />

our data suggest that c-Myc is a common transcription factor for the<br />

induction of SNAIL, ZEB1 and ZEB2 expression through its role in<br />

coordinating with epigenetic modifiers that involve an active gene<br />

state. Furthermore, UTX might be a key component for the epigenetic<br />

silencing of these EMT-TFs to antagonize c-Myc function.<br />

UTX cooperates with LSD1 demethylase and inhibits MLLmediated<br />

H3K4me to epigenetically repress EMT-TFs<br />

To further explain how UTX negatively regulates the H3K4me2 level<br />

at the EMT-TF promoters in a c-Myc-independent manner, we investigated<br />

possible histone modification enzymes targeting H3K4me<br />

that might cooperate with UTX in regulating EMT-TFs. Consistent<br />

with previous findings [2,4,17], we observed that UTX physically<br />

interacts with the H3K4-specific methyltransferase MLL4 (Fig 5A).<br />

Interestingly, UTX also bound H3K4 demethylase LSD1 and<br />

enhanced the association between LSD1 and MLL4, as confirmed by<br />

co-IP results. Furthermore, ChIP analysis revealed that MLL4 consistently<br />

bound to the EMT-TF promoters regardless of UTX expression,<br />

whereas LSD1 recruitment to these regions occurred only in<br />

the presence of UTX expression in both UTX-overexpressing and<br />

UTX-knockdown cells, thereby suggesting UTX-dependent LSD1<br />

recruitment and H3K4 demethylation at the EMT-TF promoters<br />

(Fig 5B). These data also imply that UTX is not a binding partner of<br />

MLL4 but rather inhibits H3K4 methylation by facilitating competition<br />

between MLL4 and LSD1.<br />

LSD1 role as an oncogene has been well described, despite<br />

controversy regarding in its role in breast cancer [33–35]. To define<br />

the potential role of LSD1 in the UTX-mediated repression of EMT-<br />

TFs, we inhibited LSD1 expression using lentiviral shRNA in<br />

ª 2015 The Authors EMBO reports Vol 16 | No 10 | 2015<br />

1293


EMBO reports Repression of EMT-induced CSC by UTX Hee-Joo Choi et al<br />

A<br />

B<br />

C<br />

D<br />

Figure 4.<br />

UTX suppresses EMT-TF expression by inducing c-Myc-dependent and c-Myc-independent epigenetic modification.<br />

A ChIP analysis showing the histone methylation status and recruitment of the indicated proteins within the EMT-TF promoters. *P < 0.05, **P < 0.01, ***P < 0.001<br />

versus controls (two-tailed unpaired t-test).<br />

B, C The interaction between UTX and the indicated proteins was confirmed by co-IP in 293 T cells transfected with the indicated constructs (B) and in UTXoverexpressing<br />

MDA-MB-231 cells (C).<br />

D MCF10A cells were transfected with CON si or UTX si, together with c-Myc si. The CD44 + /CD24 /ESA + population was then evaluated using flow cytometry.<br />

**P < 0.01 versus CON si; †† P < 0.01 versus UTX si (one-way ANOVA and Scheffe’s post hoc test).<br />

Data information: The results shown in (A and D) represent the means SD of three independent experiments.<br />

Source data are available online for this figure.<br />

1294<br />

EMBO reports Vol 16 | No 10 | 2015 ª 2015 The Authors


Hee-Joo Choi et al Repression of EMT-induced CSC by UTX EMBO reports<br />

A<br />

B<br />

C<br />

D<br />

E<br />

Figure 5.<br />

ª 2015 The Authors EMBO reports Vol 16 | No 10 | 2015<br />

1295


EMBO reports Repression of EMT-induced CSC by UTX Hee-Joo Choi et al<br />

◀<br />

Figure 5. UTX cooperates with LSD1 demethylase to inhibit H3K4me in EMT-TF promoters.<br />

A The binding of UTX to the indicated proteins was analysed using co-IP experiments involving lysates from 293T cells transfected with the indicated constructs.<br />

B ChIP analysis showing the recruitment of LSD1 and MLL4 to the EMT-TF promoters. *P < 0.05, **P < 0.01, ***P < 0.001 versus controls (two-tailed unpaired<br />

t-test).<br />

C, D UTX-overexpressing MDA-MB-231 cells were infected with LSD1 shRNA viruses; immunoblotting (C) and ChIP analyses (D) were then performed to confirm the<br />

expression of the indicated proteins and the epigenetic changes at the SNAIL promoters. *P < 0.05, **P < 0.01, ***P < 0.001 versus. CON/shCON; † P < 0.05,<br />

†† P < 0.01, ††† P < 0.001 versus UTX/shCON (one-way ANOVA and Scheffe’s post hoc test for c-Myc IP; Welch’s test and Dunnett’s T3 post hoc test for others).<br />

E Schematic summary of the regulation of EMT-TFs by UTX.<br />

Data information: The data shown in (B and D) represent the means SD of n = 3 independent experiments.<br />

Source data are available online for this figure.<br />

control or UTX-overexpressing MDA-MB-231 cells. Consistent with<br />

previous findings that LSD1 promotes EMT [35,36], we found that<br />

LSD1 knockdown decreased SNAIL and ZEB1 expression and<br />

increased E-cadherin expression in control MDA-MB-231 cells<br />

expressing low levels of UTX (Fig 5C). In contrast, the loss of LSD1<br />

recovered the expression of EMT-TFs that were inhibited by UTX<br />

and subsequently abolished E-cadherin expression (Fig 5C), indicating<br />

that UTX overexpression causes LSD1 to function as a negative<br />

regulator of EMT-TFs. These paradoxical results support the<br />

critical role of UTX in determining the oncogenic potential of LSD1<br />

in human breast cancer. Because LSD1 is a component of the<br />

CoREST and NuRD transcriptional repressive complexes, which<br />

contain HDACs [33,37], we assumed that UTX might recruit not<br />

only LSD1 but also LSD1-containing multiple repressive complexes<br />

to its target chromatin. Indeed, the ChIP analysis results showed<br />

that the repression of H3K4me2 and H3ac by the recruitment of<br />

LSD1 and HDAC1 was recovered in UTX-overexpressing cells by<br />

LSD1 knockdown (Figs 5D and EV4C and D), implying that LSD1 is<br />

required for the formation of a transcriptional repressive complex<br />

with UTX and HDAC1 at the EMT-TF promoters. However, LSD1<br />

could not alter the inhibitory effect of UTX on c-Myc binding to<br />

these promoters. Collectively, our results indicate that UTX epigenetically<br />

inhibits the transcription of EMT-TFs by cooperating<br />

with LSD1/HDAC1 in competition with c-Myc/MLL4 in the<br />

regulation of EMT and CSCs. Contrary to the findings of Kim and<br />

colleagues [17], these data suggest that UTX might utilize MLL4 as<br />

a recognition signal for chromatin binding but inhibits MLL4 activity<br />

towards H3K4 by recruiting the LSD1 demethylase-containing<br />

transcriptional repressive complex to silence the expression of<br />

EMT-TFs in breast cancer.<br />

In this study, we demonstrated that UTX plays an important<br />

role in breast tumour suppression as an epigenetic regulator of<br />

EMT-TFs. UTX negatively regulated EMT-induced breast CSC properties<br />

by transcriptionally suppressing SNAIL, ZEB1 and ZEB2 in<br />

an H3K27 demethylation activity-independent manner (Fig 5E). In<br />

the absence of UTX, EMT-TFs are transcriptionally activated by<br />

recruiting c-Myc/p300 to the EMT-TF promoters. MLL4, the H3K4<br />

methyltransferase, was occupied and methylated H3K4 in these<br />

regions. When UTX was present, the recruitment of c-Myc and<br />

p300 was disrupted, and LSD1/HDAC1/DNMT1 recognized the<br />

target region by promoting UTX binding to MLL4. The UTXcontaining<br />

repressive complex inhibits the H3K4me and H3ac,<br />

resulting in gene silencing of the EMT-TFs. Collectively, these findings<br />

suggest that UTX is a novel epigenetic silencer of EMT-TFs<br />

that represses EMT-associated breast CSC-like properties in an<br />

H3K27 methylation-independent manner by cooperating with H3K4<br />

demethylase LSD1 and HDAC1.<br />

Materials and Methods<br />

Flow cytometry<br />

Breast CSC populations were isolated using flow cytometry as<br />

described previously [38], based on the expression of surface markers<br />

(CD44 + /CD24 /ESA + ). To evaluate UTX expression in CSC and<br />

non-CSC populations from breast cell lines, cell populations were<br />

divided, stained using the appropriate antibodies and collected;<br />

the cells were then analysed using a FACSAria flow cytometer (BD<br />

Biosciences, Franklin Lakes, NJ, USA).<br />

Mammosphere formation<br />

MCF10A cells (1 × 10 4 cells/well), T-47D cells (5 × 10 3 cells/well),<br />

SK-BR-3 cells (1 × 10 4 cells/well) and MDA-MB-231 cells (5 × 10 3<br />

cells/well) were cultured under previously described conditions<br />

[38]. After 3, 5 and 7 days, mammosphere formation was analysed<br />

and quantified using an inverted microscope.<br />

ChIP–qPCR assay<br />

ChIP assays were performed using a ChIP assay kit according to the<br />

manufacturer’s instructions (Upstate Biotechnology, Lake Placid,<br />

NY, USA). ChIP signal enrichment was determined using real-time<br />

quantitative PCR (qPCR) (signal/input ratio) conducted using the<br />

7300 Real-Time PCR System and the SYBR Green Master Mix<br />

(Applied Biosystems, Foster City, CA, USA).<br />

Statistical analysis<br />

The statistical significance of differences between the control and<br />

experimental groups was analysed using the two-tailed unpaired<br />

t-test after confirming that data were normally distributed based on<br />

the Shapiro–Wilk test as implemented in the SPSS (version 12.0; SPSS<br />

Inc., Chicago, IL, USA). The Levene’s test was used to verify equality<br />

of variances. For multiple comparisons, one-way ANOVA followed<br />

by Scheffe’s post hoc test (for equal variances) or Welch’s test with<br />

Dunnett’s T3 post hoc test (for unequal variances) was performed. To<br />

estimate the frequency of the tumour-initiating ability, limiting<br />

dilution assay results were calculated using L-Calc software.<br />

P-values < 0.05 were considered to indicate statistical significance.<br />

For further methods, see Appendix Supplementary Methods.<br />

Expanded View for this article is available online:<br />

http://embor.embopress.org<br />

1296<br />

EMBO reports Vol 16 | No 10 | 2015 ª 2015 The Authors


Hee-Joo Choi et al Repression of EMT-induced CSC by UTX EMBO reports<br />

Acknowledgements<br />

This study was supported by the National Research Foundation of Korea<br />

(NRF), which is funded by the Korean Government (No. 2010-0020879).<br />

Author contributions<br />

GK and J-YL designed and supervised the project. H-JC, J-YL and GK wrote the<br />

manuscript. H-JC and J-HP performed the experiments, analysed and organized<br />

the data. MP, H-YW, H-SJ and C-HL conducted the experiments and<br />

analysed the data.<br />

Conflict of interest<br />

The authors declare that they have no conflict of interest.<br />

References<br />

1. Agger K, Cloos PA, Christensen J, Pasini D, Rose S, Rappsilber J, Issaeva I,<br />

Canaani E, Salcini AE, Helin K (2007) UTX and JMJD3 are histone H3K27<br />

demethylases involved in HOX gene regulation and development. Nature<br />

449: 731 – 734<br />

2. Issaeva I, Zonis Y, Rozovskaia T, Orlovsky K, Croce CM, Nakamura T,<br />

Mazo A, Eisenbach L, Canaani E (2007) Knockdown of ALR (MLL2) reveals<br />

ALR target genes and leads to alterations in cell adhesion and growth.<br />

Mol Cell Biol 27: 1889 – 1903<br />

3. Lee MG, Villa R, Trojer P, Norman J, Yan KP, Reinberg D, Di Croce L,<br />

Shiekhattar R (2007) Demethylation of H3K27 regulates polycomb<br />

recruitment and H2A ubiquitination. Science 318: 447 – 450<br />

4. Cho YW, Hong T, Hong S, Guo H, Yu H, Kim D, Guszczynski T, Dressler GR,<br />

Copeland TD, Kalkum M et al (2007) PTIP associates with MLL3- and<br />

MLL4-containing histone H3 lysine 4 methyltransferase complex. J Biol<br />

Chem 282: 20395 – 20406<br />

5. Richly H, Aloia L, Di Croce L (2011) Roles of the Polycomb group proteins<br />

in stem cells and cancer. Cell Death Dis 2: e204<br />

6. Cao Q, Yu J, Dhanasekaran SM, Kim JH, Mani RS, Tomlins SA,<br />

Mehra R, Laxman B, Cao X, Yu J et al (2008) Repression of E-cadherin<br />

by the polycomb group protein EZH2 in cancer. Oncogene 27:<br />

7274 – 7284<br />

7. Chang CJ, Yang JY, Xia W, Chen CT, Xie X, Chao CH, Woodward WA,<br />

Hsu JM, Hortobagyi GN, Hung MC (2011) EZH2 promotes expansion of<br />

breast tumor initiating cells through activation of RAF1-beta-catenin<br />

signaling. Cancer Cell 19: 86 – 100<br />

8. Wang C, Lee JE, Cho YW, Xiao Y, Jin Q, Liu C, Ge K (2012) UTX regulates<br />

mesoderm differentiation of embryonic stem cells independent of H3K27<br />

demethylase activity. Proc Natl Acad Sci USA 109: 15324 – 15329<br />

9. Mansour AA, Gafni O, Weinberger L, Zviran A, Ayyash M, Rais Y, Krupalnik<br />

V, Zerbib M, Amann-Zalcenstein D, Maza I et al (2012) The H3K27<br />

demethylase Utx regulates somatic and germ cell epigenetic reprogramming.<br />

Nature 488: 409 – 413<br />

10. Ntziachristos P, Tsirigos A, Welstead GG, Trimarchi T, Bakogianni S, Xu L,<br />

Loizou E, Holmfeldt L, Strikoudis A, King B et al (2014) Contrasting roles<br />

of histone 3 lysine 27 demethylases in acute lymphoblastic leukaemia.<br />

Nature 514: 513 – 517<br />

11. Van der Meulen J, Sanghvi V, Mavrakis K, Durinck K, Fang F, Matthijssens F,<br />

Rondou P, Rosen M, Pieters T, Vandenberghe P et al (2015) The<br />

H3K27me3 demethylase UTX is a gender-specific tumor suppressor in<br />

T-cell acute lymphoblastic leukemia. Blood 125: 13 – 21<br />

12. van Haaften G, Dalgliesh GL, Davies H, Chen L, Bignell G, Greenman C,<br />

Edkins S, Hardy C, O’Meara S, Teague J et al (2009) Somatic mutations<br />

of the histone H3K27 demethylase gene UTX in human cancer. Nat<br />

Genet 41: 521 – 523<br />

13. Gui Y, Guo G, Huang Y, Hu X, Tang A, Gao S, Wu R, Chen C, Li X, Zhou L<br />

et al (2011) Frequent mutations of chromatin remodeling genes in transitional<br />

cell carcinoma of the bladder. Nat Genet 43: 875 – 878<br />

14. Dalgliesh GL, Furge K, Greenman C, Chen L, Bignell G, Butler A, Davies H,<br />

Edkins S, Hardy C, Latimer C et al (2010) Systematic sequencing of renal<br />

carcinoma reveals inactivation of histone modifying genes. Nature 463:<br />

360 – 363<br />

15. Herz HM, Madden LD, Chen Z, Bolduc C, Buff E, Gupta R, Davuluri R,<br />

Shilatifard A, Hariharan IK, Bergmann A (2010) The H3K27me3 demethylase<br />

dUTX is a suppressor of Notch- and Rb-dependent tumors in Drosophila.<br />

Mol Cell Biol 30: 2485 – 2497<br />

16. Terashima M, Ishimura A, Yoshida M, Suzuki Y, Sugano S, Suzuki T (2010)<br />

The tumor suppressor Rb and its related Rbl2 genes are regulated by Utx<br />

histone demethylase. Biochem Biophys Res Commun 399: 238 – 244<br />

17. Kim JH, Sharma A, Dhar SS, Lee SH, Gu B, Chan CH, Lin HK, Lee MG<br />

(2014) UTX and MLL4 coordinately regulate transcriptional programs for<br />

cell proliferation and invasiveness in breast cancer cells. Cancer Res 74:<br />

1705 – 1717<br />

18. Visvader JE (2011) <strong>Cells</strong> of origin in cancer. Nature 469: 314 – 322<br />

19. Tam WL, Weinberg RA (2013) The epigenetics of epithelial-mesenchymal<br />

plasticity in cancer. Nat Med 19: 1438 – 1449<br />

20. Mani SA, Guo W, Liao MJ, Eaton EN, Ayyanan A, Zhou AY, Brooks M,<br />

Reinhard F, Zhang CC, Shipitsin M et al (2008) The epithelial-mesenchymal<br />

transition generates cells with properties of stem cells. Cell 133:<br />

704 – 715<br />

21. Scheel C, Weinberg RA (2012) Cancer stem cells and epithelialmesenchymal<br />

transition: concepts and molecular links. Semin Cancer<br />

Biol 22: 396 – 403<br />

22. Lim S, Becker A, Zimmer A, Lu J, Buettner R, Kirfel J (2013) SNAI1-<br />

mediated epithelial-mesenchymal transition confers chemoresistance<br />

and cellular plasticity by regulating genes involved in cell death and<br />

stem cell maintenance. PLoS ONE 8: e66558<br />

23. Hollier BG, Tinnirello AA, Werden SJ, Evans KW, Taube JH, Sarkar TR,<br />

Sphyris N, Shariati M, Kumar SV, Battula VL et al (2013) FOXC2 expression<br />

links epithelial-mesenchymal transition and stem cell properties in<br />

breast cancer. Cancer Res 73: 1981 – 1992<br />

24. Lin Y, Wu Y, Li J, Dong C, Ye X, Chi YI, Evers BM, Zhou BP (2010) The<br />

SNAG domain of Snail1 functions as a molecular hook for recruiting<br />

lysine-specific demethylase 1. EMBO J 29: 1803 – 1816<br />

25. Dong C, Wu Y, Yao J, Wang Y, Yu Y, Rychahou PG, Evers BM, Zhou BP<br />

(2012) G9a interacts with Snail and is critical for Snail-mediated<br />

E-cadherin repression in human breast cancer. J Clin Invest 122:<br />

1469 – 1486<br />

26. Yang F, Sun L, Li Q, Han X, Lei L, Zhang H, Shang Y (2012) SET8<br />

promotes epithelial-mesenchymal transition and confers TWIST dual<br />

transcriptional activities. EMBO J 31: 110 – 123<br />

27. Tie F, Banerjee R, Conrad PA, Scacheri PC, Harte PJ (2012) Histone<br />

demethylase UTX and chromatin remodeler BRM bind directly to CBP and<br />

modulate acetylation of histone H3 lysine 27. Mol Cell Biol 32: 2323 – 2334<br />

28. Miller SA, Mohn SE, Weinmann AS (2010) Jmjd3 and UTX play a<br />

demethylase-independent role in chromatin remodeling to regulate<br />

T-box family member-dependent gene expression. Mol Cell 40: 594 – 605<br />

29. Gyorffy B, Lanczky A, Eklund AC, Denkert C, Budczies J, Li Q, Szallasi Z<br />

(2010) An online survival analysis tool to rapidly assess the effect of<br />

22,277 genes on breast cancer prognosis using microarray data of 1,809<br />

patients. Breast Cancer Res Treat 123: 725 – 731<br />

ª 2015 The Authors EMBO reports Vol 16 | No 10 | 2015<br />

1297


EMBO reports Repression of EMT-induced CSC by UTX Hee-Joo Choi et al<br />

30. De Craene B, Berx G (2013) Regulatory networks defining EMT<br />

during cancer initiation and progression. Nat Rev Cancer 13:<br />

97 – 110<br />

31. Wang JK, Tsai MC, Poulin G, Adler AS, Chen S, Liu H, Shi Y, Chang HY<br />

(2010) The histone demethylase UTX enables RB-dependent cell fate<br />

control. Genes Dev 24: 327 – 332<br />

32. Smith AP, Verrecchia A, Faga G, Doni M, Perna D, Martinato F, Guccione<br />

E, Amati B (2009) A positive role for Myc in TGFbeta-induced Snail transcription<br />

and epithelial-to-mesenchymal transition. Oncogene 28:<br />

422 – 430<br />

33. Wang Y, Zhang H, Chen Y, Sun Y, Yang F, Yu W, Liang J, Sun L,<br />

Yang X, Shi L et al (2009) LSD1 is a subunit of the NuRD complex<br />

and targets the metastasis programs in breast cancer. Cell 138:<br />

660 – 672<br />

34. Lim S, Janzer A, Becker A, Zimmer A, Schule R, Buettner R, Kirfel J (2010)<br />

Lysine-specific demethylase 1 (LSD1) is highly expressed in ER-negative<br />

breast cancers and a biomarker predicting aggressive biology. Carcinogenesis<br />

31: 512 – 520<br />

35. Lin T, Ponn A, Hu X, Law BK, Lu J (2010) Requirement of the histone<br />

demethylase LSD1 in Snai1-mediated transcriptional repression during<br />

epithelial-mesenchymal transition. Oncogene 29: 4896 – 4904<br />

36. Ferrari-Amorotti G, Fragliasso V, Esteki R, Prudente Z, Soliera AR,<br />

Cattelani S, Manzotti G, Grisendi G, Dominici M, Pieraccioli M et al<br />

(2013) Inhibiting interactions of lysine demethylase LSD1 with snail/slug<br />

blocks cancer cell invasion. Cancer Res 73: 235 – 245<br />

37. Lee MG, Wynder C, Cooch N, Shiekhattar R (2005) An essential role for<br />

CoREST in nucleosomal histone 3 lysine 4 demethylation. Nature 437:<br />

432 – 435<br />

38. Won HY, Lee JY, Shin DH, Park JH, Nam JS, Kim HC, Kong G (2012) Loss<br />

of Mel-18 enhances breast cancer stem cell activity and tumorigenicity<br />

through activating Notch signaling mediated by the Wnt/TCF pathway.<br />

FASEB J 26: 5002 – 5013<br />

1298<br />

EMBO reports Vol 16 | No 10 | 2015 ª 2015 The Authors


Review<br />

“Histones and Chromatin” Review Series<br />

Chromatin remodeling and bivalent histone<br />

modifications in embryonic stem cells<br />

Arigela Harikumar & Eran Meshorer *<br />

Abstract<br />

Pluripotent embryonic stem cells (ESCs) are characterized by<br />

distinct epigenetic features including a relative enrichment of<br />

histone modifications related to active chromatin. Among these is<br />

tri-methylation of lysine 4 on histone H3 (H3K4me3). Several thousands<br />

of the H3K4me3-enriched promoters in pluripotent cells also<br />

contain a repressive histone mark, namely H3K27me3, a situation<br />

referred to as “bivalency”. While bivalent promoters are not unique<br />

to pluripotent cells, they are relatively enriched in these cell types,<br />

largely marking developmental and lineage-specific genes which<br />

are silent but poised for immediate action. The H3K4me3 and<br />

H3K27me3 modifications are catalyzed by lysine methyltransferases<br />

which are usually found within, although not entirely<br />

limited to, the Trithorax group (TrxG) and Polycomb group (PcG)<br />

protein complexes, respectively, but these do not provide selective<br />

bivalent specificity. Recent studies highlight the family of ATPdependent<br />

chromatin remodeling proteins as regulators of bivalent<br />

domains. Here, we discuss bivalency in general, describe the<br />

machineries that catalyze bivalent chromatin domains, and portray<br />

the emerging connection between bivalency and the action of different<br />

families of chromatin remodelers, namely INO80, esBAF, and<br />

NuRD, in pluripotent cells. We posit that chromatin remodeling<br />

proteins may enable “bivalent specificity”, often selectively acting<br />

on, or selectively depleted from, bivalent domains.<br />

Keywords chromatin; chromatin remodeling; embryonic stem cells; epigenetics;<br />

histone modifications<br />

DOI 10.15252/embr.201541011 | Received 13 July 2015 | Revised 1 October<br />

2015 | Accepted 5 October 2015 | Published online 9 November 2015<br />

EMBO Reports (2015) 16: 1609–1619<br />

See the Glossary for abbreviations used in this article.<br />

Introduction<br />

The genetic information of a living cell is stored within the DNA.<br />

However, additional layers of regulation provide the epigenetic<br />

information, which, in concert with transcription factors, enables<br />

the same primary DNA sequence to confer different identities to<br />

different cell types, developmental stages, disease states, etc.<br />

In eukaryotes, the DNA is wrapped around a histone octamer<br />

comprised of a pair of each of the core histones H2A, H2B, H3, and<br />

H4, which together form the nucleosome. Nucleosomes are the<br />

basic repeating units of chromatin, and they are arranged in a higher<br />

order chromatin structure through the binding of linker histones, H1<br />

proteins, between adjacent nucleosomes. Thus, despite having the<br />

same genetic makeup, different transcriptional outcomes of different<br />

cell types of the same organism are achieved through a variety of<br />

epigenetic modifications including DNA methylation, histone posttranslational<br />

modifications (PTMs), chromatin organization, etc. So far,<br />

several histone modifications with physiological importance have<br />

been identified, such as acetylation, methylation, phosphorylation,<br />

ubiquitylation, sumoylation, ADP ribosylation, deimination, proline<br />

isomerization, biotinylation, citrullination, and more [1–3]. Apart<br />

from influencing local chromatin structure, these modifications are<br />

also recognized by specific adaptor proteins which in turn recruit<br />

protein complexes and thereby affect gene regulation. Some of these<br />

histone marks such as H3K4me3, H3K9ac, and H3K14ac are associated<br />

with actively transcribed genes and some other modifications,<br />

for example, H3K27me3 and H3K9me3, are enriched within<br />

repressed regions. Activation and repression are believed to occur,<br />

at least partly, through charge-mediated chromatin decompaction<br />

and chromodomain-containing protein binding, respectively [3,4].<br />

Interestingly, a subset of promoters associated with both activating<br />

(H3K4me3) and repressive (H3K27me3) marks, also known as<br />

“bivalent” modifications, has been discovered in mouse embryonic<br />

stem cells (ESCs) [5,6]. Several recent reviews thoroughly covered<br />

the field of bivalency, especially in pluripotent ESCs [7–10]. Here,<br />

we focus on the emerging link between bivalent histone modifications<br />

and ATP-dependent chromatin remodeling in ESCs. The<br />

term “chromatin remodeling” is often used to describe any change<br />

or modification to chromatin including histone modification. Here,<br />

by “chromatin remodeling”, we specifically refer to the action of the<br />

family of ATP-dependent chromatin remodeling factors, described<br />

below. When other forms of chromatin structure alterations are<br />

referred to, we describe the specific mode of alteration or modification.<br />

We will briefly summarize initial and recent experiments<br />

establishing the existence and role of bivalent domains in undifferentiated<br />

ESCs and during differentiation, and will argue that both<br />

H3K4me3 and H3K27me3, and especially the cross talk between<br />

them, are intimately linked with chromatin remodeling in pluripotent<br />

ESCs.<br />

Department of Genetics, Institute of Life Sciences and The Edmond and Lily Safra Center for Brain Sciences, The Hebrew University of Jerusalem, Jerusalem, Israel<br />

*Corresponding author. Tel: +972 2 6585161; E-mail: meshorer@huji.ac.il<br />

ª 2015 The Authors EMBO reports Vol 16 | No 12 | 2015 1609


EMBO reports Chromatin remodeling, bivalent histone marks, and ESCs Arigela Harikumar & Eran Meshorer<br />

Glossary<br />

ASH2L Ash2 (absent, small, or homeotic)-like (Drosophila)<br />

protein<br />

BAF250a AT-rich interactive domain 1A (Swi1 like) protein<br />

Bmi1 B lymphoma Mo-MLV insertion region 1 protein<br />

BRG1 SWI/SNF-related, matrix-associated, actin-dependent<br />

regulator of chromatin, subfamily a, member 4<br />

Cbx7 chromobox homolog 7 protein<br />

CFP1 CXXC finger 1 (PHD domain) protein<br />

CHARGE coloboma, heart defect, atresia choanae (also known as<br />

choanal atresia), retarded growth and development,<br />

genital abnormality, and ear abnormality<br />

CHD chromodomain-helicase-DNA-binding protein<br />

ChIP chromatin immunoprecipitation<br />

EED embryonic ectoderm development protein<br />

esBAF brahma-associated factor complex associated with<br />

embryonic stem cells<br />

ESC embryonic stem cell<br />

EZH enhancer of zeste homolog protein<br />

H3K27me3 trimethylated lysine-27 on histone H3<br />

H3K4me3 trimethylated lysine-4 on histone H3<br />

HCNE highly conserved non-coding elements<br />

INO80 inositol-requiring 80<br />

iPSC induced pluripotent stem cell<br />

ISWI imitation switch<br />

LSD lysine-specific demethylase 1<br />

MLL mixed lineage leukemia protein<br />

MNase micrococcal nuclease<br />

MYC v-myc avian myelocytomatosis viral oncogene homolog<br />

NuRD nucleosome remodeling deacetylase<br />

OCT4 POU domain, class 5, transcription factor 1 protein<br />

PcG polycomb group<br />

Phc polyhomeotic-like 1 (Drosophila) protein<br />

PRC polycomb repressive complex<br />

PTMs posttranslational modifications<br />

RBBP5 retinoblastoma-binding protein 5<br />

RING1B ring finger protein 2<br />

rRNA ribosomal RNA<br />

SET SET domain containing protein<br />

SMARCD1 SWI/SNF-related, matrix-associated, actin-dependent<br />

regulator of chromatin, subfamily d, member 1 protein<br />

SUZ12 suppressor of zeste 12 homolog (Drosophila) protein<br />

SWI/SNF SWItch/Sucrose non-fermentable<br />

Tip60 K(lysine) acetyltransferase 5<br />

TrxG trithorax group<br />

TSS transcription start site<br />

WDR5 WD repeat domain 5 protein<br />

Bivalent modifications<br />

The first evidence for the existence of bivalent modifications came<br />

from studies in pluripotent mouse ESCs [5,6]. Using sequential<br />

chromatin immunoprecipitation (ChIP) and tiling arrays, highly<br />

conserved non-coding elements (HCNEs) were found to be<br />

enriched with bivalent histone modifications H3K4me3 and<br />

H3K27me3, marking lowly expressed developmental regulators<br />

[6]. Shortly thereafter, early replicating genes were similarly<br />

shown to possess bivalent domains [5]. Depletion of the EED<br />

subunit of the Polycomb repressive complex 2 (PRC2) led to an<br />

almost complete loss of H3K27me3 resulting in an upregulation of<br />

the bivalent genes analyzed. Despite the presence of the activating<br />

histone marks, the expression of the bivalent genes in both studies<br />

varied from very low to no expression, suggesting that these<br />

genes are poised for immediate activation. Supporting this notion,<br />

upon differentiation, some of these bivalent modifications were<br />

resolved, either losing the H3K27me3 mark permitting their<br />

expression, or losing the H3K4me3, rendering them stably silent<br />

(Fig 1).<br />

Bivalent histone modifications were also identified in human<br />

ESCs [11,12] and induced pluripotent stem cells (iPSCs) [13–15],<br />

marking developmentally regulated genes, similar to the situation<br />

found in mESCs. Other stem cell types, such as hematopoietic stem<br />

cells, were also shown to possess a similar bivalent chromatin architecture,<br />

containing thousands of bivalently marked, developmentally<br />

regulated promoters [16]. However, although many of these<br />

bivalent domains are resolved during differentiation, a subset of<br />

promoters retains its bivalent state following even terminal differentiation<br />

[17]. Therefore, bivalency may not merely reflect a transient,<br />

flexible chromatin state during differentiation, but rather a condition<br />

present in most or even all cell types. Supporting this idea, bivalent<br />

domains were found in a number of different non-stem-cell lines<br />

[14,18], including differentiated human T cells, where weakly<br />

expressed genes were found to possess additional acetylation marks<br />

on H3K9 and H3K14 along with H3K4me3 and H3K27me3 in their<br />

promoters [19,20]. Bivalent domains were reported in several<br />

cancer cells as well [21–24], where they were suggested to promote<br />

their plasticity and responsiveness, potentially serving as a novel<br />

unexplored therapeutic avenue [22]. These studies suggest that<br />

bivalent modifications are present in both pluripotent and nonpluripotent<br />

cells, where they maintain genes largely in a repressed<br />

state, but at the same time, keep them poised for activation until a<br />

proper signal is perceived.<br />

Despite the rapidly expanding literature reporting different<br />

aspects of bivalent chromatin, whether bivalent domains serve an<br />

actual function has recently been questioned [25,26]. In addition,<br />

the existence of bivalent domains on the same nucleosome has not<br />

been unequivocally demonstrated. Apart from the many ChIP experiments,<br />

bivalent chromatin was shown using micrococcal nuclease<br />

(MNase) digestion of chromatin followed by liquid chromatography<br />

and mass spectrometry [27], suggesting an asymmetric configuration<br />

of chromatin on opposite H3 tails. However, all evidence to<br />

date relies on population studies, and therefore, the seeming presence<br />

of two opposing marks on the same nucleosome may be the<br />

outcome of cellular heterogeneity. Single nucleosome resolution, the<br />

smoking gun of bivalent chromatin, has yet to be reported.<br />

Establishment and maintenance of bivalent modifications<br />

in ESCs<br />

Trithorax group (TrxG) proteins and Polycomb group (PcG) proteins<br />

assemble into multimeric protein complexes and are largely,<br />

although not exclusively, responsible for the deposition of H3K4me3<br />

and H3K27me3 marks, respectively (Fig 2). The exact mechanism<br />

behind the recruitment of these protein complexes to specific sites is<br />

not entirely clear; however, initial studies showed that bivalent<br />

domains are predominantly associated with CpG islands in ESCs<br />

[6]. In addition to DNA methylation, multiple studies have demonstrated<br />

that selected histone modifications, several transcription<br />

factors (TFs), and some non-coding RNAs (ncRNAs) also play a role<br />

in this process [7].<br />

1610<br />

EMBO reports Vol 16 | No 12 | 2015 ª 2015 The Authors


Arigela Harikumar & Eran Meshorer Chromatin remodeling, bivalent histone marks, and ESCs EMBO reports<br />

BIVALENT GENE<br />

SILENCED GENE<br />

H3K27me3<br />

H3K27me3<br />

ACTIVATED GENE<br />

Figure 1. The bivalency concept.<br />

A bivalent gene, depicted as a boat (top left), is ready to go (sail up: H3K4me3) but is held in check (anchor: H3K27me3). Once the sail is down (top right), the gene is stably<br />

silenced (only H3K27me3), but if instead the anchor is lifted (bottom), the gene is promptly activated (only H3K4me3). [Correction added on 2 December 2015 after first online<br />

publication: “H4K4me3” has been corrected to “H3K4me3”.]<br />

H3K4me3<br />

In mammalian systems, SETD1A, SETD1B, and MLL complexes,<br />

among others, which share several subunits including WDR5,<br />

RBBP5, ASH2L, and DPY-30, catalyze the deposition of the<br />

H3K4me3 mark [28] (Fig 2). SETD1A/B complexes seem to be<br />

responsible for global H3K4me3 deposition, whereas MLL1–MLL4<br />

complexes likely serve more specific functions. Among those,<br />

MLL2, but not MLL1 or SETD1, was shown to act as the main<br />

methyltransferase at bivalent promoters [25,26]. MLL1 and MLL2<br />

contain DNA-binding domains termed CXXC or zinc finger CXXC<br />

(ZF-CXXC) motifs, which specifically recognize unmethylated CpG<br />

islands [29,30]. These motifs act to recruit MLL complexes to chromatin<br />

templates by promoting target site recognition. Similarly,<br />

SETD1A/B complexes contain a CXXC finger protein 1 (CFP1)<br />

subunit, which includes a DNA-binding domain selectively recognizing<br />

unmethylated CpGs [31]. Loss of CFP1 most strongly affects<br />

H3K4 methylation at promoters of highly expressed genes in ESCs,<br />

but not at bivalent gene promoters [32]. Other SETD1A/B and MLL<br />

components were also shown to be important in ESCs or early ESC<br />

differentiation. Knockdown of WDR5 or ASH2L, for instance, results<br />

in aberrant expression programs and defective self-renewal and<br />

pluripotency [33–36], and knockdown of RBBP5 or DPY-30 has little<br />

effect on self-renewal but leads to improper ESC neuronal differentiation<br />

[37]. Interestingly, knockdown of DPY-30 alters H3K4 methylation<br />

specifically at bivalent domains in ESCs [38], suggesting a<br />

selective developmentally related function of this subunit. MLL2<br />

depletion also results in skewed differentiation, along all three germ<br />

layers [39], and Mll2-null mice die before embryonic day E11.5,<br />

showing drastically reduced expression of several Hox genes [40].<br />

Taken together, these studies highlight the important role that H3K4<br />

methylation and its maintenance plays in development, pluripotency,<br />

ESC biology, and early ESC differentiation.<br />

H3K27me3<br />

The PRC2 complex is responsible for the deposition of H3K27me3<br />

marks at bivalent promoters (Fig 2). The core PRC2 complex is<br />

composed of enhancer of zeste (EZH2 or EZH1), embryonic ectoderm<br />

development (EED), suppressor of zeste 12 (SUZ12), as well as<br />

RBBP4 (RbAp48) and RBBP7 (RbAp46) [41]. EZH2 is the catalytic<br />

subunit, acting as the methyltransferase of H3K27, which in turn is<br />

recognized by chromodomain-containing proteins such as CBX<br />

proteins, as well as by the EED subunit itself [42]. In mouse ESCs,<br />

CBX7, for instance, the primary CBX protein expressed in ESCs<br />

[43,44], was shown to interact with H3K27me3 thereby recruiting<br />

ª 2015 The Authors EMBO reports Vol 16 | No 12 | 2015<br />

1611


EMBO reports Chromatin remodeling, bivalent histone marks, and ESCs Arigela Harikumar & Eran Meshorer<br />

WDR5<br />

Hcf1<br />

Menin<br />

SETD1A/B<br />

MLL1,3,4<br />

ASH2<br />

MLL2<br />

MLL2<br />

complex<br />

DPY30<br />

RBBP4<br />

H3K4me3<br />

EED<br />

PRC2<br />

EZH2<br />

SUZ12<br />

PRC2<br />

complex<br />

RBBP4<br />

RBBP7<br />

H3K27me3<br />

modifications and pluripotency, the exact role that pluripotency<br />

factors play at bivalent domains remains largely unclear.<br />

Depletion of individual subunits EED or SUZ12 results in deregulation<br />

of lineage-specific genes, although with minimal impact on<br />

cell viability and self-renewal [56–59], suggesting that PcG<br />

complexes serve little function in ESCs or that in ESCs, alternative<br />

compensatory mechanisms exist. In contrast to the situation, when<br />

ESCs are kept in an undifferentiated state, ESCs deficient of PRC2<br />

components exhibit aberrant differentiation potential when differentiation<br />

is induced [56–59]. These situations parallel the postimplantation<br />

lethality phenotypes observed in PRC2 knockout mouse<br />

models [60–62]. Concomitantly, depletion of PRC1 components<br />

such as RING1B and BMI1 also impairs proper differentiation<br />

[63–66]. Simultaneous depletion of RING1B and EED in ESCs<br />

provokes an even stronger inclination toward differentiation,<br />

although self-renewal can still be preserved under careful culture<br />

conditions, and prolonged differentiation results in cell death [59].<br />

Taken together, these knockout models demonstrate that akin to<br />

TrxG proteins, PcG proteins—arguably through the control of<br />

bivalent target genes encoding developmental regulators—are vital<br />

for proper ESC differentiation.<br />

Figure 2. Main protein complexes catalyzing bivalent chromatin marks.<br />

Left: Protein complexes catalyzing H3K4 methylation (green flag). Right: The<br />

PRC2 complex catalyzing H3K27 methylation (red flag). Shown are only the main<br />

proteins and protein complexes catalyzing H3K4/H3K27 methylation. Less<br />

abundant subunits are not depicted. [Correction added on 2 December 2015 after<br />

first online publication: “H4K4” has been corrected to “H3K4”.]<br />

the PRC1 complex [43,45]. RING1B and BMI1 subunits of the PRC1<br />

complex in turn deposit the ubiquitination of H2AK119 [46,47]. It<br />

was further suggested that the deposition of PRC1 and the<br />

H2AK119ub mark (H2A ubiquitylated on lysine 119) are involved in<br />

RNA polymerase 2 (RNAPII) pausing in gene bodies, rendering their<br />

repression [48,49]. PRC1-related H2AK119ub1 was also shown to<br />

recruit PRC2 to chromatin, demonstrating a functional link between<br />

the two complexes [50]. Indeed, many developmentally regulated<br />

genes are marked with bivalent domains consisting of both PRC1<br />

and PRC2 complexes, but interestingly, a subset of these bivalent<br />

domains have been shown to be exclusively bound by PRC2 [51].<br />

Unlike the PRC1/PRC2 double positive bivalent domains, these<br />

PRC2-specific bivalent domains usually decorate promoters of genes<br />

which are not bona fide developmental genes (often encoding for<br />

membrane proteins or proteins of unknown functions) and are only<br />

weakly conserved. These findings suggest an additional mechanism<br />

of silencing. In ESCs, pluripotency factor binding sites often coincide<br />

with positioning of core subunits of MLL and PRC2 complexes on<br />

bivalent domains [6,34,52]. In addition, key pluripotency components,<br />

such as OCT4 and MYC, have been shown to interact with<br />

components of the MLL and PRC protein complexes [53,54],<br />

suggesting a tight co-regulation between bivalent domains and the<br />

pluripotency network. Indeed, depletion of OCT4 in ESCs results in<br />

a selective reduction of H3K4me3 levels on selected genes, providing<br />

evidence for the tight relationship between the pluripotency<br />

network and H3K4me3 levels [34], although whether reduced<br />

H3K4me3 levels are a cause or consequence of reduced transcription<br />

is still under question [55]. While these observations suggest a<br />

functional connection between the maintenance of bivalent<br />

Chromatin remodeling and bivalency<br />

Accumulating evidence suggests a tight interplay between ATPdependent<br />

chromatin remodeling proteins and bivalent histone<br />

modifications. While this connection likely plays a role in most, if<br />

not all, cell types, it is particularly pertinent for pluripotent stem<br />

cells, because of the relative abundance of chromatin remodelers in<br />

ESCs [67], and because of their special connection with bivalency,<br />

as described below.<br />

Chromatin remodeling proteins are ATP-dependent complexes<br />

that usually contain a catalytic ATPase subunit in addition to regulatory<br />

factors mediating protein–protein and chromatin–protein interactions.<br />

Chromatin remodelers act to alter chromatin structure by<br />

several different mechanisms, including incorporation or ejection of<br />

histone octamers, sliding nucleosomes along chromatin templates,<br />

and, by histone exchange, altering nucleosome composition [68]<br />

(Fig 3). Chromatin remodeling proteins are generally divided into<br />

four major families: SWI/SNF (switch/sucrose non-fermentable),<br />

CHD (chromodomain-helicase DNA-binding), ISWI (imitation<br />

switch), and INO80 (inositol-requiring 80), each involving different<br />

protein complexes and often different and even opposing actions. A<br />

growing number of chromatin remodeling proteins have been linked<br />

to ESC function and pluripotency and were shown to play essential<br />

roles in stemness and/or early differentiation and development [69].<br />

Interestingly, at least one member of each of these remodeler families<br />

was shown to be essential for early mouse development, before<br />

or during implantation [69], demonstrating the importance of chromatin<br />

remodelers in pluripotency and early fate decisions.<br />

One of the first clues to the involvement of chromatin remodeling<br />

proteins in regulating either H3K4 or H3K27 methylation in developmental<br />

genes came from an RNAi screen of chromatin-related<br />

proteins in mouse ESCs [70]. Among the dozens of proteins that<br />

were identified to have a potential role in maintaining the undifferentiated<br />

state in ESCs, the authors specifically identified seven<br />

subunits of the Tip60–p400 complex of the INO80 family. Using<br />

1612<br />

EMBO reports Vol 16 | No 12 | 2015 ª 2015 The Authors


Arigela Harikumar & Eran Meshorer Chromatin remodeling, bivalent histone marks, and ESCs EMBO reports<br />

INSERTION<br />

EVICTION<br />

DIMER EXCHANGE<br />

SLIDING<br />

Figure 3. The classical actions of chromatin remodeling proteins.<br />

Shown are different functional outcomes mediated by ATP-dependent chromatin remodeling proteins, including nucleosome insertion (top left), nucleosome eviction (top<br />

right), dimer exchange (bottom left), and nucleosome sliding (bottom right). The chromatin remodeling proteins themselves are not depicted.<br />

biochemical and functional assays, the authors found that Tip60–<br />

p400 significantly co-localizes with H3K4me3, especially around the<br />

transcription start site (TSS), and that it mostly acts to repress gene<br />

expression in ESCs. Knockdown of Tip–p400 resulted in 4% deregulated<br />

genes, most of which were upregulated. Interestingly, many of<br />

the upregulated genes were found to be classical bivalent early differentiation<br />

genes, which are normally silent in ESCs, reminiscent of<br />

depletion of PcG components in ESCs [70].<br />

Additionally, PcG proteins were also shown to be functionally<br />

linked to chromatin remodeling proteins in ESCs. This relationship<br />

was revealed following knockdown (KD) studies of the core component<br />

of the SWI/SNF esBAF complex, BRG1, in mouse ESCs [71].<br />

Expression analysis following BRG1 KD revealed increased transcription<br />

of several PcG subunits of both PRC1 and PRC2 complexes<br />

including Bmi1, Cbx7, Ring1, Phc1, and Phc2, promoters of which<br />

were directly bound by BRG1, as revealed by ChIP-seq experiments<br />

[71,72]. Moreover, a PRC2 component required for ESC differentiation,<br />

Jarid2, interacts with esBAF [73], counteracting PRC2’s<br />

methyltransferase activity [74,75]. These results suggest a direct<br />

association of chromatin remodeling proteins both with promoters<br />

of PcG-related genes and with the PRC2 protein complex itself.<br />

However, no co-localization of PRC2 components with BRG1 was<br />

found at chromatin on a genomewide scale [71], hinting that their<br />

association is not required for chromatin binding. When BRG1 was<br />

knocked out in ESCs, global H3K27me3 levels were not altered, but<br />

H3K27me3 displayed a selective elevation in BRG1-activated genes<br />

and a decrease in BRG1-repressed genes [76]. This indicates that<br />

BRG1 directly regulates the level and distribution of H3K27me3 at<br />

its target genes.<br />

The strong connection between SWI/SNF chromatin remodeling<br />

proteins and H3K27me3 in ESCs was recently further supported by<br />

our own studies [77]. Screening for proteins that are differentially<br />

associated with chromatin between undifferentiated and differentiated<br />

ESCs, we identified the chromatin remodeling protein<br />

SMARCD1 (BAF60a), an additional component of the esBAF<br />

complex. ChIP-seq maps of SMARCD1 in ESCs revealed a distribution<br />

not dissimilar from that of H3K27me3 around transcription start<br />

sites (TSS) and a significant enrichment in promoters of bivalent<br />

genes. Analyzing genomewide maps of H3K27me3 and H3K4me3<br />

before and after SMARCD1 depletion revealed significant redistribution<br />

of these marks. In undifferentiated ESCs, both H3K4me3 and<br />

H3K27me3 were slightly elevated around TSSs upon SMARCD1 KD.<br />

In contrast, in differentiating ESCs, H3K4me3 was still elevated<br />

around TSSs, but H3K27me3 was dramatically reduced, by more<br />

than 75% around TSSs. Interestingly, bivalent genes were relatively<br />

protected from this wave of H3K27me3 elimination, indicating selective<br />

regulation of bivalent genes, by an unknown mechanism. Once<br />

again, no apparent changes in global levels were observed, as seen<br />

in the BRG1-KO ESCs.<br />

Regulation of bivalent histone modifications by the esBAF<br />

complex was further established by an inducible knockout system<br />

of the esBAF component BAF250a in mouse ESCs [78]. Mapping<br />

of nucleosomes, bivalent histone modifications, and the PcG<br />

component SUZ12 before and after BAF250a depletion demonstrated<br />

that BAF250a mediates nucleosome occupancy and<br />

H3K27me3 levels at the upregulated, but not the downregulated<br />

genes in ESCs, and that it exerts its function likely by regulating<br />

esBAF and PRC2. BAF250a KO led to an increase in nucleosome<br />

positioning and a decrease in H3K27me3 levels, especially in bivalent<br />

and developmental gene promoters in ESCs. These alterations,<br />

accompanied by aberrant expression of developmental and<br />

pluripotency genes, resulted in differentiation defects in the<br />

BAF250a KO ESCs [78]. This study supports a synergic role for<br />

esBAF and PRC2 in ESCs.<br />

ª 2015 The Authors EMBO reports Vol 16 | No 12 | 2015<br />

1613


EMBO reports Chromatin remodeling, bivalent histone marks, and ESCs Arigela Harikumar & Eran Meshorer<br />

As indicated above, the catalytic subunit of the esBAF complex is<br />

BRG1. In human ESCs, BRG1 depletion resulted in elevated levels of<br />

the enhancer-associated histone modification H3K27ac at BRG1<br />

target gene enhancers [79]. This suggests that in addition to its function<br />

in mediating H3K27 methylation levels at promoters, BRG1<br />

may also act as a selective repressor at enhancer regions. Support<br />

for a dual role for BRG1 in regulating H3K27ac enhancer regions on<br />

one hand, and in maintaining PcG-mediated promoter repression on<br />

the other, came from a recent study of in vitro mesodermal differentiation<br />

of mouse ESCs [80]. BRG1 co-localization with H3K27ac at<br />

distal enhancers is required to maintain their H3K27 acetylation<br />

during mesoderm induction, while it is also required to maintain<br />

PcG-mediated repression of non-mesodermal developmental regulators<br />

during differentiation. Taken together, these studies demonstrate<br />

a potential role for esBAF chromatin remodeling proteins in<br />

regulating bivalent histone modifications, primarily H3K27 methylation,<br />

in ESCs and during ESC differentiation, and provide evidence<br />

that the interplay between esBAF and PcG acts both to activate and<br />

to silence gene expression programs in ESCs.<br />

Another family of chromatin remodeling proteins which was implicated<br />

in regulating H3K4/H3K27 methylation is the Chromodomainhelicase<br />

DNA-binding (CHD) family of proteins. CHD1 was found<br />

to regulate open chromatin and pluripotency in mouse ESCs by<br />

its association with H3K4me3 and counteracting heterochromatinization<br />

in pluripotent cells [81]. However, it was later concluded<br />

that its association with H3K4me3 is specific to active genes and it<br />

is in fact exclusively depleted in the dual H3K4/H3K27 regions [82],<br />

suggesting a selective role as an activator, leaving the suppressed<br />

bivalent domains intact. Since CHD1 interacts with H3K4me3, it<br />

raises the possibility that a mechanism to selectively clear CHD1<br />

from bivalent promoters exists. CHD7 on the other hand was<br />

shown, using clustering analysis, to be associated with three distinct<br />

protein complexes in ESCs: an enhancer signature cluster, a c-MYC/<br />

n-MYC-enriched cluster, and a PcG cluster, containing SUZ12,<br />

RING1B, and EZH2 [83], suggesting, among other things, a function<br />

for CHD7 in PcG-mediated gene regulation. Supporting this notion,<br />

depletion of the CHD7 homolog Kismet in Drosophila resulted in<br />

global elevation of H3K27 methylation levels, demonstrating a role<br />

for CHD7/Kismet in counteracting PcG activity [84]. CHD7 was also<br />

shown to associate with the PBAF chromatin remodeling complex<br />

during embryogenesis and human ESC differentiation to promote<br />

neural crest migration and neural crest gene expression programs<br />

[85]. CHD7 mutations or other causes of failure to activate neural<br />

crest migration have been implicated in the development of<br />

CHARGE syndrome, a complex genetic disease affecting the nervous<br />

system, heart, vision, ears, and more [86]. These studies highlight a<br />

link between CHD proteins and H3K27 methylation, thereby<br />

affecting bivalency albeit indirectly.<br />

The CHD family of proteins also includes two prominent<br />

members of the NuRD chromatin remodeling complex, namely<br />

CHD3 and CHD4. The NuRD complex was shown to be essential for<br />

proper ESC differentiation [87] and was shown to deacetylate<br />

H3K27 in ESCs, enabling the subsequent recruitment of PcG<br />

proteins and trimethylation of H3K27. It therefore controls the<br />

balance between H3K27 acetylation and methylation, thereby<br />

enabling cell differentiation [88], although it likely does not act<br />

directly at bivalent promoters since it is repelled by H3K4me3 [89].<br />

Further supporting the tight connection between NuRD and PcG<br />

during development is the association between CHD4 and the<br />

H3K27 methyltransferase EZH2 required during astroglial differentiation.<br />

CHD4 was found to be essential for EZH2 association with<br />

key astroglial gene promoters, suppressing their expression in nonglial<br />

cells, and depletion of CHD4 promotes gliogenesis in vivo [90].<br />

A similar role for NuRD, mediated by CTBP2, was also seen in differentiating<br />

ESCs during the exit from pluripotency: NuRD facilitates<br />

H3K27 deacetylation followed by recruitment of PcG and H3K27<br />

methylation [91]. Finally, NuRD was also shown to play a role in<br />

regulating H3K4me3- and H3K27me3-marked bivalent rRNA genes<br />

[92]. Although the latter study was not performed in ESCs, given<br />

the importance of the NuRD complex and rRNA transcriptional<br />

regulation in pluripotency, it is fair to speculate that it likely plays a<br />

similarly important role there too. Consistent with NuRD’s role in<br />

regulating bivalency, the complex was shown to interact not only<br />

with PcG proteins, as discussed above, but also with the H3K4<br />

methyltransferase MLL1 [93]. But perhaps even more importantly, it<br />

was also shown to interact with the H3K4 demethylase LSD1, which<br />

occupies the large majority of active genes, as well as approximately<br />

two-thirds of bivalent genes (2003 out of 3,094), in ESCs. Thus,<br />

through its association with the NuRD chromatin remodeling<br />

complex, LSD1 acts to silence pluripotency genes during early differentiation<br />

[94].<br />

Taken together, these studies demonstrate that chromatin remodeling<br />

and bivalency, and most significantly PcG-mediated H3K27<br />

methylation, are oftentimes functionally linked in ESCs and during<br />

differentiation, acting to resolve bivalent domains into stably activated<br />

or stably repressed states (Table 1; Fig 4). Because this<br />

connection between chromatin remodeling and bivalency is only<br />

beginning to emerge, the examples provided above are sometimes<br />

sketchy or indirect, with only H3K4 or H3K27 being affected.<br />

However, as more convincing data are gradually accumulating, it is<br />

becoming increasingly clear that remodelers, among other chromatin<br />

factors, act to shape and regulate bivalent chromatin<br />

Table 1. Chromatin remodelers involved in regulating bivalency.<br />

Remodeler Complex H3K4me3 H3K27me3 Bivalent References<br />

Tip60–p400 INO80 U [70]<br />

BRG1 esBAF U U [71]<br />

SMARCD1 esBAF U U U [77]<br />

BAF250a esBAF U U [78]<br />

CHD7 – U [83]<br />

CHD3/4 NuRD U U U [88]<br />

1614<br />

EMBO reports Vol 16 | No 12 | 2015 ª 2015 The Authors


Arigela Harikumar & Eran Meshorer Chromatin remodeling, bivalent histone marks, and ESCs EMBO reports<br />

MLL2<br />

NuRD<br />

PRC2<br />

Sidebar A:<br />

In need of answers<br />

?<br />

esBAF<br />

H3K4me3<br />

H3K27me3<br />

INO80<br />

(i) Is bivalency a single-cell phenomenon, or a result of population<br />

heterogeneity?<br />

(ii) If bivalency is within a single cell, are bivalent histone marks<br />

present within the same allele?<br />

(iii) If so, do bivalent histone marks appear within the same nucleosome?<br />

(iv) If they reside within same nucleosome, are the bivalent marks<br />

distributed symmetrically or asymmetrically within the histone<br />

pair of the nucleosome?<br />

To answer these questions unequivocally, bivalent histone marks must<br />

be monitored at a single-nucleosome resolution. Single nucleosomes<br />

can be reconstituted in vitro or digested using micrococcal nuclease<br />

(MNase) and immobilized for visualization.<br />

Figure 4. Chromatin remodeling complexes regulating bivalent<br />

nucleosomes.<br />

A single schematic bivalent nucleosome is shown (orange) marked with both<br />

H3K4me3 (green flag, left) and H3K27me3 (red flag, right). Chromatin remodeling<br />

complexes which were shown to regulate either or both marks are shown in<br />

green (esBAF), blue (NuRD), and mustard (INO80). Dotted arrows represent<br />

suggested regulation; dotted double lines represent potential interaction.<br />

domains. For example, non-canonical, replication-independent<br />

histone variants including H2A.Z and H3.3 have been recently<br />

reported to be associated with bivalent chromatin [95–98]. In ESCs,<br />

H2A.Z is found in active and bivalent promoters, both of which are<br />

enriched with H3K4me3, but absent from repressed H3K4me3-<br />

negative promoters [96]. H2A.Z depletion caused derepression of<br />

poised genes with concomitant loss of PcG components from<br />

bivalent promoters [95], suggesting a potential co-regulation of<br />

H2A.Z and PcG. However, a direct interaction between PcG and<br />

H2A.Z has not been reported so far. Along the same lines, H3.3 was<br />

shown to be required for the deposition of H3K27 methylation at<br />

bivalent promoters in a PRC2-dependent manner [98]. When H3.3<br />

or its chaperone, HIRA, is depleted, H3K27me3 mark is reduced at<br />

bivalent promoters, along with reduced PRC2 occupancy and<br />

reduced nucleosome turnover [98], albeit with minimal transcriptional<br />

changes. This implies that other compensatory mechanisms<br />

regulating developmental genes in ESCs are in place, but together<br />

portrays a picture which suggests that histone variants and their<br />

remodelers might also be a part of the bivalency apparatus.<br />

Concluding remarks<br />

Almost a decade has passed since the original discovery of bivalent<br />

nucleosomes in ESCs [5,6]. While it was tempting to speculate that<br />

bivalent promoters are restricted to developmentally regulated<br />

genes, enabling a quick transition to an active or a stable silent<br />

state, it is clear today that bivalency is more complicated, extending<br />

to different gene families in multiple different cell types. Furthermore,<br />

it is increasingly recognized that regulation of the bivalent<br />

state is highly complex involving a variety of different proteins and<br />

regulators. Here, we highlighted the family of ATP-dependent chromatin<br />

remodeling proteins, which is emerging as an important<br />

player in regulating bivalent domains, especially in the context of<br />

pluripotent stem cells. While TrxG and PcG proteins provide the<br />

mechanisms of action for H3K4/H3K27 methylation, we speculate<br />

(i) Does bivalency play a functional role in ESCs or in any other cell<br />

type?<br />

(ii) What is the role of bivalent promoters in terminally differentiated<br />

cells?<br />

Complete selective depletion of bivalent domains may be difficult to<br />

achieve, although a recent study [26] demonstrated little effect on<br />

early differentiation following depletion of MLL2 which conferred<br />

selective depletion of H3K4me3 on bivalent promoters. Additional<br />

studies of selective depletion of bivalent marks through the use of<br />

genetics or CRISPR/Cas9 approaches will determine the role, if any,<br />

that bivalent domains play in pluripotency and embryonic development.<br />

(i) Which of the histone variants are associated with bivalent domains?<br />

Pull-down of variant modified nucleosomes followed by MS<br />

approaches, or single nucleosome assays once single-nucleosome resolution<br />

is achieved, will determine the exact composition/modifications<br />

of histone variant-containing nucleosomes.<br />

(i) How do chromatin remodeling proteins regulate bivalent histone<br />

marks?<br />

(ii) Do pluripotency factors, in concert with chromatin remodeling<br />

complexes, regulate bivalent domains?<br />

Interaction analyses, mutational studies, rescue experiments, and<br />

functional assays will help achieve mechanistic insights into the regulation<br />

of bivalent domains by chromatin remodeling proteins and/or<br />

pluripotency factors. Understanding the mechanism by which chromatin<br />

remodelers regulate the level and distribution of bivalent chromatin<br />

domains will be key to establish a direct functional connection<br />

between chromatin remodeling proteins and bivalency.<br />

that chromatin remodeling proteins may provide the required bivalent<br />

specificity. It is important to note that bivalency is not restricted<br />

to the H3K4me3/H3K27me3 pair; several, albeit more haphazard,<br />

examples were documented. For example, trophoblast and extraembryonic<br />

endodermal stem cells were shown to contain a large fraction<br />

of H3K4me3/H3K9me3 bivalent modifications but little<br />

H3K4me3/H3K27me3 bivalency [99], and a unique H3K4me1/<br />

H3K27ac/H3K9me3 trivalent signature was observed during the<br />

transition from fibroblasts to induced neurons [100]. In both these<br />

cases, the “non-canonical” bivalent/trivalent signature enables a<br />

quick transition from a closed to an open chromatin state akin to<br />

the situation proposed for the “canonical” K4/K27 marks. The next<br />

challenge would be to identify meaningful patterns and combinations<br />

of histone modifications, decipher their potential roles, and<br />

understand whether such chromatin signatures are the cause or the<br />

consequence of their suggested function. In addition, programmable<br />

ª 2015 The Authors EMBO reports Vol 16 | No 12 | 2015<br />

1615


EMBO reports Chromatin remodeling, bivalent histone marks, and ESCs Arigela Harikumar & Eran Meshorer<br />

nucleases (ZFN, TALEN, and especially CRISPR/Cas9, and mutants<br />

thereof) are already emerging as powerful tools that enable the<br />

catalysis or removal of specific modifications at bivalent loci,<br />

making it possible to study downstream effects on the corresponding<br />

genes [101–104]. More specialized systems such as photoactivatable<br />

CRISPR switches could take this idea further even to the single<br />

cell level [105,106]. The combination of epigenetic reprogramming<br />

assays, single cell technologies, and multilevel epigenomic landscape<br />

analyses will help decipher the specific roles that multivalent<br />

domains and their connection with chromatin remodeling play in<br />

pluripotency, ESCs, and development.<br />

Acknowledgements<br />

We thank Bradley Bernstein, Cem Sievers, and Sharon Schlesinger for critical<br />

comments, and Alva Biran for help with statistics. AH was supported by the<br />

Nucleosome4D network. We thank the Israel Science Foundation (ISF 1252/12<br />

and 657/12 to E.M.) and the European Research Council (ERC-281781 to E.M.)<br />

for financial support.<br />

Conflict of interest<br />

The authors declare that they have no conflict of interest.<br />

References<br />

1. Kouzarides T (2007) Chromatin modifications and their function. Cell<br />

128: 693 – 705<br />

2. Bannister AJ, Kouzarides T (2011) Regulation of chromatin by histone<br />

modifications. Cell Res 21: 381 – 395<br />

3. Tessarz P, Kouzarides T (2014) Histone core modifications regulating<br />

nucleosome structure and dynamics. Nat Rev Mol Cell Biol 15:<br />

703 – 708<br />

4. Azzaz AM, Vitalini MW, Thomas AS, Price JP, Blacketer MJ, Cryderman<br />

DE, Zirbel LN, Woodcock CL, Elcock AH, Wallrath LL et al<br />

(2014) Human heterochromatin protein 1alpha promotes nucleosome<br />

associations that drive chromatin condensation. J Biol Chem 289:<br />

6850 – 6861<br />

5. Azuara V, Perry P, Sauer S, Spivakov M, Jorgensen HF, John RM, Gouti M,<br />

Casanova M, Warnes G, Merkenschlager M et al (2006) Chromatin<br />

signatures of pluripotent cell lines. Nat Cell Biol 8: 532 – 538<br />

6. Bernstein BE, Mikkelsen TS, Xie X, Kamal M, Huebert DJ, Cuff J, Fry B,<br />

Meissner A, Wernig M, Plath K et al (2006) A bivalent chromatin structure<br />

marks key developmental genes in embryonic stem cells. Cell 125:<br />

315 – 326<br />

7. Voigt P, Tee WW, Reinberg D (2013) A double take on bivalent promoters.<br />

Genes Dev 27: 1318 – 1338<br />

8. Vastenhouw NL, Schier AF (2012) Bivalent histone modifications in<br />

early embryogenesis. Curr Opin Cell Biol 24: 374 – 386<br />

9. Dillon N (2012) Factor mediated gene priming in pluripotent stem cells<br />

sets the stage for lineage specification. BioEssays 34: 194 – 204<br />

10. Fisher CL, Fisher AG (2011) Chromatin states in pluripotent, differentiated,<br />

and reprogrammed cells. Curr Opin Genet Dev 21: 140 – 146<br />

11. Pan G, Tian S, Nie J, Yang C, Ruotti V, Wei H, Jonsdottir GA, Stewart R,<br />

Thomson JA (2007) Whole-genome analysis of histone H3 lysine 4 and<br />

lysine 27 methylation in human embryonic stem cells. Cell <strong>Stem</strong> Cell 1:<br />

299 – 312<br />

12. Zhao XD, Han X, Chew JL, Liu J, Chiu KP, Choo A, Orlov YL, Sung WK,<br />

Shahab A, Kuznetsov VA et al (2007) Whole-genome mapping of<br />

histone H3 Lys4 and 27 trimethylations reveals distinct genomic<br />

compartments in human embryonic stem cells. Cell <strong>Stem</strong> Cell 1:<br />

286 – 298<br />

13. Maherali N, Sridharan R, Xie W, Utikal J, Eminli S, Arnold K, Stadtfeld<br />

M, Yachechko R, Tchieu J, Jaenisch R et al (2007) Directly reprogrammed<br />

fibroblasts show global epigenetic remodeling and widespread<br />

tissue contribution. Cell <strong>Stem</strong> Cell 1: 55 – 70<br />

14. Mikkelsen TS, Ku M, Jaffe DB, Issac B, Lieberman E, Giannoukos G,<br />

Alvarez P, Brockman W, Kim TK, Koche RP et al (2007) Genome-wide<br />

maps of chromatin state in pluripotent and lineage-committed cells.<br />

Nature 448: 553 – 560<br />

15. Guenther MG, Frampton GM, Soldner F, Hockemeyer D, Mitalipova M,<br />

Jaenisch R, Young RA (2010) Chromatin structure and gene expression<br />

programs of human embryonic and induced pluripotent stem cells. Cell<br />

<strong>Stem</strong> Cell 7: 249 – 257<br />

16. Cui K, Zang C, Roh TY, Schones DE, Childs RW, Peng W, Zhao K (2009)<br />

Chromatin signatures in multipotent human hematopoietic stem cells<br />

indicate the fate of bivalent genes during differentiation. Cell <strong>Stem</strong> Cell<br />

4: 80 – 93<br />

17. Abraham BJ, Cui K, Tang Q, Zhao K (2013) Dynamic regulation of epigenomic<br />

landscapes during hematopoiesis. BMC Genom 14: 193<br />

18. Mohn F, Weber M, Rebhan M, Roloff TC, Richter J, Stadler MB, Bibel M,<br />

Schubeler D (2008) Lineage-specific polycomb targets and de novo DNA<br />

methylation define restriction and potential of neuronal progenitors.<br />

Mol Cell 30: 755 – 766<br />

19. Roh TY, Cuddapah S, Cui K, Zhao K (2006) The genomic landscape of<br />

histone modifications in human T cells. Proc Natl Acad Sci USA 103:<br />

15782 – 15787<br />

20. Barski A, Cuddapah S, Cui K, Roh TY, Schones DE, Wang Z, Wei G,<br />

Chepelev I, Zhao K (2007) High-resolution profiling of histone methylations<br />

in the human genome. Cell 129: 823 – 837<br />

21. Rodriguez J, Munoz M, Vives L, Frangou CG, Groudine M, Peinado MA<br />

(2008) Bivalent domains enforce transcriptional memory of DNA<br />

methylated genes in cancer cells. Proc Natl Acad Sci USA 105:<br />

19809 – 19814<br />

22. Bapat SA, Jin V, Berry N, Balch C, Sharma N, Kurrey N, Zhang S, Fang F,<br />

Lan X, Li M et al (2010) Multivalent epigenetic marks confer microenvironment-responsive<br />

epigenetic plasticity to ovarian cancer cells. Epigenetics<br />

5: 716 – 729<br />

23. McGarvey KM, Van Neste L, Cope L, Ohm JE, Herman JG, Van Criekinge<br />

W, Schuebel KE, Baylin SB (2008) Defining a chromatin pattern that<br />

characterizes DNA-hypermethylated genes in colon cancer cells. Cancer<br />

Res 68: 5753 – 5759<br />

24. Lin B, Lee H, Yoon JG, Madan A, Wayner E, Tonning S, Hothi P, Schroeder<br />

B, Ulasov I, Foltz G et al (2015) Global analysis of H3K4me3 and<br />

H3K27me3 profiles in glioblastoma stem cells and identification of<br />

SLC17A7 as a bivalent tumor suppressor gene. Oncotarget 6: 5369 – 5381<br />

25. Denissov S, Hofemeister H, Marks H, Kranz A, Ciotta G, Singh S,<br />

Anastassiadis K, Stunnenberg HG, Stewart AF (2014) Mll2 is required<br />

for H3K4 trimethylation on bivalent promoters in embryonic stem cells,<br />

whereas Mll1 is redundant. Development 141: 526 – 537<br />

26. Hu D, Garruss AS, Gao X, Morgan MA, Cook M, Smith ER, Shilatifard A<br />

(2013) The Mll2 branch of the COMPASS family regulates bivalent<br />

promoters in mouse embryonic stem cells. Nat Struct Mol Biol 20:<br />

1093 – 1097<br />

27. Voigt P, LeRoy G, Drury WJ III, Zee BM, Son J, Beck DB, Young NL,<br />

Garcia BA, Reinberg D (2012) Asymmetrically modified nucleosomes.<br />

Cell 151: 181 – 193<br />

1616<br />

EMBO reports Vol 16 | No 12 | 2015 ª 2015 The Authors


Arigela Harikumar & Eran Meshorer Chromatin remodeling, bivalent histone marks, and ESCs EMBO reports<br />

28. Shilatifard A (2012) The COMPASS family of histone H3K4 methylases:<br />

mechanisms of regulation in development and disease pathogenesis.<br />

Annu Rev Biochem 81: 65 – 95<br />

29. Birke M, Schreiner S, Garcia-Cuellar MP, Mahr K, Titgemeyer F,<br />

Slany RK (2002) The MT domain of the proto-oncoprotein MLL binds to<br />

CpG-containing DNA and discriminates against methylation. Nucleic<br />

Acids Res 30: 958 – 965<br />

30. Bach C, Mueller D, Buhl S, Garcia-Cuellar MP, Slany RK (2009) Alterations<br />

of the CxxC domain preclude oncogenic activation of mixedlineage<br />

leukemia 2. Oncogene 28: 815 – 823<br />

31. Lee JH, Voo KS, Skalnik DG (2001) Identification and characterization of<br />

the DNA binding domain of CpG-binding protein. J Biol Chem 276:<br />

44669 – 44676<br />

32. Clouaire T, Webb S, Skene P, Illingworth R, Kerr A, Andrews R, Lee JH,<br />

Skalnik D, Bird A (2012) Cfp1 integrates both CpG content and gene<br />

activity for accurate H3K4me3 deposition in embryonic stem cells.<br />

Genes Dev 26: 1714 – 1728<br />

33. Wysocka J, Swigut T, Milne TA, Dou Y, Zhang X, Burlingame AL, Roeder<br />

RG, Brivanlou AH, Allis CD (2005) WDR5 associates with histone H3<br />

methylated at K4 and is essential for H3K4 methylation and vertebrate<br />

development. Cell 121: 859 – 872<br />

34. Ang YS, Tsai SY, Lee DF, Monk J, Su J, Ratnakumar K, Ding J, Ge Y,<br />

Darr H, Chang B et al (2011) Wdr5 mediates self-renewal and reprogramming<br />

via the embryonic stem cell core transcriptional network.<br />

Cell 145: 183 – 197<br />

35. Stoller JZ, Huang L, Tan CC, Huang F, Zhou DD, Yang J, Gelb BD, Epstein<br />

JA (2010) Ash2l interacts with Tbx1 and is required during early<br />

embryogenesis. Exp Biol Med 235: 569 – 576<br />

36. Wan M, Liang J, Xiong Y, Shi F, Zhang Y, Lu W, He Q, Yang D, Chen R,<br />

Liu D et al (2012) The trithorax group protein Ash2l is essential for<br />

pluripotency and maintaining open chromatin in embryonic stem cells.<br />

J Biol Chem 288: 5039 – 5048<br />

37. Biondi CA, Gartside MG, Waring P, Loffler KA, Stark MS, Magnuson MA,<br />

Kay GF, Hayward NK (2004) Conditional inactivation of the MEN1 gene<br />

leads to pancreatic and pituitary tumorigenesis but does not affect<br />

normal development of these tissues. Mol Cell Biol 24: 3125 – 3131<br />

38. Jiang H, Shukla A, Wang X, Chen WY, Bernstein BE, Roeder RG (2011)<br />

Role for Dpy-30 in ES cell-fate specification by regulation of H3K4<br />

methylation within bivalent domains. Cell 144: 513 – 525<br />

39. Lubitz S, Glaser S, Schaft J, Stewart AF, Anastassiadis K (2007) Increased<br />

apoptosis and skewed differentiation in mouse embryonic stem cells<br />

lacking the histone methyltransferase Mll2. Mol Biol Cell 18: 2356 – 2366<br />

40. Glaser S, Schaft J, Lubitz S, Vintersten K, van der Hoeven F, Tufteland<br />

KR, Aasland R, Anastassiadis K, Ang SL, Stewart AF (2006) Multiple<br />

epigenetic maintenance factors implicated by the loss of Mll2 in<br />

mouse development. Development 133: 1423 – 1432<br />

41. Cao R, Wang L, Wang H, Xia L, Erdjument-Bromage H, Tempst P, Jones<br />

RS, Zhang Y (2002) Role of histone H3 lysine 27 methylation in Polycomb-group<br />

silencing. Science 298: 1039 – 1043<br />

42. Margueron R, Justin N, Ohno K, Sharpe ML, Son J, Drury WJ III, Voigt P,<br />

Martin SR, Taylor WR, De Marco V et al (2009) Role of the polycomb<br />

protein EED in the propagation of repressive histone marks. Nature<br />

461: 762 – 767<br />

43. Morey L, Pascual G, Cozzuto L, Roma G, Wutz A, Benitah SA, Di Croce L<br />

(2012) Nonoverlapping functions of the Polycomb group Cbx family of<br />

proteins in embryonic stem cells. Cell <strong>Stem</strong> Cell 10: 47 – 62<br />

44. O’Loghlen A, Munoz-Cabello AM, Gaspar-Maia A, Wu HA, Banito A,<br />

Kunowska N, Racek T, Pemberton HN, Beolchi P, Lavial F et al (2012)<br />

MicroRNA regulation of Cbx7 mediates a switch of Polycomb orthologs<br />

during ESC differentiation. Cell <strong>Stem</strong> Cell 10: 33 – 46<br />

45. Morey L, Aloia L, Cozzuto L, Benitah SA, Di Croce L (2013) RYBP and<br />

Cbx7 define specific biological functions of polycomb complexes in<br />

mouse embryonic stem cells. Cell Rep 3: 60 – 69<br />

46. Wang H, Wang L, Erdjument-Bromage H, Vidal M, Tempst P, Jones RS,<br />

Zhang Y (2004) Role of histone H2A ubiquitination in Polycomb silencing.<br />

Nature 431: 873 – 878<br />

47. Cao R, Zhang Y (2004) The functions of E(Z)/EZH2-mediated methylation<br />

of lysine 27 in histone H3. Curr Opin Genet Dev 14: 155 – 164<br />

48. Brookes E, de Santiago I, Hebenstreit D, Morris KJ, Carroll T, Xie SQ,<br />

Stock JK, Heidemann M, Eick D, Nozaki N et al (2012) Polycomb associates<br />

genome-wide with a specific RNA polymerase II variant, and regulates<br />

metabolic genes in ESCs. Cell <strong>Stem</strong> Cell 10: 157 – 170<br />

49. Min IM, Waterfall JJ, Core LJ, Munroe RJ, Schimenti J, Lis JT (2011)<br />

Regulating RNA polymerase pausing and transcription elongation in<br />

embryonic stem cells. Genes Dev 25: 742 – 754<br />

50. Cooper S, Dienstbier M, Hassan R, Schermelleh L, Sharif J, Blackledge NP,<br />

De Marco V, Elderkin S, Koseki H, Klose R et al (2014) Targeting polycomb<br />

to pericentric heterochromatin in embryonic stem cells reveals a<br />

role for H2AK119u1 in PRC2 recruitment. Cell Rep 7: 1456 – 1470<br />

51. Ku M, Koche RP, Rheinbay E, Mendenhall EM, Endoh M, Mikkelsen TS,<br />

Presser A, Nusbaum C, Xie X, Chi AS et al (2008) Genomewide analysis<br />

of PRC1 and PRC2 occupancy identifies two classes of bivalent<br />

domains. PLoS Genet 4: e1000242<br />

52. Lee TI, Jenner RG, Boyer LA, Guenther MG, Levine SS, Kumar RM,<br />

Chevalier B, Johnstone SE, Cole MF, Isono K et al (2006) Control of<br />

developmental regulators by Polycomb in human embryonic stem cells.<br />

Cell 125: 301 – 313<br />

53. Thomas LR, Wang Q, Grieb BC, Phan J, Foshage AM, Sun Q, Olejniczak ET,<br />

Clark T, Dey S, Lorey S et al (2015) Interaction with WDR5 promotes target<br />

gene recognition and tumorigenesis by MYC. Mol Cell 58: 440 – 452<br />

54. Ding J, Xu H, Faiola F, Ma’ayan A, Wang J (2011) Oct4 links multiple<br />

epigenetic pathways to the pluripotency network. Cell Res 22: 155 – 167<br />

55. Henikoff S, Shilatifard A (2011) Histone modification: cause or cog?<br />

Trends Genet 27: 389 – 396<br />

56. Pasini D, Bracken AP, Hansen JB, Capillo M, Helin K (2007) The polycomb<br />

group protein Suz12 is required for embryonic stem cell differentiation.<br />

Mol Cell Biol 27: 3769 – 3779<br />

57. Chamberlain SJ, Yee D, Magnuson T (2008) Polycomb repressive<br />

complex 2 is dispensable for maintenance of embryonic stem cell<br />

pluripotency. <strong>Stem</strong> <strong>Cells</strong> 26: 1496 – 1505<br />

58. Shen X, Liu Y, Hsu YJ, Fujiwara Y, Kim J, Mao X, Yuan GC, Orkin SH<br />

(2008) EZH1 mediates methylation on histone H3 lysine 27 and<br />

complements EZH2 in maintaining stem cell identity and executing<br />

pluripotency. Mol Cell 32: 491 – 502<br />

59. Leeb M, Pasini D, Novatchkova M, Jaritz M, Helin K, Wutz A (2010)<br />

Polycomb complexes act redundantly to repress genomic repeats and<br />

genes. Genes Dev 24: 265 – 276<br />

60. Faust C, Schumacher A, Holdener B, Magnuson T (1995) The eed mutation<br />

disrupts anterior mesoderm production in mice. Development 121:<br />

273 – 285<br />

61. O’Carroll D, Erhardt S, Pagani M, Barton SC, Surani MA, Jenuwein T<br />

(2001) The polycomb-group gene Ezh2 is required for early mouse<br />

development. Mol Cell Biol 21: 4330 – 4336<br />

62. Pasini D, Bracken AP, Jensen MR, Lazzerini Denchi E, Helin K (2004)<br />

Suz12 is essential for mouse development and for EZH2 histone<br />

methyltransferase activity. EMBO J 23: 4061 – 4071<br />

ª 2015 The Authors EMBO reports Vol 16 | No 12 | 2015<br />

1617


EMBO reports Chromatin remodeling, bivalent histone marks, and ESCs Arigela Harikumar & Eran Meshorer<br />

63. Leeb M, Wutz A (2007) Ring1B is crucial for the regulation of developmental<br />

control genes and PRC1 proteins but not X inactivation in<br />

embryonic cells. J Cell Biol 178: 219 – 229<br />

64. Alkema MJ, van der Lugt NM, Bobeldijk RC, Berns A, van Lohuizen M<br />

(1995) Transformation of axial skeleton due to overexpression of bmi-1<br />

in transgenic mice. Nature 374: 724 – 727<br />

65. van der Lugt NM, Domen J, Linders K, van Roon M, Robanus-Maandag<br />

E, te Riele H, van der Valk M, Deschamps J, Sofroniew M, van Lohuizen<br />

M et al (1994) Posterior transformation, neurological abnormalities,<br />

and severe hematopoietic defects in mice with a targeted deletion of<br />

the bmi-1 proto-oncogene. Genes Dev 8: 757 – 769<br />

66. Aloia L, Di Stefano B, Di Croce L (2013) Polycomb complexes in stem<br />

cells and embryonic development. Development 140: 2525 – 2534<br />

67. Efroni S, Duttagupta R, Cheng J, Dehghani H, Hoeppner DJ, Dash C,<br />

Bazett-Jones DP, Le Grice S, McKay RD, Buetow KH et al (2008) Global<br />

transcription in pluripotent embryonic stem cells. Cell <strong>Stem</strong> Cell 2:<br />

437 – 447<br />

68. Clapier CR, Cairns BR (2009) The biology of chromatin remodeling<br />

complexes. Annu Rev Biochem 78: 273 – 304<br />

69. Gaspar-Maia A, Alajem A, Meshorer E, Ramalho-Santos M (2010) Open<br />

chromatin in pluripotency and reprogramming. Nat Rev Mol Cell Biol<br />

12: 36 – 47<br />

70. Fazzio TG, Huff JT, Panning B (2008) An RNAi screen of chromatin<br />

proteins identifies Tip60-p400 as a regulator of embryonic stem cell<br />

identity. Cell 134: 162 – 174<br />

71. Ho L, Jothi R, Ronan JL, Cui K, Zhao K, Crabtree GR (2009) An embryonic<br />

stem cell chromatin remodeling complex, esBAF, is an essential<br />

component of the core pluripotency transcriptional network. Proc Natl<br />

Acad Sci USA 106: 5187 – 5191<br />

72. Kidder BL, Palmer S, Knott JG (2009) SWI/SNF-Brg1 regulates selfrenewal<br />

and occupies core pluripotency-related genes in embryonic<br />

stem cells. <strong>Stem</strong> <strong>Cells</strong> 27: 317 – 328<br />

73. Ho L, Ronan JL, Wu J, Staahl BT, Chen L, Kuo A, Lessard J, Nesvizhskii AI,<br />

Ranish J, Crabtree GR (2009) An embryonic stem cell chromatin remodeling<br />

complex, esBAF, is essential for embryonic stem cell self-renewal<br />

and pluripotency. Proc Natl Acad Sci USA 106: 5181 – 5186<br />

74. Shen X, Kim W, Fujiwara Y, Simon MD, Liu Y, Mysliwiec MR, Yuan GC,<br />

Lee Y, Orkin SH (2009) Jumonji modulates polycomb activity and<br />

self-renewal versus differentiation of stem cells. Cell 139: 1303 – 1314<br />

75. Landeira D, Sauer S, Poot R, Dvorkina M, Mazzarella L, Jorgensen HF,<br />

Pereira CF, Leleu M, Piccolo FM, Spivakov M et al (2010) Jarid2 is a<br />

PRC2 component in embryonic stem cells required for multi-lineage<br />

differentiation and recruitment of PRC1 and RNA Polymerase II to<br />

developmental regulators. Nat Cell Biol 12: 618 – 624<br />

76. Ho L, Miller EL, Ronan JL, Ho WQ, Jothi R, Crabtree GR (2011) esBAF facilitates<br />

pluripotency by conditioning the genome for LIF/STAT3 signalling<br />

and by regulating polycomb function. Nat Cell Biol 13: 903 – 913<br />

77. Alajem A, Biran A, Harikumar A, Sailaja BS, Aaronson Y, Livyatan I,<br />

Nissim-Rafinia M, Sommer AG, Mostoslavsky G, Gerbasi VR et al (2015)<br />

Differential association of chromatin proteins identifies BAF60a/<br />

SMARCD1 as a regulator of embryonic stem cell differentiation. Cell<br />

Rep 10: 2019 – 2031<br />

78. Lei I, West J, Yan Z, Gao X, Fang P, Dennis JH, Gnatovskiy L, Wang W,<br />

Kingston RE, Wang Z (2015) BAF250a regulates nucleosome occupancy<br />

and histone modifications in priming embryonic stem cell differentiation.<br />

J Biol Chem 290: 19343 – 19352<br />

79. Zhang X, Li B, Li W, Ma L, Zheng D, Li L, Yang W, Chu M, Chen W,<br />

Mailman RB et al (2014) Transcriptional repression by the BRG1-SWI/<br />

SNF complex affects the pluripotency of human embryonic stem cells.<br />

<strong>Stem</strong> Cell Rep 3: 460 – 474<br />

80. Alexander JM, Hota SK, He D, Thomas S, Ho L, Pennacchio LA, Bruneau<br />

BG (2015) Brg1 modulates enhancer activation in mesoderm lineage<br />

commitment. Development 142: 1418 – 1430<br />

81. Gaspar-Maia A, Alajem A, Polesso F, Sridharan R, Mason MJ,<br />

Heidersbach A, Ramalho-Santos J, McManus MT, Plath K, Meshorer E<br />

et al (2009) Chd1 regulates open chromatin and pluripotency of<br />

embryonic stem cells. Nature 460: 863 – 868<br />

82. Lin JJ, Lehmann LW, Bonora G, Sridharan R, Vashisht AA, Tran N, Plath K,<br />

Wohlschlegel JA, Carey M (2011) Mediator coordinates PIC assembly<br />

with recruitment of CHD1. Genes Dev 25: 2198 – 2209<br />

83. Schnetz MP, Handoko L, Akhtar-Zaidi B, Bartels CF, Pereira CF, Fisher AG,<br />

Adams DJ, Flicek P, Crawford GE, Laframboise T et al (2010) CHD7<br />

targets active gene enhancer elements to modulate ES cell-specific<br />

gene expression. PLoS Genet 6: e1001023<br />

84. Srinivasan S, Dorighi KM, Tamkun JW (2008) Drosophila Kismet regulates<br />

histone H3 lysine 27 methylation and early elongation by RNA<br />

polymerase II. PLoS Genet 4: e1000217<br />

85. Bajpai R, Chen DA, Rada-Iglesias A, Zhang J, Xiong Y, Helms J, Chang CP,<br />

Zhao Y, Swigut T, Wysocka J (2010) CHD7 cooperates with PBAF to<br />

control multipotent neural crest formation. Nature 463: 958 – 962<br />

86. Hsu P, Ma A, Wilson M, Williams G, Curotta J, Munns CF, Mehr S<br />

(2014) CHARGE syndrome: a review. J Paediatr Child Health 50:<br />

504 – 511<br />

87. Kaji K, Caballero IM, MacLeod R, Nichols J, Wilson VA, Hendrich B<br />

(2006) The NuRD component Mbd3 is required for pluripotency of<br />

embryonic stem cells. Nat Cell Biol 8: 285 – 292<br />

88. Reynolds N, Salmon-Divon M, Dvinge H, Hynes-Allen A, Balasooriya G,<br />

Leaford D, Behrens A, Bertone P, Hendrich B (2011) NuRD-mediated<br />

deacetylation of H3K27 facilitates recruitment of Polycomb Repressive<br />

Complex 2 to direct gene repression. EMBO J 31: 593 – 605<br />

89. Zegerman P, Canas B, Pappin D, Kouzarides T (2002) Histone H3 lysine<br />

4 methylation disrupts binding of nucleosome remodeling and<br />

deacetylase (NuRD) repressor complex. J Biol Chem 277: 11621 – 11624<br />

90. Sparmann A, Xie Y, Verhoeven E, Vermeulen M, Lancini C, Gargiulo G,<br />

Hulsman D, Mann M, Knoblich JA, van Lohuizen M (2013) The chromodomain<br />

helicase Chd4 is required for Polycomb-mediated inhibition<br />

of astroglial differentiation. EMBO J 32: 1598 – 1612<br />

91. Kim TW, Kang BH, Jang H, Kwak S, Shin J, Kim H, Lee SE, Lee SM, Lee JH,<br />

Kim JH et al (2015) Ctbp2 modulates NuRD-mediated deacetylation of<br />

H3K27 and facilitates PRC2-mediated H3K27me3 in active embryonic<br />

stem cell genes during exit from pluripotency. <strong>Stem</strong> <strong>Cells</strong> 33:<br />

2442 – 2455<br />

92. Xie W, Ling T, Zhou Y, Feng W, Zhu Q, Stunnenberg HG, Grummt I,<br />

Tao W (2012) The chromatin remodeling complex NuRD establishes<br />

the poised state of rRNA genes characterized by bivalent histone modifications<br />

and altered nucleosome positions. Proc Natl Acad Sci USA 109:<br />

8161 – 8166<br />

93. Nakamura T, Mori T, Tada S, Krajewski W, Rozovskaia T, Wassell R,<br />

Dubois G, Mazo A, Croce CM, Canaani E (2002) ALL-1 is a histone<br />

methyltransferase that assembles a supercomplex of proteins involved<br />

in transcriptional regulation. Mol Cell 10: 1119 – 1128<br />

94. Whyte WA, Bilodeau S, Orlando DA, Hoke HA, Frampton GM, Foster CT,<br />

Cowley SM, Young RA (2012) Enhancer decommissioning by LSD1<br />

during embryonic stem cell differentiation. Nature 482: 221 – 225<br />

95. Creyghton MP, Markoulaki S, Levine SS, Hanna J, Lodato MA, Sha K,<br />

Young RA, Jaenisch R, Boyer LA (2008) H2AZ is enriched at polycomb<br />

1618<br />

EMBO reports Vol 16 | No 12 | 2015 ª 2015 The Authors


Arigela Harikumar & Eran Meshorer Chromatin remodeling, bivalent histone marks, and ESCs EMBO reports<br />

complex target genes in ES cells and is necessary for lineage commitment.<br />

Cell 135: 649 – 661<br />

96. Ku M, Jaffe JD, Koche RP, Rheinbay E, Endoh M, Koseki H, Carr SA,<br />

Bernstein BE (2012) H2A.Z landscapes and dual modifications in pluripotent<br />

and multipotent stem cells underlie complex genome regulatory<br />

functions. Genome Biol 13: R85<br />

97. Hu G, Cui K, Northrup D, Liu C, Wang C, Tang Q, Ge K, Levens D,<br />

Crane-Robinson C, Zhao K (2013) H2A.Z facilitates access of active and<br />

repressive complexes to chromatin in embryonic stem cell self-renewal<br />

and differentiation. Cell <strong>Stem</strong> Cell 12: 180 – 192<br />

98. Banaszynski LA, Wen D, Dewell S, Whitcomb SJ, Lin M, Diaz N,<br />

Elsasser SJ, Chapgier A, Goldberg AD, Canaani E et al (2013) Hiradependent<br />

histone H3.3 deposition facilitates PRC2 recruitment at<br />

developmental loci in ES cells. Cell 155: 107 – 120<br />

99. Rugg-Gunn PJ, Cox BJ, Ralston A, Rossant J (2010) Distinct histone<br />

modifications in stem cell lines and tissue lineages from the early<br />

mouse embryo. Proc Natl Acad Sci USA 107: 10783 – 10790<br />

100. Wapinski OL, Vierbuchen T, Qu K, Lee QY, Chanda S, Fuentes DR,<br />

Giresi PG, Ng YH, Marro S, Neff NF et al (2013) Hierarchical mechanisms<br />

for direct reprogramming of fibroblasts to neurons. Cell 155: 621 – 635<br />

101. Kleinstiver BP, Prew MS, Tsai SQ, Topkar VV, Nguyen NT, Zheng Z,<br />

Gonzales AP, Li Z, Peterson RT, Yeh JJ et al (2015) Engineered CRISPR-<br />

Cas9 nucleases with altered PAM specificities. Nature 523: 481 – 485<br />

102. Qi LS, Larson MH, Gilbert LA, Doudna JA, Weissman JS, Arkin AP, Lim<br />

WA (2013) Repurposing CRISPR as an RNA-guided platform for<br />

sequence-specific control of gene expression. Cell 152: 1173 – 1183<br />

103. Cheng AW, Wang H, Yang H, Shi L, Katz Y, Theunissen TW, Rangarajan<br />

S, Shivalila CS, Dadon DB, Jaenisch R (2013) Multiplexed activation of<br />

endogenous genes by CRISPR-on, an RNA-guided transcriptional activator<br />

system. Cell Res 23: 1163 – 1171<br />

104. Hu J, Lei Y, Wong WK, Liu S, Lee KC, He X, You W, Zhou R, Guo JT,<br />

Chen X et al (2014) Direct activation of human and mouse Oct4 genes<br />

using engineered TALE and Cas9 transcription factors. Nucleic Acids Res<br />

42: 4375 – 4390<br />

105. Nihongaki Y, Kawano F, Nakajima T, Sato M (2015) Photoactivatable<br />

CRISPR-Cas9 for optogenetic genome editing. Nat Biotechnol 33:<br />

755 – 760<br />

106. Nihongaki Y, Yamamoto S, Kawano F, Suzuki H, Sato M (2015) CRISPR-<br />

Cas9-based photoactivatable transcription system. Chem Biol 22:<br />

169 – 174<br />

ª 2015 The Authors EMBO reports Vol 16 | No 12 | 2015<br />

1619


Scientific Report<br />

TRANSPARENT<br />

PROCESS<br />

OPEN<br />

ACCESS<br />

CRISPR/Cas9‐induced disruption of gene expression<br />

in mouse embryonic brain and single neural stem<br />

cells in vivo<br />

Nereo Kalebic 1 , † , Elena Taverna 1 , † , Stefania Tavano 1 , Fong Kuan Wong 1 , Dana Suchold 1 , Sylke Winkler 1 ,<br />

Wieland B Huttner*, 1 and Mihail Sarov*, 1<br />

1<br />

Max Planck Institute of Molecular Cell Biology and Genetics (MPI‐CBG), Dresden, Germany<br />

*<br />

Corresponding author. Tel: +49 3512101500; E‐mail: huttner@mpi-cbg.de<br />

Corresponding author. Tel: +49 3512102617; E‐mail: sarov@mpi-cbg.de<br />

†<br />

These authors contributed equally to this work<br />

DOI 10.15252/embr.201541715 | Published online 12.01.2016<br />

EMBO reports (2016) 17, 338-348<br />

We have applied the CRISPR/Cas9 system in vivo to disrupt gene expression in neural stem cells in the developing<br />

mammalian brain. Two days after in utero electroporation of a single plasmid encoding Cas9 and an appropriate guide<br />

RNA (gRNA) into the embryonic neocortex of Tis21::GFP knock‐in mice, expression of GFP, which occurs specifically<br />

in neural stem cells committed to neurogenesis, was found to be nearly completely (≈90%) abolished in the progeny<br />

of the targeted cells. Importantly, upon in utero electroporation directly of recombinant Cas9/gRNA complex,<br />

near‐maximal efficiency of disruption of GFP expression was achieved already after 24 h. Furthermore, by using<br />

microinjection of the Cas9 protein/gRNA complex into neural stem cells in organotypic slice culture, we obtained<br />

disruption of GFP expression within a single cell cycle. Finally, we used either Cas9 plasmid in utero electroporation<br />

or Cas9 protein complex microinjection to disrupt the expression of Eomes/Tbr2, a gene fundamental for neocortical<br />

neurogenesis. This resulted in a reduction in basal progenitors and an increase in neuronal differentiation. Thus, the<br />

present in vivo application of the CRISPR/Cas9 system in neural stem cells provides a rapid, efficient and enduring<br />

disruption of expression of specific genes to dissect their role in mammalian brain development.<br />

Synopsis<br />

This study shows that the CRISPR/Cas9 system can be used to disrupt gene expression<br />

in single neural stem cells and their progeny in the embryonic mammalian brain in vivo.<br />

This can be achieved by electroporation of a single plasmid or of protein complexes.<br />

••<br />

In utero electroporation of a single plasmid encoding Cas9 and a gRNA into the brain of<br />

mouse embryos leads to gene inactivation in the immediate progeny of the targeted neural<br />

stem cells.<br />

••<br />

Disruption of gene expression can also be achieved by electroporating the Cas9 protein in a<br />

complex with gRNA directly into embryonic mouse brains in vivo.<br />

••<br />

Microinjection of a complex of Cas9 protein and gRNA into single neural stem cells in<br />

organotypic slice culture results in disruption of gene expression within one cell cycle.


Article<br />

SOURCE<br />

DATA<br />

OPEN<br />

ACCESS<br />

Integrative genomics positions MKRN1 as a novel<br />

ribonucleoprotein within the embryonic stem cell<br />

gene regulatory network<br />

Paul A Cassar 1 , 2 , † , Richard L Carpenedo 3 , † , Payman Samavarchi‐Tehrani 4 , Jonathan B Olsen 2 , 4 , Chang Jun<br />

Park 5 , Wing Y Chang 3 , Zhaoyi Chen 3 , 6 , Chandarong Choey 3 , Sean Delaney 3 , 6 , Huishan Guo 7 , Hongbo Guo 8 , R<br />

Matthew Tanner 3 , 6 , Theodore J Perkins 3 , 7 , Scott A Tenenbaum 9 , Andrew Emili 2 , 4 , 8 , Jeffrey L Wrana 2 , 4 , 10 , Derrick<br />

Gibbings 7 and William L Stanford*, 1 , 2 , 3 , 5 , 6 , 7 , 11<br />

1<br />

Institute of Medical Science University of Toronto, Toronto, ON, Canada 2 Collaborative Program in Genome Biology and Bioinformatics, University of Toronto,<br />

Toronto, ON, Canada 3 Sprott Centre for <strong>Stem</strong> Cell Research, Ottawa Hospital Research Institute, Ottawa, ON, Canada 4 Department of Molecular Genetics,<br />

University of Toronto, Toronto, ON, Canada 5 Institute of Biomaterials and Biomedical Engineering, University of Toronto, Toronto, ON, Canada 6 Department of<br />

Cellular and Molecular Medicine, University of Ottawa, Ottawa, ON, Canada 7 Department of Biochemistry, Microbiology & Immunology, University of Ottawa,<br />

Ottawa, ON, Canada 8 Banting and Best Department of Medical Research, Donnelly Centre, University of Toronto, Toronto, ON, Canada 9 Colleges of Nanoscale<br />

Science & Engineering SUNY Polytechnic Institute, Albany, NY, USA 10 Center for Systems Biology, Lunenfeld‐Tanenbaum Research Institute, Mount Sinai Hospital,<br />

Toronto, ON, Canada 11 Ottawa Institute of Systems Biology, Ottawa, ON, Canada<br />

*<br />

Corresponding author. Tel: +1 613 737 8899 Ext. 75495; E‐mail: wstanford@ohri.ca<br />

†<br />

These authors contributed equally to this study<br />

DOI 10.15252/embr.201540974 | Published online 11.08.2015<br />

EMBO reports (2015) 16, 1334-1357<br />

In embryonic stem cells (ESCs), gene regulatory networks (GRNs) coordinate gene expression to maintain ESC<br />

identity; however, the complete repertoire of factors regulating the ESC state is not fully understood. Our previous<br />

temporal microarray analysis of ESC commitment identified the E3 ubiquitin ligase protein Makorin‐1 (MKRN1)<br />

as a potential novel component of the ESC GRN. Here, using multilayered systems‐level analyses, we compiled a<br />

MKRN1‐centered interactome in undifferentiated ESCs at the proteomic and ribonomic level. Proteomic analyses in<br />

undifferentiated ESCs revealed that MKRN1 associates with RNA‐binding proteins, and ensuing RIP‐chip analysis<br />

determined that MKRN1 associates with mRNAs encoding functionally related proteins including proteins that<br />

function during cellular stress. Subsequent biological validation identified MKRN1 as a novel stress granule‐resident<br />

protein, although MKRN1 is not required for stress granule formation, or survival of unstressed ESCs. Thus, our<br />

unbiased systems‐level analyses support a role for the E3 ligase MKRN1 as a ribonucleoprotein within the ESC GRN.<br />

Synopsis<br />

An unbiased integrative genomics approach identified proteins associated with the E3<br />

ubiquitin ligase MKRN1 and its associated mRNAs in ESCs, indicating that MKRN1 is a<br />

RNA‐binding protein involved in the modulation of cellular stress and apoptosis.<br />

••<br />

MKRN1 is present in ribonucleoprotein particles in ESCs along with other RNA‐binding<br />

proteins that are also found in stress granules, including PABPC1, PABPC4 and YBX1.<br />

••<br />

MKRN1 associates with mRNAs encoding proteins that function during cellular stress and<br />

apoptosis.<br />

••<br />

Loss of MKRN1 augments apoptosis in ESCs recovering from stress.<br />

••<br />

However, MKRN1 is not required for survival in unstressed ESCs.


Article<br />

TRANSPARENT<br />

PROCESS<br />

SOURCE<br />

DATA<br />

OPEN<br />

ACCESS<br />

Selective influence of Sox2 on POU transcription<br />

factor binding in embryonic and neural stem cells<br />

Tapan Kumar Mistri 1 , 2 , 3 , Arun George Devasia 2 , Lee Thean Chu 2 , Wei Ping Ng 1 , Florian Halbritter 3 , Douglas<br />

Colby 3 , Ben Martynoga 4 , Simon R Tomlinson 3 , Ian Chambers*, 3 , Paul Robson*, 2 , 5 , 6 and Thorsten Wohland*, 1 , 5 , 7<br />

1<br />

Department of Chemistry, National University of Singapore, Singapore, Singapore<br />

2<br />

Developmental Cellomics Laboratory, Genome Institute of Singapore, Singapore, Singapore<br />

3<br />

MRC Centre for Regenerative Medicine, Institute for <strong>Stem</strong> Cell Research, School of Biological Sciences, University of Edinburgh, Edinburgh, UK<br />

4<br />

Division of Molecular Neurobiology, MRC‐National Institute for Medical Research, Mill Hill, London, UK<br />

5<br />

Department of Biological Sciences, National University of Singapore, Singapore, Singapore<br />

6<br />

The Jackson Laboratory for Genomic Medicine, Farmington, CT, USA<br />

7<br />

Centre for Bioimaging Sciences, National University of Singapore, Singapore, Singapore<br />

*<br />

Corresponding author. Tel: +44 131 651 9500; E‐mail: ichambers@ed.ac.uk<br />

Corresponding author. Tel: +1 207 288 6594; E‐mail: paul.robson@jax.org<br />

Corresponding author. Tel: +65 6516 1248; Fax: +65 6776 7882; E‐mail: twohland@nus.edu.sg<br />

DOI 10.15252/embr.201540467 | Published online 11.08.2015<br />

EMBO reports (2015) 16, 1177-1191<br />

Embryonic stem cell (ESC) identity is orchestrated by co‐operativity between the transcription factors (TFs) Sox2 and<br />

the class V POU‐TF Oct4 at composite Sox/Oct motifs. Neural stem cells (NSCs) lack Oct4 but express Sox2 and class<br />

III POU‐TFs Oct6, Brn1 and Brn2. This raises the question of how Sox2 interacts with POU‐TFs to transcriptionally<br />

specify ESCs versus NSCs. Here, we show that Oct4 alone binds the Sox/Oct motif and the octamer‐containing<br />

palindromic MORE equally well. Sox2 binding selectively increases the affinity of Oct4 for the Sox/Oct motif. In<br />

contrast, Oct6 binds preferentially to MORE and is unaffected by Sox2. ChIP‐Seq in NSCs shows the MORE to be the<br />

most enriched motif for class III POU‐TFs, including MORE subtypes, and that the Sox/Oct motif is not enriched.<br />

These results suggest that in NSCs, co‐operativity between Sox2 and class III POU‐TFs may not occur and that<br />

POU‐TF‐driven transcription uses predominantly the MORE cis architecture. Thus, distinct interactions between<br />

Sox2 and POU‐TF subclasses distinguish pluripotent ESCs from multipotent NSCs, providing molecular insight into<br />

how Oct4 alone can convert NSCs to pluripotency.<br />

Synopsis<br />

Sox2 and POU‐TF subclasses have distinct interactions that distinguish pluripotent ESC<br />

from multipotent NSC. The different combinations of TFs and DNA binding motifs in ESC<br />

and NSC might explain why Oct4 alone can convert NSC to pluripotency.<br />

••<br />

Oct4 binds the Sox/Oct motif and the palindromic MORE equally well, whereas Oct6 binds<br />

preferentially to MORE.<br />

••<br />

Sox2 binding selectively increases the affinity of Oct4 for the Sox/Oct motif but shows no<br />

effect on Oct6.<br />

••<br />

The MORE is the most enriched motif for class III POU‐TFs Oct6, Brn1 and Brn2 in NSC<br />

whereas the Sox/Oct motif is the most enriched motif for both Oct4 and Sox2 in ESC.


Article<br />

TRANSPARENT<br />

PROCESS<br />

Mitochondrial metabolism in hematopoietic stem<br />

cells requires functional FOXO3<br />

Pauline Rimmelé 1 , † , Raymond Liang 1 , 2 , † , Carolina L Bigarella 1 , † , Fatih Kocabas 3 , Jingjing Xie 3 , Madhavika N<br />

Serasinghe 4 , Jerry Chipuk 4 , 5 , Hesham Sadek 3 , 6 , Cheng Cheng Zhang 6 and Saghi Ghaffari*, 1 , 2 , 5 , 7 , 8<br />

1<br />

Department of Developmental & Regenerative Biology, Icahn School of Medicine at Mount Sinai, New York, NY, USA<br />

2<br />

Developmental and <strong>Stem</strong> Cell Biology Multidisciplinary Training Area, Icahn School of Medicine at Mount Sinai, New York, NY, USA<br />

3<br />

Division of Cardiology, UT Southwestern Medical Center, Dallas, TX, USA<br />

4<br />

Department of Oncological Sciences, Icahn School of Medicine at Mount Sinai, New York, NY, USA<br />

5<br />

Tisch Cancer Institute, Icahn School of Medicine at Mount Sinai, New York, NY, USA<br />

6<br />

Department of Physiology, UT Southwestern Medical Center, Dallas, TX, USA<br />

7<br />

Division of Hematology, Oncology, Department of Medicine, Icahn School of Medicine at Mount Sinai, New York, NY, USA<br />

8<br />

Black Family <strong>Stem</strong> Cell Institute, Icahn School of Medicine at Mount Sinai, New York, NY, USA<br />

*<br />

Corresponding author. Tel: +1 212 659 8271; Fax: +1 212 803 6740; E‐mail: saghi.ghaffari@mssm.edu<br />

†<br />

These authors contributed equally to this work<br />

DOI 10.15252/embr.201439704 | Published online 24.07.2015<br />

EMBO reports (2015) 16, 1164-1176<br />

Hematopoietic stem cells (HSC) are primarily dormant but have the potential to become highly active on<br />

demand to reconstitute blood. This requires a swift metabolic switch from glycolysis to mitochondrial oxidative<br />

phosphorylation. Maintenance of low levels of reactive oxygen species (ROS), a by‐product of mitochondrial<br />

metabolism, is also necessary for sustaining HSC dormancy. Little is known about mechanisms that integrate<br />

energy metabolism with hematopoietic stem cell homeostasis. Here, we identify the transcription factor FOXO3 as a<br />

new regulator of metabolic adaptation of HSC. ROS are elevated in Foxo3 –/– HSC that are defective in their activity.<br />

We show that Foxo3 –/– HSC are impaired in mitochondrial metabolism independent of ROS levels. These defects are<br />

associated with altered expression of mitochondrial/metabolic genes in Foxo3 –/– hematopoietic stem and progenitor<br />

cells (HSPC). We further show that defects of Foxo3 –/– HSC long‐term repopulation activity are independent of ROS<br />

or mTOR signaling. Our results point to FOXO3 as a potential node that couples mitochondrial metabolism with HSC<br />

homeostasis. These findings have critical implications for mechanisms that promote malignant transformation and<br />

aging of blood stem and progenitor cells.<br />

Synopsis<br />

FOXO3 regulates oxidative stress in LT‐HSC, which are highly sensitive to increased reactive<br />

oxygen species. However, while the impaired function of Foxo3 –/– LT‐HSC is associated<br />

with defective mitochondrial metabolism, it is not mediated by oxidative stress or mTOR<br />

signaling.<br />

••<br />

Mitochondrial metabolism is impaired in Foxo3 –/– LT‐HSCs.<br />

••<br />

Defects in Foxo3 –/– LT‐HSC activity in vivo or Foxo3 –/– LT‐HSC mitochondrial function are not<br />

mediated by oxidative stress.


Scientific Report<br />

TRANSPARENT<br />

PROCESS<br />

DAZL regulates Tet1 translation in murine<br />

embryonic stem cells<br />

Maaike Welling 1 , † , Hsu‐Hsin Chen 2 , 3 , † , Javier Muñoz 4 , 511 , Michael U Musheev 6 , Lennart Kester 1 , Jan Philipp<br />

Junker 1 , Nikolai Mischerikow 4 , 512 , Mandana Arbab 1 , Ewart Kuijk 1 , Lev Silberstein 2 , 3 , Peter V Kharchenko 7 ,<br />

Mieke Geens 8 , Christof Niehrs 6 , 9 , Hilde van de Velde 8 , Alexander van Oudenaarden 1 , Albert JR Heck 4 , 5 and<br />

Niels Geijsen*, 1 , 10<br />

1<br />

Hubrecht Institute–KNAW and University Medical Center Utrecht, Utrecht, The Netherlands<br />

2<br />

Center for Regenerative Medicine, Massachusetts General Hospital, Boston, MA, USA<br />

3<br />

Harvard <strong>Stem</strong> Cell Institute, Cambridge, MA, USA<br />

4<br />

Biomolecular Mass Spectrometry and Proteomics, Bijvoet Center for Biomolecular Research and Utrecht Institute for Pharmaceutical Sciences,<br />

Utrecht University, Utrecht, The Netherlands<br />

5<br />

Netherlands Proteomics Centre, Utrecht, The Netherlands<br />

6<br />

Institute of Molecular Biology, Mainz, Germany<br />

7<br />

Center for Biomedical Informatics, Harvard Medical School, Boston, MA, USA<br />

8<br />

Research Group Reproduction and Genetics, Vrije Universiteit Brussel, Brussels, Belgium<br />

9<br />

Division of Molecular Embryology, DKFZ‐ZMBH Alliance, Heidelberg, Germany<br />

10<br />

Department of Companion Animals, Faculty of Veterinary Medicine, Utrecht University, Utrecht, The Netherlands<br />

11<br />

Proteomics Unit, Spanish National Cancer Research Centre (CNIO), Madrid, Spain<br />

12<br />

Research Institute of Molecular Pathology, Mass Spectrometry & Protein Chemistry, Vienna, Austria<br />

*<br />

Corresponding author. Tel: +31 30 2121800; E‐mail: n.geijsen@hubrecht.eu<br />

†<br />

These authors contributed equally to this manuscript<br />

DOI 10.15252/embr.201540538 | Published online 15.06.2015<br />

EMBO reports (2015) 16, 791-802<br />

Embryonic stem cell (ESC) cultures display a heterogeneous gene expression profile, ranging from a pristine naïve<br />

pluripotent state to a primed epiblast state. Addition of inhibitors of GSK3β and MEK (so‐called 2i conditions)<br />

pushes ESC cultures toward a more homogeneous naïve pluripotent state, but the molecular underpinnings of this<br />

naïve transition are not completely understood. Here, we demonstrate that DAZL, an RNA‐binding protein known to<br />

play a key role in germ‐cell development, marks a subpopulation of ESCs that is actively transitioning toward naïve<br />

pluripotency. Moreover, DAZL plays an essential role in the active reprogramming of cytosine methylation. We<br />

demonstrate that DAZL associates with mRNA of Tet1, a catalyst of 5‐hydroxylation of methyl‐cytosine, and enhances<br />

Tet1 mRNA translation. Overexpression of DAZL in heterogeneous ESC cultures results in elevated TET1 protein<br />

levels as well as increased global hydroxymethylation. Conversely, null mutation of Dazl severely stunts 2i‐mediated<br />

TET1 induction and hydroxymethylation. Our results provide insight into the regulation of the acquisition of naïve<br />

pluripotency and demonstrate that DAZL enhances TET1‐mediated cytosine hydroxymethylation in ESCs that are<br />

actively reprogramming to a pluripotent ground state.<br />

Synopsis<br />

This study reports that by enhancing Tet1 mRNA translation, the RNA‐binding protein<br />

DAZL regulates TET1‐mediated cytosine hydroxymethylation in ESCs during active<br />

reprogramming to a pluripotent ground state.<br />

••<br />

DAZL marks ESCs actively transitioning to a naïve pluripotent state.<br />

••<br />

DAZL enhances Tet1 mRNA translation.<br />

••<br />

DAZL overexpression results in increased hydroxymethylation.


Scientific Report<br />

TRANSPARENT<br />

PROCESS<br />

Distinct germline progenitor subsets defined<br />

through Tsc2–mTORC1 signaling<br />

Robin M Hobbs*, 1 , 2 , Hue M La 2 , Juho‐Antti Mäkelä 2 , Toshiyuki Kobayashi 3 , Tetsuo Noda 4 and<br />

Pier Paolo Pandolfi*, 1<br />

1<br />

Cancer Research Institute, Beth Israel Deaconess Cancer Center, Department of Medicine and Pathology, Beth Israel Deaconess Medical Center Harvard Medical<br />

School, Boston, MA, USA<br />

2<br />

Australian Regenerative Medicine Institute and Department of Anatomy and Developmental Biology Monash University, Clayton, VIC, Australia<br />

3<br />

Department of Pathology and Oncology, Juntendo University School of Medicine, Tokyo, Japan<br />

4<br />

Department of Cell Biology, JFCR Cancer Institute, Tokyo, Japan<br />

*<br />

Corresponding author. Tel: +1 617 735 2145; Fax: +1 617 735 2120; E‐mail: ppandolf@bidmc.harvard.edu<br />

Corresponding author. Tel: +61 3 9902 9611; Fax: +61 3 9902 9729; E‐mail: robin.hobbs@monash.edu<br />

DOI 10.15252/embr.201439379 | Published online 19.02.2015<br />

EMBO reports (2015) 16, 467-480<br />

Adult tissue maintenance is often dependent on resident stem cells; however, the phenotypic and functional<br />

heterogeneity existing within this self‐renewing population is poorly understood. Here, we define distinct subsets of<br />

undifferentiated spermatogonia (spermatogonial progenitor cells; SPCs) by differential response to hyperactivation<br />

of mTORC1, a key growth‐promoting pathway. We find that conditional deletion of the mTORC1 inhibitor Tsc2<br />

throughout the SPC pool using Vasa‐Cre promotes differentiation at the expense of self‐renewal and leads to<br />

germline degeneration. Surprisingly, Tsc2 ablation within a subset of SPCs using Stra8‐Cre did not compromise<br />

SPC function. SPC activity also appeared unaffected by Amh‐Cre‐mediated Tsc2 deletion within somatic cells of<br />

the niche. Importantly, we find that differentiation‐prone SPCs have elevated mTORC1 activity when compared to<br />

SPCs with high self‐renewal potential. Moreover, SPCs insensitive to Tsc2 deletion are preferentially associated with<br />

mTORC1‐active committed progenitor fractions. We therefore delineate SPC subsets based on differential mTORC1<br />

activity and correlated sensitivity to Tsc2 deletion. We propose that mTORC1 is a key regulator of SPC fate and<br />

defines phenotypically distinct SPC subpopulations with varying propensities for self‐renewal and differentiation.<br />

Synopsis<br />

This study defines a cell‐autonomous role for mTORC1 signaling in balanced self‐renewal<br />

and differentiation of male germline progenitors (SPCs). Within the SPC pool, mTORC1<br />

activation is preferentially observed within differentiation‐prone fractions and promotes<br />

differentiation at the expense of self‐renewal.<br />

••<br />

SPCs exhibit heterogeneous activation of mTORC1 in vivo; cells with high self‐renewal<br />

potential have low pathway activity, while differentiation‐prone subsets have higher activity.<br />

••<br />

mTORC1 activation driven by conditional Tsc2 deletion triggers differentiation commitment<br />

and depletes the progenitor pool.<br />

••<br />

mTORC1‐active committed progenitors display increased activity of the germ cell‐specific<br />

Stra8‐Cre transgene and are insensitive to Tsc2 deletion and further mTORC1 activation.


Scientific Report<br />

TRANSPARENT<br />

PROCESS<br />

Epigenetic predisposition to reprogramming fates in<br />

somatic cells<br />

Maayan Pour 1 , † , Inbar Pilzer 1 , † , Roni Rosner 1 , Zachary D Smith 2 , Alexander Meissner 2 and Iftach Nachman*, 1<br />

1<br />

Department of Biochemistry and Molecular Biology, Tel Aviv University, Tel Aviv, Israel<br />

2<br />

Department of <strong>Stem</strong> Cell and Regenerative Biology, Harvard University, Cambridge, MA, USA<br />

*<br />

Corresponding author. Tel: +972 3 640 5900; E‐mail: iftachn@post.tau.ac.il<br />

† These authors contributed equally to this work<br />

DOI 10.15252/embr.201439264 | Published online 19.01.2015<br />

EMBO reports (2015) 16, 370-378<br />

Reprogramming to pluripotency is a low‐efficiency process at the population level. Despite notable advances to<br />

molecularly characterize key steps, several fundamental aspects remain poorly understood, including when the<br />

potential to reprogram is first established. Here, we apply live‐cell imaging combined with a novel statistical approach<br />

to infer when somatic cells become fated to generate downstream pluripotent progeny. By tracing cell lineages<br />

from several divisions before factor induction through to pluripotent colony formation, we find that pre‐induction<br />

sister cells acquire similar outcomes. Namely, if one daughter cell contributes to a lineage that generates induced<br />

pluripotent stem cells (iPSCs), its paired sibling will as well. This result suggests that the potential to reprogram<br />

is predetermined within a select subpopulation of cells and heritable, at least over the short term. We also find<br />

that expanding cells over several divisions prior to factor induction does not increase the per‐lineage likelihood<br />

of successful reprogramming, nor is reprogramming fate correlated to neighboring cell identity or cell‐specific<br />

reprogramming factor levels. By perturbing the epigenetic state of somatic populations with Ezh2 inhibitors prior<br />

to factor induction, we successfully modulate the fraction of iPSC‐forming lineages. Our results therefore suggest<br />

that reprogramming potential may in part reflect preexisting epigenetic heterogeneity that can be tuned to alter the<br />

cellular response to factor induction.<br />

Synopsis<br />

The potential of mouse embryonic fibroblasts (MEFs) to generate iPSCs is determined<br />

before OKSM induction and symmetrically maintained over the short term. Epigenetic<br />

perturbations of MEFs can alter their future response to reprogramming.<br />

••<br />

The potential to generate iPSC colonies is shared between sister lineages emanating from a<br />

pre‐induction cell division.<br />

••<br />

Cell‐specific OKSM levels or local niche effects do not explain preference toward iPSC fate.<br />

••<br />

Perturbing H3K27 or H3K4 methylation marks prior to OKSM induction increases the<br />

number of iPSC lineages.


Review<br />

Harnessing the apoptotic programs in cancer<br />

stem‐like cells<br />

Ying‐Hua Wang 1 , 2 , 3 and David T Scadden*, 1 , 2 , 3<br />

1<br />

Center for Regenerative Medicine and Cancer Center, Massachusetts General Hospital, Boston, MA, USA<br />

2<br />

Harvard <strong>Stem</strong> Cell Institute, Cambridge, MA, USA<br />

3<br />

Department of <strong>Stem</strong> Cell and Regenerative Biology, Harvard University, Cambridge, MA, USA<br />

*<br />

Corresponding author. Tel: +1 617 7265615; E‐mail: dscadden@mgh.harvard.edu<br />

DOI 10.15252/embr.201439675 | Published online 07.08.2015<br />

EMBO reports (2015) 16, 1084-1098<br />

Elimination of malignant cells is an unmet challenge for most human cancer types even with therapies targeting<br />

specific driver mutations. Therefore, a multi‐pronged strategy to alter cancer cell biology on multiple levels is<br />

increasingly recognized as essential for cancer cure. One such aspect of cancer cell biology is the relative apoptosis<br />

resistance of tumor‐initiating cells. Here, we provide an overview of the mechanisms affecting the apoptotic<br />

process in tumor cells emphasizing the differences in the tumor‐initiating or stem‐like cells of cancer. Further, we<br />

summarize efforts to exploit these differences to design therapies targeting that important cancer cell population.<br />

Synopsis<br />

Cancer stem-like cells are less sensitive to apoptotic signals and<br />

more resistant to therapy. This review discusses the regulation of<br />

apoptosis in CSCs, focusing on the therapeutical modulation of<br />

pro- and anti-apoptotic pathways to induce CSC death.


Review<br />

Effects of inflammation on stem cells:<br />

together they strive?<br />

Caghan Kizil*,1,2,†, Nikos Kyritsis2,† and Michael Brand*,2<br />

1<br />

German Centre for Neurodegenerative Diseases (DZNE) Dresden within the Helmholtz Association, Dresden, Germany<br />

2<br />

DFG‐Center for Regenerative Therapies Dresden, Cluster of Excellence (CRTD) of the Technische Universität Dresden, Dresden, Germany<br />

* Corresponding author. Tel: +49 351 458 82315; E‐mail: caghan.kizil@dzne.de<br />

Corresponding author. Tel: +49 351 458 82300; E‐mail: michael.brand@biotec.tu-dresden.de<br />

† These authors contributed equally to this work<br />

DOI 10.15252/embr.201439702 | Published online 04.03.2015<br />

EMBO reports (2015) 16, 416-426<br />

Inflammation entails a complex set of defense mechanisms acting in concert to restore the homeostatic balance in<br />

organisms after damage or pathogen invasion. This immune response consists of the activity of various immune<br />

cells in a highly complex manner. Inflammation is a double‐edged sword as it is reported to have both detrimental<br />

and beneficial consequences. In this review, we discuss the effects of inflammation on stem cell activity, focusing<br />

primarily on neural stem/progenitor cells in mammals and zebrafish. We also give a brief overview of the effects<br />

of inflammation on other stem cell compartments, exemplifying the positive and negative role of inflammation<br />

on stemness. The majority of the chronic diseases involve an unremitting phase of inflammation due to improper<br />

resolution of the initial pro‐inflammatory response that impinges on the stem cell behavior. Thus, understanding<br />

the mechanisms of crosstalk between the inflammatory milieu and tissue‐resident stem cells is an important basis<br />

for clinical efforts. Not only is it important to understand the effect of inflammation on stem cell activity for further<br />

defining the etiology of the diseases, but also better mechanistic understanding is essential to design regenerative<br />

therapies that aim at micromanipulating the inflammatory milieu to offset the negative effects and maximize the<br />

beneficial outcomes.<br />

Synopsis<br />

Inflammation is an important defense mechanism against<br />

pathogens or tissue damage, but it can also have detrimental<br />

consequences. This review focuses specifically on the effects of<br />

inflammation on stem cells.


NOTES


FURTHER further reading READING<br />

The EMBO Journal<br />

Articles<br />

Stat3 promotes mitochondrial<br />

transcription and oxidative<br />

respiration during maintenance<br />

and induction of naive<br />

pluripotency<br />

Elena Carbognin, Riccardo M Betto,<br />

Maria E Soriano, Austin G Smith,<br />

Graziano Martello<br />

DOI 10.15252/embj.201592629<br />

Published online: 22.02.2016<br />

Oct4-induced oligodendrocyte<br />

progenitor cells enhance<br />

functional recovery in spinal<br />

cord injury model<br />

Jeong Beom Kim, Hyunah Lee, Marcos J<br />

Araúzo-Bravo, Kyujin Hwang, Donggyu<br />

Nam, Myung Rae Park, Holm Zaehres,<br />

Kook In Park, Seok-Jin Lee<br />

DOI 10.15252/embj.201592652<br />

Published online: 23.10.2015<br />

Notch signaling regulates<br />

gastric antral LGR5 stem cell<br />

function<br />

Elise S Demitrack, Gail B Gifford,<br />

Theresa M Keeley, Alexis J Carulli, Kelli<br />

L VanDussen, Dafydd Thomas, Thomas<br />

J Giordano, Zhenyi Liu, Raphael Kopan,<br />

Linda C Samuelson<br />

DOI:10.15252/embj.201490583<br />

Published online: 13.08.2015<br />

The tRNA methyltransferase<br />

Dnmt2 is required for accurate<br />

polypeptide synthesis during<br />

haematopoiesis<br />

Francesca Tuorto, Friederike Herbst,<br />

Nader Alerasool, Sebastian Bender,<br />

Oliver Popp, Giuseppina Federico,<br />

Sonja Reitter, Reinhard Liebers, Georg<br />

Stoecklin, Hermann-Josef Gröne,<br />

Gunnar Dittmar, Hanno Glimm, Frank<br />

Lyko<br />

DOI:10.15252/embj.201591382<br />

Published online: 13.08.2015<br />

Robust intestinal homeostasis<br />

relies on cellular plasticity in<br />

enteroblasts mediated by miR-<br />

8–Escargot switch<br />

Zeus A Antonello, Tobias Reiff, Esther<br />

Ballesta-Illan, Maria Dominguez<br />

DOI:10.15252/embj.201591517<br />

Published online: 15.06.2015<br />

Neuropeptide Y regulates<br />

the hematopoietic stem cell<br />

microenvironment and prevents<br />

nerve injury in the bone marrow<br />

Min Hee Park, Hee Kyung Jin, Woo-Kie<br />

Min, Won Woo Lee, Jeong Eun Lee,<br />

Haruhiko Akiyama, Herbert Herzog,<br />

Grigori N Enikolopov, Edward H<br />

Schuchman, Jae-sung Bae<br />

DOI:10.15252/embj.201490174<br />

Published online: 27.04.2015<br />

Controlled induction of human<br />

pancreatic progenitors produces<br />

functional beta-like cells in vitro<br />

Holger A Russ, Audrey V Parent, Jennifer<br />

J Ringler, Thomas G Hennings, Gopika<br />

G Nair, Mayya Shveygert, Tingxia Guo,<br />

Sapna Puri, Leena Haataja, Vincenzo<br />

Cirulli, Robert Blelloch, Greg L Szot,<br />

Peter Arvan, Matthias Hebrok<br />

DOI:10.15252/embj.201591058<br />

Published online: 23.04.2015<br />

Sox2, Tlx, Gli3, and Her9<br />

converge on Rx2 to define retinal<br />

stem cells in vivo<br />

Robert Reinhardt, Lázaro Centanin,<br />

Tinatini Tavhelidse, Daigo Inoue, Beate<br />

Wittbrodt, Jean-Paul Concordet, Juan<br />

Ramón Martinez-Morales, Joachim<br />

Wittbrodt<br />

DOI:10.15252/embj.201490706<br />

Published online: 23.04.2015<br />

Telomerase abrogates<br />

aneuploidy-induced telomere<br />

replication stress, senescence<br />

and cell depletion<br />

Meena, Jitendra K.; Cerutti, Aurora;<br />

Beichler, Christine; Morita, Yohei;<br />

Bruhn, Christopher; Kumar, Mukesh;<br />

Kraus, Johann M.; Speicher, Michael<br />

R.; Wang, Zhao-Qi; Kestler, Hans A.;<br />

di Fagagna, Fabrizio d’Adda; Guenes,<br />

Cagatay; Rudolph, Karl Lenhard<br />

DOI: 10.15252/embj.201490070<br />

Published online: 27.03.2015<br />

Smg6/Est1 licenses embryonic<br />

stem cell differentiation via<br />

nonsense–mediated mRNA<br />

decay<br />

Tangliang Li, Yue Shi, Pei Wang, Luis<br />

Miguel Guachalla, Baofa Sun, Tjard<br />

Joerss, Yu-Sheng Chen, Marco Groth,<br />

Anja Krueger, Matthias Platzer, Yun-Gui<br />

Yang, Karl Lenhard Rudolph, Zhao-Qi<br />

Wang<br />

DOI:10.15252/embj.201489947<br />

Published online: 14.03.2015<br />

Snai1 regulates cell lineage<br />

allocation and stem cell<br />

maintenance in the mouse<br />

intestinal epithelium<br />

Katja Horvay, Thierry Jardé, Franca<br />

Casagranda, Victoria M Perreau,<br />

Katharina Haigh, Christian M Nefzger,<br />

Reyhan Akhtar, Thomas Gridley, Geert<br />

Berx, Jody J Haigh, Nick Barker, Jose M<br />

Polo, Gary R Hime, Helen E Abud<br />

DOI 10.15252/embj.201490881<br />

Published online: 10.03.2015<br />

Human primordial germ cell<br />

commitment in vitro associates<br />

with a unique PRDM14<br />

expression profile<br />

Fumihiro Sugawa, Marcos J Araúzo-<br />

Bravo, Juyong Yoon, Kee-Pyo Kim,<br />

Shinya Aramaki, Guangming Wu,<br />

Martin Stehling, Olympia E Psathaki,<br />

Karin Hübner, Hans R Schöler<br />

DOI 10.15252/embj.201488049<br />

Published online: 06.03.2015<br />

emboj.embopress.org<br />

EMBO Molecular Medicine<br />

Research Articles<br />

Aberrant epigenome in iPSCderived<br />

dopaminergic neurons<br />

from Parkinson’s disease<br />

patients<br />

Rubén Fernández-Santiago, Iria<br />

Carballo-Carbajal, Giancarlo Castellano,<br />

Roger Torrent, Yvonne Richaud, Adriana<br />

Sánchez-Danés, Roser Vilarrasa-Blasi,<br />

Alex Sánchez-Pla, José Luis Mosquera,<br />

Jordi Soriano, José López-Barneo, Josep<br />

M Canals, Jordi Alberch, Ángel Raya,<br />

Miquel Vila, Antonella Consiglio, José<br />

I Martín-Subero, Mario Ezquerra,<br />

Eduardo Tolosa<br />

DOI 10.15252/emmm.201505439 |<br />

Published online: 01.12.2015<br />

The MICA-129 dimorphism<br />

affects NKG2D signaling and<br />

outcome of hematopoietic stem<br />

cell transplantation<br />

Antje Isernhagen, Dörthe Malzahn,<br />

Elena Viktorova, Leslie Elsner, Sebastian<br />

Monecke, Frederike von Bonin, Markus<br />

Kilisch, Janne Marieke Wermuth, Neele<br />

Walther, Yesilda Balavarca, Christiane<br />

Stahl-Hennig, Michael Engelke, Lutz<br />

Walter, Heike Bickeböller, Dieter Kube,<br />

Gerald Wulf, Ralf Dressel<br />

DOI 10.15252/emmm.201505246 |<br />

Published online: 19.10.2015<br />

Reviews<br />

Reprogramming and<br />

transdifferentiation for<br />

cardiovascular development and<br />

regenerative medicine: where<br />

do we stand?<br />

Antje D Ebert, Sebastian Diecke, Ian Y<br />

Chen, Joseph C Wu<br />

DOI 10.15252/emmm.201504395 |<br />

Published online: 16.07.2015<br />

Lost in translation: pluripotent<br />

stem cell-derived hematopoiesis<br />

Mania Ackermann, Steffi Liebhaber,<br />

Jan-Henning Klusmann, Nico Lachmann<br />

DOI 10.15252/emmm.201505301 |<br />

Published online: 14.07.2015<br />

embomolmed.embopress.org<br />

Molecular Systems Biology<br />

Articles<br />

The bimodally expressed<br />

microRNA miR-142 gates exit<br />

from pluripotency<br />

Hanna L Sladitschek, Pierre A Neveu<br />

DOI 10.15252/msb.20156525 |<br />

Published online: 21.12.2015<br />

Hierarchical folding and<br />

reorganization of chromosomes<br />

are linked to transcriptional<br />

changes in cellular<br />

differentiation<br />

James Fraser, Carmelo Ferrai, Andrea<br />

M Chiariello, Markus Schueler, Tiago<br />

Rito, Giovanni Laudanno, Mariano<br />

Barbieri, Benjamin L Moore, Dorothee<br />

CA Kraemer, Stuart Aitken, Sheila Q<br />

Xie, KellyJ Morris, Masayoshi Itoh,<br />

Hideya Kawaji, InesJaeger, Yoshihide<br />

Hayashizaki, Piero Carninci, Alistair RR<br />

Forrest, , Colin A Semple, Josée Dostie,<br />

Ana Pombo, Mario Nicodemi<br />

DOI 10.15252/msb.20156492 |<br />

Published online: 23.12.2015<br />

Review<br />

Cell dynamics and gene<br />

expression control in tissue<br />

homeostasis and development<br />

Pau Rué, Alfonso Martinez Arias<br />

DOI 10.15252/msb.20145549 |<br />

Published online: 25.02.2015<br />

msb.embopress.org


embor.embopress.org<br />

217407 / MLEA033411

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!