Cell Sept 2011
Cell Sept 2011
Cell Sept 2011
In This Issue
Chromothripsis Moves beyond Cancer
PAGE 889
ID proteins inhibit differentiation and maintain stem cell fate and are ubiquitinated and degraded in differentiated tissues.
However, in many neoplasms, IDs appear to escape degradation. Williams et al. now identify USP1 as the deubiquitinating
enzyme controlling ID protein stability and show that, by promoting ID stability, USP1 preserves stem cell-like characteristics
of osteosarcoma tumor cells. Their findings suggest that the USP1-ID axis that normally controls bone development is
usurped to propagate osteosarcoma tumor stem cells and point to USP1 as an attractive target for differentiation therapy.
The eukaryotic replicative helicase complex CMG is loaded onto dsDNA, but it could, in principle, unwind DNA by translocating along ssDNA or dsDNA. By colliding replisomes with strand-specific roadblocks in Xenopus egg extracts, Fu et al. show
that CMG can bypass a lagging strand, but not a leading strand roadblock, strongly supporting a 3-to-5 ssDNA translocation
mechanism. Replication initiation therefore likely involves reconfiguration from a dsDNA to an ssDNA binding mode.
Sugioka et al. show that extrinsic signals that remodel the cytoskeleton lead to
asymmetric gene expression and cell fate in the dividing C. elegans embryo.
Wnt signaling generates an asymmetry in the astral microtubules that, in turn,
influences the levels of b-catenin in the nuclei of daughter cells as they separate
from each other. The findings point to a direct role of cytoskeletal dynamics in
transcriptional regulation.
Griffin et al. show that the MEX-5 protein gradient in the C. elegans zygote
arises from a localized phosphorylation/dephosphorylation cycle that regulates
the rate of MEX-5 diffusion. These results present a paradigm for protein concentration gradient formation that does not require localized protein synthesis
or degradation.
Toggling Hunger
PAGE 992
How is hunger neuronally coded? Using a combination of optogenetic, electrophysiological, and pharmacological
approaches, Yang et al. demonstrate that hunger or an appetite-stimulating hormone induces synaptic activity in the neurons
that drive feeding behavior. This activity persists due to a positive feedback loop involving AMPK and neuronal calcium
release until leptin, a hormone signaling satiety, switches it off. The findings reveal a neuronal circuit, switched on and off
by hormonal pulses, that operates as a memory storage device for a physiological state.
Choi et al. probe how the brain translates sensory stimulation such as odors to behavior. By artificially stimulating clusters of
neurons in a region where olfactory information is processed (the piriform cortex), the authors show that activating the same
arbitrarily chosen neuronal cluster can lead to opposite behavioral responses, depending on the experience (aversive or
attractive) coupled to the neuronal stimulation. The findings highlight the plasticity of responses to olfactory stimuli and
indicate that spatial order in the piriform does not inform odorant identity or behavioral output.
The diversity of histone modifications and their many functions becomes ever
more complex. Tan et al. now identify 67 histone marks, including lysine crotonylation. This modification is evolutionarily conserved, abundant in core histones, and marks either active promoters or potential enhancers. Their data
also suggest that lysine crotonlyation marks active sex chromosome genes in
male germ cells.
Genome-wide reference maps of DNA methylation are critical to an understanding of its many roles in gene regulation. Molaro et al. generate full DNA
methylation profiles of human and chimp sperm and compare these to embryonic stem cells. Their findings reveal distinct properties for the epigenetic
reprogramming that occurs in germ cells and somatic cells during mammalian
development and provide insight into the relationships between the evolution of
the genome and the epigenome.
Cell 146, September 16, 2011 2011 Elsevier Inc. 845
Leading Edge
Select
Human Evolutionary Genomics
The forces that have shaped the evolution of human genomes are now coming into sharp
focus. This issues Select explores recent discoveries that change our view of how our genomes
became what they arefrom the impact of interbreeding between ancient modern humans and
Neanderthals to evidence that rates of de novo mutation may be subject to large variation between
individuals.
African Hotspots
A more recent admixture occurred in the Americas between people of European and African ancestries. By examining
sites of recombination in the genomes of 30,000 African Americans, Hinch et al. (2011) provide a detailed map of crossover frequency with surprising results. At large size scales (>3 Mb), most of the recombination events match up with
those previously reported in studies of European genomes. However, at smaller size scales, the authors find 2,500
hotspots that are predominantly active in West Africans. A closely related study by Berg et al. (2011) similarly demonstrates the existence of these hotspots from a survey of Hap-Map genotype data and subsequent analysis of meiotic
recombination crossover events in sperm from African and European males. Both groups provide evidence that this
striking variability between African and non-African populations is due to different alleles of PRDM9, a gene encoding
a DNA-binding histone methyltransferase, which is expressed during meiosis and involved in recombination. Further
analysis of the enriched hotspots shows that they often contain a 17 bp DNA sequence, which is predicted to bind
the alleles of PRDM9 commonly found in Africans. These findings indicate that recombination frequency and location
are more variable than previously thought. Future work may confirm the prediction that the risk of disease-causing
genomic rearrangements is higher in individuals carrying these particular PRDM9 alleles or may delve further into
why the 17 bp motif is singled out by particular PRDM9 variants.
Berg, I.L., et al. (2011). Proc. Natl. Acad. Sci. USA 108, 1237812383.
Hinch, A.G., et al. (2011). Nature 476, 170175.
Cell 146, September 16, 2011 2011 Elsevier Inc. 847
Leading Edge
BenchMarks
Molecular Mechanism
of Protein Folding in the Cell
James E. Rothman1,* and Randy Schekman2,*
1Department
of Cell Biology, Yale University School of Medicine, New Haven, CT 06520-8002, USA
of Cell and Molecular Biology, HHMI, University of California, Berkeley, CA 94720-1302, USA
*Correspondence: james.rothman@yale.edu (J.E.R.), schekman@berkeley.edu (R.S.)
DOI 10.1016/j.cell.2011.08.041
2Department
F.-Ulrich Hartl and Arthur Horwich will share this years Lasker Basic Medical Science Award for the
discovery of the cells protein-folding machinery, exemplified by cage-like structures that convert
newly synthesized proteins into their biologically active forms. Their fundamental findings reveal
mechanisms that operate in normal physiologic processes and help to explain the problems that
arise in diseases of protein folding.
How is it that, even when proteins cannot
fold on their own in vitro, they somehow
fold beautifully in the much more complex
environment of a cell? As with most
breakthrough discoveries in biology, the
answer is remarkably simple: the cell provides special protein machinery, called
molecular chaperones, that surrounds
the folding protein and removes it from
the rest of the cell.
The surfaces of this remarkable machinery are very forgiving. They actually
utilize metabolic energy (i.e., ATP hydrolysis) to alternate their physical chemistry
between hydrophobic states and hydrophilic states, which restarts the folding
when it stalls and expels the protein
when folding is completed. In the case
of the heat shock protein 60 (Hsp60), a
molecular chaperone shaped like a test
tube, access to the cavity is highly selective and limited to only nascent proteins or
those purposely unfolded to repair conformational damage. Analogous specialized test tube environments, also based
on limited access to the internal cavity of
a ring-based protein structure, enable targeted proteolysis (by proteosomes),
making this strategy a general principle
in cellular biology.
These singular discoveries are celebrated in this years Lasker Basic Medical
Research Award to F.-Ulrich Hartl and
Arthur Horwich. In collaborative work,
they discovered the fundamental principle
that protein folding in the cell is generally
a facilitated process, and in independent
and often complementary work, they
and their collaborators established the
individual components of
that Horwich and Hartl stand
what we now recognize as
out (Figure 1).
the core chaperone protein
It is noteworthy that this
machinery, including phage
remarkable set of experiassembly factors and proments was possible only
teins upregulated by stress.
because of previous mechaAnd, by the late 1980s, all
nistic insights into how prothe proteins central for protein
teins encoded in the nucleus
folding were virtually known
are imported into mitochonin both structural and broad
dria. These insights emerged
phenotypic terms. In parmainly from the laboratories
ticular, John Ellis, Costa
of Walter Neupert and GottGeorgopoulos, and George
fried Schatz in the previous
Lorimer deserve special redecade. Not only was Hartl a
cognition for their genetic
protege of Neupert, but
and biochemical description
much of his early work on
of the protein-folding catathe Hsp60 chaperones was
carried out directly in Neulysts. Georgopoulos was the
perts laboratory, as reflected
first to define the genes for all
Figure 1. Chaperone Pathway for the Folding in the E. coli Cytosol
the major folding catalysts.
in Neuperts coauthorship of
The Hsp70 chaperones DnaK/DnaJ stabilize the newly synthesized protein in
a nonaggregated, folding-competent state. Transferring the protein into the
His work as a graduate student
the landmark papers. Schatz
central cavity of the Hsp60 chaperone GroEL requires the nucleotide
in the early 1970s and then
and collaborators introduced
exchange factor GrpE. GroES binds to GroEL in an ATP-dependent reaction
later as a postdoctoral fellow
the background methods inand displaces the unfolded protein into an enclosed folding cage. The protein
led to the mapping of the
volving DHFR translocation
is allowed to fold inside the cage for 10 s, the time needed for GroEL to
complete one round of ATP hydrolysis (of 7 ATPs). Binding of ATP to the
E. coli genes GroEL, GroES,
and folding.
opposite ring of GroEL induces an allosteric change that triggers the opening
DnaK, DnaJ, and GrpE.
of the folding chamber. Folded protein is released, whereas incompletely
These genes all emerged
Hsp70s and Folding
folded protein re-binds to GroEL for another round of attempted folding in the
GroEL-GroES cage.
from genetic screens for
in the Cytoplasm
mutants that affect l phage
The Hsp70 protein family
assembly and, in the case of
provides a second system of
the latter three genes, play a role in phage Hartl and Horwich isolated and char- molecular chaperones. These proteins
and chromosome replication (Georgo- acterized mutations in mitochondrial are a family of monomeric ATPases,
poulos and Herskowitz, 1971).
chaperone Hsp60/GroEL. Surprisingly, with Georgopoulos bacterial DnaK (an
In an entirely independent develop- mitochondrial proteins were synthesized Hsp70) and its cofactors DnaJ and GrpE
ment, in 1980 Ellis discovered the first and imported successfully, but they failed as the prototype of these chaperones. In
molecular chaperone (Barraclough and to fold. This was the first proof that protein all likelihood, the Hsp70 chaperones
Ellis, 1980). He found that unassembled folding required molecular assistance by supply the most general mechanism of
subunits of the Rubisco enzyme complex binding proteins (chaperones) in the facilitated protein folding, as the GroEL
in chloroplasts were associated with ano- living cell (Cheng et al., 1989). The same system is required for only a small fraction
ther protein that turned out to be a chap- team published a second landmark paper of the cells proteins (Houry et al., 1999).
erone. In 1988, Ellis and Georgopoulos (Ostermann et al., 1989) showing that
Key experiments in the laboratories of
discovered that this chaperone is related mitochondrial Hsp60 from mitochondria Hugh Pelham and Elizabeth Craig first
to the GroEL molecule involved in phage assists the folding of monomeric polypep- associated the Hsp70 proteins with proassembly (Hemmingsen et al., 1988). tide chains and requires ATP binding and tein folding in the cytoplasm and in quality
Ellis coined the term chaperonin for hydrolysis to do so. This work focused on control inspection in the secretory paththe GroE-type folding catalysts that are the folding of dihydrofolate reductase way. The DnaK proteins are related to
required to assist a variety of cellular pro- (DHFR) when its imported into isolated a family of heat-induced proteins, called
cesses within organelles and in the cyto- mitochondria, and it established the prin- the Hsp70s, which were first described
plasm of all types of cells. The following ciple of assisted folding in the cell for sin- as heat shock proteins in Drosophila.
year and in subsequent detailed enzymo- gle polypeptide chains, even those that Rothman and colleagues discovered that
logic studies, George Lorimer reconsti- can fold spontaneously and efficiently as the Hsp70s bind and hydrolyze ATP
tuted the assembly of Rubisco subunits pure proteins. This marked a radical de- (Chappell et al., 1986). With this in mind,
with isolated bacterial chaperonins (Go- parture from the earlier view that protein Pelham was the first to speculate that
loubinoff et al., 1989).
folding in the cell can be a spontaneous Hsp70s promote proper folding or reasThese foundational studies beautifully process. It is on the basis of these papers sembly of proteins by using cycles of
set the stage for Hartls and Horwichs and their future seminal contributions in ATP hydrolysis to bind and release agtwo landmark papers in 1989. In a key establishing the key concepts and molec- gregated, denatured proteins (Pelham,
genetic and biochemical experiment, ular mechanisms in cellular protein folding 1986). Pelham was also the first to
852 Cell 146, September 16, 2011 2011 Elsevier Inc.
Leading Edge
BenchMarks
Artemisinin: Discovery
from the Chinese Herbal Garden
Louis H. Miller1,* and Xinzhuan Su1
1Laboratory of Malaria and Vector Research, National Institute of Allergy and Infectious Diseases, National Institutes of Health,
Rockville, MD 20852, USA
*Correspondence: lmiller@niaid.nih.gov
DOI 10.1016/j.cell.2011.08.024
This years Lasker DeBakey Clinical Research Award goes to Youyou Tu for the discovery of artemisinin and its use in the treatment of malariaa medical advance that has saved millions of lives
across the globe, especially in the developing world.
The future benefits of many seminal discoveries in basic biomedical sciences
are not always obvious in the short run.
But for a handful of others, the impact
on human health is immediately clear.
Such is the case for the discovery by
Youyou Tu and colleagues of artemisinin
(also known as Qinghaosu) for treatment
of malaria. Artemisinin has been the frontline treatment since the late 1990s and
has saved countless lives, especially
among the worlds poorest children.
The Promise of Project 523
The story of artemisinin began in the unlikely
atmosphere of the Cultural Revolution in
China as a government initiative to aid the
North Vietnamese in their war with the United States. During the war, malaria caused
by
chloroquine-resistant
Plasmodium
falciparum was a major problem that
spurred research efforts on both sides of
the battlefield. In the US, these efforts culminated in the discovery of mefloquine, a
compound that was curative in a single
dose against chloroquine-resistant parasites (Trenholme et al., 1975). The North Vietnamese, however, lacked a research infrastructure and thus turned to China for help.
Under the instructions of Chairman Mao
and Premier Zhou, a meeting was held on
May 23, 1967 in Beijing to discuss the
problem of drug-resistant malaria parasites. This led to a secret nationwide program called project 523, involving over
500 scientists in 60 different laboratories
and institutes (Zhang et al., 2006).
Although the projects short-term goal
was to produce antimalarial drugs that
could immediately be used in the battlefield (by 1969 three treatments were
REFERENCES
Charman, S.A., Arbe-Barnes, S., Bathurst, I.C.,
Brun, R., Campbell, M., Charman, W.N., Chiu,
F.C.K., Chollet, J., Craft, J.C., Creek, D.J., et al.
(2011). Proc. Natl. Acad. Sci. USA 108, 44004405.
Dondorp, A.M., Nosten, F., Yi, P., Das, D., Phyo,
A.P., Tarning, J., Lwin, K.M., Ariey, F., Hanpithakpong, W., Lee, S.J., et al. (2009). N. Engl. J. Med.
361, 455467.
Fidock, D.A., Nomura, T., Talley, A.K., Cooper,
R.A., Dzekunov, S.M., Ferdig, M.T., Ursos, L.M.,
Sidhu, A.B., Naude, B., Deitsch, K.W., et al.
(2000). Mol. Cell 6, 861871.
Jiang, J.-B., Li, G.-Q., Guo, X.-B., Kong, Y.C., and
Arnold, K. (1982). Lancet 2, 285288.
Klayman, D.L. (1985). Science 228, 10491055.
Klonis, N., Crespo-Ortiz, M.P., Bottova, I., AbuBakar, N., Kenny, S., Rosenthal, P.J., and Tilley,
L. (2011). Proc. Natl. Acad. Sci. USA 108, 11405
11410.
Li, G.Q., Arnold, K., Guo, X.B., Jian, H.X., and Fu,
L.C. (1984). Lancet 2, 13601361.
OMeara, W.P., Mangeni, J.N., Steketee, R., and
Greenwood, B. (2010). Lancet Infect. Dis. 10,
545555.
Leading Edge
Previews
SIP-ing the Elixir of Youth
William Mair,1,2,3,4,* Kristan K. Steffen,1,2,3 and Andrew Dillin1,2,3,*
1The
AMP-activated protein kinase (AMPK) is a conserved cellular fuel gauge previously implicated in
aging. In this issue, Lu et al. (2011) describe how age-related deacetylation of Sip2, a subunit of
the AMPK homolog in yeast, acts as a life span clock that can be wound backward or forward to
modulate longevity.
From the age of Hippocrates, medicine
has long appreciated the importance of
preserving physiological homeostasis for
a healthy life. Indeed, the ancient Greeks
believed that a failure to maintain proper
balance between the four humors, or
liquids of the blood, was at the root of
human disease. Despite modern medicines departure from humorism, the idea
that a cellular collapse in homeostasis
might underlie late onset diseases still
prevails. In this issue, Lu and colleagues
(Lu et al., 2011) present new data supporting this ancient theory.
One fundamental cellular process requiring tight homeostatic control during
aging is the balance between energy
generation and usage. In eukaryotes, a
key orchestrator of energy homeostasis
is AMPK, a serine/threonine kinase that
functions as a cellular energy sensor, activating catabolic processes to generate
ATP when energy levels are low while
simultaneously shutting down energy requiring anabolic processes (Steinberg
and Kemp, 2009). The importance of
maintaining energy homeostasis throughout life is underscored by the fact that
AMPK activity mediates life span in yeast,
worms, and flies. Uncovering a novel
mode of AMPK regulation in yeast, the
work by Lu et al. (2011) demonstrates that
progressive deacetylation of the AMPK
subunit Sip2 with age limits replicative life
span (defined as the number of times that
a single cell can divide) and that blocking
this change can slow aging.
The S. cerevisiae homolog of AMPK is
a heterotrimeric complex made up of
both a single a and g subunit, Snf1 and
Snf4 (sucrose nonfermenting), respec-
Leading Edge
Previews
A New Shield for a Cytokine Storm
Akiko Iwasaki1,* and Ruslan Medzhitov1,2,*
1Department
of Immunobiology
Hughes Medical Institute
School of Medicine, Yale University New Haven, CT, 06520, USA
*Correspondence: akiko.iwasaki@yale.edu (A.I.), ruslan.medzhitov@yale.edu (R.M.)
DOI 10.1016/j.cell.2011.08.027
2Howard
Highly virulent influenza virus infection results in excessive cytokine production, recruitment of
leukocytes, and immune-mediated pulmonary injury. Teijaro et al. (2011) now demonstrate that
sphingosine-1-phosphate receptor 1 ligands suppress all features of flu-inflicted pathological
inflammation and place the endothelium at the center of this regulatory network.
Recognition and rapid clearance of pathogens by the innate immune system provide the first line of defense in metazoan
organisms. However, excessive activation of the innate immune system in
response to pathogens can lead to pathological inflammatory consequences. In
the case of highly virulent 1918 and avian
H5N1 influenza virus infections, early
recruitment of inflammatory leukocytes
to the lung, followed by excessive early
cytokine responses (known as a cytokine
storm), is considered to be the key contributor to morbidity and mortality of the
infection (Tscherne and Garca-Sastre,
2011). Likewise, the virulence of the pandemic 2009 H1N1 swine flu is associated with viral replication in the lower
respiratory tract and more severe pulmonary damage compared to seasonal flu
(Tscherne and Garca-Sastre, 2011).
However, the cell types that are responsible for the initiation and amplification of
the cytokine storms that follow virulent
influenza infection remain unclear. In this
issue, Teijaro et al. (2011) have uncovered
an unexpected role of endothelial cells in
coordinating the inflammatory sequelae
via sphingosine-1-phosphate signaling.
Sphingosine-1-phosphate is a metabolite of sphingolipid and is a ligand for a
family of five G protein-coupled receptors, S1P15 (Rosen et al., 2009). Differential expression of these receptors
enables control of angiogenesis, heart
development, and immunity in a highly
specific manner. The sphingosine analog
FTY720, a well-known immunomodulator
that has recently been approved for the
treatment of multiple sclerosis, is phosphorylated in vivo by sphingosine kinase
2 to produce a ligand for the S1P re-
Leading Edge
Previews
Synaptic Plasticity of Feeding Circuits:
Hormones and Hysteresis
Marcelo O. Dietrich1,4 and Tamas L. Horvath1,2,3,*
1Section
of Comparative Medicine
of Obstetrics, Gynecology and Reproductive Sciences
3Department of Neurobiology
Yale University School of Medicine, New Haven, CT 06520, USA
4Programa de Po
s-graduacao em Bioqumica, Department of Biochemistry, Universidade Federal do Rio Grande do Sul,
Porto Alegre RS 90035, Brazil
*Correspondence: tamas.horvath@yale.edu
DOI 10.1016/j.cell.2011.08.031
2Department
The drive to eat is controlled by neuronal circuits in the hypothalamus that respond to hormones
signaling hunger or satiety. In this issue of Cell, Yang et al. (2011) reveal an AMPK-dependent
synaptic pathway that sustains excitatory stimulation of the NPY/AgRP neurons that promote
feeding behavior until satiety signals kick in.
Metabolic hormones such a ghrelin, signaling food deprivation, and leptin, signaling satiety, stimulate synaptic activity
and plasticity within the neuronal circuits
in the hypothalamus that control feeding
behavior (Pinto et al., 2004). This process
was suggested to play an important role
in regulating metabolism (Horvath and
Diano, 2004). However, the mechanisms
that bring about these rapid changes in
synaptic connectivity and activity and
how long they persist remain ill defined.
In this issue of Cell, Yang et al. (2011)
provide a remarkable set of novel findings
pinpointing intracellular and intercellular
substrates of synaptic plasticity on the
orexigenic NPY/AgRP neurons. Their
findings demonstrate why the response
to a pulse of the appetite-stimulating hormone ghrelin can persist for hours, and
how it can be turned off in response to
a pulse of leptin. Using in vitro electrophysiological approaches to record from
neurons marked by the expression of the
neuropeptide NPY, most of which also
synthesize the neuropeptide AgRP, the
authors found that food deprivation increases the frequency of action potentials
in the these neurons (referred to as AgRP
neurons by Yang et al.). The increased
activity occurs through an AMPA-mediated increase in frequency of miniature
excitatory postsynaptic currents (mEPSCs)
onto these cells.
This increase in excitatory inputs is
driven by AMPK- and ryanodine re-
REFERENCES
Abizaid, A., Liu, Z.-W., Andrews, Z.B., Shanabrough, M., Borok, E., Elsworth, J.D., Roth, R.H.,
Sleeman, M.W., Picciotto, M.R., Tschop, M.H.,
et al. (2006). J. Clin. Invest. 116, 32293239.
Andrews, Z.B., Liu, Z.W., Walllingford, N., Erion,
D.M., Borok, E., Friedman, J.M., Tschop, M.H.,
Shanabrough, M., Cline, G., Shulman, G.I., et al.
(2008). Nature 454, 846851.
Claret, M., Smith, M.A., Batterham, R.L., Selman,
C., Choudhury, A.I., Fryer, L.G.D., Clements, M.,
Leading Edge
Perspective
DNA Demethylation Dynamics
Nidhi Bhutani,1,2 David M. Burns,1 and Helen M. Blau1,*
1Baxter Laboratory for Stem Cell Biology, Institute for Stem Cell Biology and Regenerative Medicine, Department of Microbiology and
Immunology, Stanford University School of Medicine, Stanford, CA 94305-5175, USA
2Present address: Department of Orthopaedic Surgery, Stanford University, Stanford, CA 94305-5341, USA
*Correspondence: hblau@stanford.edu
DOI 10.1016/j.cell.2011.08.042
function in higher metazoa and mediate 5mC hydroxylation (Tahiliani et al., 2009). Recombinant human TET1 was found to be
capable of converting 5mC to 5hmC in mammalian DNA,
providing evidence for its putative role in mediating DNA demethylation (Tahiliani et al., 2009). Like their human counterparts,
mouse TET1, 2, and 3 catalyze the conversion of 5mC to
5hmC (Ito et al., 2010).
Much of the initial excitement regarding the discovery of 5hmC
was the prediction that it could readily lift the repression of gene
expression imposed by 5mC at many gene promoters (Ito et al.,
2010; Tahiliani et al., 2009). However, like methylation, a high
concentration of 5hmC correlates with transcriptionally nonproductive or altogether inactive gene promoters (Ficz et al., 2011;
Pastor et al., 2011; Williams et al., 2011; Wu et al., 2011a; Xu
et al., 2011). Thus, contrary to expectations, the 5mC to 5hmC
modification is clearly not the functional equivalent of 5mC to
cytosine, which is associated with derepression of certain
gene promoters (Pastor et al., 2011).
The hypothesis that cytosine hydroxylation might play a functional role in maintaining the pluripotent state was first suggested
by Zhang and colleagues (Ito et al., 2010). They reported that
TET1 results in a loss of ESC self-renewal by reducing the
expression of the pluripotency regulator NANOG. This finding,
however, has been challenged by others who suggest that,
although levels of TET1 and TET2 (and therefore 5hmC) are
high in ESCs, these proteins largely mediate regulation of
lineage-specific genes, not the pluripotency regulator NANOG
(Ficz et al., 2011; Koh et al., 2011).
In parallel with studies in ES cells, other experiments have
been performed to determine whether the TET proteins are
involved in active DNA demethylation in early development. In
this case, the third family member, TET3, is most abundant
and plays a role in the rapid and active loss of 5mC in the male
pronucleus upon zygote formation prior to cell division. Recent
experiments have shed light on the mechanism underlying this
process. An increase in 5hmC is concomitant with a decrease
in 5mC in zygotes (as determined by immunohistochemistry),
suggesting that 5mC is converted to 5hmC (Iqbal et al., 2011;
Wossidlo et al., 2011). In addition, knocking down TET3 by
RNA interference (RNAi) led to an increase in 5mC. Thus, TET3
is responsible for 5hmC generation post-fertilization in mouse
zygotes, suggesting a potential role for TET proteins in DNA
demethylation early in development.
An unexpected complication of interpretations of experiments
regarding the effects of TET proteins on gene expression is that
TET1 plays a repressive role, independent of its enzymatic
activity as a hydroxylase (Williams et al., 2011; Wu et al.,
2011b). TET1 has been found to associate with two different
transcriptional repressor complexes containing PRC2 (polycomb repressive complex 2) or SIN3A (Swi-independent 3A)
(Williams et al., 2011). SIN3A and TET1 directly interact with
one another (as shown by coimmunoprecipitation), whereas
PRC2 and TET1 may act indirectly (Williams et al., 2011; Wu
et al., 2011b). Importantly, a high degree of target overlap is
observed between TET1 and PRC2, as well as between TET1
and SIN3A, in global chromatin immunoprecipitation (ChIP) analyses (Wu et al., 2011b; Williams et al., 2011). Furthermore, a
subset of TET1 target genes is upregulated upon loss of SIN3A
868 Cell 146, September 16, 2011 2011 Elsevier Inc.
repair mechanisms (Liu et al., 2008). Clearly, an increased understanding of how error-prone and error-free DNA repair pathways
are targeted to Ig versus non-Ig loci warrants further investigation, as AID is key to both antibody generation and DNA demethylation.
A role for AID in global DNA demethylation was first shown in
zebrafish embryos by Cairns and colleagues (Rai et al., 2008).
Upon overexpression of AID or zebrafish APOBEC deaminases
and the glycosylase MBD4, active DNA demethylation was observed in zebrafish embryos injected with a methylated linearized nonreplicating DNA. Reik and colleagues suggested a
similar role for AID in global DNA demethylation at a later stage
of embryogenesis in mice (Popp et al., 2010). Mice completely
lacking AID (AID / ) (Muramatsu et al., 2000) exhibited an
increase in genome-wide hypermethylation in their primordial
germ cells (PGCs), suggesting that AID is involved in DNA demethylation. However, if AID-mediated global DNA demethylation
plays a crucial role in early development, a more profound
phenotype would be expected in AID null mice, which are both
viable and fertile, albeit with somewhat smaller litter sizes
(Popp et al., 2010). These findings raise the possibility that, in
the absence of AID, other family members may play compensatory roles in DNA demethylation.
Studies of nuclear reprogramming provided the first evidence
that AID plays a role in active DNA demethylation in mammals
and in somatic cells (Bhutani, et al., 2010). Upon fusion of an
excess of mouse ESCs with human somatic cells (fibroblasts)
in nondividing heterokaryons, rapid demethylation was detected
at the promoters of the human pluripotency genes OCT4 and
NANOG, accompanied by their transcriptional induction. This
effect on reprogramming was dependent on AID, as a reduction
of AID by four different siRNAs completely blocked pluripotency
gene promoter demethylation and gene expression. In somatic
cells, AID does not appear to act globally as in zebrafish embryos
and mouse PGCs (Popp et al., 2010; Rai et al., 2008), but instead
it is targeted to specific loci (Bhutani et al., 2010). However, if
specific loci are involved, the question arises as to how targeting
of this enzyme is mediated. Unlike TET, which has a DNAbinding motif (Ono et al., 2002), AID does not. Future studies to
decipher the mechanism by which AID is targeted will not only
provide insights into its sites of action but also illuminate its
role in active DNA demethylation in different cell types and in
response to different stimuli.
The BER Glycosylase Family: Mediators of DNA Repair
As described above, the accumulating data from the TET and
AID/APOBEC studies suggest that active demethylation involves
cytosine replacement via DNA repair. In principle, 5hmC or 5mC
can be removed passively in the course of cell division. However,
the rapid loss of methylation that occurs independent of DNA
replication (Bhutani et al., 2010; Frank et al., 1990; Oswald
et al., 2000; Paroush et al., 1990) must be mediated by an active
mechanism. In plants, active DNA demethylation is a well-characterized process in which the accepted mechanism involves
the BER pathway. BER glycosylases mediate the first step in
the repair pathway by removing the methylated cytosine and
creating an abasic site, which is then further acted upon by other
enzymes (Gehring et al., 2009). In mammals, the process is more
complex, as no glycosylases have been identified that act
directly on 5mC or 5hmC. Recently, an intermediate step, deamination, has been suggested to precede BER in mammalian DNA
demethylation (Cortellino et al., 2011; Guo et al., 2011). The
family of glycosylases implicated in the BER pathway are
members of the uracil DNA glycosylase (UDG) family that
includes thymine-DNA glycosylase (TDG) and single-strandselective monofunctional uracil-DNA glycosylase 1 (SMUG1)
(Cortellino et al., 2011; Guo et al., 2011). Therefore, like plants,
mammalian DNA demethylation involves DNA repair pathways.
The DNA glycosylases TDG and SMUG1 are capable of
converting 5hmU to cytosine, suggesting that they act in a partnership with TET and AID/APOBEC (Cortellino et al., 2011; Guo
et al., 2011). Notably, mice lacking TDG (TDG / ) are embryonic
lethal, underscoring the significant role that BER glycosylases
play during development and DNA demethylation. A direct
physical interaction has been demonstrated between AID and
TDG by coimmunoprecipitation experiments (Cortellino et al.,
2011). In addition, recent reports suggest that 5hmC can be
further oxidized by TET proteins to form 5-formylcytosine (5fC)
and 5-carboxylcytosine (5caC) (He et al., 2011; Ito et al., 2011).
These are unstable intermediates that can be detected in
ESCs, but they are 100-fold less abundant than 5hmC. Importantly, 5caC can be repaired by TDG, providing further support
for a DNA repair pathway (He et al., 2011). However, another
study also raises the possibility that a putative decarboxylase
could directly convert 5caC to cytosine independent of DNA
repair (Ito et al., 2011). The extent to which DNA repair pathways
are involved in the removal of 5caC and its relative importance to
DNA demethylation and gene expression remain to be determined.
In summary, the BER glycosylases, along with the TET and
AID/APOBEC families of enzymes, mediate DNA demethylation
via DNA repair. It remains to be tested whether other DNA repair
pathways besides BER participate in DNA demethylation.
Examples of Active DNA Demethylation in Mammals
As described below, a body of evidence is accumulating in this
nascent and rapidly evolving field, which supports the thesis
that active DNA demethylation is more common than previously
anticipated. Examples are presented below that indicate the role
of DNA demethylation in rapid responses to changes in extrinsic
signals, in early stages of development, and in highly specialized
postmitotic cells. This is merely the tip of the iceberg, and more
studies are needed to ascertain the extent to which the methylation editors are involved in the spatial and temporal regulation of
DNA demethylation.
Rapid Active DNA Demethylation in Response to Signals
Perhaps the most striking example of active DNA demethylation in adult cells to date is the activity-dependent DNA demethylation at the promoters of brain-derived neurotrophic factor
(BDNF) and fibroblast growth factor 1 (FGF1) in postmitotic
neurons (Martinowich et al., 2003). Recent studies have elegantly elucidated the molecular mechanisms underlying this
active DNA demethylation process and revealed a partnership
between the TET, AID/APOBEC enzymes, and the BER glycosylases (Guo et al., 2011). Reconstitution in HEK293 cells and
knockdown experiments demonstrate that TET-induced conversion of 5mC to 5hmC is followed by AID/APOBEC-mediated
Cell 146, September 16, 2011 2011 Elsevier Inc. 869
fused to form nondividing heterokaryons, active DNA demethylation was observed at the human MyoD promoter, which
accompanied its activation and expression in the fibroblasts.
Remarkably, the Cedar laboratory postulated more than 20
years ago that DNA demethylation occurs by an active mechanism in muscle cells (Paroush et al., 1990; Weiss et al., 1996),
but the factors responsible were unknown. These studies
suggest that a dynamic interplay of methylation and demethylation may also function during differentiation.
Is DNA Methylation-Demethylation Bidirectional?
Although indirect evidence has been accumulating for decades,
recent advances discussed here now support the hypothesis
that DNA demethylation and methylation may be bidirectional
and dynamically regulated throughout early and late development and in certain adult tissues, especially the brain (Guo
et al., 2011; Miller and Sweatt, 2007). Much remains to be
learned, including which loci are targeted for demethylation
and how the process is spatially and temporally regulated in
diverse cell types and stages of development. The long-held
notion that DNA methylation patterns are generally maintained
in stable differentiated states is likely true. Nonetheless, nuclear
reprogramming shows that perturbations are possible (Jullien
et al., 2011; Yamanaka and Blau, 2010). As is the case for the
regulation of gene expression by transcription factors (Blau
and Baltimore, 1991; Jacob and Monod, 1961; Ptashne, 2009;
Blau, 1992), the regulation of DNA methylation may also be
continuous and dictated by a balance of enzymes and targeting
factors.
As shown in Figure 1, our current understanding of the DNA
methylation and demethylation circuitry entails members of the
following enzyme families with roles in either passive or active
DNA demethylation: (1) the DNMT family of three methyltransferases responsible for the de novo generation and maintenance
of 5mC. DNA demethylation can occur passively by a dilution or
inactivation of DNMTs; (2) the TET family of three 5mC hydroxylases, which generate 5hmC (and further oxidized intermediates)
from 5mC; (3) the AID/APOBEC family of deaminases, which
initiate an active process of demethylation by deaminating either
5mC or 5hmC generated by the TET family; (4) the family of BER
glycosylases that initiate DNA repair culminating in the replacement of methylated cytosines with unmethylated cytosines. We
have designated these enzymes as the DNA methylation editors
that are responsible for the regulation of the DNA methylome
associated with a particular cell fate. It remains to be determined
whether active DNA demethylation in different scenarios always
requires a representative member from each of these families. In
other words, does the entire TETAID/APOBECBER pathway
operate broadly, or is only a subset thereof required to achieve
active DNA demethylation in different cell contexts.
The concept of a dynamic interplay of regulators has emerged
in parallel with the demonstration of the remarkable plasticity of
cellular fates illustrated by nuclear reprogramming. When the
balance of transcription factors that recognize DNA sequence
is perturbed by either nuclear transfer, cell fusion, or defined
factors (i.e., in generating iPSCs), it leads to a dramatic shift in
cell fate. A provocative, yet perhaps overly simplistic view of
how cell fate is controlled and maintained is provided by an
Delker, R.K., Fugmann, S.D., and Papavasiliou, F.N. (2009). Nat. Immunol. 10,
11471153.
Ficz, G., Branco, M.R., Seisenberger, S., Santos, F., Krueger, F., Hore, T.A.,
Marques, C.J., Andrews, S., and Reik, W. (2011). Nature 473, 398402.
Frank, D., Lichtenstein, M., Paroush, Z., Bergman, Y., Shani, M., Razin, A., and
Cedar, H. (1990). Philos. Trans. R. Soc. Lond. B Biol. Sci. 326, 241251.
Gehring, M., Reik, W., and Henikoff, S. (2009). Trends Genet. 25, 8290.
Globisch, D., Munzel, M., Muller, M., Michalakis, S., Wagner, M., Koch, S.,
Bruckl, T., Biel, M., and Carell, T. (2010). PLoS ONE 5, e15367.
Goll, M.G., and Bestor, T.H. (2005). Annu. Rev. Biochem. 74, 481514.
Guo, J.U., Su, Y., Zhong, C., Ming, G.L., and Song, H. (2011). Cell 145,
423434.
Hajkova, P., Jeffries, S.J., Lee, C., Miller, N., Jackson, S.P., and Surani, M.A.
(2010). Science 329, 7882.
He, Y.F., Li, B.Z., Li, Z., Liu, P., Wang, Y., Tang, Q., Ding, J., Jia, Y., Chen, Z., Li,
L., et al. (2011). Science. Published online August 4, 2011.
Huang, Y., Pastor, W.A., Shen, Y., Tahiliani, M., Liu, D.R., and Rao, A. (2010).
PLoS ONE 5, e8888.
Iqbal, K., Jin, S.G., Pfeifer, G.P., and Szabo, P.E. (2011). Proc. Natl. Acad. Sci.
USA 108, 36423647.
Ito, S., DAlessio, A.C., Taranova, O.V., Hong, K., Sowers, L.C., and Zhang, Y.
(2010). Nature 466, 11291133.
Ito, S., Shen, L., Dai, Q., Wu, S.C., Collins, L.B., Swenberg, J.A., He, C., and
Zhang, Y. (2011). Science. Published online July 21, 2011.
Jacob, F., and Monod, J. (1961). J. Mol. Biol. 3, 318356.
Jaenisch, R., and Bird, A. (2003). Nat. Genet. Suppl. 33, 245254.
Jullien, J., Pasque, V., Halley-Stott, R.P., Miyamoto, K., and Gurdon, J.B.
(2011). Nat. Rev. Mol. Cell Biol. 12, 453459.
Kangaspeska, S., Stride, B., Metivier, R., Polycarpou-Schwarz, M., Ibberson,
D., Carmouche, R.P., Benes, V., Gannon, F., and Reid, G. (2008). Nature 452,
112115.
Koh, K.P., Yabuuchi, A., Rao, S., Huang, Y., Cunniff, K., Nardone, J., Laiho, A.,
Tahiliani, M., Sommer, C.A., Mostoslavsky, G., et al. (2011). Cell Stem Cell 8,
200213.
Kriaucionis, S., and Heintz, N. (2009). Science 324, 929930.
Kucharski, R., Maleszka, J., Foret, S., and Maleszka, R. (2008). Science 319,
18271830.
Law, J.A., and Jacobsen, S.E. (2010). Nat. Rev. Genet. 11, 204220.
Liu, M., and Schatz, D.G. (2009). Trends Immunol. 30, 173181.
Liu, M., Duke, J.L., Richter, D.J., Vinuesa, C.G., Goodnow, C.C., Kleinstein,
S.H., and Schatz, D.G. (2008). Nature 451, 841845.
Martinowich, K., Hattori, D., Wu, H., Fouse, S., He, F., Hu, Y., Fan, G., and Sun,
Y.E. (2003). Science 302, 890893.
Maul, R.W., and Gearhart, P.J. (2010). Adv. Immunol. 105, 159191.
Mayer, W., Niveleau, A., Walter, J., Fundele, R., and Haaf, T. (2000). Nature
403, 501502.
Metivier, R., Gallais, R., Tiffoche, C., Le Peron, C., Jurkowska, R.Z.,
Carmouche, R.P., Ibberson, D., Barath, P., Demay, F., Reid, G., et al. (2008).
Nature 452, 4550.
Miller, C.A., and Sweatt, J.D. (2007). Neuron 53, 857869.
Monk, M., Adams, R.L., and Rinaldi, A. (1991). Development 112, 189192.
Muramatsu, M., Kinoshita, K., Fagarasan, S., Yamada, S., Shinkai, Y., and
Honjo, T. (2000). Cell 102, 553563.
Nakamura, T., Arai, Y., Umehara, H., Masuhara, M., Kimura, T., Taniguchi, H.,
Sekimoto, T., Ikawa, M., Yoneda, Y., Okabe, M., et al. (2007). Nat. Cell Biol. 9,
6471.
Ono, R., Taki, T., Taketani, T., Taniwaki, M., Kobayashi, H., and Hayashi, Y.
(2002). Cancer Res. 62, 40754080.
Oswald, J., Engemann, S., Lane, N., Mayer, W., Olek, A., Fundele, R., Dean,
W., Reik, W., and Walter, J. (2000). Curr. Biol. 10, 475478.
Paroush, Z., Keshet, I., Yisraeli, J., and Cedar, H. (1990). Cell 63, 12291237.
Weiss, A., Keshet, I., Razin, A., and Cedar, H. (1996). Cell 86, 709718.
Pastor, W.A., Pape, U.J., Huang, Y., Henderson, H.R., Lister, R., Ko, M.,
McLoughlin, E.M., Brudno, Y., Mahapatra, S., Kapranov, P., et al. (2011).
Nature 473, 394397.
Williams, K., Christensen, J., Pedersen, M.T., Johansen, J.V., Cloos, P.A.,
Rappsilber, J., and Helin, K. (2011). Nature 473, 343348.
Popp, C., Dean, W., Feng, S., Cokus, S.J., Andrews, S., Pellegrini, M.,
Jacobsen, S.E., and Reik, W. (2010). Nature 463, 11011105.
Wossidlo, M., Nakamura, T., Lepikhov, K., Marques, C.J., Zakhartchenko, V.,
Boiani, M., Arand, J., Nakano, T., Reik, W., and Walter, J. (2011). Nat Commun
2, 241.
Wu, H., DAlessio, A.C., Ito, S., Wang, Z., Cui, K., Zhao, K., Sun, Y.E., and
Zhang, Y. (2011a). Genes Dev. 25, 679684.
Rougier, N., Bourchis, D., Gomes, D.M., Niveleau, A., Plachot, M., Pa`ldi, A.,
and Viegas-Pequignot, E. (1998). Genes Dev. 12, 21082113.
Wu, H., DAlessio, A.C., Ito, S., Xia, K., Wang, Z., Cui, K., Zhao, K., Sun, Y.E.,
and Zhang, Y. (2011b). Nature 473, 389393.
Ruzov, A., Tsenkina, Y., Serio, A., Dudnakova, T., Fletcher, J., Bai, Y., Chebotareva, T., Pells, S., Hannoun, Z., Sullivan, G., et al. (2011). Cell Res. Published
online July 12, 2011. 10.1038/cr.2011.113.
Song, C.X., Szulwach, K.E., Fu, Y., Dai, Q., Yi, C., Li, X., Li, Y., Chen, C.H.,
Zhang, W., Jian, X., et al. (2011). Nat. Biotechnol. 29, 6872.
Tahiliani, M., Koh, K.P., Shen, Y., Pastor, W.A., Bandukwala, H., Brudno, Y.,
Agarwal, S., Iyer, L.M., Liu, D.R., Aravind, L., and Rao, A. (2009). Science
324, 930935.
Leading Edge
Review
Basic and Therapeutic
Aspects of Angiogenesis
Michael Potente,1,2 Holger Gerhardt,3,4 and Peter Carmeliet5,6,*
1Vascular
Blood vessels form extensive networks that nurture all tissues in the body. Abnormal vessel growth
and function are hallmarks of cancer and ischemic and inflammatory diseases, and they contribute
to disease progression. Therapeutic approaches to block vascular supply have reached the clinic,
but limited efficacy and resistance pose unresolved challenges. Recent insights establish how
endothelial cells communicate with each other and with their environment to form a branched
vascular network. The emerging principles of vascular growth provide exciting new perspectives,
the translation of which might overcome the current limitations of pro- and antiangiogenic medicine.
Introduction
Blood vessels supply oxygen and nutrients and provide gateways for immune surveillance. Endothelial cells (ECs) line the
inner surface of vessels to support tissue growth and repair. As
this network nourishes all tissues, it is not surprising that structural or functional vessel abnormalities contribute to many
diseases. Inadequate vessel maintenance or growth causes
ischemia in diseases such as myocardial infarction, stroke, and
neurodegenerative or obesity-associated disorders, whereas
excessive vascular growth or abnormal remodeling promotes
many ailments including cancer, inflammatory disorders, and
eye disease (Carmeliet, 2003; Folkman, 2007). Vessels are also
used as routes for tumor cells to metastasize.
Hallmarks of Vessel Growth
In the embryo, new vessels form de novo via the assembly of
mesoderm-derived endothelial precursors (angioblasts) that
differentiate into a primitive vascular labyrinth (vasculogenesis)
(Swift and Weinstein, 2009) (Figure 1A). Subsequent vessel
sprouting (angiogenesis) creates a network that remodels into
arteries and veins (Adams and Alitalo, 2007) (Figure 1A). Recruitment of pericytes and vascular smooth muscle cells that enwrap
nascent EC tubules provides stability and regulates perfusion
(arteriogenesis) (Jain, 2003). In the adult, vessels are quiescent
and rarely form new branches. However, ECs retain high plasticity to sense and respond to angiogenic signals.
The term angiogenesis is commonly used to reference the
process of vessel growth but in the strictest sense denotes
vessel sprouting from pre-existing ones. Recent studies provided tremendous insights into fundamental aspects of angiogenesis that have led to a mechanistic model of vessel branching
(Adams and Alitalo, 2007; Carmeliet and Jain, 2011; Eilken and
Adams, 2010; Phng and Gerhardt, 2009). Attracted by proangiogenic signals, ECs become motile and invasive and protrude
filopodia (Figure 1B). These so-called tip cells spearhead new
sprouts and probe the environment for guidance cues. Following
tip cells, stalk cells extend fewer filopodia but establish a lumen
and proliferate to support sprout elongation. Tip cells anastomose with cells from neighboring sprouts to build vessel loops.
The initiation of blood flow, the establishment of a basement
membrane, and the recruitment of mural cells stabilize new
connections (Figure 1C). The sprouting process iterates until
proangiogenic signals abate, and quiescence is re-established
(Figure 1C).
Although vessels can grow via other mechanisms, such as the
splitting of pre-existing vessels through intussusception or the
stimulation of vessel expansion by circulating precursor cells
(Fang and Salven, 2011; Makanya et al., 2009), we will focus
here on the latest insights on vessel sprouting, which likely
accounts for a substantial fraction of vessel growth.
Therapeutic Expectations and Challenges
The importance of angiogenesis sparked hopes that manipulating this process could offer therapeutic opportunities
(Folkman, 1971). Despite efforts to stimulate angiogenesis therapeutically by proangiogenic factors, most trials failed to meet
these expectations. Alternative strategies, based on proangiogenic cell therapies or targeting of microRNAs, offer new opportunities but are in (pre)clinical development (Bonauer et al.,
2010).
Antiangiogenic approaches aimed at blocking vessel growth in
eye disease and cancer led to the approval of therapeutics
targeting vascular endothelial growth factor (VEGF) (Crawford
and Ferrara, 2009b). Nonetheless, only a fraction of cancer
Cell 146, September 16, 2011 2011 Elsevier Inc. 873
patients show benefit as tumors evolve mechanisms of resistance or are refractory toward VEGF (receptor) inhibitors (Bergers
and Hanahan, 2008; Crawford and Ferrara, 2009a). Conflicting
results about the benefit of VEGF blockade have kick-started a
debate on whether antiangiogenic treatment may trigger more
invasive and metastatic tumors (Ebos and Kerbel, 2011). On the
upside, sustained normalization of abnormal tumor vessels
may offer benefit for combating metastasis (Goel et al., 2011).
For antiangiogenic medicine to have an enduring impact on
cancer patient survival, an integrated understanding of the
molecular principles of vessel growth is needed. Here, we take
a cell biological perspective to explore prototypic principles
and recently discovered regulatory mechanisms, seeking to
develop a framework of the angiogenic process that might
provide the basis for novel pro- and antiangiogenic therapies.
Endothelial Differentiation
Arterial and Venous Specification
Following assembly of primitive vessels in the early embryo (such
as the dorsal aorta and cardinal vein), remodeling transforms
the plexus into a hierarchically organized network of arteries,
capillaries, and veins. Arteries form a high-pressure system,
enabling transportation of blood to capillaries, whereas veins
face low-pressure gradients. The differences in hemodynamic
load are reflected in their structures: arteries are supported by
layers of vascular smooth muscle cells and a specialized matrix,
874 Cell 146, September 16, 2011 2011 Elsevier Inc.
whereas veins are thinner and surrounded by fewer smooth muscle cells
(Gaengel et al., 2009).
Arterial and venous ECs possess
specific molecular identities (Adams and
Alitalo, 2007; Swift and Weinstein, 2009).
For instance, Notch pathway components are highly expressed in arteries
but are low in veins. Disruption of
Notch signaling causes loss of arterial
markers and re-expression of venous
signature genes, suggesting that Notch
promotes arterial specification by repressing venous identity (Gridley, 2010;
Swift and Weinstein, 2009). Notch also
controls Eph-Ephrin family members,
which configure arterio-venous boundaries. Ephrin-B2 expression in arterial ECs increases in response
to Notch, whereas its receptor EphB4 in venous ECs is repressed by Notch. In zebrafish, Sonic Hedgehog acts upstream
of Notch, where it triggers arterial differentiation by upregulating
VEGF that elevates Notch components. In mice, VEGF secreted
by nerves contributes to arterial differentiation of ECs in cotracking vessels (Carmeliet and Tessier-Lavigne, 2005). Neuropilin-1
(NRP1), a VEGF coreceptor, facilitates transduction of arterial
effects of VEGF. At the level of gene expression, the transcription
factors FOXC1 and FOXC2 drive an arterial gene signature (e.g.,
DLL4, HEY2, CXCR4) by interacting with VEGF and Notch
signaling. Although earlier proposals favored the view that the
venous fate is acquired by default, it has become clear that
venous identity requires repression of Notch signaling by the
vein-specific nuclear receptor COUP-TFII (Swift and Weinstein,
2009). In addition, hemodynamic factors such as blood pressure
and flow codetermine arterio-venous differentiation (Jones et al.,
2006).
Arterio-Venous Segregation
Zebrafish studies indicate that the cardinal vein does not form
by vasculogenesis but instead arises from a common precursor
vessel by segregation (Herbert et al., 2009) (Figure 1A). Venousfated EphB4-positive ECs migrate away from the arterial-fated
ephrinB2-positve ECs in the precursor vessel toward the location of the future cardinal vein. VEGF and Notch both restrain
ventral sprouting, whereas VEGF-C promotes segregation.
junction molecules mediate cell-cell adhesion, cytoskeletal reorganization, and intracellular signaling. VE-cadherin is a key
component of EC junctions. In complex with VEGFR2, VE-cadherin maintains EC quiescence through recruitment of phosphatases that dephosphorylate VEGFR2, thus restraining VEGF
signaling. Distinct types of VE-cadherin-based adherens junctions establish stable or transitory interactions with the cytoskeleton that either solidify EC adhesion and barrier properties
or facilitate EC separation and movement (Falk, 2010). Activation of TIE2 by ANG1 protects vessels from VEGF-induced
leakage by inhibiting VEGFs ability to induce endocytosis of
VE-cadherin.
Vessels Express Survival Signals
As endothelial proliferation decelerates during maturation, ECs
must adopt survival properties to maintain integrity of the vessel
lining. Autocrine and paracrine survival signals from endothelial
and support cells protect the vessel from environmental
stresses. One such survival factor is VEGF, which activates the
PI3K/AKT survival pathway. Interestingly, ECs themselves are
the pivotal source for VEGFs prosurvival activity. Mice lacking
VEGF in ECs suffer bleeding, microinfarcts, and EC rupture
(Warren and Iruela-Arispe, 2010). When produced by ECs as
intracrine factor, VEGF prevents EC apoptosis in nonpathological conditions (Figure 5B). This intracrine activity of VEGF
differs from its paracrine function in stimulating angiogenesis,
as loss of endothelial VEGF does not cause developmental
vascular defects (Warren and Iruela-Arispe, 2010).
Signaling by fibroblast growth factors (FGFs) has also been
implicated in maintaining vascular integrity due to their ability
to anneal adherens junctions (Beenken and Mohammadi,
2009). Inhibition of FGF signaling results in dissociation of
adherens junctions and tight junctions, subsequent loss of
ECs, and vessel disintegration (Murakami et al., 2008). Notch
signaling is critical for generating and maintaining vascular
homeostasis. A consequence of Notch activation is the establishment of mature and patent vessels that promote perfusion
and relieve tissue hypoxia. Conversely, blockade of DLL4 or
Notch1 in the adult causes vascular tumors and hemorrhage
(Liu et al., 2011; Yan et al., 2010). Similarly, endothelial inactivation of RBPj reinitiates vascular growth in adulthood (Figure 5B).
Activation of Notch in mural cells by endothelial DLL4 also
contributes to vessel stability by stimulating deposition of BM
components.
Signaling by TIE2 and ANG1 also controls survival and vessel
quiescence (Augustin et al., 2009). ANG1 clusters TIE2 junctionally at inter-EC junctions in trans to promote survival and EC
quiescence (Figure 5B). Blood flow is another important survival
cue for ECs as fluid shear stress potently inhibits EC apoptosis.
KLF2 is activated by shear stress and evokes quiescence by
upregulating endothelial nitric oxide synthase and the anticoagulant factor thrombomodulin, keeping vessels dilated, perfused,
and free of clots, and by downregulating VEGFR2, which prevents tip cell formation (Figure 4B). Other EC quiescence factors
include bone morphogenic protein 9 (BMP9) and cerebral
cavernous malformation proteins (CCM13), whose defective
signaling causes vascular malformations (Leblanc et al., 2009).
ECs in nonperfused vessels regress from their locations or
undergo apoptosis (Figure 4B).
880 Cell 146, September 16, 2011 2011 Elsevier Inc.
promotes tumorigenesis and ocular disorders such as agerelated macular degeneration. Historically, this has led to
concepts of pro- and antiangiogenic therapy, aiming to restore
adequate vessel densities. However, sprouting angiogenesis
alone might be insufficient to fully revascularize ischemic tissues,
as also collateral vessels have to enlarge to supply bulk flow
(Schaper, 2009). It has become clear that vessel densities can
no longer be considered separately from vessel function when
designing angiogenic therapeutics. We anticipate that insights
into pathological angiogenesis, guiding future diagnostic and
therapeutic approaches, will increasingly focus on the functional
quality of vessels and their effects on local metabolism rather
than on vessel quantity alone.
Tumor Vessels Are Abnormal
Tumor vessels display abnormal structure and function (Goel
et al., 2011; Jain, 2005) with seemingly chaotic organization
(Figure 7A). Highly dense regions neighbor vessel-poor areas,
and vessels vary from abnormally wide, irregular, and tortuous
serpentine-like shape to thin channels with small or compressed
lumens. Every layer of the tumor vessel wall is abnormal. ECs
lack a cobblestone appearance, are poorly interconnected,
function, so that improved perfusion and oxygenation counteract the hypoxia-driven expression of genes controlling epithelial-mesenchymal transition, invasion, and intravasation, which
prompt the metastatic switch (Goel et al., 2011; Mazzone
et al., 2009; Rolny et al., 2011) (Figure 7B). The normalized vessel
wall also restricts tumor cell intravasation (Mazzone et al., 2009),
while responses to chemo- or immunotherapy can be improved
(Goel et al., 2011; Rolny et al., 2011).
Conclusions and Perspectives
Despite progress in understanding the molecular basis of angiogenesis, and successful translation of VEGF blockade for the
treatment of age-related macular degeneration and some cancer
patients, challenges must be overcome to improve the overall
efficacy of antivascular strategies to combat cancer more efficiently. A question of high priority is whether the approved
antiangiogenic regimes are optimally used in terms of dosing,
duration, and combination therapy. The role of VEGF (receptor)
inhibitors in micrometastatic disease in adjuvant settings (e.g.,
upon resection of the primary tumor) will require further research
given the paucity of available preclinical data and suitable animal
models. Another priority is to identify predictive biomarkers,
tailored for particular tumors, stages, and treatment. Third, development of additional antiangiogenic drugs, independent of VEGF
signaling, and evaluation of their potential in clinical trials, in
particular as combination therapy with current VEGF (receptor)
inhibitors, is likely to expand the antiangiogenic armamentarium.
Fourth, the therapeutic potential of sustained vessel normalization to suppress metastasis and enhance chemotherapy will
need to be evaluated clinically, and additional studies are
required to establish how it could be combined best with available vessel pruning therapies. Also, antivascular approaches
could be beneficial for the treatment of nonsolid malignancies
(e.g., leukemias) or for the treatment of children or pregnant
women with cancer or individuals with inflammatory disorders
(e.g., arthritis) who have not been considered eligible for VEGF
blockade because of side effects. Finally, the recent molecular
breakthroughs in our understanding of vessel growth should
kindle renewed interest in developing strategies to revascularize
ischemic tissues.
ACKNOWLEDGMENTS
We acknowledge L. Notebaert and A. Truyens for help with the illustrations. We
apologize to authors whose work we could not cite because of the limit on the
number of references and therefore mostly cited overview articles. The work of
P.C. is supported by a Federal Government Belgium grant (IUAP06/30), longterm structural Methusalem funding by the Flemish Government, a Concerted
Research Activities Belgium grant (GOA2006/11), a grant from The Research
FoundationFlanders (FWO G.0673.08), and the Foundation Leducq Transatlantic Network ARTEMIS. The research of M.P. is supported by grants from the
Deutsche Forschungsgemeinschaft (DFG - SFB 834/A6 and Exc 147/1). The
research of H.G. is supported by Cancer Research UK, the Lister Institute of
Preventive Medicine, the EMBO Young Investigator Programme, and the
Foundation Leducq Transatlantic Network ARTEMIS.
REFERENCES
Adams, R.H., and Alitalo, K. (2007). Molecular regulation of angiogenesis and
lymphangiogenesis. Nat. Rev. Mol. Cell Biol. 8, 464478.
Ebos, J.M.L., and Kerbel, R.S. (2011). Antiangiogenic therapy: impact on invasion, disease progression, and metastasis. Nat. Rev. Clin. Oncol. 8, 210221.
Eilken, H.M., and Adams, R.H. (2010). Dynamics of endothelial cell behavior in
sprouting angiogenesis. Curr. Opin. Cell Biol. 22, 617625.
Augustin, H.G., Koh, G.Y., Thurston, G., and Alitalo, K. (2009). Control of
vascular morphogenesis and homeostasis through the angiopoietin-Tie
system. Nat. Rev. Mol. Cell Biol. 10, 165177.
Erez, N., Truitt, M., Olson, P., Arron, S.T., and Hanahan, D. (2010). Cancerassociated fibroblasts are activated in incipient neoplasia to orchestrate
tumor-promoting inflammation in an NF-kappaB-dependent manner. Cancer
Cell 17, 135147.
Bagri, A., Kouros-Mehr, H., Leong, K.G., and Plowman, G.D. (2010). Use of
anti-VEGF adjuvant therapy in cancer: challenges and rationale. Trends Mol.
Med. 16, 122132.
Beenken, A., and Mohammadi, M. (2009). The FGF family: biology, pathophysiology and therapy. Nat. Rev. Drug Discov. 8, 235253.
Bentley, K., Mariggi, G., Gerhardt, H., and Bates, P.A. (2009). Tipping the
balance: robustness of tip cell selection, migration and fusion in angiogenesis.
PLoS Comput. Biol. 5, e1000549.
Falk, M.M. (2010). Adherens junctions remain dynamic. BMC Biol. 8, 34.
Fang, S., and Salven, P. (2011). Stem cells in tumor angiogenesis. J. Mol. Cell.
Cardiol. 50, 290295.
Fantin, A., Maden, C.H., and Ruhrberg, C. (2009). Neuropilin ligands in
vascular and neuronal patterning. Biochem. Soc. Trans. 37, 12281232.
Fantin, A., Vieira, J.M., Gestri, G., Denti, L., Schwarz, Q., Prykhozhij, S., Peri,
F., Wilson, S.W., and Ruhrberg, C. (2010). Tissue macrophages act as cellular
chaperones for vascular anastomosis downstream of VEGF-mediated endothelial tip cell induction. Blood 116, 829840.
Ferrara, N. (2010a). Pathways mediating VEGF-independent tumor angiogenesis. Cytokine Growth Factor Rev. 21, 2126.
Bonauer, A., Boon, R.A., and Dimmeler, S. (2010). Vascular microRNAs. Curr.
Drug Targets 11, 943949.
Ferrara, N. (2010b). Role of myeloid cells in vascular endothelial growth factorindependent tumor angiogenesis. Curr. Opin. Hematol. 17, 219224.
Butler, J.M., Kobayashi, H., and Rafii, S. (2010). Instructive role of the vascular
niche in promoting tumour growth and tissue repair by angiocrine factors. Nat.
Rev. Cancer 10, 138146.
Fischer, C., Mazzone, M., Jonckx, B., and Carmeliet, P. (2008). FLT1 and its
ligands VEGFB and PlGF: drug targets for anti-angiogenic therapy? Nat.
Rev. Cancer 8, 942956.
Carmeliet, P., and Jain, R.K. (2011). Molecular mechanisms and clinical applications of angiogenesis. Nature 473, 298307.
Fraisl, P., Mazzone, M., Schmidt, T., and Carmeliet, P. (2009). Regulation of
angiogenesis by oxygen and metabolism. Dev. Cell 16, 167179.
Francia, G., Cruz-Munoz, W., Man, S., Xu, P., and Kerbel, R.S. (2011). Mouse
models of advanced spontaneous metastasis for experimental therapeutics.
Nat. Rev. Cancer 11, 135141.
Cavallaro, U., and Dejana, E. (2011). Adhesion molecule signalling: not always
a sticky business. Nat. Rev. Mol. Cell Biol. 12, 189197.
Chappell, J.C., and Bautch, V.L. (2010). Vascular development: genetic mechanisms and links to vascular disease. Curr. Top. Dev. Biol. 90, 4372.
Corada, M., Nyqvist, D., Orsenigo, F., Caprini, A., Giampietro, C., Taketo,
M.M., Iruela-Arispe, M.L., Adams, R.H., and Dejana, E. (2010). The Wnt/
beta-catenin pathway modulates vascular remodeling and specification by
upregulating Dll4/Notch signaling. Dev. Cell 18, 938949.
Crawford, Y., and Ferrara, N. (2009a). Tumor and stromal pathways mediating
refractoriness/resistance to anti-angiogenic therapies. Trends Pharmacol. Sci.
30, 624630.
Goel, S., Duda, D.G., Xu, L., Munn, L.L., Boucher, Y., Fukumura, D., and Jain,
R.K. (2011). Normalization of the vasculature for treatment of cancer and other
diseases. Physiol. Rev. 91, 10711121.
Crawford, Y., and Ferrara, N. (2009b). VEGF inhibition: insights from preclinical
and clinical studies. Cell Tissue Res. 335, 261269.
Gridley, T. (2010). Notch signaling in the vasculature. Curr. Top. Dev. Biol. 92,
277309.
De Smet, F., Segura, I., De Bock, K., Hohensinner, P.J., and Carmeliet, P.
(2009). Mechanisms of vessel branching: filopodia on endothelial tip cells
lead the way. Arterioscler. Thromb. Vasc. Biol. 29, 639649.
Desgrosellier, J.S., and Cheresh, D.A. (2010). Integrins in cancer: biological
implications and therapeutic opportunities. Nat. Rev. Cancer 10, 922.
Ding, B.S., Nolan, D.J., Butler, J.M., James, D., Babazadeh, A.O., Rosenwaks,
Z., Mittal, V., Kobayashi, H., Shido, K., Lyden, D., et al. (2010). Inductive angiocrine signals from sinusoidal endothelium are required for liver regeneration.
Nature 468, 310315.
Guarani, V., Deflorian, G., Franco, C.A., Kruger, M., Phng, L.K., Bentley, K.,
Toussaint, L., Dequiedt, F., Mostoslavsky, R., Schmidt, M.H., et al. (2011).
Acetylation-dependent regulation of endothelial Notch signalling by the
SIRT1 deacetylase. Nature 473, 234238.
Hagberg, C.E., Falkevall, A., Wang, X., Larsson, E., Huusko, J., Nilsson, I., van
Meeteren, L.A., Samen, E., Lu, L., Vanwildemeersch, M., et al. (2010). Vascular
endothelial growth factor B controls endothelial fatty acid uptake. Nature 464,
917921.
Herbert, S.P., Huisken, J., Kim, T.N., Feldman, M.E., Houseman, B.T., Wang,
R.A., Shokat, K.M., and Stainier, D.Y. (2009). Arterial-venous segregation by
selective cell sprouting: an alternative mode of blood vessel formation.
Science 326, 294298.
Eble, J.A., and Niland, S. (2009). The extracellular matrix of blood vessels.
Curr. Pharm. Des. 15, 13851400.
Huang, H., Bhat, A., Woodnutt, G., and Lappe, R. (2010). Targeting the
ANGPT-TIE2 pathway in malignancy. Nat. Rev. Cancer 10, 575585.
Iruela-Arispe, M.L., and Davis, G.E. (2009). Cellular and molecular mechanisms of vascular lumen formation. Dev. Cell 16, 222231.
Jain, R.K. (2003). Molecular regulation of vessel maturation. Nat. Med. 9,
685693.
Jain, R.K. (2005). Normalization of tumor vasculature: an emerging concept in
antiangiogenic therapy. Science 307, 5862.
Jain, R.K., Duda, D.G., Willett, C.G., Sahani, D.V., Zhu, A.X., Loeffler, J.S.,
Batchelor, T.T., and Sorensen, A.G. (2009). Biomarkers of response and resistance to antiangiogenic therapy. Nat. Rev. Clin. Oncol. 6, 327338.
Nisancioglu, M.H., Betsholtz, C., and Genove, G. (2010). The absence of pericytes does not increase the sensitivity of tumor vasculature to vascular endothelial growth factor-A blockade. Cancer Res. 70, 51095115.
Norden, A.D., Drappatz, J., and Wen, P.Y. (2009). Antiangiogenic therapies for
high-grade glioma. Nat. Rev. Neurol. 5, 610620.
Nyberg, P., Xie, L., and Kalluri, R. (2005). Endogenous inhibitors of angiogenesis. Cancer Res. 65, 39673979.
Nyberg, P., Salo, T., and Kalluri, R. (2008). Tumor microenvironment and
angiogenesis. Front. Biosci. 13, 65376553.
Jakobsson, L., Franco, C.A., Bentley, K., Collins, R.T., Ponsioen, B., Aspalter,
I.M., Rosewell, I., Busse, M., Thurston, G., Medvinsky, A., et al. (2010). Endothelial cells dynamically compete for the tip cell position during angiogenic
sprouting. Nat. Cell Biol. 12, 943953.
Padera, T.P., Kuo, A.H., Hoshida, T., Liao, S., Lobo, J., Kozak, K.R., Fukumura,
D., and Jain, R.K. (2008). Differential response of primary tumor versus
lymphatic metastasis to VEGFR-2 and VEGFR-3 kinase inhibitors cediranib
and vandetanib. Mol. Cancer Ther. 7, 22722279.
Jeansson, M., Gawlik, A., Anderson, G., Li, C., Kerjaschki, D., Henkelman, M.,
and Quaggin, S.E. (2011). Angiopoietin-1 is essential in mouse vasculature
during development and in response to injury. J. Clin. Invest. 121, 22782289.
Pardali, E., Goumans, M.J., and ten Dijke, P. (2010). Signaling by members of
the TGF-beta family in vascular morphogenesis and disease. Trends Cell Biol.
20, 556567.
Jones, E.A., le Noble, F., and Eichmann, A. (2006). What determines blood
vessel structure? Genetic prespecification vs. hemodynamics. Physiology
(Bethesda) 21, 388395.
Phng, L.K., Potente, M., Leslie, J.D., Babbage, J., Nyqvist, D., Lobov, I., Ondr,
J.K., Rao, S., Lang, R.A., Thurston, G., and Gerhardt, H. (2009). Nrarp coordinates endothelial Notch and Wnt signaling to control vessel density in angiogenesis. Dev. Cell 16, 7082.
Pietras, K., and Ostman, A. (2010). Hallmarks of cancer: interactions with the
tumor stroma. Exp. Cell Res. 316, 13241331.
Pitulescu, M.E., and Adams, R.H. (2010). Eph/ephrin moleculesa hub for
signaling and endocytosis. Genes Dev. 24, 24802492.
Qian, B.Z., and Pollard, J.W. (2010). Macrophage diversity enhances tumor
progression and metastasis. Cell 141, 3951.
Quaegebeur, A., Segura, I., and Carmeliet, P. (2010). Pericytes: blood-brain
barrier safeguards against neurodegeneration? Neuron 68, 321323.
Rolny, C., Mazzone, M., Tugues, S., Laoui, D., Johansson, I., Coulon, C., Squadrito, M.L., Segura, I., Li, X., Knevels, E., et al. (2011). HRG inhibits tumor
growth and metastasis by inducing macrophage polarization and vessel
normalization through downregulation of PlGF. Cancer Cell 19, 3144.
Roukens, M.G., Alloul-Ramdhani, M., Baan, B., Kobayashi, K., PetersonMaduro, J., van Dam, H., Schulte-Merker, S., and Baker, D.A. (2010). Control
of endothelial sprouting by a Tel-CtBP complex. Nat. Cell Biol. 12, 933942.
Sawamiphak, S., Seidel, S., Essmann, C.L., Wilkinson, G.A., Pitulescu, M.E.,
Acker, T., and Acker-Palmer, A. (2010). Ephrin-B2 regulates VEGFR2 function
in developmental and tumour angiogenesis. Nature 465, 487491.
Mazzieri, R., Pucci, F., Moi, D., Zonari, E., Ranghetti, A., Berti, A., Politi, L.S.,
Gentner, B., Brown, J.L., Naldini, L., and De Palma, M. (2011). Targeting the
ANG2/TIE2 axis inhibits tumor growth and metastasis by impairing angiogenesis and disabling rebounds of proangiogenic myeloid cells. Cancer Cell 19,
512526.
Mazzone, M., Dettori, D., Leite de Oliveira, R., Loges, S., Schmidt, T., Jonckx,
B., Tian, Y.M., Lanahan, A.A., Pollard, P., Ruiz de Almodovar, C., et al. (2009).
Heterozygous deficiency of PHD2 restores tumor oxygenation and inhibits
metastasis via endothelial normalization. Cell 136, 839851.
Miles, D., Harbeck, N., Escudier, B., Hurwitz, H., Saltz, L., Van Cutsem, E.,
Cassidy, J., Mueller, B., and Sirzen, F. (2011). Disease course patterns after
discontinuation of bevacizumab: pooled analysis of randomized phase III
trials. J. Clin. Oncol. 29, 8388.
Schmidt, T., Kharabi Masouleh, B., Loges, S., Cauwenberghs, S., Fraisl, P.,
Maes, C., Jonckx, B., De Keersmaecker, K., Kleppe, M., Tjwa, M., et al.
(2011). Loss or inhibition of stromal-derived PlGF prolongs survival of mice
with imatinib-resistant Bcr-Abl1(+) leukemia. Cancer Cell 19, 740753.
Serini, G., Maione, F., Giraudo, E., and Bussolino, F. (2009). Semaphorins and
tumor angiogenesis. Angiogenesis 12, 187193.
, B., Kucera, T., Eglinger, J., Hughes, M.R., McNagny, K.M., Tsukita, S.,
Strilic
Dejana, E., Ferrara, N., and Lammert, E. (2009). The molecular basis of
vascular lumen formation in the developing mouse aorta. Dev. Cell 17,
505515.
Swift, M.R., and Weinstein, B.M. (2009). Arterial-venous specification during
development. Circ. Res. 104, 576588.
Murakami, M., Nguyen, L.T., Zhuang, Z.W., Moodie, K.L., Carmeliet, P., Stan,
R.V., and Simons, M. (2008). The FGF system has a key role in regulating
vascular integrity. J. Clin. Invest. 118, 33553366.
Tammela, T., Zarkada, G., Wallgard, E., Murtomaki, A., Suchting, S., Wirzenius, M., Waltari, M., Hellstrom, M., Schomber, T., Peltonen, R., et al. (2008).
Blocking VEGFR-3 suppresses angiogenic sprouting and vascular network
formation. Nature 454, 656660.
Nicoli, S., Standley, C., Walker, P., Hurlstone, A., Fogarty, K.E., and Lawson,
N.D. (2010). MicroRNA-mediated integration of haemodynamics and Vegf signalling during angiogenesis. Nature 464, 11961200.
Thurston, G., Noguera-Troise, I., and Yancopoulos, G.D. (2007). The Delta
paradox: DLL4 blockade leads to more tumour vessels but less tumour
growth. Nat. Rev. Cancer 7, 327331.
Van Cutsem, E., Lambrechts, D., Prenen, H., Jain, R.K., and Carmeliet, P.
(2011). Lessons from the adjuvant bevacizumab trial on colon cancer: what
next? J. Clin. Oncol. 29, 14.
Wang, Y., Nakayama, M., Pitulescu, M.E., Schmidt, T.S., Bochenek, M.L.,
Sakakibara, A., Adams, S., Davy, A., Deutsch, U., Luthi, U., et al. (2010).
Ephrin-B2 controls VEGF-induced angiogenesis and lymphangiogenesis.
Nature 465, 483486.
Warren, C.M., and Iruela-Arispe, M.L. (2010). Signaling circuitry in vascular
morphogenesis. Curr. Opin. Hematol. 17, 213218.
Wels, J., Kaplan, R.N., Rafii, S., and Lyden, D. (2008). Migratory neighbors and
distant invaders: tumor-associated niche cells. Genes Dev. 22, 559574.
Xu, K., Sacharidou, A., Fu, S., Chong, D.C., Skaug, B., Chen, Z.J., Davis, G.E.,
and Cleaver, O. (2011). Blood vessel tubulogenesis requires Rasip1 regulation
of GTPase signaling. Dev. Cell 20, 526539.
Yan, M., Callahan, C.A., Beyer, J.C., Allamneni, K.P., Zhang, G., Ridgway, J.B.,
Niessen, K., and Plowman, G.D. (2010). Chronic DLL4 blockade induces
vascular neoplasms. Nature 463, E6E7.
Zeeb, M., Strilic, B., and Lammert, E. (2010). Resolving cell-cell junctions:
lumen formation in blood vessels. Curr. Opin. Cell Biol. 22, 626632.
SUMMARY
Complex genomic rearrangements (CGRs) consisting of two or more breakpoint junctions have been
observed in genomic disorders. Recently, a chromosome catastrophe phenomenon termed chromothripsis, in which numerous genomic rearrangements
are apparently acquired in one single catastrophic
event, was described in multiple cancers. Here, we
show that constitutionally acquired CGRs share similarities with cancer chromothripsis. In the 17 CGR
cases investigated, we observed localization and
multiple copy number changes including deletions,
duplications, and/or triplications, as well as extensive
translocations and inversions. Genomic rearrangements involved varied in size and complexities; in
one case, array comparative genomic hybridization
revealed 18 copy number changes. Breakpoint sequencing identified characteristic features, including
small templated insertions at breakpoints and microhomology at breakpoint junctions, which have been
attributed to replicative processes. The resemblance
between CGR and chromothripsis suggests similar
mechanistic underpinnings. Such chromosome catastrophic events appear to reflect basic DNA metabolism operative throughout an organisms life cycle.
INTRODUCTION
Human genomic rearrangements with two or more breakpoint
junctions are referred to as complex genomic rearrangements
(CGRs) (Zhang et al., 2009a). CGRs have been identified
frequently during characterization of nonrecurrent microduplications associated with genomic disorders. Based on the microhomologies identified at breakpoint junctions, the apparent template driven insertional complexities at breakpoints, and the
fusions of distantly distributed sequences in complex genomic
rearrangements, a replication based mechanism, the fork stalling and template switching (FoSTeS) model, has been proposed
to explain the formation of such rearrangement complexities
in the human genome (Lee et al., 2007; Slack et al., 2006).
Other similar replication based models such as microhomology
mediated break-induced replication (MMBIR) (Hastings et al.,
2009a, 2009b), microhomology/microsatellite induced replication (MMIR) (Payen et al., 2008), and microhomology-mediated
replication-dependent recombination (MMRDR) (Chen et al.,
2010) have also been proposed. Recent studies on genomic
Cell 146, 889903, September 16, 2011 2011 Elsevier Inc. 889
Table 1. Complex Genomic Rearrangements Assessed by Gains and Losses of Specific Human Genome Intervals
Subjects
(BAB#)
Sex Location
Array CGH-Inferred
Rearrangement
Sizes of Copy
Pattern
Number Changes
Size (Mb)
Clinical Indication
Patient 1 F
1q32.3-q43
del-nml-dup-nmldup-nml-del
Severe DD/ID
de novo
34.0 14.0
7.5
6.5
2920
1q43-q44
dup-del
Microcephaly,
epilepsy
mother nml
4.7
0.7
4.0
mat
4.7
3047
4q33-q34.1
trp
trp 1.2 Mb
DD
1.2
1.2
1.2
2760
4q35.1-q35.2
dup-nml-dup
Autistic spectrum, NA
DD/ID, prenatal
teratogen exposure
3.0
0.4
0.4
3012
5q31.1-q31.2
dup-trp
Mild DD, DF
mat (47%)
1.9
1.9
1.9
2778
6q27
dup-nml-terminal
dup, r(6)
Chromosomal
abnormality,
microcephaly,
speech delay
father nml
3.9
3.2
3.2
3103
7q33-ter
dup-nml-del
DF, VSD
de novo
25.9 25.9
25.6 0.3
3105
9q31.1-q33.1
dup-nml-dup-nmldup-nml-dup-nmldup-nml-dup-nmldup-nml-dup-nmldup-trp-dup-nmldup-nml-trp-dupnml-dup-nml-dupnml-dup-nml-dup
DF
de novo
51.0 11.6
11.6 0
3015
12q21.1
dup-trp
DD, DF
mat
2.3
2780
12q24
dup-nml-dup
DD
de novo
11.1 10.9
10.9 0
2785
13q34
dup-nml-dup
Moderate DD/ID,
DF
mat
1.1
1.0
3011
14q12-q21.3
dup-nml-del
DD, DF
de novo
19.9 18.2
2.8
15.4
3050
15q26.3
dup-trp-duptrp-dup
de novo
2.5
2.5
2.5
2783
18p11.32
dup-nml-dup
ID
mother nml
2.4
1.0
1.0
0.0
Chromosomal
abnormality,
epilepsy
NA
35.1 14.5
Patient 2 F
22q11.1-q13.33 del-nml-dup-nmldup-nml-delnml-dup
890 Cell 146, 889903, September 16, 2011 2011 Elsevier Inc.
2.2
1.0
2.2
12.1 2.4
Table 1. Continued
Subjects
(BAB#)
Sex Location
Array CGH-Inferred
Rearrangement
Sizes of Copy
Pattern
Number Changes
Size (Mb)
Clinical Indication
3104
NA
21.8 9.4
5.1
4.3
3032
Xq27.1
mother nml
0.7
0.7
0.0
dup-trp-duptrp-dup
0.7
Abbreviations: dup-, duplication; trp-, triplication; del-, deletion; nml-, normal; DD-, developmental delay; ID-, intellectual disability; DF-, dysmorphic
features; VSD-, ventricular septal defect; MCA-, multiple congenital anomalies; NA-, not available; mat-, maternal. See also Table S1 for detailed clinical features and Figure S5 for descriptions of low-copy repeats at breakpoints.
892 Cell 146, 889903, September 16, 2011 2011 Elsevier Inc.
(C and D) G-banded chromosome (C) and FISH (D) analyses using two probes within the two duplications show that neither of the two individual duplicated
segments appears to be simple tandem events; subsequent sequence analysis demonstrated that one duplication in BAB2780 is inserted in between the other
duplication.
(E) Interphase FISH analysis revealed three red signals, which confirmed the triplication in BAB3050. The BAC probes used were indicated.
See also Figure S1 for additional rearrangements with CMA results indicating a duplication next to a terminal deletion.
Cell 146, 889903, September 16, 2011 2011 Elsevier Inc. 893
A
BAB 3012
BAB 3032
dup
dup -
trp
BAB 2783
BAB 3015
dup
trp
dup
nml
dup
BAB 2785
BAB 3103
dup
dup
nml-del
nml
dup
BAB 2778
nml-dup-trp
BAB 3012
nml-dup-trp
BAB 3015
dup-nml-del
BAB 3103
dup
nml
dup
B
BAB 2780
dup-nml
BAB 3011
dup
dup-nml
BAB 2920
dup
del
BAB 3050
del
dup
trp
896 Cell 146, 889903, September 16, 2011 2011 Elsevier Inc.
with more severe phenotypes (Table 1). Indeed, of the ten cases
with combined copy number changes larger than 2.5 Mb, six
rearrangements were de novo events and the other four had
either no or incomplete parental studies. These de novo cases
were associated with severe DD/intellectual disability, epilepsy,
and dysmorphic features. In contrast, in the seven cases with
relatively smaller copy number changes (0.4 Mb2.2 Mb), four
rearrangements were inherited and the other three had incomplete parental information. Interestingly, in each of these
CGRs, only copy number gains were present, and most of these
patients showed a comparatively milder phenotype such as
moderate DD. It is unclear at the present time whether these relatively smaller complex rearrangements are clinically significant.
The fact that four rearrangements were maternally inherited,
but none were paternally inherited is consistent with the previous
observations in complex chromosomal rearrangements that
familial transmission is mainly observed through female carriers
(Batista et al., 1994).
As noted for cancer chromothripsis, such complex genomic
rearrangement can have multigenic consequences wherein
many genes have an altered dosage or copy number, others
are potentially disrupted, and there may be novel gene fusion
formed at the multiple breakpoints. Perhaps the observed
common clinical phenotype of neurodevelopmental delay
reflects multiple potential gene disruptions or dosage alterations
by CGR and the large number of genes in the human genome
that contribute to neurological function.
DISCUSSION
Do Somatic Chromothripsis Events Occur
in a Mechanistically Similar Manner
to Constitutional CGR?
Recently, the phenomenon of chromosome shattering termed
chromothripsis, i.e., a massively complex genomic rearrangement which occurs in a single catastrophic event involving local
apparent shattering of a chromosome and subsequent reassembly proposed to be potentially mediated by nonhomologous end
joining, was described in 2%3% of all cancers analyzed (Stephens et al., 2011). This phenomenon was also reported in the
germline (Borg et al., 2005; Kloosterman et al., 2011). The attractively simple idea that chromothripsis stems from the shattering
the copy number states for each genomic segment. The sizes of the rearrangements and normal copy number intervals are listed in megabases. The arrows
indicate translocations (upward facing) and insertions (downward facing) as indicated by FISH analysis with locus-specific BAC clones. All these additional
segments of the 5.1 Mb duplication, 2.1 Mb duplication, 0.78 Mb duplication, 1.4 Mb duplication, and 5.5 Mb triplication were translocated close to the pericentric
region, proximal to the original location of the 2.1 Mb duplication region. The orders of these additional segments are deduced from FISH results. It is unknown
whether the additional 1.4 Mb duplicated segment is located proximal to or between the two copies of the 5.5 Mb triplicate segments. Two representative FISH
images are shown. The FISH image on the left shows that the duplicated (RP11-35N6) and triplicated (RP11-18B16) segments in 9q31-q33 were translocated
close to the 9q21 region. The FISH image on the right shows that the two additional copies of the 5.5 Mb triplication marked by an arrow are in an inverted
orientation with each other. RP11-203L12 is mapped at distal end and RP11-104M22 is mapped at the proximal end of the triplicated segment. Note that the
additional complexities of the rearranged chromosome are represented as curvilinear connections of breakpoint regions. The purple curves indicate junctions
between two segments with copy number gain revealed by breakpoint sequencing.
(B) Nimblegen 4.2M aCGH plots for the trio. The displayed regions correspond to the regions in (A). Parental aCGH analyses demonstrated that all the copy
number changes were de novo. Triplication was indicated by blue and duplication was indicated by red. The pedigrees are on the bottom.
(C) Representative SNP transmission patterns that suggest a paternal interchromosome origin of the rearrangements.
(D) Zoomed-in view of the SNP array data for the 5.5 Mb triplication and the 0.78 Mb duplication in BAB3105. Five different genotypes are observed across the
entire triplicated region.
See also Figure S2.
Cell 146, 889903, September 16, 2011 2011 Elsevier Inc. 897
898 Cell 146, 889903, September 16, 2011 2011 Elsevier Inc.
of chromosomes or regions of chromosomes, and their re-ligation with scant regard for their site of origin, is based on data
obtained by next generation paired-end sequencing. This technique provides excellent data on the relative positions of
sequences, but cannot directly reveal copy number dosages.
Whereas next generation sequencing can infer copy number
dosage based on depth of read coverage, in the analysis the
raw read length data are first processed through matching to a
reference haploid human genome build; a filtering process notoriously challenged by low copy repeats, repetitive sequences,
and other sequence complexities.
By use of a combination of molecular techniques, we have obtained descriptions of chromosomes that have experienced a
chromothripsis-like process that includes full detail of copy
number, position and orientation in noncancerous patients. This
has revealed that, in addition to inversion and translocation, there
is extensive duplication and triplication of sequence. We also
show that some novel junctions have microhomology and smallscale complexity in the form of apparent templated insertion of
fragments of nearby sequence at the junctions. These features
are characteristic of events that we have previously attributed to
postulated replicative processes of chromosomal structural
changes, long-distance template-switching by FoSTeS/MMBIR;
the latter apparent insertional complexity near the junctions can
be attributed to a synthesis product resultant from a new lowprocessivity fork as proposed for the MMBIR model (Hastings
et al., 2009a). Interestingly, DNA polymerase(s) involved in
break-induced replication (BIR) may also have poor fidelity
(Deem et al., 2011; Hicks et al., 2010).
Although a duplication might potentially result from chromosome shattering and re-ligation, either because the chromosomes have replicated, or by involvement of both homologs,
explanation of triplication in this way would become complicated. Furthermore, such a shattering and re-ligation mechanism predicts the presence of a deletion reciprocal to any
duplication, and this is not observed. A more parsimonious
explanation for the origin of the extra copies is that they were
formed by replication. Although re-replication, the inappropriate
firing of replication origins, might explain some cases of overreplication (Doksani et al., 2009; Green et al., 2010), BIR could
also lead to over-replication, either by use of ectopic homology
or non-homologous processes using microhomology to anneal
single-strands that act as primers for DNA replication. Thus,
MMBIR provides an explanation for duplication and triplication
as well as deletion, inversion, and translocation. In addition,
MMBIR can explain both the observed microhomology at
selected breakpoints and the insertion of short segments flanked
by microhomology around the junctions. Can MMBIR also
breakpoint sequencing data. Specific breakpoint sequences and alignments to the reference sequences are shown below. Reference sequences are named as
PLUS, MINUS, PLU1 (plus1), and MIN1 (minus1), etc., in order to indicate the orientation of DNA strand. A transition between the plus and minus strands within
a breakpoint indicates that such a rearrangement causes an inversion. Microhomologies found at the breakpoints are boxed in red, whereas short sequences
insertions at the breakpoints with no microhomology are boxed in black.
(A) BAB3105. Six breakpoint junctions were obtained, providing information about their relative positions in the rearranged chromosome. In breakpoints #3#6,
novel sequence insertions, ranging from 541542 bp, were identified at the fusion point.
(B) BAB2870. In breakpoint junction #2, a 406 bp and a 90 bp segments copied from nearby genomic sequences are inserted. There is one 8 bp novel sequence
inserted within the 90 bp segment (green).
See also Figure S6.
Cell 146, 889903, September 16, 2011 2011 Elsevier Inc. 899
900 Cell 146, 889903, September 16, 2011 2011 Elsevier Inc.
within two years. Several lines of evidence support the contention that in the constitutional CGR cases in our cohort, in
particular the four chromothripsis-like cases, the majority of
rearranged genomic intervals also arose as one singular event.
First, the multiple rearrangements are localized to a focused
region on one homolog of a chromosome, most frequently in
the long arms. In a mutation accumulation model, the rearrangements would be expected to disperse randomly over the
genome. Second, there is no evidence of differential level of
mosaicism among individual rearrangements, which would be
the expectation when new rearrangements accumulate as cells
proliferate. Third, clustering and juxtaposition of breakpoints
indicate that these breakpoints were interrelated when they
were formed. Notably, inversion of genomic segments gives
proximity of what appears to be distant breakpoints (Branzei
and Foiani, 2010; Carvalho et al., 2011b; Futcher, 1986).
If replicative mechanisms are used, CGR can occur during
DNA replication in gametogenesis and, in some cases, as postzygotic events (Zhang et al., 2009b). Here, we provide evidence
for examples of both types (BAB3105 and 3012). Given the
resemblance of key features between constitutional CGR and
cancer chromothripsis, we propose that such chromosome
catastrophe events can occur by essentially similar mechanisms
in a variety of stages in an organisms life cycle.
In summary, genomic studies of rearrangements associated
with cancer and genomic disorders reveal unanticipated
Cell 146, 889903, September 16, 2011 2011 Elsevier Inc. 901
at least one BAC clone for one chromosome band at a resolution of 550 band
level, or by the oligo-based arrays with one probe for every 30 kb interval.
High-Density aCGH
To characterize further the complex rearrangements identified by CMA, two
customized Agilent arrays were designed including an 8X 60K array format
for investigation of the rearrangements in seven patients, BAB2760, 2778,
2785, 3012, 3015, 3047, and 3050, and a 4X 180K array format for four patients
BAB2780, 2783, 3032, and 3103. The designed probe density was two to four
oligonucleotides per kilobase for intervals with copy number changes and
breakpoint regions as assessed by transitions between two separate copy
number states. The high-density arrays also interrogate the flanking genomic
regions of up to 5 Mb in size with probe density of one to three oligos per
kilobase.
Patients BAB2780, 2920, 3011, 3050, 3104, 3105, and patients 1 and 2 were
analyzed by Nimblegen 2.1M or 4.2M arrays. Patient BAB3104 was also
analyzed by a whole-genome 244K array (Agilent Technologies); assays
were performed according to the manufacturers instructions.
Fluorescence In Situ Hybridization
After identification of a complex rearrangement by clinical CMA, confirmatory
FISH analyses were performed on Phytohemagglutinin-stimulated cultured
blood cells with standard procedures (Cheung et al., 2005). For the two-color
FISH analyses, the test probes were directly labeled by nick-translation with
rhodamine (red), while the control probes were directly labeled with FITC
(green). To confirm a deletion detected by CMA, ten metaphase cells were
examined using the centromere probe as a control. To confirm a copy number
gain detected by CMA, 50 interphase nuclei were examined. A region was
considered to be duplicated when two red signals or triplicated when three
signals were observed in more than 70% of the cells examined. Metaphase cells
were also examined to determine chromosome/genome position and investigate whether the duplicated segments were translocated to a different location.
SNP Array Analysis
SNP array analysis was performed on Illumina Infinium HD assay platform with
HumanOmni1-Quad BeadChip (Illumina) DNA whole-genome amplification,
fragmentation, hybridization, enzymatic single base extension, slides staining,
and washing were performed according to the manufacturers instructions
(Illumina). The Illumina iScan System was used for image registration, image
extraction, and data output. The GenomeStudio software (Illumina) was used
for genotyping data normalization, genotype calling, clustering, data intensity
analysis, and calculation of Log R ratio and B-allele frequency. Illumina CNV
partition statistical algorithm was used for copy number variation and copy
number neutral absence of heterozygosity analyses. The reference model file
provided by Illumina was used as a reference for data analysis. Parent of origin
was determined by comparison of probands and parental genotypes and
B-allele frequencies for SNPs in the rearranged regions on chromosome 9.
Breakpoint Analysis
Genomic coordinates for potential breakpoints were estimated from aCGH
interrogating oligonucleotides revealing gain/no gain or loss/no loss copy
number transitions. Rearrangement breakpoints located in subtelomeric or
pericentromeric regions were not included in this analysis. The presence of
LCRs was revealed by examination of the segmental_dups track in the UCSC
genome browser (assembly hg19), while the presence of both direct and inverted short repeats was revealed by examination of the self_chain track.
The breakpoint junctions were amplified by long-range PCR with the
TAKARA LA Taq kit (TAKARA Bio) as previously described (Zhang et al.,
2009b). PCR products were sequenced by the Sanger dideoxy method and
DNA sequences were compared to the human genome reference assembly
(hg19).
ACKNOWLEDGMENTS
We thank the patients and their families for participation in this research, the
Molecular Cytogenetics Laboratories in the MGL at BCM (https://www.bcm.
edu/geneticlabs/) for providing technical support, Dr. Katarzyna Derwinska
for helpful discussion, and Roche NimbleGen for supplying NimbleGen
aCGH materials. This work was supported in part by National Institute of
Neurological Disorders and Stroke, National Institutes of Health grant R01
NS058529 to J.R.L., National Institutes of General Medical Science grant
R01 GM064022 to P.J.H., and the BCM IDDRC P30HD024064 funded from
the Eunice Kennedy Shriver National Institute of Child Health and Human
Development. A.E. is supported by DK081735-02 and S.C.S.N. is supported
by a fellowship grants by the LCRC from the Osteogenesis Imperfecta Foundation and the National Urea Cycle Disorders Foundation. J.R.L. is a consultant
for Athena Diagnostics, has stock ownership in 23andMe and Ion Torrent
Systems and is a coinventor on multiple United States and European patents
for DNA diagnostics. The Department of Molecular and Human Genetics
derives revenue from clinical testing by high-resolution human genome
analysis.
Received: May 17, 2011
Revised: June 6, 2011
Accepted: July 25, 2011
Published: September 15, 2011
REFERENCES
Bailey, J.A., Gu, Z., Clark, R.A., Reinert, K., Samonte, R.V., Schwartz, S.,
Adams, M.D., Myers, E.W., Li, P.W., and Eichler, E.E. (2002). Recent
segmental duplications in the human genome. Science 297, 10031007.
Batista, D.A., Pai, G.S., and Stetten, G. (1994). Molecular analysis of a complex
chromosomal rearrangement and a review of familial cases. Am. J. Med.
Genet. 53, 255263.
Bi, W., Sapir, T., Shchelochkov, O.A., Zhang, F., Withers, M.A., Hunter, J.V.,
Levy, T., Shinder, V., Peiffer, D.A., Gunderson, K.L., et al. (2009). Increased
LIS1 expression affects human and mouse brain development. Nat. Genet.
41, 168177.
Boone, P.M., Bacino, C.A., Shaw, C.A., Eng, P.A., Hixson, P.M., Pursley, A.N.,
Kang, S.H., Yang, Y., Wiszniewska, J., Nowakowska, B.A., et al. (2010).
Detection of clinically relevant exonic copy-number changes by array CGH.
Hum. Mutat. 31, 13261342.
Borg, K., Stankiewicz, P., Bocian, E., Kruczek, A., Obersztyn, E., Lupski, J.R.,
and Mazurczak, T. (2005). Molecular analysis of a constitutional complex
genome rearrangement with 11 breakpoints involving chromosomes 3, 11,
12, and 21 and a approximately 0.5-Mb submicroscopic deletion in a patient
with mild mental retardation. Hum. Genet. 118, 267275.
Branzei, D., and Foiani, M. (2010). Leaping forks at inverted repeats. Genes
Dev. 24, 59.
Carvalho, C.M., Zhang, F., Liu, P., Patel, A., Sahoo, T., Bacino, C.A., Shaw, C.,
Peacock, S., Pursley, A., Tavyev, Y.J., et al. (2009). Complex rearrangements
in patients with duplications of MECP2 can occur by fork stalling and template
switching. Hum. Mol. Genet. 18, 21882203.
Carvalho, C., Bartnik, M., Pehlivan, D., Fang, P., Shen, J., and Lupski, J.
(2011a). Evidence for disease penetrance relating to CNV size: PelizaeusMerzbacher disease and manifesting carriers with a familial 11 Mb duplication
at Xq22. Clin. Genet. Published online May 30, 2011. 10.1111/j.1399-0004.
2011.01716.x.
SUPPLEMENTAL INFORMATION
Carvalho, C.M., Ramocki, M.B., Pehlivan, D., Franco, L.M., GonzagaJauregui, C., Fang, P., McCall, A., Pivnick, E.K., Hines-Dowell, S., Seaver,
L., et al. (2011b). Inverted genomic segments and complex triplication
rearrangements are mediated by inverted repeats in the human genome.
Nat. Genet., in press.
Supplemental Information includes six figures and two tables and can be
found with this article online at doi:10.1016/j.cell.2011.07.042.
Chen, J.M., Cooper, D.N., Ferec, C., Kehrer-Sawatzki, H., and Patrinos, G.P.
(2010). Genomic rearrangements in inherited disease and cancer. Semin.
Cancer Biol. 20, 222233.
902 Cell 146, 889903, September 16, 2011 2011 Elsevier Inc.
Cheung, S.W., Shaw, C.A., Yu, W., Li, J., Ou, Z., Patel, A., Yatsenko, S.A.,
Cooper, M.L., Furman, P., Stankiewicz, P., et al. (2005). Development and
validation of a CGH microarray for clinical cytogenetic diagnosis. Genet.
Med. 7, 422432.
Deem, A., Keszthelyi, A., Blackgrove, T., Vayl, A., Coffey, B., Mathur, R.,
Chabes, A., and Malkova, A. (2011). Break-induced replication is highly
inaccurate. PLoS Biol. 9, e1000594.
Doksani, Y., Bermejo, R., Fiorani, S., Haber, J.E., and Foiani, M. (2009).
Replicon dynamics, dormant origin firing, and terminal fork integrity after
double-strand break formation. Cell 137, 247258.
Futcher, A.B. (1986). Copy number amplification of the 2 micron circle plasmid
of Saccharomyces cerevisiae. J. Theor. Biol. 119, 197204.
Gajecka, M., Saitta, S.C., Gentles, A.J., Campbell, L., Ciprero, K., Geiger, E.,
Catherwood, A., Rosenfeld, J.A., Shaikh, T., and Shaffer, L.G. (2010).
Recurrent interstitial 1p36 deletions: Evidence for germline mosaicism and
complex rearrangement breakpoints. Am. J. Med. Genet. A. 152A, 30743083.
Green, B.M., Finn, K.J., and Li, J.J. (2010). Loss of DNA replication control is
a potent inducer of gene amplification. Science 329, 943946.
Hastings, P.J., Ira, G., and Lupski, J.R. (2009a). A microhomology-mediated
break-induced replication model for the origin of human copy number variation. PLoS Genet. 5, e1000327.
Hastings, P.J., Lupski, J.R., Rosenberg, S.M., and Ira, G. (2009b). Mechanisms of change in gene copy number. Nat. Rev. Genet. 10, 551564.
Hicks, W.M., Kim, M., and Haber, J.E. (2010). Increased mutagenesis and
unique mutation signature associated with mitotic gene conversion. Science
329, 8285.
Kloosterman, W.P., Guryev, V., van Roosmalen, M., Duran, K.J., de Bruijn, E.,
Bakker, S.C., Letteboer, T., van Nesselrooij, B., Hochstenbach, R., Poot, M.,
and Cuppen, E. (2011). Chromothripsis as a mechanism driving complex
de novo structural rearrangements in the germline. Hum. Mol. Genet. 20,
19161924.
Lee, J.A., Carvalho, C.M., and Lupski, J.R. (2007). A DNA replication mechanism for generating nonrecurrent rearrangements associated with genomic
disorders. Cell 131, 12351247.
Liu, P., Erez, A., Nagamani, S.C., Bi, W., Carvalho, C.M., Simmons, A.D.,
Wiszniewska, J., Fang, P., Eng, P.A., Cooper, M.L., et al. (2011). Copy number
gain at Xp22.31 includes complex duplication rearrangements and recurrent
triplications. Hum. Mol. Genet. 20, 19751988.
Lu, X., Shaw, C.A., Patel, A., Li, J., Cooper, M.L., Wells, W.R., Sullivan, C.M.,
Sahoo, T., Yatsenko, S.A., Bacino, C.A., et al. (2007). Clinical implementation
of chromosomal microarray analysis: summary of 2513 postnatal cases. PLoS
ONE 2, e327.
Lupski, J.R. (2010). New mutations and intellectual function. Nat. Genet. 42,
10361038.
Cell 146, 889903, September 16, 2011 2011 Elsevier Inc. 903
of Medical Oncology, Dana-Farber Cancer Institute, 450 Brookline Avenue, Boston, MA 02215, USA
of Pediatric Oncology, Dana-Farber Cancer Institute and Childrens Hospital Boston, 450 Brookline Avenue, Boston,
MA 02215, USA
3Whitehead Institute for Biomedical Research, 9 Cambridge Center, Cambridge, MA 02142, USA
4Cold Spring Harbor Laboratory, 1 Bungtown Road, Cold Spring Harbor, NY 11724, USA
5Comprehensive Cancer Center, Mayo Clinic Arizona, Scottsdale, AZ 85259, USA
6Department of Medicine, Harvard Medical School, 25 Shattuck Street, Boston, MA 02115, USA
7Department of Biology, Massachusetts Institute of Technology, Cambridge, MA 02142, USA
8Broad Institute of Harvard and MIT, 7 Cambridge Center, Cambridge, MA 02142, USA
9These authors contributed equally to this work
*Correspondence: james_bradner@dfci.harvard.edu (J.E.B.), constantine_mitsiades@dfci.harvard.edu (C.S.M.)
DOI 10.1016/j.cell.2011.08.017
2Department
SUMMARY
RESULTS
BET Bromodomains as Therapeutic Targets in MM
We first evaluated the expression of BRD2, BRD3, BRD4, and
BRDT transcripts in MM by integrating publicly available compendia of gene expression data sets. Among asymptomatic
patients with premalignant disease (Zhan et al., 2007), we
observed increasing expression of BRD4 in monoclonal gammopathy of undetermined significance (MGUS) and smoldering
MM (SMM) compared to normal bone marrow plasma cells (Figure 1B). In a second, independent data set (Mattioli et al., 2005),
we observed significantly higher expression of BRD4 in plasma
cell leukemia (PCL) compared to MM or MGUS samples (Figure 1C). Thus, BRD4 expression correlates positively with
disease progression. BRD2 and BRD3 are also expressed in
MM, but expression does not clearly correlate with stage of
disease (data not shown). BRDT, a testis-specific bromodomain-containing protein, is not expressed in MM.
Analysis of copy number polymorphism (CNP) data collected
on 254 MM patients by the Multiple Myeloma Research Consortium (MMRC) revealed that the BRD4 locus is frequently amplified in MM patient samples (Figure 1D). The majority of patient
samples exhibit broad amplification of chromosome 19p, but
focal amplification at the BRD4 locus is observed (Figure S1
available online). Among 45 established MM cell lines, expression of BRD4 was pronounced and did not correlate with amplification status (Figure 1E).
Human MM cells are highly osteotropic in vivo, and interaction
with bone marrow stromal cells (BMSCs) induces proliferation
and contributes to drug resistance (McMillin et al., 2010). Analysis of BET bromodomain expression, as influenced by MM
cell binding to BMSCs (McMillin et al., 2010), revealed marked
upregulation of BRD4 in the INA-6 human MM cell line upon
interaction with HS5 stromal cells (Figure 1F), suggesting a plausible role for BRD4 function in MM cells within the bone marrow
microenvironment.
To explore the function of BET bromodomains in MM, we
examined the effect on proliferation of small hairpin RNAs
(shRNAs) targeting each of the four BET proteins in comparison
to shRNAs targeting 1011 kinases, phosphatases, and oncogenes in a lentivirally delivered, arrayed shRNA screen in INA-6
cells. As illustrated in Figures 1G and 1H, shRNA constructs
targeting each of the expressed BET bromodomains are identified as reducing INA-6 proliferation as shown by normalized
B scores (Malo et al., 2006). Together, these data establish
Cell 146, 904917, September 16, 2011 2011 Elsevier Inc. 905
B
N O
O
O
Cl
N O
N
H
Cl
(+)-JQ1
C
BRD4
iBET
4
2
0
1.5
1.0
0.5
0.0
NPCs
BRD4
2.0
Expression (log2)
Expression (log2)
MGUS
SMM
MGUS
MM
PCL
E
Chr19
0.06
p11
q13.2
BRD4
q13.42
Normalized Expression
BRD4
0.04
*
Neutral
* *
0.02
0.00
Deleted
BRDT
Amplified
* * * *
* * *
INA-6
BRD2
BRD4
BRD2
BRD4
BRD3
BRD2
-2
-4
226054_at
226052_at
BRD3
BRD4
-2
-4
BRD3
-6
-6
202102_s_at
B-Score (Mean)
BRD3
100
300
B-Score
BRD4 Expression
INA-6
INA-6 + HS5
200
H
BRD2
400
OCI-MY5
KHM-1B
KMS-12BM
OCI-MY7
KARPAS620
XG-1
FLAM-76
LP-1
ANBL-6
SACHI
L363
CAG
OPM-2
PE1
KMS-12PE
NCI-H929
MM.1.144
KMS-18
PE2
RPMI-8226
EJM
FR4
KMS-28PE
JK-6L
OCI-MY1
INA-6
H1112
KMS-34
KMS-26
OPM-1
KMS-28BM
JJN3
ARP-1
XG-7
KHM-11
SKMM-2
KMS-11
UTMC-2
JIM3
DELTA-47
MM-M1
KMM-1
U266
XG-2
SKMM-1
p13.1
906 Cell 146, 904917, September 16, 2011 2011 Elsevier Inc.
(Figure 3A). Excellent concordance was observed between replicate measurements of expressed genes (Figure S3A). Unsupervised hierarchical clustering segregated replicate data correctly
into early- and late-treatment time points. Surprisingly, we observed immediate, progressive, and profound downregulation
of MYC transcription itself, a unique finding among all transcripts
studied (p < 0.05).
Downregulation of MYC was further confirmed by RT-PCR
and immunoblot (Figures 3B and 3C). This effect was BET bromodomain specific, supported by the nearly comparable activity
of an analogous BET inhibitor subsequently published by Glaxo
SmithKline (iBET) (Figure 1A) (Nicodeme et al., 2010) and the lack
of activity of the inactive (-)- JQ1 enantiomer, which we previously characterized as structurally incapable of inhibiting BET
bromodomains (Filippakopoulos et al., 2010) (Figure 3C). Inhibition of MYC transcription by JQ1 was observed to be dose and
time dependent, with peak inhibition at submicromolar concentrations (Figures 3D and 3E). Rapid depletion of chromatinbound c-Myc was confirmed by nuclear ELISA transcription
factor-binding assays (Figure 3F). In contrast, NF-kB and AP-1
chromatin-binding assays failed to reveal any decrease in DNA
binding within 8 hr of JQ1 treatment (Figure 3G and Figure S4A).
To assess the breadth of these findings in MM, we expanded
gene expression studies to three MM cells with distinct lesions at
the MYC locus. MM.1S cells have a complex MYC rearrangement involving an IgH insertion at the breakpoint of a derivative
chromosome der3t(3;8); KMS-11 cells have both MYC duplication and inversion; and OPM1 cells feature a der(8)t(1;8)
(Dib et al., 2008). Among 230 genes studied, MYC was one of
only four genes downregulated by treatment with JQ1, along
with MYB, TYRO3, and TERT (Figure 3H and Figure S4B).
Immunoblotting analyses confirmed the JQ1 suppression of
c-Myc protein expression in a further expanded panel of Mycdependent MM cell lines (Figure 3I). Despite the intriguing potential effect on E2F transcriptional function and MYB gene expression, JQ1 did not influence E2F or MYB protein abundance
through 24 hr of drug exposure (Figures S4C and S4D). Together,
these data support the general observation that BET inhibition
specifically suppresses MYC transcription across MM cells
with different genetic lesions affecting the MYC locus and with
striking selectivity in comparison to other oncogenic transcriptions factors with established roles in MM pathophysiology.
BRD4 Binds IgH Enhancers, Regulating MYC Expression
and Function
Based on the integrated, functional genomic analysis of BET bromodomains in MM (Figure 1), we pursued further mechanistic
studies of BRD4. Silencing of BRD4 using directed shRNAs
(E) Expression levels of BRD4 (compared to BRDT) in human MM cell lines. Asterisks denote cell lines with amplification of the BRD4 locus (19p13.1).
(F) BRD4 expression (depicted on a linear scale for three different oligonucleotide microarray probes) in INA-6 MM cells cultured in vitro in the presence or
absence of HS-5 bone marrow stromal cells.
(G) Silencing of BET bromodomains impairs proliferation in MM cells. Results of an arrayed lentiviral screen using a diverse shRNA library in INA-6 MM cells are
presented in rank order of ascending B scores. The effect of shRNAs targeting BET bromodomains on INA-6 cell viability is highlighted by red circles and
annotated by gene. Gray dots represent results for non-BET bromodomain shRNAs.
(H) Silencing of BET bromodomain family members in MM cells. Viability of INA-6 MM cells exposed to shRNAs directed against BRD2, BRD3, and BRD4 are
reported as mean B scores ( SD of the two normalized replicates).
Cell 146, 904917, September 16, 2011 2011 Elsevier Inc. 907
VEH JQ1
0.0
-0.2
-0.4
-0.6
-0.8 q < 0.0001
Enrichment Score
Enrichment Score
q < 0.0001
0.0
-0.2
-0.4
-0.6
q < 0.0001
q < 0.0001
D
V$MYCMAX_01
MYC
E2F
Enrichment Score
1
0.8
FDR
SLC38A5
ACSL5
NME1
SORD
MAP1D
PTPN22
CCR1
KCNQ5
MYB
AMPD1
SLC7A2
MORC1
MYC
ADAT2
HBD
ALDH1B1
ZNF485
POLR3G
UNQ3104
NAV1
SRM
KCNA3
MGC29506
GTF3C6
ABCC4
RNF125
MTMR2
RRS1
KAT2A
SFXN4
GALNT14
SLC16A9
MAP4K1
CDC25A
MMACHC
FKBP11
RAI14
ABLIM1
TYRO3
MANEAL
XTP3TPA
BTN3A2
MTHFD1L
ACSM3
DERL3
BDH1
FADS1
TTC27
POLE2
SLC19A1
MAP2
CALCOCO1
CNTN5
PAG1
SYT11
YPEL1
NFKBIZ
ZSWIM6
TMEM2
HEXIM1
APOLD1
STAT2
SAT1
KLHL24
PNPLA8
JARID1B
BMPR2
SCN9A
SLC12A6
ZFYVE1
SESN3
C9ORF95
SERPINI1
KIAA0913
WDR47
BNIP3L
HIST2H2BE
HHLA3
C1ORF63
C13ORF31
KIAA0825
OR2B6
USP11
DOPEY2
RNF19B
DNM3
JHDM1D
YPEL5
ZFP36
ITFG3
LYST
SATB1
C1ORF26
DNAJC28
HIST2H4A
CLDN12
LMNA
SEPP1
LGALS1
SQSTM1
Enrichment Score
KMS11
MM.1S
OPM1
KMS11
MM.1S
OPM1
0.6
0.4
0.0
-0.2
-0.4
-0.6
q < 0.0001
0.2
0
-2 -1.5 -1 -0.5
0.5
1.5
NES
Gene Set
n
SCHUHMACHER_MYC_TARGETS_UP
67
DANG_MYC_TARGETS_UP
1 27
SCHLOSSER_MYC_TARGETS_AND_SERUM_RESPONSE_UP 47
KIM_MYC_AMPLIFICATION_TARGETS_UP
153
MORI_EMU_MYC_LYMPHOMA_BY_ONSET_TIME_UP
96
RIBOSOME_BIOGENESIS_AND_ASSEMBLY
14
YU_MYC_TARGETS_UP
37
MOOTHA_GLYCOLYSIS
21
NES
-2.61
-2.29
-2.29
-2.27
-1.953
-1.78
-1.65
-1.42
FDR q-val
< 0.0001
< 0.0001
< 0.0001
< 0.0001
0.0001
0.0006
0.004
0.045
V$MYCMAX_01
V$MYCMAX_02
V$NFKAPPAB65
V$AP1_Q4
V$STAT3_02
V$MYB_Q3
V$HSF1_01
V$GR_01
V$XBP1_01
-2.078
- 1 .7 2
-1.13
-1.11
-1.04
-0.92
-0.92
-0.88
-0.89
< 0.0001
0 .0 0 1 8
NS
NS
NS
NS
NS
NS
NS
192
200
190
214
111
176
197
155
107
908 Cell 146, 904917, September 16, 2011 2011 Elsevier Inc.
(C) Quantitative comparison of all transcription factor target gene sets available from the MSigDB by GSEA for reduced expression in JQ1-treated MM cell lines.
Data are presented as scatterplot of false discovery rate (FDR) versus normalized enrichment score (NES) for each evaluated gene set. Colored dots indicate gene
sets for MYC (red), E2F (black), or other (gray) transcription factors.
(D) GSEA showing downregulation in JQ1-treated MM cells of a representative set of genes with proximal promoter regions containing Myc-Max-binding sites.
(E) Table of gene sets enriched among genes downregulated by JQ1 in MM cells (top group), highlighting the number of genes in each set (n), the normalized
enrichment score (NES), and test of statistical significance (FDR q value). The bottom group represents comparisons of top-ranking transcription factor target
gene sets of MM master regulatory proteins, enriched among genes downregulated by JQ1 in MM cells. See also Figure S2.
Cell 146, 904917, September 16, 2011 2011 Elsevier Inc. 909
910 Cell 146, 904917, September 16, 2011 2011 Elsevier Inc.
Relative mRNA
1.0
shRluc
shBRD4.533
shBRD4.602
shBRD4.1838
0.8
54%
RLuc
BRD4
533
17.9%
0.6
0.4
Count
0.2
0.0
BRD4
MYC
APC
C
Enrichment over
background region
350
Empty
VEH
JQ1
300
JQ1
Myc
250
200
Myc
150
100
-actin
50
0
NR2 NR3
E1 E2 E3 E4 E5
TS1 TS2
E1 E2 E3 E4
Negative
Region
MYC Enhancers
MYC TSS
IgH Enhancers
E
OPM1
Empty
Vehicle
OPM1
Empty
JQ1
OPM1
MYC
Vehicle
62 %
OPM1
MYC
JQ1
47 %
63 %
Count
31 %
APC
proven successful. To date, efforts to target c-Myc have identified only a small number of molecules with low biochemical
potency and limited biological characterization (Bidwell et al.,
2009; Hammoudeh et al., 2009; Jeong et al., 2010), underscoring
both the challenge of targeting c-Myc as well as the enduring
need for chemical probes of c-Myc transcriptional function.
Considering chromatin as a platform for signal transduction
(H) Heatmap of clustered gene expression data from multiplexed measurement (Nanostring) of cancer-associated genes in three human MM cell lines treated
with JQ1 or vehicle control. Among 230 genes studied (Figure S4), four genes (MYC, TERT, TYRO3, and MYB) exhibited statistically significant (p < 0.05)
downregulation. Replicate expression measurements exhibited high concordance among low and highly expressed genes (Figure S3B).
(I) Immunoblotting study of four MM lines (KMS11, LR5, OPM1, and INA-6) identifies a JQ1-induced decrease in c-Myc expression (500 nM, 24 hr).
Cell 146, 904917, September 16, 2011 2011 Elsevier Inc. 911
B
0
25
50
100
200
400
KMS-34
LR5
MOLP-8
Dox 40
INA-6
MM.1S
KMS-5
KMS-12-PE
AMO-1
MR20
OPM-1
KMS-11
KMS-12-BM
KMS-26
L363
EJM
K MM1
MM.1S-myrAkt
KMS-20
MM. 1S-Bc l-2
RPMI-8226/S
OCI-MY5
KMS-28-BM
KMS-18
MM. 1R
100
50
IC-50
800 (nM)
68
98
136
122
164
152
116
108
124
167
155
163
230
235
241
285
317
365
341
326
395
432
502
1150
1490
MYC Annotation
der8t(8;?)
der16t(16;8;22)
der16t(16;8;22)
IgH insertion at breakpoint of der3t(3;8)
IgH insertion near MYC
der16t(16;8;22)
der(8)t(1;8) with insertion of IgH
MYC insertion, duplication and inversion
IgH insertion near MYC
der(8)t(8;22)
t(5;8); der8x2+der5
Negative
MYC insertion at IgH locus
IgH insertion at breakpoint of der3t(3;8)
IgH insertion at breakpoint of der3t(3;8)
der16t(16;8;22)
der(der14)t(8;14); duplication and inversion
t(8;14)
der(6&21)t(6&21;8)
IgH insertion at breakpoint of der3t(3;8)
10
50
0
0.0
0.2
150
1.0
L-363
100
50
0
0.0
0.2
150
1.0
Dox40
50
0
0.2
1.0
150
KMS-20
100
50
0
0.0
0.2
200
1.0
AMO-1
150
100
100
0.0
RPMI-8226
Normalized Proliferation
100
150
Normalized Proliferation
MM.1S
Normalized Proliferation
150
Normalized Proliferation
Normalized Proliferation
Normalized Proliferation
100
50
0
0.0
0.2
1.0
50
0
0.0
0.2
1.0
and concomitant downregulation of the Myc-dependent transcriptional network lead to growth-inhibitory effects sharing the
specificity of phenotypes associated with prior genetic models
of Myc inhibition. These are notable observations that distinguish
the transcriptional consequences of BET inhibition from other
nonselective transcriptional inhibitors, such as actinomycin D,
a-amanitin, and flavopiridol.
A compelling finding is the observed, direct interaction of
BRD4 with IgH enhancers in MM cells possessing IgH rearrangement into the MYC locus and the depletion of BRD4 binding by
JQ1. This suggests BET inhibition as a strategy for targeting
other structural rearrangements in cancer involving IgH or other
A
VEH
Count
S 40%
G1 42%
JQ1 (24h)
JQ1 (48h)
S 5%
G1 80%
S 5%
G1 81%
PI
B
VEH
12%
12%
JQ1 (24h)
3%
23%
5%
PI
3%
JQ1 (48h)
AV
D
Enrichment Score
C
FRIDMAN_SENESCENCE_UP
0.4
0.3
0.2
0.1
0.0
Vehicle (40x)
strong enhancers and has potential implications for the modulation of immunoglobulin gene expression in autoimmune
diseases.
An unexpected finding was the pronounced and concordant
suppression of multiple E2F-dependent transcriptional signatures. In this instance, E2F1 protein and transcript levels were
not affected by BET inhibition, suggesting either an unrecognized
function of BET bromodomains in E2F transcriptional complexes
or a dominant effect of Myc downregulation causing cell-cycle
arrest in G1 leading to silencing of E2F. These observations are
also compatible with the known role of Myc and E2F1 as transcriptional collaborators in cell-cycle progression and tumor
cell survival (Matsumura et al., 2003; Trimarchi and Lees, 2002).
Insights provided by our study identify rational strategies for
combination therapeutic approaches warranting exploration in
MM. MYC activation is commonly accompanied by antiapoptotic signaling in human cancer. In MM, constitutive or microenvironment-inducible activation of antiapoptotic Bcl-2 proteins
has been reported (Harada et al., 1998; Legartova et al., 2009).
Thus, Myc pathway inhibition by JQ1 may demonstrate synergism with targeted proapoptotic agents (e.g., ABT-737) (Oltersdorf et al., 2005; Trudel et al., 2007). Additionally, the selective
effect of JQ1 on Myc and E2F1 transcriptional programs provides an opportunity to combine BET inhibitors with pathwaydirected antagonists of the NF-kB, STAT3, XBP1, or HSF1 transcriptional programs.
Direct inhibition of c-Myc remains a central challenge in the
discipline of ligand discovery. Inhibition of MYC expression and
function, demonstrated herein, presents an immediate opportunity to study and translate the concept of c-Myc inhibition more
Cell 146, 904917, September 16, 2011 2011 Elsevier Inc. 913
B
100
% Survival
Pt#1
Pt#2
Pt#3
Pt#4
Pt#5
50
c-Myc
GAPDH
0
0
200
400
JQ1 (nM)
600
800
D
Bioluminescence
(ph/s/cm 2/sr)
2.5 x 1010
VEH
JQ1
JQ1
Vehicle
1.5 x 1010
5.0 x 10 9
p = 0.0002
0
0
Luminescence (x106)
10
15
20
25
Day of Treatment
0.5
1.0
1.5
2.0
2.5
3.0
Percent survival
100
Mouse 1
Day
JQ1
80
Mouse 2
14
14
Vehicle
60
40
20
p < 0.0001
0
0
10
M Protein
20
30
40
Day of Treatment
EXPERIMENTAL PROCEDURES
Gene Expression Analysis
MM cells treated with JQ1 (500 nM, 24 h) were processed for transcriptional
profiling using Affymetrix Human Gene 1.0 ST microarrays. Expression of individual genes was assessed in the context of dose- and time-ranging experiments by real-time quantitative polymerase chain reaction, multiplexed direct
detection (Nanostring), and immunoblotting using antibodies as described in
the Extended Experimental Procedures.
Chromatin Immunoprecipitation
ChIP was performed on MM.1S cells cultured in the presence or absence of
JQ1 (500 nM, 24 hr). Specific antibodies, detailed methods, and primer sequences for MYC and IgH enhancers, as well as the MYC TSS, are described
in the Extended Experimental Procedures.
In Vitro and In Vivo MM Studies
The impact of JQ1 on cell viability, proliferation, and cell cycle was assessed in
human MM cells as documented in the Extended Experimental Procedures.
In vivo efficacy studies were performed with protocols approved by Institutional Animal Care and Use Committees at the DFCI or Mayo Clinic Arizona.
JQ1 was administered by intraperitoneal injection into SCID-beige mice with
MM lesions established after subcutaneous or intravenous injections and in
nonimmunocompromised tumor-bearing Vk*myc mice. Tumor burden in these
models was quantified by caliper measurement, whole-body bioluminescence
imaging, and serum protein electrophoresis, respectively, as detailed in the
Extended Experimental Procedures.
ACCESSION NUMBERS
Oligonucleotide microarray data have been deposited in the Gene Expression
Omnibus under the accession number GSE31365.
SUPPLEMENTAL INFORMATION
Supplemental Information includes Extended Experimental Procedures,
seven figures, and one table and can be found with this article online at
doi:10.1016/j.cell.2011.08.017.
ACKNOWLEDGMENTS
We are grateful to S. Lowe for sharing unpublished information; A. Azab, D.
McMillin, C. Ott, and A. Roccaro for technical support; E. Fox for microarray
data; J. Daley and S. Lazo-Kallanian for flow cytometry; the MMRF, MMRC,
and Broad Institute for establishing the MM Genomics Portal (http://www.
broadinstitute.org/mmgp/). This research was supported by NIHK08CA128972 (J.E.B.), NIH-R01CA050947 (C.S.M.), NIH-R01HG002668
(R.A.Y.), and NIH-R01CA46455 (R.A.Y.); the Chambers Medical Foundation
(P.G.R., C.S.M.); the Stepanian Fund for Myeloma Research (P.G.R.,
C.S.M.); and the Richard J. Corman Foundation (P.G.R., C.S.M.); an American
Cancer Society Postdoctoral Fellowship, 120272-PF-11-042-01-DMC
(P.B.R.); the Burroughs-Wellcome Fund, the Smith Family Award, the
Damon-Runyon Cancer Research Foundation, and the MMRF (to J.E.B.).
J.E.B. and C.S.M. designed the study, analyzed data, and prepared the
manuscript. J.E.D., H.M.J., and E.K. assayed MM drug sensitivity. G.C.I.
and J.E.D. assessed the effects of JQ1 on Myc expression. J.Q. performed
scaling synthesis and purification of JQ1. P.G.R. and K.C.A. provided primary
MM samples. P.B.R. and T.G. conducted ChIP experiments, and P.B.R. and
R.A.Y. contributed to their interpretation. R.M.P., T.P.H., and M.R.M. performed RNA expression analysis. I.M.G. and K.C.A. provided support and interpreted cellular data. A.C.S. and W.C.H. designed and performed shRNA
screens. M.E.L. analyzed expression array data. J.S. and C.R.V. performed
Myc rescue experiments. A.L.K. supervised in vivo efficacy and biostatistical
studies. M.C. and P.L.B. performed in vivo GEMM studies. J.E.B. and
C.S.M. supervised the research. All authors edited the manuscript.
Cell 146, 904917, September 16, 2011 2011 Elsevier Inc. 915
Frank, S.R., Parisi, T., Taubert, S., Fernandez, P., Fuchs, M., Chan, H.M.,
Livingston, D.M., and Amati, B. (2003). MYC recruits the TIP60 histone acetyltransferase complex to chromatin. EMBO Rep. 4, 575580.
Frye, S.V. (2010). The art of the chemical probe. Nat. Chem. Biol. 6, 159161.
Mattioli, M., Agnelli, L., Fabris, S., Baldini, L., Morabito, F., Bicciato, S.,
Verdelli, D., Intini, D., Nobili, L., Cro, L., et al. (2005). Gene expression profiling
of plasma cell dyscrasias reveals molecular patterns associated with distinct
IGH translocations in multiple myeloma. Oncogene 24, 24612473.
Fukazawa, T., Maeda, Y., Matsuoka, J., Yamatsuji, T., Shigemitsu, K., Morita,
I., Faiola, F., Durbin, M.L., Soucek, L., and Naomoto, Y. (2010). Inhibition of
Myc effectively targets KRAS mutation-positive lung cancer expressing high
levels of Myc. Anticancer Res. 30, 41934200.
McMillin, D.W., Delmore, J., Weisberg, E., Negri, J.M., Geer, D.C., Klippel, S.,
Mitsiades, N., Schlossman, R.L., Munshi, N.C., Kung, A.L., et al. (2010). Tumor
cell-specific bioluminescence platform to identify stroma-induced changes to
anticancer drug activity. Nat. Med. 16, 483489.
Gomi, M., Moriwaki, K., Katagiri, S., Kurata, Y., and Thompson, E.B. (1990).
Glucocorticoid effects on myeloma cells in culture: correlation of growth inhibition with induction of glucocorticoid receptor messenger RNA. Cancer Res.
50, 18731878.
Mitsiades, C.S., Mitsiades, N.S., McMullan, C.J., Poulaki, V., Shringarpure, R.,
Akiyama, M., Hideshima, T., Chauhan, D., Joseph, M., Libermann, T.A., et al.
(2004). Inhibition of the insulin-like growth factor receptor-1 tyrosine kinase
activity as a therapeutic strategy for multiple myeloma, other hematologic
malignancies, and solid tumors. Cancer Cell 5, 221230.
Hammoudeh, D.I., Follis, A.V., Prochownik, E.V., and Metallo, S.J. (2009).
Multiple independent binding sites for small-molecule inhibitors on the oncoprotein c-Myc. J. Am. Chem. Soc. 131, 73907401.
Harada, N., Hata, H., Yoshida, M., Soniki, T., Nagasaki, A., Kuribayashi, N.,
Kimura, T., Matsuzaki, H., and Mitsuya, H. (1998). Expression of Bcl-2 family
of proteins in fresh myeloma cells. Leukemia 12, 18171820.
Harris, A.W., Pinkert, C.A., Crawford, M., Langdon, W.Y., Brinster, R.L., and
Adams, J.M. (1988). The E mu-myc transgenic mouse. A model for highincidence spontaneous lymphoma and leukemia of early B cells. J. Exp.
Med. 167, 353371.
Haynes, S.R., Dollard, C., Winston, F., Beck, S., Trowsdale, J., and Dawid, I.B.
(1992). The bromodomain: a conserved sequence found in human, Drosophila
and yeast proteins. Nucleic Acids Res. 20, 2603.
Hideshima, T., Mitsiades, C., Tonon, G., Richardson, P.G., and Anderson, K.C.
(2007). Understanding multiple myeloma pathogenesis in the bone marrow to
identify new therapeutic targets. Nat. Rev. Cancer 7, 585598.
Hurt, E.M., Wiestner, A., Rosenwald, A., Shaffer, A.L., Campo, E., Grogan, T.,
Bergsagel, P.L., Kuehl, W.M., and Staudt, L.M. (2004). Overexpression of
c-maf is a frequent oncogenic event in multiple myeloma that promotes proliferation and pathological interactions with bone marrow stroma. Cancer Cell 5,
191199.
Jain, M., Arvanitis, C., Chu, K., Dewey, W., Leonhardt, E., Trinh, M., Sundberg,
C.D., Bishop, J.M., and Felsher, D.W. (2002). Sustained loss of a neoplastic
phenotype by brief inactivation of MYC. Science 297, 102104.
Jeong, K.C., Ahn, K.O., and Yang, C.H. (2010). Small-molecule inhibitors of
c-Myc transcriptional factor suppress proliferation and induce apoptosis of
promyelocytic leukemia cell via cell cycle arrest. Mol. Biosyst. 6, 15031509.
Keats, J.J., Fonseca, R., Chesi, M., Schop, R., Baker, A., Chng, W.J., Van Wier,
S., Tiedemann, R., Shi, C.X., Sebag, M., et al. (2007). Promiscuous mutations
activate the noncanonical NF-kappaB pathway in multiple myeloma. Cancer
Cell 12, 131144.
Kim, J., Chu, J., Shen, X., Wang, J., and Orkin, S.H. (2008). An extended transcriptional network for pluripotency of embryonic stem cells. Cell 132, 1049
1061.
Kim, Y.H., Girard, L., Giacomini, C.P., Wang, P., Hernandez-Boussard, T.,
Tibshirani, R., Minna, J.D., and Pollack, J.R. (2006). Combined microarray
analysis of small cell lung cancer reveals altered apoptotic balance and distinct
expression signatures of MYC family gene amplification. Oncogene 25,
130138.
Leder, A., Pattengale, P.K., Kuo, A., Stewart, T.A., and Leder, P. (1986).
Consequences of widespread deregulation of the c-myc gene in transgenic
mice: multiple neoplasms and normal development. Cell 45, 485495.
Legartova, S., Krejci, J., Harnicarova, A., Hajek, R., Kozubek, S., and Bartova,
E. (2009). Nuclear topography of the 1q21 genomic region and Mcl-1 protein
levels associated with pathophysiology of multiple myeloma. Neoplasma 56,
404413.
Malo, N., Hanley, J.A., Cerquozzi, S., Pelletier, J., and Nadon, R. (2006). Statistical practice in high-throughput screening data analysis. Nat. Biotechnol. 24,
167175.
Matsumura, I., Tanaka, H., and Kanakura, Y. (2003). E2F1 and c-Myc in cell
growth and death. Cell Cycle 2, 333338.
916 Cell 146, 904917, September 16, 2011 2011 Elsevier Inc.
Mitsiades, N., Mitsiades, C.S., Poulaki, V., Chauhan, D., Fanourakis, G., Gu, X.,
Bailey, C., Joseph, M., Libermann, T.A., Treon, S.P., et al. (2002). Molecular
sequelae of proteasome inhibition in human multiple myeloma cells. Proc.
Natl. Acad. Sci. USA 99, 1437414379.
Nair, S.K., and Burley, S.K. (2003). X-ray structures of Myc-Max and Mad-Max
recognizing DNA. Molecular bases of regulation by proto-oncogenic transcription factors. Cell 112, 193205.
Nicodeme, E., Jeffrey, K.L., Schaefer, U., Beinke, S., Dewell, S., Chung, C.W.,
Chandwani, R., Marazzi, I., Wilson, P., Coste, H., et al. (2010). Suppression of
inflammation by a synthetic histone mimic. Nature 468, 11191123.
Oltersdorf, T., Elmore, S.W., Shoemaker, A.R., Armstrong, R.C., Augeri, D.J.,
Belli, B.A., Bruncko, M., Deckwerth, T.L., Dinges, J., Hajduk, P.J., et al.
(2005). An inhibitor of Bcl-2 family proteins induces regression of solid
tumours. Nature 435, 677681.
Palumbo, A.P., Pileri, A., Dianzani, U., Massaia, M., Boccadoro, M., and
Calabretta, B. (1989). Altered expression of growth-regulated protooncogenes
in human malignant plasma cells. Cancer Res. 49, 47014704.
Pomerantz, M.M., Ahmadiyeh, N., Jia, L., Herman, P., Verzi, M.P., Doddapaneni, H., Beckwith, C.A., Chan, J.A., Hills, A., Davis, M., et al. (2009a). The
8q24 cancer risk variant rs6983267 shows long-range interaction with MYC
in colorectal cancer. Nat. Genet. 41, 882884.
Pomerantz, M.M., Beckwith, C.A., Regan, M.M., Wyman, S.K., Petrovics, G.,
Chen, Y., Hawksworth, D.J., Schumacher, F.R., Mucci, L., Penney, K.L.,
et al. (2009b). Evaluation of the 8q24 prostate cancer risk locus and MYC
expression. Cancer Res. 69, 55685574.
Rahl, P.B., Lin, C.Y., Seila, A.C., Flynn, R.A., McCuine, S., Burge, C.B., Sharp,
P.A., and Young, R.A. (2010). c-Myc regulates transcriptional pause release.
Cell 141, 432445.
Rahman, S., Sowa, M.E., Ottinger, M., Smith, J.A., Shi, Y., Harper, J.W., and
Howley, P.M. (2011). The Brd4 extraterminal domain confers transcription
activation independent of pTEFb by recruiting multiple proteins, including
NSD3. Mol. Cell. Biol. 31, 26412652.
Schlosser, I., Holzel, M., Hoffmann, R., Burtscher, H., Kohlhuber, F., Schuhmacher, M., Chapman, R., Weidle, U.H., and Eick, D. (2005). Dissection of
transcriptional programmes in response to serum and c-Myc in a human
B-cell line. Oncogene 24, 520524.
Schreiber, S.L., and Bernstein, B.E. (2002). Signaling network model of chromatin. Cell 111, 771778.
Schuhmacher, M., Kohlhuber, F., Holzel, M., Kaiser, C., Burtscher, H., Jarsch,
M., Bornkamm, G.W., Laux, G., Polack, A., Weidle, U.H., and Eick, D. (2001).
The transcriptional program of a human B cell line in response to Myc. Nucleic
Acids Res. 29, 397406.
Shaffer, A.L., Emre, N.C., Lamy, L., Ngo, V.N., Wright, G., Xiao, W., Powell, J.,
Dave, S., Yu, X., Zhao, H., et al. (2008). IRF4 addiction in multiple myeloma.
Nature 454, 226231.
Shou, Y., Martelli, M.L., Gabrea, A., Qi, Y., Brents, L.A., Roschke, A., Dewald,
G., Kirsch, I.R., Bergsagel, P.L., and Kuehl, W.M. (2000). Diverse karyotypic
abnormalities of the c-myc locus associated with c-myc dysregulation and
tumor progression in multiple myeloma. Proc. Natl. Acad. Sci. USA 97,
228233.
Soucek, L., Helmer-Citterich, M., Sacco, A., Jucker, R., Cesareni, G., and Nasi,
S. (1998). Design and properties of a Myc derivative that efficiently homodimerizes. Oncogene 17, 24632472.
Soucek, L., Jucker, R., Panacchia, L., Ricordy, R., Tato`, F., and Nasi, S. (2002).
Omomyc, a potential Myc dominant negative, enhances Myc-induced
apoptosis. Cancer Res. 62, 35073510.
Stewart, T.A., Pattengale, P.K., and Leder, P. (1984). Spontaneous mammary
adenocarcinomas in transgenic mice that carry and express MTV/myc fusion
genes. Cell 38, 627637.
Trimarchi, J.M., and Lees, J.A. (2002). Sibling rivalry in the E2F family. Nat. Rev.
Mol. Cell Biol. 3, 1120.
Trudel, S., Stewart, A.K., Li, Z., Shu, Y., Liang, S.B., Trieu, Y., Reece, D.,
Paterson, J., Wang, D., and Wen, X.Y. (2007). The Bcl-2 family protein inhibitor,
ABT-737, has substantial antimyeloma activity and shows synergistic effect
with dexamethasone and melphalan. Clin. Cancer Res. 13, 621629.
Vervoorts, J., Luscher-Firzlaff, J.M., Rottmann, S., Lilischkis, R., Walsemann,
G., Dohmann, K., Austen, M., and Luscher, B. (2003). Stimulation of c-MYC
Cell 146, 904917, September 16, 2011 2011 Elsevier Inc. 917
of Physiological Chemistry
of Translational Oncology
3Department of Pathology
4Department of Bioinformatics
5Department of Biomedical Imaging
Genentech, Inc., 1 DNA Way, South San Francisco, CA 94080, USA
*Correspondence: dixit@gene.com
DOI 10.1016/j.cell.2011.07.040
2Department
SUMMARY
Inhibitors of DNA binding (IDs) antagonize basichelix-loop-helix (bHLH) transcription factors to inhibit
differentiation and maintain stem cell fate. ID ubiquitination and proteasomal degradation occur in differentiated tissues, but IDs in many neoplasms appear
to escape degradation. We show that the deubiquitinating enzyme USP1 promotes ID protein stability
and stem cell-like characteristics in osteosarcoma.
USP1 bound, deubiquitinated, and thereby stabilized
ID1, ID2, and ID3. A subset of primary human osteosarcomas coordinately overexpressed USP1 and ID
proteins. USP1 knockdown in osteosarcoma cells
precipitated ID protein destabilization, cell-cycle
arrest, and osteogenic differentiation. Conversely,
ectopic USP1 expression in mesenchymal stem cells
stabilized ID proteins, inhibited osteoblastic differentiation, and enhanced proliferation. Consistent with
USP1 functioning in normal mesenchymal stem cells,
USP1-deficient mice were osteopenic. Our observations implicate USP1 in preservation of the stem
cell state that characterizes osteosarcoma and identify USP1 as a target for differentiation therapy.
INTRODUCTION
Basic-helix-loop-helix (bHLH) transcription factors comprise the
third-largest family of recognized transcription factors in the
human genome (Tupler et al., 2001) and are essential regulators
of development and differentiation through binding DNA elements termed E boxes (Massari and Murre, 2000). Class I
bHLH homodimers are expressed broadly and promote expression of antiproliferative genes such as CDKN1A, CDKN2A, and
CDKN2B (Yokota and Mori, 2002). Class II bHLH proteins
show more restricted expression and form heterodimers with
918 Cell 146, 918930, September 16, 2011 2011 Elsevier Inc.
nated ID2 purified from 293T cells was incubated in vitro with
either wild-type USP1 or USP1 C90S purified separately from
293T cells. Ubiquitinated ID2 was decreased by wild-type
USP1 but not USP1 C90S (Figure 1D), indicating that deubiquitination was unlikely a consequence of a coeluted protease. The
decrease in ID2 ubiquitination also was sensitive to N-ethylmaleimide, confirming the involvement of a cysteine protease
(Figure 1D). Consistent with ubiquitinated ID2 being a USP1
substrate, deletion mutant USP1D260300 interacted poorly
with ID2 (Figure S1C) and did not enhance ID2 abundance
(Figure S1D).
USP1 and ID2 Are Coordinately Overexpressed in
a Subset of Primary Osteosarcoma Tumors
To identify a biological context in which USP1 deubiquitinates ID
proteins, we examined USP1 expression patterns. Microarray
analyses of healthy and diseased human tissues revealed that
osteosarcoma tumors expressed more USP1 mRNA than
healthy or osteoarthritic bone biopsies (Figure 2A). Western blotting of a separate set of primary human osteosarcoma biopsies
found that USP1 was elevated in 7 of 14 osteosarcomas when
compared to 3 normal primary human osteoblast samples (Figure 2B). Strikingly, ID2 protein abundance in these primary
human tumor samples correlated well with USP1 abundance.
One anomalous sample contained abundant USP1 but little
ID2 (Figure 2B, lane 6), perhaps due to poor expression of the
USP1 cofactor WDR48. Another sample contained abundant
ID2 and little USP1 (Figure 2B, lane 16), which may reflect
reduced ID2 ubiquitination or that other DUBs are active.
The amount of USP1 protein in the primary osteosarcomas
correlated largely with USP1 mRNA abundance (Figure 2C),
suggesting that elevated USP1 in osteosarcoma is due to transcriptional upregulation. In contrast, ID2 protein and mRNA
levels correlated poorly (Figure 2D). The coincident overexpression of USP1 and ID2 in primary osteosarcoma was confirmed by
immunohistochemistry (Figures 2E2G). These results strongly
suggest that USP1 modifies ID proteins posttranslationally in
osteosarcoma.
USP1 Stabilizes ID Proteins in Osteosarcoma
We also assessed USP1 abundance and ID2 stability in human
osteosarcoma cell lines and in primary osteoblasts (Figure S2A).
In U2-OS osteosarcoma cells, USP1 was elevated, and the normally labile ID2 was stable (Figures S2A and S2B). Knockdown of
USP1 with two distinct USP1 shRNAs caused a reduction in ID1,
ID2, and ID3 but had no effect on ID4 (Figure 3A). ID1, ID2, and
ID3 mRNAs were not reduced, excluding decreased transcription as the reason for the drop in ID protein abundance (Figure S3I). USP1 knockdown specificity was confirmed with
shRNA-resistant USP1, which restored ID1, ID2, and ID3 to basal
levels. USP1 catalytic activity was essential for ID stability
because shRNA-resistant USP1 C90S did not restore ID protein
levels. Similar results were observed in osteosarcoma cell lines
HOS, SAOS, and SJSA (Figure S2C). USP1 knockdown did not
impact ID2 abundance in MG-63 osteosarcoma cells, likely
because these cells express very little WDR48 (Figure S2C).
Consistent with WDR48 deficiency limiting USP1 activity in
MG-63 cells, ectopic WDR48 increased ID2 (Figure S2D).
Cell 146, 918930, September 16, 2011 2011 Elsevier Inc. 919
40
80
120
Time (min)
D
WDR48
ID2-FLAG
USP1
T
90
S
USP1
USP1
T
90
S
USP1
IB-FLAG
WT
C90S
USP1
ID3-FLAG
CTL
ID2-FLAG
17 18 19 20 21 22 23 24
25%
9 10 11 12 13 14 15 16
MG-132
WT
C90S
CTL
WT
C90S
CTL
USP1
MG-132
WT
C90S
CTL
ID1-FLAG
MG-132
MG-132
50%
- - - + + +
+ + + + + +
188
WB: Flag
(ID/IB)
WDR48
ID2-Ub
NEM
T
90
S
W
T
WB: Actin
75%
- + + +
+ + + +
- - - +
C
WB: USP1
CTL
USP1 WT
USP1 C90S
100%
WB: ID2
USP1 C90S
0
7
15
30
60
12
0
24
0
48
0
0
7
15
30
60
12
0
24
0
48
0
CHX (min)
USP1 WT
0
7
15
30
60
12
0
24
0
48
0
CTL
98
IP: FLAG/
62
WB:
HA-Ub
WB: USP1
Poly-Ub
49
38
WB: Tub
1
10 11 12
13
14
15
28
16
WB:
HA-Ub
5
4
3
17
IP: FLAG/
WB:
ID2-FLAG
WB: USP1
WB: USP1
WB: WDR48
WB: WDR48
WB: ID2-FLAG
WB: ID2
1
WB: Tub
1
6000
A B C
5000
Primary Osteosarcomas
28271
28272
28273
28274
28275
28276
28277
7000
4000
WB: USP1
3000
WB: ID2
2000
WB: WDR48
1000
Osteoblastic
Osteosarcoma
Osteoarthritis
10
11 12 13 14 15 16 17
8
7
6
5
4
3
2
1
0
A B C
A B C
293T-ID2shRNA
293T-ID2
Normal
Bone
WB: Actin
1
2915
4384
4898
4931
10047
10049
10050
28271
28272
28273
28274
28275
28276
28277
2915
4384
4898
4931
10047
10049
10050
28271
28272
28273
28274
28275
28276
28277
8000
2915
4384
4898
4931
10047
10049
10050
9000
Primary
Osteoblasts
Control
USP1
ID2
Figure 2. USP1 Is Overexpressed in Osteosarcoma and Correlates with ID2 Protein Expression
(A) Box and whisker plots of USP1 mRNA expression in primary human bone biopsies from normal and diseased tissue.
(B) Western blot (WB) analysis of USP1 and ID2 protein expression in primary human osteoblasts and osteosarcoma tumor samples.
(C and D) RT-PCR quantification of USP1 (C) and ID2 (D) expression in the samples in (B). Bars represent the mean SD of triplicate observations.
(E and F) Immunohistochemical detection of ID2 in 293T cells transfected with an ID2 expression vector (top panel) or an ID2 shRNA (bottom panel) (E) or in
a primary human osteosarcoma biopsy (F).
(G) Immunohistochemical staining of USP1 and ID2 in serial sections from primary osteosarcoma tissue. Control staining was with an isotype-control antibody.
shUSP1-B
shUSP1-A
shCTL
shRes
USP1
CTL
WT
- - + - - +
- + - - + + - - + - -
shRes
USP1
C90S
- - +
- + + - -
C
14
WB: ID1
WB: ID2
WB: ID3
WB: ID4
WB: E47
WB: USP1
CTL
USP1-WT
USP1-C90S
12
MG-132
10
WB: ID1
WB: ID2
WB: ID3
WB: USP1
WB: Tub
WB: Tub
MG-132 -
IP
IP: ID2-FLAG
IB: ID2-FLAG
- + +
IP: ID2-FLAG
IB: HA-Ub
WB: ID2
WB: USP1
WB: USP1
WB: ID2
IP
-
IgG
ID2
Input
IgG
USP1
Input
shCTL
shUSP1
shCTL
shUSP1
sh
sh
sh USP1 USP1
B
CTL
A
- + - + - +
WB: GAPDH
WB: Tub
1
1
IB: USP1
IB: ID2-FLAG
IB: GAPDH
1
of CDKN1A, was not increased by USP1 knockdown, suggesting that increased p21 was p53 independent.
p21 is a potent inhibitor of cell cycle progression (Polyak et al.,
1996), so we assessed the proliferative capacity of U2-OS cells
following USP1 knockdown. Consistent with increased p21,
USP1 knockdown reduced U2-OS cell proliferation (Figure 4B
and Figure S3A). shRNA-resistant wild-type USP1, but neither
USP1 C90S nor USP1D260300, restored cell proliferation
(Figures S3B and S3C), indicating that both USP1 catalytic activity
and ID substrate recognition are required to maintain U2-OS cell
proliferation. USP1 knockdown similarly reduced proliferation in
HOS, SAOS, and SJSA, but not MG-63 osteosarcoma cells (Figure S3D). Flow cytometric analysis of the DNA content in U2-OS
cells after USP1 knockdown revealed a moderate increase in cells
922 Cell 146, 918930, September 16, 2011 2011 Elsevier Inc.
CTL
+
+
-
+
-
+
-
+
-
+
-
+
-
100
shUSP1-B
shUSP1-A
shCtl
shRes
shRes
USP1 WT USP1 C90S
WB: p21
WB: p53
WB: USP1
WB: Tub
80
60
40
Ctl
USP1
WT
siCTL
sip21
50
si-CTL + - + - + si-p21 - + - + - +
40
% Cells in S-phase
WB: p53
WB: ID1
WB: ID2
WB: ID3
G1
G2/M
shCTL shUSP1
WB: ID2
20
WB: ID3
10
WB: USP1
shCTL
shUSP1
shID1,
ID2, ID3
WB: Actin
1
CTL
USP1
ID1,2,3
30
% Cells in S-phase
USP1
C90S
WB: ID1
30
WB: Actin
2
20
WB: p21
WB: USP1
30
10
WB: p21
40
sh
sh
shID1,
CTL USP1 ID2, ID3
shCTL
shUSP1-A
shUSP1-B
60
50
20
1
C
shCTL
shUSP1-A
shUSP1-B
% Cell Cycle Distribution
20
10
shCTL
shUSP1
Cell 146, 918930, September 16, 2011 2011 Elsevier Inc. 923
these data suggest that ID protein stabilization by USP1 in osteosarcoma blocks a normal osteogenic differentiation program.
The potential of USP1 inhibition as a tumor differentiation
strategy was investigated in the 143B osteosarcoma xenograft
model. A doxycycline-induced USP1 shRNA suppressed USP1
expression and reduced ID1 and ID2 in the xenografts (Figure 5C
and Figure S4D). ID3 was not detectable in this setting (data
not shown). USP1 knockdown also reduced 143B tumor
growth (Figure 5D), promoted OSTEONECTIN, RUNX2, SPP1/
OSTEOPONTIN, OSTERIX, and BGLAP/OSTEOCALCIN expression (Figure 5E and Figure S4E), and enhanced ALP activity
(Figure 5F). Remarkably, four of ten USP1-deficient xenograft
tumors achieved stasis and differentiation in situ, displaying
markedly altered cellular morphology and accumulation of
acellular collagenous masses consistent with proto-ossification
(Figure 5G). The tumors that continued to proliferate showed
evidence of escape from knockdown, presumably due to loss
or silencing of the shRNA (Figure S4F). These data indicate
that reducing USP1 is sufficient to initiate an osteogenic differentiation program in osteosarcoma.
Dysregulated USP1 Expression Inhibits hMSC
Differentiation
Next, we determined if USP1 stabilization of the IDs contributes
to normal mesenchymal stem cell maintenance. USP1 was expressed in primary hMSCs but declined steadily as the cells
were cultured in conditions favoring osteoblastic differentiation
(Figure 6A). Consistent with a previous study (Peng et al., 2003),
ID1 and ID2 were induced transiently and then declined as well.
ID3 was not detected (data not shown). These data, together
with a study showing that misregulated ID expression inhibits
osteogenic differentiation (Peng et al., 2004), prompted us
to investigate whether USP1 overexpression disrupts hMSC
differentiation. hMSCs overexpressing USP1 and cultured in
osteogenic differentiation medium expressed abnormally high
levels of ID1 and ID2 (Figure 6B), exhibited low ALP activity (Figure 6C), showed minimal induction of RUNX2, OSTERIX, and
OSTEONECTIN (Figure 6D), and stained poorly with alizarin
red, which reveals mineral deposition that is a classic marker of
osteoblast activity (Figure 6E). These data imply that the hMSCs
overexpressing USP1 failed to differentiate. A similar differentiation defect was observed in hMSCs overexpressing ID2, whereas
hMSCs overexpressing USP1 C90S differentiated similarly to
control cells. Thus, the catalytic activity of USP1 was necessary
and ID stabilization sufficient to inhibit osteogenic differentiation.
Coincident with their apparent failure to differentiate, hMSCs
overexpressing USP1 or ID2 proliferated significantly in the presence of excess osteogenic differentiation factors (Figure 6F). In
contrast, proliferation of control hMSCs, or those expressing
USP1 C90S, slowed as they differentiated in culture. Collectively, our observations suggest that overexpression of USP1
or ID2 is sufficient to block osteoblastic differentiation, promote
retention of stem-like features, and render cells resistant to
differentiation cues.
USP1 Promotes Transformation and Tumor Formation
The ability of USP1 to inhibit mesenchymal stem cell differentiation and sustain proliferation of osteosarcoma cell lines
B
shCTL
WB: E-cadherin
C
-DOX
shUSP1
Fibronectin
+DOX
IHC: USP1
WB: N-cadherin
N-cadherin
WB: Fibronectin
IHC: ID2
WB: USP1
1400
shUSP1
shCTL
E-cadherin
WB: ID2
-DOX
600
400
200
0
12 15 18
Days
16
2
1.5
12
0.5
- + - +
USP1 ID2
- + - + - + - + DOX
ON RX2 OSX OP
+DOX
2
ALP Activity
(nmol pNPP/g protein/hr)
800
+DOX
1000
WB: GAPDH
-DOX
1200
H&E
1.5
0.5
Trichrome
shCTL shUSP1
0 1 3 6 9 Un
WB: USP1
WB: ID2
WB: GAPDH
WB: GAPDH
1
hMSC
hMSC +ODM
hMSC USP1 WT +ODM
hMSC USP1 C90S +ODM
hMSC ID2 +ODM
10
Runx2
Osterix
20
0
CTL
50
CTL
40
USP1 WT
USP1 C90S
CTL
USP1 USP1
WT
C90S
+ODM
ID2
hMSC
hMSC +ODM
hMSC USP1 WT +ODM
hMSC USP1 C90S +ODM
hMSC ID2 +ODM
30
20
10
1
0
40
60
F
ODM
15
20
+ + + +
WB: ID2
WB: ID1
25
ODM
WB: ID1
WB: USP1
+ODM (days)
USP1
ALP Activity
(nmoles pNPP/g protein/hr)
80
ID2
0
Osteonectin
3
6
Culture Period (Days)
(Figure 7D and Figure S6J). These data indicate that the USP1-ID
axis regulating differentiation in osteosarcoma is recapitulated
in normal skeletal development.
DISCUSSION
In this study we show that ID stabilization by USP1 sustains a
significant fraction of human osteosarcomas. USP1 was overexpressed frequently in primary osteosarcomas and osteosarcoma
cell lines (Figure 2), and by deubiquitinating the ID proteins
(Figures 1 and 3), inhibited bHLH-dependent expression of
CDKI p21 (Figure 4) resulting in unchecked cell proliferation
(Figure 5). USP1 overexpression not only was necessary for
the proliferation of several osteosarcoma cell lines, it also was
sufficient to prevent normal mesenchymal cell differentiation,
capturing the cells in a stem-like state (Figure 6). By contrast,
USP1 knockdown in osteosarcoma cell lines reduced expression of mesenchymal stem cell markers and initiated an osteogenic development program (Figure 5). USP1 deficiency in
mice impaired normal osteogenesis and resulted in pronounced
350
340
Usp1-/-
WT
BMD (mg/cm3)
330
320
310
300
290
280
10 mm
Usp1-/-
WT
E
BALP (U/L)
WB: ID2
1 mm
1
0.5
WT
Usp1-/-
12
10
WB: ID1
1 mm
1.5
Usp1-/-
WT
10 mm
2.5
WB: USP1
WB: WDR48
8
6
4
2
WB: GAPDH
1
WT
Usp1-/-
Figure S3). Thus, USP1 overexpression perturbs normal osteoblast differentiation, which is characterized by p53-independent
upregulation of multiple CDKIs (Funato et al., 2001; Kenner et al.,
2004; Matsumoto et al., 1998; Yan et al., 1997; Zhang et al.,
1997). CDKI function often is compromised in osteosarcomas;
CDKN2A/p16INK4a and CDKN2B/p15INK4b gene deletions are
common (Miller et al., 1996; Nielsen et al., 1998), as is gene
inactivation due to promoter methylation (Oh et al., 2006). In
contrast, CDK4, a target of CDKIs, is frequently overexpressed
in osteosarcoma due to gene amplification (Ozaki et al., 2003).
ID-mediated transcriptional repression of p21 represents an
additional oncogenic mechanism in osteosarcoma.
ID protein overexpression has been observed in various
human cancers but has been attributed largely to increased ID
transcription (Perk et al., 2005). For example ID2 is transcriptionally upregulated by the EWS-Ets translocation in Ewings
sarcoma (Nishimori et al., 2002), which is an osteoid tumor
bearing strong resemblance to osteosarcoma. Patients with a
disrupted copy of the RB1 gene are strongly sensitized to development of osteosarcoma (Friend et al., 1986), RB being able to
sequester and inactivate ID2 (Iavarone et al., 1994; Lasorella
et al., 2000). Our study reveals an additional mechanism by
which ID proteins and, in turn, CDKIs can be dysregulated in
osteosarcoma.
928 Cell 146, 918930, September 16, 2011 2011 Elsevier Inc.
were from CytoMix, LLC, and the CHTN. DT-40 cells from K. Patel (MRC-LMB)
were cultured in RPMI with 7% FBS and 3% chicken serum (GIBCO). Where
indicated, cells were treated with 10 mM MG-132 (Calbiochem) or 25 mg/ml
cycloheximide (Sigma-Aldrich). Complete experimental procedures are
available in the Extended Experimental Procedures.
Vectors
USP1 variants, ID1, ID2, and ID3 were in pRK2001 with or without a C-terminal
Flag epitope. A WDR48 expression vector was from OriGene. pMACS was
from Miltenyi Biotec. ID2 and USP1 variants also were subcloned into pQCXIP
(Clontech) or pHUSH.Lenti.Puro (Genentech). Knockdown experiments used
pRS-based shRNA vectors (OriGene) or the pTRIPZ shRNA system (Open
Biosystems).
Transfections
Osteosarcoma cells were transfected with the plasmids indicated in combination with the marker plasmid pMACS using FuGENE 6 (Sigma-Aldrich).
Transfected cells were sorted with MACS-select H-2Kk microbeads (Miltenyi
Biotec). Proteins were extracted in lysis buffer (1% NP-40, 120 mM NaCl,
50 mM Tris-HCl [pH 7.4], and 1 mM EDTA) containing protease inhibitor
cocktail 1 and phosphatase inhibitor cocktails 1 and 2 (Calbiochem). For
dual shRNA/siRNA experiments, cells transfected with plasmids in FuGENE
6 then underwent nucleofection. DT-40 cells were transfected by
nucleofection.
Antibodies
Rat 5E10 anti-USP1 and 9F10 anti-WDR48 antibodies (Genentech) were
raised against the C-terminal 100 amino acids of the human proteins. Other
antibodies recognized ID1, ID2, ID3, E47, and p53 (Santa Cruz Biotechnology),
GAPDH (Assay Designs), Flag, HA, tubulin, and actin (Sigma-Aldrich), and p21
(Cell Signaling Technologies). Immunoprecipitations included 10 mM MG-132
and used protein A/G agarose beads (Pierce). Abcam AB52093 anti-ID2
antibody was used for immunohistochemistry in Figures 2E and 2F.
Gene Expression
RNA was extracted with RNeasy kits (QIAGEN) and amplified with the QuantiTect SYBR Green RT-PCR system (QIAGEN). Primary tumor RNA data were
obtained from GeneLogic microarray analyses (Ocimum Biosolutions) with
HGU133P Affymetrix chips.
Flow Cytometry
Cells staining with FITC-conjugated anti-H-2Kk antibodies (Miltenyi Biotec)
were stained with propidium iodide (Sigma-Aldrich) in the presence of RNase
A (Sigma-Aldrich), and their DNA content was measured in a FACSCalibur
flow cytometer (BD Biosciences). Other antibodies used recognized CD90
(Chemicon), CD105 (R&D Systems), CD106 (SouthernBiotech), and CD144
(eBioscience).
Immunohistochemistry
Formalin-fixed, paraffin-embedded tissue sections were incubated in target
retrieval solution (Dako) at 99 C for 20 min and cooled to 74 C for 20 min.
Sections were blocked at room temperature in avidin/biotin blocking buffer
(Vector Labs) and then 3% BSA for 30 min. Staining with primary antibody
was at room temperature for 60 min. Sections were rinsed twice in Dako
wash buffer and then incubated in VECTASTAIN reagents (Vector Labs) for
30 min. Staining was revealed with peroxidase substrate buffer (Pierce), then
sections were counterstained with Mayers Hematoxylin. TRAP staining
(Sigma-Aldrich) was on 5 mm sections from formalin-fixed, paraffin-embedded
P12 mouse femurs. TRAP-positive osteoclasts were counted in ten fields.
Alizarin red staining was performed according to manufacturer instructions
(Ricca Chemical).
Immunofluorescence Staining
Stably transduced U2-OS cells treated with 3 mg/ml doxycycline (Clontech) for
14 days were fixed in 1% PFA in PBS and stained with antibodies recognizing
E-cadherin, N-cadherin (BD), or fibronectin (Calbiochem).
In Vivo Deubiquitination
Transfected 293T cells were cultured with 10 mM MG-132 for 30 min prior to
extraction in lysis buffer containing 10 mM MG-132 and 10 mM N-ethylmaleimide (Sigma-Aldrich). Soluble lysates were denatured with 1% SDS at 95 C
for 5 min, then diluted 20-fold in lysis buffer prior to immunoprecipitation
with M2 anti-Flag agarose (Sigma-Aldrich).
In Vitro Deubiquitination
293T cells were transfected with USP1-Flag, USP1-C90S-Flag, or ID2-Flag
and HA-ubiquitin. ID2-Flag was immunoprecipitated from SDS-denatured
cell lysate. USP1 was immunoprecipitated from regular cell lysate. Elutions
were with 500 mg/ml 3X Flag peptide (Sigma-Aldrich). Samples were combined
in DUB buffer (20 mM HEPES [pH 8.3], 20 mM NaCl, 100 mg/ml BSA, 500 mM
EDTA, 1 mM DTT) at room temperature.
Osteosarcoma Differentiation
U2-OS, HOS, or SAOS cells transfected serially with pRS.shUSP1 (or shCTL)
were selected based on H-2Kk expression and cultured for 12 days. 143B cells
transduced with pTRIPZ inducible USP1 shRNA were selected in puromycin.
ALP activity was determined in cell lysates lacking phosphatase inhibitors after
normalizing for protein content. p-nitrophenyl phosphatase cleavage was
analyzed by colorimetric assay (Sigma-Aldrich).
3T3 Transformation
Transduced 3T3 cells were plated in DMEM/0.5% low-melting agar on 1%
agar. Colonies of eight or more cells were scored visually after 21 days. For
ID knockdown, USP1-expressing 3T3 cells were transduced with pTRIPZ
inducible shRNAs and cultured with 3 mg/ml doxycycline for 72 hr prior to
embedding in agar containing 3 mg/ml doxycycline.
Mice
Eight-week-old female NCr nude mice (Taconic) or C.B-17 SCID.bg mice
(CRL) were injected subcutaneously in the right hind flank with 1 3 106 3T3
cells or 2.5 3 106 143B cells in 100 ml HBSS and provided 1 mg/ml doxycycline
in 5% sucrose water. Usp1 gene-targeted C57BL/6 ES cells were from the
Knockout Mouse Project Repository (Davis, CA). Microcomputed tomography
was a mCT 40 (SCANCO Medical). Amniotic fluid from E18.5 embryos was
assayed for deoxypyridinoline (TSZ ELISA), BALP (EIAab & USCNLIFE), and
creatinine (R&D Systems). The Genentech Institutional Animal Care and Use
Committee approved all animal studies.
ACCESSION NUMBERS
mRNA expression data of USP1 were extracted from the Gene Logic expression database of Affymetrix Human Genome U133 Plus 2.0 data. The probe set
202412_s_at was chosen to represent the expression of USP1.
REFERENCES
Bounpheng, M.A., Dimas, J.J., Dodds, S.G., and Christy, B.A. (1999).
Degradation of Id proteins by the ubiquitin-proteasome pathway. FASEB J.
13, 22572264.
Ciarapica, R., Annibali, D., Raimondi, L., Savino, M., Nasi, S., and Rota, R.
(2009). Targeting Id protein interactions by an engineered HLH domain induces
human neuroblastoma cell differentiation. Oncogene 28, 18811891.
Cohn, M.A., Kowal, P., Yang, K., Haas, W., Huang, T.T., Gygi, S.P., and
DAndrea, A.D. (2007). A UAF1-containing multisubunit protein complex
regulates the Fanconi anemia pathway. Mol. Cell 28, 786797.
DAndrea, A.D. (2010). Susceptibility pathways in Fanconis anemia and breast
cancer. N. Engl. J. Med. 362, 19091919.
Di Fiore, R., Santulli, A., Ferrante, R.D., Giuliano, M., De Blasio, A., Messina, C.,
Pirozzi, G., Tirino, V., Tesoriere, G., and Vento, R. (2009). Identification and
expansion of human osteosarcoma-cancer-stem cells by long-term 3-aminobenzamide treatment. J. Cell. Physiol. 219, 301313.
Dupont, S., Mamidi, A., Cordenonsi, M., Montagner, M., Zacchigna, L.,
Adorno, M., Martello, G., Stinchfield, M.J., Soligo, S., Morsut, L., et al.
(2009). FAM/USP9x, a deubiquitinating enzyme essential for TGFbeta
signaling, controls Smad4 monoubiquitination. Cell 136, 123135.
Friend, S.H., Bernards, R., Rogelj, S., Weinberg, R.A., Rapaport, J.M., Albert,
D.M., and Dryja, T.P. (1986). A human DNA segment with properties of the
gene that predisposes to retinoblastoma and osteosarcoma. Nature 323,
643646.
Funato, N., Ohtani, K., Ohyama, K., Kuroda, T., and Nakamura, M. (2001).
Common regulation of growth arrest and differentiation of osteoblasts by
helix-loop-helix factors. Mol. Cell. Biol. 21, 74167428.
Hanahan, D., and Weinberg, R.A. (2000). The hallmarks of cancer. Cell 100,
5770.
Huang, T.T., Nijman, S.M., Mirchandani, K.D., Galardy, P.J., Cohn, M.A., Haas,
W., Gygi, S.P., Ploegh, H.L., Bernards, R., and DAndrea, A.D. (2006).
Regulation of monoubiquitinated PCNA by DUB autocleavage. Nat. Cell
Biol. 8, 339347.
Iavarone, A., Garg, P., Lasorella, A., Hsu, J., and Israel, M.A. (1994). The
helix-loop-helix protein Id-2 enhances cell proliferation and binds to the
retinoblastoma protein. Genes Dev. 8, 12701284.
Kastan, M.B., Onyekwere, O., Sidransky, D., Vogelstein, B., and Craig, R.W.
(1991). Participation of p53 protein in the cellular response to DNA damage.
Cancer Res. 51, 63046311.
Kenner, L., Hoebertz, A., Beil, T., Keon, N., Karreth, F., Eferl, R., Scheuch, H.,
Szremska, A., Amling, M., Schorpp-Kistner, M., et al. (2004). Mice lacking
JunB are osteopenic due to cell-autonomous osteoblast and osteoclast
defects. J. Cell Biol. 164, 613623.
Kim, D., Peng, X.C., and Sun, X.H. (1999). Massive apoptosis of thymocytes in
T-cell-deficient Id1 transgenic mice. Mol. Cell. Biol. 19, 82408253.
SUPPLEMENTAL INFORMATION
Kim, J.M., Parmar, K., Huang, M., Weinstock, D.M., Ruit, C.A., Kutok, J.L., and
DAndrea, A.D. (2009). Inactivation of murine Usp1 results in genomic instability and a Fanconi anemia phenotype. Dev. Cell 16, 314320.
Lasorella, A., Noseda, M., Beyna, M., Yokota, Y., and Iavarone, A. (2000). Id2 is
a retinoblastoma protein target and mediates signalling by Myc oncoproteins.
Nature 407, 592598.
ACKNOWLEDGMENTS
Lasorella, A., Uo, T., and Iavarone, A. (2001). Id proteins at the cross-road of
development and cancer. Oncogene 20, 83268333.
We thank Kim Newton for editorial assistance. All authors were employees of
Genentech, Inc.
Lasorella, A., Stegmuller, J., Guardavaccaro, D., Liu, G., Carro, M.S.,
Rothschild, G., de la Torre-Ubieta, L., Pagano, M., Bonni, A., and Iavarone,
A. (2006). Degradation of Id2 by the anaphase-promoting complex couples
cell cycle exit and axonal growth. Nature 442, 471474.
Lassar, A.B., Davis, R.L., Wright, W.E., Kadesch, T., Murre, C., Voronova, A.,
Baltimore, D., and Weintraub, H. (1991). Functional activity of myogenic HLH
proteins requires hetero-oligomerization with E12/E47-like proteins in vivo.
Cell 66, 305315.
Cell 146, 918930, September 16, 2011 2011 Elsevier Inc. 929
Luo, X., Chen, J., Song, W.X., Tang, N., Luo, J., Deng, Z.L., Sharff, K.A., He, G.,
Bi, Y., He, B.C., et al. (2008). Osteogenic BMPs promote tumor growth of
human osteosarcomas that harbor differentiation defects. Lab. Invest. 88,
12641277.
Peng, Y., Kang, Q., Luo, Q., Jiang, W., Si, W., Liu, B.A., Luu, H.H., Park, J.K., Li,
X., Luo, J., et al. (2004). Inhibitor of DNA binding/differentiation helix-loop-helix
proteins mediate bone morphogenetic protein-induced osteoblast differentiation of mesenchymal stem cells. J. Biol. Chem. 279, 3294132949.
Lyden, D., Young, A.Z., Zagzag, D., Yan, W., Gerald, W., OReilly, R., Bader,
B.L., Hynes, R.O., Zhuang, Y., Manova, K., and Benezra, R. (1999). Id1 and
Id3 are required for neurogenesis, angiogenesis and vascularization of tumour
xenografts. Nature 401, 670677.
Maeda, Y., Tsuji, K., Nifuji, A., and Noda, M. (2004). Inhibitory helix-loop-helix
transcription factors Id1/Id3 promote bone formation in vivo. J. Cell. Biochem.
93, 337344.
Massari, M.E., and Murre, C. (2000). Helix-loop-helix proteins: regulators of
transcription in eucaryotic organisms. Mol. Cell. Biol. 20, 429440.
Polyak, K., Waldman, T., He, T.C., Kinzler, K.W., and Vogelstein, B. (1996).
Genetic determinants of p53-induced apoptosis and growth arrest. Genes
Dev. 10, 19451952.
Prabhu, S., Ignatova, A., Park, S.T., and Sun, X.H. (1997). Regulation of the
expression of cyclin-dependent kinase inhibitor p21 by E2A and Id proteins.
Mol. Cell. Biol. 17, 58885896.
Matsumoto, T., Sowa, Y., Ohtani-Fujita, N., Tamaki, T., Takenaka, T.,
Kuribayashi, K., and Sakai, T. (1998). p53-independent induction of WAF1/
Cip1 is correlated with osteoblastic differentiation by vitamin D3. Cancer
Lett. 129, 6168.
Roa, S., Avdievich, E., Peled, J.U., Maccarthy, T., Werling, U., Kuang, F.L.,
Kan, R., Zhao, C., Bergman, A., Cohen, P.E., et al. (2008). Ubiquitylated
PCNA plays a role in somatic hypermutation and class-switch recombination
and is required for meiotic progression. Proc. Natl. Acad. Sci. USA 105,
1624816253.
Miller, C.W., Aslo, A., Campbell, M.J., Kawamata, N., Lampkin, B.C., and
Koeffler, H.P. (1996). Alterations of the p15, p16,and p18 genes in osteosarcoma. Cancer Genet. Cytogenet. 86, 136142.
Rogakou, E.P., Boon, C., Redon, C., and Bonner, W.M. (1999). Megabase
chromatin domains involved in DNA double-strand breaks in vivo. J. Cell
Biol. 146, 905916.
Nielsen, G.P., Burns, K.L., Rosenberg, A.E., and Louis, D.N. (1998). CDKN2A
gene deletions and loss of p16 expression occur in osteosarcomas that lack
RB alterations. Am. J. Pathol. 153, 159163.
Sharff, K.A., Song, W.X., Luo, X., Tang, N., Luo, J., Chen, J., Bi, Y., He, B.C.,
Huang, J., Li, X., et al. (2009). Hey1 basic helix-loop-helix protein plays an
important role in mediating BMP9-induced osteogenic differentiation of
mesenchymal progenitor cells. J. Biol. Chem. 284, 649659.
Nijman, S.M., Huang, T.T., Dirac, A.M., Brummelkamp, T.R., Kerkhoven, R.M.,
DAndrea, A.D., and Bernards, R. (2005). The deubiquitinating enzyme USP1
regulates the Fanconi anemia pathway. Mol. Cell 17, 331339.
Nishimori, H., Sasaki, Y., Yoshida, K., Irifune, H., Zembutsu, H., Tanaka, T.,
Aoyama, T., Hosaka, T., Kawaguchi, S., Wada, T., et al. (2002). The Id2 gene
is a novel target of transcriptional activation by EWS-ETS fusion proteins in
Ewing family tumors. Oncogene 21, 83028309.
Oestergaard, V.H., Langevin, F., Kuiken, H.J., Pace, P., Niedzwiedz, W.,
Simpson, L.J., Ohzeki, M., Takata, M., Sale, J.E., and Patel, K.J. (2007).
Deubiquitination of FANCD2 is required for DNA crosslink repair. Mol. Cell
28, 798809.
Oh, J.H., Kim, H.S., Kim, H.H., Kim, W.H., and Lee, S.H. (2006). Aberrant
methylation of p14ARF gene correlates with poor survival in osteosarcoma.
Clin. Orthop. Relat. Res. 442, 216222.
Ozaki, T., Neumann, T., Wai, D., Schafer, K.L., van Valen, F., Lindner, N.,
Scheel, C., Bocker, W., Winkelmann, W., Dockhorn-Dworniczak, B., et al.
(2003). Chromosomal alterations in osteosarcoma cell lines revealed by
comparative genomic hybridization and multicolor karyotyping. Cancer Genet.
Cytogenet. 140, 145152.
Palombella, V.J., Rando, O.J., Goldberg, A.L., and Maniatis, T. (1994). The
ubiquitin-proteasome pathway is required for processing the NF-kappa B1
precursor protein and the activation of NF-kappa B. Cell 78, 773785.
Parmar, K., Kim, J., Sykes, S.M., Shimamura, A., Stuckert, P., Zhu, K.,
Hamilton, A., Deloach, M.K., Kutok, J.L., Akashi, K., et al. (2010). Hematopoietic stem cell defects in mice with deficiency of Fancd2 or Usp1. Stem Cells 28,
11861195.
Peng, Y., Kang, Q., Cheng, H., Li, X., Sun, M.H., Jiang, W., Luu, H.H., Park,
J.Y., Haydon, R.C., and He, T.C. (2003). Transcriptional characterization of
bone morphogenetic proteins (BMPs)-mediated osteogenic signaling. J.
Cell. Biochem. 90, 11491165.
930 Cell 146, 918930, September 16, 2011 2011 Elsevier Inc.
Soignet, S.L., Maslak, P., Wang, Z.G., Jhanwar, S., Calleja, E., Dardashti, L.J.,
Corso, D., DeBlasio, A., Gabrilove, J., Scheinberg, D.A., et al. (1998). Complete
remission after treatment of acute promyelocytic leukemia with arsenic
trioxide. N. Engl. J. Med. 339, 13411348.
Stock, C., Kager, L., Fink, F.M., Gadner, H., and Ambros, P.F. (2000). Chromosomal regions involved in the pathogenesis of osteosarcomas. Genes
Chromosomes Cancer 28, 329336.
Suh, J.H., Lee, H.W., Lee, J.W., and Kim, J.B. (2008). Hes1 stimulates
transcriptional activity of Runx2 by increasing protein stabilization during
osteoblast differentiation. Biochem. Biophys. Res. Commun. 367, 97102.
Tang, N., Song, W.X., Luo, J., Haydon, R.C., and He, T.C. (2008). Osteosarcoma development and stem cell differentiation. Clin. Orthop. Relat. Res.
466, 21142130.
Thiery, J.P., Acloque, H., Huang, R.Y., and Nieto, M.A. (2009). Epithelialmesenchymal transitions in development and disease. Cell 139, 871890.
Tupler, R., Perini, G., and Green, M.R. (2001). Expressing the human genome.
Nature 409, 832833.
Weintraub, H., Genetta, T., and Kadesch, T. (1994). Tissue-specific gene activation by MyoD: determination of specificity by cis-acting repression
elements. Genes Dev. 8, 22032211.
Yan, Y., Frisen, J., Lee, M.H., Massague, J., and Barbacid, M. (1997). Ablation
of the CDK inhibitor p57Kip2 results in increased apoptosis and delayed
differentiation during mouse development. Genes Dev. 11, 973983.
Yokota, Y., and Mori, S. (2002). Role of Id family proteins in growth control. J.
Cell. Physiol. 190, 2128.
Zhang, P., Liegeois, N.J., Wong, C., Finegold, M., Hou, H., Thompson, J.C.,
Silverman, A., Harper, J.W., DePinho, R.A., and Elledge, S.J. (1997). Altered
cell differentiation and proliferation in mice lacking p57KIP2 indicates a role
in Beckwith-Wiedemann syndrome. Nature 387, 151158.
of Biological Chemistry and Molecular Pharmacology, Harvard Medical School, Boston, MA 02115, USA
of Pharmacological Sciences and Chemistry, Stony Brook University, Stony Brook, NY 11794, USA
3Program of Molecular Biology, Memorial Sloan Kettering Cancer Center, 1275 York Avenue, New York, NY 10021
4Present address: Tecan Group Ltd., 8708 Ma
nnedorf, Switzerland
5Present address: Helsinn Therapeutics (U.S.) Inc., Bridgewater, NJ 08807, USA
6Present address: The Zernike Institute for Advanced Materials, University of Groningen, 9747 AG Groningen, The Netherlands
7These authors contributed equally to this work
*Correspondence: johannes_walter@hms.harvard.edu
DOI 10.1016/j.cell.2011.07.045
2Department
SUMMARY
The eukaryotic replicative DNA helicase, CMG, unwinds DNA by an unknown mechanism. In some
models, CMG encircles and translocates along one
strand of DNA while excluding the other strand. In
others, CMG encircles and translocates along duplex
DNA. To distinguish between these models, replisomes were confronted with strand-specific DNA
roadblocks in Xenopus egg extracts. An ssDNA translocase should stall at an obstruction on the translocation strand but not the excluded strand, whereas
a dsDNA translocase should stall at obstructions on
either strand. We found that replisomes bypass large
roadblocks on the lagging strand template much
more readily than on the leading strand template.
Our results indicate that CMG is a 30 to 50 ssDNA translocase, consistent with unwinding via steric exclusion. Given that MCM2-7 encircles dsDNA in G1, the
data imply that formation of CMG in S phase involves
remodeling of MCM2-7 from a dsDNA to a ssDNA
binding mode.
INTRODUCTION
In eukaryotic cells, DNA replication initiates at many chromosomal locations called origins. At each origin, two sister replisomes are assembled that move away from the origin in opposite
directions. An essential component of the replisome is the replicative DNA helicase, which unwinds parental DNA, generating
substrates for leading and lagging strand DNA polymerases.
Current evidence indicates that the eukaryotic replicative DNA
helicase contains at least three components, a heterohexameric
ATPase called MCM2-7 and two cofactors, Cdc45 and GINS
(Bochman and Schwacha, 2009; Ilves et al., 2010; Moyer et al.,
pICLInter
Stu I (-268)
0
10
20
30
5-CCCTCTTCCGCTCTTCTTTCGTGCGCGGCCGCGATCCGCTGCATT
GAGAAGAAAGCACGCGCCGGCGCTAGGCGACGTAA
AGGCCT-3
TCCGGA
3-GGGAGAAGGCGAGAAGAAAGCACGCGCCGGCGCTAGGCGACGTAA
pICLIntra
10
20
TCCGGA-5
Stu I (-268)
30
5-CCCTCTCCTTGGTTCTTCTCGTGCGCGGCCGCGATCCGCTGCATT
CAAGAAGAGCACGCGCCGGCGCTAGGCGACGTAA
AGGCCT-3
TCCGGA
3-GGGAGAGGAACCAAGAAGAGCACGCGCCGGCGCTAGGCGACGTAA
pICLInter
pICLIntra
Primer M
TCCGGA-5
0
-1
10
20
-20
30
40
-40
50
6 7
9 10 11 12 13 14 15 16
17 18 19 20
Relative amount
-20 Arrest
100
-1 Arrest
Cell 146, 931941, September 16, 2011 2011 Elsevier Inc. 933
To test this idea, we performed chromatin immunoprecipitation (ChIP) for Mcm7, a CMG subunit. pICLInter was replicated,
and at different times, the reaction was crosslinked with formaldehyde, sonicated, immunoprecipitated with Mcm7 antibody,
and the recovered DNA amplified with ICL-proximal and control
primer pairs (Figure 1D). ChIP revealed that soon after replication
began, Mcm7 was depleted from the control region, as expected if CMG travels toward the ICL while displacing any latent
MCM2-7 complexes (Figure 1D, purple triangles). In contrast,
Mcm7 initially accumulated at the ICL, concurrent with the arrival
of leading strands at the 20 position (Figure 1D, compare pink
circles and blue diamonds). The subsequent disappearance of
Mcm7 from the ICL after 10 min correlated with the disappearance of the 20 to 40 leading strand cluster (Figure 1D), and
the advance of the leading strand to the 1 position (Figure 1D,
gray squares). These data indicate that arrest of the leading
strand 20-40 nt from a cisplatin ICL is caused by the footprint
of CMG on DNA (and any dead volume of the DNA polymerase)
and that dissociation of CMG facilitates resumption of leading
strand synthesis toward the ICL.
Converging Replisomes Do Not Interfere with Each
Other at an ICL
We postulated that the wide range of leading strand arrest points
near a DNA inter-strand crosslink (Figure 1C, red bracket) could
reflect heterogeneity in the CMG footprint and/or interference by
converging CMG complexes. In the latter view, the first CMG to
arrive at the ICL would impose a more distal stoppage point on
the second replisome, giving rise to the distribution of leading
strand products. Such interference might be particularly pronounced if CMG travels along dsDNA, since an ICL might
enter deep into the central channel of CMG, allowing the first
replisome to prevent approach of the second replisome
(Figure S2A, top).
To test whether replisome interference occurs, we designed
a plasmid, pICLLead/Lag, in which the arrival of replication forks
on one side of an ICL can be blocked. pICLLead/Lag contains
a nitrogen mustard (NM)-like ICL [which causes a 24 position
arrest (Raschle et al., 2008)], as well as four biotinylated thymidine nucleotides placed 34-40 nt to the right of the ICL, two on
the leading strand template and two on the lagging strand
template of the leftward replication fork (Figure 2A). In the presence of SA, the leftward DNA replication fork should not be able
to reach the ICL. To test whether this is the case, pICLLead/Lag
was preincubated with and without SA and then replicated in
the presence of [a32P]dATP. At different times, DNA was digested with StuI and nascent strands of the leftward replication
fork were visualized on sequencing gels (Figure 2B). In the
absence of SA, the leftward leading strand initially advanced to
within 24 nt of the ICL (Figure 2B, lane 1, red bracket), after which
it crept forward, nearly reaching the ICL (Figure 2B, lanes 5, 7, 9,
11, 13, green arrow), as expected based on our previous results
(Raschle et al., 2008). In the presence of SA, the 24 cluster was
largely absent, and we observed a new set of products starting
at the 70 position (Figure 2B, lanes 2, 4, 6, 8, 10, 12, orange
bracket), which is 30 nt from the outermost biotin-SA complex
at the 40 position (Figure 2B, bottom white arrow on DNA
sequencing ladder). The arrest of the leading strand 30 nt from
934 Cell 146, 931941, September 16, 2011 2011 Elsevier Inc.
AflIII (151)
Stu I (-309)
ICL
-35 -40
-70
Primer M
rS
Rightward
Fork
AflIII (-581)
Prime
Leftward
Fork
-34 -39
Biotin-SA
Primer M
G A T C
10
15
20
25
30
60
: Time (min)
: SA
-1
ICL 0
-10
-20
-24
-30
-40
-41
-50
-50
-60
-70
-70
-80
-80
-90
1 2
3 4
5 6
7 8
9 10
11 12
13 14
10
15
20
25
30
60
Primer S
G A T C
: Time (min)
+ : SA
-1
ICL 0
10
20
-24
30
40
50
-50
1
3 4
5 6
9 10
11 12
13 14
Cell 146, 931941, September 16, 2011 2011 Elsevier Inc. 935
3 to 5 ssDNA translocation
Position (nt): 0
-24
-40
dsDNA translocation
Bio-SA
A
pICLLead/Lag
-24
-70
pICLLead/Lag
-70
Bio-SA
ICL
-40
ICL
pICLLead
pICLLead
ICL
ICL
pICLLag
pICLLag
ICL
10
15
ICL
20
30
: Time (min)
: Plasmid
Primer M
G A T
-1
ICL 0
: SA
-10
-20
- 24
-30
-40
- 41
-50
- 50
-60
-70
- 70
-80
- 80
-90
9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24
Figure 3. Biotin-SA Complexes Located on the Leading Strand Template But Not on the Lagging Strand Template Arrest the Replisome
(AF) The 30 to 50 ssDNA translocation (AC) and dsDNA translocation models (DF) for CMG make different predictions regarding how the leftward
moving replisome (CMG, green; DNA polymerase, gray) will interact with SA molecules bound to pICLLead/Lag, pICL Lead, or pICLLag (see main text). On all
plasmids, the rightward replisome (not depicted) will be prevented from approaching the biotin-SA complexes by the ICL. The yellow line in (DF) represents
the ploughshare postulated to split the duplex as it emerges from the central channel. (G) pICLLead/Lag, pICL Lead, or pICLLag was preincubated with buffer
or streptavidin, as indicated, and replicated in the presence of [a32P]dATP as in Figure 2B. SA was not displaced from pICLLag during replication
(Figure S3).
Side View
Latent MCM2-7
double hexamer
End-on view
(left hexamer)
G1 Phase
(1)
Split Dimer
(2)
(3)
Extrude strand
S Phase
(4)
Activate ATPase
with GINS/Cdc45
DNA unwinding
EXPERIMENTAL PROCEDURES
Preparation of Plasmids
To make pICLLead/Lag, pICLLead, and pICLLag, various ICL-biotin oligonucleotides (see Supplementary Methods) were purified by polyacrylamide gel
electrophoresis, and ligated into pSVRLuc to form the three plasmids (see Figure 3) (Guainazzi et al., 2010; Raschle et al., 2008). pICLControl contains the
identical sequence as pICLLead/Lag, except for the four biotinylated thymidine
nucleotides. pICLInter and pICLIntra were constructed as previously described
(Raschle et al., 2008; Tremeau-Bravard et al., 2004). Compared to pICLInter
and pICLIntra, the plasmids pICLControl, pICLLead/Lag, pICLLead, and pICLLag
contain a 41 nt insert near the ICL.
Xenopus Egg Extracts and Replication
The preparation of Xenopus egg extracts (NPE and HSS) was as described
(Walter et al., 1998). For DNA replication, the plasmids (75ng/ml) were first
incubated with an equal volume of Streptavidin (5 mg/ml) (SouthernBiotech,
Birmingham, AL, USA) or buffer for 1 hr at room temperature (RT), after which
this mixture was added to HSS for 5 min at 22 C (1215 ng/ml final plasmid
concentration), followed by addition of two volumes of NPE containing
[a32P]dATP. Replication was stopped with Stop Solution (0.5% SDS,
25mM EDTA, 50 mM Tris, [pH 7.5]) at different time points. The purification
of DNA replication products was as described (Raschle et al., 2008).
Nascent Strand Analysis
The nascent strand analysis was carried out as described (Raschle et al.,
2008). Briefly, purified DNA replication products were isolated and digested
with the indicated enzymes. Restriction fragments were separated on 5% or
7% polyacrylamide sequencing gels. Gels were transferred to filter paper,
dried, and nascent strands visualized with a phosphorimager. Sequencing
ladders using primers S and M (see Figure 2A) were generated using the
Cell 146, 931941, September 16, 2011 2011 Elsevier Inc. 939
Cycle Sequencing kit from USB (USB Corporation, Cleveland, OH, USA).
Quantification was performed by Image Gauge V4.22 (Fuji Photo Film Corporation, Tokyo, Japan).
Chromatin Immunoprecipitation
Chromatin immunoprecipitation (ChIP) was modified from our existing procedure (Pacek et al., 2006). After the addition of NPE, aliquots of the reaction
were crosslinked through the addition of 1% formaldehyde in ELB. After
10 min incubation at RT, glycine was added to stop the crosslinking reaction.
The crosslinked material was spun through Micro Bio-Spin 6 Chromatography
Columns (BIO-RAD, Hercules, CA, USA) to remove formaldehyde. The flowthrough was diluted with sonication buffer and subjected to sonication. Immonoprecipitation with the indicated antibodies and quantitative real-time PCR
were performed as described (Pacek et al., 2006). Anti-Mcm7 antibody was
previously described (Walter and Newport, 2000); anti-Streptavidin antibody
was purchased from Spring Bioscience (Spring Bioscience, Pleasanton,
CA, USA).
Single-Molecule Assays
A QDot was attached to the top strand 15 kb from the right end of l DNA
or to the bottom strand 19 kb away from the end of l DNA through modification of a previous procedure (Kochaniak et al., 2009; Kuhn and FrankKamenetskii, 2008). Anti-digoxigenin antibody (Roche Applied Sciences,
Indianapolis, IN, USA) was attached to QDot605 using Invitrogen QDot
Antibody Conjugation Kit. After immobilizing l DNA in the flow cell, 1 nM
of anti-digoxigenin conjugated QDot was injected. After 10-20 min incubation, the flow cell was washed with buffer and extracts were introduced.
Replication of surface immobilized l DNA from single initiations in Xenopus
egg extracts and visualization of replicated products was performed as
described (Yardimci et al., 2010). SYTOX, fluorescein labeled anti-dig, and
the QDot were imaged using 568 nm, 488 nm, and 405 nm laser light,
respectively. To assess replication fork arrest near QDots, replication forks
were labeled with dig-dUTP for 25 min. Since the average fork rate in
these experiments was 268 bp/minute (Yardimci et al., 2010), uninterrupted
replication would yield average dig-dUTP tracts of 6.7 kb (25 min 3
0.268 kb/minute). Therefore, only dig-dUTP tracts that were significantly
shorter than expected (%4.5 kb), which ended at a QDot, were considered
to represent stalled forks.
SUPPLEMENTAL INFORMATION
Supplemental Information contains Extended Experimental Procedures and
four figures and can be found with this article online at doi:10.1016/j.cell.
2011.07.045.
ACKNOWLEDGMENTS
We thank Milica Enoiu for the gift of pICLIntra, Anna Loveland, Puck Knipscheer,
and Tatsuro Takahashi for helpful discussions, and the members of our laboratory for valuable feedback on the manuscript. J.C.W was supported by
National Institutes of Health (NIH) grant GM62267, by a Leukemia and
Lymphoma Scholar Award, and by the American Cancer Society grant RSG08-234-01-GMC. Y.V.F. was supported by a Human Frontier Science Program
Long-Term Fellowship (LT000307/2009). D.T.L was supported by a postdoctoral fellowship (PF-10-146-01-DMC) from the American Cancer Society.
O.D.S. was supported by grants from the New York State Office of Science
and Technology and Academic Research NYSTAR (C040069) and the NIH
(GM08454 and CA092584). A.M.v.O. acknowledges support from the American Cancer Society grant RSG-08-234-01 and Searle Scholarship 05-L-104.
J.H. was supported by NIH grant GM034559 and American Cancer Society
grant RSG-08-234-01-GMC.
Received: December 15, 2010
Revised: May 17, 2011
Accepted: July 29, 2011
Published: September 15, 2011
940 Cell 146, 931941, September 16, 2011 2011 Elsevier Inc.
REFERENCES
Angelov, T., Guainazzi, A., and Scharer, O.D. (2009). Generation of DNA
interstrand cross-links by post-synthetic reductive amination. Org. Lett. 11,
661664.
Arias, E.E., and Walter, J.C. (2004). Initiation of DNA replication in Xenopus egg
extracts. Front. Biosci. 9, 30293045.
Arias, E.E., and Walter, J.C. (2007). Strength in numbers: preventing rereplication via multiple mechanisms in eukaryotic cells. Genes Dev. 21, 497518.
Bochman, M.L., and Schwacha, A. (2008). The Mcm2-7 complex has in vitro
helicase activity. Mol. Cell 31, 287293.
Bochman, M.L., and Schwacha, A. (2009). The Mcm complex: unwinding the
mechanism of a replicative helicase. Microbiol. Mol. Biol. Rev. 73, 652683.
Bowers, J.L., Randell, J.C., Chen, S., and Bell, S.P. (2004). ATP hydrolysis by
ORC catalyzes reiterative Mcm2-7 assembly at a defined origin of replication.
Mol. Cell 16, 967978.
Brewster, A.S., Wang, G., Yu, X., Greenleaf, W.B., Carazo, J.M., Tjajadia, M.,
Klein, M.G., and Chen, X.S. (2008). Crystal structure of a near-full-length
archaeal MCM: functional insights for an AAA+ hexameric helicase. Proc.
Natl. Acad. Sci. USA 105, 2019120196.
Costa, A., Ilves, I., Tamberg, N., Petojevic, T., Nogales, E., Botchan, M.R., and
Berger, J.M. (2011). The structural basis for MCM2-7 helicase activation by
GINS and Cdc45. Nat. Struct. Mol. Biol. 18, 471477.
Enemark, E.J., and Joshua-Tor, L. (2006). Mechanism of DNA translocation in
a replicative hexameric helicase. Nature 442, 270275.
Evrin, C., Clarke, P., Zech, J., Lurz, R., Sun, J., Uhle, S., Li, H., Stillman, B., and
Speck, C. (2009). A double-hexameric MCM2-7 complex is loaded onto origin
DNA during licensing of eukaryotic DNA replication. Proc. Natl. Acad. Sci. USA
106, 2024020245.
Fletcher, R.J., Bishop, B.E., Leon, R.P., Sclafani, R.A., Ogata, C.M., and Chen,
X.S. (2003). The structure and function of MCM from archaeal M. Thermoautotrophicum. Nat. Struct. Biol. 10, 160167.
Gambus, A., Jones, R.C., Sanchez-Diaz, A., Kanemaki, M., van Deursen, F.,
Edmondson, R.D., and Labib, K. (2006). GINS maintains association of
Cdc45 with MCM in replisome progression complexes at eukaryotic DNA
replication forks. Nat. Cell Biol. 8, 358366.
Gambus, A., Khoudoli, G.A., Jones, R.C., and Blow, J.J. (2011). MCM2-7 form
double hexamers at licensed origins in Xenopus egg extract. J. Biol. Chem.
286, 1185511864.
Geraghty, D.S., Ding, M., Heintz, N.H., and Pederson, D.S. (2000). Premature
structural changes at replication origins in a yeast minichromosome maintenance (MCM) mutant. J. Biol. Chem. 275, 1801118021.
Guainazzi, A., Campbell, A.J., Angelov, T., Simmerling, C., and Scharer, O.D.
(2010). Synthesis and molecular modeling of a nitrogen mustard DNA interstrand crosslink. Chemistry (Easton) 16, 1210012103.
Ilves, I., Petojevic, T., Pesavento, J.J., and Botchan, M.R. (2010). Activation of
the MCM2-7 helicase by association with Cdc45 and GINS proteins. Mol. Cell
37, 247258.
Ishimi, Y. (1997). A DNA helicase activity is associated with an MCM4, 6,
and 7 protein complex. J. Biol. Chem. 272, 2450824513.
Kanter, D.M., and Kaplan, D.L. (2010). Sld2 binds to origin single-stranded
DNA and stimulates DNA annealing. Nucleic Acids Res. 39, 25802592.
Kaplan, D.L., Davey, M.J., and ODonnell, M. (2003). Mcm4,6,7 uses a pump
in ring mechanism to unwind DNA by steric exclusion and actively translocate
along a duplex. J. Biol. Chem. 278, 4917149182.
Kelman, Z., Lee, J.K., and Hurwitz, J. (1999). The single minichromosome
maintenance protein of Methanobacterium thermoautotrophicum DeltaH
contains DNA helicase activity. Proc. Natl. Acad. Sci. USA 96, 1478314788.
Knipscheer, P., Raschle, M., Smogorzewska, A., Enoiu, M., Ho, T.V., Scharer,
O.D., Elledge, S.J., and Walter, J.C. (2009). The Fanconi anemia pathway
promotes replication-dependent DNA interstrand cross-link repair. Science
326, 16981701.
Kochaniak, A.B., Habuchi, S., Loparo, J.J., Chang, D.J., Cimprich, K.A.,
Walter, J.C., and van Oijen, A.M. (2009). Proliferating cell nuclear antigen
uses two distinct modes to move along DNA. J. Biol. Chem. 284, 17700
17710.
Kuhn, H., and Frank-Kamenetskii, M.D. (2008). Labeling of unique sequences
in double-stranded DNA at sites of vicinal nicks generated by nicking endonucleases. Nucleic Acids Res. 36, e40.
Labib, K. (2010). How do Cdc7 and cyclin-dependent kinases trigger the initiation of chromosome replication in eukaryotic cells? Genes Dev. 24, 1208
1219.
Laskey, R.A., and Madine, M.A. (2003). A rotary pumping model for helicase
function of MCM proteins at a distance from replication forks. EMBO Rep. 4,
2630.
Li, D., Zhao, R., Lilyestrom, W., Gai, D., Zhang, R., DeCaprio, J.A., Fanning, E.,
Jochimiak, A., Szakonyi, G., and Chen, X.S. (2003). Structure of the replicative
helicase of the oncoprotein SV40 large tumour antigen. Nature 423, 512518.
Lucas, I., Chevrier-Miller, M., Sogo, J.M., and Hyrien, O. (2000). Mechanisms
ensuring rapid and complete DNA replication despite random initiation in
Xenopus early embryos. J. Mol. Biol. 296, 769786.
Mendez, J., and Stillman, B. (2003). Perpetuating the double helix: molecular
machines at eukaryotic DNA replication origins. Bioessays 25, 11581167.
Moyer, S.E., Lewis, P.W., and Botchan, M.R. (2006). Isolation of the Cdc45/
Mcm2-7/GINS (CMG) complex, a candidate for the eukaryotic DNA replication
fork helicase. Proc. Natl. Acad. Sci. USA 103, 1023610241.
Pacek, M., Tutter, A.V., Kubota, Y., Takisawa, H., and Walter, J.C. (2006).
Localization of MCM2-7, Cdc45, and GINS to the site of DNA unwinding during
eukaryotic DNA replication. Mol. Cell 21, 581587.
Remus, D., Beuron, F., Tolun, G., Griffith, J.D., Morris, E.P., and Diffley, J.F.
(2009). Concerted loading of Mcm2-7 double hexamers around DNA during
DNA replication origin licensing. Cell 139, 719730.
Shechter, D.F., Ying, C.Y., and Gautier, J. (2000). The intrinsic DNA helicase
activity of Methanobacterium thermoautotrophicum delta H minichromosome
maintenance protein. J. Biol. Chem. 275, 1504915059.
Smith, S.B., Cui, Y., and Bustamante, C. (1996). Overstretching B-DNA: the
elastic response of individual double-stranded and single-stranded DNA
molecules. Science 271, 795799.
Takahashi, T.S., and Walter, J.C. (2005). Cdc7-Drf1 is a developmentally regulated protein kinase required for the initiation of vertebrate DNA replication.
Genes Dev. 19, 22952300.
Takahashi, T.S., Wigley, D.B., and Walter, J.C. (2005). Pumps, paradoxes and
ploughshares: mechanism of the MCM2-7 DNA helicase. Trends Biochem.
Sci. 30, 437444.
Tremeau-Bravard, A., Riedl, T., Egly, J.M., and Dahmus, M.E. (2004). Fate of
RNA polymerase II stalled at a cisplatin lesion. J. Biol. Chem. 279, 77517759.
Tye, B.K. (1999). MCM proteins in DNA replication. Annu. Rev. Biochem. 68,
649686.
van Oijen, A.M. (2007). Honey, I shrunk the DNA: DNA length as a probe for
nucleic-acid enzyme activity. Biopolymers 85, 144153.
Walter, J., and Newport, J. (2000). Initiation of eukaryotic DNA replication:
origin unwinding and sequential chromatin association of Cdc45, RPA, and
DNA polymerase alpha. Mol. Cell 5, 617627.
Walter, J., Sun, L., and Newport, J. (1998). Regulated chromosomal DNA
replication in the absence of a nucleus. Mol. Cell 1, 519529.
Pacek, M., and Walter, J.C. (2004). A requirement for MCM7 and Cdc45 in
chromosome unwinding during eukaryotic DNA replication. EMBO J. 23,
36673676.
Warren, E.M., Vaithiyalingam, S., Haworth, J., Greer, B., Bielinsky, A.K.,
Chazin, W.J., and Eichman, B.F. (2008). Structural basis for DNA binding by
replication initiator Mcm10. Structure 16, 18921901.
Pape, T., Meka, H., Chen, S., Vicentini, G., van Heel, M., and Onesti, S. (2003).
Hexameric ring structure of the full-length archaeal MCM protein complex.
EMBO Rep. 4, 10791083.
Wessel, R., Schweizer, J., and Stahl, H. (1992). Simian virus 40 T-antigen DNA
helicase is a hexamer which forms a binary complex during bidirectional
unwinding from the viral origin of DNA replication. J. Virol. 66, 804815.
Randell, J.C., Fan, A., Chan, C., Francis, L.I., Heller, R.C., Galani, K., and Bell,
S.P. (2010). Mec1 Is One of Multiple Kinases that Prime the Mcm2-7 Helicase
for Phosphorylation by Cdc7. Mol. Cell 40, 353363.
Wohlschlegel, J.A., Dhar, S.K., Prokhorova, T.A., Dutta, A., and Walter, J.C.
(2002). Xenopus Mcm10 binds to origins of DNA replication after Mcm2-7
and stimulates origin binding of Cdc45. Mol. Cell 9, 233240.
Raschle, M., Knipscheer, P., Enoiu, M., Angelov, T., Sun, J., Griffith, J.D.,
Ellenberger, T.E., Scharer, O.D., and Walter, J.C. (2008). Mechanism of replication-coupled DNA interstrand crosslink repair. Cell 134, 969980.
Yardimci, H., Loveland, A.B., Habuchi, S., van Oijen, A.M., and Walter, J.C.
(2010). Uncoupling of sister replisomes during eukaryotic DNA replication.
Mol. Cell 40, 834840.
Cell 146, 931941, September 16, 2011 2011 Elsevier Inc. 941
for Cell Fate Decision, RIKEN Center for Developmental Biology, Kobe 650-0047, Japan
of Biology, Graduate School of Science, Kobe University, Kobe 657-8501, Japan
3Multicellular Organization Laboratory, National Institute of Genetics, 1111 Yata, Mishima 411-8540, Japan
4Present address: Department of Biological Sciences and Neurosciences Program, Stanford University, Stanford, CA 94305, USA
*Correspondence: hisawa@lab.nig.ac.jp
DOI 10.1016/j.cell.2011.07.043
2Department
SUMMARY
Extrinsic signals received by a cell can induce remodeling of the cytoskeleton, but the downstream effects
of cytoskeletal changes on gene expression have not
been well studied. Here, we show that during telophase of an asymmetric division in C. elegans, extrinsic Wnt signaling modulates spindle structures
through APR-1/APC, which in turn promotes asymmetrical nuclear localization of WRM-1/b-catenin
and POP-1/TCF. APR-1 that localized asymmetrically
along the cortex established asymmetric distribution
of astral microtubules, with more microtubules found
on the anterior side. Perturbation of the Wnt signaling
pathway altered this microtubule asymmetry and led
to changes in nuclear WRM-1 asymmetry, gene expression, and cell-fate determination. Direct manipulation of spindle asymmetry by laser irradiation altered the asymmetric distribution of nuclear WRM-1.
Moreover, laser manipulation of the spindles rescued
defects in nuclear POP-1 asymmetry in wnt mutants.
Our results reveal a mechanism in which the nuclear
localization of proteins is regulated through the
modulation of microtubules.
INTRODUCTION
Extrinsic signals can regulate gene transcription in the nuclei as
well as the cytoskeletal network in the cytoplasm. Regulation of
the cytoskeleton is generally considered to be a process independent from gene transcription or a consequencerather
than causeof gene transcription. One example counter to
this prevailing notion is the ability of actin polymerization triggered by extrinsic signals to directly influence gene expression
and differentiation (Miralles et al., 2003). These effects are elicited by the MAL transcription factor, which binds to unpolymerized actin in the cytoplasm and is transported to the nucleus
upon actin polymerization where it regulates target genes (Miralles et al., 2003). The effects of regulating other major cytoskel942 Cell 146, 942954, September 16, 2011 2011 Elsevier Inc.
in the embryonic EMS cell, which also undergoes Wnt-dependent asymmetric division (Lin et al., 1995; Nakamura et al.,
2005; Walston et al., 2004). Experiments in which WRM-1 was
expressed uniformly on the T cell cortex showed that WRM-1
on the anterior cortex not only recruits APR-1 to the cortex, but
also inhibits WRM-1 itself from localizing to the anterior nucleus,
so that it accumulates at the posterior nucleus. In addition,
APR-1 is required for nuclear asymmetry and efficient nuclear
export of WRM-1 (Mizumoto and Sawa, 2007a). However, the
mechanism by which cortical WRM-1 and APR-1 regulate
nuclear WRM-1 asymmetry remains elusive.
Focusing on the EMS cell division, we found that microtubules
and kinesins are required for nuclear WRM-1 and POP-1 asymmetry. During this division, the number of astral microtubules
became higher in the anterior than posterior side. This spindle
asymmetry was disrupted in mutants of apr-1, as well as in those
of mom-2/wnt and wrm-1, which also showed defects in cortical
APR-1 localization. To determine the importance of spindle
asymmetry for WRM-1 asymmetry, we disrupted or enhanced
the spindle asymmetry by laser irradiation of the centrosomes,
and we found that the nuclear WRM-1 asymmetry was lost and
enhanced, respectively. Moreover, in wnt mutants, restoration
of the spindle asymmetry with laser treatment rescued the
nuclear POP-1 asymmetry. These results strongly suggest that
the nuclear WRM-1 and POP-1 asymmetries are regulated by
spindle asymmetry. Our study reveals a microtubule-mediated
link between extrinsic signals and gene expression.
RESULTS
Nuclear WRM-1 Asymmetry Requires Microtubules
In the early embryogenesis of C. elegans, asymmetric EMS cell
division is regulated by mom-2/wnt expressed in a P2 cell that
attaches to the EMS cell posterior cortex. After the division,
the anterior MS and posterior E daughters produce muscle
and gut, respectively. However, in mom-2 mutants, the posterior
E daughter has levels of nuclear WRM-1 and POP-1 similar to the
anterior MS daughter (Nakamura et al., 2005; Rocheleau et al.,
1997) and adopts an MS-like fate, resulting in the gutless phenotype (Rocheleau et al., 1997; Thorpe et al., 1997). To study the
role of microtubules in nuclear WRM-1 and POP-1 asymmetry,
we treated EMS cells with vinblastine or nocodazole at anaphase
to destabilize the microtubules without disrupting the cell division itself (see the Experimental Procedures; data not shown).
After division, we found that nuclear WRM-1 was symmetric
(Figures 1C and 1G). In addition, POP-1 signal in the anterior
nucleus was significantly decreased, resulting in more symmetric localization (Figures 1D and 1H). These results suggest
that microtubules are necessary to establish the nuclear WRM-1
and POP-1 asymmetries and are consistent with a previous
observation that a microtubule destabilizing drug inhibits gut
production (Goldstein, 1995a).
Astral Microtubule Numbers Become Asymmetric
during EMS Division
To determine how microtubules regulate the WRM-1 and POP-1
asymmetry, we examined whether the microtubule organization
is itself asymmetric. In embryos immunostained for endogenous
a-tubulin, we manually counted the number of astral microtubules emanating from each anterior or posterior centrosome
during EMS division. We found that the numbers of astral microtubules were asymmetric after metaphase and that the anterior
spindle pole had more astral microtubules than the posterior pole (Figure 2A and Figure S1 available online). This microtubule number asymmetry (hereafter called spindle asymmetry)
became most prominent at telophase (Figure 2A). Analyses of
the number of astral microtubules in living embryos expressing
GFP::b-tubulin confirmed the asymmetry to be strongest at telophase (Figure 2B; 0240 s). Consistent with this observation, the
amount of a-tubulin around each centrosome was asymmetric at
telophase, as measured by immunostaining and normalized to
the amount of g-tubulin (Figure S2B). To further confirm the
spindle asymmetry, we quantified GFP::b-tubulin intensities in
the entire anterior or posterior halves of the EMS cell and mathematically estimated the GFP signals corresponding to astral
microtubules. Astral microtubule intensity was defined as the
signals found between the strong signals of the centrosomal
regions and the weak background signals (estimated from signals in regions that do not contain microtubules). We found
that the numbers of pixels corresponding to microtubules are
larger in the anterior than posterior sides (Figure S2A). Based
on these results, we concluded that the astral microtubule
numbers are asymmetric during EMS division, most prominently
at telophase.
The Asymmetric Spindle Is Regulated by Wnt Signaling
Because the mitotic spindle in the EMS cell is oriented along
the anterior-posterior axis through the functions of MOM-2/
Wnt by metaphase (Goldstein, 1995b; Schlesinger et al., 1999),
we analyzed whether various Wnt signaling components are
involved in establishing the spindle asymmetry. In mom-2 or
mom-5/frizzled embryos, there appeared to be fewer microtubules in the anterior and more microtubules in the posterior
halves, albeit not significantly, than those in the wild-type.
Accordingly, the difference between the number of anterior
and posterior microtubules was significantly decreased in both
embryos (p value = 0.03 and 0.05 in mom-2 and mom-5, respectively), thereby disrupting the spindle asymmetry (Figures 2C and
2D and Figure S2C). In dsh-2(RNAi)mig-5(RNAi) [hereafter called
dsh(RNAi)], wrm-1(RNAi), and apr-1(RNAi) embryos, the anterior
microtubule numbers were significantly decreased compared to
those of control embryos, resulting in nearly symmetric microtubule numbers (Figure 2D and Figure S2C). Because apr-1 and
wrm-1 are not involved in spindle orientation (Schlesinger
et al., 1999), our results indicate that spindle asymmetry is regulated independently of spindle orientation by the Wnt/b-catenin
asymmetry pathway.
Spindle Asymmetry Is Likely to Be Regulated by Cortical
APR-1/APC
How does Wnt signaling regulate spindle asymmetry? APR-1/
APC may be a key regulator, since APC is known to regulate
microtubules (Munemitsu et al., 1994). In postembryonic seam
cells, Wnt signaling promotes the asymmetrical localization
of APR-1 (Mizumoto and Sawa, 2007a). Similarly, in the EMS
cell, we found that a GFP fusion protein of APR-1 localized
Cell 146, 942954, September 16, 2011 2011 Elsevier Inc. 943
GFP::WRM-1/-catenin
GFP::POP-1/TCF
0.2% DMSO
MS
MS
E
100 nM vinblastine
MS
10 M ATA
10 M ATA
MS
H
200
(A.U.)
n.s.
n.s.
150
100
50
A P
DMSO
(n = 10)
A P
vinblastine
(n = 10)
A
P
ATA
(n = 10)
GFP::POP-1 in
intensity
GFP::WRM-1 intensity
**
G
(A.U.)
E
100 nM vinblastine
MS
MS
0.2% DMSO
2000
**
**
1500
1000
500
A
P
DMSO
(n = 10)
reduction in the anterior GFP::APR-1 levels (Figure 3B and Figure S3B). Although
wrm-1(RNAi) embryos still showed significant asymmetry in GFP::APR-1 levels,
the intensity of the anterior signal was
significantly reduced. As with the localization kinetics, the effect of these mutations
n.s.
*
or RNAi on GFP::APR-1 localization strongly correlated with their effects on spindle
asymmetry (compare Figure 2D and Figure 3B). Our results strongly suggest that
spindle asymmetry depends on the cortical GFP::APR-1 level.
A
P
A P
The number of microtubules depends
vinblastine
ATA
on
microtubule nucleation rate and
(n = 10)
(n = 8)
stability. To analyze which of the factors
is the cause for spindle asymmetry,
we used living embryos expressing
GFP::EBP-1. EBP-1 is a C. elegans homolog of the mammalian
EB1 protein, which binds to microtubule plus ends and was
used to analyze microtubule dynamics (Mimori-Kiyosue et al.,
2000b; Srayko et al., 2005). We measured the number and speed
of GFP::EBP-1 comets around the centrosomes for 5 s to
measure microtubule nucleation and polymerization rates,
respectively, and we found that these factors were neither significantly asymmetric nor affected by apr-1 knockdown (Figures
S3DS3F). Then, we examined the possibility that APR-1 on the
cell cortex affects the local microtubule stability similar to
mammalian APC. We measured the EBP-1 residency time on
the cortex and found that EBP-1 comets persisted for a longer
time on the anterior than on the posterior cortex (67% and
27% comets persisted longer than 5 s on the anterior and posterior cortex, respectively) (Figures 3C and 3D). In apr-1(RNAi),
EBP-1 comets on the anterior cortex rapidly detached themselves from the cortex (27% comets persisted longer than 5 s)
similar to the posterior cortex (Figures 3C and 3D). These results
E
B
80
**
n.s.
60
40
20
0
A P
prophase
(n = 4)
A P
metaphase
(n = 8)
telophase
20
A P
telophase
(n = 7)
15
Anterior
Posterior
10
5
(n = 10)
0
C
-tubulin
-tubulin
merge
WT
WT
WT
mom-2
mom-2
mom-2
D
80
**
n.s.
n.s.
n.s.
n.s.
n.s.
60
40
20
0
A P
WT
(n = 7)
A P
A P
A P
mom-2(or309) mom-5(ne12) dsh(RNAi)
(n = 6)
(n = 7)
(n = 5)
A P
A P
wrm-1(RNAi) apr-1(RNAi)
(n = 5)
(n = 11)
Cell 146, 942954, September 16, 2011 2011 Elsevier Inc. 945
GFP::APR-1
cortical plane
mid-plane
interphase
GFP::EBP-1
control
anterior
posterior
cell cortex
cell cortex
0:00
0:00
apr-1(RNAi)
anterior
cell cortex
0:00
0:01
0:01
0:01
0:02
0:02
0:02
0:03
0:03
0:03
0:04
0:04
0:04
0:05
0:05
0:05
0:06
0:06
0:06
prophase
metaphase
anaphase
telophase
B (A.U.)
n.s.
n.s.
3500
3000
n.s.
2500
n.s.
2000
1500
1000
500
0
genotypes
transgene
A P A P A P A P A P A P
WT mom-2 mom-5 dsh wrm-1 WT
GFP::APR-1
none
(n = 5) (n = 5) (n = 5) (n = 5) (n = 5) (n = 3)
Frequency (%)
4000
70
60
50
40
30
20
10
0
Figure 3. Wnt-Dependent Asymmetric APR-1 Localization Is Required for Asymmetric Microtubule Stability on the Cell Cortex
(A) GFP::APR-1 localization in mid-planes (left) and cortical planes (right) of living embryos during EMS division at each cell-cycle phase. Arrowheads indicate the
boundary of EMS cell.
(B) Cortical GFP::APR-1 signal intensities in the anterior and posterior sides of the EMS cell with the indicated genotypes. Embryos without the GFP::APR-1
transgene were also quantified as background signals. Error bars indicate 95% confidence intervals. Asterisk (*), double asterisk (**), and n.s. indicate the
significance of differences in intensity between the anterior and posterior sides (Welchs t test; *p < 0.05; **p < 0.01; n.s., p > 0.05). Red circles indicate significant
reductions in intensity compared to wild-type (p < 0.05).
(C) Examples of time lapse images of cortical GFP::EBP-1 movement during telophase of EMS. The left, center, and right panels show the anterior cell cortex
(control), posterior cell cortex (control), and anterior cell cortex [apr-1(RNAi)], respectively. Arrows and arrowheads with distinct colors indicate individual EBP-1
comets and the positions where the comets disappeared, respectively.
946 Cell 146, 942954, September 16, 2011 2011 Elsevier Inc.
(D) Frequency of GFP::EBP-1 residency time on the cortex [CtrlA, control anterior; CtrlP, control posterior; aprA, apr-1(RNAi) anterior; and aprP, apr-1(RNAi)
posterior cortex].
Scale bars represent 5 mm. See also Figure S3.
Cell 146, 942954, September 16, 2011 2011 Elsevier Inc. 947
GFP::WRM-1
control irradiation
anaphase
1 2
GFP::POP-1
control irradiation
1. Anterior irradiation
2. Control irradiation
3. Posterior irradiation
E (A.U.)
D (A.U.)
posterior irradiation
5000
GFP::WRM-1
4000
12000
3000
2000
1000
0
-1000
-2000
*
Ctrl A-ir P-ir
posterior irradiation
anterior irradiation
anterior irradiation
8000
6000
anti-POP-1
DAPI
4000
2000
0
-2000
(n = 7) (n = 8)(n = 11)
GFP::POP-1
10000
(n = 10)(n = 8) (n = 9)
mom-2(or309)
no irradiation
mom-2(or309)
posterior irradiation
Wild type
(A.U.)
6000
endogenous POP-1
5000
4000
3000
2000
1000
0
-1000
(n = 9)
Wild type
(n = 10)
(n = 5)
Posterior
No
irradiation irradiation
mom-2(or309)
Figure 5. Effect of Centrosomal Irradiation at Anaphase on the Nuclear Localization of WRM-1 and POP-1 at Telophase
(A and B) Wild-type or mom-2(or309) embryos irradiated at anaphase were immunostained for GFP::WRM-1 (A, left), GFP::POP-1 (A, right) or endogenous POP-1
(B) 5 min after furrowing onset. White lines and white dotted lines outline EMS daughter cells and their nuclei, respectively. Scale bars represent 5 mm.
(C) Schematic view of the EMS cell at anaphase with irradiation locations. Dark circles and white ellipses represent centrosomes and chromosomes, respectively.
(DF) Bar graphs represent the results of centrosomal irradiation using GFP::WRM-1 (D), GFP::POP-1 (E), and endogenous POP-1 staining (F) as shown in (A) and
(B). Differences in signal intensities of GFP::WRM-1 (D), GFP::POP-1 (E), or POP-1 (F) between the anterior and posterior nuclei are shown. Error bars indicate
95% confidence intervals. Asterisks indicate significant differences compared to the control irradiation (D and E) or mom-2(or309) embryos with no irradiation
(F) (Welchs t test; p < 0.05). A.U., arbitrary units.
See also Figure S4.
uniform after the late six-cell stage (Walston et al., 2004), DSH-2
might be enriched at the EMS cell posterior cortex during telophase, as seen in postembryonic seam cells (Mizumoto and
Sawa, 2007a). However, dsh(RNAi) experiments revealed that
its effects were observed more strongly in the anterior cortex
than in the posterior cortex. Therefore, DSH-2 appears to function at the anterior cortex to recruit APR-1 to the cortex and
this action may be inhibited in the posterior cortex, probably
by MOM-2. Another possibility is that DSH-2 has both positive
and negative roles in APR-1 localization. While DSH-2 functions
to localize APR-1 to the cortex, the DSH-2 that is recruited to the
E (A.U.)
Wildtype
Wildtype
cyk-4(ts)
cyk-4(ts)
zen-4(ts)
GFP::POP-1 intensity
400
WRM-1::GFP
22.5C
350
300
250
**
200
**
150
100
50
0
zen-4(ts)
control apr-1
AP
AP
(A.U.)
250
GFP::POP-1 intensity
F
B
AJM-1::GFP
zen-4(ts)
cyk-4(ts)
ZEN-4::GFP
metaphase
telophase
(n = 9) (n = 13)
15C
25C
200
150
**
**
100
50
0
zen-4 background
(n = 7)
D
control
apr-1(RNAi)
kap-1(ok676)
zen-4(or153ts)
klp-18(RNAi)
zen-4;klp-18
GFP::POP-1
(n = 8)
(n = 8)
(n = 6) (n = 6)
klp-18(RNAi) , anterior
-tubulin/-tubulin/DAPI
klp-18(RNAi) , posterior
bmk-1(RNAi)
zen-4;bmk-1
H
80
70
60
50
40
30
20
10
0
22.5C
15C
25C
**
A P d
control
(n = 10)
A P d
klp-18
(n = 5)
A P d
A P d
zen-4
zen-4;bmk-1
(n = 7)
(n = 7)
950 Cell 146, 942954, September 16, 2011 2011 Elsevier Inc.
B
C
WRM-1
KLP-18/kinesin
P
N
MT
[WRM-1N]
[WRM-1C]
[WRM-1P]
Kinesin
(C) Localization of ZEN-4::GFP during T cell division. Arrowheads and arrows indicate T cell outlines and centrosomes, respectively.
(DF) GFP::POP-1 localization in kinesin gene mutants or RNAi embryos. Quantified fluorescence intensities were shown in (E) and (F) (A.U., arbitrary unit). zen4(or153ts) mutants and animals used in (F) were cultured at 15 C and shifted to 25 C just before EMS division. Other animals were cultured at 22.5 C. kap-1 is
a C. elegans homolog of mammalian KAP3.
(G) Immunostaining of a-tubulin (red) and g-tubulin (green) in klp-18(RNAi) embryos. Blue is DAPI.
(H) Number of astral microtubules in the anterior (A) or posterior (P) sides with their difference (d) in kinesin mutants or RNAi animals.
Error bars in (E), (F), and (H) indicate 95% confidence intervals. Asterisks in (E), (F), and (H) indicate the significant difference compared to control (Welchs t test;
p < 0.01). Lines and dotted lines in (A), (D), and (G) indicate cell boundaries and nuclei, respectively. Scale bars represent 10 mm. See also Table S1.
Cell 146, 942954, September 16, 2011 2011 Elsevier Inc. 951
containing less than 0.2% DMSO), were introduced into the space to displace
the egg salt buffer (Edgar, 1995). To introduce the drugs without affecting cell
division, we produced a hole on the eggshell by irradiating the trypan blue
particles on it specifically, when the dividing EMS cell was at anaphase. Permeabilization of the eggshell was confirmed by the slight swelling or shrinking
of the embryo.
EXPERIMENTAL PROCEDURES
Immunostaining
Embryos were freeze-cracked and fixed in 20 C methanol for a-tubulin,
g-tubulin, and GFP::APR-1 staining or, for GFP::WRM-1, GFP::POP-1 and
20 C acetone. For
LacZ staining, in
20 C methanol followed by
GFP::WRM-1 staining, after being fixed and washed, the embryos were incubated with Image-iT FX Signal Enhancer (Invitrogen) at room temperature for
30 min. The embryos were stained in the presence of 1% BSA (except for
GFP::WRM-1 staining) with the following primary antibodies: anti-a-tubulin
mouse monoclonal (clone DM1a; 1:1000; Sigma-Aldrich), anti-g-tubulin rabbit
polyclonal (LL-17; 1:1000; Sigma-Aldrich), anti-GFP rabbit polyclonal (1:1000;
Invitrogen), or anti-LacZ mouse monoclonal (Z378; 1:1000; Promega). Primary antibodies were detected by goat anti-mouse Rhodamine-X (1:1000;
Invitrogen), goat anti-rabbit Fluorescein (1:1000; Invitrogen), or goat antimouse Alexa 568 highly cross-adsorbed (1:1000; Invitrogen). For endogenous
POP-1 staining, embryos were fixed and immunostained as described previously (Sugioka and Sawa, 2010) with an anti-POP-1 antibody (Lin et al.,
1998).
952 Cell 146, 942954, September 16, 2011 2011 Elsevier Inc.
REFERENCES
activating protein (GAP) required for central spindle formation and cytokinesis.
J. Cell Biol. 149, 13911404.
Batut, J., Howell, M., and Hill, C.S. (2007). Kinesin-mediated transport of
Smad2 is required for signaling in response to TGF-b ligands. Dev. Cell 12,
261274.
Jimbo, T., Kawasaki, Y., Koyama, R., Sato, R., Takada, S., Haraguchi, K., and
Akiyama, T. (2002). Identification of a link between the tumour suppressor APC
and the kinesin superfamily. Nat. Cell Biol. 4, 323327.
Labbe, J.C., McCarthy, E.K., and Goldstein, B. (2004). The forces that position
a mitotic spindle asymmetrically are tethered until after the time of spindle
assembly. J. Cell Biol. 167, 245256.
Lalli, G., Gschmeissner, S., and Schiavo, G. (2003). Myosin Va and microtubule-based motors are required for fast axonal retrograde transport of tetanus
toxin in motor neurons. J. Cell Sci. 116, 46394650.
Lee, J.Y., and Goldstein, B. (2003). Mechanisms of cell positioning during
C. elegans gastrulation. Development 130, 307320.
Lin, R., Thompson, S., and Priess, J.R. (1995). pop-1 encodes an HMG box
protein required for the specification of a mesoderm precursor in early
C. elegans embryos. Cell 83, 599609.
Lin, R., Hill, R.J., and Priess, J.R. (1998). POP-1 and anterior-posterior fate
decisions in C. elegans embryos. Cell 92, 229239.
Lo, M.C., Gay, F., Odom, R., Shi, Y., and Lin, R. (2004). Phosphorylation by the
b-catenin/MAPK complex promotes 14-3-3-mediated nuclear export of TCF/
POP-1 in signal-responsive cells in C. elegans. Cell 117, 95106.
Maduro, M.F., Lin, R., and Rothman, J.H. (2002). Dynamics of a developmental
switch: recursive intracellular and intranuclear redistribution of Caenorhabditis
elegans POP-1 parallels Wnt-inhibited transcriptional repression. Dev. Biol.
248, 128142.
Maduro, M.F., Kasmir, J.J., Zhu, J., and Rothman, J.H. (2005). The Wnt
effector POP-1 and the PAL-1/Caudal homeoprotein collaborate with SKN-1
to activate C. elegans endoderm development. Dev. Biol. 285, 510523.
Mimori-Kiyosue, Y., Shiina, N., and Tsukita, S. (2000a). Adenomatous polyposis coli (APC) protein moves along microtubules and concentrates at their
growing ends in epithelial cells. J. Cell Biol. 148, 505518.
Mimori-Kiyosue, Y., Shiina, N., and Tsukita, S. (2000b). The dynamic behavior
of the APC-binding protein EB1 on the distal ends of microtubules. Curr. Biol.
10, 865868.
Fagotto, F., Gluck, U., and Gumbiner, B.M. (1998). Nuclear localization
signal-independent and importin/karyopherin-independent nuclear import of
b-catenin. Curr. Biol. 8, 181190.
Miralles, F., Posern, G., Zaromytidou, A.I., and Treisman, R. (2003). Actin
dynamics control SRF activity by regulation of its coactivator MAL. Cell 113,
329342.
Giannakakou, P., Sackett, D.L., Ward, Y., Webster, K.R., Blagosklonny, M.V.,
and Fojo, T. (2000). p53 is associated with cellular microtubules and is transported to the nucleus by dynein. Nat. Cell Biol. 2, 709717.
Mishima, M., Kaitna, S., and Glotzer, M. (2002). Central spindle assembly and
cytokinesis require a kinesin-like protein/RhoGAP complex with microtubule
bundling activity. Dev. Cell 2, 4154.
Goldstein, B. (1995b). Cell contacts orient some cell division axes in the
Caenorhabditis elegans embryo. J. Cell Biol. 129, 10711080.
Mizumoto, K., and Sawa, H. (2007a). Cortical b-catenin and APC regulate
asymmetric nuclear b-catenin localization during asymmetric cell division in
C. elegans. Dev. Cell 12, 287299.
Goldstein, B., Takeshita, H., Mizumoto, K., and Sawa, H. (2006). Wnt signals
can function as positional cues in establishing cell polarity. Dev. Cell 10,
391396.
Mizumoto, K., and Sawa, H. (2007b). Two betas or not two betas: regulation of
asymmetric division by b-catenin. Trends Cell Biol. 17, 465473.
Grill, S.W., Howard, J., Schaffer, E., Stelzer, E.H., and Hyman, A.A. (2003). The
distribution of active force generators controls mitotic spindle position.
Science 301, 518521.
Motegi, F., Velarde, N.V., Piano, F., and Sugimoto, A. (2006). Two phases of
astral microtubule activity during cytokinesis in C. elegans embryos. Dev.
Cell 10, 509520.
Gundersen, G.G., Gomes, E.R., and Wen, Y. (2004). Cortical control of microtubule stability and polarization. Curr. Opin. Cell Biol. 16, 106112.
Munemitsu, S., Souza, B., Muller, O., Albert, I., Rubinfeld, B., and Polakis, P.
(1994). The APC gene product associates with microtubules in vivo and
promotes their assembly in vitro. Cancer Res. 54, 36763681.
Herman, M.A., Vassilieva, L.L., Horvitz, H.R., Shaw, J.E., and Herman, R.K.
(1995). The C. elegans gene lin-44, which controls the polarity of certain asymmetric cell divisions, encodes a Wnt protein and acts cell nonautonomously.
Cell 83, 101110.
Hopkins, S.C., Vale, R.D., and Kuntz, I.D. (2000). Inhibitors of kinesin activity
from structure-based computer screening. Biochemistry 39, 28052814.
Nakamura, K., Kim, S., Ishidate, T., Bei, Y., Pang, K., Shirayama, M., Trzepacz,
C., Brownell, D.R., and Mello, C.C. (2005). Wnt signaling drives WRM-1/
b-catenin asymmetries in early C. elegans embryos. Genes Dev. 19, 1749
1754.
Hyman, A.A., and White, J.G. (1987). Determination of cell division axes in the
early embryogenesis of Caenorhabditis elegans. J. Cell Biol. 105, 21232135.
Nathke, I.S., Adams, C.L., Polakis, P., Sellin, J.H., and Nelson, W.J. (1996). The
adenomatous polyposis coli tumor suppressor protein localizes to plasma
membrane sites involved in active cell migration. J. Cell Biol. 134, 165179.
Jantsch-Plunger, V., Gonczy, P., Romano, A., Schnabel, H., Hamill, D., Schnabel, R., Hyman, A.A., and Glotzer, M. (2000). CYK-4: A Rho family gtpase
Phillips, B.T., and Kimble, J. (2009). A new look at TCF and b-catenin through
the lens of a divergent C. elegans Wnt pathway. Dev. Cell 17, 2734.
Cell 146, 942954, September 16, 2011 2011 Elsevier Inc. 953
Praitis, V., Casey, E., Collar, D., and Austin, J. (2001). Creation of low-copy
integrated transgenic lines in Caenorhabditis elegans. Genetics 157, 1217
1226.
Srayko, M., Kaya, A., Stamford, J., and Hyman, A.A. (2005). Identification and
characterization of factors required for microtubule growth and nucleation in
the early C. elegans embryo. Dev. Cell 9, 223236.
Purro, S.A., Ciani, L., Hoyos-Flight, M., Stamatakou, E., Siomou, E., and Salinas, P.C. (2008). Wnt regulates axon behavior through changes in microtubule
growth directionality: a new role for adenomatous polyposis coli. J. Neurosci.
28, 86448654.
Rocheleau, C.E., Downs, W.D., Lin, R., Wittmann, C., Bei, Y., Cha, Y.H., Ali, M.,
Priess, J.R., and Mello, C.C. (1997). Wnt signaling and an APC-related gene
specify endoderm in early C. elegans embryos. Cell 90, 707716.
Takeshita, H., and Sawa, H. (2005). Asymmetric cortical and nuclear localizations of WRM-1/b-catenin during asymmetric cell division in C. elegans. Genes
Dev. 19, 17431748.
Roth, D.M., Moseley, G.W., Glover, D., Pouton, C.W., and Jans, D.A. (2007). A
microtubule-facilitated nuclear import pathway for cancer regulatory proteins.
Traffic 8, 673686.
Thorpe, C.J., Schlesinger, A., Carter, J.C., and Bowerman, B. (1997). Wnt
signaling polarizes an early C. elegans blastomere to distinguish endoderm
from mesoderm. Cell 90, 695705.
Sawa, H., Kouike, H., and Okano, H. (2000). Components of the SWI/SNF
complex are required for asymmetric cell division in C. elegans. Mol. Cell 6,
617624.
Schlesinger, A., Shelton, C.A., Maloof, J.N., Meneghini, M., and Bowerman, B.
(1999). Wnt pathway components orient a mitotic spindle in the early
Caenorhabditis elegans embryo without requiring gene transcription in the
responding cell. Genes Dev. 13, 20282038.
Schlessinger, K., McManus, E.J., and Hall, A. (2007). Cdc42 and noncanonical
Wnt signal transduction pathways cooperate to promote cell polarity. J. Cell
Biol. 178, 355361.
Schneider, S.Q., and Bowerman, B. (2007). b-Catenin asymmetries after all
animal/vegetal- oriented cell divisions in Platynereis dumerilii embryos
mediate binary cell-fate specification. Dev. Cell 13, 7386.
954 Cell 146, 942954, September 16, 2011 2011 Elsevier Inc.
Walston, T., Tuskey, C., Edgar, L., Hawkins, N., Ellis, G., Bowerman, B., Wood,
W., and Hardin, J. (2004). Multiple Wnt signaling pathways converge to orient
the mitotic spindle in early C. elegans embryos. Dev. Cell 7, 831841.
Wiechens, N., and Fagotto, F. (2001). CRM1- and Ran-independent nuclear
export of b-catenin. Curr. Biol. 11, 1827.
Yokoya, F., Imamoto, N., Tachibana, T., and Yoneda, Y. (1999). b-catenin can
be transported into the nucleus in a Ran-unassisted manner. Mol. Biol. Cell 10,
11191131.
Zumbrunn, J., Kinoshita, K., Hyman, A.A., and Nathke, I.S. (2001). Binding of
the adenomatous polyposis coli protein to microtubules increases microtubule
stability and is regulated by GSK3 b phosphorylation. Curr. Biol. 11, 4449.
SUMMARY
Protein concentration gradients encode spatial information across cells and tissues and often depend
on spatially localized protein synthesis. Here, we report that a different mechanism underlies the MEX-5
gradient. MEX-5 is an RNA-binding protein that
becomes distributed in a cytoplasmic gradient
along the anterior-to-posterior axis of the one-cell
C. elegans embryo. We demonstrate that the MEX5 gradient is a direct consequence of an underlying
gradient in MEX-5 diffusivity. The MEX-5 diffusion
gradient arises when the PAR-1 kinase stimulates
the release of MEX-5 from slow-diffusive, RNA-containing complexes in the posterior cytoplasm. PAR1 directly phosphorylates MEX-5 and is antagonized
by the spatially uniform phosphatase PP2A. Mathematical modeling and in vivo observations demonstrate that spatially segregated phosphorylation
and dephosphorylation reactions are sufficient to
generate stable protein concentration gradients in
the cytoplasm. The principles demonstrated here
apply to any spatially segregated modification cycle
that affects protein diffusion and do not require
protein synthesis or degradation.
INTRODUCTION
Protein gradients are an efficient way to encode spatial information within cells and across tissues. The mechanisms that
generate and maintain protein gradients have been the subject
of extensive theoretical and experimental analyses (Wartlick
et al., 2009). Most studies have emphasized the role of a localized
protein source as the foundational asymmetry underlying
gradient formation. For example, in Drosophila embryos, the
Bicoid protein is synthesized at one end of the egg from a localized pool of bicoid mRNA. Diffusion away from the local source
and uniform protein degradation across the egg generate a
concentration gradient over the course of 2 hr (Ephrussi and
help to specify their distinct fates (anterior/somatic and posterior/germline). Mutations in the PARs cause MEX-5 (and its
targets) to remain symmetrically distributed (Schubert et al.,
2000; Tenlen et al., 2008), but the mechanisms linking PAR
asymmetry to the MEX-5 gradient are not known.
Fluorescence recovery after photobleaching (FRAP) and
fluorescence correlation spectroscopy (FCS) experiments have
shown that, in polarized zygotes, GFP::MEX-5 diffuses faster
in the posterior cytoplasm, where MEX-5 protein concentration
is lowest (Daniels et al., 2010; Tenlen et al., 2008). Fast diffusion
requires par-1 activity and a C-terminal serine in MEX-5 (S458),
which is phosphorylated in a par-1- and par-4-dependent
manner in vivo (Tenlen et al., 2008). Phosphorylation of S458,
however, does not correlate with gradient formation or fast
diffusion, suggesting that other mechanisms regulate MEX-5
asymmetry (Tenlen et al., 2008). Two speculative models have
been proposed. The first model invokes dynamic binding of
MEX-5 to cytoskeletal elements asymmetrically distributed in
the cytoplasm (Tenlen et al., 2008). In this model, the PARs
localize MEX-5 indirectly by localizing factors, such as myosin,
that retard MEX-5 diffusion in the anterior cytoplasm (Tenlen
et al., 2008). A second model proposes that the PARs regulate
MEX-5 distribution by forming reactive surfaces in the anterior
and posterior cortices, which locally decrease and increase,
respectively, the rate of MEX-5 diffusion (Daniels et al., 2010).
How the PARs modify MEX-5 diffusion, and how differences
originated at the cortex are propagated through the cytoplasm,
however, is not known.
In this study, we present evidence that the MEX-5 gradient
arises as a direct consequence of a complementary PAR-1
kinase activity gradient in the cytoplasm. We demonstrate
that MEX-5 is a substrate of PAR-1 and identify PP2A as the
opposing phosphatase in the cytoplasm. Our findings reveal
an unexpected direct patterning role for PAR-1 in the cytoplasm
and provide experimental evidence for the theoretical model of
Lipkow and Odde (2008).
RESULTS
A MEX-5 Diffusion Gradient Underlies the MEX-5
Concentration Gradient
To examine MEX-5 dynamics in live zygotes, we generated
a Dendra::MEX-5 fusion. Dendra is a photoactivatable fluorescent protein that is photoconverted irreversibly from green to
red fluorescence by exposure to 405 nm light (Gurskaya et al.,
2006). Unlike FRAP, photoconversion is a positive marking technique that can be used to measure rates of protein degradation
and diffusion, without interference from new protein synthesis
(Lippincott-Schwartz and Patterson, 2008). We first photoconverted Dendra::MEX-5 throughout the zygote before polarization
(prior to appearance of the pronuclei). We found that photoconverted Dendra::MEX-5 (DendraR::MEX-5) formed an 3-fold
anterior-posterior gradient by nuclear envelope breakdown
(NEBD, first mitotic division), as is observed for endogenous
MEX-5 (Figures 1A and 1B). Total levels of DendraR::MEX-5 did
not change during gradient formation: levels increased in the
anterior half and decreased in the posterior half by 25% (Figure 1C). We conclude that formation of the MEX-5 gradient
956 Cell 146, 955968, September 16, 2011 2011 Elsevier Inc.
5m
Endogenous MEX-5
DendraR::MEX-5
Anterior
Posterior
After PN
0%
100%
Embryo Length
Before PN
Formation
Relative Concentration
(AU)
NEBD
DendraR::MEX-5
PN
Meeting
NEBD
2 second
Relative Intensity
(AU)
Before PN
D
1 second
DendraR::MEX-5
Anterior
Total
Posterior
Time (minutes)
3 second
12
Anterior
Posterior
Before PN
10%
10%
50%
After PN
NEBD
90%
90%
Anterior
10%
Posterior
30%
50%
70%
90%
Embryo Length
Embryo Length
Long Axis
Short Axis
Wild-type
PKC-3
MEX-5
Gradient:
pkc-3(RNAi)
par-1(it51)
par-1(b274)
par-2(RNAi), PN meeting
No
No
No
Yes
PAR-1
Yes
PAR-1
it51
R409K
b274
PKC-3
phosphorylation
site
Q819STOP
T983
1192
Kinase
Cortical localization
domain (aa965-1192)
Before PN Formation
KA1 domain
NEBD
Anterior
Posterior
WT
par-1(b274)
Wild-type
After PN Formation
WT
pkc-3(RNAi)
par-1(it51)
Wild-type
PN Meeting
par-1(it51)
pkc-3(RNAi)
par-1(b274)
par-1(b274)
par-1(RNAi)
par-1(b274)
pkc-3(RNAi)
par-2(RNAi)
par-2(RNAi)
PN Meeting
GFP::PAR-1
DendraR::MEX-5
958 Cell 146, 955968, September 16, 2011 2011 Elsevier Inc.
migration, GFP::PAR-1 levels remained low in the anteriorperipheral cytoplasm but increased in the posterior cytoplasm
and on the posterior cortex. By NEBD, GFP::PAR-1 was
enriched on the posterior cortex and formed a 3-fold anteriorlow/posterior-high gradient in the cytoplasm, paralleling the
gradient in MEX-5 diffusivity (Figure 2D and Figure S1C). Immunostaining of wild-type (WT) embryos with an anti-PAR-1 antibody confirmed the presence of a PAR-1 gradient in the cytoplasm of zygotes at NEBD (Figure S1D). We conclude that
PAR-1 dynamics in the cytoplasm correlate with MEX-5 diffusion
dynamics and that MEX-5 responds quickly to changes in PAR-1
distribution.
To explore whether cytoplasmic PAR-1 is sufficient to stimulate MEX-5 diffusion, we analyzed the par-1 allele b274.
par-1(b274) zygotes do not localize PAR-1 to the cortex and
do not segregate MEX-5 but are positive for pS458, suggesting
that this allele retains some par-1 kinase activity (Figure S1B)
(Guo and Kemphues, 1995; Tenlen et al., 2008). We sequenced
par-1(b274) and found a premature stop codon at residue
Q819 between the kinase domain and the domain that localizes
PAR-1 to the cortex (Figure 2B). Western blotting and immunofluorescence analyses confirmed the presence of a truncated
PAR-1 protein, expressed at 14% of wild-type levels and
uniformly cytoplasmic (Figures S1B and S1E) (Hurd and Kemphues, 2003). Before pronuclear formation, DendraR::MEX-5
mobility was uniformly low in par-1(b274) zygotes, as in wildtype and par-1(it51) zygotes. By NEBD, however, the apparent
Dc of DendraR::MEX-5 had increased throughout the cytoplasm
to a value intermediate between that of par-1(it51) and
pkc-3(RNAi) zygotes (Figure 2C). In par-1(b274) zygotes,
PKC-3 became enriched on the anterior cortex as in wild-type,
whereas DendraR::MEX-5 remained symmetrically distributed
(Figure S1B) (Tenlen et al., 2008). The intermediate DendraR::
MEX-5 diffusion rate in par-1(b274) zygotes was dependent on
PAR-1 but not on PKC-3 (Figure 2C). We conclude that PAR-1
kinase activity in the cytoplasm is sufficient to increase MEX-5
diffusivity after pronuclear formation.
We also examined the distribution of PAR-1 and MEX-5 in
par-2 zygotes, which localize anterior PARs to the anterior cortex
before, but not after, NEBD and which never enrich PAR-1 on the
posterior cortex (Boyd et al., 1996; Cuenca et al., 2003). We
found that GFP::PAR-1 still formed a cytoplasmic gradient by
pronuclear meeting in par-2(RNAi) zygotes (Figure 2D and Figure S1C). The GFP::PAR-1 gradient was transient and became
less pronounced following NEBD (Figure S1C). Remarkably,
DendraR::MEX-5 also formed a gradient by pronuclear meeting,
which weakened following NEBD (Figure 2D and Figure S1F).
The diffusivity of DendraR::MEX-5 was also asymmetric in par2(RNAi) zygotes (Figure 2C). We conclude that formation of
a cytoplasmic PAR-1 gradient is sufficient to change MEX-5
diffusion and drive the formation of a complementary MEX-5
gradient.
PAR-1 Phosphorylates MEX-5 on Two Residues: S458
and S404
Phosphorylation of S458 depends on par-1 activity in vivo,
raising the possibility that MEX-5 is a PAR-1 substrate (Tenlen
et al., 2008). To test this possibility directly, we expressed the
S458A
S404A
WT
MBP
S404A,
S458A
MBP:MEX-5
B
WT
MBP:PIE-1
par-1(RNAi)
IP: anti-pS404
Blot: anti-MEX-5
P-ATP
Input
32
Coomassie
MEX-5
PAR-1
D
Anterior
Posterior
Ratio (anterior/posterior)
WT
S458A
S404A
WT
let-92(RNAi)
WT
S458A
S404A
+ Embryonic Extract
Relative Immunoblot Signal
+ MBP::PAR-1
S404A
S404A
WT
S404A
pkc-3(RNAi) par-1(b274) let-92(RNAi) let-92(RNAi)
pS458
pS404
pS458 + OA
pS458
pS404 + OA
pS404
01 5
20
60
120 min
01 5
20
60
120 min
PAR-1
Phosphorylation Sites
271
S458
S404
ZF2
K318
R274
MEX-5
M288
F294
ZF1
Y333
F339
342
468
Asymmetry
Diffusion
Anterior
Posterior
Full Length
Yes
1.03 (.11)
2.68 (.24)
aa1-355
No
1.02 (.07)
0.82 (.07)
aa1-245
No
13.3 (1.2)
12.3 (1.6)
aa245-468
Weak
3.22 (.30)
5.93 (.87)
aa345-468
No
> 10
> 10
C
Before PN Formation
NEBD
Ratio (anterior/posterior)
2.5
2.0
1.5
1.0
WT
WT
R274E,K318E R274E,K318E M288E,F294N, M288E,F294N,
mex-5/6(RNAi)
mex-5/6(RNAi) Y333E,F339N Y333E,F339N
mex-5/6(RNAi)
Anterior
Posterior
WT
R274E, K318E
WT
WT
Lat A
observations suggest that, in addition to RNA binding, interactions among MEX-5 and MEX-6 molecules can also contribute
to MEX-50 s diffusive behavior.
The actin cytoskeleton, which becomes enriched in the anterior cytoplasm during polarization, has been proposed as
another candidate for retarding MEX-5 mobility (Tenlen et al.,
2008). To test this possibility, we treated zygotes with Latrunculin
A, which depolymerizes F-actin and blocks polarization of the
PARs (Severson and Bowerman, 2003). Latrunculin A treatment
resulted in uniformly slow MEX-5 diffusion and blocked DendraR::MEX-5 gradient formation (Figure 4C and data not shown),
indicating that F-actin is not essential to retard MEX-5 mobility.
% of Total Dendra::MEX-5
40S
60S
80S
Fraction
WT
par-1(it51)
S404A
% Component
Component
Fast Slow
Fast Slow
Anterior Posterior
Fast Slow
Fast Slow
Anterior Posterior
Fast Slow
Fast Slow
Anterior Posterior
PAR-1
MEX-5 pS404
MEX-5
(5 m2/sec)
(0.07 m2/sec)
PP2A
Cytoplasmic PAR-1 +
kphos 0.1s-1
B
6
fast
slow
total
PAR-1
0.11s-1
PAR-1
0.02s-1
10
15
20
25
30
P O S IT IO N (m )
35
40
45
50
Cortical PAR-1 +
kphos 0.1s-1
PAR-1
0.02s-1
0
10
6
fast
slow
total
15
20
25
30
P O S IT IO N (m )
35
40
45
fast
slow
total
C O N C E N T R AT IO N (AU )
50
Cortical PAR-1 +
kphos 0.01s-1
C O N C E N T R AT IO N (AU )
PAR-1
0.11s-1
fast
slow
total
5
C O N C E N T R AT IO N (AU )
C O N C E N T R AT IO N (AU )
Cytoplasmic PAR-1 +
kphos 0.01s-1
10
15
20
25
30
35
40
45
50
P O S IT IO N (m )
10
15
20
25
30
P O S IT IO N (m )
35
40
45
50
Symbol
Value
Units
Slow diffusion
coefficient
Dslow
0.07
mm2/s
Notes
Fast diffusion
coefficient
Dfast
mm2/s
Kinase (PAR-1)
rate constant
kkin(x)
0.020.11
s1
Linear rise
along A/P axis
Phosphatase rate
constant
kphos
0.1
s1
Uniform along
A/P axis
Embryo length
50
mm
Additional notes:
(1) Kinase and phosphatase rates can be varied coordinately over a range
of values (e.g., kkin(x) = 0.21.1 s1 and kphos = 1 s1 yields similar results).
Rate constants must be approximately equal in the posterior region to
obtain 1:1 slow:fast diffusing species, and kphos > kkin to obtain >1:1
slow:fast in the anterior region.
(2) Kinase rate constant gradient needs to be larger than the MEX-5
gradient. Here it is assumed that the PAR-1 activity gradient is 5.5-fold,
resulting in 2.9-fold MEX-5 concentration gradient.
(3) Posterior cortical-only PAR-1 case modeled with instantaneous
kinase reaction at the right boundary (i.e., x = L), kkin(x) = 0 s1 and kphos =
0.1 s1 (Figure 6D) or kphos=0.01s1 (Figure 6E).
be applied broadly to understanding rapid changes in the distribution of cytoplasmic proteins in a variety of cell types.
EXPERIMENTAL PROCEDURES
Detailed experimental procedures are described in the Extended Experimental
Procedures.
C. elegans Strains
Transgenic worms used in this study are listed in Table S1.
Determination of DendraR::MEX-5 Diffusion Coefficients
Dendra::MEX-5 was photoconverted in a stripe with UV light and imaged on
a spinning disk confocal microscope. Intensity values were fit to Gaussian
distributions for each time point (GraphPad Prism), and the change in variance
over time was used to calculate Dc (Berg, 1993).
Recombinant Protein Purification, Kinase Assays,
and Dephosphorylation Assays
MBP:MEX-5 and MBP:PAR-1 (1492, T325E) were partially purified from
E. coli and incubated at 30 C in the presence of [32P]-ATP or cold ATP. For
nonisotopic phosphorylation and dephosphorylation assays, kinase reactions
were terminated with 20 nM staurosporine before embryonic extract was
added.
Immunoprecipitations
MEX-5 pS404 phosphospecific antibodies coupled to ProteinG dynabeads
were used to immunoprecipitate from whole worm extracts.
Sucrose Gradient Fractionation
Cycloheximide-treated whole worm extracts were fractionated over 10%
45% linear sucrose gradients at 39,000 rpm for 3 hr. Fractions were collected
after passing the gradient through a UV detector, and the distribution of
Dendra::MEX-5 was determined by western blot with anti-Dendra antibodies
(Axxora).
Fluorescence Correlation Spectroscopy
GFP::MEX-5 levels were reduced by partial GFP RNAi depletion prior to
imaging. Embryos were imaged on a Zeiss LSM 510 Confocal microscope
equipped with a Confocor 3 FCS. Autocorrelation curves were analyzed within
the Zeiss Confocor 3 software package.
Modeling of the MEX-5 Gradient
Parameters used in the models are listed in Table 1. A detailed description
of model and the contribution of individual parameters to the steady-state
and unsteady-state models are provided in the Extended Experimental
Procedures.
SUPPLEMENTAL INFORMATION
Supplemental Information includes Extended Experimental Procedures, six
figures, and one table and can be found with this article online at doi:10.
1016/j.cell.2011.08.012.
ACKNOWLEDGMENTS
We thank F. Motegi for providing the Dendra2 construct, S. He for help
with sucrose gradients, and N. Perkins and E. Pryce at the JHU Integrated
Imaging Center for assistance with FCS. We thank K. Kemphues for providing
strains and antibodies. We thank A. Cuenca for generation of the GFP::PAR-1
transgenic strain. We thank S. Kuo for helpful conversations and members of
the Seydoux lab for helpful comments on the manuscript. This work was
supported by the National Institutes of Health (Grants R01HD37047 [G.S.]
and R01 GM71522 [D.J.O.]) and the American Cancer Society (Grant PF-08158-01-DDC [E.E.G.]). G.S. is an Investigator of the Howard Hughes Medical
Institute.
Kao, G., Tuck, S., Baillie, D., and Sundaram, M.V. (2004). C. elegans SUR-6/
PR55 cooperates with LET-92/protein phosphatase 2A and promotes Raf
activity independently of inhibitory Akt phosphorylation sites. Development
131, 755765.
Kemphues, K. (2000). PARsing embryonic polarity. Cell 101, 345348.
Krahn, M.P., Egger-Adam, D., and Wodarz, A. (2009). PP2A antagonizes phosphorylation of Bazooka by PAR-1 to control apical-basal polarity in dividing
embryonic neuroblasts. Dev. Cell 16, 901908.
Lai, W.S., Kennington, E.A., and Blackshear, P.J. (2002). Interactions of CCCH
zinc finger proteins with mRNA: non-binding tristetraprolin mutants exert
an inhibitory effect on degradation of AU-rich element-containing mRNAs.
J. Biol. Chem. 277, 96069613.
Lin, J., Hou, K.K., Piwnica-Worms, H., and Shaw, A.S. (2009). The polarity
protein Par1b/EMK/MARK2 regulates T cell receptor-induced microtubuleorganizing center polarization. J. Immunol. 183, 12151221.
Lipkow, K., and Odde, D.J. (2008). Model for protein concentration gradients in
the cytoplasm. Cell. Mol. Bioeng. 1, 8492.
Cheeks, R.J., Canman, J.C., Gabriel, W.N., Meyer, N., Strome, S., and Goldstein, B. (2004). C. elegans PAR proteins function by mobilizing and stabilizing
asymmetrically localized protein complexes. Curr. Biol. 14, 851862.
Chekulaeva, M., Hentze, M.W., and Ephrussi, A. (2006). Bruno acts as a dual
repressor of oskar translation, promoting mRNA oligomerization and formation
of silencing particles. Cell 124, 521533.
Coppey, M., Boettiger, A.N., Berezhkovskii, A.M., and Shvartsman, S.Y.
(2008). Nuclear trapping shapes the terminal gradient in the Drosophila
embryo. Curr. Biol. 18, 915919.
Cuenca, A.A., Schetter, A., Aceto, D., Kemphues, K., and Seydoux, G. (2003).
Polarization of the C. elegans zygote proceeds via distinct establishment and
maintenance phases. Development 130, 12551265.
Daniels, B.R., Dobrowsky, T.M., Perkins, E.M., Sun, S.X., and Wirtz, D. (2010).
MEX-5 enrichment in the C. elegans early embryo mediated by differential
diffusion. Development 137, 25792585.
Ephrussi, A., and St Johnston, D. (2004). Seeing is believing: the bicoid
morphogen gradient matures. Cell 116, 143152.
Fuller, B.G., Lampson, M.A., Foley, E.A., Rosasco-Nitcher, S., Le, K.V., Tobelmann, P., Brautigan, D.L., Stukenberg, P.T., and Kapoor, T.M. (2008). Midzone
activation of aurora B in anaphase produces an intracellular phosphorylation
gradient. Nature 453, 11321136.
Gallo, C.M., Munro, E., Rasoloson, D., Merritt, C., and Seydoux, G. (2008).
Processing bodies and germ granules are distinct RNA granules that interact
in C. elegans embryos. Dev. Biol. 323, 7687.
Golding, I., and Cox, E.C. (2004). RNA dynamics in live Escherichia coli cells.
Proc. Natl. Acad. Sci. USA 101, 1131011315.
Goldstein, B., and Macara, I.G. (2007). The PAR proteins: fundamental players
in animal cell polarization. Dev. Cell 13, 609622.
Guo, S., and Kemphues, K.J. (1995). par-1, a gene required for establishing
polarity in C. elegans embryos, encodes a putative Ser/Thr kinase that is
asymmetrically distributed. Cell 81, 611620.
Gurskaya, N.G., Verkhusha, V.V., Shcheglov, A.S., Staroverov, D.B., Chepurnykh, T.V., Fradkov, A.F., Lukyanov, S., and Lukyanov, K.A. (2006).
Engineering of a monomeric green-to-red photoactivatable fluorescent protein
induced by blue light. Nat. Biotechnol. 24, 461465.
Hudson, B.P., Martinez-Yamout, M.A., Dyson, H.J., and Wright, P.E. (2004).
Recognition of the mRNA AU-rich element by the zinc finger domain of
TIS11d. Nat. Struct. Mol. Biol. 11, 257264.
Hurd, D.D., and Kemphues, K.J. (2003). PAR-1 is required for morphogenesis
of the Caenorhabditis elegans vulva. Dev. Biol. 253, 5465.
Hurov, J.B., Watkins, J.L., and Piwnica-Worms, H. (2004). Atypical PKC
phosphorylates PAR-1 kinases to regulate localization and activity. Curr.
Biol. 14, 736741.
Kalab, P., Weis, K., and Heald, R. (2002). Visualization of a Ran-GTP gradient
in interphase and mitotic Xenopus egg extracts. Science 295, 24522456.
Cell 146, 955968, September 16, 2011 2011 Elsevier Inc. 967
St Johnston, D., and Ahringer, J. (2010). Cell polarity in eggs and epithelia:
parallels and diversity. Cell 141, 757774.
Su, T.T., Sprenger, F., DiGregorio, P.J., Campbell, S.D., and OFarrell, P.H.
(1998). Exit from mitosis in Drosophila syncytial embryos requires proteolysis
and cyclin degradation, and is associated with localized dephosphorylation.
Genes Dev. 12, 14951503.
Tenlen, J.R., Molk, J.N., London, N., Page, B.D., and Priess, J.R. (2008).
MEX-5 asymmetry in one-cell C. elegans embryos requires PAR-4- and
PAR-1-dependent phosphorylation. Development 135, 36653675.
968 Cell 146, 955968, September 16, 2011 2011 Elsevier Inc.
Vaknin, A., and Berg, H.C. (2004). Single-cell FRET imaging of phosphatase
activity in the Escherichia coli chemotaxis system. Proc. Natl. Acad. Sci.
USA 101, 1707217077.
Wartlick, O., Kicheva, A., and Gonzalez-Gaitan, M. (2009). Morphogen
gradient formation. Cold Spring Harb. Perspect. Biol. 1, a001255.
Yoder, J.H., Chong, H., Guan, K.L., and Han, M. (2004). Modulation of KSR
activity in Caenorhabditis elegans by Zn ions, PAR-1 kinase and PP2A phosphatase. EMBO J. 23, 111119.
SUMMARY
Sip2
415
Ac
335
154
N-Myr
AcAcAc
247
C
HDAC
TSA
Rpd3 Hda1 -
Rpd3 Hda1
+
+
IP: -AcK
Probe: -GST
GBD
Input:
-GST-Sip2p
Snf1 interacon
D
Sip2p-TAP
WT
GST-Sip2p
Sip2p-TAP
Sip2p-TAP
esa1
WT esa1 rpd3 rpd3
WT
rpd3
sip2- sip23KR 4KR
IP: -AcK
Probe: -TAP
Input:
-Sip2p-TAP
Figure 1. Sip2 Is Acetylated at Four Lysine Sites by Esa1, and Rpd3 Is the Counteracting Deacetylase
(A) Cartoon of Sip2 structural domains (adapted from Hedbacker and Carlson, 2008). Mass spectrometry identified four acetylated lysine residues of Sip2, K12,
K16, K17, and K256. Numbers indicate amino acid residues. (Vertical straight line) Myristoylation site; (downward arrows) acetylation sites; (right-left arrow)
regions mapped as sufficient for Snf1 interaction by deletion analysis (Amodeo et al., 2007). Ac, acetylation; N-Myr, N-myristoylation; GBD, glycogen-binding
domain.
(B) Chromosomally integrated sip2-3KR and sip2-4KR, but not sip2-G2A, mutants are hypoacetylated in vivo. The sip2-G2A mutant blocks myristoylation.
sip2-3KR, sip2-K12/16/17R; sip2-4KR, sip2-K12/16/17/256R.
(C) Rpd3, but not Hda1, removes Sip2 acetylation in vitro. Deacetylation reaction is inhibited by addition of trichostatin A (TSA). Asterisk indicates Hda1-TAP;
arrowhead indicates Rpd3-TAP.
(D) Sip2 is hypoacetylated in strains carrying the Ts allele of ESA1, esa1-531, but is hyperacetylated in rpd3D; deletion of RPD3 rescues the acetylation defect of
esa1-531.
(E) Increased Sip2 acetylation in rpd3D is blocked when lysines are mutated to arginines.
See also Figure S1 and Figure S2.
constructs at these four sites in various combinations by sitedirected mutagenesis and introduced these mutant SIP2
constructs into the endogenous chromosomal locus (Toulmay
and Schneiter, 2006). Using a previously described reverse IP
approach (Lin et al., 2009), we showed that Sip2 is hypoacetylated when the first three (K12, K16, and K17 [3KR]) or all
four (K12, K16, K17, and K256 [4KR]) lysines are mutated
(Sip2-3KR or -4KR; Figures S1D, Figure 1B). We also created
a chromosomally integrated SIP2 glycine 2-to-alanine mutant
(sip2-G2A) to mimic the short-lived, non-N-myristoylated
species of Sip2 (Ashrafi et al., 2000). We found that sip2-G2A
did not have acetylation defects (Figure 1B). An in vitro deacetylation reaction carried out with purified Rpd3-TAP and Hda1TAP (Figure S2A) revealed that Rpd3 treatment abolished the
acetylation signal of Sip2 (Figure 1C). We demonstrated the
activity of purified Hda1-TAP by showing that it deacetylated
lysine 14 of FLAG-Htz1 purified from hda1D strain (Figure S2B)
(Lin et al., 2008). As will be shown below, mutagenesis of the
Sip2 acetylation sites affects replicative aging. The findings are
consistent with a previous report that Rpd3, but not Hda1, is
involved in replicative life span regulation (Kim et al., 1999).
Previously we have shown that the acetylation signals of Sip2
are virtually completely dependent on NuA4/Esa1 both in vivo
and in vitro (Lin et al., 2009). Importantly, simultaneous deletion
Figure 3. Glucose Limitation and Replicative Life Span in SIP2 Acetylation Mutants
Survival curves of WT (A), sip2D (B), sip2-4KR (C),
and sip2-4KQ (D) grown in normal (2%, NG) versus
low glucose (0.05%, LG) media. The fractions of
live cells are plotted as a function of age in
generations. Median life span is shown in parentheses. Glucose limitation significantly increases
life span in WT, sip2D, and sip2-4KR, but not in
sip2-4KQ, as determined by Mantel-Cox log-rank
test. The statistical results were summarized in
Table S1. See also Figure S3.
that, although the interaction between Sip2 and Snf1 was readily
seen in young cells, it became undetectable in old cells
(Figure 4B).
To determine whether Sip2 acetylation dictates physical interaction between Sip2 and Snf1, we examined their association in
SIP2 acetylation mutants. Physical interaction between Sip2 and
Snf1 significantly decreased in esa1-531 but increased in rpd3D
(Figure 4C). More importantly, whereas the sip2-3KR and
sip2-4KR mutations almost abolished the interaction with Snf1,
the sip2-3KQ and sip2-4KQ mutations markedly enhanced it
(Figure 4D), supporting the idea Sip2 acetylation is required for
Snf1 interaction and defining the N terminus of Sip2 as an
enhancer of Snf1 interaction (Figure 1A). In contrast, the myristoylation null mutation sip2-G2A (Lin et al., 2003) did not affect
the physical interaction between Sip2 and Snf1 (Figure 4D), suggesting that myristoylation regulates Snf1 activity via a different
mechanism. Because Sip2 is a negative regulator of Snf1 (Ashrafi et al., 2000), these results at least partially explain why
Snf1 activity, which is stimulated under carbon stress, increases
with age and hence has detrimental effects on life span, even in
the presence of abundant ambient glucose (Lin et al., 2003).
To establish the connection between aging and Sip2 acetylation status in cells, we measured trehalose levels, a general
stress indicator in yeast, in young and old cells. This is based
on the previous findings that, during the quiescent phase, trehalose is stored in favor of glycogen presumably to fulfill its numerous stress-protectant functions (Shi et al., 2010). Indeed,
we found that young cells contained significantly less trehalose
than old cells (Figure 4E). Glucose limitation, a form of nutrient
stress, also significantly increased trehalose levels as reported
before (Figure S3D) (Pluskal et al., 2011). Measurement of trehalose levels revealed that esa1-531 contained elevated trehalose
levels that were partially reversed by concomitant RPD3 deletion
(Figure S3E). Sip2-4KR rpd3D and sip2D rpd3D increased treha972 Cell 146, 969979, September 16, 2011 2011 Elsevier Inc.
lose levels compared to rpd3D (Figure S3F). Also, sip2-4KR, sip2-G2A, and
sip2D mutants had increased trehalose
levels compared to WT, whereas sip24KQ had a lower level of trehalose (Figure 4F), supporting a role of Ac-Sip2 in
counteracting stress. We further showed
that sip2-4KR and sip2D mutants were
much more sensitive to hydrogen peroxide (H2O2), a form of oxidative stress
and a mediator of aging (Giorgio et al., 2007), when compared
to WT (Figure 4G); on the contrary, sip2-4KQ was resistant to
H2O2 (Figure 4G).
Consistent with previous findings that trehalose plays a key
role in fueling cell-cycle progression during growth and division
(Shi et al., 2010), we found that growth rates significantly
decreased in the rpd3D strain (Figure S4A) and the sip2-4KQ
strain on various genetic backgrounds (Figures S4B and S4C).
Conversely, sip2-4KR and sip2D mutants partially rescued the
rpd3D growth defect (Figure S4A). Cell volumes were similarly
affected by the mutations; all of the slower-growing variants
were smaller than normal cells (Figure S4D).
Sch9 Is the Downstream Target of Sip2-Snf1
Based on the above results and literature information, we
proposed that Snf1 might have a downstream target that,
when phosphorylated by Snf1, fuels rapid cell growth but results
in shortened life span. In a previous phosphorylation proteome
microarray study, 80 in vitro Snf1 substrates were identified (Ptacek et al., 2005). Among these, Sch9 (an Akt/S6K homolog)
kinase was a promising candidate because it had already been
shown to play an important role in yeast aging (Kaeberlein
et al., 2005). In addition to Snf1, Sch9 is the in vitro substrate
of two additional kinases (of 87 kinases tested): Tos3, the
upstream kinase of Snf1 (Kim et al., 2005) and Pho85 (Ptacek
et al., 2005). Sch9 is also a well-known in vivo kinase substrate
of target of rapamycin complex 1 (TORC1). Similar to TORC1,
it negatively regulates life span in response to nutrient availability
(Kaeberlein et al., 2005; Urban et al., 2007). A kinase itself, Sch9
phosphorylates a critically important ribosomal protein, Rps6
(Urban et al., 2007); rps6D mutants grow slowly and are small,
but they have increased replicative life span through general
inhibition of translation machinery (Chiocchetti et al., 2007) and
protein synthesis (Huber et al., 2009).
Sip2-TAP
Young Old
Young Old
IP
IgG HC
Input
WT
esa1 rpd3
IP: TAP
-HA
Input:
-Snf1-HA
IP: TAP
-HA
Input:
-Snf1-HA
Input:
-Sip2-TAP
Input:
-Sip2-TAP
+
+
+
+
+
+
+
+
+
+
+
+
Trehalose
Input:
-Sip2-TAP
Input:
-Snf1-HA
12.5
10.0
7.5
5.0
2.5
0.0
-Tubulin
G0
3
2
1
Fracon survival
Trehalose
10
**
ns
G7
***
1
0.1
0.01
H2O2
0 min
60 min
0.001
Figure 4. Sip2 Acetylation and Physical Interaction between Sip2 and Snf1 Decrease as Cells Age
(A) Sip2 acetylation is significantly decreased in old cells.
(B and C) Physical interaction between Sip2 and Snf1 assessed by coimmunoprecipitation is decreased in old cells (B) and esa1-531 strain but increased in rpd3D
strain (C).
(D) Interaction between Sip2 and Snf1 is decreased in sip2D, and deacetylation mimetics (sip2-3KR and sip2-4KR) but is not changed in sip2-G2A mutants.
The interaction significantly increases in acetylation mimetics (sip2-3KQ and sip2-4KQ mutants).
(E) Cellular trehalose is significantly increased when cells age. Error bars show standard error of the mean. n = 3.
(F) Trehalose levels are significantly higher in sip2-4KR, sip2-G2A, and sip2D when compared to WT but are lower in sip2-4KQ. Error bars show standard error of
the mean. n = 3.
(G) Fractions of survival of WT, sip2-4KR, sip2-4KQ, and sip2D before and after 1 hr of treatment with 1.5 mM H2O2, are plotted on a log scale. Error bars
show standard error of the mean. n = 2. Statistical significance was assessed by two-way ANOVA with post-hoc test. ns, nonsignificant; *p < 0.05; **p < 0.01;
***p < 0.001.
See also Figure S3 and Figure S4.
Sch9-HA phosphorylation markedly decreased in snf1D especially after treatment with rapamycin to block TORC1 activity
(Figure 5B and whole gel images in Figure S5B). Moreover, a
mutation that abolishes the kinase activity of Snf1 (snf1-K84R)
fails to revert the decreased phosphorylation of endogenous
Sch9-HA in a snf1D strain (Figure S5C), which suggests that
Sch9 is a bona fide in vivo substrate of Snf1.
To detect Sch9 phosphorylation states in different SIP2
mutants, we purified GST-Sch9 proteins from WT, sip2-4KR,
sip2-4KQ, and sip2D strains and assessed their phosphorylation
signals. Both serine and threonine phosphorylation increased
slightly in sip2-4KR and sip2D compared to WT and sip2-4KQ
Cell 146, 969979, September 16, 2011 2011 Elsevier Inc. 973
Snf1
ATP
GST-Sch9
+
+
-
Sch9-HA
+
+
Rapa
IP: HA
-P-Ser
IP: HA
-P-Thr
-P-Ser
-P-Thr
WT snf1 WT
+
-
In vitro
GST-Sch9 (+ Rapamycin)
snf1
+
-HA
-GST
-GST
-Tubulin
In vivo
Sch9-HA
Young Old
IP: HA
-P-Ser
IP: HA
-P-Thr
-HA
YPD
YPD + Rapamycin
WT
sch9
sch9
sip2-4KR
sch9
sip2
WT (24)
sip2 sch9 (43.5)
1.0
0.8
0.6
sch9 (51.5)
0.4
0.2
0.0
0
25
50
75
100
Figure 5. Acetylation of Sip2 Affects Snf1 Kinase Activity and Results in Hypophosphorylation of Sch9
(A) Snf1 phosphorylates GST-Sch9 in vitro at both serine and threonine sites. In vitro kinase assays were performed by incubating purified GST-Sch9 with or
without Snf1 or ATP as indicated and analyzed by phosphoserine antibody (a-P-Ser) and phosphothreonine antibody (a-P-Thr) to detect Sch9p phosphorylation.
(B) Endogenous Sch9-HA phosphorylation decreases in snf1D mutants after rapamycin (200 ng/ml) treatment to suppress the TOR pathway. Arrowheads
indicate the Sch9-HA band.
(C) GST-Sch9 phosphorylation increases in sip2-4KR and sip2D but decreases in sip2-4KQ, compared to WT after rapamycin (200 ng/ml) treatment.
(D) Endogenous Sch9-HA phosphorylation significantly increases in the old cells. Arrowheads indicate the Sch9-HA band.
(E) SCH9 is epistatic and thus downstream to SIP2 in regulating cellular growth. Ten-fold dilutions of the indicated strains were spotted and grown on YPD plates
without (2 days, 30 C) or with rapamycin (25 ng/ml, 4 days, 30 C).
(F) Deletion of SCH9 rescues the life span shortening of sip2D and sip2-4KR.
See also Figures S1 and Figure S5.
We thank Sheng-Ce Tao and Chien-Sheng Chen for their early contribution
to this work. We are grateful to Dr. Brian K. Kennedy for providing sch9D
strain, Dr. Robbie Loewith for providing SCH9 phosphorylation mutants, and
Dr. Marian B. Carlson for providing snf1-K84R construct. We thank
Dr. Fang-Jen Lee for kindly sharing the laboratory space and reagents. We
thank Department of Medical Science, National Taiwan University for technical
help with sequencing. We thank Dr. Yi-Juang Chern for critical suggestions on
this work. Work was supported by National Science Council (NSC 98-2314-B002-031-MY3 to J.-Y.L.), National Taiwan University Hospital (099-001376,
to J.-Y.L.), National Taiwan University (99C101-603 to J.-Y.L., Y.-Y.L., and
L.-M.C.), Liver Disease Prevention & Treatment Research Foundation (Taiwan)
(to J.-Y.L. and Y.-Y.L.), and NIH Common Fund grant (USA) (U54-RR020839
to H.Z. and J.D.B.).
Received: January 18, 2011
Revised: May 2, 2011
Accepted: July 29, 2011
Published online: September 8, 2011
REFERENCES
Amodeo, G.A., Rudolph, M.J., and Tong, L. (2007). Crystal structure of the
heterotrimer core of Saccharomyces cerevisiae AMPK homologue SNF1.
Nature 449, 492495.
Cell 146, 969979, September 16, 2011 2011 Elsevier Inc. 977
Ashrafi, K., Lin, S.S., Manchester, J.K., and Gordon, J.I. (2000). Sip2p and its
partner snf1p kinase affect aging in S. cerevisiae. Genes Dev. 14, 18721885.
Barker, M.G., Brimage, L.J., and Smart, K.A. (1999). Effect of Cu,Zn superoxide dismutase disruption mutation on replicative senescence in Saccharomyces cerevisiae. FEMS Microbiol. Lett. 177, 199204.
Bitterman, K.J., Medvedik, O., and Sinclair, D.A. (2003). Longevity regulation in
Saccharomyces cerevisiae: linking metabolism, genome stability, and heterochromatin. Microbiol. Mol. Biol. Rev. 67, 376399.
Chang, C.S., and Pillus, L. (2009). Collaboration between the essential Esa1
acetyltransferase and the Rpd3 deacetylase is mediated by H4K12 histone
acetylation in Saccharomyces cerevisiae. Genetics 183, 149160.
Chang, K.T., and Min, K.T. (2002). Regulation of lifespan by histone deacetylase. Ageing Res. Rev. 1, 313326.
Chiocchetti, A., Zhou, J., Zhu, H., Karl, T., Haubenreisser, O., Rinnerthaler, M.,
Heeren, G., Oender, K., Bauer, J., Hintner, H., et al. (2007). Ribosomal proteins
Rpl10 and Rps6 are potent regulators of yeast replicative life span. Exp.
Gerontol. 42, 275286.
Close, P., Creppe, C., Gillard, M., Ladang, A., Chapelle, J.P., Nguyen, L., and
Chariot, A. (2010). The emerging role of lysine acetylation of non-nuclear
proteins. Cell. Mol. Life Sci. 67, 12551264.
Cohen, E., and Dillin, A. (2008). The insulin paradox: aging, proteotoxicity and
neurodegeneration. Nat. Rev. Neurosci. 9, 759767.
Colman, R.J., and Anderson, R.M. (2011). Nonhuman primate calorie restriction. Antioxid. Redox Signal. 14, 229239.
Dang, W., Steffen, K.K., Perry, R., Dorsey, J.A., Johnson, F.B., Shilatifard, A.,
Kaeberlein, M., Kennedy, B.K., and Berger, S.L. (2009). Histone H4 lysine 16
acetylation regulates cellular lifespan. Nature 459, 802807.
Dirks, A.J., and Leeuwenburgh, C. (2006). Caloric restriction in humans:
potential pitfalls and health concerns. Mech. Ageing Dev. 127, 17.
Lin, S.S., Manchester, J.K., and Gordon, J.I. (2003). Sip2, an N-myristoylated
beta subunit of Snf1 kinase, regulates aging in Saccharomyces cerevisiae by
affecting cellular histone kinase activity, recombination at rDNA loci, and
silencing. J. Biol. Chem. 278, 1339013397.
Ehrentraut, S., Weber, J.M., Dybowski, J.N., Hoffmann, D., and EhrenhoferMurray, A.E. (2010). Rpd3-dependent boundary formation at telomeres by
removal of Sir2 substrate. Proc. Natl. Acad. Sci. USA 107, 55225527.
Lin, Y.Y., Qi, Y., Lu, J.Y., Pan, X., Yuan, D.S., Zhao, Y., Bader, J.S., and Boeke,
J.D. (2008). A comprehensive synthetic genetic interaction network governing
yeast histone acetylation and deacetylation. Genes Dev. 22, 20622074.
Evans, D.S., Kapahi, P., Hsueh, W.C., and Kockel, L. (2011). TOR signaling
never gets old: aging, longevity and TORC1 activity. Ageing Res. Rev. 10,
225237.
Lin, Y.Y., Lu, J.Y., Zhang, J., Walter, W., Dang, W., Wan, J., Tao, S.C., Qian, J.,
Zhao, Y., Boeke, J.D., et al. (2009). Protein acetylation microarray reveals that
NuA4 controls key metabolic target regulating gluconeogenesis. Cell 136,
10731084.
Giorgio, M., Migliaccio, E., Orsini, F., Paolucci, D., Moroni, M., Contursi, C.,
Pelliccia, G., Luzi, L., Minucci, S., Marcaccio, M., et al. (2005). Electron transfer
between cytochrome c and p66Shc generates reactive oxygen species that
trigger mitochondrial apoptosis. Cell 122, 221233.
Giorgio, M., Trinei, M., Migliaccio, E., and Pelicci, P.G. (2007). Hydrogen
peroxide: a metabolic by-product or a common mediator of ageing signals?
Nat. Rev. Mol. Cell Biol. 8, 722728.
Guarente, L., and Kenyon, C. (2000). Genetic pathways that regulate ageing in
model organisms. Nature 408, 255262.
Haigis, M.C., and Yankner, B.A. (2010). The aging stress response. Mol. Cell
40, 333344.
Hedbacker, K., and Carlson, M. (2008). SNF1/AMPK pathways in yeast. Front.
Biosci. 13, 24082420.
Huber, A., Bodenmiller, B., Uotila, A., Stahl, M., Wanka, S., Gerrits, B.,
Aebersold, R., and Loewith, R. (2009). Characterization of the rapamycinsensitive phosphoproteome reveals that Sch9 is a central coordinator of
protein synthesis. Genes Dev. 23, 19291943.
Liu, F., Benashski, S.E., Persky, R., Xu, Y., Li, J., and McCullough, L.D. (2011).
Age-related changes in AMP-activated protein kinase after stroke. Age
(Dordr). Published online March 1, 2011. 10.1007/s11357-011-9214-8.
Madia, F., Wei, M., Yuan, V., Hu, J., Gattazzo, C., Pham, P., Goodman, M.F.,
and Longo, V.D. (2009). Oncogene homologue Sch9 promotes age-dependent
mutations by a superoxide and Rev1/Polzeta-dependent mechanism. J. Cell
Biol. 186, 509523.
Mair, W., Morantte, I., Rodrigues, A.P., Manning, G., Montminy, M., Shaw,
R.J., and Dillin, A. (2011). Lifespan extension induced by AMPK and calcineurin
is mediated by CRTC-1 and CREB. Nature 470, 404408.
Marino, G., Ugalde, A.P., Salvador-Montoliu, N., Varela, I., Quiros, P.M.,
Cadinanos, J., van der Pluijm, I., Freije, J.M., and Lopez-Otn, C. (2008).
Premature aging in mice activates a systemic metabolic response involving
autophagy induction. Hum. Mol. Genet. 17, 21962211.
Migliaccio, E., Giorgio, M., Mele, S., Pelicci, G., Reboldi, P., Pandolfi, P.P.,
Lanfrancone, L., and Pelicci, P.G. (1999). The p66shc adaptor protein controls
oxidative stress response and life span in mammals. Nature 402, 309313.
Imai, S., Armstrong, C.M., Kaeberlein, M., and Guarente, L. (2000). Transcriptional silencing and longevity protein Sir2 is an NAD-dependent histone
deacetylase. Nature 403, 795800.
Nakamura, A., Kawakami, K., Kametani, F., Nakamoto, H., and Goto, S. (2010).
Biological significance of protein modifications in aging and calorie restriction.
Ann. N Y Acad. Sci. 1197, 3339.
Jablonowski, D., Fichtner, L., Stark, M.J., and Schaffrath, R. (2004). The yeast
elongator histone acetylase requires Sit4-dependent dephosphorylation for
toxin-target capacity. Mol. Biol. Cell 15, 14591469.
978 Cell 146, 969979, September 16, 2011 2011 Elsevier Inc.
Pluskal, T., Hayashi, T., Saitoh, S., Fujisawa, A., and Yanagida, M. (2011).
Specific biomarkers for stochastic division patterns and starvation-induced
quiescence under limited glucose levels in fission yeast. FEBS J. 278, 1299
1315.
Thomson, D.M., Brown, J.D., Fillmore, N., Ellsworth, S.K., Jacobs, D.L.,
Winder, W.W., Fick, C.A., and Gordon, S.E. (2009). AMP-activated protein
kinase response to contractions and treatment with the AMPK activator AICAR
in young adult and old skeletal muscle. J. Physiol. 587, 20772086.
Powers, T. (2007). TOR signaling and S6 kinase 1: Yeast catches up. Cell
Metab. 6, 12.
Toulmay, A., and Schneiter, R. (2006). A two-step method for the introduction
of single or multiple defined point mutations into the genome of Saccharomyces cerevisiae. Yeast 23, 825831.
Ptacek, J., Devgan, G., Michaud, G., Zhu, H., Zhu, X., Fasolo, J., Guo, H., Jona,
G., Breitkreutz, A., Sopko, R., et al. (2005). Global analysis of protein phosphorylation in yeast. Nature 438, 679684.
Puig, O., Caspary, F., Rigaut, G., Rutz, B., Bouveret, E., Bragado-Nilsson, E.,
Wilm, M., and Seraphin, B. (2001). The tandem affinity purification (TAP)
method: a general procedure of protein complex purification. Methods 24,
218229.
Rundlett, S.E., Carmen, A.A., Kobayashi, R., Bavykin, S., Turner, B.M., and
Grunstein, M. (1996). HDA1 and RPD3 are members of distinct yeast histone
deacetylase complexes that regulate silencing and transcription. Proc. Natl.
Acad. Sci. USA 93, 1450314508.
Sanz, P. (2003). Snf1 protein kinase: a key player in the response to cellular
stress in yeast. Biochem. Soc. Trans. 31, 178181.
Sanz, P. (2008). AMP-activated protein kinase: structure and regulation. Curr.
Protein Pept. Sci. 9, 478492.
Schmidt, M.C., and McCartney, R.R. (2000). beta-subunits of Snf1 kinase are
required for kinase function and substrate definition. EMBO J. 19, 49364943.
Urban, J., Soulard, A., Huber, A., Lippman, S., Mukhopadhyay, D., Deloche,
O., Wanke, V., Anrather, D., Ammerer, G., Riezman, H., et al. (2007). Sch9 is
a major target of TORC1 in Saccharomyces cerevisiae. Mol. Cell 26, 663674.
Vijg, J., and Campisi, J. (2008). Puzzles, promises and a cure for ageing.
Nature 454, 10651071.
Wang, J., Jiang, J.C., and Jazwinski, S.M. (2010). Gene regulatory changes in
yeast during life extension by nutrient limitation. Exp. Gerontol. 45, 621631.
Wang, W., Yang, X., Lopez de Silanes, I., Carling, D., and Gorospe, M. (2003).
Increased AMP:ATP ratio and AMP-activated protein kinase activity during
cellular senescence linked to reduced HuR function. J. Biol. Chem. 278,
2701627023.
Wang, Y., Liang, Y., and Vanhoutte, P.M. (2011). SIRT1 and AMPK in regulating
mammalian senescence: a critical review and a working model. FEBS Lett.
585, 986994.
Scott, J.W., Oakhill, J.S., and van Denderen, B.J. (2009). AMPK/SNF1 structure: a menage a trois of energy-sensing. Front. Biosci. 14, 596610.
Williams, D.S., Cash, A., Hamadani, L., and Diemer, T. (2009). Oxaloacetate
supplementation increases lifespan in Caenorhabditis elegans through an
AMPK/FOXO-dependent pathway. Aging Cell 8, 765768.
Shi, L., Sutter, B.M., Ye, X., and Tu, B.P. (2010). Trehalose is a key determinant
of the quiescent metabolic state that fuels cell cycle progression upon return to
growth. Mol. Biol. Cell 21, 19821990.
Silva, H., and Conboy, I.M. (2008). Aging and stem cell renewal (Cambridge,
MA: In StemBook).
Zhou, J., Zhou, B.O., Lenzmeier, B.A., and Zhou, J.Q. (2009). Histone
deacetylase Rpd3 antagonizes Sir2-dependent silent chromatin propagation.
Nucleic Acids Res. 37, 36993713.
Zu, Y., Liu, L., Lee, M.Y., Xu, C., Liang, Y., Man, R.Y., Vanhoutte, P.M., and
Wang, Y. (2010). SIRT1 promotes proliferation and prevents senescence
through targeting LKB1 in primary porcine aortic endothelial cells. Circ. Res.
106, 13841393.
Cell 146, 969979, September 16, 2011 2011 Elsevier Inc. 979
SUMMARY
1200
VEH
AAL-R
CYM-5442
***
900
600
***
300
40
*
***
*
***
20
0
***
**
**
IFN-
CCL2
IL-6
TNF-
IFN-
Cytokines/chemokines
3
***
VEH
AAL-R
CYM-5442
2
1
**
0.8
***
0.6
**
0.4
***
0.2
0
*
Macrophages/
monocytes
Neutrophils
NK cells
Cell Populations
CD69 MFI on
macrophages/monocytes
C
% of Max
4000
***
**
3000
2000
1000
D
CD69 MFI on NK cells
1500
% of Max
RESULTS
4000
**
3000
2000
1000
0
VEH
AAL-R CYM-5442
CD69
Figure 1. S1P1 Receptor Agonism Suppresses Early Proinflammatory Cytokine and Chemokine Production and Innate
Immune Cell Recruitment during Influenza Virus Infection
Mice were infected with 1 3 104 PFU WSN influenza virus, and vehicle (water),
AAL-R (0.2 mg/kg) (1 hr postinfection), or CYM-5442 (2 mg/kg) (1,13, 25, and
37 hr postinfection) were administered i.t. to mice.
(A) Proinflammatory cytokines and chemokines were measured 48 hr postinfection in BALF by ELISA.
(B) Total numbers of innate immune cells were quantified from collagenasedigested lungs by flow cytometry at 48 hr postinfluenza virus infection.
(C) Histograms (left) and mean fluorescent intensity (MFI) (right) of CD69
expression on macrophages was quantified on vehicle, AAL-R-, or CYM-5442treated mice 48 hr postinfluenza virus infection by flow cytometry staining.
(D) Histograms (left) and MFI (right) of NK cell CD69 expression quantified as
in (C).
Data represent average SEM from four to mice per group. *p < 0.05;
**p < 0.005; ***p < 0.0005. Results are representative of greater than six
independent experiments. See also Figure S1 and Figure S2.
Cell 146, 980991, September 16, 2011 2011 Elsevier Inc. 981
2000
Vehicle
CYM-5442
1500
1000
500
400
300
200
100
0
**
**
**
**
***
1.0
**
0
CD69+ Macrophages
& monocytes
Cytokines/Chemokines
CD69+ NK cells
Vehicle
RP-002
1500
1000
500
400
300
200
100
0
***
**
*
**
**
**
IL-1
Cytokines/Chemokines
D
2000
0.5
IL-1
Vehicle
CYM-5442
1.5
2.0
Vehicle
RP-002
1.5
1.0
**
0.5
0
CD69+ Macrophages/
monocytes
CD69+ NK cells
Figure 2. S1P1 Receptor Agonism Suppresses Early Proinflammatory Cytokine and Chemokine Production and Recruitment of Activated
Innate Immune Cells during Human Pathogenic H1N1:2009 Swine Influenza Virus Infection
(AD) Mice were infected with 1 3 105 PFU A/Wisconsin/WSLH34939/09 influenza virus, and either vehicle (water), CYM-5442 (2 mg/kg) (1, 13, 25, and 37 hr
postinfection), or RP-002 (2 mg/kg on 1 and 25 hr postinfection) were administered i.t. to mice. Proinflammatory cytokines and chemokines were measured 48 hr
postinfection in BALF by ELISA in either CYM-5442- (A) or RP-002-treated mice (C). Total numbers of innate immune cells were quantified from collagenasedigested lungs by flow cytometry at 48 hr postinfluenza virus infection in mice treated with either CYM-5442 (B) or RP-002 (D). Data represent average SEM from
four to five mice per group. *p < 0.05; **p < 0.005; ***p < 0.0005. Results are representative of two independent experiments. See also Figure S3.
CYM-5442, RP-002 treatment significantly inhibited the production of multiple proinflammatory cytokine and chemokines
(Figure 2C) and suppressed the accumulation of activated
(CD69+) macrophages/monocytes and NK cells in the infected
lung 48 hr postinfection (Figure 2D). Moreover, RP-002-mediated suppression of innate immune cell recruitment and cytokine/chemokine production occurred without altering lung viral
titers (Figure S3), demonstrating that S1P1 agonist-mediated
suppression of cytokines and chemokines is not due to direct
effects on influenza virus replication.
S1P1 Agonist-Mediated Suppression of Early Innate
Immune Responses Results in Protection to Human
Pathogenic Influenza Virus Challenge in Mice
Early dysregulated innate immune responses in the lung have
been associated with morbidity and mortality during infection
with highly pathogenic strains of influenza virus (Cilloniz et al.,
2009; Kobasa et al., 2007). Therefore, we asked whether blunting
early innate cytokine/chemokine responses using an S1P1
agonist could protect mice from lethal infection with a virulent
human isolate of pandemic 2009 H1N1 that had not been
passaged in mice (A/Wisconsin/WSLH/34939/09). In order not
to impose additional stress on the infected lung and because
oral delivery is a popular route for drug administration, we
orally administered the S1P1 agonist, RP-002, at 6 mg/kg
by gavage to C57BL/6J mice infected with 1 3 105 PFU of
800
Vehicle
RP-002
600
400
200
150
100
**
50
**
IFN-
CCL5
IL-1
IL-6
TNF-
Cytokines/chemokines
B
Total # of cells in the lung
of C57Bl/6J mice (x105)
10
Vehicle
RP-002
8
6
**
4
2
***
0
CD69+ Macrophages/
monocytes
C
Percent Survival
CD69+ NK cells
100
75
50
Vehicle
RP-002
6 mg/kg p.o.
25
0
0
10
12
14
16
Days post-infection
Figure 3. Oral Administration of an S1P1 Agonist Suppresses Early
Innate Cytokine and Chemokine Production and Significantly
Improves Survival to Lethal Infection with H1N1 2009 Swine Influenza Virus
Mice were infected with 2 3 105 PFU A/Wisconsin/WSLH34939/09 and treated
by gavage with either vehicle (water) or RP-002 (6 mg/kg on 1 and 25 hr
postinfection for cytokine and cell recruitment assays at 48 hr, with a third dose
administered at 49 hr for the 16 day survival experiment).
(A) Levels of proinflammatory cytokines and chemokines were analyzed in the
BALF 48 hr postinfection by ELISA.
(B) Total numbers of activated (CD69+) macrophages/monocytes and NK cells
were detected in collagenase-digested lungs 48 hr postinfection by flow
cytometry.
(C) Mice were monitored daily for survival for 16 days postinfection.
For (A) and (B), data represent the average SEM of four to five mice per group,
whereas (C) had ten mice per group. *p < 0.05; **p < 0.005; ***p < 0.0005.
Data are representative of three independent experiments. See also Figure S3.
Cell 146, 980991, September 16, 2011 2011 Elsevier Inc. 983
Lymphatic Endothelium
Vascular Endothelium
Epithelium
4.59
0.95
S1P1-eGFP
expression
% of Max
5.47
S1P1-eGFP
>
B
CD4 T Cells
CD8 T Cells
3.00
B Cells
1.77
S1P1-eGFP
expression
% of Max
3.98
C57BL/6J
S1P1 - eGFP
S1P1-eGFP
>
C
Macrophage/Monocytes
Alveolar Macrophages
0.96
1.20
NK Cells
1.29
1.12
S1P1-eGFP
expression
% of Max
1.44
Neutrophils
Dendritic Cells
S1P1-eGFP
>
Relative S1P1-eGFP
expression
8
Prior to Infection
48-Hours post-infection
6
4
2
lls
ce
ils
N
ph
ro
eu
t
ph
ag
e
ce
lls
M
ac
ro
8
D
C
ce
ce
l
ls
lls
m
el
iu
ith
Ep
ph
m
Ly
Va
sc
ul
ar
at
ic
Cell Types
Figure 4. S1P1 Receptor Is Expressed on Pulmonary Endothelium and Lymphocytes, but Not on Pulmonary Epithelial Cells
(AC) Flow cytometry histograms showing eGFP fluorescence from heterozygous S1P1-eGFP on lung endothelial and epithelial cells. (A) CD4 and CD8 T cells and
B cells. (B) Macrophages, neutrophils, dendritic cells, and NK cells (C) prior to influenza virus infection.
(D) Relative expression of S1P1-eGFP on lung cell populations before or 48 hr after influenza virus infection. The numbers in the upper-right corner of each plot
represent the relative eGFP expression as calculated by dividing the MFI of eGFP on S1P1-eGFP transgenic mice over eGFP MFI of C57BL/6J littermate controls.
Data represent average SD from two to five mice per group. Results are representative of three independent experiments. See also Figure S4.
984 Cell 146, 980991, September 16, 2011 2011 Elsevier Inc.
A
pg/mL in the BALF of
RAG2-/- mice
800
Vehicle
CYM-5442
600
400
*
200
***
***
**
0
IFN-
CCL2
IL-6
TNF-
IFN-
Chemokine/Cytokine
B
Number of cells in
6
RAG2 -/- mice (x10 )
Vehicle
CYM-5442
**
*
0
Macrophages/
monocytes
Neutrophils
NK cells
Cell Types
Figure 5. S1P1 Receptor Agonism Suppresses Early Proinflammatory Immune Responses Independently of Lymphocytes
Rag2 / mice were infected and treated with either vehicle or CYM-5442, as in
Figure 1.
(A) Proinflammatory cytokines and chemokines were measured in BALF by
ELISA 48 hr postinfluenza infection.
(B) Total numbers of innate immune cell recruitment were quantified from
collagenase-digested lungs by flow cytometry at 48 hr postinfluenza virus
infection.
Data represent average SEM from five mice per group. *p < 0.05; **p < 0.005;
***p < 0.0005. Results are representative of two independent experiments. See
also Figure S1 and Figure S5.
50
40
30
20
10
Vehicle
CYM-5442
6
5
4
3
2
1
CCL2
CCL5
Lymphatic
Endothelium
Vascular
Endothelium
CXCL10
Cytokines/Chemokines
25
25
20
20
**
15
15
10
10
800
1200
600
800
400
***
400
200
0
**
VEH
CYM-5442
VEH
CYM-5442
Treatment
Treatment
E
C
Harvest lung
tissue 2 days
post-infection
and run FACS
48 hours post
infection
LysM-GFP
C57Bl/6
2.5
***
2.0
1.5
Vehicle
CYM -5442
Uninfected Vehicle
Unifected CYM -5442
1.0
***
0.5
0
Granulocytes
Macrophages/
Monocytes
**
VEH
CYM-5442
CYM-5442 + CCL2
***
**
**
Macrophages/
monocytes
***
NK cells
Neutrophils
Cell Types
3000
VEH
CYM-5442
CYM-5442 + CCL2
2000
1000
400
300
***
**
*
***
*
200
100
0
Cytokines/chemokines
Figure 6. S1P1 Receptor Agonism Actively Suppresses Recruitment of Innate Immune Cells through Downregulation of Chemokine
Expression on Lung Endothelial Cells
(A) Relative mRNA expression of chemokines from purified lung endothelial cell populations at 36 hr postinfluenza virus infection.
(B) Protein expression of CCL2 and CXCL10 measured by ELISA on FACS-purified vascular and lymphatic lung endothelial cells 48 hr postinfluenza virus
infection.
(C) LysM-GFP mice were infected with 1 3 104 PFU of influenza virus. At 2 days later, bone marrow cells were harvested from infected LysM-GFP mice and
transferred into C57BL/6J mice that were either left uninfected or infected with 1 3 104 PFU of influenza virus 2 hr prior. Mice receiving bone marrow cells from LysMGFP-positive mice were treated with either vehicle or CYM-5442 (2 mg/kg 1, 13, 25, and 37 hr postinfection), and at 48 hr postinfection, total lung cells were harvested
from lung homogenates and the total numbers of GFP-expressing granulocytes (CD11b+, Ly6G+) and macrophages/monocytes (CD11b+F480+) were quantified.
(D) Total number of LysM-GFP-positive granulocytes and macrophages/monocytes per lung of uninfected or infected mice treated with either vehicle or
CYM-5442.
(E) Exogenous intratracheal administration of chemokine restores recruitment of innate immune cells. However, it does not resurrect cytokine/chemokine
production after CYM-5442 treatment. Mice were infected with 1 3 104 PFU WSN influenza virus, and either vehicle (water) or CYM-5442 (2 mg/kg 1, 13, 25, and
37 hr postinfection) were administered to mice i.t. Following administration of CYM-5442, recombinant mouse CCL2 was administered (50 mg) i.t. directly into the
lungs. The bar graph shows the total numbers of macrophages/monocytes, neutrophils, and NK cells in the lung 48 hr postinfluenza virus infection.
(F) Proinflammatory cytokines and chemokines were measured in BALF 48 hr postinfection by ELISA in mice treated with vehicle, CYM-5442, or CYM-5442 +
rmCCL2, as in (E).
Data represent average SEM from five mice/group. *p < 0.05; **p < 0.005; ***p < 0.0005. Results are representative of two independent experiments. See also
Figure S1 and Figure S6.
986 Cell 146, 980991, September 16, 2011 2011 Elsevier Inc.
Figure 7. Proinflammatory Cytokine Responses Are Independent of Innate Immune Cell Recruitment and Dependent on Type I Interferon
Signaling
(A) Total numbers of innate immune cells were quantified from lung digests by flow cytometry at 48 hr postinfluenza virus infection in mice treated 1 hr postinfection with vehicle or CYM-5442 in the presence of either anti-CD11b or isotype control antibody (M7/80) (0.5 mg/mouse) 0 and 24 hr postinfection.
(B) Proinflammatory cytokines and chemokines were measured 48 hr postinfection by ELISA in BALF in mice treated as in (A).
(C) C57BL/6J or IFN ab receptor-deficient mice were infected with 1 3 104 PFU of influenza virus and treated with either vehicle or CYM-5442 (2 mg/kg 1, 13, 25,
and 36 hr postinfection. Total numbers of innate immune cells were quantified from lung digests by flow cytometry at 48 hr postinfluenza virus infection, and
(D) proinflammatory cytokines and chemokines were measured 48 hr postinfection by ELISA in BALF fluid.
Data represent average SEM from five mice per group. *p < 0.05; **p < 0.005; ***p < 0.0005. Results are representative of two independent experiments. See
also Figure S1 and Figure S7.
Cell 146, 980991, September 16, 2011 2011 Elsevier Inc. 989
endothelial cells were FACS purified to > 90%95% purity, and mRNA was
purified using the RNeasy mini kit from QIAGEN. Prior to real-time PCR,
genomic DNA was digested and cDNA was made using the RT2 First Strand
Kit (C-03) according to manufacturers instructions (SA Biosciences, Frederick, MD). cDNA (200 ng) was added to individual wells and quantified using
the RT2 profiler PCR Array according to the manufacturers instructions (SA
Biosciences, Frederick, MD).
Cellular Analysis by Flow Cytometry
Lungs were harvested from PBS-perfused mice and mechanically diced into
small tissue pieces using surgical scissors. Diced lungs were suspended in
4 ml of CDTI buffer (0.5 mg/ml collagenase from Clostridium histolyticum
type IV [Sigma], 0.1 mg/ml Dnase I from bovine pancreas grade II [Roche],
1 mg/ml trypsin inhibitor type Ii-s [Sigma] in DMEM) for 1 hr at 37 C. Lung
was then disrupted mechanically through a 100 mm filter, and red blood cells
were lysed using red blood cell lysis buffer (0.02 Tris-HCL and 0.14 NH4Cl).
Inflammatory cells were purified by centrifugation in 35% PBS-buffered Percoll
(GE Healthcare Life Sciences) at 1,500 rpm for 15 min. Cell pellets were resuspended in staining buffer, and Fc receptors were blocked using 25 mg/ml antimouse CD16/32 (BD Biosciences). Cells were stained with the following
anti-mouse antibodies: AlexaFluor 488-conjugated gp38 (eBioscience; clone
eBio8.1.1), PE-conjugated (BioLegend, Inc.; clone ME13.3) and APCconjugated (eBioscience; clone 390) CD31, PE-Cy7-conjugated EpCAM
(BioLegend, Inc.; clone 68.8), Pacific blue-conjugated CD45.2 (BioLegend,
Inc.; clone 104), PerCP-Cy5.5-conjugated NK1.1 (BD Biosciences; clone
PK136), PE-Cy7-conjugated CD3e (eBioscience; clone 145-2C11), e450conjugated CD4 (eBioscience; clone L3T4), PE-conjugated CD8a (BD Biosciences; clone 53-6.1), Pacific blue-conjugated B220 (BD Biosciences; clone
RA3-6B2), PE-conjugated CD19 (BD Biosciences, clone 1D3), PE-Cy7conjugated CD11b (eBiosciences; clone M1/70), PerCP-Cy5.5-conjugated
CD11c (eBiosciences; clone N418), APC-conjugated Gr-1 (BD Biosciences;
clone RB6-8C5), Pacific blue and PE-conjugated Ly6G (BD Biosciences; clone
IA8), APC-conjugated F480 (eBioscience; clone BM8), PE-conjugated 7/4
(AbD Serotec; clone 7/4), PE-conjugated I/A-I/E (BD Biosciences; clone M5/
114.15.2), PE-Cy7 conjugated CD205 (eBiosciences; clone 205yekta), Fitcconjugated CD69 (BD Biosciences; clone H1.2F3), and APC-conjugated
CD25 (eBiosciences; clone PC61.5). Flow cytometry acquisition was performed with BD FACSDiva-driven BD LSR II flow cytometer (Becton, Dickinson
and Company). Data were then analyzed with FlowJo software (Treestar Inc.).
Western Blot
FACS-purified lung cell populations (1 3 105 cells) using the antibodies
described above were homogenized in RIPA buffer supplemented with
protease inhibitors (Pierce). Lysates were centrifuged at 50,000 3 g for
30 min, and the protein concentration in the supernatant was determined by
BCA assay (Pierce). Equal amounts of protein from cell lystates were loaded
in nondenaturing conditions and separated by SDS-PAGE in 4%12%
NuPAGE (Novex) Bis-Tris gels. Gels were transferred to PVDF membranes
followed by probing for GFP using an anti-GFP antibody (Abcam) and an
anti-rabbit Ig light-chain HRP secondary (ELC Biosciences) using chemiluminescence autoradiography.
SUPPLEMENTAL INFORMATION
Supplemental Information includes seven figures and can be found with this
article online at doi:10.1016/j.cell.2011.08.015.
ACKNOWLEDGMENTS
This is Publication Number 21112 from the Department of Immunology and
Microbial Science and the Department of Chemical Physiology and The
Scripps Research Institute Molecular Screening Center, The Scripps Research
Institute (TSRI). This work was supported, in part, by USPHS grants AI074564
(M.B.A.O., H.R., K.B.W., and J.R.T.), AI009484 (M.B.A.O.), AI05509 (H.R.),
MH084512 (H.R.), and NIH training grants NS041219 (K.B.W.), AI007244
(K.B.W.), and AI007364 (J.R.T). We thank Marcus Boehm, Li-ming Huang,
and Bryan Clemons (Receptos, Inc.) for helping provide RP-002 as a chemical
990 Cell 146, 980991, September 16, 2011 2011 Elsevier Inc.
Kawane, K., Tanaka, H., Kitahara, Y., Shimaoka, S., and Nagata, S. (2010).
Cytokine-dependent but acquired immunity-independent arthritis caused by
DNA escaped from degradation. Proc. Natl. Acad. Sci. USA 107, 19432
19437.
Kobasa, D., Jones, S.M., Shinya, K., Kash, J.C., Copps, J., Ebihara, H., Hatta,
Y., Kim, J.H., Halfmann, P., Hatta, M., et al. (2007). Aberrant innate immune
response in lethal infection of macaques with the 1918 influenza virus. Nature
445, 319323.
Kobasa, D., Takada, A., Shinya, K., Hatta, M., Halfmann, P., Theriault, S.,
Suzuki, H., Nishimura, H., Mitamura, K., Sugaya, N., et al. (2004). Enhanced
virulence of influenza A viruses with the haemagglutinin of the 1918 pandemic
virus. Nature 431, 703707.
La Gruta, N.L., Kedzierska, K., Stambas, J., and Doherty, P.C. (2007).
A question of self-preservation: immunopathology in influenza virus infection.
Immunol. Cell Biol. 85, 8592.
Link, H. (1998). The cytokine storm in multiple sclerosis. Mult. Scler. 4, 1215.
Marsolais, D., Hahm, B., Edelmann, K.H., Walsh, K.B., Guerrero, M., Hatta, Y.,
Kawaoka, Y., Roberts, E., Oldstone, M.B., and Rosen, H. (2008). Local not
systemic modulation of dendritic cell S1P receptors in lung blunts virusspecific immune responses to influenza. Mol. Pharmacol. 74, 896903.
Marsolais, D., Hahm, B., Walsh, K.B., Edelmann, K.H., McGavern, D., Hatta,
Y., Kawaoka, Y., Rosen, H., and Oldstone, M.B. (2009). A critical role for the
sphingosine analog AAL-R in dampening the cytokine response during
influenza virus infection. Proc. Natl. Acad. Sci. USA 106, 15601565.
Rosen, H., and Gordon, S. (1987). Monoclonal antibody to the murine type 3
complement receptor inhibits adhesion of myelomonocytic cells in vitro and
inflammatory cell recruitment in vivo. J. Exp. Med. 166, 16851701.
Rosen, H., and Liao, J. (2003). Sphingosine 1-phosphate pathway therapeutics: a lipid ligand-receptor paradigm. Curr. Opin. Chem. Biol. 7, 461468.
Rosen, H., Sanna, M.G., Cahalan, S.M., and Gonzalez-Cabrera, P.J. (2007).
Tipping the gatekeeper: S1P regulation of endothelial barrier function. Trends
Immunol. 28, 102107.
Rosen, H., Gonzalez-Cabrera, P., Marsolais, D., Cahalan, S., Don, A.S., and
Sanna, M.G. (2008). Modulating tone: the overture of S1P receptor immunotherapeutics. Immunol. Rev. 223, 221235.
Rosen, H., Gonzalez-Cabrera, P.J., Sanna, M.G., and Brown, S. (2009). Sphingosine 1-phosphate receptor signaling. Annu. Rev. Biochem. 78, 743768.
Sanchez, T., Estrada-Hernandez, T., Paik, J.H., Wu, M.T., Venkataraman, K.,
Brinkmann, V., Claffey, K., and Hla, T. (2003). Phosphorylation and action of
the immunomodulator FTY720 inhibits vascular endothelial cell growth
factor-induced vascular permeability. J. Biol. Chem. 278, 4728147290.
Sanna, M.G., Wang, S.K., Gonzalez-Cabrera, P.J., Don, A., Marsolais, D.,
Matheu, M.P., Wei, S.H., Parker, I., Jo, E., Cheng, W.C., et al. (2006). Enhancement of capillary leakage and restoration of lymphocyte egress by a chiral
S1P1 antagonist in vivo. Nat. Chem. Biol. 2, 434441.
Martinborough, E., Boehm, M., Yeager, A., Yamiyo, J., Huang, L., Brahmachary, E., Moorjani, M., Timony, G., Brooks, J., Peach, R., et al. (2011). Selective
sphingosine 1-phosphate receptor modulators and methods of chiral
synthesis, World International Patent Organization: WO/2011/060392 p. 92.
Stacey, A.R., Norris, P.J., Qin, L., Haygreen, E.A., Taylor, E., Heitman, J.,
Lebedeva, M., DeCamp, A., Li, D., Grove, D., et al. (2009). Induction of a striking
systemic cytokine cascade prior to peak viremia in acute human immunodeficiency virus type 1 infection, in contrast to more modest and delayed
responses in acute hepatitis B and C virus infections. J. Virol. 83, 37193733.
Thiel, V., and Weber, F. (2008). Interferon and cytokine responses to SARScoronavirus infection. Cytokine Growth Factor Rev. 19, 121132.
Rivera, J., Proia, R.L., and Olivera, A. (2008). The alliance of sphingosine1-phosphate and its receptors in immunity. Nat. Rev. Immunol. 8, 753763.
Cell 146, 980991, September 16, 2011 2011 Elsevier Inc. 991
SUMMARY
Synaptic plasticity in response to changes in physiologic state is coordinated by hormonal signals across
multiple neuronal cell types. Here, we combine celltype-specific electrophysiological, pharmacological,
and optogenetic techniques to dissect neural circuits
and molecular pathways controlling synaptic plasticity onto AGRP neurons, a population that regulates
feeding. We find that food deprivation elevates
excitatory synaptic input, which is mediated by a
presynaptic positive feedback loop involving AMPactivated protein kinase. Potentiation of glutamate
release was triggered by the orexigenic hormone
ghrelin and exhibited hysteresis, persisting for hours
after ghrelin removal. Persistent activity was reversed by the anorexigenic hormone leptin, and optogenetic photostimulation demonstrated involvement of opioid release from POMC neurons. Based
on these experiments, we propose a memory storage
device for physiological state constructed from
bistable synapses that are flipped between two
sustained activity states by transient exposure to
hormones signaling energy levels.
INTRODUCTION
Neurons that express Agouti-related protein (Agrp) are a molecularly defined population localized in the hypothalamic arcuate
nucleus, and increasing their electrical activity is sufficient to
rapidly induce voracious feeding behavior (Aponte et al.,
2011). They are intermingled with a separate and functionally
opposed population that is delineated by expression of Proopiomelanocortin (Pomc), which inhibits feeding (Aponte et al.,
2011). These neurons form a core circuit regulating food intake
and energy expenditure in which AGRP neurons inhibit POMC
neurons (Cowley et al., 2001). Both populations behave as interoceptive sensory neurons, modulating their electrical activity
in response to hormonal signals of metabolic state. In addition,
AGRP and POMC neurons receive synaptic inputs that are also
sensitive to the adipocyte-derived anorexigenic hormone leptin
992 Cell 146, 9921003, September 16, 2011 2011 Elsevier Inc.
(Hahn et al., 1998) located in the arcuate nucleus of NpysapphireFP transgenic mice (Pinto et al., 2004). Because two
measures of postsynaptic plasticity, AMPA-R/NMDA-R synaptic current ratio (Figure S1A available online) and the rectification of glutamatergic synaptic currents (Figure S1B), did not
show significant differences between fed and food-deprived
mice, we focused on presynaptic plasticity.
We examined presynaptic function by measuring the frequency of miniature excitatory postsynaptic currents (fmEPSC).
mEPSCs reflect spontaneous neurotransmitter release from
the presynaptic terminal (Fatt and Katz, 1952) under pharmacologic conditions that block action potentials and, by extension,
network activity (tetrodotoxin, TTX, 1 mM). For AGRP neurons,
we found that fmEPSC was approximately doubled after deprivation or when measured at the beginning of the dark period (DP)
in fed mice (fed: 1.4 0.1 s 1, n = 62; deprived: 3.0 0.2 s 1,
n = 62; fed(DP): 3.1 0.3 s 1, n = 26; F2,147 = 30.5, p < 0.001;
Figures 1A and 1B), whereas mEPSC amplitudes were not significantly different (F2,145 = 2.2, p = 0.11; Figure S1C). Thus, in
AGRP neurons, fmEPSC increased in association with the animals
tendency to consume food, either after food deprivation or at the
start of the dark period. Furthermore, deprivation-induced
synaptic upregulation was cell type selective as POMC neurons,
which are intermingled with AGRP neurons (Figure 1C) but
are functionally opposed, experienced marked reduction in
fmEPSC with deprivation (fed: 3.3 0.3 s 1, n = 30; deprived:
1.8 0.2 s 1, n = 22; unpaired t test, p < 0.001; Figure 1D),
consistent with previous experiments measuring synaptic inputs
to these neurons (Sternson et al., 2005).
Activation of AGRP neurons increases feeding behavior
(Aponte et al., 2011), so we considered the impact of elevated
spontaneous synaptic input on AGRP neuron firing rate.
Because neuron firing is measured in the absence of TTX, we
first recorded the frequency of spontaneous excitatory synaptic
currents under this condition, which showed a significant deprivation-induced increase (fed: 2.4 0.3 s 1, n = 9; deprived: 5.4
0.7 s 1, n = 10; unpaired t test, p = 0.009). In line with elevated
excitatory synaptic input, AGRP neurons from deprived mice
also have a higher spontaneous firing rate (fAP) than neurons
from fed mice. This difference was eliminated by blocking excitatory synaptic activity with the glutamate receptor antagonist
CNQX (2 mM) (CNQX: F1,21 = 15.8, p < 0.001; fed/dep: F1,21 =
3.2, p = 0.09; interaction: F1,21 = 6.2, p = 0.02; Figure 1E). A
previous report showed that AGRP neuron firing rate was
elevated after food deprivation in the absence of all synaptic
input (Takahashi and Cone, 2005). For pharmacological conditions similar to those used in this prior study, we find comparable fAP in slices from deprived mice (fAP, CNQX/picrotoxin:
1.2 0.4 s 1, n = 10); however, our results indicate that in the
presence of synaptic inhibition, a more physiologically relevant
condition, excitatory synaptic upregulation is necessary for
increased electrical activity. Together with AGRP neuron photostimulation-evoked feeding (Aponte et al., 2011), these experiments suggest a relationship between synaptic activity, AGRP
neuron firing, and feeding behavior. Because of the potential
behavioral importance of this synaptic control point, we further
investigated synaptic regulation of these neurons.
Synaptic Upregulation Requires Ryanodine-Sensitive
Calcium Stores
To probe the mechanism of synaptic upregulation at AGRP
neurons, we first manipulated intracellular Ca2+ buffering in brain
slices with the membrane-permeable Ca2+ buffer, BAPTA-AM
(25 mM). This treatment reduced fmEPSC in slices from deprived
mice to the level observed in BAPTA-AM-treated slices from
fed mice (fed: 1.1 0.1 s 1, n = 24; deprived: 1.4 0.1 s 1,
n = 27; Figure 2A, left). These data implicate Ca2+-dependent
processes in synaptic plasticity observed after food deprivation.
One potential pathway for Ca2+ is through voltage-gated Ca2+
channels (VGCCs). Blockade of VGCCs with CdCl2 (200 mM)
reduced fmEPSC in both fed and deprived mice (fed: 0.7
0.1 s 1, n = 26; deprived: 1.8 0.3 s 1, n = 21; Figure 2A, right);
however, fmEPSC in AGRP neurons from deprived mice was still
significantly greater than that observed in fed mice. Thus,
VGCCs contribute to fmEPSC in AGRP neurons, but they are not
required for deprivation-induced synaptic upregulation. Moreover, in the presence of CdCl2, fmEPSC in AGRP neurons from
deprived mice was still reduced by BAPTA-AM to the level
Cell 146, 9921003, September 16, 2011 2011 Elsevier Inc. 993
observed in slices from fed mice (0.6 0.2 s 1, n = 11; Figure 2A,
right), which indicated that a different source of Ca2+ was
responsible for elevated fmEPSC in deprived mice.
We next perturbed Ca2+ release from internal stores by blocking ryanodine receptors (RyR) with ryanodine (10 mM). Under
these conditions, fmEPSC in AGRP neurons from deprived mice
was reduced to the level observed for fed mice (fed: 1.5
0.1 s 1, n = 26; deprived: 1.6 0.1 s 1, n = 21; Figure 2A, left),
and this was also the case in the presence of CdCl2 (deprived:
0.8 0.1 s 1, n = 10; Figure 2A, right). This indicates that internal
stores serve as a source of Ca2+ required for elevated fmEPSC in
deprived mice. In addition, caffeine (10 mM), which activates
RyR-mediated Ca2+ release, increased fmEPSC in AGRP neurons
from fed mice to the level observed after deprivation, and this
was blocked by ryanodine pretreatment (Figure 2B). It has
been shown previously that increased spontaneous neurotransmitter release is associated with Ca2+ release from internal
stores in the presynaptic terminal (Emptage et al., 2001). Our
experiments are consistent with a role for Ca2+ release from
internal stores in deprivation-induced plasticity.
994 Cell 146, 9921003, September 16, 2011 2011 Elsevier Inc.
996 Cell 146, 9921003, September 16, 2011 2011 Elsevier Inc.
electrophysiology in D-Lys3-GHRP6-containing solution (Figure 4C). Remarkably, AGRP neurons from these slices still
showed elevated fmEPSC levels that were not significantly
998 Cell 146, 9921003, September 16, 2011 2011 Elsevier Inc.
baseline opioid tone in fed mice can reverse this after ghrelin
levels fall.
POMC Neurons Release an Opioid that Resets
Persistent Synaptic Activity
Although opioid receptors are involved in the leptin-mediated
inactivation of presynaptic AMPK, it was unclear whether
POMC neurons could serve as the source of these opioids. We
tested this using an optogenetic approach. We expressed the
light-activated cation channel channelrhodopsin-2 (ChR2) (Boyden et al., 2005) in POMC neurons, rendering them selectively
photoexcitable. POMC neurons were targeted using the Cre recombinase (Cre)-dependent viral vector rAAV-FLEX-rev-ChR2tdtomato, which we have described previously for POMC neuron
photostimulation (Aponte et al., 2011; Atasoy et al., 2008).
In brain slices from food-deprived Pomc-Cre;Npy-sapphireFP
double-transgenic mice expressing ChR2, we photostimulated
the area containing the arcuate nucleus with focused laser light
in a 4 3 10 spatial pattern of stimulation sites (Figure 6A). Our
expectation was that a neuromodulator would be released that
could potentially downregulate AMPK signaling. Next, the slice
was treated with TTX to record mEPSCs and CdCl2 in order to
eliminate any contribution of VGCCs. POMC neuron photostimulation led to a significant reduction in fmEPSC in AGRP neurons,
and this effect was blocked if photostimulation was performed
in the presence of NTX (500 nM) (control: 1.3 0.2 s 1, n = 15;
NTX: 2.2 0.2 s 1, n = 14; unpaired t test, p = 0.003; Figure 6B).
When NTX was present during photostimulation, AGRP neurons
from deprived mice maintained insensitivity to AICAR (Figure 6C).
However, after photostimulation in the absence of NTX, fmEPSC
could be increased in response to subsequent AICAR administration (AICAR: F1,8 = 43.3, p < 0.001; NTX: F1,8 = 3.6, p =
0.096; interaction: F1,8 = 36.8, p < 0.001; Figure 6C). These
results indicate that a POMC neuron-derived opioid, likely
b-endorphin, reverses AMPK-mediated persistent synaptic
upregulation.
DISCUSSION
Homeostasis requires increased food intake as energy stores
are depleted. In this study, we provide evidence for a dynamic
neural circuit (Figure 7A) with a reversible memory storage
capacity that regulates feeding behavior in response to energy
1000 Cell 146, 9921003, September 16, 2011 2011 Elsevier Inc.
1002 Cell 146, 9921003, September 16, 2011 2011 Elsevier Inc.
ACKNOWLEDGMENTS
This research was funded by the Howard Hughes Medical Institute. We thank
L. Scheffer for discussion of digital electronic circuits; J. Cox for mouse
husbandry; Y. Aponte for assistance with animal studies; Y. Li and Z. Zhang
for assistance with biochemistry; and K. Moses, S. Eddy, K. Svoboda,
G. Murphy, and C. Zuker for comments on the manuscript. Leptin was
provided by Amylin Pharmaceuticals. Y.Y. performed the experiments. H.S.
performed western blots. D.A. made the initial finding of deprivation-induced
plasticity. Y.Y. and S.M.S. designed the study, analyzed the data, and wrote
the paper.
Received: November 10, 2010
Revised: June 15, 2011
Accepted: July 15, 2011
Published: September 15, 2011
REFERENCES
Anderson, K.A., Ribar, T.J., Lin, F., Noeldner, P.K., Green, M.F., Muehlbauer,
M.J., Witters, L.A., Kemp, B.E., and Means, A.R. (2008). Hypothalamic
CaMKK2 contributes to the regulation of energy balance. Cell Metab. 7,
377388.
Andrews, Z.B., Liu, Z.W., Walllingford, N., Erion, D.M., Borok, E., Friedman,
J.M., Tschop, M.H., Shanabrough, M., Cline, G., Shulman, G.I., et al. (2008).
UCP2 mediates ghrelins action on NPY/AgRP neurons by lowering free
radicals. Nature 454, 846851.
Aponte, Y., Atasoy, D., and Sternson, S.M. (2011). AGRP neurons are sufficient
to orchestrate feeding behavior rapidly and without training. Nat. Neurosci. 14,
351355.
Appleyard, S.M., Hayward, M., Young, J.I., Butler, A.A., Cone, R.D.,
Rubinstein, M., and Low, M.J. (2003). A role for the endogenous opioid
beta-endorphin in energy homeostasis. Endocrinology 144, 17531760.
Atasoy, D., Aponte, Y., Su, H.H., and Sternson, S.M. (2008). A FLEX switch
targets Channelrhodopsin-2 to multiple cell types for imaging and long-range
circuit mapping. J. Neurosci. 28, 70257030.
Bhalla, U.S., and Iyengar, R. (1999). Emergent properties of networks of biological signaling pathways. Science 283, 381387.
Boelen, A., Kwakkel, J., Vos, X.G., Wiersinga, W.M., and Fliers, E. (2006).
Differential effects of leptin and refeeding on the fasting-induced decrease
of pituitary type 2 deiodinase and thyroid hormone receptor beta2 mRNA
expression in mice. J. Endocrinol. 190, 537544.
Boyden, E.S., Zhang, F., Bamberg, E., Nagel, G., and Deisseroth, K. (2005).
Millisecond-timescale, genetically targeted optical control of neural activity.
Nat. Neurosci. 8, 12631268.
Cannon, W.B. (1929). Organization for physiological homeostasis. Physiol.
Rev. 9, 399431.
Corton, J.M., Gillespie, J.G., Hawley, S.A., and Hardie, D.G. (1995). 5-aminoimidazole-4-carboxamide ribonucleoside. A specific method for activating
AMP-activated protein kinase in intact cells? Eur. J. Biochem. 229, 558565.
Cowley, M.A., Smart, J.L., Rubinstein, M., Cerdan, M.G., Diano, S., Horvath,
T.L., Cone, R.D., and Low, M.J. (2001). Leptin activates anorexigenic POMC
neurons through a neural network in the arcuate nucleus. Nature 411, 480484.
Cowley, M.A., Smith, R.G., Diano, S., Tschop, M., Pronchuk, N., Grove, K.L.,
Strasburger, C.J., Bidlingmaier, M., Esterman, M., Heiman, M.L., et al.
(2003). The distribution and mechanism of action of ghrelin in the CNS
demonstrates a novel hypothalamic circuit regulating energy homeostasis.
Neuron 37, 649661.
DeFalco, J., Tomishima, M., Liu, H., Zhao, C., Cai, X., Marth, J.D., Enquist, L.,
and Friedman, J.M. (2001). Virus-assisted mapping of neural inputs to
a feeding center in the hypothalamus. Science 291, 26082613.
Dobrunz, L.E., and Stevens, C.F. (1997). Heterogeneity of release probability,
facilitation, and depletion at central synapses. Neuron 18, 9951008.
Emptage, N.J., Reid, C.A., and Fine, A. (2001). Calcium stores in hippocampal
synaptic boutons mediate short-term plasticity, store-operated Ca2+ entry,
and spontaneous transmitter release. Neuron 29, 197208.
Endo, K., and Yawo, H. (2000). mu-Opioid receptor inhibits N-type Ca2+
channels in the calyx presynaptic terminal of the embryonic chick ciliary
ganglion. J. Physiol. 524, 769781.
Evans, A.M., Mustard, K.J., Wyatt, C.N., Peers, C., Dipp, M., Kumar, P.,
Kinnear, N.P., and Hardie, D.G. (2005). Does AMP-activated protein kinase
couple inhibition of mitochondrial oxidative phosphorylation by hypoxia to
calcium signaling in O2-sensing cells? J. Biol. Chem. 280, 4150441511.
Fatt, P., and Katz, B. (1952). Spontaneous subthreshold activity at motor nerve
endings. J. Physiol. 117, 109128.
Ferrell, J.E., Jr. (2002). Self-perpetuating states in signal transduction: positive
feedback, double-negative feedback and bistability. Curr. Opin. Cell Biol. 14,
140148.
Gordon, G.R., and Bains, J.S. (2006). Can homeostatic circuits learn and
remember? J. Physiol. 576, 341347.
Grandison, L., and Guidotti, A. (1977). Stimulation of food intake by muscimol
and beta endorphin. Neuropharmacology 16, 533536.
Hahn, T.M., Breininger, J.F., Baskin, D.G., and Schwartz, M.W. (1998).
Coexpression of Agrp and NPY in fasting-activated hypothalamic neurons.
Nat. Neurosci. 1, 271272.
Hawley, S.A., Pan, D.A., Mustard, K.J., Ross, L., Bain, J., Edelman, A.M.,
Frenguelli, B.G., and Hardie, D.G. (2005). Calmodulin-dependent protein
kinase kinase-beta is an alternative upstream kinase for AMP-activated
protein kinase. Cell Metab. 2, 919.
Holm, S.A. (1979). A simple sequentially rejective multiple test procedure.
Scand. J. Stat. 6, 6570.
Holst, B., Cygankiewicz, A., Jensen, T.H., Ankersen, M., and Schwartz, T.W.
(2003). High constitutive signaling of the ghrelin receptoridentification of
a potent inverse agonist. Mol. Endocrinol. 17, 22012210.
Horowitz, P., and Hill, W. (1989). The Art of Electronics, Second Edition
(Cambridge, UK: Cambridge University Press).
Howard, A.D., Feighner, S.D., Cully, D.F., Arena, J.P., Liberator, P.A.,
Rosenblum, C.I., Hamelin, M., Hreniuk, D.L., Palyha, O.C., Anderson, J.,
et al. (1996). A receptor in pituitary and hypothalamus that functions in growth
hormone release. Science 273, 974977.
Hurley, R.L., Anderson, K.A., Franzone, J.M., Kemp, B.E., Means, A.R., and
Witters, L.A. (2005). The Ca2+/calmodulin-dependent protein kinase kinases
are AMP-activated protein kinase kinases. J. Biol. Chem. 280, 2906029066.
Lopez, M., Lage, R., Saha, A.K., Perez-Tilve, D., Vazquez, M.J., Varela, L.,
Sangiao-Alvarellos, S., Tovar, S., Raghay, K., Rodrguez-Cuenca, S., et al.
(2008). Hypothalamic fatty acid metabolism mediates the orexigenic action
of ghrelin. Cell Metab. 7, 389399.
Luquet, S., Phillips, C.T., and Palmiter, R.D. (2007). NPY/AgRP neurons are not
essential for feeding responses to glucoprivation. Peptides 28, 214225.
Martianez, T., France`s, S., and Lopez, J.M. (2009). Generation of digital
responses in stress sensors. J. Biol. Chem. 284, 2390223911.
Minokoshi, Y., Alquier, T., Furukawa, N., Kim, Y.B., Lee, A., Xue, B., Mu, J.,
Foufelle, F., Ferre, P., Birnbaum, M.J., et al. (2004). AMP-kinase regulates
food intake by responding to hormonal and nutrient signals in the hypothalamus. Nature 428, 569574.
Nakazato, M., Murakami, N., Date, Y., Kojima, M., Matsuo, H., Kangawa, K.,
and Matsukura, S. (2001). A role for ghrelin in the central regulation of feeding.
Nature 409, 194198.
Pinto, S., Roseberry, A.G., Liu, H., Diano, S., Shanabrough, M., Cai, X.,
Friedman, J.M., and Horvath, T.L. (2004). Rapid rewiring of arcuate nucleus
feeding circuits by leptin. Science 304, 110115.
Silva, R.M., Hadjimarkou, M.M., Rossi, G.C., Pasternak, G.W., and Bodnar,
R.J. (2001). Beta-endorphin-induced feeding: pharmacological characterization using selective opioid antagonists and antisense probes in rats.
J. Pharmacol. Exp. Ther. 297, 590596.
Smith, R.G. (2005). Development of growth hormone secretagogues. Endocr.
Rev. 26, 346360.
Sternson, S.M., Shepherd, G.M., and Friedman, J.M. (2005). Topographic
mapping of VMH > arcuate nucleus microcircuits and their reorganization
by fasting. Nat. Neurosci. 8, 13561363.
Takahashi, K.A., and Cone, R.D. (2005). Fasting induces a large, leptin-dependent increase in the intrinsic action potential frequency of orexigenic arcuate
nucleus neuropeptide Y/Agouti-related protein neurons. Endocrinology 146,
10431047.
Tanaka, K., and Augustine, G.J. (2008). A positive feedback signal transduction loop determines timing of cerebellar long-term depression. Neuron 59,
608620.
Tschop, M., Smiley, D.L., and Heiman, M.L. (2000). Ghrelin induces adiposity
in rodents. Nature 407, 908913.
Koshland, D.E., Jr., Goldbeter, A., and Stock, J.B. (1982). Amplification and
adaptation in regulatory and sensory systems. Science 217, 220225.
Woods, A., Dickerson, K., Heath, R., Hong, S.P., Momcilovic, M., Johnstone,
S.R., Carlson, M., and Carling, D. (2005). Ca2+/calmodulin-dependent protein
kinase kinase-beta acts upstream of AMP-activated protein kinase in mammalian cells. Cell Metab. 2, 2133.
Li, C., Chen, P., and Smith, M.S. (1999). Identification of neuronal input to the
arcuate nucleus (ARH) activated during lactation: implications in the activation
of neuropeptide Y neurons. Brain Res. 824, 267276.
Zigman, J.M., Jones, J.E., Lee, C.E., Saper, C.B., and Elmquist, J.K. (2006).
Expression of ghrelin receptor mRNA in the rat and the mouse brain.
J. Comp. Neurol. 494, 528548.
Cell 146, 9921003, September 16, 2011 2011 Elsevier Inc. 1003
SUMMARY
Talbot and Marshall, 1941; Woolsey and Walzl, 1942) has led
to the view that early sensory processing is mediated by developmentally programmed neural circuits. In contrast, olfactory
features cannot be meaningfully represented along continuous
dimensions in the physical world and are not topographically
organized in the olfactory sensory epithelium (Ressler et al.,
1993; Vassar et al., 1993). Olfactory information from the nose
is transmitted to the olfactory bulb and then to higher centers.
The vast majority of odors drive behavior only after learning,
but the brain regions responsible for these learned behaviors
remain elusive. The piriform cortex receives extensive input
from the bulb (Price and Powell, 1970) and projects to areas
implicated in behavioral output (Schwabe et al., 2004), posing
the question as to whether the piriform is a substrate for olfactory
learning. We have developed an experimental strategy that
permits us to ask whether exogenous activation of the same subpopulation of piriform neurons can be entrained to elicit distinct
behavioral responses depending on the learning paradigm.
Olfactory perception is initiated by the recognition of odorants
by a large repertoire of receptors in the sensory epithelium. Individual sensory neurons in mice express only one of 1500 different receptor genes (Buck and Axel, 1991; Godfrey et al.,
2004; Zhang and Firestein, 2002). An odorant can interact with
multiple distinct receptors resulting in the activation of an ensemble of sensory neurons (Araneda et al., 2004; Malnic et al.,
1999; Oka et al., 2006). Discrimination among odorants then
requires that the brain determine which of the sensory neurons
have been activated by a given odorant. Neurons expressing
a given receptor are distributed within zones of the epithelium
but project with precision to two spatially invariant glomeruli in
the olfactory bulb (Mombaerts et al., 1996; Ressler et al., 1993,
1994; Vassar et al., 1994, 1993). Thus a transformation in the
representation of olfactory information is apparent in the bulb
where the dispersed population of active neurons in the sense
organ is consolidated into a discrete spatial map of glomerular
activity.
If an odorant activates a unique ensemble of glomeruli, then
the recognition of an odor requires integration of information
from multiple glomeruli by higher olfactory centers. The projection neurons of the olfactory bulb, mitral and tufted cells, extend
an apical dendrite into a single glomerulus and send axons to
several telencephalic areas, including a significant input to the
results in the activation of ChR2-expressing neurons with millisecond precision in awake, behaving animals (Aravanis et al.,
2007; Boyden et al., 2005). In initial experiments, channelrhodopsin expression was driven by the human Synapsin1 promoter
(Kugler et al., 2003) in both excitatory and inhibitory cells in the
piriform cortex (Figure 1A). In a second set of experiments,
a ChR2 gene in the reverse orientation was flanked by pairs of
loxP sites, such that expression of ChR2 was dependent upon
Cre recombinase (Atasoy et al., 2008). Injection of this virus
into the piriform cortex of Emx1-IRES-Cre mice in which Cre recombinase is restricted to excitatory neurons (Gorski et al.,
2002), resulted in ChR2 expression in pyramidal neurons but
not inhibitory interneurons (Figure 1B). These two strategies for
ChR2 delivery resulted in abundant expression of ChR2 in
over 50% of the neurons in an injection site that ranged from
500 mm-1000 mm in diameter (Figures 1A and 1B).
We have achieved sparser distribution of active neurons
by coinjection of high-titer Cre-dependent ChR2 virus with
a range of dilutions of lentivirus encoding Cre. Under these
conditions (dual virus strategy), only a small subset of the cells
were coinfected with both viruses and high-level ChR2 expression was therefore sparse, with about 10% (8.25 0.33%,
n = 3) of neurons expressing ChR2 at the injection site (Figure 1C). Analysis of c-Fos expression, a marker of neuronal activation (Morgan and Curran, 1991), revealed that all three genetic
approaches to effect channelrhodopsin expression resulted in
robust neural activation upon exposure to light (Figures 1A, 1B,
and 1D). We observed a threefold (3.06 0.46, n = 4) increase
in the number of c-Fos+ cells at injection sites using the first
two expression strategies that produced dense populations of
ChR2-expressing neurons. In the third expression strategy that
generated sparse populations of ChR2-expressing neurons,
the coexpression of nuclear Cherry (Figure 1C) allowed us to
identify the incidence of c-Fos expression among ChR2+ cells.
We observed that about 40% of cells expressing ChR2 also expressed c-Fos (37.76 11.7%, n = 6), whereas 5% of cells
were positive for c-Fos expression in uninjected control hemispheres (6.09 0.36%, n = 3).
1006 Cell 146, 10041015, September 16, 2011 2011 Elsevier Inc.
ure 3C and Table S1 available online). The observation that aversive behavior can be entrained at multiple loci distributed across
the piriform indicates that valence of behavioral output is not
spatially segregated in the piriform cortex.
We have compared the efficiency with which odor and photostimulation served as a conditioned stimulus to elicit flight
behavior. Pairing of odor exposure with foot shock resulted in
a consistent aversive response to odor alone after 10 CS-US
pairings (10.25 0.96 pairings, n = 4). Approximately twice as
many pairings were required to elicit flight behavior in mice expressing ChR2 in a subpopulation of piriform neurons (18
4.02 pairings, n = 11) (Figure 3D). Thus, the activation of an
ensemble of 500 piriform neurons approaches the efficacy of
odor activation of 100,000 neurons (Stettler and Axel, 2009), in
eliciting conditioned aversion.
An Ensemble of Neurons Trained to Elicit Appetitive
Behavior
We next asked whether the photostimulation of an ensemble of
piriform neurons expressing ChR2 could elicit appetitive behavioral responses if paired with a rewarding US. We modified
Cell 146, 10041015, September 16, 2011 2011 Elsevier Inc. 1007
Piriform
Neurons
asked whether the same population of ChR2-expressing neurons in piriform could be entrained to sequentially elicit behaviors
of different valence (Figure 6A). Mice expressing ChR2 in piriform
were first trained in the appetitive paradigm and consistently exhibited conditioned licking responses upon photostimulation
(Figure 4B). The animals were then conditioned in the aversive
foot shock paradigm and now displayed robust flight behavior
(Figure 6B). Sequentially trained animals acquired the aversive
behavior as quickly as naive animals (data not shown) and exhibited flight behavior in 90% of the trials (ChR2+ animals =
88.57 25.55%, n = 5; control animals = 3.33 8.16%, n = 6).
We next asked whether the response to photostimulation in
these sequentially trained animals was context-dependent.
Sequentially trained animals were returned to the appetitive
conditioning context. Upon photostimulation, they no longer exhibited appetitive responses in anticipation of water reward.
Rather, the level of licking to the CS+ approached that of the
CS- (Figure 6C). Moreover, photostimulation occasionally
1010 Cell 146, 10041015, September 16, 2011 2011 Elsevier Inc.
Figure 6. The Same Ensemble of ChR2-Expressing Neurons Can Be Entrained to Elicit Appetitive and Aversive Behaviors
(A) A schematic of the sequential training of ChR2-expressing animals to produce appetitive and aversive behaviors.
(B) A subset of mice shown in Figure 4B, which were trained in an appetitive water reward behavior, was subsequently trained in the aversive foot shock paradigm.
The percentage of trials in which animals exhibited flight behavior in response to photostimulation alone during the testing phase is plotted for ChR2-expressing
(Emx1: n = 1, dual virus with > 300 ChR2+ neurons: n = 4) and control animals (n = 6).
(C) The average lick number over the last three training blocks of sequentially trained animals (from Figure 4B and Figure 6B) before and after aversive conditioning
with the same ensemble. ChR2 before aversive: number of licks following CS+ and CS- for ChR2-expressing mice during initial appetitive conditioning (CS+ =
58.68 9.76 licks and CS- = 14.43 9.99 licks, n = 5). ChR2 After Aversive: number of licks following CS+ and CS- for these mice after sequential appetitiveaversive conditioning (CS+ = 12.81 15.96 licks and CS- = 11.39 8.21 licks, n = 5). Control After Aversive: number of licks following CS+ and CS- for control
animals after sequential appetitive-aversive conditioning (CS+ = 35.58 13.05 licks and CS- = 34.29 4.94 licks, n = 5).
network is then subject to selection by experience over the life of an organism. Selection reinforces connections from neurons that represent
sensory objects of behavioral significance. The
ability to entrain populations of piriform neurons
to elicit specific behavioral responses is in accord with these
models.
A similar conceptual organization in which the random convergence of entorhinal inputs creates a distributed ensemble may
also be operative for hippocampal place cells. Place cells in
the hippocampus, a three-layered cortical structure like piriform,
exhibit no apparent relationship between their positions in brain
space and their firing field in the external world (OKeefe et al.,
1998; Redish et al., 2001). As a consequence, the spatial map
in the hippocampus is likely to differ in different individuals
residing in the same environment. Moreover, place cells remap:
they alter their firing properties in response to changes in spatial
environment (Colgin et al., 2008). This suggests models in which
inputs to individual place cells are randomly chosen during
development such that a given location is represented by a
distributed ensemble of active neurons.
The exogenous activation of neurons in other sensory cortices, either by microstimulation (Doty, 1969; Murphey and
Maunsell, 2007; Yang et al., 2008) or photostimulation (Huber
et al., 2008), has been shown to elicit behaviors following
training. Microstimulation of loci within visual, auditory, and
somatosensory cortex suggests that neuronal activation at
many neocortical levels of sensory processing is capable of influencing perceptual tasks. In more recent experiments, photostimulation of sparse ensembles of ChR2-expressing neurons in
somatosensory cortex was conditioned to drive appetitive
behavior (Huber et al., 2008). These sensory neocortices maintain topographic order that represents stimulus features in at
least one dimension across the cortex. As a consequence, microstimulation or photostimulation of a locus in these cortices
results in the activation of a topographically constrained subpopulation of neurons that is likely to encode specific features
of a sensory stimulus. In these experiments, focal activation
Cell 146, 10041015, September 16, 2011 2011 Elsevier Inc. 1013
for help with behavioral assays; Monica Mendelsohn, Jennifer Kirkland, and
Nataliya Zabello for help with the mice; Anmo Kim for help with laser and olfactometer machines; Phyllis Kisloff for assistance in preparation of the manuscript; David J. Anderson, Larry Abbott, Thomas Jessell, Eric R. Kandel, and
members of the Axel lab for critical reading of the manuscript and discussions;
and Miriam Gutierrez for general laboratory support. This work was supported
by the Howard Hughes Medical Institute and the Mathers Foundation. G.B.C.
was supported by the Damon Runyon Cancer Research Foundation Postdoctoral Fellowship. A.F. was supported by long-term postdoctoral fellowships
from EMBO and the Human Frontiers Science Program.
Received: March 4, 2011
Revised: May 26, 2011
Accepted: July 21, 2011
Published: September 15, 2011
REFERENCES
Abraham, N.M., Spors, H., Carleton, A., Margrie, T.W., Kuner, T., and
Schaefer, A.T. (2004). Maintaining accuracy at the expense of speed: stimulus
similarity defines odor discrimination time in mice. Neuron 44, 865876.
Araneda, R.C., Peterlin, Z., Zhang, X., Chesler, A., and Firestein, S. (2004). A
pharmacological profile of the aldehyde receptor repertoire in rat olfactory
epithelium. J. Physiol. 555, 743756.
Aravanis, A.M., Wang, L.P., Zhang, F., Meltzer, L.A., Mogri, M.Z., Schneider,
M.B., and Deisseroth, K. (2007). An optical neural interface: in vivo control of
rodent motor cortex with integrated fiberoptic and optogenetic technology.
J. Neural Eng. 4, S143S156.
Atasoy, D., Aponte, Y., Su, H.H., and Sternson, S.M. (2008). A FLEX switch
targets Channelrhodopsin-2 to multiple cell types for imaging and long-range
circuit mapping. J. Neurosci. 28, 70257030.
Bodyak, N., and Slotnick, B. (1999). Performance of mice in an automated
olfactometer: odor detection, discrimination and odor memory. Chem. Senses
24, 637645.
Boyden, E.S., Zhang, F., Bamberg, E., Nagel, G., and Deisseroth, K. (2005).
Millisecond-timescale, genetically targeted optical control of neural activity.
Nat. Neurosci. 8, 12631268.
Brunet, L.J., Gold, G.H., and Ngai, J. (1996). General anosmia caused by
a targeted disruption of the mouse olfactory cyclic nucleotide-gated cation
channel. Neuron 17, 681693.
Buck, L., and Axel, R. (1991). A novel multigene family may encode odorant
receptors: a molecular basis for odor recognition. Cell 65, 175187.
Changeux, J.P., Courrege, P., and Danchin, A. (1973). A theory of the epigenesis of neuronal networks by selective stabilization of synapses. Proc. Natl.
Acad. Sci. USA 70, 29742978.
Colgin, L.L., Moser, E.I., and Moser, M.B. (2008). Understanding memory
through hippocampal remapping. Trends Neurosci. 31, 469477.
Davison, I.G., and Ehlers, M.D. (2011). Neural circuit mechanisms for pattern
detection and feature combination in olfactory cortex. Neuron 70, 8294.
Dittgen, T., Nimmerjahn, A., Komai, S., Licznerski, P., Waters, J., Margrie,
T.W., Helmchen, F., Denk, W., Brecht, M., and Osten, P. (2004). Lentivirusbased genetic manipulations of cortical neurons and their optical and electrophysiological monitoring in vivo. Proc. Natl. Acad. Sci. USA 101, 1820618211.
Gorski, J.A., Talley, T., Qiu, M., Puelles, L., Rubenstein, J.L., and Jones, K.R.
(2002). Cortical excitatory neurons and glia, but not GABAergic neurons, are
produced in the Emx1-expressing lineage. J. Neurosci. 22, 63096314.
Hubel, D.H., and Wiesel, T.N. (1959). Receptive fields of single neurones in the
cats striate cortex. J. Physiol. 148, 574591.
Huber, D., Petreanu, L., Ghitani, N., Ranade, S., Hromadka, T., Mainen, Z., and
Svoboda, K. (2008). Sparse optical microstimulation in barrel cortex drives
learned behaviour in freely moving mice. Nature 451, 6164.
Illig, K.R., and Haberly, L.B. (2003). Odor-evoked activity is spatially distributed
in piriform cortex. J. Comp. Neurol. 457, 361373.
Kobayakawa, K., Kobayakawa, R., Matsumoto, H., Oka, Y., Imai, T., Ikawa, M.,
Okabe, M., Ikeda, T., Itohara, S., Kikusui, T., et al. (2007). Innate versus learned
odour processing in the mouse olfactory bulb. Nature 450, 503508.
Kugler, S., Kilic, E., and Bahr, M. (2003). Human synapsin 1 gene promoter
confers highly neuron-specific long-term transgene expression from an
adenoviral vector in the adult rat brain depending on the transduced area.
Gene Ther. 10, 337347.
Malnic, B., Hirono, J., Sato, T., and Buck, L.B. (1999). Combinatorial receptor
codes for odors. Cell 96, 713723.
Marin, E.C., Jefferis, G.S., Komiyama, T., Zhu, H., and Luo, L. (2002). Representation of the glomerular olfactory map in the Drosophila brain. Cell 109,
243255.
Marshall, W.H., Woolsey, C.N., and Bard, P. (1941). Observations on cortical
somatic sensory mechanisms of cat and monkey. J. Neurophysiol. 4, 124.
Miyamichi, K., Amat, F., Moussavi, F., Wang, C., Wickersham, I., Wall, N.R.,
Taniguchi, H., Tasic, B., Huang, Z.J., He, Z., et al. (2011). Cortical representations of olfactory input by trans-synaptic tracing. Nature 472, 191196.
Mombaerts, P., Wang, F., Dulac, C., Chao, S.K., Nemes, A., Mendelsohn, M.,
Edmondson, J., and Axel, R. (1996). Visualizing an olfactory sensory map. Cell
87, 675686.
Morgan, J.I., and Curran, T. (1991). Stimulus-transcription coupling in the
nervous system: involvement of the inducible proto-oncogenes fos and jun.
Annu. Rev. Neurosci. 14, 421451.
Mountcastle, V.B., Davies, P.W., and Berman, A.L. (1957). Response properties of neurons of cats somatic sensory cortex to peripheral stimuli.
J. Neurophysiol. 20, 374407.
Murphey, D.K., and Maunsell, J.H. (2007). Behavioral detection of electrical
microstimulation in different cortical visual areas. Curr. Biol. 17, 862867.
Murthy, M., Fiete, I., and Laurent, G. (2008). Testing odor response stereotypy
in the Drosophila mushroom body. Neuron 59, 10091023.
Nadler, J.J., Moy, S.S., Dold, G., Trang, D., Simmons, N., Perez, A., Young,
N.B., Barbaro, R.P., Piven, J., Magnuson, T.R., et al. (2004). Automated apparatus for quantitation of social approach behaviors in mice. Genes Brain
Behav. 3, 303314.
OKeefe, J., Burgess, N., Donnett, J.G., Jeffery, K.J., and Maguire, E.A. (1998).
Place cells, navigational accuracy, and the human hippocampus. Philos.
Trans. R. Soc. Lond. B Biol. Sci. 353, 13331340.
Oka, Y., Katada, S., Omura, M., Suwa, M., Yoshihara, Y., and Touhara, K.
(2006). Odorant receptor map in the mouse olfactory bulb: in vivo sensitivity
and specificity of receptor-defined glomeruli. Neuron 52, 857869.
Poo, C., and Isaacson, J.S. (2009). Odor representations in olfactory cortex:
sparse coding, global inhibition, and oscillations. Neuron 62, 850861.
Price, J.L., and Powell, T.P. (1970). The mitral and short axon cells of the olfactory bulb. J. Cell Sci. 7, 631651.
Redish, A.D., Battaglia, F.P., Chawla, M.K., Ekstrom, A.D., Gerrard, J.L., Lipa,
P., Rosenzweig, E.S., Worley, P.F., Guzowski, J.F., McNaughton, B.L., et al.
(2001). Independence of firing correlates of anatomically proximate hippocampal pyramidal cells. J. Neurosci. 21, RC134.
Ghosh, S., Larson, S.D., Hefzi, H., Marnoy, Z., Cutforth, T., Dokka, K., and
Baldwin, K.K. (2011). Sensory maps in the olfactory cortex defined by longrange viral tracing of single neurons. Nature 472, 217220.
Godfrey, P.A., Malnic, B., and Buck, L.B. (2004). The mouse olfactory receptor
gene family. Proc. Natl. Acad. Sci. USA 101, 21562161.
1014 Cell 146, 10041015, September 16, 2011 2011 Elsevier Inc.
Rennaker, R.L., Chen, C.F., Ruyle, A.M., Sloan, A.M., and Wilson, D.A. (2007).
Spatial and temporal distribution of odorant-evoked activity in the piriform
cortex. J. Neurosci. 27, 15341542.
Ressler, K.J., Sullivan, S.L., and Buck, L.B. (1993). A zonal organization of
odorant receptor gene expression in the olfactory epithelium. Cell 73,
597609.
Vassar, R., Chao, S.K., Sitcheran, R., Nunez, J.M., Vosshall, L.B., and Axel, R.
(1994). Topographic organization of sensory projections to the olfactory bulb.
Cell 79, 981991.
Ressler, K.J., Sullivan, S.L., and Buck, L.B. (1994). Information coding in the
olfactory system: evidence for a stereotyped and highly organized epitope
map in the olfactory bulb. Cell 79, 12451255.
Vassar, R., Ngai, J., and Axel, R. (1993). Spatial segregation of odorant
receptor expression in the mammalian olfactory epithelium. Cell 74, 309318.
Schwabe, K., Ebert, U., and Loscher, W. (2004). The central piriform cortex:
anatomical connections and anticonvulsant effect of GABA elevation in the
kindling model. Neuroscience 126, 727741.
Shepherd, G.M. (2004). The Synaptic Organization of the Brain (New York:
Oxford University Press).
Sosulski, D.L., Lissitsyna Bloom, M., Cutforth, T., Axel, R., and Datta, S.R.
(2011). Distinct representations of olfactory information in different cortical
centres. Nature 472, 213216.
Stettler, D.D., and Axel, R. (2009). Representations of odor in the piriform
cortex. Neuron 63, 854864.
Sugai, T., Miyazawa, T., Fukuda, M., Yoshimura, H., and Onoda, N. (2005).
Odor-concentration coding in the guinea-pig piriform cortex. Neuroscience
130, 769781.
Talbot, S.A., and Marshall, W.H. (1941). Physiological studies on neural mechanisms of visual localization and discrimination. Am. J. Ophthalmol. 24, 1255
1263.
Wong, A.M., Wang, J.W., and Axel, R. (2002). Spatial representation of the
glomerular map in the Drosophila protocerebrum. Cell 109, 229241.
Woolsey, C.N., and Walzl, E.M. (1942). Topical projection of nerve fibers from
local regions of the cochlea to the cerebral cortex of the cat. Bull. Johns
Hopkins Hosp. 71, 315344.
Yan, Z., Tan, J., Qin, C., Lu, Y., Ding, C., and Luo, M. (2008). Precise circuitry
links bilaterally symmetric olfactory maps. Neuron 58, 613624.
Yang, Y., DeWeese, M.R., Otazu, G.H., and Zador, A.M. (2008). Millisecondscale differences in neural activity in auditory cortex can drive decisions.
Nat. Neurosci. 11, 12621263.
Zhan, C., and Luo, M. (2011). Diverse patterns of odor representation by
neurons in the anterior piriform cortex of awake mice. J. Neurosci. 30,
1666216672.
Zhang, X., and Firestein, S. (2002). The olfactory receptor gene superfamily of
the mouse. Nat. Neurosci. 5, 124133.
Cell 146, 10041015, September 16, 2011 2011 Elsevier Inc. 1015
Resource
May Department of Cancer Research, The University of Chicago, Chicago, IL 60637, USA
Institute for Cancer Research and Department of Cellular and Molecular Medicine, University of California San Diego School of
Medicine, 9500 Gilman Drive, La Jolla, CA 92093, USA
3INSERM, U823; Universite
Joseph Fourier - Grenoble 1; Institut Albert Bonniot, Faculte de Medecine, Domaine de la Merci,
38706 La Tronche Cedex, France
4Shanghai Institute of Materia Medica, Chinese Academy of Sciences, 555 Zu Chong Zhi Road, Shanghai 201203, P.R. China
5Department of Chemical and Systems Biology, Stanford University School of Medicine, Stanford, CA 94305, USA
6These authors contributed equally to this work
*Correspondence: yingming.zhao@uchicago.edu
DOI 10.1016/j.cell.2011.08.008
2Ludwig
SUMMARY
We report the identification of 67 previously undescribed histone modifications, increasing the current
number of known histone marks by about 70%.
We further investigated one of the marks, lysine crotonylation (Kcr), confirming that it represents an
evolutionarily-conserved histone posttranslational
modification. The unique structure and genomic
localization of histone Kcr suggest that it is mechanistically and functionally different from histone
lysine acetylation (Kac). Specifically, in both human
somatic and mouse male germ cell genomes, histone
Kcr marks either active promoters or potential enhancers. In male germinal cells immediately following meiosis, Kcr is enriched on sex chromosomes
and specifically marks testis-specific genes, including a significant proportion of X-linked genes
that escape sex chromosome inactivation in haploid
cells. These results therefore dramatically extend
the repertoire of histone PTM sites and designate
Kcr as a specific mark of active sex chromosomelinked genes in postmeiotic male germ cells.
INTRODUCTION
Mounting evidence suggests that histone PTMs play a crucial
role in diverse biological processes, such as cell differentiation
and organismal development, and that aberrant modification of
histones contributes to diseases such as cancer (Berdasco
and Esteller, 2010; Fullgrabe et al., 2011). At least eleven types
of PTMs have been reported at over 60 different amino acid residues on histones, including histone methylation, acetylation,
1016 Cell 146, 10161028, September 16, 2011 2011 Elsevier Inc.
propionylation, butyrylation, formylation, phosphorylation, ubiquitylation, sumoylation, citrullination, proline isomerization, and
ADP ribosylation (Martin and Zhang, 2007; Ruthenburg et al.,
2007).
Histone PTMs are thought to contribute to the regulation of
chromatin-templated processes via two major mechanisms
(Kouzarides, 2007; Ruthenburg et al., 2007). First, histone
PTMs can directly modulate the packaging of chromatin either
by altering the net charge of histone molecules or by altering
inter-nucleosomal interactions, thereby regulating chromatin
structure and the access of DNA-binding proteins such as transcription factors. Second, histone PTMs regulate chromatin
structure and function by recruiting PTM-specific binding proteins, which recognize modified histones via specialized structural folds such as bromo-, chromo- and PHD domains (Wysocka
et al., 2005, 2006; Zeng and Zhou, 2002). Alternatively, histone
PTMs can also function by inhibiting the interaction of specific
binders with chromatin. PTM-induced changes in protein interactions between chromatin and its binding partners are in turn
translated into biological outcomes (Margueron et al., 2005).
While the majority of known histone PTMs are located within
the N-terminal tail domain of core histones, PTMs of crucial
importance for histone-DNA and histone-histone interactions
have also been found in the globular domain of core histones
(Cosgrove et al., 2004; Garcia et al., 2007c; Mersfelder and Parthun, 2006). Novel PTM sites occurring outside of the N-terminal
tails continue to be discovered, generally with the aid of sequence
and modification-specific antibodies or by unbiased mass spectrometry (MS) methods (Chu et al., 2006; Garcia et al., 2007b;
Johnson et al., 2004; Wisniewski et al., 2007). The recent
discovery of O-GlcNAc modification (Sakabe et al., 2010)
suggests that additional histone PTMs may yet be discovered.
Here, we used an integrated, mass spectrometry-based
proteomics approach, which takes advantage of in vitro propionylation, efficient peptide separation using isoelectric focusing
Histones
PTMs
Linker
histone
Trypsin
Propionylation
Trypsin
Trypsin
yp
Propionylation
Method I
In-gel digestion
In-sol digestion
Method II
Method IV
Method III
IEF fractionation
12 fractions
Validation
# of
identified
sites
# of novel
sites
SDS-PAGE
In-sol digestion
Kme
Core histones
Total
H1.2
H2A
H2B
H3
H4
31
12
31
33
23
17
10
19
10
11
67
18
130
Kme2
Kfo
4
2
Protein sequence
alignment
Kac
Rme
Yoh
Kcr
28
HPLC-MS/MS
me
fo
Nac
H1.2
fo
me
fo me
ac fo
me
oh
fo
me
fo
ac
fo
me
me
me
me
me
me
me
me
SKASGPKKKKYKKKKLNKKKKKKKKK
1
36
54
48
65 66
me
ac
73
oh
87
83
92
99
131
150
me
fo
me
me
108
170
201
186
226
me
me
ac ac
fo
ac
H2B
H3
39
118 119
125
me
me
ac
me
ac ac
88
42
oh
fo
me
fo
me me
oh
me
me
fo
fo
fo
...KKKGSKKAVTKAQKK..Y..K..K..RYNKR..KKAVTK
11 12
1516
20
ac
me
me2
me
me
ac
34
23
me
ac
me
me2
me3
ac
me
ac
46
37
57
79
83
me
fo
me2
me ac
me fo
me2 ac
85
99 108
me
fo
ac
116
120
: Known sites
X : Novel sites
me
ARTKQTARKKKKAARKK..K..R..K...KDIQLR
4
ac
14
me
ac
fo
ac
ac
23
18
27
me
me
fo
36
56
122
63 79
oh
fo
me
me2 me
me
128
me
ox
fo
fo
ac
H4 SGRGKK..K...KRHRKK..RYEETRK..RKRKYALK
5
12
16
20
Kcr: lysine
y
crotonylation
y
Mouse:
Human:
31
35
51
55
59
67
77
79
88
91
Kac: lysine
y
acetylation
y
Kcr
Kcr
Kcr
Kcr
K cr
Kcr
Kcr
Kcr
K cr
Kcr
Kcr
K cr
63
33
Mouse:
Human:
89
96
Kcr
Kcr
158
167
Kcr
KcrKcr
Kcr
118 119
125
Kac
Mouse:
Human:
13
Kac
Kcr
36
Kac Kac
KcrKcr
KcrKcr
Kcr
15
Kac
Kcr Kcr
Kcr Kcr
Kcr
Kcr
Kcr
Kcr
Kcr
K cr
Kac
Mouse:
Human:
Kcr
Kcr
Kac
15 16
20
Kac
23 24
Kac
Kcr
Kcr
Kcr
Kcr
34
Kac
Kcr
Kcr
Kcr
Kcr
Kcr
Kcr
H3 Human: NH2-ARTKQTARKSTGGKAPRKQLATKAARKSRYQKSTRIRGERA-COOH
4
Mouse:
Human:
14
Kac
Kac
Kcr
Kcr
Kcr
Kcr
18
Kac
23
Kac
Kac
27
Kac
56
Kac
K cr
Kcr
H4 Human: NH2-SGRGKGGKGLGKGGAKRHRKVLRDNIQGTLYGFGG-COOH
5
Kac
Kac
12
Kac
16
Kac
20
Kac
Figure 1. Experimental Strategy and Results for Identified Histone PTM Sites
(A) Schematic diagram of the experimental design for comprehensive mapping of PTM sites in linker and core histones from HeLa cells. Histone extracts were
in-solution trypticly digested without chemical propionylation (Method I), chemically propionylated after in-solution tryptic digestion (Method II), chemically
propionylated before in-solution tryptic digestion (Method III), and in-gel digested after SDS-PAGE gel separation. Samples from Methods I and II were further
subjected to IEF fractionation to generate 12 fractions.
(B) Peptide sequence coverage of linker and core histones in each of the four methods is shown.
(C) A table summarizing all the PTM sites identified by this study. Abbreviations: me, monomethylation; me2, dimethylation; me3, trimethylation; fo, formylation;
ac, acetylation; oh, hydroxylation; and cr, crotonylation.
(D) A diagram showing sites of histone PTMs other than Kcr identified in this study. Amino acid residue number is indicated below its sequence. Gray and blank
boxes indicate N-terminal and globular core domains, respectively.
1018 Cell 146, 10161028, September 16, 2011 2011 Elsevier Inc.
+ X-CoA
Crotonyllysine
X = acetyl, crotonyl
Lysine
Acetyllysine
Crotonyllysine
Acetyllysine
C
Multiple steps
(E) Illustrations of histone Kcr sites in human HeLa cells and mouse MEF cells. All Kcr sites are shown in red and underlined. Previously reported Kac
sites are shown in blue. See also Figure S1 and Supplemental Information 1, Supplemental Information 2, Supplemental Information 3, and Supplemental
Information 4.
Cell 146, 10161028, September 16, 2011 2011 Elsevier Inc. 1019
y 10 y 9 y 8
y 9++
580.8181
z=2
100
y7
y6 y5
y4 y3
y2
P E P A Kcr S A P A P K
467.7691
b2
b4
b5
b6 b7
b9
In Vivo
y6
580 582
934.5306
570.3210
b2
y2
227.1013
244.1640
b4
y4
395.1901
y5
412.2520 483.2897
y3
100
y7
766.4409
500
y8
b9
917.4662
837.4792
678.3420
200
100
b7
b 6 749.3759
b5
591.3103
315.2015
y9
y 10
1063.5884
800
1100
467.7692
580.8174
z=2
Synthetic
0
580 582
227.1013
244.1644
934.5308
570.3215
412.2527
395.1904
483.2896
315.2018
0
200
500
1063.5768
800
1100
467.7691
580.8176
z=2
100
766.4418
591.3106
678.3412 749.3792
917.4660
837.4786
Mixture
934.5307
580 582
570.3215
412.2524
591.3104
483.2893
395.1922
227.1014
244.1642
766.4417
749.3783
917.4668
837.4816
678.3448
0
200
500
1063.5651
800
1100
m/z
D
61.82
61.12
60.08
In Vivo
40
80
Synthetic
40
80
Mixture
40
80
Retention Time (min)
with a mass shift of + 68.0230 Da, the synthetic Kcr peptide with
the same peptide sequence (PEPAKcrSAPAPK), and the mixture
of the two peptides exhibited almost identical parent masses
1020 Cell 146, 10161028, September 16, 2011 2011 Elsevier Inc.
C
Competition
Kcr
H1
Peptide
Kac
Kpr
Kbu
WB: anti-Kcr
Kcr
Core Histones
1 ng
Crotonate (mM)
5 ng
50
100
WB: anti-Kcr
H1
25 ng
Blue staining
Blue staining
Core Histones
D
y 10 y 9 y 8
y 9++
100
b2
581
b2
y4
412.2546
y5
570.3226
y2
b9
y8
b7
y9
938.5569
b5
y 7 841.4987 b 9
753.4058
595.3361 b
921.4927
770.4672
6
682.3664
483.2908
0
200
b5 b6 b7
y6
227.1019
y2
y6 y5 y4
D 4- crotonate labeling
585
244.1656
y7
P E P A KD4-cr S A P A P K
469.7821
582.8303
z=2
500
800
y10
1067.5991
1100
m/z
Competition
Kcr
WB: anti-Kcr
Blue staining
1022 Cell 146, 10161028, September 16, 2011 2011 Elsevier Inc.
A
Input
H3K4me3
H3K4me1
Kcr
H3K4 1
H3K4me1
Others
15%
TSS
23%
Gene
body
17%
Enhancer
45%
60
50
H3K4me3
Kcr
Crotonylation
40
30
20
10
0
-10
Genomic distribution of
84 435 Crotonylation
84,435
Croton lation peaks
-2000
lowest 20%
H3K4me3
Kcr
Crotonylation
6
4
2
0
-2
50
H3K4me1
2000
-1000
0
1000
Relative distance to TSS (bp)
2nd 20%
40
3rd 20%
4th 20%
30
top 20%
20
10
-4000
-2000
2000
4000
-2000
-1000
1000
2000
Figure 6. Correlation of Histone Kcr with Gene Expression in Meiotic and Postmeiotic Male Germ Cells and in Tissues
(A) Hypercrotonylation wave in elongating spermatids. Kcr was detected on paraffin mouse testis tubule sections representing different stages of spermatogenesis by immunohistochemistry (IH) using an anti-Kcr antibody. Pre-meiotic spermatogonia (Spg) and meiotic spermatocytes (Spc) cells are present at the
periphery and middle of the tubule sections, whereas postmeiotic round (RS), elongating (ES) and condensing (CS) spermatids are near the lumen. The nuclei of
ES are positive for Kcr.
(B and C) Genes associated with higher Kcr in RS than Spc are mostly postmeiotically activated and show a predominant expression in the testis. The genes
associated with Kcr peaks were divided into three categories according to their Kcr levels in Spc and RS: (1) Spc = RS, similar Kcr levels between Spc and RS; (2)
Spc > RS, lower Kcr levels in RS than Spc (fold change > = 2); (3) Spc < RS, higher Kcr levels in RS than Spc (fold change > = 2). The expression of these genes in
male germ cells (B) and tissues (C) was then compared among the three categories. (B) Expression in male germ cells. Left panel: respective proportions of genes
(y axis) with higher expression either in Spc or RS among the three gene categories (x axis). Right panel: heatmap showing the expression of the third category of
genes (Kcr, Spc < RS) in Spc (4 samples) and in RS (4 samples). Color scale showing low expression in green to high expression in red. (C) Expression of genes in
tissues. Left panel: pie charts showing the respective proportions of genes with the highest level of expression in the indicated tissues. Genes with the highest
tissue-specific expression are those whose levels of expression, in the indicated tissue, are elevated by at least two standard deviations above the mean
expression in all tissues. Right panel: heatmap showing the expression of the third category of genes in the indicated tissues.
1024 Cell 146, 10161028, September 16, 2011 2011 Elsevier Inc.
(D) The X-linked genes are highly and specifically marked by histone Kcr in RS. The respective proportion (%) of genes associated with Kcr (left panel) or Kac
peaks (right panel) among chromosomes in male germ cells is shown. See also Figure S3.
Cell 146, 10161028, September 16, 2011 2011 Elsevier Inc. 1025
Third, our results show that KATs and HDACs exert very different
effects upon Kac and Kcr. Fourth, the status of short-chain lysine
acylation may be modulated in response to CoA concentration.
In the case of crotonyl-CoA, it can be generated either from
butyryl-CoA by short chain acyl-CoA dehydrogenase, or from
glutaryl-CoA by glutaryl-CoA dehydrogenase. Once formed,
crotonyl-CoA can be converted to acetyl-CoA for the TCA cycle.
Finally, histone Kac and Kcr mark different sets of gene in some
differentiated cells, as we report here in mouse sperm cells.
Strikingly, in both human somatic and mouse male germ cell
genomes, histone Kcr specifically labels enhancers and, most
precisely, the TSS of active genes. In postmeiotic male germ
cells, a gain in histone Kcr is a consistent indicator of an X-linked
haploid cell-specific gene expression program. Indeed, the identification of a subset of genes presenting increased Kcr in round
spermatids allowed us to show that such genes are enriched on
the X chromosome and are mostly predominantly expressed in
the testis with a postmeiotic pattern of expression.
Meiosis is known to be associated with the inactivation of sex
chromosomes, called meiotic sex chromosome inactivation
(MSCI). MSCI initiates in pachytene spermatocytes, and sex
chromosome gene silencing continues after meiosis in round
spermatids (RS) until the general shutdown of transcription
(Namekawa et al., 2006; Turner, 2007). However, in round spermatids, a significant number of X-linked genes have been shown
to be specifically reactivated (Mueller et al., 2008; Namekawa
et al., 2006). Interestingly, our results showed that histone Kcr
specifically marks X-linked genes that are postmeiotically
expressed. Consistent with this observation, our in situ experiment uncovered a remarkable postmeiotic labeling of Kcr on
sex chromosomes. In addition, our result indicates that the addi1026 Cell 146, 10161028, September 16, 2011 2011 Elsevier Inc.
EXPERIMENTAL PROCEDURES
In-Solution Proteolytic Digestion and Chemical Derivatization
of Histone Proteins
In-solution tryptic digestion of histone samples was carried out using a
protocol previously described (Kim et al., 2006; Luo et al., 2008). In vitro lysine
propionylation of histone extract and tryptic histone peptides was performed
as previously described (Garcia et al., 2007a). Histone extracts were in-solution digested without chemical propionylation, chemically propionylated after
in-solution digestion, or chemically propionylated before in-solution digestion.
Isoelectric Focusing Fractionation
The histone proteolytic peptides were separated using an Agilent 3100
OFFGEL Fractionator (Agilent, Santa Clara, CA) according to the manufacturers instructions. Twelve fractions were obtained from each IEF fractionation experiment.
Nano-HPLC/Mass Spectrometric Analysis
The tryptic digests were injected into a NanoLC-1D plus HPLC system
(Eksigent Technologies, Dublin, CA), and analyzed by an LTQ-Orbitrap Velos
mass spectrometer (Thermo Fisher Scientific, Waltham, MA). Full scan MS
spectra from m/z 3501400 were acquired in the Orbitrap. Twenty of the
most intense ions were isolated for MS/MS analysis.
Protein Sequencing Alignment
All MS/MS spectra were searched against the NCBInr human protein
sequence database using Mascot and PTMap software (Chen et al., 2009).
For histone samples that were generated by tryptic digestion of propionylated
histones, the specific parameters included lysine propionylmethylation
(+ 70.04187 Da) and lysine propionylation as variable modifications. For
histone samples propionylated after trypsin digestion, N-terminal propionylation was included as a fixed modification. All the identified peptides were
manually verified according to the rules described previously (Chen et al.,
2005).
Generation of Pan Anti-Kcr Antibody
The pan anti-Kcr antibody was generated and purified from rabbit with
lysine-crotonylated bovine serum albumin (BSA) as an antigen. For more
details, see Extended Experimental Procedures.
ChIP-Seq
ChIP-seq for histone Kcr or Kac was carried out as previously described
with 500 mg IMR90 chromatin (or 100 mg of fractionated germ cells chromatin)
and 5 mg pan anti-Kcr or anti-Kac antibody (Hawkins et al., 2010). ChIP-seq
libraries for sequencing were prepared following Illumina protocols (Illumina,
San Diego, CA) with minor modifications. Libraries for input samples were
generated using 20 ng corresponding input chromatin. Briefly, ChIPed DNA
was first blunted with END-IT DNA repair kit (Epicenter Biotechnology, Madison, WI) and then incubated with Klenow (exo-) (New England Biolabs, MA)
and dATP to generate single base 30 -dA overhang. Illumina sequencing
adaptor was then ligated to the resulting DNA, and followed by size selection
(180-400bp) from a 8% acrylamide gel. This size-selection step was repeated
after PCR amplification with DNA primers supplied by Illumina. Libraries were
sequenced using Illumina GAII or HiSeq machine as per manufacturers protocols. Following sequencing cluster imaging, base calling were conducted
using the Illumina pipeline. Reads were mapped to human hg18 (for IMR90
data) or mouse mm9 (for sperm cell data) genome build with a bowtie software
package. Total mapped tags were paired down to unique, monoclonal tags.
These are tags that mapped to one location in the genome and each sequence
is represented once.
For additional experimental materials and methods, see Extended Experimental Procedures, including methods for preparation of histones from
HeLa cells, in-solution proteolytic digestion and chemical derivatization of
histone proteins, HPLC/MS/MS analysis and protein sequence database
searching, verification of lysine crotonylated peptides by HPLC/MS/MS analysis, synthesis of BSA derivatives, conjugation of Kcr-immobilized agarose
beads, generation of pan anti-Kac and anti-Kcr antibodies, western blotting
Cell 146, 10161028, September 16, 2011 2011 Elsevier Inc. 1027
Garcia, B.A., Pesavento, J.J., Mizzen, C.A., and Kelleher, N.L. (2007b).
Pervasive combinatorial modification of histone H3 in human cells. Nat.
Methods 4, 487489.
Garcia, B.A., Shabanowitz, J., and Hunt, D.F. (2007c). Characterization of
histones and their post-translational modifications by mass spectrometry.
Curr. Opin. Chem. Biol. 11, 6673.
Gaucher, J., Reynoird, N., Montellier, E., Boussouar, F., Rousseaux, S., and
Khochbin, S. (2010). From meiosis to postmeiotic events: the secrets of
histone disappearance. FEBS J. 277, 599604.
Govin, J., Caron, C., Lestrat, C., Rousseaux, S., and Khochbin, S. (2004). The
role of histones in chromatin remodelling during mammalian spermiogenesis.
Eur. J. Biochem. 271, 34593469.
Hawkins, R.D., Hon, G.C., Lee, L.K., Ngo, Q., Lister, R., Pelizzola, M., Edsall,
L.E., Kuan, S., Luu, Y., Klugman, S., et al. (2010). Distinct epigenomic landscapes of pluripotent and lineage-committed human cells. Cell Stem Cell 6,
479491.
Hazzouri, M., Pivot-Pajot, C., Faure, A.K., Usson, Y., Pelletier, R., Sele, B.,
Khochbin, S., and Rousseaux, S. (2000). Regulated hyperacetylation of core
histones during mouse spermatogenesis: involvement of histone deacetylases. Eur. J. Cell Biol. 79, 950960.
Heintzman, N.D., Stuart, R.K., Hon, G., Fu, Y., Ching, C.W., Hawkins, R.D.,
Barrera, L.O., Van Calcar, S., Qu, C., Ching, K.A., et al. (2007). Distinct and
predictive chromatin signatures of transcriptional promoters and enhancers
in the human genome. Nat. Genet. 39, 311318.
Johnson, L., Mollah, S., Garcia, B.A., Muratore, T.L., Shabanowitz, J., Hunt,
D.F., and Jacobsen, S.E. (2004). Mass spectrometry analysis of Arabidopsis
histone H3 reveals distinct combinations of post-translational modifications.
Nucleic Acids Res. 32, 65116518.
Kim, S.C., Sprung, R., Chen, Y., Xu, Y., Ball, H., Pei, J., Cheng, T., Kho, Y.,
Xiao, H., Xiao, L., et al. (2006). Substrate and functional diversity of lysine
acetylation revealed by a proteomics survey. Mol. Cell 23, 607618.
Wysocka, J., Swigut, T., Xiao, H., Milne, T.A., Kwon, S.Y., Landry, J., Kauer,
M., Tackett, A.J., Chait, B.T., Badenhorst, P., et al. (2006). A PHD finger of
NURF couples histone H3 lysine 4 trimethylation with chromatin remodelling.
Nature 442, 8690.
Luger, K., Mader, A.W., Richmond, R.K., Sargent, D.F., and Richmond, T.J.
(1997). Crystal structure of the nucleosome core particle at 2.8 A resolution.
Nature 389, 251260.
Zee, B.M., Levin, R.S., Xu, B., LeRoy, G., Wingreen, N.S., and Garcia, B.A.
(2010). In vivo residue-specific histone methylation dynamics. J. Biol. Chem.
285, 33413350.
Luo, H., Li, Y., Mu, J.J., Zhang, J., Tonaka, T., Hamamori, Y., Jung, S.Y., Wang,
Y., and Qin, J. (2008). Regulation of intra-S phase checkpoint by ionizing
radiation (IR)-dependent and IR-independent phosphorylation of SMC3. J.
Biol. Chem. 283, 1917619183.
Margueron, R., Trojer, P., and Reinberg, D. (2005). The key to development:
interpreting the histone code? Curr. Opin. Genet. Dev. 15, 163176.
1028 Cell 146, 10161028, September 16, 2011 2011 Elsevier Inc.
Zhao, M., Shirley, C.R., Hayashi, S., Marcon, L., Mohapatra, B., Suganuma, R.,
Behringer, R.R., Boissonneault, G., Yanagimachi, R., and Meistrich, M.L.
(2004). Transition nuclear proteins are required for normal chromatin condensation and functional sperm development. Genesis 38, 200213.
Resource
Hughes Medical Institute, Cold Spring Harbor Laboratory, 1 Bungtown Road, Cold Spring Harbor, NY 11724, USA
and Computational Biology, University of Southern California, Los Angeles, CA 90089, USA
3These authors contributed equally to this work
*Correspondence: hannon@cshl.edu (G.J.H.), andrewds@usc.edu (A.D.S.)
DOI 10.1016/j.cell.2011.08.016
2Molecular
SUMMARY
Sample
Mapped
Distinct
Mismatches
BS Conversion
Methylation
CpG Coverage
CpGs Covered
Human
sperm (1)
609,127,589
388,835,058
1.58
0.992
0.724
8.8
0.96
sperm (2)
588,920,777
316,860,245
1.84
0.983
0.674
7.3
0.94
sperm (both)
1,198,048,366
705,695,303
1.70
0.988
0.701
16.1
0.96
0.93
Chimp
ESCs
940,731,922
366,844,212
0.64
0.988
0.663
14.1
sperm (1)
459,258,834
255,193,493
1.87
0.985
0.665
6.2
0.95
sperm (2)
520,905,232
327,796,614
1.70
0.984
0.672
7.4
0.94
sperm (both)
980,164,066
582,990,107
1.78
0.985
0.669
13.6
0.96
Mapped: reads mapping optimally to a single location in the reference genome. Distinct: number of genomic locations to which a read maps; when
multiple reads map to the same position, one with the best mapping score was selected at random, and all others discarded. Mismatches: average
number of mismatches for the reads indicated in the distinct fragments column. Bisulfite (BS) conversion rate was calculated at non-CpG cytosines.
Methylation: proportion of Cs in reads mapping over CpG dinucleotides.
RESULTS
Promoter
Repeat
an
u m (1)
a
C n(
hi
2
m )
p
C
(
1
hi
m )
p
H
um ( 2 )
a
H n(
u m 1)
a
C n(
hi
2
m )
p
C
(
1
hi
m )
p
H
um (2)
a
H n(
um 1)
a
C n(
hi
2
m )
p
C
(
1
hi
m )
p
(2
)
Genome-wide
ESC
Chimp (2)
Chimp (1)
Human (2)
um
0.64 0.70 0.70 0.70 0.67 0.73 0.66 0.65 0.26 0.32 0.25 0.22
0.87 0.86 0.91
0.93 0.91 0.95
0.69 0.68 0.80
0.88 0.87
0.93 0.91
0.70 0.68
0.89
0.94
0.78
Human (1)
Human (2)
Chimp (1)
Chimp (2)
ESC
CGI
0.95
0.85
0.84
0.64
0.79
um
H
C
G
Repeat
an
um (1)
a
C n(
hi
2
m )
p
C
(
1
hi
m )
p
ES ( 2
C )
C
G
I
H
um
a
H n(
u m 1)
a
C n(
hi
2
m )
p
C
(
1
hi
m )
p
ES ( 2
C )
C
G
I
(2
p
C
ES
m
hi
(1
Promoter
(2
0.82 0.62
0.68
0.81
0.80 0.90
0.64 0.58
0.79 0.75
hi
an
um
H
um
H
an
(1
Genome-wide
0.99 0.95
0.95
0.98
0.98 0.99
0.96 0.93
0.96 0.92
0.78 0.53
0.59
0.74
0.73 0.86
0.45 0.35
0.81 0.72
1.0
0.8
Sperm-specific hypo
(1,336 promoters)
ESC-specific hypo
(201 promoters)
0.6
Methylated in both
(5,380 promoters)
0.4
0.0
0.2
0.4
0.6
0.8
0.0
0.2
1.0
Homophilic
cell adhesion
Protein-DNA
complex assembly
0.8
0.6
0.4
0.2
0.0
1.0
Human sperm
Chimp sperm
Human ESC
CpG count
1k
100
10
10
100
1k
Promoter
10k
100k
100
1k
10k
HMR size (bp)
Repeat (non-promoter)
100k
100
Other
1k
10% CpG
10k
100k
1% CpG
C
ESC meth.
0.6
0.4
0.6
0.2
0.0
0.2
0.0
-4k
-3k
-2k
-1k
1k
2k
3k
4k
-5 -4 -3 -2 -1 +1 +2 +3 +4 +5
0.6
ESC
0.4
0.2
0.0
Schematic
Sperm
0.4
0.8
Methylation
Methylation
Methylation
0.8
Sperm meth.
0.8
-5 -4 -3 -2 -1 +1 +2 +3 +4 +5
CpG O/E
D
0.8 Methylated
in both
0.6
0.4
Nested HMR
(ESC+sperm)
0.2
-500
500
-500
240
Inter-CpG distance
Extended HMRs
500
180
60
180
120
E-Box O/E
1.5
Methylated
Extended HMR
0
-5 -4 -3 -2 -1 +1 +2 +3 +4 +5
1.0
1.0
0.5
0.5
0.0
ESC
60
Sperm
120
-5 -4 -3 -2 -1 +1 +2 +3 +4 +5 -5 -4 -3 -2 -1 +1 +2 +3 +4 +5
0.0
Cell 146, 10291041, September 16, 2011 2011 Elsevier Inc. 1033
density /10kb
0.6
0.5
Fraction hypomethylated
Satellite
methylation (sperm)
Human sperm
0.4
0.3
p12.3
chr12
Chimp sperm
p12.1
12q12
2e07
q14.1
4e07
q15 21.1
6e07
21.2
q21.31
Human ESC
0.1
q23.3
1e08
24.31
1e08
2Mb
Sperm
HMR 1 _
Meth.
0.2
q22 q23.1
8e07
0_
N
tR
N
R
N
D
N
LI
LT
R
llit
SI
te
Sa
SV
A
0.0
ESC
HMR 1 _
Meth.
Unassembled
Centromere
0_
Transposon
Satellite
D
250
HMR overlap
CpG density
0.2
0.1
ESC
L1PA2
200
Sperm
Promoter
150
100
50
0
-1K
0
1K
5UTR
2K
3K
4K
5K
6K
0
Hypo Hyper Hypo Hyper Hypo Hyper Hypo Hyper
LINE
LTR
SINE
SVA
HMR overlap
2000
LTR12C
1500
1000
500
0
-2K
-1K
0
1K
LTR
2K
3K
4K
Figure 4. Differential Repeat Methylation during Male Germ Cell and Somatic Reprogramming
(A) For each repeat class, the proportion of elements that overlap HMRs is shown for human sperm (red), chimp sperm (orange), and ESCs (blue).
(B) Upper: Average methylation level (red) and satellite density (blue) in 10 kb sliding windows across chromosome 12. Lower: Chromosome 12 centromeric
region with HMRs (blue) and methylation level (orange) for human sperm and ESCs.
(C) CpG densities of hypomethylated repeat copies (red) and methylated repeat copies (yellow) for LINEs, LTRs, SINEs, and SVAs.
(D) HMR overlap distribution around full-length L1PA2 and LTR12 ERV9 elements for human sperm (blue) and ESCs (red).
See also Figure S3 and Table S4.
1034 Cell 146, 10291041, September 16, 2011 2011 Elsevier Inc.
Hypomethylated copies
A
80%
Human
specific
70%
60%
50%
Human sperm
40%
Chimp sperm
30%
Human ESC
20%
10%
0%
C
D
E
SVA Family
C
40
HMR
20
10
20
15
SVA
Human (1)
CpG Methylation
number of copies
30
TLR1
10
5
0
0.0
0.2
0.4
0.6
0.8
1.0
methylation level
Human (2)
0
1
Chimp (1)
0
1
Chimp (2)
0
1
ESC
0.6
0.4
0.2
0.0
HMRs Promoter Nested &
HMRs
extended
HMRs
Cellspecific
HMRs
0.8
0.6
r2 = 0.92
Human
r2 = 0.87
0.0
CGI
0.4
0.8
Chimp
0.2
HMM
0.8 0.0
methylated
0.6
GG&F
1.0
CpG Observed/Expected
sperm
0.4
ESC
0.2
0.0
0.2
0.4
0.6
0.8
1.0
D
0.4
Derived allele:
G
A
0.3
Fraction of sequences
C
T
0.2
0.1
0.0
0.0
0.2
0.4
0.6
0.8
1.0
25%
Sequences diverged at
non-CpG sites in:
Human
Chimp
20%
15%
10%
5%
0%
Common
HMRs
Chimpspecific
HMRs
Humanspecific
HMRs
E
HMR
1
HTR3E (serotonin)
Human (1)
CpG Methylation
0
1
Human (2)
0
1
Chimp (1)
0
1
Chimp (2)
0
1
ESC
DISCUSSION
Sperm Methylation Patterns Are Conserved
Overall, sperm methylation patterns were highly similar in all our
samples. However, there were differences, even among individuals. There has been much discussion regarding the role of
germline transmission of epigenetic marks in interindividual variation (Curley et al., 2011). Changes in epigenetic state could
allow flexibility in phenotype that could be reverted over short
time spans if a trait became disadvantageous. Erosion of CpG
content provides a mechanism to allow fixation of a positive trait
in the long run. Thus, changes in DNA methylation patterns
preceding changes in DNA sequence presents an attractive
model for at least one mode of adaptation. Although evaluating
such hypotheses will require many more datasets, the work presented here builds a firm foundation for such studies.
Most Promoters Have HMRs in Sperm
Global resetting of DNA methylation patterns happens twice
during mammalian development: once during germ cell development and once early in embryogenesis. Our data permit a
genome-scale analysis of these two events. Although high
genome-wide levels of methylation are re-established during
both waves of epigenetic remodeling, some regions are protected and establish HMR boundaries that appear relevant
even in fully differentiated somatic cells (Hodges et al., 2011).
A few promoters showed selective hypomethylation in sperm,
and these are strongly enriched for annotations related to germ
cell processes. Far fewer were selectively hypomethylated in
ESCs, and these were not enriched in any particular annotation
category. Promoters of genes retaining nucleosomes have
recently been shown to be hypomethylated in human sperm
(Hammoud et al., 2009), and both of these features have
been proposed to aid rapid activation during development. We
find that gene-associated hypomethylation in sperm can be
extended to more than 70% of all annotated genes in both
human and chimp. Among these we failed to find any enrichment
for regulators of early development. Instead, it seems that
promoter regions are generally identified and bookmarked in
sperm (see Zaidi et al., 2010).
Distinct Processes of HMR Formation Shape Germ Cell
and ESC Methylomes
Genome-wide, CpG sites seem to adopt a methylated state by
default (Edwards et al., 2010). This raises the problem of
precisely how regions that become HMRs are identified as
such. Regions of hypomethylation at promoters have been
correlated with regulatory DNA in various developmental
contexts (Illingworth et al., 2008; Laurent et al., 2010; Rollins
et al., 2006; Straussman et al., 2009). Based upon analysis of
histone marks and on the proposed binding properties of
DNMT3s (Dhayalan et al., 2010; Zhang et al., 2010), active transcription and accompanying methylation of K4 on histone H3 are
thought to locally inhibit the methylation machinery. This could
enable large-scale recognition of promoter regions if widespread
transcription occurs during fetal germ cell development as
genomic methylation patters are erased and reset. It is also plausible that specific protein/DNA complexes act locally even in the
1038 Cell 146, 10291041, September 16, 2011 2011 Elsevier Inc.
by an increased sequence divergence even at non-CpG dinucleotides. One interpretation is that most species-specific HMRs
have arisen newly along one lineage with these novel functional
elements showing signs of recent adaptation. On the other hand,
if this accelerated sequence change were more a reflection of
relaxed selective pressure, we would expect species-specific
HMRs to more frequently result from loss of functional elements
along the opposite lineage. Resolution of these questions can
only come from a broadening to many more species of the
studies reported herein.
EXPERIMENTAL PROCEDURES
Detailed methods can be found in the Extended Experimental Procedures.
Sperm Collection
Two anonymous human donors were used and data pooled after sequencing.
Two chimp donors were used. Semen was collected at the New Iberia
Research Center (New Liberia, LA) or the Southwest National Primate
Research Center (San Antonio, TX, USA). Coagulated semen was separated
from the liquid phase manually. Both human and chimp samples were diluted
(1:1) in HBS buffer (0.01M HEPES, ph 7.4; 150 mM NaCl) and passed though
a silica-based gradient, SpermFilter (Cryobiosystems), by centrifugation
(according to manufacturers instructions).
Library Preparation
DNA from 100 million cells was extracted and sheared to a size of 150200
nt by sonication. Double-stranded DNA fragments were end repaired, A-tailed,
and ligated to methylated Illumina adaptors. Ligated fragments were bisulfite
converted using the EZ-DNA Methylation-Gold Kit (Zymo research). Following
PCR enrichment, fragments of 340 to 360 bp were size selected and
sequenced.
Computational Methods
Reads were mapped with RMAPBS (Smith et al., 2009). The accuracy of our
mapping method is discussed in the Extended Experimental Procedures.
Mapped reads were used to infer the methylation frequency at each CpG
dinucleotide. These frequencies, along with the number of reads contributing
to each frequency estimate, were supplied to a segmentation algorithm used
to identify HMRs. Ortholog mapping between human and chimp was done with
the liftOver tool available through the UCSC Genome Browser. Sequence
conservation between human, chimp, and was measured based on MULTIZ
44-way vertebrate alignments, also available through the UCSC Genome
Browser. Complete details of all computational methods are provided in the
Extended Experimental Procedures.
ACCESSION NUMBERS
Data analyzed herein have been deposited in GEO with accession GSE30340.
SUPPLEMENTAL INFORMATION
Doi, A., Park, I.H., Wen, B., Murakami, P., Aryee, M.J., Irizarry, R., Herb, B.,
Ladd-Acosta, C., Rho, J., Loewer, S., et al. (2009). Differential methylation of
tissue- and cancer-specific CpG island shores distinguishes human induced
pluripotent stem cells, embryonic stem cells and fibroblasts. Nat. Genet. 41,
13501353.
ACKNOWLEDGMENTS
Duncan, B.K., and Miller, J.H. (1980). Mutagenic deamination of cytosine residues in DNA. Nature 287, 560561.
We thank Michelle Rooks, Pramod Thekkat, and Colin Malone for help with
experimental procedures and Assaf Gordon, Luigi Manna, and the CSHL
and USC High Performance Computing Centers for computational support.
We thank Babette Fontenot (New Iberia Research Center) and Jerilyn Pecotte
(Southwest National Primate Center) for help with chimp sperm collection. We
thank Sergey Nuzhdin, Ed Green, Peter Calabrese, Maren Friesen, Magnus
Norborg, and Marie-Stanislas Remigereau for helpful discussions. This work
1040 Cell 146, 10291041, September 16, 2011 2011 Elsevier Inc.
Edwards, J.R., ODonnell, A.H., Rollins, R.A., Peckham, H.E., Lee, C., Milekic,
M.H., Chanrion, B., Fu, Y., Su, T., Hibshoosh, H., et al. (2010). Chromatin and
sequence features that define the fine and gross structure of genomic methylation patterns. Genome Res. 20, 972980.
Ehrlich, M., Zhang, X.Y., and Inamdar, N.M. (1990). Spontaneous deamination
of cytosine and 5-methylcytosine residues in DNA and replacement of 5-methylcytosine residues with cytosine residues. Mutat. Res. 238, 277286.
Enard, W., Fassbender, A., Model, F., Adorjan, P., Paabo, S., and Olek, A.
(2004). Differences in DNA methylation patterns between humans and chimpanzees. Curr. Biol. 14, R148R149.
Frescas, D., Guardavaccaro, D., Kuchay, S.M., Kato, H., Poleshko, A., Basrur,
V., Elenitoba-Johnson, K.S., Katz, R.A., and Pagano, M. (2008). KDM2A
represses transcription of centromeric satellite repeats and maintains the
heterochromatic state. Cell Cycle 7, 35393547.
Gardiner-Garden, M., and Frommer, M. (1987). CpG islands in vertebrate
genomes. J. Mol. Biol. 196, 261282.
Gaszner, M., and Felsenfeld, G. (2006). Insulators: exploiting transcriptional
and epigenetic mechanisms. Nat. Rev. Genet. 7, 703713.
Goodier, J.L., and Kazazian, H.H., Jr. (2008). Retrotransposons revisited: the
restraint and rehabilitation of parasites. Cell 135, 2335.
Hammoud, S.S., Nix, D.A., Zhang, H., Purwar, J., Carrell, D.T., and Cairns, B.R.
(2009). Distinctive chromatin in human sperm packages genes for embryo
development. Nature 460, 473478.
Hodges, E., Molaro, A., Dos Santos, C.O., Thekkat, P., Song, Q., Uren, P.,
Park, J., Butler, J., Rafii, S., McCombie, W.R., Smith, A.D., and Hannon,
G.J. (2011). Directional DNA methylation changes and complex intermediate
states accompany lineage specificity in the adult hematopoietic compartment.
Mol. Cell. Published online September 15 2011. 10.1016/j.cell.2008.06.028.
Parelho, V., Hadjur, S., Spivakov, M., Leleu, M., Sauer, S., Gregson, H.C., Jarmuz, A., Canzonetta, C., Webster, Z., Nesterova, T., et al. (2008). Cohesins
functionally associate with CTCF on mammalian chromosome arms. Cell
132, 422433.
Popp, C., Dean, W., Feng, S., Cokus, S.J., Andrews, S., Pellegrini, M., Jacobsen, S.E., and Reik, W. (2010). Genome-wide erasure of DNA methylation in
mouse primordial germ cells is affected by AID deficiency. Nature 463,
11011105.
Probst, A.V., Okamoto, I., Casanova, M., El Marjou, F., Le Baccon, P., and
Almouzni, G. (2010). A strand-specific burst in transcription of pericentric
satellites is required for chromocenter formation and early mouse development. Dev. Cell 19, 625638.
Rollins, R.A., Haghighi, F., Edwards, J.R., Das, R., Zhang, M.Q., Ju, J., and
Bestor, T.H. (2006). Large-scale structure of genomic methylation patterns.
Genome Res. 16, 157163.
Sasaki, H., and Matsui, Y. (2008). Epigenetic events in mammalian germ-cell
development: reprogramming and beyond. Nat. Rev. Genet. 9, 129140.
Schmid, C.W. (1991). Human Alu subfamilies and their methylation revealed by
blot hybridization. Nucleic Acids Res. 19, 56135617.
Illingworth, R., Kerr, A., Desousa, D., Jrgensen, H., Ellis, P., Stalker, J., Jackson, D., Clee, C., Plumb, R., Rogers, J., et al. (2008). A novel CpG island set
identifies tissue-specific methylation at developmental gene loci. PLoS Biol.
6, e22.
Shen, L., Wu, L.C., Sanlioglu, S., Chen, R., Mendoza, A.R., Dangel, A.W.,
Carroll, M.C., Zipf, W.B., and Yu, C.Y. (1994). Structure and genetics of the
partially duplicated gene RP located immediately upstream of the complement
C4A and the C4B genes in the HLA class III region. Molecular cloning, exonintron structure, composite retroposon, and breakpoint of gene duplication.
J. Biol. Chem. 269, 84668476.
Khan, H., Smit, A., and Boissinot, S. (2006). Molecular evolution and tempo of
amplification of human LINE-1 retrotransposons since the origin of primates.
Genome Res. 16, 7887.
Smith, A.D., Chung, W.Y., Hodges, E., Kendall, J., Hannon, G., Hicks, J., Xuan,
Z., and Zhang, M.Q. (2009). Updates to the RMAP short-read mapping software. Bioinformatics 25, 28412842.
Kochanek, S., Renz, D., and Doerfler, W. (1993). DNA methylation in the Alu
sequences of diploid and haploid primary human cells. EMBO J. 12, 1141
1151.
Straussman, R., Nejman, D., Roberts, D., Steinfeld, I., Blum, B., Benvenisty,
N., Simon, I., Yakhini, Z., and Cedar, H. (2009). Developmental programming
of CpG island methylation profiles in the human genome. Nat. Struct. Mol.
Biol. 16, 564571.
Lander, E.S., Linton, L.M., Birren, B., Nusbaum, C., Zody, M.C., Baldwin, J.,
Devon, K., Dewar, K., Doyle, M., FitzHugh, W., et al; International Human
Genome Sequencing Consortium. (2001). Initial sequencing and analysis of
the human genome. Nature 409, 860921.
Thomson, J.P., Skene, P.J., Selfridge, J., Clouaire, T., Guy, J., Webb, S., Kerr,
A.R., Deaton, A., Andrews, R., James, K.D., et al. (2010). CpG islands influence
chromatin structure via the CpG-binding protein Cfp1. Nature 464, 10821086.
Laurent, L., Wong, E., Li, G., Huynh, T., Tsirigos, A., Ong, C.T., Low, H.M., Kin
Sung, K.W., Rigoutsos, I., Loring, J., et al. (2010). Dynamic changes in the
human methylome during differentiation. Genome Res. 20, 320331.
Walsh, C.P., Chaillet, J.R., and Bestor, T.H. (1998). Transcription of IAP endogenous retroviruses is constrained by cytosine methylation. Nat. Genet. 20,
116117.
Lee, S.H., Cho, S.Y., Shannon, M.F., Fan, J., and Rangasamy, D. (2010). The
impact of CpG island on defining transcriptional activation of the mouse L1
retrotransposable elements. PLoS ONE 5, e11353.
Wang, H., Xing, J., Grover, D., Hedges, D.J., Han, K., Walker, J.A., and Batzer,
M.A. (2005). SVA elements: a hominid-specific retroposon family. J. Mol. Biol.
354, 9941007.
Li, E., Bestor, T.H., and Jaenisch, R. (1992). Targeted mutation of the DNA
methyltransferase gene results in embryonic lethality. Cell 69, 915926.
Weber, M., Hellmann, I., Stadler, M.B., Ramos, L., Paabo, S., Rebhan, M., and
Schubeler, D. (2007). Distribution, silencing potential and evolutionary impact
of promoter DNA methylation in the human genome. Nat. Genet. 39, 457466.
Liu, W.M., Maraia, R.J., Rubin, C.M., and Schmid, C.W. (1994). Alu transcripts:
cytoplasmic localisation and regulation by DNA methylation. Nucleic Acids
Res. 22, 10871095.
Mayer, W., Niveleau, A., Walter, J., Fundele, R., and Haaf, T. (2000). Demethylation of the zygotic paternal genome. Nature 403, 501502.
Mills, R.E., Bennett, E.A., Iskow, R.C., Luttig, C.T., Tsui, C., Pittard, W.S., and
Devine, S.E. (2006). Recently mobilized transposons in the human and chimpanzee genomes. Am. J. Hum. Genet. 78, 671679.
Okano, M., Bell, D.W., Haber, D.A., and Li, E. (1999). DNA methyltransferases
Dnmt3a and Dnmt3b are essential for de novo methylation and mammalian
development. Cell 99, 247257.
Ooi, S.K., Qiu, C., Bernstein, E., Li, K., Jia, D., Yang, Z., Erdjument-Bromage,
H., Tempst, P., Lin, S.P., Allis, C.D., et al. (2007). DNMT3L connects unmethylated lysine 4 of histone H3 to de novo methylation of DNA. Nature 448,
714717.
Ostertag, E.M., Goodier, J.L., Zhang, Y., and Kazazian, H.H., Jr. (2003).
SVA elements are nonautonomous retrotransposons that cause disease in
humans. Am. J. Hum. Genet. 73, 14441451.
Yamagata, K., Yamazaki, T., Miki, H., Ogonuki, N., Inoue, K., Ogura, A., and
Baba, T. (2007). Centromeric DNA hypomethylation as an epigenetic signature
discriminates between germ and somatic cell lineages. Dev. Biol. 312,
419426.
Yu, X., Zhu, X., Pi, W., Ling, J., Ko, L., Takeda, Y., and Tuan, D. (2005). The long
terminal repeat (LTR) of ERV-9 human endogenous retrovirus binds to NF-Y
in the assembly of an active LTR enhancer complex NF-Y/MZF1/GATA-2.
J. Biol. Chem. 280, 3518435194.
Zaidi, S.K., Young, D.W., Montecino, M.A., Lian, J.B., van Wijnen, A.J., Stein,
J.L., and Stein, G.S. (2010). Mitotic bookmarking of genes: a novel dimension
to epigenetic control. Nat. Rev. Genet. 11, 583589.
Zemach, A., McDaniel, I.E., Silva, P., and Zilberman, D. (2010). Genome-wide
evolutionary analysis of eukaryotic DNA methylation. Science 328, 916919.
Zhang, Y., Jurkowska, R., Soeroes, S., Rajavelu, A., Dhayalan, A., Bock, I.,
Rathert, P., Brandt, O., Reinhardt, R., Fischle, W., and Jeltsch, A. (2010). Chromatin methylation activity of Dnmt3a and Dnmt3a/3L is guided by interaction
of the ADD domain with the histone H3 tail. Nucleic Acids Res. 38, 42464253.
Cell 146, 10291041, September 16, 2011 2011 Elsevier Inc. 1041
Correction
1042 Cell 146, 1042, September 16, 2011 2011 Elsevier Inc.
Correction
1042 Cell 146, 1042, September 16, 2011 2011 Elsevier Inc.
Erratum
Cell 146, 1043, September 16, 2011 2011 Elsevier Inc. 1043
SnapShot: HighThroughput
Sequencing Applications
Hong Han,1 Razvan Nutiu,1 Jason Moffat,1 and Benjamin J. Blencowe1
Banting and Best Department of Medical Research, University of Toronto, Toronto, ON M5S 3E1, Canada
Genome Sequencing/Resequencing
Genome-wide polymorphism and mutation mapping
Genome assembly
CpG
METHYLATION
G
3
Pol ymer as e II
com pl ex
en
ic
TF
A SNP
Relative enrichment
Chromosome
H3K36me3
TF
TF
H3K36me3
H2K27ac
Exon
5 UTR
Polymerase II
Pol II
TF
RBP
Transcription
factors
Constitutively
spliced exon
TF
RBP
TF
Nucleosome
3 UTR
Cap
RBP
TRANSCRIPTION
RB
Intron Alt.
Exon
RBP
N as cen t R N A
SPLICING
Intron
Alternatively
spliced exon
RBP
RNA-binding
proteins
Transcriptome Sequencing/RNA-Seq
AAAAAAAA
AAAA
Poly(A) tail
ACTIVE TRANSLATION
Protein
m R N A EX P O R T
AAA
AAAAAAA
RNA-induced silencing complex
(RISC)
microRNA
60S
60S
Ribosome
40S
Ribosome Profiling
60S
5 UTR
AGO
mRNA
DEGRADATION
40S
40S
TRANSLATION
SUPPRESSION
1044
DOI 10.1016/j.cell.2011.09.002
Start
codon
CDS
CDS
Stop
codon 3 UTR
SnapShot: HighThroughput
Sequencing Applications
Hong Han,1 Razvan Nutiu,1 Jason Moffat,1 and Benjamin J. Blencowe1
Banting and Best Department of Medical Research, University of Toronto, Toronto, ON M5S 3E1, Canada
Amplification; Sequencing
Chemistries
Detection
Read lengtha
SE-single end; PE-paired end)
Run timeb
Fluorescence
100 bp (PE)
50 106 / 8 (X2)
11 days
Luminescence
400 bp (SE)
1 106
10 hr
Life/APGs SOLiD 3
www.appliedbiosystems.com
Fluorescence
50 bp (PE)
40 106 / 8 (X2)
14 days
Polonator G.007
http://www.polonator.org
Fluorescence
13 bp (PE)
10 106 / 8 (X2)
4 days
No amplification; Synthesis
Fluorescence
35 bp (SE)
20 106 / 25 (X2)
8 days
Pacific Biosciences
http://www.pacificbiosciences.com
No amplification; Synthesis
Fluorescence
>1000 bp (SE)
N/A
Ion Torrent
http://www.iontorrent.com
Change in pH
200 bp (SE)
Variable
<2 hr
Fluorescence
High-throughput, next-generation sequencing (NGS) technologies have revolutionized genomics, epigenomics and transcriptomics studies by allowing massively parallel
sequencing at a relatively low cost. In this SnapShot, we highlight the increasingly diverse applications of NGS, including genome sequencing/resequencing, transcriptome
sequencing, small RNA sequencing, analysis of DNA/RNA-protein interactions, and ribosome profiling. In addition, we provide a quick guide (Table 1) to the currently available
NGS platforms, together with their underlying methodologies and unique features.
Genome Sequencing/Resequencing
Whole-genome sequencing/resequencing and targeted genome resequencing have been used extensively for sequence polymorphism discovery and mutation mapping. These
applications are rapidly advancing our understanding of human health and disease and are also facilitating the de novo assembly of uncharacterized genomes.
DNA-Protein Interactions and Epigenome Sequencing
Chromatin immunoprecipitation coupled with high-throughput sequencing (ChIP-Seq) is a powerful technique for genome-wide profiling of DNA-protein interactions and epigenetic marks. It has facilitated a wide range of biological studies, including transcription factor binding, RNA polymerase occupancy, nucleosome positioning and histone
modifications. Complementary methods being used to study chromatin structure and composition are Methyl-Seq and DNase-Seq for profiling DNA methylation and DNasehypersensitive sites, respectively.
Transcriptome Sequencing/RNA-Seq
The introduction of transcriptome sequencing/RNA-Seq has provided a new approach for characterizing and quantifying transcripts. In general, total RNA, rRNA-depleted total
RNA, or poly(A)-selected RNA are converted to double-stranded cDNA fragments that are then subjected to high-throughput sequencing. This strategy has been applied for
profiling mRNA and noncoding RNA expression, alternative splicing, trans-splicing, and alternative polyadenylation and for mapping transcription initiation, termination, and RNA
editing sites. Related applications include targeted RNA-Seq, direct RNA-Seq, strand-specific RNA-Seq, and nascent RNA-Seq (e.g., global run-on sequencing, GRO-Seq, and
native elongating transcript sequencing, NET-Seq).
RNA-Protein Interactions
CLIP-Seq, also known as HITS-CLIP, is a method employing in vivo crosslinking of RNA to protein followed by immunoprecipitation and high-throughput RNA sequencing to
generate transcriptome-wide RNA-protein interaction maps. Modified CLIP-Seq technologies, such as PAR-CLIP (photoactivatable ribonucleoside-enhanced CLIP) and iCLIP
(individual nucleotide resolution CLIP), have been applied to increase crosslinking efficiency and resolution.
Small RNA Sequencing
Similar to RNA-Seq, sequencing of size-selected short RNA provides insight into small RNA populations in different organisms, tissue and cell types, developmental stages,
and disease states. It has greatly contributed to our understanding of the functions and regulatory mechanisms of different classes of small RNAs, such as microRNAs (miRNAs)
and Piwi-interacting RNAs (piwiRNAs). With the recent development of Argonaute (Ago) HITS-CLIP, it is possible to simultaneously detect Ago-bound microRNAs and mRNA
segments, which enables the large-scale mapping of in vivo miRNA-mRNA interactions.
Ribosome Profiling
In addition to the profound impact of NGS on transcriptomic studies, the development of methods enabling high-throughput sequencing of ribosome-protected mRNA fragments
has provided a powerful tool for the analysis of translationally engaged mRNA on a genome-wide scale.
Additional Applications and Future Directions
High-throughput sequencing is a rapidly evolving technology and will likely continue to change the face of omics studies in the years to come. Although NGS technologies
power a wide spectrum of current research applications, new innovations are continually being developed. These include barcode sequencing strategies for multiplexing the
analysis of samples, metagenomic analyses, protein-protein interactome mapping (Stitch-Seq), and high-definition measurement of DNA-affinity landscapes (HiTS-FLIP).
Future technical advances and applications are expected to further revolutionize our understanding of evolutionary biology and genotype-phenotype relationships and ultimately
to bring personalized medicine into the clinic.
1044.e1 Cell 146, September 16, 2011 2011 Elsevier Inc. DOI 10.1016/j.cell.2011.09.002
SnapShot: HighThroughput
Sequencing Applications
Hong Han,1 Razvan Nutiu,1 Jason Moffat,1 and Benjamin J. Blencowe1
Banting and Best Department of Medical Research, University of Toronto, Toronto, ON M5S 3E1, Canada
References
Chi, S.W., Zang, J.B., Mele, A., and Darnell, R.B. (2009). Argonaute HITS-CLIP decodes microRNA-mRNA interaction maps. Nature 460, 479486.
Core, L.J., Waterfall, J.J., and Lis, J.T. (2008). Nascent RNA sequencing reveals widespread pausing and divergent initiation at human promoters. Science 322, 18451848.
Ingolia, N.T., Ghaemmaghami, S., Newman, J.R., and Weissman, J.S. (2009). Genome-wide analysis in vivo of translation with nucleotide resolution using ribosome profiling. Science
324, 218223.
Licatalosi, D.D., Mele, A., Fak, J.J., Ule, J., Kayikci, M., Chi, S.W., Clark, T.A., Schweitzer, A.C., Blume, J.E., Wang, X., et al. (2008). HITS-CLIP yields genome-wide insights into brain
alternative RNA processing. Nature 456, 464469.
Metzker, M.L. (2010). Sequencing technologies - the next generation. Nat. Rev. Genet. 11, 3146.
Nutiu, R., Friedman, R.C., Luo, S., Khrebtukova, I., Silva, D., Li, R., Zhang, L., Schroth, G.P., and Burge, C.B. (2011). Direct measurement of DNA affinity landscapes on a high-throughput sequencing instrument. Nat. Biotechnol. 29, 659664.
Pan, Q., Shai, O., Lee, L.J., Frey, B.J., and Blencowe, B.J. (2008). Deep surveying of alternative splicing complexity in the human transcriptome by high-throughput sequencing.
Nat. Genet. 40, 14131415.
Park, P.J. (2009). ChIP-seq: advantages and challenges of a maturing technology. Nat. Rev. Genet. 10, 669680.
Smith, A.M., Heisler, L.E., St Onge, R.P., Farias-Hesson, E., Wallace, I.M., Bodeau, J., Harris, A.N., Perry, K.M., Giaever, G., Pourmand, N., and Nislow, C. (2010). Highly-multiplexed
barcode sequencing: an efficient method for parallel analysis of pooled samples. Nucleic Acids Res. 38, e142.
Yu, H., Tardivo, L., Tam, S., Weiner, E., Gebreab, F., Fan, C., Svrzikapa, N., Hirozane-Kishikawa, T., Rietman, E., Yang, X., et al. (2011). Next-generation sequencing to generate interactome datasets. Nat. Methods 8, 478480.
1044.e2 Cell 146, September 16, 2011 2011 Elsevier Inc. DOI 10.1016/j.cell.2011.09.002