You are on page 1of 25

Leading Edge

Review

EMT: 2016
M. Angela Nieto,1,6,* Ruby Yun-Ju Huang,2,3,6 Rebecca A. Jackson,4,6 and Jean Paul Thiery3,4,5,6,*
1Instituto de Neurociencias CSIC-UMH, Avda. Ramon y Cajal s/n, 03550 San Juan de Alicante, Spain
2Department of Obstetrics & Gynaecology, National University Hospital, Singapore 119228, Singapore
3Cancer Science Institute of Singapore, National University of Singapore, Singapore 117599, Singapore
4Department of Biochemistry, Yong Loo Lin School of Medicine, National University of Singapore, Singapore 117596, Singapore
5Institute of Molecular and Cell Biology, A-STAR, Singapore 138673, Singapore
6Co-first authors

*Correspondence: anieto@umh.es (M.A.N.), bchtjp@nus.edu.sg (J.P.T.)


http://dx.doi.org/10.1016/j.cell.2016.06.028

The significant parallels between cell plasticity during embryonic development and carcinoma pro-
gression have helped us understand the importance of the epithelial-mesenchymal transition (EMT)
in human disease. Our expanding knowledge of EMT has led to a clarification of the EMT program
as a set of multiple and dynamic transitional states between the epithelial and mesenchymal phe-
notypes, as opposed to a process involving a single binary decision. EMT and its intermediate
states have recently been identified as crucial drivers of organ fibrosis and tumor progression,
although there is some need for caution when interpreting its contribution to metastatic coloniza-
tion. Here, we discuss the current state-of-the-art and latest findings regarding the concept of
cellular plasticity and heterogeneity in EMT. We raise some of the questions pending and identify
the challenges faced in this fast-moving field.

Introduction transformation to transition more than a decade ago (first


During embryonic development, cells can transition between meeting of The Epithelial-Mesenchymal Transition International
epithelial and mesenchymal states in a highly plastic and dy- Association [TEMTIA]) already conveyed the concept of plas-
namic manner. A shift toward the mesenchymal state, in a pro- ticity and transitional states. It is therefore not surprising that
cess referred to as epithelial-mesenchymal transition (EMT), more recent work points to a greater flexibility in this transitional
modifies the adhesion molecules expressed by the cell, allow- process, and cells are no longer thought to oscillate between the
ing it to adopt a migratory and invasive behavior. The reverse full epithelial and full mesenchymal states, but rather, they move
of this processmesenchymal-epithelial transition (MET)is through a spectrum of intermediary phases. This purported plas-
associated with a loss of this migratory freedom, with cells ticity means that cells could linger in intermediary stages and
adopting an apico-basal polarization and expressing the junc- that they may frequently undergo a partial EMT program
tional complexes that are hallmarks of epithelial tissues (Thiery (Figure 1). Much of the work into the mechanisms regulating
et al., 2009). EMT has been carried out on cultured cells. However, the iden-
The EMT is executed in response to pleiotropic signaling fac- tification of hybrid intermediate EMT states during organ fibrosis
tors that induce the expression of specific transcription factors and in circulating tumor cells provides evidence that the spec-
(TFs) called EMT-TFs (e.g., Snail, Zeb, Twist, and others) and trum of intermediary phases previously observed in cultured
miRNAs together with epigenetic and post-translational regula- cells reflects what happens in vivo.
tors, many of which are involved in embryonic development, In this review, we will discuss these intermediate states of
wound healing, fibrosis, and cancer metastasis. Indeed, the sig- EMT, assessing whether the acquisition of stemness depends
nificant parallels between cell plasticity in embryonic develop- on EMT and exploring some of the latest findings regarding
ment and carcinoma progression led to the proposition that EMT in development and disease. We will also deliberate on
EMT is a central driver of epithelial-derived tumor malignancies the impetus to exploit EMT for therapeutic gain and pose some
(Cano et al., 2000; Thiery, 2002). EMT has since been shown to pending questions and challenges in this expanding field.
trigger the dissociation of carcinoma cells from primary carci-
nomas, which subsequently migrate and disseminate to distant Broadening the Reductionists View: Intermediate EMT
sites. Significantly, it is the MET that is then believed to trigger Phenotypes
the cessation of migration, inducing the same cells to proliferate From a reductionists viewpoint, the operational definition of EMT
and seed the new tumor. describes the loss and gain of certain characteristics during a sin-
The classic description of EMT as the transformation of epithe- gle transition between two states. As such, most experimental
lial cells into mesenchymal cells (Hay, 1995) may have invoked models tend to consider a dramatic change in the expression
the perception of this process as a shift between two alternative of selected epithelial and mesenchymal markers to confirm
states, mesenchymal or epithelial. However, the change from EMT (Thiery and Sleeman, 2006): E-cadherin, occludins, and

Cell 166, June 30, 2016 2016 Elsevier Inc. 21


Figure 1. Transitions through the Different
States along the EMT Spectrum
EMT can be considered as a continuum, whereby
cells exhibit epithelial (E), intermediate (EM), and
mesenchymal (M) phenotypes. As cells make the
transition from left to right, they sequentially lose
apico-basal polarity and cell-cell adhesions and
they gain front-back polarity and enhanced cell-
matrix interactions. The colored spectrum de-
notes hypothetical transitions (x axis). Different
models predict that both stability and metastability
could occur during the phase transitions where
there is a point of low energy or a net thermody-
namic equilibrium (y axis), in function of the
intrinsic, multi-parametric EMT regulators (z axis).
These regulators include transcription factors
(SNAI1/2, ZEB, TWIST1, GRHL2, OVOL1/2, and
PRRX1), post-transcriptional regulators (miRNAs),
and the proposed epigenetic controls at the pro-
moters of epithelial genes (Tam and Weinberg,
2013). When the system is perturbed in response
to EMT signals (e.g., hypoxia, growth factors, etc.),
cells could pass through a thermodynamic hump
that leads to a distinct phenotype (e.g., EM1;
EM3). The backward arrows represent reversal to
more epithelial states. It is worth noting that the
existence of discrete metastable states (e.g., EM1,
EM3) has not been proven and that the existence
of quasi-stable transitional states is also predicted
and found in different contexts, including fibrosis
(e.g., EM2). Importantly, although full reversibility
is evident during embryonic development, it is not
clear whether there is a point of no return in the
EMT undertaken by adult cells. Finally, it seems
that, although the mesenchymal state prevents
metastatic outgrowth, a full reversion to the
epithelial phenotype is not required for metastatic
colonization. TJ, tight junction; AJ, adherens
junction; DS, desmosome.

cytokeratins are the most commonly used markers for the epithe- linear energy gradient based on epigenetic changes between
lial trait and N-cadherin and vimentin for the mesenchymal (Thiery different EMT states has been hypothesized (Figure 1) (Tam
et al., 2009). This oversimplification ultimately led to some confu- and Weinberg, 2013). More recently, the free-energy landscape
sion and considerable debate as to whether EMT actually occurs of EMT was modeled, proposing three thermodynamically and
in some circumstances, such as during carcinoma progression kinetically stable states for EMT: the epithelial, maturation, and
(Tarin et al., 2005). However, the definition of EMT has now mesenchymal states (Zadran et al., 2014). Energy is released af-
been broadened, based on many observations and particularly ter exiting the epithelial state as the cell moves to a maturation
those regarding the so-called partial EMT. A partial EMT state state, which corresponds to an intermediate EMT state (Zadran
has been noted in association with many developmental, wound et al., 2014), and this energy is then consumed for the cell to
healing, fibrosis, and cancer processes (Arnoux et al., 2008; reach the mesenchymal state. Thus, the intermediate EMT state
Blanco et al., 2007; Futterman et al., 2011; Grande et al., 2015; in this model is not equivalent to the metastable state described
Leroy and Mostov, 2007; Lovisa et al., 2015; Grigore et al., with biophysical models. Computational modeling, including
2016), evident through the existence of intermediate hybrid those that consider the mutual inhibitory loops between several
epithelial and mesenchymal phenotypes (Abell et al., 2011; microRNAs (miRNAs) and EMT transcriptional drivers like Snail1
Huang et al., 2013; Jordan et al., 2011; Yu et al., 2013). and Zeb1, also accepts an intermediate hybrid EMT state that
Cells bearing this hybrid phenotype have been referred to as could favor the progress of developmental programs and meta-
metastable (Lee et al., 2006; Tam and Weinberg, 2013), re- static potential (Jolly et al., 2015; Lu et al., 2013; Tian et al., 2013;
flecting the flexibility of these cells to induce or reverse the Zhang et al., 2014). The inclusion of additional reciprocal inhibi-
EMT process (Rosano` et al., 2006). In physics and chemistry, tory loops that involve other transcription factors (e.g., Zeb1 with
the metastable state denotes an isolated system that is main- Ovol2 and Grhl2) and the description of these as phenotypic sta-
tained in a delicate equilibrium and not in a state of least energy bility factors indicates that the network is capable of generating
(Cheng and Keller, 1998). Thus, metastability suggests a dy- additional intermediate stabilized states that, therefore, are not
namic phase transition along the energy gradient. Adapting necessarily metastable (Hong et al., 2015; Jolly et al., 2016).
this concept of metastability to EMT probably fails to accurately Figure 1 depicts putative intermediate states in a hypothetical
reflect thermodynamic laws. Direct evidence of the energy tran- diagram to highlight the transitional nature of the process, which
sition that occurs during the EMT process is limited, although a can include stable and metastable states.

22 Cell 166, June 30, 2016


Figure 2. The Core Regulatory Machinery
of EMT
Although multiple non-coding RNAs control EMT,
two regulatory networks have been described
that can be considered as the core regulatory
machinery: the miR34-SNAI1 and miR200-ZEB1
axes. These two miR-transcription factor (TF) axes
employ a double-negative feedback mechanism
in which miR34-SNAI1 and miR200-ZEB1 repress
each other. SNAI1 and ZEB1 further repress
miR200 and miR34, respectively. At this level,
other TFs, such OVOL1/2 and GRHL2, further feed
into this core regulatory machinery to protect
the epithelial phenotype. The miR34-SNAI1 and
miR200-ZEB1 core not only contributes to the
epigenetic control of EMT, but are also targets
for epigenetic modifications. Downstream of the
miR34-SNAI1 and miR200-ZEB1 axes, regulation
of transcript processing would shape the land-
scape of epithelial and mesenchymal effectors,
for example, through alternative splicing. These
EMT effectors may also be subject to epigenetic
modifications. How other TFs, like TWIST1 and
PRRX1, affect the network and its epigenetic
control requires further study, although it is clear
that, unlike Snail and Zeb, they are much more
potent mesenchymal promoters than epithelial
repressors.

Evidence of transitional EMT states comes from multiple con- skeleton, the cellular machinery driving migration and invasion.
texts. Whereas many studies use the co-expression of epithelial This wider view of EMT offers a more dynamic interpretation of
and mesenchymal markers to define the hybrid state (Huang the fluidity and plasticity of this phenomenon (Figure 1).
et al., 2013; Yu et al., 2013), cells do not need to gain mesen-
chymal traits in a partial EMT. Indeed, early EMT might only Complex Regulatory Networks of EMT: Beyond
involve a dampening of epithelial properties like apico-basal Transcriptional Control
polarity and a remodeling of junctional complexes in favor of EMT is modulated at different levels by integrating epigenetic
cell-substrate adhesions (Huang et al., 2012). Even when a modifications, transcriptional control, alternative splicing, pro-
mesenchymal trait is gained, there is still much heterogeneity tein stability, and subcellular localization. Thus, the mechanisms
in terms of the mesenchymal characteristic induced. In an anal- that govern the intermediary phases of EMT are non-linear.
ysis of 43 ovarian cancer cell lines, only 50% of cells with an in- Despite the efforts to identify common regulatory networks in
termediate phenotype (co-expressing cytokeratin and vimentin) normal development and disease, it is clear that some pathways
induced N-cadherin (Huang et al., 2013). This heterogeneity may may be unique to a given EMT event in a particular tissue or
reflect different biological consequences of the intermediate tumor subset. The regulation of classical EMT centers on the
states, such that cells losing E-cadherin that do not gain N-cad- transcriptional suppression of the prototypic adhesion molecule,
herin will most likely behave distinctly during migration and inva- E-cadherin (Lamouille et al., 2014; Thiery et al., 2009). The
sion to those that do gain N-cadherin, given their different adhe- activities of major EMT-TFs, such as SNAI1, SNAI2, ZEB1,
sive behavior (Chu et al., 2006; Halbleib and Nelson, 2006). ZEB1, and TWIST1, have been described and reviewed exten-
It is also worth noting that the partial EMT seems to be a final sively (Lamouille et al., 2014; Peinado et al., 2007). With these
state in some cases, as described during organ fibrosis. Dedif- EMT-TFs taking center stage over the past decade, our under-
ferentiated renal epithelial cells do not seem to undergo a full standing of EMT has evolved to explain how their transcripts
EMT nor do they undergo reversal during the course of the dis- are expressed and processed by miRNAs (Lamouille et al.,
ease. Importantly, they do not engage in the invasion program 2013) or alternative splicing (Warzecha and Carstens, 2012).
but rather they remain integrated in the tubules (Grande et al., Multiple miRNAs are thought to govern EMT, as reviewed else-
2015). Hence, it is important to consider not only epithelial or where (Daz-Lopez et al., 2014). As a prototypic regulatory
mesenchymal traits during epithelial transitions but also other model, we focus here on the more elaborate networks involving
programs associated with EMT, such as invasion, increased miR-200 and miR-34, together with ZEB1 and SNAI1. This regu-
survival or decreased proliferation. latory network (Figure 2) establishes a negative feedback that is
In sum, intermediate states probably reflect the delicate thought to maintain epithelial homeostasis under normal condi-
balance of transcriptional drivers and suppressors of EMT. tions (Puisieux et al., 2014). Indeed, various epithelial markers
This equilibrium is inevitably affected by epigenetic changes are downstream targets of this regulatory loop, including E-cad-
(Tam and Weinberg, 2013) and by the main effectors in the cyto- herin, claudins, and occludins (Lamouille et al., 2014). Alternative

Cell 166, June 30, 2016 23


splicing mediated by epithelial-specific regulatory protein 1 and et al., 2012). ZEB proteins are required for the co-repression
2 (ESRP1 and ESRP2) also affects the maintenance of epithelial in Jurkat cells mediated by a histone acetyltransferase (HAT)
features (Warzecha and Carstens, 2012). A panel of transcripts domain-containing protein, Tip60 (Hlubek et al., 2001), and
involved in cell-cell adhesion, cell motility, and cell-matrix adhe- ZEB1 itself is affected by the recruitment of PRC2 by the histone
sion are regulated by these ESRPs, including fibroblast growth methyltransferase, PRDM14 (Chan et al., 2013). Both inactive
factor receptor 2 (FGFR2), p120-catenin, CD44, and Mena (War- (H3K27Me3) and active (H3K4Me3) histone marks can be found
zecha and Carstens, 2012). Conversely, splicing programs pro- at the ZEB1 promoter region, presumably to facilitate a switch
moted by Quaking (QKI), SRSF2, and RBFOX2 are upregulated during EMT (Chaffer et al., 2013; Spaderna et al., 2008). There-
during EMT (Bonomi et al., 2013; Braeutigam et al., 2014; fore, the consequences of the epigenetic changes initiated by
Conn et al., 2015). QKI promotes the generation of circular these EMT-TFs during EMT are expected to confer dynamic
RNAs (circRNAs) by non-sequential back-splicing of pre- control over chromatin and nuclear organization during the tran-
mRNAs, and RBFOX2 has multiple targets that include the re- scription of downstream EMT effector genes, thereby contrib-
ceptor tyrosine kinase Ron, cortactin, and dynamin (Figure 2). uting to the plasticity evident during transitions.
Therefore, the maintenance or loss of epithelial features is gov- The miR-200 family is also subjected to epigenetic modifica-
erned by how transcripts are processed to their mature forms tions. The expression of miR-200 family members is regulated
prior to translation. This systems view provides an alternative by a histone demethylase, KDM5B (Enkhbaatar et al., 2013).
way to describe the EMT spectrum in terms of regulatory net- The miR-200b-200a-429 cluster is regulated by histone modifi-
works, both quantitatively and qualitatively. cations mediated by the polycomb group and the EMT suppres-
Importantly, the miRNA-TF regulatory circuits are not linear in sor GRHL2, and the miR-200c-141 cluster is regulated by DNA
EMT, as there is substantial diversity in the regulation exerted by methylation (Lim et al., 2013; Chung et al., 2016). This epigenetic
different miRNAs in EMT. Within the miR-200 family, miR-200a/ regulation of these EMT miRNA-TF networks further complicates
b/c, miR-141, and miR-429 all target ZEB1, yet they probably the entire regulatory landscape of EMT, especially given that
inhibit ZEB1 translation to a different extent. It is unclear how many EMT effector genes can also be directly regulated by his-
other novel TFs recently implicated in EMT or MET affect the tone modifications and methylation, including E-cadherin
interaction between the miR34/SNAIL and miR200/ZEB regula- (Figure 2) (Wu et al., 2012). Histone acetylation is altered in
tory circuits, such as PRRX1 (Ocana et al., 2012), grainyhead- epithelial, intermediate, and mesenchymal trophoblast stem
like 2 (GRHL2) (Cieply et al., 2012), or connective tissue growth cells, and these changes have been correlated with the expres-
factor (CTGF) (Chang et al., 2013). GRHL2 and OVOL1/2 form sion of different TF combinations (Abell et al., 2011; Jordan et al.,
a double-negative feedback loop with ZEB1 (Figure 2) (Hong 2011). Accordingly, an EMT acetylome could be derived by
et al., 2015), and GRHL2 directly regulates the transcription of crossing these changes in acetylation with known EMT signa-
miR-200 and OVOL1/2 upstream of ZEB1 (Chung et al., 2016; tures (Jordan et al., 2011).
Cieply et al., 2012). PRRX1 induces EMT in a SNAI1-independent At the protein level, EMT is also regulated by post-translational
manner, yet cooperates with TWIST1 to drive invasiveness modifications including the ubiquitin-mediated degradation of
(Ocana et al., 2012). Both PRRX1 and TWIST1 are more potent important effectors, such as that following the phosphorylation
mesenchymal inducers than epithelial repressors, and SNAI1 of Snail1 and Twist1 by GSK3b or MAP kinases, respectively
and ZEB1 are strong epithelial repressors and weaker mesen- (Hong et al., 2011; Zhou et al., 2004). F-box proteins also target
chymal promoters. In contrast to SNAIL and ZEB family mem- these and other EMT-TFs to the proteasome (Diaz and de Her-
bers, there is still little information regarding the regulation of reros, 2016). Phosphorylation also impinges into Snail1 subcellu-
TWIST and PRRX by miRNAs or by other non-transcriptional lar localization (Du et al., 2010; Zhang et al., 2012a), in which nu-
mechanisms (Figure 2). clear import and export pathways have also been characterized
Upstream and downstream epigenetic controls largely affect (Domnguez et al., 2003; Mingot et al., 2009, 2013). As such, it
the EMT miRNA-SNAIL-ZEB networks, as these EMT transcrip- seems to be a daunting task to construct a systems landscape
tional drivers are involved in a complex epigenetic regulatory that encompasses the regulatory networks involving miRNA-TF
loop (Bedi et al., 2014). Interestingly, their regulation might loops, transcript processing, and the epigenetic control of pro-
involve transitions through intermediate states (Jordan et al., tein stability to determine what controls the continuous transi-
2011). SNAI1 recruits chromatin modifiers, including SIN3A, tions through the complete EMT spectrum. However, the recent
histone deacetylase 1 (HDAC1), HDAC2 (Peinado et al., 2004), demonstration that a generic EMT signature is common to
lysine-specific demethylase1 (LSD1) (Lin et al., 2010), compo- various cancers (Tan et al., 2014) and other advances in our un-
nents of the polycomb repressor-2 (PRC2) complex (Herranz derstanding of EMT in physiological and pathological conditions
et al., 2008), and the G9a and Suv39H1 histone methyltrans- makes it reasonable to believe that such broad systems analyses
ferases (Dong et al., 2013; Lin et al., 2014). SNAI1 binds tran- are not so far away.
siently to its target promoters, triggering both transient and
long-lasting chromatin changes (Javaid et al., 2013). In addition, New Insights for EMT in Development
SNAI1 can interact with a lysyl oxidase, LOXL2 (Peinado et al., EMT drives important aspects of embryonic development. This
2005), and regulate heterochromatin transcription (Millanes-Ro- phenomenon became apparent to developmental biologists in
mero et al., 2013). On the other hand, ZEB1 represses its target the late nineteenth century, presenting clear histological evi-
genes by recruiting the LSD1-containing co-repressor complex dence that, during gastrulation, mesenchymal cells originated
(Wang et al., 2007), as well as HDAC1 and HDAC2 (Aghdassi from the adjacent epiblast. EMT was established experimentally

24 Cell 166, June 30, 2016


by showing that embryonic and adult epithelia embedded in 3D with the phenotype observed in mouse embryos, where Snail1
collagen gels converted into migratory, invasive fibroblast-like mutant embryos do not survive gastrulation (Carver et al.,
cells (Greenburg and Hay, 1982). This seminal study provided 2001). However, the specific conditional deletion of Snail1 in
evidence that, contrary to popular belief, embryonic and also the epiblast affects mesoderm migration, but not its initial forma-
adult epithelial tissues were less stable than expected. Different tion (Lomel et al., 2009; Murray and Gridley, 2006).
model systems were explored to define the molecular mecha- Strong gastrulation defects are observed in FGFR1 knockout
nisms driving this remarkable conversion. One important finding mice (Ciruna and Rossant, 2001). FGF signaling, together with
was that rapid changes in epithelial cell shape were driven by the Nodal and Wnt3 signaling, is required for the generation of
complementary modulation of cell-cell adhesion and cell-sub- migratory mesenchymal mesodermal cells (Lim and Thiery,
strate interactions, mediated by N-CAM and E-cadherin, as 2012). FGF signaling maintains Snail1 expression, and Snail1
well as extracellular matrix (ECM) proteins such as fibronectin and Sox2/3 mutually repress each other, with Sox3 preventing
and integrin receptors, respectively (Boucaut et al., 1984; Edel- epithelial cell ingression and Snail1 promoting it by inhibiting
man et al., 1983; Takeichi, 1988). Another landmark was the dis- Sox3 in the primitive streak in chick and mouse embryos (Aclo-
covery that an ortholog of the Drosophila snail gene, expressed que et al., 2011).
during the formation of the ventral furrow (Leptin and Grunewald, The winged helix-turn-helix Ets2 transcription factor is ex-
1990), was critical to the formation of the primary mesenchyme pressed in extraembryonic ectoderm trophoblasts, and it plays
during gastrulation in the chicken embryo (Nieto et al., 1994). a role in the formation and elongation of the primitive streak in
The first signal-transduction pathways leading to EMT were mouse embryos (Polydorou and Georgiades, 2013). Indeed,
then defined from genetic studies in Drosophila and in vertebrate Ets2 promotes the expression of Elf5 and Cdx2 in order to main-
embryos. Here, we integrate some recent findings on EMT in tain trophoblast stemness, and it also induces bone morphoge-
development with the knowledge accumulated over the past netic protein 4 (BMP4) expression, which acts on the epiblast in a
two decades. paracrine manner to induce Wnt3. Wnt3 elevates Nodal and in-
Gastrulation duces Snail1, promoting primitive streak elongation and EMT,
Genetic studies in Drosophila revealed how dorsoventral polar- respectively. These findings add another layer of complexity to
ity, mediated by Toll receptor activation, engages a program EMT signaling in mouse gastrulation.
orchestrated by Twist and Snail. This program involves cell-cycle Neural Crest and Mesectoderm
repression by Tribbles, and the activation of actomyosin At the neurula stage, a group of stem/progenitor cells appears at
contractility by T48 and Fog through RhoGEF2. EMT of the the dorsal border of the neural plate of vertebrates, the neural
invaginated epithelium in the ventral furrow is then promoted crest (NC) (Baggiolini et al., 2015; Buitrago-Delgado et al.,
by the activation of heartless, an FGFR tyrosine kinase (Lim 2015), which generates glial cells, most neurons in the peripheral
and Thiery, 2012). Snail promotes an E- to N-cadherin switch nervous system, some endocrine and paraendocrine cells, and
(Oda et al., 1998) that, in conjunction with FGF signaling medi- all melanocytes. NC cells share intercellular adhesion systems
ated by heartless activation, initiates EMT in the nascent fly with adjacent neural epithelial cells that are destined to give
mesoderm. Recent findings question whether the repression of rise to the CNS, yet they engage in EMT to form a population
E-cadherin during the invagination process and/or the gain of of migratory cells that populate distant territories. The mecha-
N-cadherin are in fact key events during fly gastrulation (Schafer nisms driving EMT induction in NC cells are only partially shared
et al., 2014), with the suggestion that other mechanisms control across species, although a generic gene regulatory network has
the changes in morphology and the individualization of the cells been established (Simoes-Costa and Bronner, 2015), whereby
in the mesoderm. One possibility to reconcile earlier data with the EMT signaling module is believed to be embedded in a
these findings could be that E-cadherin turnover is accentuated rather complex network. Multiple feed-back loops regulate NC
upon forced overexpression in order to reduce the strength of development and guarantee the robustness of the system,
intercellular adhesion, giving rise to a much weaker phenotype anticipating compensatory regulatory mechanisms. Epigenetic
than that anticipated. Clearly, more work is required to clarify regulation also occurs during NC specification and migration,
how this intricate network is autoregulated when challenged, controlling the expression of important EMT regulators (Hu
leading to unpredictable and weak phenotypes and emphasizing et al., 2014).
the robustness of gastrulation in Drosophila. In the chick embryo, the NC territory at the trunk level is
The analysis of the gene regulatory networks in sea urchin induced by reciprocal rostro-caudal gradients of retinoic acid
gastrulation also indicated that Snail and Twist execute the (RA) and FGF. The territorys borders are refined by downstream
EMT process in primary mesenchyme formation (Saunders and signaling that involves opposing gradients of BMP4/noggin and
McClay, 2014). However, a detailed analysis of blastula epithelial Wnt1, as well as epigenetic controls that establish poised chro-
cell behavior defined five distinct sub-circuits involved in EMT: matin domains of genes like Snail, which are activated before cell
basement membrane remodeling, motility, apical constriction, emigration. In the NC territory, BMP4 induces N-cadherin cleav-
loss of apico-basal polarity, and changes in intercellular adhe- age by ADAM10 (Shoval et al., 2007), and it creates two
sion. Thirteen transcription factors contribute to the execution opposing gradients to control the downregulation of N-cadherin
of EMT, including Snail and Twist, but no single transcription and the expression of cadherin-6 (Park and Gumbiner, 2010),
factor is expressed in all five sub-circuits, highlighting the events that transiently mark the territory from which the NC dis-
complexity and robustness of the system even in sea urchin sociates. Cadherin-6 is also cleaved by ADAM10 and g-secre-
(Saunders and McClay, 2014). These findings are consistent tase (Schiffmacher et al., 2014), and thus, it is downregulated

Cell 166, June 30, 2016 25


following crest cell emigration when these cells acquire cad- The OFT obeys the same controls in the formation of the
herin-7. In zebrafish, cadherin-6 acts autonomously in NC pro- conotruncal endocardial cushion, albeit with the additional
genitors to localize activated RhoA in the apical domain. This complexity that the cells arise from the second heart field and
restricted distribution promotes actomyosin contractility and the NC. EndMT operates under the Tbx2-BMP2 axis, and
constriction of the apical domain, releasing the apical junctional when deregulated, retinoic acid (RA) can directly target Tbx2
complexes of NC cells to allow them to emigrate from the neural transcription and reduce BMP2 levels. Indeed, ectopic RA
fold (Clay and Halloran, 2014). In summary, the loss of apico- causes defects in the OFT cushion (Sakabe et al., 2012).
basal polarity, the modulation of cadherins and other adhesion Mesothelial cells of the sinus venosus located at the posterior
molecules, and the profound remodeling of the actin cytoskel- end of the embryonic heart form the proepicardium anlage.
eton allow EMT to occur and cells to acquire a mesenchymal Bmp2 expression in this anlage is weak, inducing the expression
migratory morphology. of Wilms tumor gene 1 (Wt1) and Tbx18 to sustain epicardial
At cranial levels, mesectodermal cells can also give rise identity while the cells migrate to the heart primordium to form
to mesodermal derivatives that include bone, cartilage, and the primitive epicardium. Pre-epicardial cells migrate and
meninges. Believed to be part of the NC, recent findingsstill assemble as a single-layered epithelial sheet around the
subjected to debatepropose that this may be a distinct yet myocardium through a MET-like process. Subsequently, the
adjacent ectodermal cell population that also undergoes EMT. epicardial layer undergoes EMT and gives rise to resident inter-
These cells would downregulate E-cadherin rather than N-cad- stitial fibroblasts within the myocardium, as well as smooth
herin as they dissociate from the non-neural epithelium of the muscle cells in the coronary arteries and at least 20% of the
neural fold (Lee et al., 2013; Weston and Thiery, 2015). However, endothelium (Cano et al., 2016). This process has recently
in Xenopus embryos, migratory cranial cells still express N-cad- been considered as an extended process of the EMT of meso-
herin during migration as they pursue collective cell migration, thelial cells, as the primitive epicardium never undergoes a full
having undergone only a partial EMT (Scarpa et al., 2015). MET (Ariza et al., 2016). Epicardial EMT is controlled by the
Heart neurofibromatosis type 1 (NF1) gene, as its deletion induces pre-
The heart develops from a group of mesodermal mesenchymal mature and more extensive EMT in the epicardium (Baek and
cells during gastrulation that assembles as two cardiogenic Tallquist, 2012). Wt1 is also essential for this process, inducing
mesodermal layers (the primary and secondary heart fields), the expression of either Snail1 or b-catenin to promote cellular
completing a first round of EMT and MET. The second round is dissociation and migration in the subepicardial layer. The
activated in the splanchnopleuric mesoderm to form the inner basic-loop-helix TCF21 also influences the EMT and lineage
endothelial layer, and it is followed by a third round to form the specification of epicardial cells (Acharya et al., 2012). On the
cardiac valves that arise from the dissociation of endothelial cells other hand, the Notch pathway promotes EMT by activating
in the atrio-ventricular region and outflow track (OFT). This third platelet-derived growth factor receptor (PDGFR) and producing
round should perhaps more precisely be termed an endothelial- TGFb in cooperation with Wt1, which can induce RA production
mesenchymal transition (EndMT). by the Raldh2 enzyme (Lim and Thiery, 2012; Perez-Pomares
Quite distinct mechanisms drive the successive rounds of and de la Pompa, 2011).
EMT/EndMT and MET during heart development (Lim and Together, these findings show that proper heart development
Thiery, 2012; von Gise and Pu, 2012). Whereas the first round requires a plethora of growth factors and pathways to drive
is basically a subprogram of gastrulation that depends on Wnt, epicardium-myocardium interactions, which are still to be
FGF, and TGFb signaling, the third is initially driven by the comprehensively integrated (as required in other developmental
BMP2 produced by myocardial cells. Nevertheless, BMP null- processes). It is widely accepted that pathways converge on the
homozygotes do not exhibit valve defects, indicating that activation of different combinations of EMT-TFs and that an
compensatory mechanisms exist similar to those seen in gastru- EMT code determines cell behavior, producing heterogeneity
lation and NC development. However, altering downstream ca- and tissue specificity. Establishing such gene regulatory net-
nonical signaling, such as interfering with Smad4, leads to a works and feedback loops ensures the robustness of pivotal
loss of the cardiac cushion. BMP2 can drive local EMT in the developmental processes, as observed during gastrulation and
atrioventricular region through the activity of its downstream NC development.
target Tbx2. The atrioventricular endocardium responds to the Although EMT in development has inspired cancer re-
myocardium through secreted BMP2 and TGFb1/2, triggering searchers to identify similar mechanisms in the progression of
canonical Smad pathways. All four ErbB receptors are important carcinomas, it is now clear that the regulation observed in normal
for cardiac valve formation, and while ErbB2-ErbB3 hetero- development would not only be used, but also expanded on in
dimers are activated by neuregulin in the endocardium in an au- cancer, and that additional effort is needed to describe the
tocrine manner (Lim and Thiery, 2012), ErbB3 signaling is also signaling networks that drive invasion and metastasis.
evident in the myocardium and is driven by the hyaluronic acid
produced by hyaluronan synthase (HAS)-2, a Tbx2 target. Notch Protection of the Epithelial Phenotype
signaling is another critical pathway activated in the endocar- Despite the high degree of epithelial plasticity observed during
dium (Luxan et al., 2016), and all three pathways, Notch, BMP/ embryonic development, it is appropriate to stress that, in
TGFb, and ErbB, contribute to endocardial EndMT to establish normal adult epithelial tissues, various mechanisms guarantee
the valve interstitial cells (Moskowitz et al., 2011; von Gise and epithelial homeostasis and protect the epithelial phenotype in or-
Pu, 2012). der to maintain tissue integrity and organ function. In addition to

26 Cell 166, June 30, 2016


the aforementioned global program driven by epithelial-specific, The Partial EMT Associated with Wound Healing
ESRP-mediated splicing (Warzecha and Carstens, 2012), the During epithelial wound healing, cells at the edge of the wound
epithelial-specific H2BK5Ac epigenetic mark also exists (Abell move into damaged area to re-establish normal tissue architec-
et al., 2011). Moreover, Ovol zinc-finger TFs appear to act as pro- ture in a process referred to as re-epithelialization. Since epithe-
tectors of the epithelial phenotype, with Ovol1/Ovol2 inhibiting lial cells are fairly immotile, they must undergo behavioral
the cells capacity to engage into EMT by activating epithelial changes to move as a coordinated group of cells and heal
genes and repressing multiple EMT drivers and mesenchymal the wound, ultimately reverting to their epithelial phenotype to
genes (Lee et al., 2014; Kitazawa et al., 2016). Ovol2 deletion reconstitute epithelial sheet integrity. Both basal and suprabasal
dramatically augments the EMT phenotype in epithelial cells of cells are involved in the repair process that involves a partial
the mammary gland, provoking considerable structural pertur- EMT (Shaw and Martin, 2016). The partial EMT is thought to
bation during extension of the terminal end buds (Watanabe rely on Snail2 under the control of the epidermal growth factor
et al., 2014). Elf5 shares similar properties in protecting the receptor (EGFR)-extracellular signal-regulated kinase 5 pathway
epithelial state by suppressing Snail2 during mammary gland in mouse skin explants (Arnoux et al., 2008). As such, wound
development and metastasis (Chakrabarti et al., 2012). healing is defective in adult mice deficient for Snail2 (Hudson
New functions for p53 extend its role of guardian of the et al., 2009). EGFR signaling also promotes re-epithelialization
genome (Lane, 1992) to include guardian of the epithelial in N-acetylglucosaminyltransferase V (GnT-V) transgenic mice
phenotype, as p53 also safeguards epithelial homeostasis by through the upregulation of Snail and Twist (Terao et al., 2011).
directly activating the miR-200 family. In mammary epithelial In addition, axon guidance molecules play an important role in
cells, the loss of p53 causes a decrease in miR-200c thereby this partial transition during wound healing in the mouse skin.
increasing EMT, with a concomitant increase in the stem cell pop- As such, upregulation of EphrinB1 in the basal and suprabasal
ulation (Chang et al., 2011). The latter is in part mediated by the cells downregulate components of the cell-cell adhesion com-
upregulation of Bmi1 (Shimono et al., 2009), which is expressed plexes, loosening the junctions to release tension and enable
by cisplatin-resistant bladder cancer cells that behave as cancer cell movement (Nunan et al., 2015). In fly larvae, Netrin A controls
stem cell (CSC)-like cells, undergoing EMT with the potential to EMT in the squamous epithelium by downregulating Frazzled, a
form spheres (Zhang et al., 2012b). EMT is also prevented by member of the deleted in colorectal carcinoma (DCC) family of
miR-200b, which inhibits mammosphere formation through tar- surface receptors, which are known to stabilize adherens junc-
geting Suz12, a subunit of the polycomb PRC2 complex (Iliopou- tions through the phosphorylation of moesin (Manhire-Heath
los et al., 2010). Another PRC2 component, EZH2, supresses et al., 2013). These studies offer insight into epithelial breakdown
EMT, invasion, and metastasis by repressing Snail1 in colorectal during EMT, which is important given that Netrin is overex-
cancer cells (Wang et al., 2015). Other protectors of the epithelial pressed in various cancers, and it induces cancer cell prolifera-
phenotype include proteins associated with DNA damage repair, tion and migration through the activation of YAP signaling (Qi
such as the ubiquitin ligases FBXW7 and Siah, which prevent et al., 2015).
EMT by targeting Zeb1 for degradation (Chen et al., 2015; Yang In the lung, wound healing is also associated with an EMT
et al., 2015), as well as ataxia-telangiectasia-mutated (ATM) ki- phenotype, typified by a reduction in cell-cell junctions, cell flat-
nase, which represses Sox9 expression (Russell et al., 2015). In tening, the acquisition of migratory properties, and the expres-
some cell contexts, Sox9 behaves as an important transcription sion of vimentin. Yet, this transient phenotypic change is limited
factor associated with EMT and stemness. to the basal cells of the airway and there is no evidence that
The programs implemented to protect the epithelial pheno- type II alveolar cells, which are responsible for the production
type are counterbalanced by other programs that promote and secretion of surfactant to reduce surface tension, undergo
EMT. As such, epigenetically driven global reprogramming of EMT or adopt mesenchymal features. By contrast, the club (or
specific chromatin domains and specific splicing programs clara) cells in the bronchiole airways appear to undergo a tran-
occur during the induction of EMT (Bonomi et al., 2013; Braeuti- sient EMT program during the regeneration of the bronchiolar
gam et al., 2014; McDonald et al., 2011). While the final response epithelium. Club cells act as stem cells, and they can terminally
would depend on the relative contribution of these positive and differentiate into ciliated cells during lung tissue repair (Vaughan
negative EMT programs, in the healthy adult, the protection of and Chapman, 2013).
epithelial homeostasis will generally prevail. Nevertheless, the Wound healing is often investigated using the dorsal closure
remarkable epithelial plasticity observed under physiological model in Drosophila, as EMT- and MET-like events are coordi-
conditions in certain tissues is worth noting, such as in the mam- nated during this process. During dorsal closure, epidermal
mary gland during puberty and pregnancy. Ductal elongation of flanksamnioserosamigrate to extend over the gap in the
the mammary gland during puberty is driven by terminal end epidermis, and as the epidermis progressively expands on the
buds. These multi-layered structures exhibit significant epithelial outer surface of the dorsal territory, the transient amnioserosa
cell plasticity to allow fat pad invasion through the collective epithelium becomes internalized. Epidermal pioneer cells at the
migration of cells with an intermediate EMT phenotype (Kurley leading edge undergo a partial EMT, forming filopodia and
et al., 2012; Shamir and Ewald, 2015). As the reactivation of lamellipodia that actively promote cell migration through
EMT programs in the adult go far beyond this process, the Cdc42/Rac-mediated activation of actomyosin contractility.
following sections discuss the implication of EMT in wound heal- The Scribble apico-basal polarity complex recruits p21-activated
ing, fibrosis, and cancer, with a focus on partial phenotypes in kinase (PAK) to the leading edge for polarized cell migration, and
terms of both physiology and pathology. when closure is complete, PAK recruits the Scrib complex to

Cell 166, June 30, 2016 27


restore septate junctions and apico-basal polarity (Bahri et al., et al., 2015; Lovisa et al., 2015). It seems that, although renal
2010). GRH, the homolog of mammalian GRHL-1, -2 and -3, reg- epithelial cells do not directly generate myofibroblasts, deletion
ulates wound healing in Drosophila amnioserosa through the of Twist or Snail in renal epithelial cells significantly attenuates
establishment of septate junctions (Narasimha et al., 2008). interstitial fibrosis in mouse models of TIF: unilateral ureteral
GRHL3 is also involved in epithelial homeostasis in the mouse obstruction (UUO), folic acid administration (FA), or nephro-
(Ting et al., 2005), where it contributes to wound healing by acti- toxic-serum-induced nephritis (NTN) (Grande et al., 2015; Lovisa
vating transglutaminase 1 and RhoGEF19 (Darido and Jane, et al., 2015).
2010). However, GRHL2 can compensate for the loss of Yet, how can EMT-TFs be so relevant for the development of
GRHL3 to heal the wounds of hind limb amputated mouse fibrosis if kidney epithelial cells are not the source of myofibro-
embryos (Boglev et al., 2011), suggesting redundancy is at play. blasts? These two recent studies both suggest that renal epithe-
It is important to note that, while EMT in cancer is deleterious, lial cells undergo a partial EMT: one pointing to perturbations in
wound-healing-driven EMT induced in response to injury is transporter proteins and cell-cycle regulation (Lovisa et al., 2015)
beneficial. Therefore, analyzing the similarities and differences and the other showing that epithelial cells lose epithelial markers
between these two types of EMT will be essential to better and apico-basal polarity, but they remain integrated in the tu-
design specific therapies (Leopold et al., 2012). Nonetheless, bules (Grande et al., 2015). Importantly, the injured epithelial
some EMT-healing responses may also be deleterious, such cells relay signals to the interstitium to promote the differentia-
as opacification of the posterior capsular (PCO), a major compli- tion of fibroblasts into myofibroblasts and the subsequent fibro-
cation after surgery for cataract treatment (Xiao et al., 2015), or genesis. The cells that underwent a partial EMT also secrete exo-
the exaggerated healing responses that lead to fibrosis and scar- somes and cytokines that help recruit (1) bone-marrow-derived
ring. Interestingly, defective healing and increased scarring was mesenchymal cells that also differentiate into myofibroblasts to
recently associated with prolonged inflammation (Qian et al., promote fibrogenesis and (2) macrophages that sustain inflam-
2016), highlighting the tight connection between a sustained mation (Borges et al., 2013; Grande et al., 2015; Lovisa et al.,
inflammatory response and the progression of fibrosis (as dis- 2015). Interestingly, although Snail factors were considered
cussed below). purely as transcriptional repressors, Snail is converted into an
activator upon its binding to Twist (Rembold et al., 2014) and
Reconciling the Role of EMT in Fibrosis also upon binding to CREB-binding protein (CBP). The latter al-
Fibrosis, or tissue scarring, is a hallmark of many chronic degen- lows Snail1 to directly activate the transcription of inflammatory
erative disorders and is associated with reduced organ function cytokines genes in cancer cells and during fibrosis (Grande et al.,
and eventual organ failure. Fibrotic disease is on the increase; for 2015; Hsu et al., 2014; Lovisa et al., 2015).
example, idiopathic pulmonary fibrosis (IPF) is estimated to Thus, even though kidney tubular cells do not become
match the incidence of stomach, liver, testicular, and cervical collagen-producing fibroblasts, they undergo a partial EMT,
malignancies (Hutchinson et al., 2015). Fibrosis typically ensues and in a paracrine manner, they contribute significantly to fibro-
through the progressive accumulation of ECM deposited by my- genesis and inflammation, two hallmarks of fibrosis (Figure 3).
ofibroblasts. The origin of these myofibroblasts has been Is There a Role for EMT in Liver and Lung Fibrosis?
debated for many years, with EMT being considered by some Evidence to clarify the role of EMT in liver fibrosis has been less
authors to drive the transformation of epithelial and endothelial forthcoming. Hepatocytes do not seem to undergo EMT in vivo,
cells into myofibroblasts (Iwano et al., 2002; Zeisberg et al., and they are not the source of collagen-producing cells in liver
2007b), whereas lineage tracing in transgenic mice indicates fibrosis (Taura et al., 2010). In fact, the transdifferentiation of he-
that the contribution of those cells to the population of myofibro- patic stellate cells are reportedly responsible for 90% of the
blasts is negligible (Humphreys et al., 2010; Li et al., 2010a; myofibroblast population in transgenic mice (Mederacke et al.,
LeBleu et al., 2013). Recent advances in fibrosis-EMT research 2013), with neither hepatocytes nor cholangiocytes (the epithe-
has helped to shed light on some of these issues. lial cells of the bile duct) seeming to undergo EMT (Chu et al.,
The EMT Debate in Renal Fibrosis May Be Explained by 2011; Scholten et al., 2010). These findings contrast with earlier
Partial Activation of the Program lineage-tracing studies showing that fibroblasts were derived
Early insight into the potential benefit of abrogating EMT in from hepatocytes that underwent EMT (Zeisberg et al.,
fibrotic conditions came from the finding that mice lacking 2007b). Others have reported that hepatic stellate cells secrete
Smad3 were protected against tubular interstitial fibrosis (TIF) collagen I to trigger EMT in epithelial hepatoma cells (Yang
(Sato et al., 2003). Smad3 acts downstream of TGFb, one of et al., 2014), that fibrosis can be repressed by inhibiting TGFb
the main EMT inducers, and indeed, Smad2/3 activation medi- signaling (Park et al., 2015), and that miRNAs can modulate
ates renal fibrosis in response to TGFb and angiotensin II EMT in cultured hepatocytes through TGFb inhibition (Brockhau-
(Yang et al., 2010a). Furthermore, reactivation of Snail and sen et al., 2015; Zhao et al., 2014; Zhao et al., 2015). Importantly,
EMT in renal epithelial cells can induce renal fibrosis and renal Snail deletion in hepatocytes attenuates liver fibrosis in adult
failure (Boutet et al., 2006). However, there is no evidence of transgenic mice (Rowe et al., 2011), also suggesting that hepa-
the migratory potential of renal epithelial cells in vivo despite their tocyte-EMT may play a role in liver fibrosis.
accumulation of mesenchymal markers when observed in cul- In the light of these studies and the new data garnered from
ture upon treatment with TGFb (Humphreys et al., 2010; LeBleu renal fibrosis (Grande et al., 2015; Lovisa et al., 2015), it is
et al., 2013). New evidence may help to understand the mecha- tempting to speculate that EMT contributes to the development
nisms behind renal fibrosis and resolve this controversy (Grande of liver fibrosis. In parallel with that observed in the kidney,

28 Cell 166, June 30, 2016


Figure 3. EMT and Fibrosis
Following one of unilateral ureteral obstruction
(UUO), folic acid treatment (FA), or nephrotoxic
serum-induced nephritis (NTN), renal epithelial
cells undergo a partial EMT characterized by the
loss of epithelial and specific differentiation
markers and accompanied by the attenuation of
their proliferative capacity. These dedifferentiated
damaged renal epithelial cells persist in the tu-
bules, relaying instructive signals to the inter-
stitium to promote myofibroblast differentiation
and the recruitment of immune cells, thereby
promoting fibrogenesis and sustaining inflamma-
tion. All these effects are mainly due to the acti-
vation of EMT-TFs, including Snail and Twist, as
their specific deletion in renal epithelial cells
strongly attenuates the development of the
hallmarks of fibrosis (adapted from Grande et al.,
2015).

STAT3 is a downstream mediator of


TGFb signaling and induces EMT in coop-
eration with active K-Ras by increasing
Snail expression in cancer cells (Saitoh
et al., 2016). STAT3 is also reported to
be elevated in lung biopsies from patients
with IPF and in mice with fibrotic lungs.
Inhibiting STAT3 with a synthetic small-
molecule inhibitor (C-188-9) attenuates
fibrosis in bleomycin-treated mice (Pe-
droza et al., 2016). Agonists of another
endogenous inhibitor of TGFb signal-
ing, the orphan nuclear receptor 4A1
(NR4A1), can also inhibit fibrosis in skin,
liver, lung, and kidney (Palumbo-Zerr
et al., 2015), and this is consistent with
the inhibition of TGFb-induced EMT and
metastasis by the expression of NR4A1
itself (Zhou et al., 2014). Others have tar-
damaged hepatocytes could secrete signals after undergoing a geted TGFb signaling in mouse models by modulating the activ-
partial EMT, which promotes stellate cell transdifferentiation ity of BMP components, including administration of BMP7 (Zeis-
and enhances inflammation. This would be consistent with berg et al., 2003) and the use of small peptides to specifically
data obtained in mouse models of fibrosis in Snail-deficient he- inhibit the BMP-Alk3 receptor interaction (Sugimoto et al., 2012).
patocytes (Rowe et al., 2011). This partial EMT hypothesis is also The prevention or reversal of fibrosis requires long-lasting sup-
supported by findings in lung fibrosis, in which a subset of pression and presumably, tissue remodeling. Epigenetic modifi-
epithelial cells in patients with IPF expresses both epithelial cations can perpetuate fibroblast activation in renal fibrosis
and mesenchymal markers (Marmai et al., 2011). (Bechtel et al., 2010) and, given that some epigenetic modifica-
Therapeutic Strategies in Fibrosis tions can be reversible, epigenetic modulators such as miRNAs
As well as being a very potent EMT inducer, aberrant TGFb are increasingly being targeted for the prevention or reversal of
signaling is the main cause of fibrosis, and it provokes a massive fibrosis, although there are still concerns about the off-target
accumulation of ECM. How or why TGFb signaling becomes or toxic side effects associated with the delivery of miRNA inhib-
constitutive is unknown (Marmai et al., 2011), but nevertheless, itors (Pottier et al., 2014). Examples of putative candidates are
attenuating TGFb signaling has been the main strategy to fight miR-21 and miR-181, linked with EMT-driven fibrosis in epicar-
fibrosis (as in colorectal cancer, in which TGFb inhibitors dial mesothelial cells (Brnnum et al., 2013), hepatic stellate
block the crosstalk between the stroma and cancer cells to cells, and hepatocytes (Zhao et al., 2014; Brockhausen et al.,
halt disease progression) (Calon et al., 2015). We describe 2015). Other miRNAs have been proposed as beneficial against
below different strategies proposed to inhibit the TGFb pathway fibrosis, after showing that they inhibit EMT: miR-200b and miR-
as a proof of concept of the benefits of targeting EMT, albeit 34c inhibit EMT in TIF (Morizane et al., 2014; Oba et al., 2010;
some targets may not fulfill the criteria for the optimal targeting Tang et al., 2013); the miR-106b-25 cluster (miR-106b, miR-93,
in patients. and miR-25) is markedly downregulated in TGFb1-induced

Cell 166, June 30, 2016 29


EMT in renal HK2 cells (Tang et al., 2014); the loss of miR-101 Despite the vast body of literature regarding the role of EMT in
promotes EMT and fibrosis in hepatocytes through TGFb activa- cancer, its applicability in cancer diagnosis and treatment has
tion (Zhao et al., 2015); and overexpression of let-7 miRNA in pri- remained limited, in part due to the intrinsic heterogeneity of
mary fibroblasts causes changes consistent with the inhibition of the actual tumor cells and tumor microenvironment when
EMT (Huleihel et al., 2014). These miRNAs could be considered compared with the relative homogeneity of cell culture models.
in anti-fibrotic therapeutic strategies that seek to restore their Thus, there are frequent debates as to whether EMT is truly rele-
function. As such, the intravenous injection of synthetic RNA du- vant to cancer in vivo. The execution of EMT in cancer is not ho-
plexes that mimic miR-29 can recover its function and even mogeneous, lending further support to the need to view EMT as a
block pulmonary fibrosis in mice (Montgomery et al., 2014). spectrum of intermediate states. Indeed, the description of the
Heat shock proteins (HSPs), molecular chaperones that assist tumor invasive front as being functionally distinct from the
in protein folding, are overexpressed in numerous cancers (Lia- main tumor bulk (Brabletz et al., 2001) points to the heteroge-
nos et al., 2015), and they have been linked with fibrosis (Bellaye neous nature of the EMT program executed within the tumor
et al., 2014). HSP27 inhibition (OGX-427) destabilizes Snail and microenvironment. The tumor invasive front, in which the EMT
inhibits TGFb-induced EMT and fibrosis (Wettstein et al., program is executed, has the hallmarks of a mesenchymal
2013). In fact, blocking HSP27 with a synthetic tetrapeptide phenotype, with a weakened cell adhesion system. By contrast,
(Ac-SDKP) or with the OGX-427 antisense oligonucleotide (an the main tumor bulk remains largely epithelial (Figure 4). Thus,
anticancer agent in phase II clinical trials) can impair EMT there is probably an EMT gradient from full, (some focal events
(Deng et al., 2014, Wettstein et al., 2013). at the tumor invasive front) to partial, to no (main tumor bulk)
Fibrosis is associated with most cardiac diseases, leading to EMT (Huang et al., 2013). Furthermore, this gradient may be
stiffness of the cardiac tissue and dysfunctional cardiac contrac- more or less steep in different tumors, reflecting the intrinsic
tion (Krenning et al., 2010). Patient survival rates are reported to molecular heterogeneity arising from diverse driver mutation
deteriorate as the proportion of fibroblasts rises (Gulati et al., profiles or altered signaling networks. For example, the circu-
2013). In the heart, a subset of endothelial cells undergo TGF- lating tumor cells (CTCs) in breast cancers with a triple-negative
b-mediated EndMT, and these fibroblasts are associated with molecular subtype (ER/PR/HER2) tend to have a more mesen-
heart fibrosis (Zeisberg et al., 2007a). In addition, recent work chymal phenotype. Therefore, the overall biology of the tumor
shows that approximately 35% of cardiac fibroblasts undergo will likely dictate the frequency of identifying CTC subpopula-
transdifferentiation into endothelial cells in a p53-dependent tions with a partial or full EMT.
manner through a mesenchymal to endothelial transition The tissue of origin may also influence the heterogeneity in
(MEndoT), contributing to neovascularization after ischemia the execution of EMT. A quantitative EMT scoring system was
(Ubil et al., 2014). Again, p53 could be a beneficial agent, and recently established based on gene expression profiles, ranking
reversion to the epithelial or endothelial phenotype should be the EMT state from 1.0 to +1.0 (Tan et al., 2014). This system
the preferred strategy in anti-fibrotic therapies. was used to highlight that the developmental lineage of each
In summary, the extensive exploration of EMT in fibrotic dis- cancer type is defined by both a micro- and macro-EMT
eases has helped identify therapeutic targets that predominantly gradient, and each cancer type has a distinct propensity to
hinder TGFb signaling (Chen et al., 2013; Wang et al., 2013). How- exhibit diverse EMT states. Expectedly, cancers that arise from
ever, inhibiting TGFb will also impede its beneficial effects (Mehal the mesoderm or NC lineages consistently show higher EMT
et al., 2011) and it may therefore be more desirable to target scores in both in vivo tumors and cultured cell lines. In these can-
its principal downstream pro-fibrotic effectors (e.g., STAT3 cers (e.g., osteosarcoma and malignant melanoma), the range of
or NR4A1), or alternatively, restore the expression of epithelial intrinsic EMT gradients is more restricted than in solid tumors of
miRNAs. These potential therapeutic targets can ameliorate epithelial origin, such as ovarian, breast, and lung carcinomas.
fibrosis by attenuating EMT and EndMT, offering an opportunity These different EMT gradients might help explain the inconclu-
to reverse fibrosis via MET. As such, inactivating Snail1 in mice sive clinical significance of EMT found in the literature. Given
with established renal fibrosis improves organ morphology and the wide range of EMT states in ovarian, breast, and lung carci-
function, significantly attenuating the disease (Grande et al., 2015). nomas, their clinical significance should be considered in
conjunction with their intrinsic molecular profiles, such as the
EMT in Cancer gene expression-based molecular subtype (GEMS) for ovarian
EMT is thought to be activated in cancer cells, linked to their and breast cancers and the EGFR mutational status or ALK
dissociation from the primary tumor and their intravasation into fusion transcript status for lung cancer. In ovarian cancer, the
blood vessels (Thiery et al., 2009). However, the impact of the EMT status is closely linked to the reported GEMS, with worse
EMT in cancer progression and patient survival is still far from GEMS prognosis correlated with higher EMT scores (Tan et al.,
fully understood. From skepticism (Tarin et al., 2005) to funda- 2013). Furthermore, the EMT status consistently predicts over-
mental support (Lamouille et al., 2014; Thiery et al., 2009; Ye all- (OS) and disease-free survival (DFS) (Tan et al., 2014). Collec-
and Weinberg, 2015), more recent data suggest notes of caution tively, these findings offer a new perspective on the translational
on the true role of EMT in cancer progression (Fischer et al., applicability of EMT in prognosis.
2015; Zheng et al., 2015). Nevertheless, studies in the last In the absence of cell lineage analyses, the expression of
decade have fueled the interest in EMT in the cancer field, mak- epithelial and mesenchymal markers in primary tumors is a
ing it necessary to discuss recent advances regarding this com- good indicator of a potential EMT. The existence of intermediate
plex biological process. phenotypes should also allow cells in the stromal compartment

30 Cell 166, June 30, 2016


Figure 4. The Metastatic Cascade
This scheme aims to illustrate the complexity of the metastatic cascade from the point of view of cell plasticity, from carcinoma cell heterogeneity in the primary
tumor to the different strategies that cancer cells adopt to survive in the bloodstream and colonize distant sites. EMT is a limited focal event in the primary tumor
that could occur upon the interaction of carcinoma cells with the microenvironment, in which cancer-associated fibroblasts (CAFs) and immune cells can be
found, including different types of natural killer T and B lymphocytes and inflammatory cells (not shown). The stroma also includes tumor-associated macro-
phages (TAMs) and other bone-marrow-derived cells. Some cells are possibly recruited prior to the establishment of primary and metastatic malignancies
(Barcellos-Hoff et al., 2013). CAFs are like myofibroblasts (Lau et al., 2016) and TAMs are M2-type macrophages (Hsu et al., 2014), indicating that the micro-
environment in the primary tumor is similar to that found in fibrosis, including an important inflammatory component. Other stromal cells also play a major role in
tumor progression (Gentles et al., 2015). As expected from a focal event, the proportion of carcinoma cells with EMT features in primary breast tumors does not
exceed 3% in estrogen receptor (ER)-positive tumors and 10% to 15% in ER-negative tumors. However, the majority of circulating tumor cells (CTCs) exhibits an
intermediate EMT phenotype. Cells that undergo EMT in the primary tumor can also help epithelial cancer cells to invade. CTCs may be derived from carcinoma
cells that undergo EMT in situ in the primary tumor or they may acquire intermediate EMT phenotypes in circulation, particularly when in clusters exposed to high
concentrations of TGFb originating from associated platelets. Recent data suggest that microemboli may be able to extravasate and that they exhibit higher
clonogenic potential (Aceto et al., 2014; Au et al., 2016). Cancer cells in metastatic outgrowths are clearly epithelial-like, and they can be identified morpho-
logically and molecularly as having been derived from the primary tumor, again highlighting the plasticity of tumor cells. This is also consistent with data showing
that mesenchymal cells must revert to the epithelial phenotype to colonize their new destination. The plasticity of carcinoma cells is therefore of upmost
importance to escape death during the different stages of tumor progression. Mesenchymal-like cells are better adapted to cell deformation, resisting shear
stress and being more resistant to drugs. Tumor dissemination in the lymph nodes is much less documented, albeit it is a major aspect of tumor staging (not
shown). Histopathological examination of proximal (sentinel) lymph nodes show either dispersed carcinoma single cells, micrometastases (<2 mm), or massive
invasion. The latter is compatible with findings indicating that lymph nodes can be colonized after collective cell migration (Giampieri et al., 2009).

(which maintain the epithelial markers of their tissue of origin) to HGF (Bai et al., 2015a; Bai et al., 2015b). Nevertheless, it is
be identified. The distribution of a number of epithelial and crucial to consider the nature of CTCs as indicators of the next
mesenchymal markers was first established in a breast cancer step in the metastatic cascade.
tissue microarray (Sarrio et al., 2008). More recently, mesen- EMT and Circulating Tumor Cells
chymal markers were found in 12% of the basal molecular sub- Most CTCs express both epithelial and mesenchymal markers
types of breast carcinoma. This percentage is likely to be higher (Khoo et al., 2015; Yu et al., 2013), suggesting that an active
in the claudin-low molecular subtype. Furthermore, cells ex- EMT process is likely to operate during the dissemination of car-
pressing mesenchymal markers are also likely to exist among cinoma cells (Figure 4) (Thiery and Lim, 2013). In addition, after
epithelial carcinoma cells in ErbB2 and luminal molecular transient amplification of breast cancer CTCs, the cells exhibit
subtypes; albeit, to a lesser extent (Tan et al., 2014; Yu et al., a full range of EMT phenotypes (Khoo et al., 2015), with evidence
2013). It is also interesting to note that, in cancer, as in embryonic indicating that CTCs express ECM components potentially
development, EMT may be a focal rather than a global event, involved in the successful colonization of distant organs (Ting
probably a response of cancer cells to their local microenviron- et al., 2014).
ment (Figure 4). Macrophages participate in EMT induction A landmark study revealed that CTCs in blood from metastatic
within primary tumors in transgenic mouse models, orthotopic breast cancer patients can serve to evaluate prognosis and
xenografts (Wyckoff et al., 2004), and primary human breast car- response to treatment (Cristofanilli et al., 2004), and this was
cinomas, interacting with carcinoma cells adjacent to the endo- recently confirmed when CTCs quantified in metastatic patients
thelial cells (Robinson et al., 2009). The latter suggests that were shown to predict OS and progression-free survival (Bidard
macrophages may be also critical for local intravasation. M2a et al., 2014). However, a recent clinical trial could not identify an
macrophages migrate toward carcinoma aggregates to induce effective second-line chemotherapy strategy for patients whose
the dissociation and migration of single carcinoma cells through CTC numbers were not diminished by first-line chemotherapy
the local production of TGFb and direct ICAM-1-b2 integrin- (Smerage et al., 2014). Clearly, there is an urgent need to under-
mediated cell-cell contact. Endothelial cells also contribute to stand more about the molecular attributes of CTCs, particularly
their dissociation through diffusible scatter factors, such as how their presence and EMT status reflect treatment response

Cell 166, June 30, 2016 31


and malignant potential. Several studies have addressed the data are consistent with the lack of metastatic outgrowth from
value of CTC monitoring and phenotyping during neoadjuvant cells with constitutive mesenchymal traits, and it highlights the
and adjuvant therapies (Alix-Panabie`res and Pantel, 2014; transient nature of EMT and the cells need to revert to a more
Khoo et al., 2015) and their use in drug sensitivity testing (Yu epithelial phenotype for colonization (Ocana et al., 2012; Tsai
et al., 2013). Importantly, cancer therapy affects CTC number et al., 2012).
and phenotype, with refractory patients having more mesen- To better understand the contribution of EMT to metastatic
chymal-like CTCs and patients who are responding to treatment colonization, Fsp-1 or vimentin GFP-reporter lines were used
demonstrating significantly fewer CTCs with a more epithelial- to monitor the expression of mesenchymal markers in a
like phenotype (Yu et al., 2013). These findings are in agreement MMTV-PyMT breast cancer model (Fischer et al., 2015). While
with recent studies highlighting the importance of EMT in confer- this provided evidence of EMT in a small proportion of cells in
ring chemoresistance in breast and pancreatic cancer models the primary tumor, these GFP-positive cellseven though pro-
(Fischer et al., 2015; Zheng et al., 2015). portionally enriched in the CTC populationdid not seem to
Intriguingly, the EMT program may be engaged at premalig- contribute significantly to the formation of metastatic out-
nant stages, as shown in an experimental model of pancreatic growths. Others show that pancreas-specific loss of either Snail
cancer, especially at sites of inflammation (Rhim et al., 2012). or Twist does not block systemic dissemination or the formation
This suggests that EMT may be a very early event in tumorigen- of metastasis in pancreatic ductal adenoma carcinoma (PDAC)
esis and is consistent with the presence of disseminated tumor (Zheng et al., 2015). Together, these findings prompted the au-
cells (DTCs) at very early tumor stages in breast cancer patients thors to suggest that EMT may be dispensable for carcinoma
(Bidard et al., 2008; Husemann et al., 2008). CTCs are mainly progression to a metastatic state.
identified with the EpCAM antibody, which may exclude cells Although these data are provocative, EMT cannot yet be ruled
that have undergone a more complete EMT and have lost this out as a driver of cancer progression (Maheswaran and Haber,
epithelial marker. Thus, lineage tracing experiments in animal 2015). First, the cooperation between the multiple EMT-TFs sug-
models are required to identify CTCs in different states, and gests that loss of individual factors may be insufficient to prevent
this may also offer a way to explore whether CTCs only undergo cell invasion and dissemination, particularly in PDAC, in which
transitions after reaching the metastatic site or if MET can also even normal cells do not exhibit a strong epithelial phenotype.
occur in the bloodstream. Interestingly, circulating tumor Furthermore, Fsp-1 is not a universal marker of EMT and it would
clusters are more effective in colonizing secondary sites than not be activated in the primary tumors of Neu and PyMT breast
single mesenchymal CTCs (Aceto et al., 2014). These clusters cancer models (Trimboli et al., 2008), albeit EMT has been
comprise distinct clonal carcinoma cells which are held together described in both (Ye et al., 2015). Second, the few cells that un-
through plakoglobin-mediated intercellular adhesion. Again, this derwent EMT in the primary tumor (Fischer et al., 2015) are
highlights the requirement for mesenchymal cancer cells to at consistent with the numbers observed in other studies (Tan
least partially reverse to the epithelial state for metastatic growth et al., 2014). Third, the relative enrichment of mesenchymal
(Nieto, 2013). CTC clusters formed through platelet interactions markers in CTCs confirms their enhanced ability to invade and
are also correlated with a worse prognosis, and mesenchymal intravasate. Fourth, the comparatively low contribution of the
CTCs may interact with epithelial-like carcinoma cells and vimentin-expressing cells to metastases calls for an analysis of
drag them along to secondary sites (Figure 4). It was unclear other markers that may better assess intermediate EMT pheno-
how such aggregates can reach secondary sites through the types and for the evaluation of a possible cooperation between
510-mm lumens of blood capillaries in the lung, but recent mesenchymal and epithelial cells (Tsuji et al., 2009). Indeed,
data show that clusters of %20 cells can traverse these constric- EMT may facilitate the invasion and intravasation of other cells
tions (Au et al., 2016). Extravasation is a very efficient process that retain their epithelial character (Figure 4). As such, even if
and even untransformed epithelial mammary cells can reach EMT were more limited than anticipated, it would still be required
the lung after being injected into the bloodstream (Podsypanina for tumor progression to the metastatic state (Li and Kang, 2016).
et al., 2008). If epithelial cells, and cell clusters, can efficiently A recent study using intravital microscopy in mice has detected a
disseminate and extravasate, the EMT/MET pathway may pool of breast tumor cells that spontaneously undergo EMT,
not be necessary in all cells for secondary tumor formation. become motile, and disseminate and then reverse to the epithe-
This directly impinges on the requirement for EMT in tumor lial state upon metastatic outgrowth (Beerling et al., 2016). More
progression. studies like these are needed in other types of cancer in order to
Is EMT an Essential Element of the Metastatic Cascade? assess the contribution of EMT as a crucial requirement for the
Most cancer EMT studies, including mechanistic analyses, are progression of primary tumors toward the metastatic state, as
conducted on cultured cells, and the vast majority of in vivo consistently suggested in different models and patient datasets.
studies are carried out after injection or xenografting of parental EMT and Cancer Stem Cells
or manipulated cancer cells, which are likely to show distinct The induction of both EMT and stemness at the invasive fronts
EMT features. Some mouse breast cancer reporter models of tumors was first suggested to explain how CSCs disseminate
have been used to follow the endogenous activation of Snail and seed fully heterogeneous tumors at secondary sites (Bra-
TFs (Tran et al., 2014; Ye et al., 2015), showing that Snail1 acti- bletz et al., 2005). Indeed, the EMT-derived stem cell phenotype
vation and endogenous EMT occurs in the primary tumor, and in the mammary glandcapable of forming mammospheres
that, although important for invasion and the formation of following EMT induction and expressing a CD44highCD24low pro-
CTCs, EMT is not required for metastatic colonization. These file (Mani et al., 2008; Morel et al., 2008)was reminiscent of that

32 Cell 166, June 30, 2016


of CSCs. This profile has since been identified in numerous CSC plasticity in highly metastatic neuroblastoma cells, cells that
subpopulations (Gupta et al., 2009a), with other groups reporting strongly express EMT genes and that display a stem-like pheno-
high intracellular aldehyde dehydrogenase 1 (ALDH1) activity as type, with high SOX2 and NANOG expression (Pandian et al.,
a marker of stemness (Ginestier et al., 2007). Recent indications 2015). Collectively, these studies highlight the regulatory effects
submit that CD44high and ALDH1 activity do not generally coexist of the microenvironment on cell fate.
in CSCs (Liu et al., 2013a), suggesting the existence of different Uncoupling EMT and Stemness
CSC types. The gene signature identified for the induction of Despite the evidence linking stemness and the EMT, recent
pluripotency is also relevant to the persistence of CSCs, with studies consider that these two aspects of CSC development
Oct4 and Nanog ectopically inducing CSC-like properties may occur in parallel rather than through the activation of the
and enhanced features of EMT (Kong et al., 2010). This is similar same pathway. It is now widely accepted that, although EMT is
to other genes involved in maintaining self-renewal capacity, instigated in CSCs to drive their dissemination, they undergo
including b-catenin, Bmi1, Gli1, and Pou5f1. Within the MET on reaching a desirable metastatic niche to seed a new
CD44highCD24low population of triple-negative breast cancers, tumor, presumably by suppressing driver EMT-TFs (Nieto,
there are discrete epithelial and mesenchymal cell types that 2013). The MET at the new metastatic site is surely influenced
are differentiated by their expression of a6-integrin splice vari- by CAFs and a specific type of metastasis-associated macro-
antsa6A (epithelial) and a6B (mesenchymal)with the a6B- phages (MAMs) (Qian et al., 2015), as well as mesenchymal
b1-integrin noted as essential for CSC function, mammosphere stem cells, immunological mediators, and other factors involved
formation, and anchorage-independent growth (Goel et al., in hypoxia and angiogenesis (Plaks et al., 2015). As such, recent
2014). Both a6-integrin and CD44 are alternatively spliced by data indicate that MAMs promote the conversion of hepatic stel-
ESRPs, which as already discussed, protect the epithelial late cells into myofibroblasts, resulting in a fibrotic microenviron-
phenotype and, hence, promote the presence of epithelial a6 ment that sustains metastatic tumor growth in the liver (Nielsen
and CD44 isoforms in these cells (Goel et al., 2014; Warzecha et al., 2016). Mesenchymal DTCs activate CAFs, which in turn,
et al., 2009). Accordingly, it would be interesting to correlate promote the transition of DTCs to a more epithelial and prolifer-
the combinations of integrins and CD44 isoforms with ALDH1 ative phenotype that favors metastatic colonization (Del Pozo
activity in primary tumors and their metastatic outgrowths. Martin et al., 2015).
EMT involves the expression of one or several classical EMT- Parallels can again be seen between the induction of tumor-
TFs, and since the expression of these TFs often accompanies initiating capacities in cancer cells and the acquisition of plurip-
features of stemness, the retention of this latter state has been otency during fibroblast reprogramming to induced pluripotent
linked to EMT. ZEB1 has been specifically detected in poorly stem cells. The latter also requires a MET (Li et al., 2010b; Sama-
differentiated pancreatic carcinomas, and it is thought to repress varchi-Tehrani et al., 2010), compatible with the fact that embry-
the expression of stemness-inhibiting miRNAs to maintain a onic stem cells are epithelial-like. Recent data indicate that the
stem-like phenotype (Wellner et al., 2009). CSC expansion is serial delivery of four transcription factors in a specific order
also thought to occur after Snail1 mediates a shift from asym- Oct-4 with Klf4, followed by c-Myc, followed by Sox2 (Liu
metric (one stem cell, one differentiating cell) to symmetric et al., 2013b)induces a higher rate of pluripotency. Oct4 first
(two stem cells) cell division, implying a role for EMT in increasing promotes fibroblasts with an even more mesenchymal-like
the stem cell pool within the tumor (Hwang et al., 2014). Similarly, phenotype through the upregulation of Snail (Oct4), such that
other inducers of EMT, such as the homeobox protein Six1 or these cells are then more responsive to Sox2, which is needed
ladybird homeobox-1 protein (LBX1), also induce stemness to repress Snail and induce MET in embryos and mouse fibro-
and mammary spheroidogenesis when ectopically expressed blasts (Acloque et al., 2011; Li et al., 2010b). This delay in MET
in the mouse mammary gland, presumably through their effects is thought to provide time for epigenetic reorganization before
on one of their transcriptional targets (ZEB1, ZEB2, SNAIL1, and pluripotency is induced (Gaeta et al., 2013). Interestingly, cells
TGFb2) (McCoy et al., 2009; Yu et al., 2009). Nonetheless, it is engaged in intermediate reprogramming closely resemble
noteworthy that the EMT program implicated in normal stem those found during mesendoderm formation, which is driven
cells and CSCs may differ, as recently shown in the murine mam- by a MET-like process in mesenchymal mesodermal cells in
mary gland, in which they involve Snail2 and Snail1, respectively the primitive steak (Takahashi et al., 2014). Again this highlights
(Ye et al., 2015). the similarities with embryonic development and the importance
Much of the plasticity associated with CSCs arises from the of intermediate phenotypes in the induction of EMT. A prototyp-
local niche, both at the primary and secondary tumor sites (Nieto, ical example of this is the loss of an epigenetic mark in tropho-
2013), and the presumed CSC reservoir is thought to be respon- blast epithelial cells that allows them to undergo a partial EMT
sible for tumor relapse and poor clinical outcome. TGFb induces while maintaining their self-renewal capacity and multipotency
a program in stromal cells that correlates with poor prognosis (Abell et al., 2011).
in colorectal cancer patients, in which cancer-associated fibro- Further evidence for uncoupling EMT and stemness derives
blasts (CAFs) have been linked with an increase in tumor-initi- from studies of the homeobox transcription factor, PRRX1, a
ating cells (Calon et al., 2015). Similarly, in hepatocellular carci- potent EMT inducer in both embryos and cancer cells. In human
noma that usually develops in the context of fibrosis, CAFs breast cancer BT-549 cells, downregulation of PRRX1 is linked
with characteristics of myofibroblasts promote the appearance with MET and increased proliferation, but also with a gain in
of liver-tumor-initiating cells (Lau et al., 2016). Ongoing molecu- mammosphere formation and the emergence of a predominant
lar rearrangements were recently proposed to drive adaptive CD44high cell population (Ocana et al., 2012), all favoring

Cell 166, June 30, 2016 33


metastatic colonization. This again suggests that EMT is not chymal genes without affecting the expression of the epithelial
necessarily associated with stemness. Interestingly, CTGF and genes (Drasin et al., 2015).
the inhibitor of differentiation (Id) both induce MET and the Hybrid, partial transitions in cells lead to an expression of
expression of pluripotency genes (Chang et al., 2013; Stankic both epithelial and mesenchymal traits, which can potentially
et al., 2013), which suggest that the morphological phenotype endow cells with migratory potential while retaining some degree
and stemness can be regulated independently. Indeed, TGFb of cell-cell adhesion, promoting a more efficient coordinated
is essential for the initial EMT, after which the EGFR-Ras migration, as observed during embryonic development (Theve-
pathway independently drives stem cell properties (Voon et al., neau and Mayor, 2011). Stemness is regulated by a LIN28/
2013). This is also consistent with Snail- or Twist-induced let-7 inhibitory loop (Yang et al., 2010b), with a balanced, inter-
expression of miR-424, which promotes mesenchymal traits mediate level of the two helping to maintain an intermediate level
while inhibiting tumor initiation (Drasin et al., 2015). In a more of Oct4 for stemness (Niwa et al., 2000). Building on this, a
extreme case, the mutual exclusivity of EMT and stem cell-like stemness window that regulates the interplay between
traits is evident in prostate and bladder cancer cell lines, in which LIN28 and Let-7 has been proposed to control stemness and
constitutive Snail1 activation suppresses stemness (Celia`-Ter- EMT (Jolly et al., 2015). Through the use of theoretical framework
rassa et al., 2012). Furthermore, in the skin, low Twist levels modeling, it was shown that strong inhibition of LIN28 by miR-
induce tumor initiation and confer stem cell properties while 200 pushes the stemness window toward the mesenchymal
higher levels are required for the induction of EMT (Beck et al., end of the EMT axis, whereas strong inhibition of ZEB by Let-7
2015). Consistently, Twist-induced tumor-initiating capacity is forces it in the other direction. From this, it was suggested that
only manifested after Twist downregulation at the metastatic stemness can be gained by cells with different phenotypes
site (Schmidt et al., 2015), and decreasing Snail1, Twist1, or based on the actions of Let-7 and Lin28 but also in response
Prrx1 expression is a requisite for EMT attenuation to allow to changes in other signals like Snail1 and, presumably, Twist
CSCs to populate a new metastatic site (Ocana et al., 2012; or other prominent inducers of EMT.
Tran et al., 2014; Tsai et al., 2012). A sub-population of metastasis stem cells was recently
Can Partial EMT States Explain the Stemness-EMT-MET described among the CTC population (Baccelli et al., 2013),
Link? although it remains unclear whether CTCs harbor a small fraction
The aforementioned data show that EMT and stemness are not of CSCs or if CTCs derive from cells with the potential to become
necessarily linked and that the acquisition of both traits can be CSCs, themselves a separate entity. Resistance to treatment
uncoupled. However, they also suggest that an intermediate may be acquired at metastatic sites in micrometastatic stages.
EMT state may make cells more prone to exhibiting CSC-like Therefore, it will be necessary to determine whether stemness
properties. Transient Twist1 activation in mammary epithelial is a property of subsets of CTCs or whether EMT is solely a
cells induces their long-term invasive potential, as well as their mechanism critical for dissemination, with stemness re-acquired
capacity to colonize distant organs and tendency to express at the site of arrest in the parenchyma of the target organ through
traits typical of both epithelial and mesenchymal cells. By MET (Beck et al., 2015; Brabletz, 2012; Jung and Yang, 2015;
contrast, constitutive Twist1 activation induces a migratory but Polyak and Weinberg, 2009). Stemness, or the attainment of
not a metastatic phenotype, as cells do not demonstrate tu- stem cell properties, is no longer considered a fixed trait,
mor-initiating capacities (Schmidt et al., 2015). This is consistent with such cells displaying a high degree of plasticity (Jolly
with the finding that a complete EMT is not required for Twist1- et al., 2015). Stemness is well defined in embryonic development
induced invasion (Shamir et al., 2014) and that plasticity at the and it is a hallmark of cancer within a particular window during
metastatic site does not imply reversion to the fully differentiated disease progression. Like other hallmarks of cancer, stemness
epithelial phenotype (Nieto, 2013). In summary, stemness seems can be lost or acquired, and it may be affected by the expression
to be better manifested at intermediate epitheloid states, consis- of certain phenotypic traits as cancer cells and tumors progress
tent with those observed in cultured mammospheres (Mani et al., toward the metastatic disease (Antoniou et al., 2013).
2008; Ocana et al., 2012) and in embryonic stem cells.
The intermediate EMT state associated with the stem cell The Spectrum of EMT Plasticity in Cancer: Therapeutic
properties of trophoblast cells (Abell et al., 2011) is similar to Challenges
that observed in claudin-low breast cancer cells (Jordan et al., Re-differentiation represents an attractive therapeutic option to
2011; Prat et al., 2010). This is also the basis of the proposed ex- reverse EMT (Beug, 2009; Huang et al., 2012; Nieto, 2013; Thiery
istence of migratory CSCs at the invasive front of tumors (Bra- and Sleeman, 2006), a feature that provides a unique way to tap
bletz et al., 2005). Indeed, transitions between EMT and MET are into current mainstream therapies (Chua et al., 2011). However,
likely to be controlled by many factors that sway the balance there are several concerns that must first be considered. Given
from one extreme to the other. Thus, a tissue could display the existence of EMT gradients, we have yet to determine how
more mesenchymal or epithelial characteristics at any given far the cells should be reversed. Furthermore, it remains unclear
time, yet still with evidence of the other. Indeed, pancreatic can- what specific EMT score would be effective to abolish metas-
cer cells are shown to have stem-like features and an intermedi- tasis and/or enhance drug sensitivity. Evidence indicates that a
ate EMT phenotype in vivo, with low levels of E-cadherin and full EMT reversal might not be optimal for metastatic coloniza-
concomitant mesenchymal features (Rhim et al., 2012). Along tion, and thus, the effective shift is likely to be context dependent
similar lines, the upregulation of miR-424 during Twist1- or and based on the different patterns of EMT in distinct cancer
Snai1-induced EMT in breast cancer cells can induce mesen- types. Indeed, the re-expression of E-cadherin in fibroblasts

34 Cell 166, June 30, 2016


does not reverse them to the epithelial state (Navarro et al., immune cells. Finding a way to eradicate these epithelial cells
1993), and while restoring E-cadherin to ovarian cancer cells in would identify a mechanism through which the lethality of the
culture abolishes their invasive properties, their mesenchymal cells can be enhanced following EMT reversal. One example
traits remain unaffected (Huang et al., 2013). would be the use of epigenetic modifiers, such as histone de-
The drug discovery platform for EMT reversal is limited, acetylase (HDAC) inhibitors, to reverse ZEB1-regulated EMT
although it is feasible to conceive drug pipelines that preferen- and sensitize pancreatic cancers to gemcitabine (Meidhof
tially induce cytotoxicity in cells that underwent EMT. Lead com- et al., 2015). This strategy could be applied in the front-line or
pounds to target CSCs in breast cancers have been identified in chemoresistance settings.
in screening programs based on EMT models, such as human Finally, anti-EMT therapeutics should be carefully defined and
mammary epithelial cells bearing shRNA against E-cadherin positioned. Should we reverse or kill the cells that underwent
(Carmody et al., 2012; Gupta et al., 2009b). There are also pipe- EMT? Are we comfortable with reversing EMT and maintaining
lines for anti-cancer treatments that target specific components a stable disease status with these cells regaining an epithelial-
of signaling pathways, including HGFR, insulin-like growth fac- like phenotype since they are less migratory and invasive? An
tor-1 receptor (IGF-1R), EGFR, PDGFR, TGFbR, phosphatidyli- important note of caution here comes from the evidence that
nositol 3-kinase (PI3K)/Akt, and ERK/MAPK (Lamouille et al., reversal is also required for metastatic colonization in vivo (Alder-
2014). While these treatments could be considered as anti- ton, 2013; Brabletz, 2012; Ocana et al., 2012; Tsai et al., 2012;
EMT drugs, these pipelines were not designed with a specific Beerling et al., 2016)). Therefore, promoting EMT reversal in-
anti-EMT concept in mind, and therefore, the anti-proliferative cludes the potential risk of enhancing the establishment of sec-
and EMT-specific effects of these inhibitors might be difficult ondary metastasis by DTCs (Nieto, 2013). Thus, it will be imper-
to assess independently. Few platforms have been developed ative to either define a proper therapeutic window in which to
that screen for non-cytotoxic agents that reverse EMT. A administer an EMT-reversing agent, or alternatively, the goal
screening model based on a growth-factor-induced scatter should be to eradicate the cells that underwent EMT. If effective,
assay (including HGF, IGF-1, and EGF) in NBT-2 rat bladder car- specifically eliminating those cells appears to be the optimal
cinoma cells has been adapted for this purpose (Chua et al., strategy. Even in the absence of metastatic disease, most locally
2012). Its use highlighted compounds that target activin A recep- advanced patients still relapse, and it is known that recurrence is
tor type II-like kinase (ALK5)/TGFbR1, MAPK, Src, and PI3K, accompanied by EMT, with Snail1 proving sufficient to promote
none of which exhibited significant effects against proliferation. mammary tumor recurrence in vivo (Moody et al., 2005). Interest-
However, this model requires additional validation in a second- ingly, specifically targeting cells that underwent EMT might also
ary screen on human carcinomas. Recent data indicate that target dormant cells in distant organs, as they are likely to be in a
activation of the PKA pathway induces epigenetic changes mesenchymal state. As such, NR2F1 has recently been consid-
that can modulate EMT together with the loss of tumor-initiating ered a critical node in the induction of dormancy in DTCs (Sosa
capacities in a xenograft breast cancer model (Pattabiraman et al., 2015). Interestingly, NR2F1 is known to lie at the core of the
et al., 2016). EMT regulatory chromatin landscape in the neural crest (Rada-
TGFb inhibitors are the most intensively investigated anti- Iglesias et al., 2012).
EMT compounds, and recent Phase I studies tested the use of
LY2157299 as a TGFb inhibitor in glioblastoma and hepatocellu- EMT Confers Resistance to Chemotherapy and
lar carcinoma (Giannelli et al., 2014; Rodon et al., 2015). How- Immunotherapy
ever, the effect of LY2157299 on EMT reversal warrants further EMT confers resistance to cell death induced through various
study. It is worth noting here that TGFb can also act as a tumor means both in embryos and in cancer cells (Vega et al., 2004;
suppressor through the activation of a particular type of lethal Thiery et al., 2009), including chemotherapy (Singh and Settle-
EMT in pancreatic cancer cells (David et al., 2016). Src kinase in- man, 2010). Even in studies that find a limited contribution of
hibitors also clearly reverse EMT (Huang et al., 2013; Vultur et al., cells that have undergone EMT to established metastases, the
2008), yet their clinical effects as a monotherapy or in combina- role of EMT in conferring chemoresistance is clear (Fischer
tion with other agents have been disappointing (Kim et al., 2009; et al., 2015; Zheng et al., 2015). Therapeutic strategies should
Puls et al., 2011). An inhibitor of the closely related focal adhe- therefore encompass an arm that targets network rewiring
sion kinase (FAK), PF-00562271, is also being tested in a when resistance emerges. For instance, a rewiring event occurs
Phase I dose-escalation trial against solid tumors with promising in EGFR-mutated non-small-cell lung carcinoma (NSCLC) that
results (Infante et al., 2012). These Src and FAK inhibitors must switches EGFR to Axl receptor tyrosine kinase in association
undergo lethality screens to determine the appropriate combi- with a mesenchymal phenotype (Byers et al., 2013). One can
nations of the compounds to induce EMT reversal. These drug therefore propose targeting these erlotinib-resistant mesen-
screening platforms suffer from the same lack of microenviron- chymal NSCLC cells with potent inhibitors of Axl (Sheridan,
mental modeling as cell culture models. However, microfluidic 2013), although it might also be envisaged that resistance to
co-culture systems have recently been developed to mimic Axl will be acquired when network rewiring repeats. Neverthe-
the tumor microenvironment and induce the formation of less, there is increased interest in targeting Axl signaling (Gjer-
3-dimensional tumor spheroids in order to screen for EMT- drum et al., 2010) as a first-in-line class of therapeutic agent
blocking agents (Aref et al., 2013). This type of model more for EMT (clinicaltrials.gov; NCT02488408).
accurately assesses the efficacy of EMT reversal agents to Immunotherapy for cancer treatment has generated much in-
combat the appearance of sprouting endothelial cells and/or terest in recent years, with notable successes against several

Cell 166, June 30, 2016 35


tumors, including melanoma, NSCLCs, and renal carcinoma. Challenges, Future Directions, and Perspectives
These approaches are based on the blockade of immune inhib- Originally described as a major event in embryonic morphogen-
itory checkpoints, such as the use of monoclonal antibodies to esis, EMT is successfully recapitulated in various normal and
programmed death-1 (PD1) or to cytotoxic T lymphocyte-asso- abnormal processes, a phenomenon that has become well
ciated antigen 4 (CTL4), two inhibitory T cell receptors (Okazaki accepted since it was first proposed in the early 2000s. Yet,
et al., 2013). Further improvement in these therapeutic interven- despite extensive forays into the role of EMT in disease progres-
tions would permit longer survival than that achieved using tar- sion, we are still far from reaching a consensus as to the true
geted therapeutics. Numerous somatic mutationssome of role EMT plays in these processes. While is seems clear that
which are considered to be trunk and branch drivershave reversing EMT is beneficial for the treatment of fibrosis, the
been described in many tumors. However, new, dominating mu- inherent complexity of cancer suggests that a better strategy
tations are often induced by conventional and targeted thera- may be to eradicate the cells that underwent EMT in the carci-
peutics, and they have the potential to make cells refractory to noma. However, this still requires the development of better
these treatments (McGranahan and Swanton, 2015). In principle, animal models together with technological advances to better
although such rapid drift in mutation profiles may not affect CTL visualize the EMT in vivo at the organ, tissue, cellular, and sub-
cancer cell lysis, EMT of the target cancer cell may potentially cellular levels including advanced genomics in FACS-enriched
abrogate CTL lysis. cell populations. These approaches will directly test some of
Melanoma cells that undergo Snail-mediated EMT are much the concepts discussed here and help comprehend the dy-
more metastatic than their parental cells, which was attributed namics of EMT to better design anticancer therapies.
to the emergence of T regulatory (Treg) CD4 cells expressing An emerging challenge will be to decipher inter- and intratu-
Foxp3, a known inducer of immunosuppression in Treg cells moral heterogeneity, including the full spectrum of intermediate
(Kudo-Saito et al., 2009). The emergence of these Treg cells EMT states and how cells make the transitions between these
was in part driven by the TGFb and thrombospondin secreted states. Targeted therapeutics face the formidable challenge
by these Snail1-expressing melanoma cells. Furthermore, a of rapid tumor evolution, as revealed by the genomic land-
Snail small interfering RNA (siRNA) injected in vivo reduced the scape of distinct biopsies in primary and metastatic tumors
immunosuppressive and metastatic potential of melanoma cells (McGranahan and Swanton, 2015). It is also important to eval-
(Kudo-Saito et al., 2009). uate to what extent stochastic events could generate cells with
Plasticity of carcinoma cell phenotypes can be achieved by a stem-cell-like phenotype and how the cell of origin influences
conventional and targeted therapies and through the tumor the susceptibility to undergo EMT. Another important issue is
microenvironment (Holzel et al., 2013). Thus, as well as abro- to better understand the mechanisms for lymphatic dissemina-
gating immune suppression in CTLs, one might consider modi- tion in different cancer types. Although the EMT does not seem
fying the EMT phenotype of carcinoma cells. Indeed, CTL lysis to be required for lymph node colonization (Giampieri et al.,
of carcinoma cells was considerably reduced when Snail-medi- 2009), its relative contribution versus collective cell migration
ated EMT of MCF7 cells occurred in response to chronic expo- deserves further investigation.
sure to TNFa due to the induction of autophagy (Akalay et al., Bioinformatics algorithms must be developed to go far beyond
2013). EMT of the target cells affects the maturation of the gene ontology, principal component analyses, and the currently
immune synapse, although defining the mechanism that drives available networks. Most importantly, the predictions from math-
this plasticity in the cell cortex requires further study. It must ematical models and hypothetical gene regulatory networks
also be ascertained what causes the defects in the organization need to be tested in appropriate animal models. Furthermore,
and positioning of T cells, and of the adhesive receptors ar- powerful bioinformatics analysis can also help in the new ave-
ranged as concentric rings in the immune synapse (Chouaib nues that link EMT with metabolic changes (Choudhary et al.,
et al., 2014). Intriguingly, tumors that respond best to CTL-A4, 2016; Jiang et al., 2015; Mathow et al., 2015; Shaul et al., 2014).
and more recently to PD1 and PD1L antibodies, exhibit a higher Intravasation has been the least studied process in the meta-
EMT score (Tan et al., 2014). These tumors include melanomas static cascade, and in light of recent data that suggests that EMT
and renal, bladder, and NSCL carcinomas (Topalian et al., may be more limited than previously anticipated (Fischer et al.,
2012; Motzer et al., 2015; Rizvi et al., 2015). Thus, these anti- 2015; Zheng et al., 2015), attention should be paid to the possi-
bodies could specifically restore a functional immune synapse bility of cancer epithelial cells being able to efficiently intrava-
in carcinomas that exhibit a mesenchymal- rather than an epithe- sate. Do just a few mesenchymal cells help epithelial cancer cells
lial-like phenotype (Engl et al., 2015). This proposal is supported intravasate? Is the leakiness of the tumor vessels sufficient to
by recent data indicating that an increase in target cell tension allow intravasation of epithelial cell clusters? As intravasation is
potentiates killing by CTLs (Basu et al., 2016). clearly a crucial step in dissemination, further attention should
In addition to immunotherapy, a vaccine was recently devel- be now paid to this process, particularly probing the recent pro-
oped against a mesenchymal-associated T-box TF, brachyury posal that TAMs and vascular mimicry acquired by cancinoma
(Hamilton et al., 2013), which is currently undergoing clinical tri- cells help them to intravasate (Harney et al., 2015; Wagenblast
als for the treatment of solid tumors. In general, these fast-devel- et al., 2015).
oping fields in immunotherapy and cancer vaccine development It will also be relevant to define at what point in the lifespan of a
will most likely provide additional therapeutic options and in- CTC the cell undergoes a MET-like process: before it reaches the
sights into how to target and eradicate the population of cells metastatic site (i.e., in the bloodstream) or after the metastasis
that have undergone EMT in cancer. has started to grow at a new site. Answering this question will

36 Cell 166, June 30, 2016


surely be aided by identifying better markers for CTCs, particu- ence Institute National University of Singapore and Institute of Molecular Cell
larly as EpCAM fails to detect all CTC states. It is also important Biology A*STAR Singapore to J.P.T. The Instituto de Neurociencias is a Centre
of Excellence Severo Ochoa. This review is dedicated to the late Professor
to ask whether CSCs travel with CTCs in collective migration, the
Gerald Maurice Edelman who was very influential on the research carried
extent to which CSCs convert into non-CSCs (and vice versa) out by J.P.T. and to Mr. Angel Nieto who passed away recently and always
(Chaffer et al., 2011; Schwitalla et al., 2013), and how is such was a role model for M.A.N.
plasticity controlled.
Cancer stroma crosstalk appears to be context specific, and REFERENCES
thus, we must better understand the nature of the stroma that
respectively promotes or represses EMT in the primary tumor Abell, A.N., Jordan, N.V., Huang, W., Prat, A., Midland, A.A., Johnson, N.L.,
or in the metastatic niche. In relation to this, the description of Granger, D.A., Mieczkowski, P.A., Perou, C.M., Gomez, S.M., et al. (2011).
a pre-metastatic niche has been an important development (Ka- MAP3K4/CBP-regulated H2B acetylation controls epithelial-mesenchymal
transition in trophoblast stem cells. Cell Stem Cell 8, 525537.
plan et al., 2006), as has the role of membrane vesicles derived
Aceto, N., Bardia, A., Miyamoto, D.T., Donaldson, M.C., Wittner, B.S.,
from tumor cells, called exosomes, which can induce vascular
Spencer, J.A., Yu, M., Pely, A., Engstrom, A., Zhu, H., et al. (2014). Circulating
leakiness and bone marrow progenitor mobilization to educate
tumor cell clusters are oligoclonal precursors of breast cancer metastasis. Cell
the niche (Peinado et al., 2012). Exosomes derived from pancre- 158, 11101122.
atic ductal carcinoma can initiate the pre-metastatic niche in the Acharya, A., Baek, S.T., Huang, G., Eskiocak, B., Goetsch, S., Sung, C.Y.,
liver (Costa-Silva et al., 2015). The target cells are the Kupffer Banfi, S., Sauer, M.F., Olsen, G.S., Duffield, J.S., et al. (2012). The bHLH tran-
cells, which instruct stellate cells to transdifferentiate into myofi- scription factor Tcf21 is required for lineage-specific EMT of cardiac fibroblast
broblasts and generate a fibrotic microenvironment that favors progenitors. Development 139, 21392149.
the formation of the metastatic niche. This is reminiscent of the Acloque, H., Ocana, O.H., Matheu, A., Rizzoti, K., Wise, C., Lovell-Badge, R.,
situation in renal fibrosis, where damaged epithelial cells secrete and Nieto, M.A. (2011). Reciprocal repression between Sox3 and snail transcrip-
exosomes that initiate the differentiation of fibroblasts into tion factors defines embryonic territories at gastrulation. Dev. Cell 21, 546558.
myofibroblasts (Borges et al., 2013). Importantly, the homing Aghdassi, A., Sendler, M., Guenther, A., Mayerle, J., Behn, C.O., Heidecke,
specificity of exosomes directed by the expression of different C.D., Friess, H., Buchler, M., Evert, M., Lerch, M.M., and Weiss, F.U. (2012).
Recruitment of histone deacetylases HDAC1 and HDAC2 by the transcrip-
integrins defines the target tissue where cancer cells will prefer-
tional repressor ZEB1 downregulates E-cadherin expression in pancreatic
entially form metastatic outgrowths (Hoshino et al., 2015). Thus, cancer. Gut 61, 439448.
a deeper analysis of the metastatic niche and its interaction with Akalay, I., Janji, B., Hasmim, M., Noman, M.Z., Andre, F., De Cremoux, P.,
exosomes, DTCs, and dormant cancer cells is now warranted. Bertheau, P., Badoual, C., Vielh, P., Larsen, A.K., et al. (2013). Epithelial-to-
Most intriguing is how we can target the EMT program for ther- mesenchymal transition and autophagy induction in breast carcinoma pro-
apeutic gain. Specific deletion of the cells that have undergone mote escape from T-cell-mediated lysis. Cancer Res. 73, 24182427.
EMT would be optimal to prevent the colonization of distant sites Alderton, G.K. (2013). Metastasis: Epithelial to mesenchymal and back again.
by newly generated DTCs or long-standing dormant cells. This Nat. Rev. Cancer 13, 3.
implies specifically rendering those cells sensitive to cytotoxic Alix-Panabie`res, C., and Pantel, K. (2014). Technologies for detection of circu-
drugs, which is a challenge if we consider that they are resistant lating tumor cells: facts and vision. Lab Chip 14, 5762.
to multiple death-inducing signals and also to chemo- and Antoniou, A., Hebrant, A., Dom, G., Dumont, J.E., and Maenhaut, C. (2013).
immunotherapy (Fischer et al., 2015; Kudo-Saito et al., 2009; Cancer stem cells, a fuzzy evolving concept: a cell population or a cell prop-
erty? Cell Cycle 12, 37433748.
Voon et al., 2013; Zheng et al., 2015). Many of these challenges
will require innovative approaches. Aref, A.R., Huang, R.Y., Yu, W., Chua, K.N., Sun, W., Tu, T.Y., Bai, J., Sim,
W.J., Zervantonakis, I.K., Thiery, J.P., and Kamm, R.D. (2013). Screening ther-
We are at the dawn of a new era of personalized medicine and
apeutic EMT blocking agents in a three-dimensional microenvironment. Integr
the recent advances in pharmacogenomics will provide clear-cut Biol (Camb) 5, 381389.
opportunities to tailor individualized cancer treatment strategies. Ariza, L., Carmona, R., Canete, A., Cano, E., and Munoz-Chapuli, R. (2016).
With better treatments for cancer should come additional in- Coelomic epithelium-derived cells in visceral morphogenesis. Dev. Dyn. 245,
sights into how cancer progresses, and knowledge as to which 307322.
are the most relevant drivers, pathways, and modes of dissem- Arnoux, V., Nassour, M., LHelgoualch, A., Hipskind, R.A., and Savagner, P.
ination. These advances may help clarify the extent to which (2008). Erk5 controls Slug expression and keratinocyte activation during
EMT is involved in the various disease states and point to wound healing. Mol. Biol. Cell 19, 47384749.
avenues through which our current understanding of the EMT Au, S.H., Storey, B.D., Moore, J.C., Tang, Q., Chen, Y.-L., Javaid, S., Sarioglu,
pathway and transitional events can be exploited to target tu- A.F., Sullivan, R., Madden, M.W., OKeefe, R., et al. (2016). Clusters of circu-
mors and/or make them more susceptible to treatment regimes. lating tumor cells traverse capillary-sized vessels. Proc. Natl. Acad. Sci. USA
113, 49474952.
Baccelli, I., Schneeweiss, A., Riethdorf, S., Stenzinger, A., Schillert, A., Vogel,
ACKNOWLEDGMENTS
V., Klein, C., Saini, M., Bauerle, T., Wallwiener, M., et al. (2013). Identification of
a population of blood circulating tumor cells from breast cancer patients that
We sincerely apologize to colleagues whose work has been omitted from this
initiates metastasis in a xenograft assay. Nat. Biotechnol. 31, 539544.
review owing to space limitations. We thank Stuart Ingham from the Instituto
de Neurociencias for his help with the figures. This work has been supported Baek, S.T., and Tallquist, M.D. (2012). Nf1 limits epicardial derivative expan-
by grants from the Spanish Ministry of Economy and Competitiveness sion by regulating epithelial to mesenchymal transition and proliferation.
(MINECO-BFU2014-53128-R), the Generalitat Valenciana (Prometeo II/2013/ Development 139, 20402049.
002 and ISIC 2012/010), and the European Research Council (ERC Baggiolini, A., Varum, S., Mateos, J.M., Bettosini, D., John, N., Bonalli, M., Zie-
AdG322694) to M.A.N. and from Department of Biochemistry and Cancer Sci- gler, U., Dimou, L., Clevers, H., Furrer, R., and Sommer, L. (2015). Premigratory

Cell 166, June 30, 2016 37


and migratory neural crest cells are multipotent in vivo. Cell Stem Cell 16, exosomes from injured epithelial cells activate fibroblasts to initiate tissue
314322. regenerative responses and fibrosis. J. Am. Soc. Nephrol. 24, 385392.
Bahri, S., Wang, S., Conder, R., Choy, J., Vlachos, S., Dong, K., Merino, C., Boucaut, J.C., Darribe`re, T., Boulekbache, H., and Thiery, J.P. (1984). Preven-
Sigrist, S., Molnar, C., Yang, X., et al. (2010). The leading edge during dorsal tion of gastrulation but not neurulation by antibodies to fibronectin in
closure as a model for epithelial plasticity: Pak is required for recruitment amphibian embryos. Nature 307, 364367.
of the Scribble complex and septate junction formation. Development 137, Boutet, A., De Frutos, C.A., Maxwell, P.H., Mayol, M.J., Romero, J., and Nieto,
20232032. M.A. (2006). Snail activation disrupts tissue homeostasis and induces fibrosis
Bai, J., Adriani, G., Dang, T.M., Tu, T.Y., Penny, H.X., Wong, S.C., Kamm, R.D., in the adult kidney. EMBO J. 25, 56035613.
and Thiery, J.P. (2015a). Contact-dependent carcinoma aggregate dispersion Brabletz, T. (2012). EMT and MET in metastasis: where are the cancer stem
by M2a macrophages via ICAM-1 and b2 integrin interactions. Oncotarget 6, cells? Cancer Cell 22, 699701.
2529525307.
Brabletz, T., Jung, A., Reu, S., Porzner, M., Hlubek, F., Kunz-Schughart, L.A.,
Bai, J., Tu, T.Y., Kim, C., Thiery, J.P., and Kamm, R.D. (2015b). Identification of Knuechel, R., and Kirchner, T. (2001). Variable beta-catenin expression in
drugs as single agents or in combination to prevent carcinoma dissemination colorectal cancers indicates tumor progression driven by the tumor environ-
in a microfluidic 3D environment. Oncotarget 6, 3660336614. ment. Proc. Natl. Acad. Sci. USA 98, 1035610361.
Barcellos-Hoff, M.H., Lyden, D., and Wang, T.C. (2013). The evolution of the Brabletz, T., Jung, A., Spaderna, S., Hlubek, F., and Kirchner, T. (2005).
cancer niche during multistage carcinogenesis. Nat. Rev. Cancer 13, 511518. Opinion: migrating cancer stem cells - an integrated concept of malignant
Basu, R., Whitlock, B.M., Husson, J., Le Floch, A., Jin, W., Oyler-Yaniv, A., tumour progression. Nat. Rev. Cancer 5, 744749.
Dotiwala, F., Giannone, G., Hivroz, C., Biais, N., et al. (2016). Cytotoxic T Cells Braeutigam, C., Rago, L., Rolke, A., Waldmeier, L., Christofori, G., and Winter,
Use Mechanical Force to Potentiate Target Cell Killing. Cell 165, 100110. J. (2014). The RNA-binding protein Rbfox2: an essential regulator of EMT-
Bechtel, W., McGoohan, S., Zeisberg, E.M., Muller, G.A., Kalbacher, H., driven alternative splicing and a mediator of cellular invasion. Oncogene 33,
Salant, D.J., Muller, C.A., Kalluri, R., and Zeisberg, M. (2010). Methylation 10821092.
determines fibroblast activation and fibrogenesis in the kidney. Nat. Med. Brockhausen, J., Tay, S.S., Grzelak, C.A., Bertolino, P., Bowen, D.G., dAvig-
16, 544550. dor, W.M., Teoh, N., Pok, S., Shackel, N., Gamble, J.R., et al. (2015). miR-181a
Beck, B., Lapouge, G., Rorive, S., Drogat, B., Desaedelaere, K., Delafaille, S., mediates TGF-b-induced hepatocyte EMT and is dysregulated in cirrhosis and
Dubois, C., Salmon, I., Willekens, K., Marine, J.C., and Blanpain, C. (2015). hepatocellular cancer. Liver Int. 35, 240253.
Different levels of Twist1 regulate skin tumor initiation, stemness, and progres- Brnnum, H., Andersen, D.C., Schneider, M., Sandberg, M.B., Eskildsen, T.,
sion. Cell Stem Cell 16, 6779. Nielsen, S.B., Kalluri, R., and Sheikh, S.P. (2013). miR-21 promotes fibrogenic
Bedi, U., Mishra, V.K., Wasilewski, D., Scheel, C., and Johnsen, S.A. (2014). epithelial-to-mesenchymal transition of epicardial mesothelial cells involving
Epigenetic plasticity: a central regulator of epithelial-to-mesenchymal transi- Programmed Cell Death 4 and Sprouty-1. PLoS ONE 8, e56280.
tion in cancer. Oncotarget 5, 20162029. Buitrago-Delgado, E., Nordin, K., Rao, A., Geary, L., and LaBonne, C. (2015).
Beerling, E., Seinstra, D., de Wit, E., Kester, L., van der Velden, D., Maynard, NEURODEVELOPMENT. Shared regulatory programs suggest retention of
C., Schafer, R., van Diest, P., Voest, E., van Oudenaarden, A., et al. (2016). blastula-stage potential in neural crest cells. Science 348, 13321335.
Plasticity between Epithelial and Mesenchymal States Unlinks EMT from Byers, L.A., Diao, L., Wang, J., Saintigny, P., Girard, L., Peyton, M., Shen, L.,
Metastasis-Enhancing Stem Cell Capacity. Cell Rep. 14, 22812288. Fan, Y., Giri, U., Tumula, P.K., et al. (2013). An epithelial-mesenchymal transi-
Bellaye, P.S., Burgy, O., Causse, S., Garrido, C., and Bonniaud, P. (2014). Heat tion gene signature predicts resistance to EGFR and PI3K inhibitors and iden-
shock proteins in fibrosis and wound healing: good or evil? Pharmacol. Ther. tifies Axl as a therapeutic target for overcoming EGFR inhibitor resistance. Clin.
143, 119132. Cancer Res. 19, 279290.
Beug, H. (2009). Breast cancer stem cells: eradication by differentiation ther- Calon, A., Lonardo, E., Berenguer-Llergo, A., Espinet, E., Hernando-Mom-
apy? Cell 138, 623625. blona, X., Iglesias, M., Sevillano, M., Palomo-Ponce, S., Tauriello, D.V., Byrom,
Bidard, F.C., Vincent-Salomon, A., Gomme, S., Nos, C., de Rycke, Y., Thiery, D., et al. (2015). Stromal gene expression defines poor-prognosis subtypes in
J.P., Sigal-Zafrani, B., Mignot, L., Sastre-Garau, X., and Pierga, J.Y.; Institut colorectal cancer. Nat. Genet. 47, 320329.
Curie Breast Cancer Study Group (2008). Disseminated tumor cells of breast Cano, A., Perez-Moreno, M.A., Rodrigo, I., Locascio, A., Blanco, M.J., del Bar-
cancer patients: a strong prognostic factor for distant and local relapse. rio, M.G., Portillo, F., and Nieto, M.A. (2000). The transcription factor snail con-
Clin. Cancer Res. 14, 33063311. trols epithelial-mesenchymal transitions by repressing E-cadherin expression.
Bidard, F.C., Peeters, D.J., Fehm, T., Nole, F., Gisbert-Criado, R., Mavroudis, Nat. Cell Biol. 2, 7683.
D., Grisanti, S., Generali, D., Garcia-Saenz, J.A., Stebbing, J., et al. (2014). Cano, E., Carmona, R., Ruiz-Villalba, A., Rojas, A., Chau, Y.-Y., Wagner, N.,
Clinical validity of circulating tumour cells in patients with metastatic breast Hastie, N.D., Munoz-Chapuli, R., and Perez-Pomares, J.M. (2016). Extracar-
cancer: a pooled analysis of individual patient data. Lancet Oncol. 15, diac septum transversum/proepicardial endothelial cells pattern embryonic
406414. coronary arterio-venous connections. Proc. Natl. Acad. Sci. USA 113,
Blanco, M.J., Barrallo-Gimeno, A., Acloque, H., Reyes, A.E., Tada, M., Allende, 656661.
M.L., Mayor, R., and Nieto, M.A. (2007). Snail1a and Snail1b cooperate in Carmody, L.C., Germain, A.R., VerPlank, L., Nag, P.P., Munoz, B., Perez, J.R.,
the anterior migration of the axial mesendoderm in the zebrafish embryo. and Palmer, M.A. (2012). Phenotypic high-throughput screening elucidates
Development 134, 40734081. target pathway in breast cancer stem cell-like cells. J. Biomol. Screen. 17,
Boglev, Y., Wilanowski, T., Caddy, J., Parekh, V., Auden, A., Darido, C., Hislop, 12041210.
N.R., Cangkrama, M., Ting, S.B., and Jane, S.M. (2011). The unique and coop- Carver, E.A., Jiang, R., Lan, Y., Oram, K.F., and Gridley, T. (2001). The mouse
erative roles of the Grainy head-like transcription factors in epidermal develop- snail gene encodes a key regulator of the epithelial-mesenchymal transition.
ment reflect unexpected target gene specificity. Dev. Biol. 349, 512522. Mol. Cell. Biol. 21, 81848188.
Bonomi, S., di Matteo, A., Buratti, E., Cabianca, D.S., Baralle, F.E., Ghigna, C., Celia`-Terrassa, T., Meca-Cortes, O., Mateo, F., Martnez de Paz, A., Rubio, N.,
and Biamonti, G. (2013). HnRNP A1 controls a splicing regulatory circuit pro- Arnal-Estape, A., Ell, B.J., Bermudo, R., Daz, A., Guerra-Rebollo, M., et al.
moting mesenchymal-to-epithelial transition. Nucleic Acids Res. 41, 8665 (2012). Epithelial-mesenchymal transition can suppress major attributes of
8679. human epithelial tumor-initiating cells. J. Clin. Invest. 122, 18491868.
Borges, F.T., Melo, S.A., Ozdemir, B.C., Kato, N., Revuelta, I., Miller, C.A., Chaffer, C.L., Brueckmann, I., Scheel, C., Kaestli, A.J., Wiggins, P.A., Ro-
Gattone, V.H., 2nd, LeBleu, V.S., and Kalluri, R. (2013). TGF-b1-containing drigues, L.O., Brooks, M., Reinhardt, F., Su, Y., Polyak, K., et al. (2011). Normal

38 Cell 166, June 30, 2016


and neoplastic nonstem cells can spontaneously convert to a stem-like state. Clay, M.R., and Halloran, M.C. (2014). Cadherin 6 promotes neural crest cell
Proc. Natl. Acad. Sci. USA 108, 79507955. detachment via F-actin regulation and influences active Rho distribution dur-
Chaffer, C.L., Marjanovic, N.D., Lee, T., Bell, G., Kleer, C.G., Reinhardt, F., ing epithelial-to-mesenchymal transition. Development 141, 25062515.
DAlessio, A.C., Young, R.A., and Weinberg, R.A. (2013). Poised chromatin Conn, S.J., Pillman, K.A., Toubia, J., Conn, V.M., Salmanidis, M., Phillips, C.A.,
at the ZEB1 promoter enables breast cancer cell plasticity and enhances Roslan, S., Schreiber, A.W., Gregory, P.A., and Goodall, G.J. (2015). The RNA
tumorigenicity. Cell 154, 6174. binding protein quaking regulates formation of circRNAs. Cell 160, 11251134.
Chakrabarti, R., Hwang, J., Andres Blanco, M., Wei, Y., Lukac isin, M., Costa-Silva, B., Aiello, N.M., Ocean, A.J., Singh, S., Zhang, H., Thakur, B.K.,
Romano, R.A., Smalley, K., Liu, S., Yang, Q., Ibrahim, T., et al. (2012). Elf5 in- Becker, A., Hoshino, A., Mark, M.T., Molina, H., et al. (2015). Pancreatic cancer
hibits the epithelial-mesenchymal transition in mammary gland development exosomes initiate pre-metastatic niche formation in the liver. Nat. Cell Biol. 17,
and breast cancer metastasis by transcriptionally repressing Snail2. Nat. 816826.
Cell Biol. 14, 12121222.
Cristofanilli, M., Budd, G.T., Ellis, M.J., Stopeck, A., Matera, J., Miller, M.C.,
Chan, Y.S., Goke, J., Lu, X., Venkatesan, N., Feng, B., Su, I.H., and Ng, H.H. Reuben, J.M., Doyle, G.V., Allard, W.J., Terstappen, L.W., and Hayes, D.F.
(2013). A PRC2-dependent repressive role of PRDM14 in human embryonic (2004). Circulating tumor cells, disease progression, and survival in metastatic
stem cells and induced pluripotent stem cell reprogramming. Stem Cells 31, breast cancer. N. Engl. J. Med. 351, 781791.
682692.
Darido, C., and Jane, S.M. (2010). Grhl3 and GEF19 in the front rho. Small
Chang, C.J., Chao, C.H., Xia, W., Yang, J.Y., Xiong, Y., Li, C.W., Yu, W.H., Reh- GTPases 1, 104107.
man, S.K., Hsu, J.L., Lee, H.H., et al. (2011). p53 regulates epithelial-mesen-
David, C.J., Huang, Y.H., Chen, M., Su, J., Zou, Y., Bardeesy, N., Iacobuzio-
chymal transition and stem cell properties through modulating miRNAs. Nat.
Donahue, C.A., and Massague, J. (2016). TGF-b Tumor Suppression through
Cell Biol. 13, 317323.
a Lethal EMT. Cell 164, 10151030.
Chang, C.C., Hsu, W.H., Wang, C.C., Chou, C.H., Kuo, M.Y., Lin, B.R., Chen,
Del Pozo Martin, Y., Park, D., Ramachandran, A., Ombrato, L., Calvo, F., Chak-
S.T., Tai, S.K., Kuo, M.L., and Yang, M.H. (2013). Connective tissue growth
ravarty, P., Spencer-Dene, B., Derzsi, S., Hill, C.S., Sahai, E., and Malanchi, I.
factor activates pluripotency genes and mesenchymal-epithelial transition in
(2015). Mesenchymal Cancer Cell-Stroma Crosstalk Promotes Niche Activa-
head and neck cancer cells. Cancer Res. 73, 41474157.
tion, Epithelial Reversion, and Metastatic Colonization. Cell Rep. 13, 2456
Chen, Y.L., Zhang, X., Bai, J., Gai, L., Ye, X.L., Zhang, L., Xu, Q., Zhang, Y.X., 2469.
Xu, L., Li, H.P., and Ding, X. (2013). Sorafenib ameliorates bleomycin-induced
pulmonary fibrosis: potential roles in the inhibition of epithelial-mesenchymal Deng, H., Yang, F., Xu, H., Sun, Y., Xue, X., Du, S., Wang, X., Li, S., Liu, Y., and
transition and fibroblast activation. Cell Death Dis. 4, e665. Wang, R. (2014). Ac-SDKP suppresses epithelial-mesenchymal transition in
A549 cells via HSP27 signaling. Exp. Mol. Pathol. 97, 176183.
Chen, A., Wong, C.S., Liu, M.C., House, C.M., Sceneay, J., Bowtell, D.D.,
Thompson, E.W., and Moller, A. (2015). The ubiquitin ligase Siah is a novel Diaz, V.M., and de Herreros, A.G. (2016). F-box proteins: Keeping the epithe-
regulator of Zeb1 in breast cancer. Oncotarget 6, 862873. lial-to-mesenchymal transition (EMT) in check. Semin. Cancer Biol. 36, 7179.

Cheng, S.Z.D., and Keller, A. (1998). The role of metastable states in polymer Daz-Lopez, A., Moreno-Bueno, G., and Cano, A. (2014). Role of microRNA in
phase transitions: Concepts, principles, and experimental observations. Annu. epithelial to mesenchymal transition and metastasis and clinical perspectives.
Rev. Mater. Sci. 28, 533562. Cancer Manag. Res. 6, 205216.
Chouaib, S., Janji, B., Tittarelli, A., Eggermont, A., and Thiery, J.P. (2014). Tumor Domnguez, D., Montserrat-Sents, B., Virgos-Soler, A., Guaita, S., Grueso, J.,
plasticity interferes with anti-tumor immunity. Crit. Rev. Immunol. 34, 91102. Porta, M., Puig, I., Baulida, J., Franc, C., and Garca de Herreros, A. (2003).
Phosphorylation regulates the subcellular location and activity of the snail tran-
Choudhary, K.S., Rohatgi, N., Halldorsson, S., Briem, E., Gudjonsson, T., Gud-
scriptional repressor. Mol. Cell. Biol. 23, 50785089.
mundsson, S., and Rolfsson, O. (2016). EGFR Signal-Network Reconstruction
Demonstrates Metabolic Crosstalk in EMT. PLoS Comput. Biol. 12, e1004924. Dong, C., Wu, Y., Wang, Y., Wang, C., Kang, T., Rychahou, P.G., Chi, Y.I.,
Chu, Y.S., Eder, O., Thomas, W.A., Simcha, I., Pincet, F., Ben-Zeev, A., Perez, Evers, B.M., and Zhou, B.P. (2013). Interaction with Suv39H1 is critical for
E., Thiery, J.P., and Dufour, S. (2006). Prototypical type I E-cadherin and type II Snail-mediated E-cadherin repression in breast cancer. Oncogene 32, 1351
cadherin-7 mediate very distinct adhesiveness through their extracellular do- 1362.
mains. J. Biol. Chem. 281, 29012910. Drasin, D.J., Guarnieri, A.L., Neelakantan, D., Kim, J., Cabrera, J.H., Wang,
Chu, A.S., Diaz, R., Hui, J.J., Yanger, K., Zong, Y., Alpini, G., Stanger, B.Z., and C.A., Zaberezhnyy, V., Gasparini, P., Cascione, L., Huebner, K., et al. (2015).
Wells, R.G. (2011). Lineage tracing demonstrates no evidence of cholangio- TWIST1-Induced miR-424 Reversibly Drives Mesenchymal Programming
cyte epithelial-to-mesenchymal transition in murine models of hepatic fibrosis. while Inhibiting Tumor Initiation. Cancer Res. 75, 19081921.
Hepatology 53, 16851695. Du, C., Zhang, C., Hassan, S., Biswas, M.H., and Balaji, K.C. (2010). Protein
Chua, K.N., Poon, K.L., Lim, J., Sim, W.J., Huang, R.Y., and Thiery, J.P. (2011). kinase D1 suppresses epithelial-to-mesenchymal transition through phos-
Target cell movement in tumor and cardiovascular diseases based on the phorylation of snail. Cancer Res. 70, 78107819.
epithelial-mesenchymal transition concept. Adv. Drug Deliv. Rev. 63, 558567. Edelman, G.M., Gallin, W.J., Delouvee, A., Cunningham, B.A., and Thiery, J.P.
Chua, K.N., Sim, W.J., Racine, V., Lee, S.Y., Goh, B.C., and Thiery, J.P. (2012). (1983). Early epochal maps of two different cell adhesion molecules. Proc.
A cell-based small molecule screening method for identifying inhibitors of Natl. Acad. Sci. USA 80, 43844388.
epithelial-mesenchymal transition in carcinoma. PLoS ONE 7, e33183. Engl, W., Viasnoff, V., and Thiery, J.P. (2015). Epithelial Mesenchymal Transi-
Chung, V.Y., Tan, T.Z., Tan, M., Wong, M.K., Kuay, K.T., Yang, Z., Ye, J., tion Influence on CTL Activity. In Resistance of Cancer Cells to CTL-Mediated
Muller, J., Koh, C.M.-Y., Guccione, E., et al. (2016). GRHL2-miR-200-ZEB1 Immunotherapy, B. Bonavida and S. Chouaib, eds. (Springer International
maintains the epithelial status of ovarian cancer through transcriptional regu- Publishing), pp. 267284.
lation and histone modification. Sci. Rep. 6, 19943. http://dx.doi.org/10.1038/ Enkhbaatar, Z., Terashima, M., Oktyabri, D., Tange, S., Ishimura, A., Yano, S.,
srep19943. and Suzuki, T. (2013). KDM5B histone demethylase controls epithelial-mesen-
Cieply, B., Riley, P., 4th, Pifer, P.M., Widmeyer, J., Addison, J.B., Ivanov, A.V., chymal transition of cancer cells by regulating the expression of the micro-
Denvir, J., and Frisch, S.M. (2012). Suppression of the epithelial-mesenchymal RNA-200 family. Cell Cycle 12, 21002112.
transition by Grainyhead-like-2. Cancer Res. 72, 24402453. Fischer, K.R., Durrans, A., Lee, S., Sheng, J., Li, F., Wong, S.T., Choi, H., El
Ciruna, B., and Rossant, J. (2001). FGF signaling regulates mesoderm cell fate Rayes, T., Ryu, S., Troeger, J., et al. (2015). Epithelial-to-mesenchymal transi-
specification and morphogenetic movement at the primitive streak. Dev. Cell tion is not required for lung metastasis but contributes to chemoresistance.
1, 3749. Nature 527, 472476.

Cell 166, June 30, 2016 39


Futterman, M.A., Garca, A.J., and Zamir, E.A. (2011). Evidence for partial Hlubek, F., Lohberg, C., Meiler, J., Jung, A., Kirchner, T., and Brabletz, T.
epithelial-to-mesenchymal transition (pEMT) and recruitment of motile blasto- (2001). Tip60 is a cell-type-specific transcriptional regulator. J. Biochem.
derm edge cells during avian epiboly. Dev. Dyn. 240, 15021511. 129, 635641.
Gaeta, X., Xie, Y., and Lowry, W.E. (2013). Sequential addition of reprogram- Holzel, M., Bovier, A., and Tuting, T. (2013). Plasticity of tumour and immune
ming factors improves efficiency. Nat. Cell Biol. 15, 725727. cells: a source of heterogeneity and a cause for therapy resistance? Nat.
Gentles, A.J., Newman, A.M., Liu, C.L., Bratman, S.V., Feng, W., Kim, D., Nair, Rev. Cancer 13, 365376.
V.S., Xu, Y., Khuong, A., Hoang, C.D., et al. (2015). The prognostic landscape Hong, J., Zhou, J., Fu, J., He, T., Qin, J., Wang, L., Liao, L., and Xu, J. (2011).
of genes and infiltrating immune cells across human cancers. Nat. Med. 21, Phosphorylation of serine 68 of Twist1 by MAPKs stabilizes Twist1 protein and
938945. promotes breast cancer cell invasiveness. Cancer Res. 71, 39803990.
Giampieri, S., Manning, C., Hooper, S., Jones, L., Hill, C.S., and Sahai, E. Hong, T., Watanabe, K., Ta, C.H., Villarreal-Ponce, A., Nie, Q., and Dai, X.
(2009). Localized and reversible TGFbeta signalling switches breast cancer (2015). An Ovol2-Zeb1 Mutual Inhibitory Circuit Governs Bidirectional and
cells from cohesive to single cell motility. Nat. Cell Biol. 11, 12871296. Multi-step Transition between Epithelial and Mesenchymal States. PLoS Com-
Giannelli, G., Villa, E., and Lahn, M. (2014). Transforming growth factor-b as a put. Biol. 11, e1004569.
therapeutic target in hepatocellular carcinoma. Cancer Res. 74, 18901894. Hoshino, A., Costa-Silva, B., Shen, T.L., Rodrigues, G., Hashimoto, A., Tesic
Ginestier, C., Hur, M.H., Charafe-Jauffret, E., Monville, F., Dutcher, J., Brown, Mark, M., Molina, H., Kohsaka, S., Di Giannatale, A., Ceder, S., et al. (2015).
M., Jacquemier, J., Viens, P., Kleer, C.G., Liu, S., et al. (2007). ALDH1 is a Tumour exosome integrins determine organotropic metastasis. Nature 527,
marker of normal and malignant human mammary stem cells and a predictor 329335.
of poor clinical outcome. Cell Stem Cell 1, 555567. Hsu, D.S., Wang, H.J., Tai, S.K., Chou, C.H., Hsieh, C.H., Chiu, P.H., Chen,
Gjerdrum, C., Tiron, C., Hiby, T., Stefansson, I., Haugen, H., Sandal, T., Col- N.J., and Yang, M.H. (2014). Acetylation of snail modulates the cytokinome
lett, K., Li, S., McCormack, E., Gjertsen, B.T., et al. (2010). Axl is an essential of cancer cells to enhance the recruitment of macrophages. Cancer Cell 26,
epithelial-to-mesenchymal transition-induced regulator of breast cancer 534548.
metastasis and patient survival. Proc. Natl. Acad. Sci. USA 107, 11241129. Hu, N., Strobl-Mazzulla, P.H., and Bronner, M.E. (2014). Epigenetic regulation
Goel, H.L., Gritsko, T., Pursell, B., Chang, C., Shultz, L.D., Greiner, D.L., No- in neural crest development. Dev. Biol. 396, 159168.
rum, J.H., Toftgard, R., Shaw, L.M., and Mercurio, A.M. (2014). Regulated Huang, R.Y., Guilford, P., and Thiery, J.P. (2012). Early events in cell adhesion
splicing of the a6 integrin cytoplasmic domain determines the fate of breast and polarity during epithelial-mesenchymal transition. J. Cell Sci. 125, 4417
cancer stem cells. Cell Rep. 7, 747761. 4422.
Grande, M.T., Sanchez-Laorden, B., Lopez-Blau, C., De Frutos, C.A., Boutet, Huang, R.Y., Wong, M.K., Tan, T.Z., Kuay, K.T., Ng, A.H., Chung, V.Y., Chu,
A., Arevalo, M., Rowe, R.G., Weiss, S.J., Lopez-Novoa, J.M., and Nieto, M.A. Y.S., Matsumura, N., Lai, H.C., Lee, Y.F., et al. (2013). An EMT spectrum de-
(2015). Snail1-induced partial epithelial-to-mesenchymal transition drives fines an anoikis-resistant and spheroidogenic intermediate mesenchymal
renal fibrosis in mice and can be targeted to reverse established disease. state that is sensitive to e-cadherin restoration by a src-kinase inhibitor, sara-
Nat. Med. 21, 989997. catinib (AZD0530). Cell Death Dis. 4, e915.
Greenburg, G., and Hay, E.D. (1982). Epithelia suspended in collagen gels can Hudson, L.G., Newkirk, K.M., Chandler, H.L., Choi, C., Fossey, S.L., Parent,
lose polarity and express characteristics of migrating mesenchymal cells. A.E., and Kusewitt, D.F. (2009). Cutaneous wound reepithelialization is
J. Cell Biol. 95, 333339. compromised in mice lacking functional Slug (Snai2). J. Dermatol. Sci. 56,
Grigore, A.D., Jolly, M.K., Jia, D., Farach-Carson, M.C., and Levine, H. (2016). 1926.
Tumor Budding: The Name is EMT. Partial EMT. J. Clin. Med. 5, E51. Huleihel, L., Ben-Yehudah, A., Milosevic, J., Yu, G., Pandit, K., Sakamoto, K.,
Gulati, A., Jabbour, A., Ismail, T.F., Guha, K., Khwaja, J., Raza, S., Morarji, K., Yousef, H., LeJeune, M., Coon, T.A., Redinger, C.J., et al. (2014). Let-7d
Brown, T.D., Ismail, N.A., Dweck, M.R., et al. (2013). Association of fibrosis microRNA affects mesenchymal phenotypic properties of lung fibroblasts.
with mortality and sudden cardiac death in patients with nonischemic dilated Am. J. Physiol. Lung Cell. Mol. Physiol. 306, L534L542.
cardiomyopathy. JAMA 309, 896908.
Humphreys, B.D., Lin, S.L., Kobayashi, A., Hudson, T.E., Nowlin, B.T.,
Gupta, P.B., Chaffer, C.L., and Weinberg, R.A. (2009a). Cancer stem cells: Bonventre, J.V., Valerius, M.T., McMahon, A.P., and Duffield, J.S. (2010).
mirage or reality? Nat. Med. 15, 10101012. Fate tracing reveals the pericyte and not epithelial origin of myofibroblasts in
Gupta, P.B., Onder, T.T., Jiang, G., Tao, K., Kuperwasser, C., Weinberg, R.A., kidney fibrosis. Am. J. Pathol. 176, 8597.
and Lander, E.S. (2009b). Identification of selective inhibitors of cancer stem Husemann, Y., Geigl, J.B., Schubert, F., Musiani, P., Meyer, M., Burghart, E.,
cells by high-throughput screening. Cell 138, 645659. Forni, G., Eils, R., Fehm, T., Riethmuller, G., and Klein, C.A. (2008). Systemic
Halbleib, J.M., and Nelson, W.J. (2006). Cadherins in development: cell adhe- spread is an early step in breast cancer. Cancer Cell 13, 5868.
sion, sorting, and tissue morphogenesis. Genes Dev. 20, 31993214. Hutchinson, J., Fogarty, A., Hubbard, R., and McKeever, T. (2015). Global
Hamilton, D.H., Litzinger, M.T., Jales, A., Huang, B., Fernando, R.I., Hodge, incidence and mortality of idiopathic pulmonary fibrosis: a systematic review.
J.W., Ardiani, A., Apelian, D., Schlom, J., and Palena, C. (2013). Immunological Eur. Respir. J. 46, 795806.
targeting of tumor cells undergoing an epithelial-mesenchymal transition via a Hwang, W.L., Jiang, J.K., Yang, S.H., Huang, T.S., Lan, H.Y., Teng, H.W.,
recombinant brachyury-yeast vaccine. Oncotarget 4, 17771790. Yang, C.Y., Tsai, Y.P., Lin, C.H., Wang, H.W., and Yang, M.H. (2014).
Harney, A.S., Arwert, E.N., Entenberg, D., Wang, Y., Guo, P., Qian, B.Z., Oktay, MicroRNA-146a directs the symmetric division of Snail-dominant colorectal
M.H., Pollard, J.W., Jones, J.G., and Condeelis, J.S. (2015). Real-Time Imag- cancer stem cells. Nat. Cell Biol. 16, 268280.
ing Reveals Local, Transient Vascular Permeability, and Tumor Cell Intravasa- Iliopoulos, D., Lindahl-Allen, M., Polytarchou, C., Hirsch, H.A., Tsichlis, P.N.,
tion Stimulated by TIE2hi Macrophage-Derived VEGFA. Cancer Discov. 5, and Struhl, K. (2010). Loss of miR-200 inhibition of Suz12 leads to poly-
932943. comb-mediated repression required for the formation and maintenance of
Hay, E.D. (1995). An overview of epithelio-mesenchymal transformation. Acta cancer stem cells. Mol. Cell 39, 761772.
Anat. (Basel) 154, 820. Infante, J.R., Camidge, D.R., Mileshkin, L.R., Chen, E.X., Hicks, R.J., Rischin,
Herranz, N., Pasini, D., Daz, V.M., Franc, C., Gutierrez, A., Dave, N., Escriva`, D., Fingert, H., Pierce, K.J., Xu, H., Roberts, W.G., et al. (2012). Safety,
M., Hernandez-Munoz, I., Di Croce, L., Helin, K., et al. (2008). Polycomb com- pharmacokinetic, and pharmacodynamic phase I dose-escalation trial of
plex 2 is required for E-cadherin repression by the Snail1 transcription factor. PF-00562271, an inhibitor of focal adhesion kinase, in advanced solid tumors.
Mol. Cell. Biol. 28, 47724781. J. Clin. Oncol. 30, 15271533.

40 Cell 166, June 30, 2016


Iwano, M., Plieth, D., Danoff, T.M., Xue, C., Okada, H., and Neilson, E.G. Lee, R.T., Nagai, H., Nakaya, Y., Sheng, G., Trainor, P.A., Weston, J.A., and
(2002). Evidence that fibroblasts derive from epithelium during tissue fibrosis. Thiery, J.P. (2013). Cell delamination in the mesencephalic neural fold and
J. Clin. Invest. 110, 341350. its implication for the origin of ectomesenchyme. Development 140, 4890
Javaid, S., Zhang, J., Anderssen, E., Black, J.C., Wittner, B.S., Tajima, K., Ting, 4902.
D.T., Smolen, G.A., Zubrowski, M., Desai, R., et al. (2013). Dynamic chromatin Lee, B., Villarreal-Ponce, A., Fallahi, M., Ovadia, J., Sun, P., Yu, Q.C., Ito, S.,
modification sustains epithelial-mesenchymal transition following inducible Sinha, S., Nie, Q., and Dai, X. (2014). Transcriptional mechanisms link epithelial
expression of Snail-1. Cell Rep. 5, 16791689. plasticity to adhesion and differentiation of epidermal progenitor cells. Dev.
Jiang, L., Xiao, L., Sugiura, H., Huang, X., Ali, A., Kuro-o, M., Deberardinis, Cell 29, 4758.
R.J., and Boothman, D.A. (2015). Metabolic reprogramming during TGFb1- Leopold, P.L., Vincent, J., and Wang, H. (2012). A comparison of epithelial-
induced epithelial-to-mesenchymal transition. Oncogene 34, 39083916. to-mesenchymal transition and re-epithelialization. Semin. Cancer Biol. 22,
Jolly, M.K., Boareto, M., Huang, B., Jia, D., Lu, M., Ben-Jacob, E., Onuchic, 471483.
J.N., and Levine, H. (2015). Implications of the Hybrid Epithelial/Mesenchymal Leptin, M., and Grunewald, B. (1990). Cell shape changes during gastrulation
Phenotype in Metastasis. Front. Oncol. 5, 155. in Drosophila. Development 110, 7384.
Jolly, M.K., Tripathi, S.C., Jia, D., Mooney, S.M., Celiktas, M., Hanash, S.M., Leroy, P., and Mostov, K.E. (2007). Slug is required for cell survival during
Mani, S.A., Pienta, K.J., Ben-Jacob, E., and Levine, H. (2016). Stability of the partial epithelial-mesenchymal transition of HGF-induced tubulogenesis.
hybrid epithelial/mesenchymal phenotype. Oncotarget. http://dx.doi.org/10. Mol. Biol. Cell 18, 19431952.
18632/oncotarget.8166. Li, W., and Kang, Y. (2016). Probing the fifty shades of EMT in metastasis.
Jordan, N.V., Johnson, G.L., and Abell, A.N. (2011). Tracking the intermediate Trends Cancer 2, 6567.
stages of epithelial-mesenchymal transition in epithelial stem cells and cancer. Li, L., Zepeda-Orozco, D., Black, R., and Lin, F. (2010a). Autophagy is a
Cell Cycle 10, 28652873. component of epithelial cell fate in obstructive uropathy. Am. J. Pathol. 176,
Jung, H.Y., and Yang, J. (2015). Unraveling the TWIST between EMT and can- 17671778.
cer stemness. Cell Stem Cell 16, 12. Li, R., Liang, J., Ni, S., Zhou, T., Qing, X., Li, H., He, W., Chen, J., Li, F., Zhuang,
Kaplan, R.N., Rafii, S., and Lyden, D. (2006). Preparing the soil: the preme- Q., et al. (2010b). A mesenchymal-to-epithelial transition initiates and is
tastatic niche. Cancer Res. 66, 1108911093. required for the nuclear reprogramming of mouse fibroblasts. Cell Stem Cell
Khoo, B.L., Lee, S.C., Kumar, P., Tan, T.Z., Warkiani, M.E., Ow, S.G., Nandi, 7, 5163.
S., Lim, C.T., and Thiery, J.P. (2015). Short-term expansion of breast circu- Lianos, G.D., Alexiou, G.A., Mangano, A., Mangano, A., Rausei, S., Boni, L.,
lating cancer cells predicts response to anti-cancer therapy. Oncotarget 6, Dionigi, G., and Roukos, D.H. (2015). The role of heat shock proteins in cancer.
1557815593. Cancer Lett. 360, 114118.
Kim, L.C., Song, L., and Haura, E.B. (2009). Src kinases as therapeutic targets Lim, J., and Thiery, J.P. (2012). Epithelial-mesenchymal transitions: insights
for cancer. Nat. Rev. Clin. Oncol. 6, 587595. from development. Development 139, 34713486.
Kitazawa, K., Hikichi, T., Nakamura, T., Mitsunaga, K., Tanaka, A., Nakamura, Lim, Y.Y., Wright, J.A., Attema, J.L., Gregory, P.A., Bert, A.G., Smith, E.,
M., Yamakawa, T., Furukawa, S., Takasaka, M., Goshima, N., et al. (2016). Thomas, D., Lopez, A.F., Drew, P.A., Khew-Goodall, Y., and Goodall, G.J.
OVOL2 Maintains the Transcriptional Program of Human Corneal Epithelium (2013). Epigenetic modulation of the miR-200 family is associated with transi-
by Suppressing Epithelial-to-Mesenchymal Transition. Cell Rep. 15, 13591368. tion to a breast cancer stem-cell-like state. J. Cell Sci. 126, 22562266.
Kong, D., Banerjee, S., Ahmad, A., Li, Y., Wang, Z., Sethi, S., and Sarkar, F.H. Lin, Y., Wu, Y., Li, J., Dong, C., Ye, X., Chi, Y.I., Evers, B.M., and Zhou, B.P.
(2010). Epithelial to mesenchymal transition is mechanistically linked with stem (2010). The SNAG domain of Snail1 functions as a molecular hook for recruiting
cell signatures in prostate cancer cells. PLoS ONE 5, e12445. lysine-specific demethylase 1. EMBO J. 29, 18031816.
Krenning, G., Zeisberg, E.M., and Kalluri, R. (2010). The origin of fibroblasts Lin, Y., Dong, C., and Zhou, B.P. (2014). Epigenetic regulation of EMT: the Snail
and mechanism of cardiac fibrosis. J. Cell. Physiol. 225, 631637. story. Curr. Pharm. Des. 20, 16981705.
Kudo-Saito, C., Shirako, H., Takeuchi, T., and Kawakami, Y. (2009). Cancer Liu, S., Cong, Y., Wang, D., Sun, Y., Deng, L., Liu, Y., Martin-Trevino, R.,
metastasis is accelerated through immunosuppression during Snail-induced Shang, L., McDermott, S.P., Landis, M.D., et al. (2013a). Breast cancer stem
EMT of cancer cells. Cancer Cell 15, 195206. cells transition between epithelial and mesenchymal states reflective of their
Kurley, S.J., Bierie, B., Carnahan, R.H., Lobdell, N.A., Davis, M.A., Hofmann, I., normal counterparts. Stem Cell Reports 2, 7891.
Moses, H.L., Muller, W.J., and Reynolds, A.B. (2012). p120-catenin is essential Liu, X., Sun, H., Qi, J., Wang, L., He, S., Liu, J., Feng, C., Chen, C., Li, W., Guo,
for terminal end bud function and mammary morphogenesis. Development Y., et al. (2013b). Sequential introduction of reprogramming factors reveals a
139, 17541764. time-sensitive requirement for individual factors and a sequential EMT-MET
Lamouille, S., Subramanyam, D., Blelloch, R., and Derynck, R. (2013). Regula- mechanism for optimal reprogramming. Nat. Cell Biol. 15, 829838.
tion of epithelial-mesenchymal and mesenchymal-epithelial transitions by Lomel, H., Starling, C., and Gridley, T. (2009). Epiblast-specific Snai1 deletion
microRNAs. Curr. Opin. Cell Biol. 25, 200207. results in embryonic lethality due to multiple vascular defects. BMC Res. Notes
Lamouille, S., Xu, J., and Derynck, R. (2014). Molecular mechanisms of epithe- 2, 22.
lial-mesenchymal transition. Nat. Rev. Mol. Cell Biol. 15, 178196. Lovisa, S., LeBleu, V.S., Tampe, B., Sugimoto, H., Vadnagara, K., Carstens,
Lane, D.P. (1992). Cancer. p53, guardian of the genome. Nature 358, 1516. J.L., Wu, C.C., Hagos, Y., Burckhardt, B.C., Pentcheva-Hoang, T., et al.
Lau, E.Y., Lo, J., Cheng, B.Y., Ma, M.K., Lee, J.M., Ng, J.K., Chai, S., Lin, C.H., (2015). Epithelial-to-mesenchymal transition induces cell cycle arrest and
Tsang, S.Y., Ma, S., et al. (2016). Cancer-Associated Fibroblasts Regulate parenchymal damage in renal fibrosis. Nat. Med. 21, 9981009.
Tumor-Initiating Cell Plasticity in Hepatocellular Carcinoma through c-Met/ Lu, M., Jolly, M.K., Levine, H., Onuchic, J.N., and Ben-Jacob, E. (2013).
FRA1/HEY1 Signaling. Cell Rep. 15, 11751189. MicroRNA-based regulation of epithelial-hybrid-mesenchymal fate determina-
LeBleu, V.S., Taduri, G., OConnell, J., Teng, Y., Cooke, V.G., Woda, C., tion. Proc. Natl. Acad. Sci. USA 110, 1814418149.
Sugimoto, H., and Kalluri, R. (2013). Origin and function of myofibroblasts in Luxan, G., DAmato, G., MacGrogan, D., and de la Pompa, J.L. (2016). Endo-
kidney fibrosis. Nat. Med. 19, 10471053. cardial Notch Signaling in Cardiac Development and Disease. Circ. Res. 118,
Lee, J.M., Dedhar, S., Kalluri, R., and Thompson, E.W. (2006). The epithelial- e1e18.
mesenchymal transition: new insights in signaling, development, and disease. Maheswaran, S., and Haber, D.A. (2015). Cell fate: Transition loses its invasive
J. Cell Biol. 172, 973981. edge. Nature 527, 452453.

Cell 166, June 30, 2016 41


Manhire-Heath, R., Golenkina, S., Saint, R., and Murray, M.J. (2013). Netrin- Murray, S.A., and Gridley, T. (2006). Snail family genes are required for left-
dependent downregulation of Frazzled/DCC is required for the dissociation right asymmetry determination, but not neural crest formation, in mice. Proc.
of the peripodial epithelium in Drosophila. Nat. Commun. 4, 2790. Natl. Acad. Sci. USA 103, 1030010304.
Mani, S.A., Guo, W., Liao, M.J., Eaton, E.N., Ayyanan, A., Zhou, A.Y., Brooks, M., Narasimha, M., Uv, A., Krejci, A., Brown, N.H., and Bray, S.J. (2008). Grainy
Reinhard, F., Zhang, C.C., Shipitsin, M., et al. (2008). The epithelial-mesen- head promotes expression of septate junction proteins and influences epithe-
chymal transition generates cells with properties of stem cells. Cell 133, 704715. lial morphogenesis. J. Cell Sci. 121, 747752.
Marmai, C., Sutherland, R.E., Kim, K.K., Dolganov, G.M., Fang, X., Kim, S.S., Navarro, P., Lozano, E., and Cano, A. (1993). Expression of E- or P-cadherin is
Jiang, S., Golden, J.A., Hoopes, C.W., Matthay, M.A., et al. (2011). Alveolar not sufficient to modify the morphology and the tumorigenic behavior of mu-
epithelial cells express mesenchymal proteins in patients with idiopathic pul- rine spindle carcinoma cells. Possible involvement of plakoglobin. J. Cell
monary fibrosis. Am. J. Physiol. Lung Cell. Mol. Physiol. 301, L71L78. Sci. 105, 923934.
Mathow, D., Chessa, F., Rabionet, M., Kaden, S., Jennemann, R., Sandhoff,
Nielsen, S.R., Quaranta, V., Linford, A., Emeagi, P., Rainer, C., Santos, A.,
R., Grone, H.J., and Feuerborn, A. (2015). Zeb1 affects epithelial cell adhesion
Ireland, L., Sakai, T., Sakai, K., Kim, Y.S., et al. (2016). Macrophage-secreted
by diverting glycosphingolipid metabolism. EMBO Rep. 16, 321331.
granulin supports pancreatic cancer metastasis by inducing liver fibrosis. Nat.
McCoy, E.L., Iwanaga, R., Jedlicka, P., Abbey, N.S., Chodosh, L.A., Heich- Cell Biol. 18, 549560.
man, K.A., Welm, A.L., and Ford, H.L. (2009). Six1 expands the mouse mam-
Nieto, M.A. (2013). Epithelial plasticity: a common theme in embryonic and
mary epithelial stem/progenitor cell pool and induces mammary tumors that
cancer cells. Science 342, 1234850.
undergo epithelial-mesenchymal transition. J. Clin. Invest. 119, 26632677.
McDonald, O.G., Wu, H., Timp, W., Doi, A., and Feinberg, A.P. (2011). Nieto, M.A., Sargent, M.G., Wilkinson, D.G., and Cooke, J. (1994). Control
Genome-scale epigenetic reprogramming during epithelial-to-mesenchymal of cell behavior during vertebrate development by Slug, a zinc finger gene.
transition. Nat. Struct. Mol. Biol. 18, 867874. Science 264, 835839.

McGranahan, N., and Swanton, C. (2015). Biological and therapeutic impact of Niwa, H., Miyazaki, J., and Smith, A.G. (2000). Quantitative expression of
intratumor heterogeneity in cancer evolution. Cancer Cell 27, 1526. Oct-3/4 defines differentiation, dedifferentiation or self-renewal of ES cells.
Nat. Genet. 24, 372376.
Mederacke, I., Hsu, C.C., Troeger, J.S., Huebener, P., Mu, X., Dapito, D.H.,
Pradere, J.P., and Schwabe, R.F. (2013). Fate tracing reveals hepatic stellate Nunan, R., Campbell, J., Mori, R., Pitulescu, M.E., Jiang, W.G., Harding, K.G.,
cells as dominant contributors to liver fibrosis independent of its aetiology. Adams, R.H., Nobes, C.D., and Martin, P. (2015). Ephrin-Bs Drive Junctional
Nat. Commun. 4, 2823. Downregulation and Actin Stress Fiber Disassembly to Enable Wound Re-
Mehal, W.Z., Iredale, J., and Friedman, S.L. (2011). Scraping fibrosis: epithelialization. Cell Rep. 13, 13801395.
expressway to the core of fibrosis. Nat. Med. 17, 552553. Oba, S., Kumano, S., Suzuki, E., Nishimatsu, H., Takahashi, M., Takamori, H.,
Meidhof, S., Brabletz, S., Lehmann, W., Preca, B.T., Mock, K., Ruh, M., Schu- Kasuya, M., Ogawa, Y., Sato, K., Kimura, K., et al. (2010). miR-200b precursor
ler, J., Berthold, M., Weber, A., Burk, U., et al. (2015). ZEB1-associated drug can ameliorate renal tubulointerstitial fibrosis. PLoS ONE 5, e13614.
resistance in cancer cells is reversed by the class I HDAC inhibitor mocetino- Ocana, O.H., Corcoles, R., Fabra, A., Moreno-Bueno, G., Acloque, H., Vega,
stat. EMBO Mol. Med. 7, 831847. S., Barrallo-Gimeno, A., Cano, A., and Nieto, M.A. (2012). Metastatic coloniza-
Millanes-Romero, A., Herranz, N., Perrera, V., Iturbide, A., Loubat-Casanovas, tion requires the repression of the epithelial-mesenchymal transition inducer
J., Gil, J., Jenuwein, T., Garca de Herreros, A., and Peiro, S. (2013). Regulation Prrx1. Cancer Cell 22, 709724.
of heterochromatin transcription by Snail1/LOXL2 during epithelial-to-mesen- Oda, H., Tsukita, S., and Takeichi, M. (1998). Dynamic behavior of the
chymal transition. Mol. Cell 52, 746757. cadherin-based cell-cell adhesion system during Drosophila gastrulation.
Mingot, J.M., Vega, S., Maestro, B., Sanz, J.M., and Nieto, M.A. (2009). Char- Dev. Biol. 203, 435450.
acterization of Snail nuclear import pathways as representatives of C2H2 zinc
Okazaki, T., Chikuma, S., Iwai, Y., Fagarasan, S., and Honjo, T. (2013).
finger transcription factors. J. Cell Sci. 122, 14521460.
A rheostat for immune responses: the unique properties of PD-1 and their
Mingot, J.M., Vega, S., Cano, A., Portillo, F., and Nieto, M.A. (2013). eEF1A advantages for clinical application. Nat. Immunol. 14, 12121218.
mediates the nuclear export of SNAG-containing proteins via the Exportin5-
Palumbo-Zerr, K., Zerr, P., Distler, A., Fliehr, J., Mancuso, R., Huang, J.,
aminoacyl-tRNA complex. Cell Rep. 5, 727737.
Mielenz, D., Tomcik, M., Furnrohr, B.G., Scholtysek, C., et al. (2015). Orphan
Montgomery, R.L., Yu, G., Latimer, P.A., Stack, C., Robinson, K., Dalby, C.M., nuclear receptor NR4A1 regulates transforming growth factor-b signaling
Kaminski, N., and van Rooij, E. (2014). MicroRNA mimicry blocks pulmonary and fibrosis. Nat. Med. 21, 150158.
fibrosis. EMBO Mol. Med. 6, 13471356.
Pandian, V., Ramraj, S., Khan, F.H., Azim, T., and Aravindan, N. (2015). Meta-
Moody, S.E., Perez, D., Pan, T.C., Sarkisian, C.J., Portocarrero, C.P., Sterner,
static neuroblastoma cancer stem cells exhibit flexible plasticity and adaptive
C.J., Notorfrancesco, K.L., Cardiff, R.D., and Chodosh, L.A. (2005). The tran-
stemness signaling. Stem Cell Res. Ther. 6, 2.
scriptional repressor Snail promotes mammary tumor recurrence. Cancer Cell
8, 197209. Park, K.S., and Gumbiner, B.M. (2010). Cadherin 6B induces BMP signaling
and de-epithelialization during the epithelial mesenchymal transition of the
Morel, A.P., Lie`vre, M., Thomas, C., Hinkal, G., Ansieau, S., and Puisieux, A.
neural crest. Development 137, 26912701.
(2008). Generation of breast cancer stem cells through epithelial-mesen-
chymal transition. PLoS ONE 3, e2888. Park, J.H., Yoon, J., Lee, K.Y., and Park, B. (2015). Effects of geniposide on
Morizane, R., Fujii, S., Monkawa, T., Hiratsuka, K., Yamaguchi, S., Homma, K., hepatocytes undergoing epithelial-mesenchymal transition in hepatic fibrosis
and Itoh, H. (2014). miR-34c attenuates epithelial-mesenchymal transition and by targeting TGFb/Smad and ERK-MAPK signaling pathways. Biochimie
kidney fibrosis with ureteral obstruction. Sci. Rep. 4, 4578. 113, 2634.

Moskowitz, I.P., Wang, J., Peterson, M.A., Pu, W.T., Mackinnon, A.C., Pattabiraman, D.R., Bierie, B., Kober, K.I., Thiru, P., Krall, J.A., Zill, C.,
Oxburgh, L., Chu, G.C., Sarkar, M., Berul, C., Smoot, L., et al. (2011). Tran- Reinhardt, F., Tam, W.L., and Weinberg, R.A. (2016). Activation of PKA leads
scription factor genes Smad4 and Gata4 cooperatively regulate cardiac valve to mesenchymal-to-epithelial transition and loss of tumor-initiating ability.
development. [corrected]. Proc. Natl. Acad. Sci. USA 108, 40064011. Science 351, aad3680.
Motzer, R.J., Rini, B.I., McDermott, D.F., Redman, B.G., Kuzel, T.M., Harrison, Pedroza, M., Le, T.T., Lewis, K., Karmouty-Quintana, H., To, S., George, A.T.,
M.R., Vaishampayan, U.N., Drabkin, H.A., George, S., Logan, T.F., et al. Blackburn, M.R., Tweardy, D.J., and Agarwal, S.K. (2016). STAT-3 contributes
(2015). Nivolumab for Metastatic Renal Cell Carcinoma: Results of a Random- to pulmonary fibrosis through epithelial injury and fibroblast-myofibroblast
ized Phase II Trial. J. Clin. Oncol. 33, 14301437. differentiation. FASEB J. 30, 129140.

42 Cell 166, June 30, 2016


Peinado, H., Ballestar, E., Esteller, M., and Cano, A. (2004). Snail mediates Robinson, B.D., Sica, G.L., Liu, Y.F., Rohan, T.E., Gertler, F.B., Condeelis, J.S.,
E-cadherin repression by the recruitment of the Sin3A/histone deacetylase 1 and Jones, J.G. (2009). Tumor microenvironment of metastasis in human
(HDAC1)/HDAC2 complex. Mol. Cell. Biol. 24, 306319. breast carcinoma: a potential prognostic marker linked to hematogenous
Peinado, H., Portillo, F., and Cano, A. (2005). Switching on-off Snail: LOXL2 dissemination. Clin. Cancer Res. 15, 24332441.
versus GSK3beta. Cell Cycle 4, 17491752. Rodon, J., Carducci, M.A., Sepulveda-Sanchez, J.M., Azaro, A., Calvo, E.,
Peinado, H., Olmeda, D., and Cano, A. (2007). Snail, Zeb and bHLH factors in Seoane, J., Brana, I., Sicart, E., Gueorguieva, I., Cleverly, A.L., et al. (2015).
tumour progression: an alliance against the epithelial phenotype? Nat. Rev. First-in-human dose study of the novel transforming growth factor-b receptor
Cancer 7, 415428. I kinase inhibitor LY2157299 monohydrate in patients with advanced cancer
and glioma. Clin. Cancer Res. 21, 553560.
Peinado, H., Alec kovic
, M., Lavotshkin, S., Matei, I., Costa-Silva, B., Moreno-
Bueno, G., Hergueta-Redondo, M., Williams, C., Garca-Santos, G., Ghajar, Rosano`, L., Spinella, F., Di Castro, V., Decandia, S., Nicotra, M.R., Natali, P.G.,
C., et al. (2012). Melanoma exosomes educate bone marrow progenitor cells and Bagnato, A. (2006). Endothelin-1 is required during epithelial to mesen-
toward a pro-metastatic phenotype through MET. Nat. Med. 18, 883891. chymal transition in ovarian cancer progression. Exp. Biol. Med. (Maywood)
231, 11281131.
Perez-Pomares, J.M., and de la Pompa, J.L. (2011). Signaling during epicar-
dium and coronary vessel development. Circ. Res. 109, 14291442. Rowe, R.G., Lin, Y., Shimizu-Hirota, R., Hanada, S., Neilson, E.G., Greenson,
J.K., and Weiss, S.J. (2011). Hepatocyte-derived Snail1 propagates liver
Plaks, V., Kong, N., and Werb, Z. (2015). The cancer stem cell niche: how fibrosis progression. Mol. Cell. Biol. 31, 23922403.
essential is the niche in regulating stemness of tumor cells? Cell Stem Cell
16, 225238. Russell, R., Perkhofer, L., Liebau, S., Lin, Q., Lechel, A., Feld, F.M., Hessmann,
E., Gaedcke, J., Guthle, M., Zenke, M., et al. (2015). Loss of ATM accelerates
Podsypanina, K., Du, Y.C., Jechlinger, M., Beverly, L.J., Hambardzumyan, D., pancreatic cancer formation and epithelial-mesenchymal transition. Nat.
and Varmus, H. (2008). Seeding and propagation of untransformed mouse Commun. 6, 7677.
mammary cells in the lung. Science 321, 18411844.
Saitoh, M., Endo, K., Furuya, S., Minami, M., Fukasawa, A., Imamura, T., and
Polyak, K., and Weinberg, R.A. (2009). Transitions between epithelial and Miyazawa, K. (2016). STAT3 integrates cooperative Ras and TGF-beta signals
mesenchymal states: acquisition of malignant and stem cell traits. Nat. Rev. that induce Snail expression. Oncogene. 35, 10491057.
Cancer 9, 265273.
Sakabe, M., Kokubo, H., Nakajima, Y., and Saga, Y. (2012). Ectopic retinoic
Polydorou, C., and Georgiades, P. (2013). Ets2-dependent trophoblast signal- acid signaling affects outflow tract cushion development through suppression
ling is required for gastrulation progression after primitive streak initiation. Nat. of the myocardial Tbx2-Tgfb2 pathway. Development 139, 385395.
Commun. 4, 1658.
Samavarchi-Tehrani, P., Golipour, A., David, L., Sung, H.K., Beyer, T.A., Datti,
Pottier, N., Cauffiez, C., Perrais, M., Barbry, P., and Mari, B. (2014). FibromiRs: A., Woltjen, K., Nagy, A., and Wrana, J.L. (2010). Functional genomics reveals
translating molecular discoveries into new anti-fibrotic drugs. Trends Pharma- a BMP-driven mesenchymal-to-epithelial transition in the initiation of somatic
col. Sci. 35, 119126. cell reprogramming. Cell Stem Cell 7, 6477.
Prat, A., Parker, J.S., Karginova, O., Fan, C., Livasy, C., Herschkowitz, J.I., He, Sarrio, D., Rodriguez-Pinilla, S.M., Hardisson, D., Cano, A., Moreno-Bueno,
X., and Perou, C.M. (2010). Phenotypic and molecular characterization of the G., and Palacios, J. (2008). Epithelial-mesenchymal transition in breast cancer
claudin-low intrinsic subtype of breast cancer. Breast Cancer Res. 12, R68. relates to the basal-like phenotype. Cancer Res. 68, 989997.
Puisieux, A., Brabletz, T., and Caramel, J. (2014). Oncogenic roles of EMT- Sato, M., Muragaki, Y., Saika, S., Roberts, A.B., and Ooshima, A. (2003).
inducing transcription factors. Nat. Cell Biol. 16, 488494. Targeted disruption of TGF-beta1/Smad3 signaling protects against
Puls, L.N., Eadens, M., and Messersmith, W. (2011). Current status of SRC renal tubulointerstitial fibrosis induced by unilateral ureteral obstruction.
inhibitors in solid tumor malignancies. Oncologist 16, 566578. J. Clin. Invest. 112, 14861494.
Qi, Q., Li, D.Y., Luo, H.R., Guan, K.L., and Ye, K. (2015). Netrin-1 exerts onco- Saunders, L.R., and McClay, D.R. (2014). Sub-circuits of a gene regulatory
genic activities through enhancing Yes-associated protein stability. Proc. Natl. network control a developmental epithelial-mesenchymal transition. Develop-
Acad. Sci. USA 112, 72557260. ment 141, 15031513.
Qian, B.Z., Zhang, H., Li, J., He, T., Yeo, E.J., Soong, D.Y., Carragher, N.O., Scarpa, E., Szabo, A., Bibonne, A., Theveneau, E., Parsons, M., and Mayor, R.
Munro, A., Chang, A., Bresnick, A.R., et al. (2015). FLT1 signaling in metas- (2015). Cadherin Switch during EMT in Neural Crest Cells Leads to Contact
tasis-associated macrophages activates an inflammatory signature that pro- Inhibition of Locomotion via Repolarization of Forces. Dev. Cell 34, 421434.
motes breast cancer metastasis. J. Exp. Med. 212, 14331448. Schafer, G., Narasimha, M., Vogelsang, E., and Leptin, M. (2014). Cadherin
Qian, L.W., Fourcaudot, A.B., Yamane, K., You, T., Chan, R.K., and Leung, switching during the formation and differentiation of the Drosophila
K.P. (2016). Exacerbated and prolonged inflammation impairs wound healing mesoderm - implications for epithelial-to-mesenchymal transitions. J. Cell
and increases scarring. Wound Repair Regen. 24, 2634. Sci. 127, 15111522.
Rada-Iglesias, A., Bajpai, R., Prescott, S., Brugmann, S.A., Swigut, T., and Wy- Schiffmacher, A.T., Padmanabhan, R., Jhingory, S., and Taneyhill, L.A. (2014).
socka, J. (2012). Epigenomic annotation of enhancers predicts transcriptional Cadherin-6B is proteolytically processed during epithelial-to-mesenchymal
regulators of human neural crest. Cell Stem Cell 11, 633648. transitions of the cranial neural crest. Mol. Biol. Cell 25, 4154.
Rembold, M., Ciglar, L., Yanez-Cuna, J.O., Zinzen, R.P., Girardot, C., Jain, A., Schmidt, J.M., Panzilius, E., Bartsch, H.S., Irmler, M., Beckers, J., Kari, V.,
Welte, M.A., Stark, A., Leptin, M., and Furlong, E.E. (2014). A conserved role for Linnemann, J.R., Dragoi, D., Hirschi, B., Kloos, U.J., et al. (2015). Stem-cell-
Snail as a potentiator of active transcription. Genes Dev. 28, 167181. like properties and epithelial plasticity arise as stable traits after transient
Rhim, A.D., Mirek, E.T., Aiello, N.M., Maitra, A., Bailey, J.M., McAllister, F., Twist1 activation. Cell Rep. 10, 131139.
Reichert, M., Beatty, G.L., Rustgi, A.K., Vonderheide, R.H., et al. (2012). Scholten, D., Osterreicher, C.H., Scholten, A., Iwaisako, K., Gu, G., Brenner,
EMT and dissemination precede pancreatic tumor formation. Cell 148, D.A., and Kisseleva, T. (2010). Genetic labeling does not detect epithelial-to-
349361. mesenchymal transition of cholangiocytes in liver fibrosis in mice. Gastroen-
Rizvi, N.A., Mazie`res, J., Planchard, D., Stinchcombe, T.E., Dy, G.K., Antonia, terology 139, 987998.
S.J., Horn, L., Lena, H., Minenza, E., Mennecier, B., et al. (2015). Activity and Schwitalla, S., Fingerle, A.A., Cammareri, P., Nebelsiek, T., Goktuna, S.I.,
safety of nivolumab, an anti-PD-1 immune checkpoint inhibitor, for patients Ziegler, P.K., Canli, O., Heijmans, J., Huels, D.J., Moreaux, G., et al. (2013).
with advanced, refractory squamous non-small-cell lung cancer (CheckMate Intestinal tumorigenesis initiated by dedifferentiation and acquisition of
063): a phase 2, single-arm trial. Lancet Oncol. 16, 257265. stem-cell-like properties. Cell 152, 2538.

Cell 166, June 30, 2016 43


Shamir, E.R., and Ewald, A.J. (2015). Adhesion in mammary development: Tang, O., Chen, X.M., Shen, S., Hahn, M., and Pollock, C.A. (2013). MiRNA-
novel roles for E-cadherin in individual and collective cell migration. Curr. 200b represses transforming growth factor-b1-induced EMT and fibronectin
Top. Dev. Biol. 112, 353382. expression in kidney proximal tubular cells. Am. J. Physiol. Renal Physiol.
Shamir, E.R., Pappalardo, E., Jorgens, D.M., Coutinho, K., Tsai, W.T., Aziz, K., 304, F1266F1273.
Auer, M., Tran, P.T., Bader, J.S., and Ewald, A.J. (2014). Twist1-induced Tang, Q., Zhong, H., Xie, F., Xie, J., Chen, H., and Yao, G. (2014). Expression of
dissemination preserves epithelial identity and requires E-cadherin. J. Cell miR-106b-25 induced by salvianolic acid B inhibits epithelial-to-mesenchymal
Biol. 204, 839856. transition in HK-2 cells. Eur. J. Pharmacol. 741, 97103.
Shaul, Y.D., Freinkman, E., Comb, W.C., Cantor, J.R., Tam, W.L., Thiru, P., Tarin, D., Thompson, E.W., and Newgreen, D.F. (2005). The fallacy of epithelial
Kim, D., Kanarek, N., Pacold, M.E., Chen, W.W., et al. (2014). Dihydropyrimi- mesenchymal transition in neoplasia. Cancer Res. 65, 59966001.
dine accumulation is required for the epithelial-mesenchymal transition. Cell Taura, K., Miura, K., Iwaisako, K., Osterreicher, C.H., Kodama, Y., Penz-Oster-
158, 10941109. reicher, M., and Brenner, D.A. (2010). Hepatocytes do not undergo epithelial-
Shaw, T.J., and Martin, P. (2016). Wound repair: a showcase for cell plasticity mesenchymal transition in liver fibrosis in mice. Hepatology 51, 10271036.
and migration. Curr. Opin. Cell Biol. 42, 2937.
Terao, M., Ishikawa, A., Nakahara, S., Kimura, A., Kato, A., Moriwaki, K.,
Sheridan, C. (2013). First Axl inhibitor enters clinical trials. Nat. Biotechnol. 31, Kamada, Y., Murota, H., Taniguchi, N., Katayama, I., and Miyoshi, E. (2011).
775776. Enhanced epithelial-mesenchymal transition-like phenotype in N-acetylgluco-
Shimono, Y., Zabala, M., Cho, R.W., Lobo, N., Dalerba, P., Qian, D., Diehn, M., saminyltransferase V transgenic mouse skin promotes wound healing. J. Biol.
Liu, H., Panula, S.P., Chiao, E., et al. (2009). Downregulation of miRNA-200c Chem. 286, 2830328311.
links breast cancer stem cells with normal stem cells. Cell 138, 592603. Theveneau, E., and Mayor, R. (2011). Can mesenchymal cells undergo collec-
Shoval, I., Ludwig, A., and Kalcheim, C. (2007). Antagonistic roles of full-length tive cell migration? The case of the neural crest. Cell Adhes. Migr. 5, 490498.
N-cadherin and its soluble BMP cleavage product in neural crest delamination. Thiery, J.P. (2002). Epithelial-mesenchymal transitions in tumour progression.
Development 134, 491501. Nat. Rev. Cancer 2, 442454.
Simoes-Costa, M., and Bronner, M.E. (2015). Establishing neural crest identity: Thiery, J.P., and Lim, C.T. (2013). Tumor dissemination: an EMT affair. Cancer
a gene regulatory recipe. Development 142, 242257. Cell 23, 272273.
Singh, A., and Settleman, J. (2010). EMT, cancer stem cells and drug resis- Thiery, J.P., and Sleeman, J.P. (2006). Complex networks orchestrate epithe-
tance: an emerging axis of evil in the war on cancer. Oncogene 29, 47414751. lial-mesenchymal transitions. Nat. Rev. Mol. Cell Biol. 7, 131142.
Smerage, J.B., Barlow, W.E., Hortobagyi, G.N., Winer, E.P., Leyland-Jones, Thiery, J.P., Acloque, H., Huang, R.Y., and Nieto, M.A. (2009). Epithelial-
B., Srkalovic, G., Tejwani, S., Schott, A.F., ORourke, M.A., Lew, D.L., et al. mesenchymal transitions in development and disease. Cell 139, 871890.
(2014). Circulating tumor cells and response to chemotherapy in metastatic
Tian, X.J., Zhang, H., and Xing, J. (2013). Coupled reversible and irreversible
breast cancer: SWOG S0500. J. Clin. Oncol. 32, 34833489.
bistable switches underlying TGFb-induced epithelial to mesenchymal transi-
Sosa, M.S., Parikh, F., Maia, A.G., Estrada, Y., Bosch, A., Bragado, P., Ekpin, tion. Biophys. J. 105, 10791089.
E., George, A., Zheng, Y., Lam, H.M., et al. (2015). NR2F1 controls tumour
Ting, S.B., Caddy, J., Hislop, N., Wilanowski, T., Auden, A., Zhao, L.L., Ellis, S.,
cell dormancy via SOX9- and RARb-driven quiescence programmes. Nat.
Kaur, P., Uchida, Y., Holleran, W.M., et al. (2005). A homolog of Drosophila
Commun. 6, 6170.
grainy head is essential for epidermal integrity in mice. Science 308, 411413.
Spaderna, S., Schmalhofer, O., Wahlbuhl, M., Dimmler, A., Bauer, K., Sultan,
Ting, D.T., Wittner, B.S., Ligorio, M., Vincent Jordan, N., Shah, A.M.,
A., Hlubek, F., Jung, A., Strand, D., Eger, A., et al. (2008). The transcriptional
Miyamoto, D.T., Aceto, N., Bersani, F., Brannigan, B.W., Xega, K., et al.
repressor ZEB1 promotes metastasis and loss of cell polarity in cancer.
(2014). Single-cell RNA sequencing identifies extracellular matrix gene
Cancer Res. 68, 537544.
expression by pancreatic circulating tumor cells. Cell Rep. 8, 19051918.
Stankic, M., Pavlovic, S., Chin, Y., Brogi, E., Padua, D., Norton, L., Massague,
Topalian, S.L., Hodi, F.S., Brahmer, J.R., Gettinger, S.N., Smith, D.C.,
J., and Benezra, R. (2013). TGF-b-Id1 signaling opposes Twist1 and promotes
McDermott, D.F., Powderly, J.D., Carvajal, R.D., Sosman, J.A., Atkins, M.B.,
metastatic colonization via a mesenchymal-to-epithelial transition. Cell Rep. 5,
et al. (2012). Safety, activity, and immune correlates of anti-PD-1 antibody in
12281242.
cancer. N. Engl. J. Med. 366, 24432454.
Sugimoto, H., LeBleu, V.S., Bosukonda, D., Keck, P., Taduri, G., Bechtel, W.,
Tran, H.D., Luitel, K., Kim, M., Zhang, K., Longmore, G.D., and Tran, D.D.
Okada, H., Carlson, W., Jr., Bey, P., Rusckowski, M., et al. (2012). Activin-like
(2014). Transient SNAIL1 expression is necessary for metastatic competence
kinase 3 is important for kidney regeneration and reversal of fibrosis. Nat. Med.
in breast cancer. Cancer Res. 74, 63306340.
18, 396404.
Trimboli, A.J., Fukino, K., de Bruin, A., Wei, G., Shen, L., Tanner, S.M.,
Takahashi, K., Tanabe, K., Ohnuki, M., Narita, M., Sasaki, A., Yamamoto, M.,
Creasap, N., Rosol, T.J., Robinson, M.L., Eng, C., et al. (2008). Direct evidence
Nakamura, M., Sutou, K., Osafune, K., and Yamanaka, S. (2014). Induction of
for epithelial-mesenchymal transitions in breast cancer. Cancer Res. 68,
pluripotency in human somatic cells via a transient state resembling primitive
937945.
streak-like mesendoderm. Nat. Commun. 5, 3678.
Tsai, J.H., Donaher, J.L., Murphy, D.A., Chau, S., and Yang, J. (2012). Spatio-
Takeichi, M. (1988). The cadherins: cell-cell adhesion molecules controlling
temporal regulation of epithelial-mesenchymal transition is essential for squa-
animal morphogenesis. Development 102, 639655.
mous cell carcinoma metastasis. Cancer Cell 22, 725736.
Tam, W.L., and Weinberg, R.A. (2013). The epigenetics of epithelial-mesen-
chymal plasticity in cancer. Nat. Med. 19, 14381449. Tsuji, T., Ibaragi, S., and Hu, G.F. (2009). Epithelial-mesenchymal transition
and cell cooperativity in metastasis. Cancer Res. 69, 71357139.
Tan, T.Z., Miow, Q.H., Huang, R.Y., Wong, M.K., Ye, J., Lau, J.A., Wu, M.C.,
Bin Abdul Hadi, L.H., Soong, R., Choolani, M., et al. (2013). Functional geno- Ubil, E., Duan, J., Pillai, I.C., Rosa-Garrido, M., Wu, Y., Bargiacchi, F., Lu, Y.,
mics identifies five distinct molecular subtypes with clinical relevance and Stanbouly, S., Huang, J., Rojas, M., et al. (2014). Mesenchymal-endothelial
pathways for growth control in epithelial ovarian cancer. EMBO Mol. Med. 5, transition contributes to cardiac neovascularization. Nature 514, 585590.
983998. Vaughan, A.E., and Chapman, H.A. (2013). Regenerative activity of the lung
Tan, T.Z., Miow, Q.H., Miki, Y., Noda, T., Mori, S., Huang, R.Y., and Thiery, J.P. after epithelial injury. Biochim. Biophys. Acta 1832, 922930.
(2014). Epithelial-mesenchymal transition spectrum quantification and its Vega, S., Morales, A.V., Ocana, O.H., Valdes, F., Fabregat, I., and Nieto, M.A.
efficacy in deciphering survival and drug responses of cancer patients. (2004). Snail blocks the cell cycle and confers resistance to cell death. Genes
EMBO Mol. Med. 6, 12791293. Dev. 18, 11311143.

44 Cell 166, June 30, 2016


von Gise, A., and Pu, W.T. (2012). Endocardial and epicardial epithelial to Yang, M.C., Wang, C.J., Liao, P.C., Yen, C.J., and Shan, Y.S. (2014). Hepatic
mesenchymal transitions in heart development and disease. Circ. Res. 110, stellate cells secretes type I collagen to trigger epithelial mesenchymal transi-
16281645. tion of hepatoma cells. Am. J. Cancer Res. 4, 751763.
Voon, D.C., Wang, H., Koo, J.K., Chai, J.H., Hor, Y.T., Tan, T.Z., Chu, Y.S., Yang, H., Lu, X., Liu, Z., Chen, L., Xu, Y., Wang, Y., Wei, G., and Chen, Y.
Mori, S., and Ito, Y. (2013). EMT-induced stemness and tumorigenicity are (2015). FBXW7 suppresses epithelial-mesenchymal transition, stemness and
fueled by the EGFR/Ras pathway. PLoS ONE 8, e70427. metastatic potential of cholangiocarcinoma cells. Oncotarget 6, 63106325.
Vultur, A., Buettner, R., Kowolik, C., Liang, W., Smith, D., Boschelli, F., and Ye, X., and Weinberg, R.A. (2015). Epithelial-Mesenchymal Plasticity: A Central
Jove, R. (2008). SKI-606 (bosutinib), a novel Src kinase inhibitor, suppresses Regulator of Cancer Progression. Trends Cell Biol. 25, 675686.
migration and invasion of human breast cancer cells. Mol. Cancer Ther. 7,
Ye, X., Tam, W.L., Shibue, T., Kaygusuz, Y., Reinhardt, F., Ng Eaton, E., and
11851194.
Weinberg, R.A. (2015). Distinct EMT programs control normal mammary
Wagenblast, E., Soto, M., Gutierrez-Angel, S., Hartl, C.A., Gable, A.L., Maceli, stem cells and tumour-initiating cells. Nature 525, 256260.
A.R., Erard, N., Williams, A.M., Kim, S.Y., Dickopf, S., et al. (2015). A model of
breast cancer heterogeneity reveals vascular mimicry as a driver of metas- Yu, M., Smolen, G.A., Zhang, J., Wittner, B., Schott, B.J., Brachtel, E.,
tasis. Nature 520, 358362. Ramaswamy, S., Maheswaran, S., and Haber, D.A. (2009). A developmentally
regulated inducer of EMT, LBX1, contributes to breast cancer progression.
Wang, J., Scully, K., Zhu, X., Cai, L., Zhang, J., Prefontaine, G.G., Krones, A.,
Genes Dev. 23, 17371742.
Ohgi, K.A., Zhu, P., Garcia-Bassets, I., et al. (2007). Opposing LSD1 com-
plexes function in developmental gene activation and repression programmes. Yu, M., Bardia, A., Wittner, B.S., Stott, S.L., Smas, M.E., Ting, D.T., Isakoff,
Nature 446, 882887. S.J., Ciciliano, J.C., Wells, M.N., Shah, A.M., et al. (2013). Circulating breast
tumor cells exhibit dynamic changes in epithelial and mesenchymal composi-
Wang, C., Song, X., Li, Y., Han, F., Gao, S., Wang, X., Xie, S., and Lv, C. (2013).
tion. Science 339, 580584.
Low-dose paclitaxel ameliorates pulmonary fibrosis by suppressing TGF-b1/
Smad3 pathway via miR-140 upregulation. PLoS ONE 8, e70725. Zadran, S., Arumugam, R., Herschman, H., Phelps, M.E., and Levine, R.D.
Wang, J., Zhu, X., Hu, J., He, G., Li, X., Wu, P., Ren, X., Wang, F., Liao, W., (2014). Surprisal analysis characterizes the free energy time course of cancer
Liang, L., and Ding, Y. (2015). The positive feedback between Snail and cells undergoing epithelial-to-mesenchymal transition. Proc. Natl. Acad. Sci.
DAB2IP regulates EMT, invasion and metastasis in colorectal cancer. Onco- USA 111, 1323513240.
target 6, 2742727439. Zeisberg, M., Hanai, J., Sugimoto, H., Mammoto, T., Charytan, D., Strutz, F.,
Warzecha, C.C., and Carstens, R.P. (2012). Complex changes in alternative and Kalluri, R. (2003). BMP-7 counteracts TGF-beta1-induced epithelial-
pre-mRNA splicing play a central role in the epithelial-to-mesenchymal transi- to-mesenchymal transition and reverses chronic renal injury. Nat. Med. 9,
tion (EMT). Semin. Cancer Biol. 22, 417427. 964968.
Warzecha, C.C., Sato, T.K., Nabet, B., Hogenesch, J.B., and Carstens, R.P. Zeisberg, E.M., Tarnavski, O., Zeisberg, M., Dorfman, A.L., McMullen, J.R.,
(2009). ESRP1 and ESRP2 are epithelial cell-type-specific regulators of Gustafsson, E., Chandraker, A., Yuan, X., Pu, W.T., Roberts, A.B., et al.
FGFR2 splicing. Mol. Cell 33, 591601. (2007a). Endothelial-to-mesenchymal transition contributes to cardiac
Watanabe, K., Villarreal-Ponce, A., Sun, P., Salmans, M.L., Fallahi, M., Ander- fibrosis. Nat. Med. 13, 952961.
sen, B., and Dai, X. (2014). Mammary morphogenesis and regeneration require Zeisberg, M., Yang, C., Martino, M., Duncan, M.B., Rieder, F., Tanjore, H., and
the inhibition of EMT at terminal end buds by Ovol2 transcriptional repressor. Kalluri, R. (2007b). Fibroblasts derive from hepatocytes in liver fibrosis via
Dev. Cell 29, 5974. epithelial to mesenchymal transition. J. Biol. Chem. 282, 2333723347.
Wellner, U., Schubert, J., Burk, U.C., Schmalhofer, O., Zhu, F., Sonntag, A., Zhang, K., Rodriguez-Aznar, E., Yabuta, N., Owen, R.J., Mingot, J.M., Nojima,
Waldvogel, B., Vannier, C., Darling, D., zur Hausen, A., et al. (2009). The H., Nieto, M.A., and Longmore, G.D. (2012a). Lats2 kinase potentiates Snail1
EMT-activator ZEB1 promotes tumorigenicity by repressing stemness-inhibit- activity by promoting nuclear retention upon phosphorylation. EMBO J. 31,
ing microRNAs. Nat. Cell Biol. 11, 14871495. 2943.
Weston, J.A., and Thiery, J.P. (2015). Pentimento: Neural Crest and the origin Zhang, Y., Wang, Z., Yu, J., Shi, Jz., Wang, C., Fu, Wh., Chen, Zw., and Yang,
of mesectoderm. Dev. Biol. 401, 3761. J. (2012b). Cancer stem-like cells contribute to cisplatin resistance and pro-
Wettstein, G., Bellaye, P.S., Kolb, M., Hammann, A., Crestani, B., Soler, P., gression in bladder cancer. Cancer Lett. 322, 7077.
Marchal-Somme, J., Hazoume, A., Gauldie, J., Gunther, A., et al. (2013). Inhi-
Zhang, J., Tian, X.J., Zhang, H., Teng, Y., Li, R., Bai, F., Elankumaran, S., and
bition of HSP27 blocks fibrosis development and EMT features by promoting
Xing, J. (2014). TGF-b-induced epithelial-to-mesenchymal transition proceeds
Snail degradation. FASEB J. 27, 15491560.
through stepwise activation of multiple feedback loops. Sci. Signal. 7, ra91.
Wu, C.Y., Tsai, Y.P., Wu, M.Z., Teng, S.C., and Wu, K.J. (2012). Epigenetic
Zhao, J., Tang, N., Wu, K., Dai, W., Ye, C., Shi, J., Zhang, J., Ning, B., Zeng, X.,
reprogramming and post-transcriptional regulation during the epithelial-
and Lin, Y. (2014). MiR-21 simultaneously regulates ERK1 signaling in HSC
mesenchymal transition. Trends Genet. 28, 454463.
activation and hepatocyte EMT in hepatic fibrosis. PLoS ONE 9, e108005.
Wyckoff, J., Wang, W., Lin, E.Y., Wang, Y., Pixley, F., Stanley, E.R., Graf, T.,
Pollard, J.W., Segall, J., and Condeelis, J. (2004). A paracrine loop between Zhao, S., Zhang, Y., Zheng, X., Tu, X., Li, H., Chen, J., Zang, Y., and Zhang, J.
tumor cells and macrophages is required for tumor cell migration in mammary (2015). Loss of MicroRNA-101 Promotes Epithelial to Mesenchymal Transition
tumors. Cancer Res. 64, 70227029. in Hepatocytes. J. Cell. Physiol. 230, 27062717.

Xiao, W., Chen, X., Li, W., Ye, S., Wang, W., Luo, L., and Liu, Y. (2015). Zheng, X., Carstens, J.L., Kim, J., Scheible, M., Kaye, J., Sugimoto, H., Wu,
Quantitative analysis of injury-induced anterior subcapsular cataract in the C.C., LeBleu, V.S., and Kalluri, R. (2015). Epithelial-to-mesenchymal transition
mouse: a model of lens epithelial cells proliferation and epithelial-mesen- is dispensable for metastasis but induces chemoresistance in pancreatic
chymal transition. Sci. Rep. 5, 8362. cancer. Nature 527, 525530.
Yang, F., Huang, X.R., Chung, A.C., Hou, C.C., Lai, K.N., and Lan, H.Y. (2010a). Zhou, B.P., Deng, J., Xia, W., Xu, J., Li, Y.M., Gunduz, M., and Hung, M.C.
Essential role for Smad3 in angiotensin II-induced tubular epithelial-mesen- (2004). Dual regulation of Snail by GSK-3beta-mediated phosphorylation in
chymal transition. J. Pathol. 221, 390401. control of epithelial-mesenchymal transition. Nat. Cell Biol. 6, 931940.
Yang, X., Lin, X., Zhong, X., Kaur, S., Li, N., Liang, S., Lassus, H., Wang, L., Kat- Zhou, F., Drabsch, Y., Dekker, T.J., de Vinuesa, A.G., Li, Y., Hawinkels, L.J.,
saros, D., Montone, K., et al. (2010b). Double-negative feedback loop between Sheppard, K.A., Goumans, M.J., Luwor, R.B., de Vries, C.J., et al. (2014).
reprogramming factor LIN28 and microRNA let-7 regulates aldehyde dehydro- Nuclear receptor NR4A1 promotes breast cancer invasion and metastasis
genase 1-positive cancer stem cells. Cancer Res. 70, 94639472. by activating TGF-b signalling. Nat. Commun. 5, 3388.

Cell 166, June 30, 2016 45

You might also like