Mohamed 2015

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Biochimica et Biophysica Acta 1856 (2015) 121–129

Contents lists available at ScienceDirect

Biochimica et Biophysica Acta

journal homepage: www.elsevier.com/locate/bbacan

Review

The Hippo signal transduction pathway in soft tissue sarcomas


Abdalla D. Mohamed a, Annie M. Tremblay c,d,e, Graeme I. Murray b, Henning Wackerhage a,⁎
a
School of Medical Sciences, University of Aberdeen, AB25 2ZD Scotland, UK
b
School of Medicine and Dentistry, University of Aberdeen, AB25 2ZD Scotland, UK
c
Stem Cell Program, Children's Hospital, Boston, MA 02115, USA
d
Department of Stem Cell and Regenerative Biology, Harvard University, Cambridge, MA 02138, USA
e
Harvard Stem Cell Institute, Cambridge, MA 02138, USA

a r t i c l e i n f o a b s t r a c t

Article history: Sarcomas are rare cancers (≈1% of all solid tumours) usually of mesenchymal origin. Here, we review evidence
Received 10 April 2015 implicating the Hippo pathway in soft tissue sarcomas. Several transgenic mouse models of Hippo pathway
Received in revised form 27 May 2015 members (Nf2, Mob1, LATS1 and YAP1 mutants) develop various types of sarcoma. Despite that, Hippo member
Accepted 28 May 2015
genes are rarely point mutated in human sarcomas. Instead, WWTR1-CAMTA1 and YAP1-TFE3 fusion genes are
Available online 4 June 2015
found in almost all cases of epithelioid haemangioendothelioma. Also copy number gains of YAP1 and other
Keywords:
Hippo members occur at low frequencies but the most likely cause of perturbed Hippo signalling in sarcoma is
Sarcoma the cross-talk with commonly mutated cancer genes such as KRAS, PIK3CA, CTNNB1 or FBXW7. Current Hippo
Rhabdomyosarcoma pathway-targeting drugs include compounds that target the interaction between YAP and TEAD G protein-
Hippo pathway coupled receptors (GPCR) and the mevalonate pathway (e.g. statins). Given that many Hippo pathway-
YAP modulating drugs are already used in patients, this could lead to early clinical trials testing their efficacy in differ-
TAZ ent types of sarcoma.
Fusion genes Crown Copyright © 2015 Published by Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
2. Pathology of soft tissue sarcomas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
3. Hippo pathway & cancer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
4. Hippo mouse models that develop sarcomas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
5. Genetics of soft tissue sarcoma and the Hippo pathway. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
5.1. Hippo fusion genes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
5.2. Hippo copy number alterations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
5.3. Epigenetic regulation of Hippo genes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
5.4. Hippo signalling and the genomic landscape of sarcoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
6. Hippo-targeted therapies in soft tissue sarcoma?. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
6.1. Drugs that target YAP-TEAD interaction or TEADs: verteporfin, cyclic peptides, VGLL4-mimicking peptides . . . . . . . . . . . . . . . . 126
6.2. G protein-coupled receptors (GPCRs) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
6.3. Statins and bisphosphonates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
7. Summary and perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
Transparency document . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
Funding. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction be conserved in mammals. In flies, inactivation of Hippo members re-


sulted in an overgrowth that resembles the skin of a Hippopotamus,
The Hippo pathway was discovered as a result of tumour suppressor thereby naming the pathway [1]. Hippo pathway members are rarely
screens in the fruit fly (Drosophila melanogaster) and was later found to point mutated in cancer [2,3] but the experimental mutation of Hippo
members usually causes overgrowth in fruit flies [4] and tumours in
⁎ Corresponding author. mice, including sarcomas [2,3,5]. Here, we review the role of the Hippo
E-mail address: h.wackerhage@abdn.ac.uk (H. Wackerhage). pathway specifically in soft tissue sarcomas excluding osteosarcomas

http://dx.doi.org/10.1016/j.bbcan.2015.05.006
0304-419X/Crown Copyright © 2015 Published by Elsevier B.V. All rights reserved.
122 A.D. Mohamed et al. / Biochimica et Biophysica Acta 1856 (2015) 121–129

and viral-mediated sarcomas such as Kaposi's sarcoma. After introducing embryonal subtype, tends to occur predominantly in young children
sarcomas and the Hippo pathway, we review animal models where [21]. Most soft tissue sarcomas arise sporadically but muscle injury,
Hippo pathway-related transgenesis results in sarcomas. We then dis- most likely by increasing the number of activated satellite cells, en-
cuss findings and hypotheses implicating the Hippo pathway in sarcoma hances the penetrance and shortens the latency of rhabdomyosarcoma
signal transduction, development and pathology. Finally, we review the phenotypes in mice [18,22]. In addition, known risk factors for
current possibilities for Hippo pathway-targeting treatments in sarcoma the development of specific types of sarcoma in humans include expo-
and highlight areas for future research. sure to radiation (e.g. post-radiation angiosarcoma) or exposure to en-
vironmental/occupational carcinogens (e.g. vinyl chloride-associated
angiosarcoma). Sarcomas, in contrast to carcinomas, have a predilection
2. Pathology of soft tissue sarcomas for showing a vascular pattern of spread with the lungs generally being
one of the predominant sites of metastasis.
Soft tissue sarcomas (from Greek sarx flesh) represent a biologically, A combination of surgery, cytotoxic chemotherapy and radiotherapy
clinically, pathologically, and genetically [6] diverse array of malignant are the mainstays of current treatment with some types of sarcoma
tumours arising mostly from mesenchyme-derived tissues. Sarcomas being treated with neoadjuvant chemo-radiotherapy prior to definitive
can occur in both adults and children. They are rare cancers and account surgery [23–25]. Immunotherapy and targeted, novel biologic therapies
for about 1% of solid tumours in adults and for a significantly higher pro- are also now being evaluated in the treatment of specific types of sarco-
portion of cases in children (Cancer Research UK, 2015). In 2014, it was ma [26–29] and as we show below, targeting the Hippo pathway in sar-
estimated that 790 new cases of rhabdomyosarcoma and bone tumours, coma should already be possible with existing drugs.
including osteosarcomas and Ewing sarcomas, would be diagnosed in The prognosis and outcome of soft tissue sarcomas depend on a
children 0–14 years old, representing 7% of all childhood cancers range of factors including the particular subtype of tumour, the anatom-
(American Cancer Society, 2014). The most common types of soft tissue ical location, size and grade as well as tumour stage at time of diagnosis
sarcomas include leiomyosarcoma, liposarcoma, fibrosarcoma, rhabdo- [7,30,31]. The overall 5-year survival rate for soft tissue sarcomas has
myosarcoma and angiosarcoma (see Fig. 1 [7]). gradually been improving, especially in children, but the outcome varies
The current classification of soft tissue sarcomas is based on a com- with the specific type of soft tissue sarcoma and depends on the prog-
bination of tumour morphology, immunophenotype and molecular pa- nostic factors that have been outlined above. Overall the 5-year survival
thology. Genetically, sarcomas can show either aberrant, chimeric rate for adults with soft tissue sarcomas is approximately 60% while it is
transcription regulators as a consequence of fusion genes such as higher in children, reaching approximately 70% (Cancer Research UK,
PAX3/7-FOXO1 [6,8], somatic point mutations of well-known cancer 2015).
genes such as oncogenic RAS isoforms, PIK3CA or TP53, or DNA copy
number gains or losses [6,9–12]. As we will show later, there is no evi- 3. Hippo pathway & cancer
dence for recurrent Hippo gene point mutations in sarcomas, while
there is abundant evidence for mutations of typical cancer genes [13, The fruit fly (D. melanogaster) has been used over many decades as a
14] that can cross-talk to the main members of the Hippo pathway. Ad- model organism to identify genes whose knockout results in cancerous
ditionally, fusion genes involving WWTR1 or to a lesser extent YAP1 are growth [32]. This research has led to the discovery of a set of genes that
found in nearly all cases of epithelioid haemangioendothelioma [15,16], encode two interacting kinases and auxiliary proteins, now defined as
whereas YAP1 and VGLL3 copy number gains have been reported for the core Hippo pathway (see Fig. 2). The main function of the Hippo
some types of sarcoma, especially rhabdomyosarcoma [17–19]. It is un- pathway is to inhibit proliferation and to promote apoptosis, thereby
clear to this date whether VGLL3, which is associated with tumour sup- limiting organ growth [4]. In the conserved mammalian Hippo pathway,
pression in ovarian cancer [20], is a bona fide Hippo pathway member. the STE20-like protein kinases 1 and 2 (MST1 and MST2, gene symbols:
The increasing use of immunohistochemistry and the application of STK4 and STK3) regulate the large tumour suppressor kinases 1 and 2
sophisticated molecular techniques combined with a better under- (protein name and gene symbol: LATS1 and LATS2), through phosphor-
standing of soft tissue sarcoma biology have led to a continued refine- ylation [33]. Active LATS1 and LATS2 then interact through their PPxY
ment of the classification of soft tissue sarcomas. Some previously motifs with the WW domains of the transcriptional co-factors YAP
recognised types of soft tissue sarcoma, most notably malignant fibrous (gene symbol YAP1) or TAZ (gene symbol WWTR1; note that the gene
histiocytoma, are now being reclassified. Most tumours previously clas- TAZ encodes a protein termed Tafazzin which is not part of the path-
sified as malignant fibrous histiocytomas but showing no specific way) [34]. This physical contact allows LATS1 and LATS2 to inhibit
immunophenotype would now be regarded as undifferentiated pleo- YAP [35] and TAZ [36] through the phosphorylation of multiple
morphic sarcomas. The majority of soft tissue sarcomas arise in the HXRXXS amino acid motifs. The phosphorylation of these motifs pro-
limbs of older people although rhabdomyosarcoma, especially the motes the inactivation of YAP and TAZ through translocation from the
nucleus into the cytosol and degradation. Additionally, YAP can be phos-
phorylated at Tyr357 by the tyrosine kinase YES1, which has resulted in
Dedifferentiated
Pleomorphic Myxoid its name Yes-associated protein (YAP) [37]. Nuclear and active YAP,
Well differentiated
which was first discovered by Marius Sudol [38,39], and its paralogue
Liposarcoma Biphasic TAZ is believed to exert their tumourigenic functions mainly via the
Gastrointestinal
(adipocytes) Synovial sarcoma TEAD transcription factors [40]. Specifically, YAP and presumably TAZ
stromal tumour (GIST) Spindle cell
Epithelioid haemangio-
de-repress and activate the TEAD transcription factors that otherwise
endothelioma recruit transcriptional repressors [41]. Additionally, YAP and TAZ are ca-
Leiomyosarcoma Soft tissue sarcoma Vascular
(smooth muscle) Angiosarcoma pable of co-regulating other transcription factors including those be-
Kaposi’s sarcoma
longing to the Smad family [42] and Tbx5 in some contexts [37,43]. In
Rhabdomyosarcoma
(myoblasts)
addition to the Hippo kinases, extensive cross-talk mechanisms also
Fibrosarcoma regulate the activity of YAP and TAZ, notably mechanotransduction
(fibroblasts)
Low-grade [44], WNT signalling [45,46], and G protein-coupled receptors [47].
Embryonal fibromyxoid sarcoma
Spindle cell The early studies in fruit flies and subsequent studies in mammals
Alveolar
Pleomorphic Myxofibrosaroma demonstrate that the upstream proliferation-inhibiting Hippo proteins
and the proliferation-promoting Hippo transcriptional regulators
Fig. 1. Classification of soft tissue sarcomas on the basis of their differentiation. act as potent tumour suppressors and oncogenes, respectively. For
A.D. Mohamed et al. / Biochimica et Biophysica Acta 1856 (2015) 121–129 123

Fig. 2. Schematic drawing depicting proteins associated with the Hippo pathway in sarcoma. A The Hippo kinases MST1/2 and LATS1/2 together with the auxiliary proteins SAV1 and Mob1
inhibit the transcriptional co-factors YAP and TAZ by phosphorylation of HXRXXS motifs such as Ser127 on YAP. B YAP and TAZ dephosphorylation by phosphatases results in their nuclear
translocation, allowing the de-repression or activation of TEAD1-4 transcription factors. In some contexts, VGLL4 acts as a co-repressor of TEAD1-4 [41] and other VGLL isoforms are likely
to be capable of interaction with TEADs [105]. C–G Links between the Hippo pathway and proteins whose genes are recurrently mutated in sarcoma (shown in red) C In epithelioid
haemangioendothelioma, mostly the WWTR1-CAMTA1 [16] but also YAP1-TFE3 [15] fusion genes (encoding TAZ-CAMTA1 and YAP-TFE3 fusion proteins) presumably regulate the TEAD
transcription factors. D The Wnt-regulator β-CATENIN (CTNNB1) can also interact with YAP and TAZ. It has been proposed that β-CATENIN, YAP, the tyrosine kinase YES1 and the tran-
scription factor TBX5 form complexes of functional importance in cancer. In this context, the tyrosine kinase YES1 can activate YAP by Tyr357 phosphorylation [37]. E KRAS and PI3K
can enhance the co-activator function of YAP, or their activity can be substituted by YAP to overcome oncogene addiction, identifying YAP as a modulator of drug resistance. F In alveolar
rhabdomyosarcoma the PAX3-FOXO1 fusion proteins/genes drive the expression of RASSF4 which inhibits MST1 [84]. Not shown: Others suggest that YAP and TAZ are included in the Wnt
destruction complex [46]. G The SCF (βTRCP) ubiquitin ligase regulates YAP [35] and especially TAZ [94] degradation via phosphodegron-mediated mechanisms. Recently another ubiq-
uitin ligase named FBXW7, which is frequently mutated in rhabdomyosarcoma [73], was identified as part of a complex that mediates ubiquitin-dependent destruction of YAP and TAZ and
its loss was associated with increased YAP abundance [95]. Drugs and drug targets that modulate the activity of the Hippo pathway are shown as red text. Verteporfin [97] and cyclic pep-
tides [106] inhibit the binding of YAP to TEAD transcription factors. VGLL4-mimicking peptides repress TEAD transcription factors [79]. Specific G protein-coupled receptors (GPCRs) signal
to YAP [47]. Given that GPCR-targeting drugs are the largest class of currently used drugs some might be applicable for Hippo activity modulation in sarcomas. The activity of YAP and TAZ is
regulated by the mevalonate pathway [102,103], which can itself be targeted for example by statins. For more Hippo pathway related drugs see [107].

example, YAP hyperactivity as a consequence of the sole expression of found in early replicating chromatin [55]. Thus, Hippo genes may lie in
the constitutively active YAP1 S127A mutant in mice alone is sufficient chromatin with a low mutation rate perhaps due to their functional im-
to initiate hyperplasia and over time cause cancer in the liver [48,49] portance. Regarding the second question, alternative mechanisms of
and skin [50]. Also, along with injury in the skeletal muscle, YAP1 YAP and TAZ activation in cancer include Hippo fusion genes, copy num-
S127A expression in satellite cells alone is sufficient to causes rhabdo- ber alterations of Hippo genes, hypermethylation of Hippo kinase gene
myosarcoma [18]. These studies identify YAP as an unusually potent on- promoters and cross-talk from significantly mutated non-Hippo cancer
cogene. Thus it would be intuitive that YAP-activating mutations such as genes. Most prominently, Hippo fusion genes have been identified for
YAP1 S127A occur by chance and be a driver of tumourigenesis. Howev- almost all cases of epithelioid haemangioendothelioma [15,16], a vascu-
er, this is not the case. So far, even if the LATS kinases appear to be mu- lar sarcoma that will be discussed below. Copy number gains have been
tated in some rare cases [51], recurrent mutations of Hippo pathway reported for YAP1 and VGLL3 [17,18,56] and frequently mutated cancer
genes have not been reported [2,3,5,13]. It is even more surprising genes such as CTNNB1 (protein: β-catenin [37] or KRAS [57–59] have
that not a single YAP1 S127A mutation was found in 21,441 tumour been shown to affect YAP activity and require YAP for some aspects of
DNA samples from 91 cancer studies that are compiled in the cBIOPortal tumourigenesis.
database [52,53].
From this conundrum two questions arise: First, why are Hippo 4. Hippo mouse models that develop sarcomas
genes, unlike other well established cancer-associated genes such as
TP53, PIK3CA or KRAS [13], rarely point mutated in cancer? Second, if Important direct evidence for a role of the Hippo pathway in sarco-
point mutations are not major the source of tumourigenic Hippo signal- ma stems from transgenic Hippo mouse models that develop sarcoma.
ling, what else is causing the high YAP activity frequently seen in Such models are summarised in Table 1 and discussed in greater detail
cancers? A potential answer to the first question is that the somatic below.
(i.e. cancer cell) mutation frequency differs at least 100-fold in- As discussed earlier, the Hippo pathway kinase Lats1 inhibits YAP
between different types of cancer and also varies across the genome and TAZ by phosphorylation and restricts their nuclear localization
from ≈ 0.1 to ≈ 10 mutations per 100,000,000 bases depending on and transcriptional activity with the TEAD factors. Most Lats−/− mice
DNA replication timing and transcriptional activity [54,55]. There is ev- die in utero, and only ~8% survive. The surviving mice display a strong
idence that copy number gains and other rearrangements are especially growth retardation phenotype, severe fertility defects and pituitary
124 A.D. Mohamed et al. / Biochimica et Biophysica Acta 1856 (2015) 121–129

Table 1
Hippo transgenic animal models that lead to the development of sarcomas.

Genotype Sarcoma type Incidence Latency Reference

Lats1−/− Skin fibrosarcoma 14% 4–10 months [60]


Lats1−/− and mutagen (DMBA+ UVB exposure) Skin fibrosarcoma 83% 10–14 weeks after birth and treatment [60]
Nf2/Merlin +/− Osteosarcoma*** 63% 10–30 months [61]
Fibrosarcoma* 9% 10–30 months
Chondrosarcoma Low frequency Not determined
Uterine sarcoma Low frequency Not determined
Nf2/Merlin +/− p53 +/− (cis)** Osteosarcoma*** 77% b5 months [61]
Fibrosarcoma* 32% b5 months
Nf2/Merlin +/− p53 +/− (trans)** Osteosarcoma*** 74% 4–21 months [61]
Fibrosarcoma* 8% 4–21 months
Mob1aΔ/+1btr/tr or Mob1aΔ/Δ1btr/+ Extraskeletal osteosarcoma, fibrosarcoma† 24% 25–70 weeks [62]
22% 25–70 weeks
Pax7CreER TetO-YAP1 S127A and cardiotoxin injury Embryonal rhabdomyosarcoma 100% 10–11 weeks after cardiotoxin injury [18]

*Of note: In this study, the tumours were reported to also contain areas presenting features of rhabdomyosarcoma within the same tumours, but were classified predominantly as
fibrosarcomas.
**p53 mutations were introduced on the same allele as the NF2/Merlin mutation (in cis), or on the other allele (in trans).
***In nearly all sarcomas, loss of the other wild-type NF2/Merlin allele was observed in the tumours (Loss of Heterozygosity, LOH);

or myofibrosarcoma.

hyperplasia leading to hormonal impairment. Indeed, Lats1−/− Mob1b (Mob1aΔ/Δ1btr/+) survive. Of these heterozygotes, 52% develop
females have reduced levels of luteinizing hormone, prolactin and dental malocclusion, 25% have a disorganization of the inner ear hair
growth hormone. All Lats1−/− females have underdeveloped mammary bundles in the cochlear organ of Corti resulting in a disequilibrium
glands and develop ovarian stromal cell tumours by 3 months of age. and 31% display increased trabeculae in femurs [62]. In addition, 100%
Large skin fibrosarcomas develop in 14% of all Lats1−/− females be- of the Mob1aΔ/+1btr/tr and Mob1aΔ/Δ1btr/+ heterozygotes mice sponta-
tween 4-10 months of age. The incidence of skin fibrosarcomas in- neously develop various types of tumours, while only 4% of the control
creases to 83% when LATS1−/− mice are subjected to a mutagen mice (heterozygotes for both genes; Mob1aΔ/+1btr/+) developed tu-
treatment consisting of a single application of 9,10-dimethyl-1,2-benz- mours [62]. All single heterozygote mice developed skin cancer
anthracene (DMBA) followed by repeated exposure to UVB, whereas all (100%), 92% developed benign bony overgrowths called exostoses,
Lats1+/− animals remained tumour-free [60]. 24% developed extraskeletal osteosarcomas and 22% developed subcu-
Neurofibromatosis type II is a dominantly inherited disorder caused taneous fibrosarcomas or myofibrosarcomas. These mice also devel-
by mutations in the neurofibromin 2 (gene symbol Nf2) gene, encoding oped other tumour types at lower incidence, such as breast (16%),
a protein called Merlin, also known as schwannomin. In both drosophila lung (5%) and salivary gland (5%) tumours. Interestingly, while 50% of
and mammals, Nf2/Merlin acts upstream of the Hippo kinases to control the Mob1aΔ/Δ1btr/+ mice developed liver tumours, not a single liver tu-
the transcriptional activity of Yap [63,64]. Mutations in the Nf2 gene mour was found in the Mob1aΔ/+1btr/tr mice [62].
identified in Neurofibromatosis type II patients lead to the production YAP, along with its paralogue TAZ, is the main transducer of the
of shorter protein products that have lost the tumour suppressor Hippo pathway activity via its de-repression and activation of TEAD
function associated with full-length Merlin. Consequently, neurofibro- transcription factors. Specifically in activated but not quiescent satellite
matosis type II patients are predisposed to developing tumours affecting cells, the overexpression of YAP1 S127A alone is sufficient to cause em-
the nervous system, primarily schwannomas, meningiomas and bryonal rhabdomyosarcomas in 100% of the mice [18], in line with the
ependymomas [61]. However, mice expressing a mutated Nf2 allele, pro-proliferative and anti-differentiation roles of YAP in myoblasts
originally intended to mimic human Neurofibromatosis type II in and primary satellite cells in culture [66,67]. The mouse tumours
mice, do not develop schwannomas, meningiomas or ependymomas. match human ERMS well as judged by their pathology and gene expres-
Instead, they principally develop osteosarcomas (63%), fibrosarcomas sion profile. Conversely, YAP1 knockdown in the human RD ERMS cell
(9%) and hepatocellular carcinomas (9%) occurring between 10 and line reduces their proliferation, soft agar growth and xenotransplant
30 months of age [61]. Notably, human neurofibromatosis type II pa- burden as well as promotes myogenic differentiation [18]. The fact
tients do not develop osteosarcomas, fibrosarcomas or hepatocellular that YAP1 S127A drives tumourigenesis only in activated but not quies-
carcinomas at higher incidence than the normal population [61]. Also, cent satellite cells matches clinical observations. First, ERMS occurs es-
chondrosarcomas and uterine sarcomas arose at low frequency in pecially in infants and young children where activated satellite cells
Nf2+/− mice, and in some cases loss of the wild-type Nf2 allele was de- support longitudinal muscle growth [21]. Additionally, mdx mice,
tected [61]. However, nearly all of the osteosarcomas and fibrosarcomas where many satellite cells are activated to regenerate damaged muscle
displayed a loss of the wild-type Nf2 allele (LOH), supporting a role for [68] spontaneously develop rhabdomyosarcomas [69]. Dystrophin and
the complete loss of Merlin function in the aetiology of those tumour dysferlin double-mutants are also emerging as potential models for
types. These tumours also displayed an enhanced metastatic potential rhabdomyosarcoma [70].
[61]. An additional hemizygous inactivation of p53 produced a similar
tumour spectrum but enhanced the incidence of sarcomas (osteosarco- 5. Genetics of soft tissue sarcoma and the Hippo pathway
mas, 77%; fibrosarcomas, 32%) and significantly reduced tumour-free
survival [61]. In this section we will review Hippo-related genetic changes in sar-
Mob1 proteins are core components of the Hippo tumour suppressor comas. Specifically, we discuss Hippo fusion genes, Hippo copy number
pathway that are required for activation of the LATS kinases by the MST changes, epigenetic alterations of Hippo genes and mutations that may
kinases [65]. Mice bearing a null mutation of the Mps one binder (Mob) affect Hippo members through cross-talk in sarcomas.
kinase activator 1A (Mob1aΔ/Δ) or a trapped mutation of the Mob1b
gene (Mob1btr/tr) show no abnormalities in morphology, body weight, 5.1. Hippo fusion genes
histology, or life span [62]. While the complete loss of both genes
(Mob1a and Mob1b; Mob1aΔ/Δ1btr/tr) is embryonic lethal, double- Fusion products involving Hippo genes have been associated
mutant mice retaining one allele of either Mob1a (Mob1aΔ/+1btr/tr) or especially with epithelioid haemangioma, a vascular sarcoma. In this
A.D. Mohamed et al. / Biochimica et Biophysica Acta 1856 (2015) 121–129 125

tumour, the WWTR1-CAMTA (WWTR1 is the gene symbol for TAZ) fu- of sarcomas can only partially explain why the balance is tipped to-
sion product was found in 17 cases independent of location [16]. A wards tumourigenesis.
follow-up analysis on the 10% of epithelioid haemangiomas that were
negative for the WWTR1-CAMTA identified those as a pathologically dis- 5.3. Epigenetic regulation of Hippo genes
tinct subset. In eight of these cases a YAP1-TFE3 fusion product was
found [15]. These findings were confirmed by a study on 35 epithelioid Epigenetically mediated loss of the expression of the Hippo ki-
haemangioendotheliomas in which WWTR1-CAMTA fusions were found nases or other Hippo pathway components has been associated
in 33 cases and YAP1-TFE3 fusion genes in the remaining 2 cases [71]. In with certain cancers [80,81]. In primary soft tissue sarcomas, the
this type of tumour, fusion products involving Hippo genes appear to be CpG island in of MST1 and MST2 kinase promoters were found to be
as specific as the PAX3/7-FOXO1 fusions are for alveolar rhabdomyosar- significantly hypermethylated. MST1 was methylated in liposarcoma
comas [72,73]. Apart from epithelioid haemangioma, a single case of (25%), leiomyosarcoma (50%), malignant fibrous histiocytoma and
spindle cell rhabdomyosarcoma was associated with a TEAD1-NCOA1 myxofibrosarcoma (17%) as well as in rhabdomyosarcoma (80%). In
fusion protein [74]. How these Hippo fusion genes contribute to contrast MST2 was found methylated principally in liposarcoma (15%),
tumourigenesis and why epithelioid haemangioendotheliomas appear malignant fibrous histiocytoma and myxofibrosarcoma (31%) as well
to depend almost in every case on WWTR1-CAMTA or YAP1-TFE3 fusion as synovial sarcoma (50%). MST1 methylation was associated with a de-
genes is unknown. Recently, the WWTR1-CAMTA fusion was shown to creased expression of the MST1 mRNA, and surprisingly, a significantly
inhibit the tumour suppressive function of the CAMTA transcriptional increased risk for tumour-related death was found for patients with
regulator and to confer a TAZ-like transcriptional programme in cells an unmethylated MST1 promoter. The tumour suppressor kinase
expressing the fusion, thereby favouring oncogenic transformation LATS1 was also methylated in 7% of all sarcomas tested and the LATS2
and resistance to anoikis [75]. and WW45/Salvador genes were not found methylated in sarcomas.
The promoter of the tumour suppressor gene RASSF1A has been shown
5.2. Hippo copy number alterations to be methylated in liposarcoma (18%), leiomyosarcoma (39%), malig-
nant fibrous histiocytoma and myxofibrosarcoma (6%) and in neurogenic
Generally, specific types of cancer are either dominated by high sarcomas (50%), which is associated with an unfavourable prognosis for
levels of somatic point mutations (termed M type) or by copy number cancer patients [82]. These results, obtained with a small sample size,
gains with TP53 mutations (termed C type), or by intermediate but are a proof-of-principle for the epigenetic regulation of Hippo genes. It
not high levels of both [76]. This may reflect different causes of muta- remains to be determined if the methylation status of the MST kinases
genesis and also be related to chromatin features, as point mutations is linked with YAP activity in sarcomas. Indeed, the MST kinases have
and copy number losses are associated with late replicating chromatin been shown to be dispensable for Hippo pathway activity in some cell
whereas copy number gains and rearrangements are associated with types, notably in mouse embryonic fibroblasts and skin [83].
early replicating chromatin [54,55]. Generally, Hippo genes appear far
more affected by copy number gains and rearrangements than by 5.4. Hippo signalling and the genomic landscape of sarcoma
point mutations as evidenced from Hippo genes analyses using the
cBioPortal. This supports the idea that Hippo genes are possibly located So far two studies have reported the sequencing of multiple types of
in early replicating chromatin, consistent with the elevated copy num- sarcoma [77] and of rhabdomyosarcoma, a sarcoma where skeletal
ber alterations found in C type cancers and the existence of YAP1 and muscle-related genes are expressed [73]. Additionally, preliminary
WWTR1 fusion genes. Indeed, sporadic Hippo copy number gains are data from a TCGA sarcoma study are accessible through cBIO but these
found in different types of sarcoma (see Fig. S1 for a re-analysis of the data are under publication embargo until 2016. In the earliest large-
data of Barretina et al. 2010 using the cBIOPortal platform [77]). Copy scale sequencing study 207 high grade, seven types of soft-tissue sarco-
number gains of YAP1 (11q) were reported in 22 cases and copy number mas were analysed for the expression of 722 protein-coding and
gains of VGLL3 (3p) were found in 19 cases of soft tissue sarcoma sub- microRNA genes [77]. A limitation of this study was the subjective selec-
types [17]. YAP1 copy number gains were additionally observed in em- tion of the genes sequenced and the relatively small sample number for
bryonal but not alveolar rhabdomyosarcoma [18] and high level VGLL3 such an analysis. However, the mutated genes CTNNB1, PIK3CA, NF1 and
amplifications were found in 10 out of 12 cases of myxoinflammatory, PIK3CA identified in this study were also identified later in a larger unbi-
fibroblastic sarcoma and hemosiderotic, fibrolipomatous tumours [78]. ased study involving next generation sequencing of rhabdomyosar-
Additionally the re-analysis of cBioPortal-deposited sequencing data comas [73].
for different types of sarcoma [77] reveals YAP1, WWTR1, VGLL1-4 or In rhabdomyosarcoma, the most distinct subgroup is alveolar rhab-
TEAD1-4 copy number gains but no losses in 17% of the cases (YAP1 domyosarcoma (ARMS). This subtype is in most cases marked by
3%, WWTR1 2%, VGLL4 2%, VGLL3 5%; Fig. S1). However, it is currently un- PAX3-FOXO1 or PAX7-FOXO1 fusion genes with rare alternative PAX fu-
known, although seemingly unlikely, whether all of these copy number sions [73]. A recent study has linked PAX3-FOXO1 to the increased ex-
gains in Hippo genes are bona fide cancer drivers and whether their ef- pression of RASSF4 which inhibits the Hippo kinase MST1 [84]. Such
fect is a major one. Indeed, copy number gains of YAP1, VGLL3, VGLL4 or an inhibition of a Hippo kinase should result in an increased transcrip-
TEAD1-4 appear to co-exist even though VGLL4 has been associated tional output downstream of the Hippo signalling cascade, but the
with a disruption of TEAD/YAP transcriptional activity and tumour sup- exact mechanism could not be identified. Nevertheless, this suggest
pression [41,79]. Second, high YAP activity appears to result in the acti- cross-talk between the ARMS PAX3/7-FOXO1 fusions and Hippo signal-
vation of a negative feedback loop [67]. In line with this, a higher ling. Conversely, the rhabdomyosarcomas that are negative for PAX3/
concentration of YAP (or elevated YAP nuclear localization and tran- 7-FOXO1 fusions, which represent mostly embryonal rhabdomyosar-
scriptional activity) is generally associated with an increase in the levels comas and 20% of the alveolar RMS cases [85], carry point mutations
of inhibitory YAP Ser127 phosphorylation because of the increased ex- of cancer genes that are frequently mutated pan-cancer [13] including
pression of several members of the Hippo kinase cascade that inhibits KRAS, HRAS, NRAS, NF1, PIK3CA, CTNNB1 and FBXW7 [73]. Many of
YAP through phosphorylation [67]. This is functionally supported by these mutated cancer genes are likely to cross-talk to Hippo members
the fact that the expression of a wild-type YAP1 does not prevent (described in more detail below). Additionally, the genomes of ERMS
myogenic differentiation whereas the expression of constitutively ac- are complex and are aneuploid (i.e. changed chromosome numbers)
tive YAP1 S127A does prevent differentiation [66]. Collectively, this sug- and display segmental aberrations whereas the genomes of ARMS
gests that recurrent copy number alterations, especially gains, in the only show few of such changes [73,86]. While the genes affected in
TEAD co-regulator genes YAP1, WWTR1, VGLL3 and VGLL4 in a subset ERMS vary greatly, 171 out of 196 ERMS tumours were positive for
126 A.D. Mohamed et al. / Biochimica et Biophysica Acta 1856 (2015) 121–129

YAP by immunohistochemistry and in 143 out of these 171 cases YAP Generally, these drugs can be sub-divided into either drugs that target
was nuclear [18]. Additionally, the expression of constitutively active the binding of YAP to TEAD transcription factors or into drugs that mod-
YAP1 S127A in mice resulted in the formation of ERMS-like tumours ulate other members of the wider Hippo signal transduction network,
with short latency in every case [18]. This suggests that several of the such as G protein-coupled receptors or the mevalonate pathway. Sever-
mutated ERMS genes [73], together with the YAP1 copy number gains al potential Hippo drugs such as verteporfin or statins are already wide-
[17,18], converge to towards increasing YAP activity in ERMS, which is ly used in patients to treat other conditions and this can pave the way
likely driving the disease. for early clinical trials involving such compounds. Key obstacles for sys-
There is evidence that several of the somatically point mutated temic and long-term Hippo-targeting treatments for sarcoma, and for
ERMS and sarcoma genes can cross-talk to YAP. Several recent reports cancer in general, are that the Hippo pathway is ubiquitously expressed
have linked KRAS to YAP in cancer. To identify genes that can substitute (i.e. the drugs may perturb signalling in other tissues and cause side ef-
for oncogenic RAS, 15,294 genes were expressed in KRAS-dependent fects), although, in many tissues, the loss-of-function of YAP is incon-
colon cancer cells. This non-biased analysis revealed that YAP could res- spicuous under homeostatic conditions until tissue regeneration is
cue the viability of KRAS-addicted cells and that YAP was required for required. Here we discuss the different Hippo treatment options that
KRAS-induced cell transformation [57]. In another study, the authors ob- might become useful for the treatment of sarcomas that result from dys-
served in a mouse model of pancreatic ductal adenocarcinoma that regulated Hippo signalling.
some KRAS G12D-driven tumours relapsed after KRAS inhibition. The re-
sultant tumours exhibited an amplification and overexpression of YAP1 6.1. Drugs that target YAP-TEAD interaction or TEADs: verteporfin, cyclic
[58]. YAP1 was also identified in a non-biased screen using 27,500 peptides, VGLL4-mimicking peptides
shRNAs to knockdown 5046 signalling genes as the top hit for a gene re-
sponsible for resistance to RAF or MEK inhibitor therapy [87]. Also, in Verteporfin, a member of the porphyrin family, contains aromatic
pancreatic ductal adenocarcinoma, it was found that YAP1 expression heterocyclic molecules formed of four modified pyrrole units intercon-
was increased and that the deletion of YAP1 prevented the progression nected at their carbon atoms by methane bridges. Verteporfin is already
of early lesions [59]. Taken together, it seems likely that similar mecha- a clinically approved photosensitizer for vascular macular degeneration
nisms operate in ERMS where oncogenic RAS mutations (NRAS, KRAS, [96]. Verteporfin binds to YAP and changes its conformation thereby
HRAS or mutations of the RAS inhibitor NF1) are relatively common inhibiting its interaction with TEAD2 and, presumably, with other
[73]. In line with this, the knockdown of YAP in the RD ERMS cell line, TEAD isoforms. Furthermore, verteporfin decreased liver overgrowth
which carries a NRAS Q61H mutation (accompanied by MYC amplifica- and blocked liver tumourigenesis induced by YAP overexpression or en-
tion and TP53 mutation with a karyotype of 51-hyperdiploidy [88]), re- dogenous YAP activation [97]. Moreover, uveal melanoma is a cancer
duces proliferation, soft agar growth, and xenotransplant tumour resulting from gain-of-function mutations in the GNAQ and GNA11 on-
burden while promoting the differentiation of these cells [18]. cogenes encoding the heterotrimeric Gαq family that stimulate YAP ac-
Somatic mutations of PIK3CA, which encodes phosphoinositide 3- tivity. Verteporfin has also been shown to reduce the tumour formation
kinase (PI3K), have been reported in 14% of all cases in a large-scale potential of GNAQ/GNA11 mutated uveal melanoma cells in xenograft
pan-cancer study [13] showing that PI3K is a frequently mutated gene assays [98,99].
pan-cancer. It has been demonstrated that epidermal growth factor
(EGF) inhibits the Hippo pathway through PI3K, resulting in YAP nucle- 6.2. G protein-coupled receptors (GPCRs)
ar localisation transcriptional activity [89]. In this study, it has been sug-
gested that this effect is independent of PKB/AKT, which is a key kinase G protein-coupled receptors (GPCRs) are the largest family of cell
downstream of PI3K. However, several studies have demonstrated surface receptors. Deep sequencing studies have shown an association
cross-talk between Hippo signalling and AKT1, which can phosphory- between aberrant expression and activity of GPCRs and tumourigenesis
late MST1 at Thr120 and Thr387 leading to the inhibition of Mst1 in nearly 20% of human cancers carrying mutations in GPCRs [100]. Re-
activity [90–92]. Thus, AKT1-MST1 cross-talk could be an alternative cent studies have revealed that the Hippo pathway can be broadly reg-
mechanism by which PIK3CA mutations affect the activity of YAP or TAZ. ulated by different GPCRs. Serum-borne lysophosphatidic acid (LPA)
Mutations of the Wnt gene CTNNB1 (β-catenin) occur in some ERMS and sphingosine 1-phosphate (S1P) inhibit the Hippo kinases LATS1/2
cases [73]. We have already mentioned the close integration of Wnt and by acting through G12/G13 coupled receptors and Rho leading to induc-
Hippo signalling. Specifically, it was demonstrated that YAP was crucial tion of YAP/TAZ transcription activity [47]. Therefore, the drugs
for the proliferation and soft agar growth of cancer cell lines with high targeting the GPCR isoforms affecting Hippo signalling represent a po-
β-catenin activity [37]. It was shown that β-catenin can form complexes tential therapeutic approach to treat sarcomas driven by high activity
with YES1, YAP and TBX5 [37]. Intriguingly, a loss-of-function screen of YAP or TAZ.
implicated YES1 in the survival of rhabdomyosarcoma cells [93], per-
haps suggesting that similar mechanisms operate in sarcomas. 6.3. Statins and bisphosphonates
The SCF (β-TrCP) ubiquitin ligase regulates YAP [35] and TAZ [94]
degradation via well-defined phosphodegron-mediated mechanisms. Aberrant activity of the mevalonate pathway can promote tumour
FBXW7 is another E3 ubiquitin ligase that was suggested to regulate progression in vivo and anchorage independent growth of cells
the degradation of YAP [95]. Indeed, an inverse correlation was reported in vitro. Also, high mRNA levels of hydroxymethylglutaryl coenzyme A
between FBXW7 and YAP protein abundance in hepatocellular carcino- reductase (HMGCR) and mevalonate pathway genes are associated
ma. Furthermore, YAP overexpression could counteract the apoptosis with poor prognosis and reduced survival in breast cancer patients
and growth arrest caused by FBXW7 [95]. This effect of FBXW7 on YAP [101]. Two recent studies have shown that YAP/TAZ activity can be reg-
levels is consistent with the notion that a loss of FBXW7 in ERMS could ulated by the mevalonate pathway [102,103]. Mechanistically, the
favour a higher level of YAP activity. Collectively, this also implies that geranylgeranyl phosphate produced by the mevalonate cascade acti-
Hippo signalling plays an important but different role in both ARMS vates Rho GTPases, which in turn increases YAP/TAZ activity by
[84] and ERMS downstream of the mutated genes [18]. inhibiting their phosphorylation and enhancing their nuclear accumula-
tion. Interestingly, statins and bisphosphonates that inhibit distinct en-
6. Hippo-targeted therapies in soft tissue sarcoma? zymes of the mevalonate pathway have a marked effect on YAP/TAZ
transcriptional activity. Taken together these findings suggest that
Before 2012, the Hippo pathway was considered un-druggable, but YAP/TAZ are presumably mediating the oncogenic responses associated
since 2012 several compounds and treatment strategies have emerged. with the dysregulation of the mevalonate pathway. Moreover, statins
A.D. Mohamed et al. / Biochimica et Biophysica Acta 1856 (2015) 121–129 127

and bisphosphonates are potential treatments to target the malignant Supplementary data to this article can be found online at http://dx.
effects of YAP and TAZ in cancer cells [102,103]. In accordance with doi.org/10.1016/j.bbcan.2015.05.006.
this, a number of recent studies have reported a possible preventive or
therapeutic role of statins in cancer, and further studies are ongoing to Transparency document
validate and characterize the anticancer effects of statins in various con-
texts [104]. The Transparency document associated with this article can be
found, in the online version.
7. Summary and perspectives

In summary, the Hippo pathway comprises potent tumour suppres- Funding


sors and oncogenes that are rarely somatically point mutated in sarco-
mas or in other types of human cancer. The involvement of Hippo Work in the University of Aberdeen laboratories is supported by
members, especially YAP and TAZ, in human epithelioid haemangioma grants from Sarcoma UK (reference 004.2012), Friends of Anchor (refer-
occurs in the form of WWTR1-CAMTA and YAP1-TFE3 fusion genes, ences 14004 and 14005), and the Medical Research Council (grant num-
whereas copy number alterations of YAP1 and other Hippo genes occur ber 99477). Annie M. Tremblay is recipient of a postdoctoral fellowship
generally at a low frequency in other types of sarcomas. However, it is from the Canadian Institutes of Health Research (CIHR).
unclear whether these copy number alterations are important or even
essential for sarcomagenesis in the presence of functional Hippo kinases. References
Thus, the majority of the changes in the activity of Hippo genes, and spe-
[1] R.S. Udan, M. Kango-Singh, R. Nolo, C. Tao, G. Halder, Hippo promotes proliferation ar-
cifically YAP1, in sarcoma probably result from the cross-talk with sever- rest and apoptosis in the Salvador/Warts pathway, Nat. Cell Biol. 5 (2003) 914–920.
al genes that are frequently mutated in sarcoma and across cancers. For [2] K.F. Harvey, X. Zhang, D.M. Thomas, The Hippo pathway and human cancer, Nat.
instance, the PAX3-FOXO1 fusion gene driving alveolar RMS can inter- Rev. Cancer 13 (2013) 246–257.
[3] T. Moroishi, C.G. Hansen, K.L. Guan, The emerging roles of YAP and TAZ in cancer,
play with the Hippo pathway via the upregulation of RASSF4, a negative nature reviews, Cancer 15 (2015) 73–79.
regulator of the MST1 kinase, and perhaps also by modulating YAP1 ac- [4] K. Harvey, N. Tapon, The Salvador–Warts–Hippo pathway — an emerging tumour-
tivity [84]. In other sarcoma subtypes like embryonal RMS, somatic gain- suppressor network, Nat. Rev. Cancer 7 (2007) 182–191.
[5] D. Pan, The hippo signaling pathway in development and cancer, Dev. Cell 19
of-function mutations in KRAS, PIK3CA or CTNNB1 and loss-of-function
(2010) 491–505.
mutations in FBXW7 could affect Hippo pathway activity. Perturbed [6] B.S. Taylor, J. Barretina, R.G. Maki, C.R. Antonescu, S. Singer, M. Ladanyi, Advances in
Hippo signalling can clearly drive sarcomagenesis as evidenced from sarcoma genomics and new therapeutic targets, nature reviews, Cancer 11 (2011)
541–557.
mouse models expressing YAP1 S127A (gain-of-function) or from mice
[7] J.R.F. Goldblum, A.L., S. Weiss, Enzinger and Weiss's Soft Tissue Tumors, 6th ed.
harbouring LATS1, Nf2 or Mob1a/b loss-of-function mutations. Also, the Elsevier, 2014.
near 100% penetrance of WWTR1-CAMTA or YAP1-TFE3 fusion genes in [8] J.L. Anderson, C.T. Denny, W.D. Tap, N. Federman, Pediatric sarcomas: translating
epithelioid haemangioendotheliomas suggests that these Hippo fusion molecular pathogenesis of disease to novel therapeutic possibilities, Pediatr. Res.
72 (2012) 112–121.
genes are driving sarcomagenesis. [9] B.S. Fletcher, J.A. Bridge, P.C.D. Hogendoorn, F. Mertens, World Health Organisation,
Given the apparent importance of Hippo signal transduction path- Classification of tumours: Tumours of Soft Tissue and Bone, 4th ed. WHO Press,
way members in sarcoma it is welcome news that many different vali- 2013.
[10] C.D. Fletcher, The evolving classification of soft tissue tumours — an update based
dated or putative Hippo treatments have emerged since 2012. They on the new 2013 WHO classification, Histopathology 64 (2014) 2–11.
can be subdivided into drugs that directly target the Hippo transcrip- [11] C. Fisher, Immunohistochemistry in diagnosis of soft tissue tumours, Histopathol-
tional regulators and drugs that target other members of the wider ogy 58 (2011) 1001–1012.
[12] C. Fisher, Dataset for Histopathology Reporting of Soft Tissue Sarcomas, in: T.R.C.o.
Hippo signal transduction network. Some of these drugs are already Pathologists (Ed.), 2014.
widely used for the treatment of other diseases and this could facilitate [13] M.S. Lawrence, P. Stojanov, C.H. Mermel, J.T. Robinson, L.A. Garraway, T.R. Golub,
the progression towards preclinical studies aimed at inhibiting the ac- M. Meyerson, S.B. Gabriel, E.S. Lander, G. Getz, Discovery and saturation analysis
of cancer genes across 21 tumour types, Nature 505 (2014) 495–501.
tivity of YAP or TAZ in sarcomas.
[14] B. Vogelstein, N. Papadopoulos, V.E. Velculescu, S. Zhou, L.A. Diaz Jr., K.W. Kinzler,
Several lines of evidence now support the notion that the Hippo Cancer genome landscapes, Science 339 (2013) 1546–1558.
pathway might be a major player in sarcomagenesis, but they also [15] C.R. Antonescu, F. Le Loarer, J.M. Mosquera, A. Sboner, L. Zhang, C.L. Chen, H.W.
Chen, N. Pathan, T. Krausz, B.C. Dickson, I. Weinreb, M.A. Rubin, M. Hameed, C.D.
raise a number of questions yet to be answered:
Fletcher, Novel YAP1-TFE3 fusion defines a distinct subset of epithelioid
hemangioendothelioma, Genes Chromosom. Cancer 52 (2013) 775–784.
– Can the expression, abundance or transcriptional activity of YAP (as [16] C. Errani, L. Zhang, Y.S. Sung, M. Hajdu, S. Singer, R.G. Maki, J.H. Healey, C.R.
assessed by gene expression signatures [18]) be used in the clinic for Antonescu, A novel WWTR1-CAMTA1 gene fusion is a consistent abnormality in
epithelioid hemangioendothelioma of different anatomic sites, Genes Chromosom.
classification of sarcomas and prediction of clinical outcome? Cancer 50 (2011) 644–653.
– Does the Hippo pathway mostly regulate generic functions such as [17] Z. Helias-Rodzewicz, G. Perot, F. Chibon, C. Ferreira, P. Lagarde, P. Terrier, J.M.
proliferation and apoptosis or does it consistently interfere with Coindre, A. Aurias, YAP1 and VGLL3, encoding two cofactors of TEAD transcription
factors, are amplified and overexpressed in a subset of soft tissue sarcomas, Genes
lineage-specific fate determination gene expression as in ERMS Chromosom. Cancer 49 (2010) 1161–1171.
where it mainly suppresses the expression of differentiated muscle [18] A.M. Tremblay, E. Missiaglia, G.G. Galli, S. Hettmer, R. Urcia, M. Carrara, R.N. Judson,
genes [18]? K. Thway, G. Nadal, J.L. Selfe, G. Murray, R.A. Calogero, C. De Bari, P.S. Zammit, M.
Delorenzi, A.J. Wagers, J. Shipley, H. Wackerhage, F.D. Camargo, The Hippo trans-
– Why is there a nearly 100% penetrance of WWTR1-CAMTA or YAP1-
ducer YAP1 transforms activated satellite cells and is a potent effector of embryo-
TFE3 fusion genes in epithelioid haemangioendothelioma? What nal rhabdomyosarcoma formation, Cancer Cell 26 (2014) 273–287.
do these fusion genes do and why are they so important in this [19] K.H. Hallor, R. Sciot, J. Staaf, M. Heidenblad, A. Rydholm, H.C. Bauer, K. Astrom, H.A.
type of sarcoma? Domanski, J.M. Meis, L.G. Kindblom, I. Panagopoulos, N. Mandahl, F. Mertens, Two
genetic pathways, t(1;10) and amplification of 3p11-12, in myxoinflammatory fi-
broblastic sarcoma, haemosiderotic fibrolipomatous tumour, and morphologically
similar lesions, J. Pathol. 217 (2009) 716–727.
[20] K. Gambaro, M.C. Quinn, P.M. Wojnarowicz, S.L. Arcand, M. de Ladurantaye, V.
Even if the role of the Hippo pathway in cancer is now well
Barres, J.S. Ripeau, A.M. Killary, E.C. Davis, J. Lavoie, D.M. Provencher, A.M. Mes-
established, Hippo research is still in its infancy compared with other tu- Masson, M. Chevrette, P.N. Tonin, VGLL3 expression is associated with a tumor
mour suppressor and oncogenic proteins and pathways such as P53, suppressor phenotype in epithelial ovarian cancer, Mol. Oncol. 7 (2013) 513–530.
MYC or oncogenic RAS, and there is still much left to understand before [21] D.M. Parham, D.A. Ellison, Rhabdomyosarcomas in adults and children: an update,
Arch. Pathol. Lab. Med. 130 (2006) 1454–1465.
achieving full control over this intriguing pathway to specifically target [22] D. Van Mater, L. Ano, J.M. Blum, M.T. Webster, W. Huang, N. Williams, Y. Ma, D.M.
sarcomas and other cancers. Cardona, C.M. Fan, D.G. Kirsch, Acute tissue injury activates satellite cells and
128 A.D. Mohamed et al. / Biochimica et Biophysica Acta 1856 (2015) 121–129

promotes sarcoma formation via the HGF/c-MET signaling pathway, Cancer Res. 75 [53] E. Cerami, J. Gao, U. Dogrusoz, B.E. Gross, S.O. Sumer, B.A. Aksoy, A. Jacobsen, C.J.
(2015) 605–614. Byrne, M.L. Heuer, E. Larsson, Y. Antipin, B. Reva, A.P. Goldberg, C. Sander, N.
[23] D.A. Liebner, The indications and efficacy of conventional chemotherapy in prima- Schultz, The cBio cancer genomics portal: an open platform for exploring multidi-
ry and recurrent sarcoma, J. Surg. Oncol. 111 (2015) 622–631. mensional cancer genomics data, Cancer Discov. 2 (2012) 401–404.
[24] A. Gronchi, C. Colombo, C.P. Raut, Surgical management of localized soft tissue tu- [54] M.S. Lawrence, P. Stojanov, P. Polak, G.V. Kryukov, K. Cibulskis, A. Sivachenko, S.L.
mors, Cancer 120 (2014) 2638–2648. Carter, C. Stewart, C.H. Mermel, S.A. Roberts, A. Kiezun, P.S. Hammerman, A.
[25] M. Linch, A.B. Miah, K. Thway, I.R. Judson, C. Benson, Systemic treatment of soft- McKenna, Y. Drier, L. Zou, A.H. Ramos, T.J. Pugh, N. Stransky, E. Helman, J. Kim, C.
tissue sarcoma-gold standard and novel therapies, nature reviews, Clin. Oncol. 11 Sougnez, L. Ambrogio, E. Nickerson, E. Shefler, M.L. Cortes, D. Auclair, G. Saksena, D.
(2014) 187–202. Voet, M. Noble, D. DiCara, P. Lin, L. Lichtenstein, D.I. Heiman, T. Fennell, M.
[26] S.P. D'Angelo, W.D. Tap, G.K. Schwartz, R.D. Carvajal, Sarcoma immunotherapy: Imielinski, B. Hernandez, E. Hodis, S. Baca, A.M. Dulak, J. Lohr, D.A. Landau, C.J. Wu, J.
past approaches and future directions, Sarcoma 2014 (2014) 391967. Melendez-Zajgla, A. Hidalgo-Miranda, A. Koren, S.A. McCarroll, J. Mora, R.S. Lee, B.
[27] S. Radaelli, S. Stacchiotti, P.G. Casali, A. Gronchi, Emerging therapies for adult soft Crompton, R. Onofrio, M. Parkin, W. Winckler, K. Ardlie, S.B. Gabriel, C.W. Roberts,
tissue sarcoma, Expert. Rev. Anticancer. Ther. 14 (2014) 689–704. J.A. Biegel, K. Stegmaier, A.J. Bass, L.A. Garraway, M. Meyerson, T.R. Golub, D.A.
[28] T. Al-Zaid, N. Somaiah, A.J. Lazar, Targeted therapies for sarcomas: new roles for the Gordenin, S. Sunyaev, E.S. Lander, G. Getz, Mutational heterogeneity in cancer and
pathologist, Histopathology 64 (2014) 119–133. the search for new cancer-associated genes, Nature 499 (2013) 214–218.
[29] C. Forscher, M. Mita, R. Figlin, Targeted therapy for sarcomas, Biologics Targets [55] J. Sima, D.M. Gilbert, Complex correlations: replication timing and mutational land-
Ther. 8 (2014) 91–105. scapes during cancer and genome evolution, Curr. Opin. Genet. Dev. 25C (2014)
[30] A. Neuville, F. Chibon, J.M. Coindre, Grading of soft tissue sarcomas: from histolog- 93–100.
ical to molecular assessment, Pathology 46 (2014) 113–120. [56] M. Overholtzer, J. Zhang, G.A. Smolen, B. Muir, W. Li, D.C. Sgroi, C.X. Deng, J.S.
[31] S.B. Edge, D.R., C.C. Compton, A.G. Fritz, F.L. Greene, A. Trotti, AJCC Cancer Staging Brugge, D.A. Haber, Transforming properties of YAP, a candidate oncogene on
Manual, 7th ed. Springer, 2010. the chromosome 11q22 amplicon, Proc. Natl. Acad. Sci. U. S. A. 103 (2006)
[32] E. Gateff, Malignant neoplasms of genetic origin in Drosophila melanogaster, Sci- 12405–12410.
ence 200 (1978) 1448–1459. [57] D.D. Shao, W. Xue, E.B. Krall, A. Bhutkar, F. Piccioni, X. Wang, A.C. Schinzel, S. Sood,
[33] J. Avruch, D. Zhou, J. Fitamant, N. Bardeesy, F. Mou, L.R. Barrufet, Protein kinases of the J. Rosenbluh, J.W. Kim, Y. Zwang, T.M. Roberts, D.E. Root, T. Jacks, W.C. Hahn, KRAS
Hippo pathway: regulation and substrates, Semin. Cell Dev. Biol. 23 (2012) 770–784. and YAP1 converge to regulate EMT and tumor survival, Cell 158 (2014) 171–184.
[34] T. Oka, V. Mazack, M. Sudol, Mst2 and Lats kinases regulate apoptotic function of [58] A. Kapoor, W. Yao, H. Ying, S. Hua, A. Liewen, Q. Wang, Y. Zhong, C.J. Wu, A.
Yes kinase-associated protein (YAP), J. Biol. Chem. 283 (2008) 27534–27546. Sadanandam, B. Hu, Q. Chang, G.C. Chu, R. Al-Khalil, S. Jiang, H. Xia, E. Fletcher-
[35] B. Zhao, L. Li, K. Tumaneng, C.Y. Wang, K.L. Guan, A coordinated phosphorylation by Sananikone, C. Lim, G.I. Horwitz, A. Viale, P. Pettazzoni, N. Sanchez, H. Wang, A.
Lats and CK1 regulates YAP stability through SCF(beta-TRCP), Genes Dev. 24 Protopopov, J. Zhang, T. Heffernan, R.L. Johnson, L. Chin, Y.A. Wang, G. Draetta,
(2010) 72–85. R.A. DePinho, Yap1 activation enables bypass of oncogenic kras addiction in pan-
[36] Q.Y. Lei, H. Zhang, B. Zhao, Z.Y. Zha, F. Bai, X.H. Pei, S. Zhao, Y. Xiong, K.L. Guan, TAZ creatic cancer, Cell 158 (2014) 185–197.
promotes cell proliferation and epithelial-mesenchymal transition and is inhibited [59] W. Zhang, N. Nandakumar, Y. Shi, M. Manzano, A. Smith, G. Graham, S. Gupta, E.E.
by the hippo pathway, Mol. Cell. Biol. 28 (2008) 2426–2436. Vietsch, S.Z. Laughlin, M. Wadhwa, M. Chetram, M. Joshi, F. Wang, B. Kallakury, J.
[37] J. Rosenbluh, D. Nijhawan, A.G. Cox, X. Li, J.T. Neal, E.J. Schafer, T.I. Zack, X. Wang, A. Toretsky, A. Wellstein, C. Yi, Downstream of mutant KRAS, the transcription regu-
Tsherniak, A.C. Schinzel, D.D. Shao, S.E. Schumacher, B.A. Weir, F. Vazquez, G.S. lator YAP is essential for neoplastic progression to pancreatic ductal adenocarcino-
Cowley, D.E. Root, J.P. Mesirov, R. Beroukhim, C.J. Kuo, W. Goessling, W.C. Hahn, ma, Sci. Signal. 7 (2014) ra42.
beta-Catenin-driven cancers require a YAP1 transcriptional complex for survival [60] M.A. St John, W. Tao, X. Fei, R. Fukumoto, M.L. Carcangiu, D.G. Brownstein, A.F.
and tumorigenesis, Cell 151 (2012) 1457–1473. Parlow, J. McGrath, T. Xu, Mice deficient of Lats1 develop soft-tissue sarcomas,
[38] M. Sudol, Yes-associated protein (YAP65) is a proline-rich phosphoprotein that ovarian tumours and pituitary dysfunction, Nat. Genet. 21 (1999) 182–186.
binds to the SH3 domain of the Yes proto-oncogene product, Oncogene 9 (1994) [61] A.I. McClatchey, I. Saotome, K. Mercer, D. Crowley, J.F. Gusella, R.T. Bronson, T. Jacks,
2145–2152. Mice heterozygous for a mutation at the Nf2 tumor suppressor locus develop a
[39] M. Sudol, P. Bork, A. Einbond, K. Kastury, T. Druck, M. Negrini, K. Huebner, D. range of highly metastatic tumors, Genes Dev. 12 (1998) 1121–1133.
Lehman, Characterization of the mammalian YAP (Yes-associated protein) gene [62] M. Nishio, K. Hamada, K. Kawahara, M. Sasaki, F. Noguchi, S. Chiba, K. Mizuno, S.O.
and its role in defining a novel protein module, the WW domain, J. Biol. Chem. Suzuki, Y. Dong, M. Tokuda, T. Morikawa, H. Hikasa, J. Eggenschwiler, N. Yabuta, H.
270 (1995) 14733–14741. Nojima, K. Nakagawa, Y. Hata, H. Nishina, K. Mimori, M. Mori, T. Sasaki, T.W. Mak,
[40] B. Zhao, X. Ye, J. Yu, L. Li, W. Li, S. Li, J. Yu, J.D. Lin, C.Y. Wang, A.M. Chinnaiyan, Z.C. T. Nakano, S. Itami, A. Suzuki, Cancer susceptibility and embryonic lethality in
Lai, K.L. Guan, TEAD mediates YAP-dependent gene induction and growth control, Mob1a/1b double-mutant mice, J. Clin. Invest. 122 (2012) 4505–4518.
Genes Dev. 22 (2008) 1962–1971. [63] F. Hamaratoglu, M. Willecke, M. Kango-Singh, R. Nolo, E. Hyun, C. Tao, H. Jafar-
[41] L.M. Koontz, Y. Liu-Chittenden, F. Yin, Y. Zheng, J. Yu, B. Huang, Q. Chen, S. Wu, D. Nejad, G. Halder, The tumour-suppressor genes NF2/merlin and expanded act
Pan, The Hippo effector yorkie controls normal tissue growth by antagonizing through Hippo signalling to regulate cell proliferation and apoptosis, Nat. Cell
scalloped-mediated default repression, Dev. Cell 25 (2013) 388–401. Biol. 8 (2006) 27–36.
[42] E. Aragon, N. Goerner, A.I. Zaromytidou, Q. Xi, A. Escobedo, J. Massague, M.J. Macias, [64] N. Zhang, H. Bai, K.K. David, J. Dong, Y. Zheng, J. Cai, M. Giovannini, P. Liu, R.A.
A Smad action turnover switch operated by WW domain readers of a Anders, D. Pan, The Merlin/NF2 tumor suppressor functions through the YAP
phosphoserine code, Genes Dev. 25 (2011) 1275–1288. oncoprotein to regulate tissue homeostasis in mammals, Dev. Cell 19 (2010)
[43] M. Murakami, M. Nakagawa, E.N. Olson, O. Nakagawa, A WW domain protein TAZ 27–38.
is a critical coactivator for TBX5, a transcription factor implicated in Holt–Oram [65] A. Hergovich, MOB control: reviewing a conserved family of kinase regulators, Cell.
syndrome, Proc. Natl. Acad. Sci. U. S. A. 102 (2005) 18034–18039. Signal. 23 (2011) 1433–1440.
[44] S. Dupont, L. Morsut, M. Aragona, E. Enzo, S. Giulitti, M. Cordenonsi, F. Zanconato, [66] K.I. Watt, R. Judson, P. Medlow, K. Reid, T.B. Kurth, J.G. Burniston, A. Ratkevicius, C.
D.J. Le, M. Forcato, S. Bicciato, N. Elvassore, S. Piccolo, Role of YAP/TAZ in De Bari, H. Wackerhage, Yap is a novel regulator of C2C12 myogenesis, Biochem.
mechanotransduction, Nature 474 (2011) 179–183. Biophys. Res. Commun. 393 (2010) 619–624.
[45] L. Azzolin, F. Zanconato, S. Bresolin, M. Forcato, G. Basso, S. Bicciato, M. Cordenonsi, [67] R.N. Judson, A.M. Tremblay, P. Knopp, R.B. White, R. Urcia, B.C. De, P.S. Zammit, F.D.
S. Piccolo, Role of TAZ as mediator of Wnt signaling, Cell 151 (2012) 1443–1456. Camargo, H. Wackerhage, The Hippo pathway member Yap plays a key role in
[46] L. Azzolin, T. Panciera, S. Soligo, E. Enzo, S. Bicciato, S. Dupont, S. Bresolin, C. influencing fate decisions in muscle satellite cells, J. Cell Sci. 125 (2012)
Frasson, G. Basso, V. Guzzardo, A. Fassina, M. Cordenonsi, S. Piccolo, YAP/TAZ incor- 6009–6019.
poration in the beta-catenin destruction complex orchestrates the Wnt response, [68] G. Pallafacchina, S. Francois, B. Regnault, B. Czarny, V. Dive, A. Cumano, D.
Cell 158 (2014) 157–170. Montarras, M. Buckingham, An adult tissue-specific stem cell in its niche: a gene
[47] F.X. Yu, B. Zhao, N. Panupinthu, J.L. Jewell, I. Lian, L.H. Wang, J. Zhao, H. Yuan, K. profiling analysis of in vivo quiescent and activated muscle satellite cells, Stem
Tumaneng, H. Li, X.D. Fu, G.B. Mills, K.L. Guan, Regulation of the Hippo-YAP path- Cell Res. 4 (2010) 77–91.
way by G-protein-coupled receptor signaling, Cell 150 (2012) 780–791. [69] J.S. Chamberlain, J. Metzger, M. Reyes, D. Townsend, J.A. Faulkner, Dystrophin-
[48] J. Dong, G. Feldmann, J. Huang, S. Wu, N. Zhang, S.A. Comerford, M.F. Gayyed, R.A. deficient mdx mice display a reduced life span and are susceptible to spontaneous
Anders, A. Maitra, D. Pan, Elucidation of a universal size-control mechanism in Dro- rhabdomyosarcoma, FASEB J. 21 (2007) 2195–2204.
sophila and mammals, Cell 130 (2007) 1120–1133. [70] V. Hosur, A. Kavirayani, J. Riefler, L.M. Carney, B. Lyons, B. Gott, G.A. Cox, L.D. Shultz,
[49] F.D. Camargo, S. Gokhale, J.B. Johnnidis, D. Fu, G.W. Bell, R. Jaenisch, T.R. Dystrophin and dysferlin double mutant mice: a novel model for rhabdomyosarco-
Brummelkamp, YAP1 increases organ size and expands undifferentiated progeni- ma, Cancer Genet. 205 (2012) 232–241.
tor cells, Curr. Biol. 17 (2007) 2054–2060. [71] U. Flucke, R.J. Vogels, N. de Saint Aubain Somerhausen, D.H. Creytens, R.G. Riedl,
[50] K. Schlegelmilch, M. Mohseni, O. Kirak, J. Pruszak, J.R. Rodriguez, D. Zhou, B.T. J.M. van Gorp, A.N. Milne, C.J. Huysentruyt, M.A. Verdijk, M.M. van Asseldonk, A.J.
Kreger, V. Vasioukhin, J. Avruch, T.R. Brummelkamp, F.D. Camargo, Yap1 acts Suurmeijer, J. Bras, G. Palmedo, P.J. Groenen, T. Mentzel, Epithelioid
downstream of alpha-catenin to control epidermal proliferation, Cell 144 (2011) hemangioendothelioma: clinicopathologic, immunhistochemical, and molecular
782–795. genetic analysis of 39 cases, Diagn. Pathol. 9 (2014) 131.
[51] T. Yu, J. Bachman, Z.C. Lai, Mutation analysis of large tumor suppressor genes LATS1 [72] E. Missiaglia, D. Williamson, J. Chisholm, P. Wirapati, G. Pierron, F. Petel, J.P.
and LATS2 supports a tumor suppressor role in human cancer, Protein Cell 6 Concordet, K. Thway, O. Oberlin, K. Pritchard-Jones, O. Delattre, M. Delorenzi, J.
(2015) 6–11. Shipley, PAX3/FOXO1 fusion gene status is the key prognostic molecular marker
[52] J. Gao, B.A. Aksoy, U. Dogrusoz, G. Dresdner, B. Gross, S.O. Sumer, Y. Sun, A. in rhabdomyosarcoma and significantly improves current risk stratification, J.
Jacobsen, R. Sinha, E. Larsson, E. Cerami, C. Sander, N. Schultz, Integrative analysis Clin. Oncol. 30 (2012) 1670–1677.
of complex cancer genomics and clinical profiles using the cBioPortal, Sci. Signal. 6 [73] J.F. Shern, L. Chen, J. Chmielecki, J.S. Wei, R. Patidar, M. Rosenberg, L. Ambrogio, D.
(2013) l1. Auclair, J. Wang, Y.K. Song, C. Tolman, L. Hurd, H. Liao, S. Zhang, D. Bogen, A.S.
A.D. Mohamed et al. / Biochimica et Biophysica Acta 1856 (2015) 121–129 129

Brohl, S. Sindiri, D. Catchpoole, T. Badgett, G. Getz, J. Mora, J.R. Anderson, S.X. [88] A.R. Hinson, R. Jones, L.E. Crose, B.C. Belyea, F.G. Barr, C.M. Linardic, Human rhabdo-
Skapek, F.G. Barr, M. Meyerson, D.S. Hawkins, J. Khan, Comprehensive genomic myosarcoma cell lines for rhabdomyosarcoma research: utility and pitfalls, Front.
analysis of rhabdomyosarcoma reveals a landscape of alterations affecting a com- Oncol. 3 (2013) 183.
mon genetic axis in fusion-positive and fusion-negative tumors, Cancer Discov. 4 [89] R. Fan, N.G. Kim, B.M. Gumbiner, Regulation of Hippo pathway by mitogenic
(2014) 216–231. growth factors via phosphoinositide 3-kinase and phosphoinositide-dependent
[74] J.M. Mosquera, A. Sboner, L. Zhang, N. Kitabayashi, C.L. Chen, Y.S. Sung, L.H. Wexler, kinase-1, Proc. Natl. Acad. Sci. U. S. A. 110 (2013) 2569–2574.
M.P. LaQuaglia, M. Edelman, C. Sreekantaiah, M.A. Rubin, C.R. Antonescu, Recurrent [90] Z. Yuan, D. Kim, S. Shu, J. Wu, J. Guo, L. Xiao, S. Kaneko, D. Coppola, J.Q. Cheng,
NCOA2 gene rearrangements in congenital/infantile spindle cell rhabdomyosarco- Phosphoinositide 3-kinase/Akt inhibits MST1-mediated pro-apoptotic signaling
ma, Genes Chromosom. Cancer 52 (2013) 538–550. through phosphorylation of threonine 120, J. Biol. Chem. 285 (2010) 3815–3824.
[75] M.R. Tanas, S. Ma, F.O. Jadaan, C.K. Ng, B. Weigelt, J.S. Reis-Filho, B.P. Rubin, Mech- [91] S.W. Jang, S.J. Yang, S. Srinivasan, K. Ye, Akt phosphorylates MstI and prevents its
anism of action of a WWTR1(TAZ)-CAMTA1 fusion oncoprotein, Oncogene (2015) proteolytic activation, blocking FOXO3 phosphorylation and nuclear translocation,
(in press). J. Biol. Chem. 282 (2007) 30836–30844.
[76] G. Ciriello, M.L. Miller, B.A. Aksoy, Y. Senbabaoglu, N. Schultz, C. Sander, Emerging [92] F.K. Collak, K. Yagiz, D.J. Luthringer, B. Erkaya, B. Cinar, Threonine-120 phosphory-
landscape of oncogenic signatures across human cancers, Nat. Genet. 45 (2013) lation regulated by phosphoinositide-3-kinase/Akt and mammalian target of
1127–1133. rapamycin pathway signaling limits the antitumor activity of mammalian sterile
[77] J. Barretina, B.S. Taylor, S. Banerji, A.H. Ramos, M. Lagos-Quintana, P.L. Decarolis, K. 20-like kinase 1, J. Biol. Chem. 287 (2012) 23698–23709.
Shah, N.D. Socci, B.A. Weir, A. Ho, D.Y. Chiang, B. Reva, C.H. Mermel, G. Getz, Y. [93] C.L. Yeung, V.N. Ngo, P.J. Grohar, F.I. Arnaldez, A. Asante, X. Wan, J. Khan, S.M.
Antipin, R. Beroukhim, J.E. Major, C. Hatton, R. Nicoletti, M. Hanna, T. Sharpe, T.J. Hewitt, C. Khanna, L.M. Staudt, L.J. Helman, Loss-of-function screen in rhabdomyo-
Fennell, K. Cibulskis, R.C. Onofrio, T. Saito, N. Shukla, C. Lau, S. Nelander, S.J. sarcoma identifies CRKL-YES as a critical signal for tumor growth, Oncogene 32
Silver, C. Sougnez, A. Viale, W. Winckler, R.G. Maki, L.A. Garraway, A. Lash, H. (2013) 5429–5438.
Greulich, D.E. Root, W.R. Sellers, G.K. Schwartz, C.R. Antonescu, E.S. Lander, H.E. [94] C.Y. Liu, Z.Y. Zha, X. Zhou, H. Zhang, W. Huang, D. Zhao, T. Li, S.W. Chan, C.J. Lim, W.
Varmus, M. Ladanyi, C. Sander, M. Meyerson, S. Singer, Subtype-specific genomic Hong, S. Zhao, Y. Xiong, Q.Y. Lei, K.L. Guan, The hippo tumor pathway promotes
alterations define new targets for soft-tissue sarcoma therapy, Nat. Genet. 42 TAZ degradation by phosphorylating a phosphodegron and recruiting the
(2010) 715–721. SCF{beta}-TrCP E3 ligase, J. Biol. Chem. 285 (2010) 37159–37169.
[78] C.R. Antonescu, L. Zhang, G.P. Nielsen, A.E. Rosenberg, P. Dal Cin, C.D. Fletcher, Con- [95] K. Tu, W. Yang, C. Li, X. Zheng, Z. Lu, C. Guo, Y. Yao, Q. Liu, Fbxw7 is an independent
sistent t(1;10) with rearrangements of TGFBR3 and MGEA5 in both prognostic marker and induces apoptosis and growth arrest by regulating YAP
myxoinflammatory fibroblastic sarcoma and hemosiderotic fibrolipomatous abundance in hepatocellular carcinoma, Mol. Cancer 13 (2014) 110.
tumor, Genes Chromosom. Cancer 50 (2011) 757–764. [96] K. Brodowska, A. Al-Moujahed, A. Marmalidou, M. Meyer Zu Horste, J. Cichy, J.W.
[79] S. Jiao, H. Wang, Z. Shi, A. Dong, W. Zhang, X. Song, F. He, Y. Wang, Z. Zhang, W. Miller, E. Gragoudas, D.G. Vavvas, The clinically used photosensitizer Verteporfin
Wang, X. Wang, T. Guo, P. Li, Y. Zhao, H. Ji, L. Zhang, Z. Zhou, A peptide mimicking (VP) inhibits YAP-TEAD and human retinoblastoma cell growth in vitro without
VGLL4 function acts as a YAP antagonist therapy against gastric cancer, Cancer Cell light activation, Exp. Eye Res. 124 (2014) 67–73.
25 (2014) 166–180. [97] Y. Liu-Chittenden, B. Huang, J.S. Shim, Q. Chen, S.J. Lee, R.A. Anders, J.O. Liu, D. Pan,
[80] Y. Takahashi, Y. Miyoshi, C. Takahata, N. Irahara, T. Taguchi, Y. Tamaki, S. Noguchi, Genetic and pharmacological disruption of the TEAD-YAP complex suppresses the
Down-regulation of LATS1 and LATS2 mRNA expression by promoter hypermethy- oncogenic activity of YAP, Genes Dev. 26 (2012) 1300–1305.
lation and its association with biologically aggressive phenotype in human breast [98] F.X. Yu, J. Luo, J.S. Mo, G. Liu, Y.C. Kim, Z. Meng, L. Zhao, G. Peyman, H. Ouyang, W.
cancers, Clin. Cancer Res. 11 (2005) 1380–1385. Jiang, J. Zhao, X. Chen, L. Zhang, C.Y. Wang, B.C. Bastian, K. Zhang, K.L. Guan, Mutant
[81] Z. Jiang, X. Li, J. Hu, W. Zhou, Y. Jiang, G. Li, D. Lu, Promoter hypermethylation- Gq/11 promote uveal melanoma tumorigenesis by activating YAP, Cancer Cell 25
mediated down-regulation of LATS1 and LATS2 in human astrocytoma, Neurosci. (2014) 822–830.
Res. 56 (2006) 450–458. [99] X. Feng, M.S. Degese, R. Iglesias-Bartolome, J.P. Vaque, A.A. Molinolo, M. Rodrigues,
[82] C. Seidel, U. Schagdarsurengin, K. Blumke, P. Wurl, G.P. Pfeifer, S. Hauptmann, H. M.R. Zaidi, B.R. Ksander, G. Merlino, A. Sodhi, Q. Chen, J.S. Gutkind, Hippo-
Taubert, R. Dammann, Frequent hypermethylation of MST1 and MST2 in soft tissue independent activation of YAP by the GNAQ uveal melanoma oncogene through
sarcoma, Mol. Carcinog. 46 (2007) 865–871. a trio-regulated rho GTPase signaling circuitry, Cancer Cell 25 (2014) 831–845.
[83] A.M. Tremblay, F.D. Camargo, Hippo signaling in mammalian stem cells, Semin. [100] M. O'Hayre, M.S. Degese, J.S. Gutkind, Novel insights into G protein and G protein-
Cell Dev. Biol. 23 (2012) 818–826. coupled receptor signaling in cancer, Curr. Opin. Cell Biol. 27 (2014) 126–135.
[84] L.E. Crose, K.A. Galindo, J.G. Kephart, C. Chen, J. Fitamant, N. Bardeesy, R.C. Bentley, [101] J.W. Clendening, A. Pandyra, P.C. Boutros, S. El Ghamrasni, F. Khosravi, G.A. Trentin,
R.L. Galindo, J.T. Ashley Chi, C.M. Linardic, Alveolar rhabdomyosarcoma-associated A. Martirosyan, A. Hakem, R. Hakem, I. Jurisica, L.Z. Penn, Dysregulation of the
PAX3-FOXO1 promotes tumorigenesis via Hippo pathway suppression, J. Clin. In- mevalonate pathway promotes transformation, Proc. Natl. Acad. Sci. U. S. A. 107
vest. 124 (2014) 285–296. (2010) 15051–15056.
[85] D. Williamson, E. Missiaglia, R.A. de, G. Pierron, B. Thuille, G. Palenzuela, K. Thway, [102] G. Sorrentino, N. Ruggeri, V. Specchia, M. Cordenonsi, M. Mano, S. Dupont, A.
D. Orbach, M. Lae, P. Freneaux, K. Pritchard-Jones, O. Oberlin, J. Shipley, O. Delattre, Manfrin, E. Ingallina, R. Sommaggio, S. Piazza, A. Rosato, S. Piccolo, G. Del Sal, Met-
Fusion gene-negative alveolar rhabdomyosarcoma is clinically and molecularly in- abolic control of YAP and TAZ by the mevalonate pathway, Nat. Cell Biol. 16 (2014)
distinguishable from embryonal rhabdomyosarcoma, J. Clin. Oncol. 28 (2010) 357–366.
2151–2158. [103] Z. Wang, Y. Wu, H. Wang, Y. Zhang, L. Mei, X. Fang, X. Zhang, F. Zhang, H. Chen, Y.
[86] X. Chen, E. Stewart, A.A. Shelat, C. Qu, A. Bahrami, M. Hatley, G. Wu, C. Bradley, J. Liu, Y. Jiang, S. Sun, Y. Zheng, N. Li, L. Huang, Interplay of mevalonate and Hippo
McEvoy, A. Pappo, S. Spunt, M.B. Valentine, V. Valentine, F. Krafcik, W.H. Lang, M. pathways regulates RHAMM transcription via YAP to modulate breast cancer cell
Wierdl, L. Tsurkan, V. Tolleman, S.M. Federico, C. Morton, C. Lu, L. Ding, J. Easton, motility, Proc. Natl. Acad. Sci. U. S. A. 111 (2014) E89–E98.
M. Rusch, P. Nagahawatte, J. Wang, M. Parker, L. Wei, E. Hedlund, D. Finkelstein, [104] A.K. Altwairgi, Statins are potential anticancerous agents (Review), Oncol. Rep. 33
M. Edmonson, S. Shurtleff, K. Boggs, H. Mulder, D. Yergeau, S. Skapek, D.S. (2015) 1019–1039.
Hawkins, N. Ramirez, P.M. Potter, J.A. Sandoval, A.M. Davidoff, E.R. Mardis, R.K. [105] A.V. Pobbati, S.W. Chan, I. Lee, H. Song, W. Hong, Structural and functional
Wilson, J. Zhang, J.R. Downing, M.A. Dyer, P. St, Jude Children's Research similarity between the Vgll1-TEAD and the YAP-TEAD complexes, Structure 20
Hospital-Washington University Pediatric Cancer Genome, targeting oxidative (2012) 1135–1140.
stress in embryonal rhabdomyosarcoma, Cancer Cell 24 (2013) 710–724. [106] Z. Zhou, T. Hu, Z. Xu, Z. Lin, Z. Zhang, T. Feng, L. Zhu, Y. Rong, H. Shen, J.M. Luk, X.
[87] L. Lin, A.J. Sabnis, E. Chan, V. Olivas, L. Cade, E. Pazarentzos, S. Asthana, D. Neel, J.J. Zhang, N. Qin, Targeting Hippo pathway by specific interruption of YAP-TEAD in-
Yan, X. Lu, L. Pham, M.M. Wang, N. Karachaliou, M.G. Cao, J.L. Manzano, J.L. teraction using cyclic YAP-like peptides, FASEB J. 29 (2015) 724–732.
Ramirez, J.M. Torres, F. Buttitta, C.M. Rudin, E.A. Collisson, A. Algazi, E. Robinson, [107] R. Johnson, G. Halder, The two faces of Hippo: targeting the Hippo pathway for re-
I. Osman, E. Munoz-Couselo, J. Cortes, D.T. Frederick, Z.A. Cooper, M. McMahon, generative medicine and cancer treatment, Nat. Rev. Drug Discov. 13 (2014) 63–79.
A. Marchetti, R. Rosell, K.T. Flaherty, J.A. Wargo, T.G. Bivona, The Hippo effector
YAP promotes resistance to RAF- and MEK-targeted cancer therapies, Nat. Genet.
47 (2015) 250–256.

You might also like