Você está na página 1de 14

Journal of Pathology

J Pathol 2011; 223: 205218


Published online 18 October 2010 in Wiley Online Library
(wileyonlinelibrary.com) DOI: 10.1002/path.2785

INVITED REVIEW

The dynamic roles of TGF- in cancer


Erik Meulmeester1 and Peter ten Dijke1,2 *
1 Department of Molecular Cell Biology and Centre for Biomedical Genetics, Leiden University Medical Center, Postbus 9600, 2300 RC, Leiden,
The Netherlands
2 Uppsala University and Ludwig Institute for Cancer Research, Box 595, 75124, Uppsala, Sweden

*Correspondence to: Peter ten Dijke, Department of Molecular Cell Biology and Centre for Biomedical Genetics, Leiden University Medical Center,
Building 2, Room R-02-022, Postzone S-1-P, Postbus 9600, 2300 RC, Leiden, The Netherlands e-mail: p.ten_dijke@lumc.nl

Abstract
The transforming growth factor- (TGF-) signalling pathway plays a critical and dual role in the progression of
human cancer. During the early phase of tumour progression, TGF- acts as a tumour suppressor, exemplified
by deletions or mutations in the core components of the TGF- signalling pathway. On the contrary, TGF- also
promotes processes that support tumour progression such as tumour cell invasion, dissemination, and immune
evasion. Consequently, the functional outcome of the TGF- response is strongly context-dependent including
cell, tissue, and cancer type. In this review, we describe the molecular signalling pathways employed by TGF-
in cancer and how these, when perturbed, may lead to the development of cancer. Concomitantly with our
increased appreciation of the molecular mechanisms that govern TGF- signalling, the potential to therapeutically
target specific oncogenic sub-arms of the TGF- pathway increases. Indeed, clinical trials with systemic TGF-
signalling inhibitors for treatment of cancer patients have been initiated. However, considering the important
role of TGF- in cardiovascular and many other tissues, careful screening of patients is warranted to minimize
unwanted on-target side effects.
Copyright 2010 Pathological Society of Great Britain and Ireland. Published by John Wiley & Sons, Ltd.

Keywords: BMP; cancer; epithelial to mesenchymal transition; EMT; metastasis; signal transduction; Smad; TGF-; ubiquitin

Received 3 August 2010; Revised 18 August 2010; Accepted 1 September 2010

No conflicts of interest were declared.

Introduction
Transforming growth factor- (TGF-) emerged with
the evolution of multi-cellular organisms, where it
plays an essential role in the development of the
body plan during embryogenesis and is crucial for tissue homeostasis. The TGF- pathway mediates such
regulation through control of proliferation, differentiation, apoptosis, adhesion, invasion, and cellular microenvironment [14]. Consequently, malfunctioning of
this pathway has adverse effects and inactivation of
components in this signalling pathway is central to
many diseases including the development of tumourigenesis and tumour progression. In several types of
human carcinomas, mutations or loss of heterozygosity (LOH) in central components of the TGF-
pathway has been observed [5]. These observations
support the idea that the TGF- pathway has a tumoursuppressive role that needs circumvention to allow
tumourigenesis. Also, due to its growth-suppressive
effects, in the past TGF- itself has been regarded
as an attractive cytokine for the treatment of cancer. Therefore, studies were initiated in which TGF-
was explored as an adjuvant for chemotherapy. Indeed,
TGF- was able to protect normal cells and sensitize
tumour cells towards standard chemotherapy in some
Copyright 2010 Pathological Society of Great Britain and Ireland.
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk

pre-clinical models [6,7]. Nonetheless, issues with dosing and timing of TGF- treatment, among others,
halted the translation of these findings towards clinical
application.
Importantly, analysis of human tumour samples
has also suggested an active role for the TGF-
pathway in tumour progression. Immuno-staining for
TGF- correlated with metastasis in breast, colon,
and prostate cancer [810]. In addition, the intensity
of TGF- staining in invading lymph node metastases was higher in breast and colon cancers than
in the primary tumour [11,12]. Moreover, TGF-
increases the motility and invasion of certain cancer cells, demonstrating that these cells, while having lost sensitivity for TGF--induced growth arrest,
remained TGF--responsive [13]. Furthermore, TGF secreted by tumour cells stimulates stroma formation and immune evasion of tumour cells [14,15].
Thus, TGF- has both tumour-suppressive and tumourpromoting functions [16]. In this review, we will
elaborate on the dynamic role that TGF- plays in
cancer biology; how perturbation of the TGF- signalling pathways contributes to cancer; and how new
insights provide opportunities for targeted therapy of
cancer patients.
J Pathol 2011; 223: 205218
www.thejournalofpathology.com

206

Molecular mechanism of TGF- signalling


Canonical signalling via Smad proteins
TGF- is the founding member of a larger family
of secreted dimeric cytokines that comprise TGFs, activins, bone morphogenetic proteins (BMPs),
and growth and differentiation factors (GDFs) [4]. In
mammals, three alternative TGF- isoforms exist, ie
TGF-1, TGF-2, and TGF-3. The bioactive ligands
are composed of homo- or hetero-dimers of polypeptides that are synthesized as precursor molecules and
matured by proteolytic cleavage by endoproteases such
as Furin [1719]. The cleaved N-terminal peptide, also
known as latency-associated peptide (LAP), forms a
non-covalent interaction with the C-terminal peptide
dimer (TGF-). Subsequent binding to the latent TGF-binding protein (LTBP) allows the formation of the
large latency complex (LLC) that facilitates secretion
into the extracellular matrix [20]. By release from this
complex, TGF- is activated.
Active TGF- dimers mediate signalling through the
TGF- type I and type II receptors (TGF-RI and
TGF-RII, respectively) that are endowed with serine/threonine kinase activity [2123]. The membraneanchored proteoglycan betaglycan (TGF-RIII) assists
TGF- binding to TGF-RII [24]. High affinity binding of TGF- to TGF-RII leads to heterotetrameric
complex formation with TGF-RI, which results in the
phosphorylation of TGF-RI by TGF-RII (Figure 1).
In most cell types, TGF-RI (also known as activin
receptor-like kinase 5; ALK5) transduces signalling,
while in certain cell types ALK1 or other type I receptors can mediate signalling responses [25,26]. The type
I receptor propagates signalling by recruitment and
phosphorylation of receptor-regulated Smad (R-Smad)
proteins. Whereas ALK5 signalling is mediated by
phosphorylation of Smad2 and Smad3 proteins, ALK1
signalling is mediated by Smad1, Smad5, and Smad8.
Activated Smads form a complex with the common
Smad (co-Smad; Smad4 in mammals) and shuttle into
the nucleus [21,27]. Since activated Smad complexes
have a weak binding affinity for DNA, additional
DNA-binding transcription factors are required to regulate high affinity interaction and specificity [28,29].
A variety of transcription factor families have been
described to act in concert with Smad proteins such
as p300/CBP, Forkhead, homeobox, zinc-finger, AP1,
Ets, and basic helix-loop-helix families [30,31]. The
diversity of R-Smad/co-Smad/co-factor combinations
regulates the transcription of a vast amount of target
genes. The cell type-specific responses observed upon
TGF- stimulation can at least in part be explained by
the differential expression of these regulators in such
cell types [32].

Regulation of TGF- Smad signalling


To control the intensity and duration of TGF- signalling, each step in this signalling pathway is under
Copyright 2010 Pathological Society of Great Britain and Ireland.
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk

E Meulmeester and P ten Dijke

intense regulation [33]. Presentation and sequestration


of R-Smads to TGF- receptor is controlled by SARA
and TMEPAI, respectively [34,35]. Phosphorylation
of Smad proteins is reversed by phosphatases (eg
PPM1A, PDP, and SCP1, 2 and 3), thereby creating
a rapid activationdeactivation cycle [3638]. Furthermore, activated Smad2/3 proteins are counteracted
by ubiquitin E3 ligases that target them for proteasomal degradation [39,40]. In addition, the interaction
between phosphorylated Smad2 and Smad4 is impeded
by mono-ubiquitination of Smad4. Ectodermin is a
ubiquitin E3 ligase that targets Smad4 for monoubiquitination, which is counteracted by the deubiquitinating enzyme USP9x [41]. Moreover, the inhibitory
Smads (I-Smads), Smad6 and Smad7, are transcriptionally induced upon BMP and TGF- signalling [4245].
While Smad6 mainly inhibits BMP signalling, Smad7
has the capability to prevent both BMP and TGF-
signalling. Mechanistically, it has been proposed that
I-Smads repress signalling by competing with R-Smads
for receptor binding [43,44,46]. Secondly, it was shown
that overexpression of Smad6 prevents the formation of
Smad1Smad4 complexes, by binding to Smad1 [47].
Smad7 has been implicated in the recruitment of the
E3 ubiquitin ligases Smurf1 and Smurf2 to activated
TGF- receptors, which in turn leads to their degradation via the ubiquitin proteasome system [4850].
The importance of Smad7-mediated inhibition of TGF signalling is underscored by the observation that
Smad7 is overexpressed in endometrial carcinomas and
thyroid follicular cell lines [51,52]. Another point of
regulation of TGF- signalling is by the transcriptional
repressors Ski, SnoN, and members of the HDAC family that interact with Smad2/3 and Smad4 [5356].
Even though Smad signalling is required for the
majority of TGF--mediated signalling, not all responses to TGF- are solely dependent on Smad4
[57,58]. For example, transcriptional intermediary factor 1 (TIF1) selectively binds receptor-phosphorylated Smad2/3 in competition with Smad4, in the
control of haematopoietic cell fate by TGF- [59]. In
addition, IKK was identified as a critical co-regulator
of Smad2/3 in a Smad4-independent manner, which
controls keratinocyte differentiation [60].

Non-canonical signalling
While canonical signalling directly regulates the transcription of Smad-dependent target genes, Smad
proteins have also been shown to participate in sequestration, recruitment, and enzyme activation [6163].
R-Smads (independent of Smad4) are also involved in
the regulation of miRNA maturation in the nucleus
[64]. Moreover, alternative signalling modules are
present in parallel with Smad signalling that are
also responsive to TGF-. The existence of Smadindependent signalling is supported by the identification of the TRAF6TAK1p38/JNK pathway as
a TGF- signalling module downstream of TGF-
receptors [6567]. Furthermore, TGF- signalling is
J Pathol 2011; 223: 205218
www.thejournalofpathology.com

TGF- in cancer

207

Figure 1. The TGF- signalling pathway. Transforming growth factor- (TGF-) present in the extracellular matrix resides in the large
latent complex (LLC). Upon release, active TGF- weakly interacts with the large betaglycan membrane protein that is present in vast
excess compared with TGF-RII. Subsequently, TGF- is presented to TGF-RII, which leads to the formation of a heterotetrameric
complex between the serine/threonine kinases TGF-RI and TGF-RII. The constitutive active TGF-RII phosphorylates (P) TGF-RI, which
in turn recruits, phosphorylates, and activates Smad transcription factors (Smad2/3). Phosphorylated Smad2/3 form a complex with the
co-Smad (Smad4), translocate into the nucleus, and further build transcription complexes with additional co-repressors or co-activators to
regulate the expression of a wide variety of genes. The inhibitory Smad (Smad7) reduces further signalling by preventing phosphorylation
of Smad2/3. Besides Smad-mediated signalling, TGF- also activates several other signalling cascades such as TRAF6TAK1p38/JNK,
RhoARhock1, and Par6.

engaged in RhoARock1 signalling that is required


for the epithelial-to-mesenchymal transition (EMT)
[68,69]. The TGF- receptor also activates Erk-MAP
kinase signalling through direct phosphorylation of Shc
on tyrosine and serine residues by TGF-RI [70]. TGFRII can also signal independently of TGF-RI by
direct phosphorylation of, for example, Par6 that contributes to EMT by a loss of tight junctions [71]. While
an increasing number of proteins have been identified
to interact with the TGF- receptors [72,73], novel
mechanisms of Smad-independent signalling will most
likely be uncovered in the near future.
Copyright 2010 Pathological Society of Great Britain and Ireland.
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk

Mechanisms of TGF- signalling in tumour


suppression
In several cancers, deletion or mutations of TGF- signalling components alleviate the tumour-suppressive
effects of this pathway. In colon cancer with microsatellite instability, TGF-RII is subjected to accumulation
of replication errors, leading to inactivation of the
receptor and thus the TGF- pathway [74]. Similarly,
mutations in TGF-RII have been described in gastric
tumours, gliomas, and colorectal cancer [75,76]. While
mutations in TGF-RI are less frequent, these have
J Pathol 2011; 223: 205218
www.thejournalofpathology.com

208

been observed in pancreas, ovarian, and breast cancer [7779]. Downstream of the receptors, inactivating
mutations or deletions of SMAD4 are present in half
of the pancreatic tumours and to a lesser extent in gastrointestinal tumours [75,78]. However, tissue-specific
inactivation of Smad4 or TGF-RII alone in mouse
models rarely leads to spontaneous tumour formation, suggesting that the outcome of interference with
TGF- signalling is strongly dependent on the context
of individual tumours [8085]. In fact, the tumoursuppressive role of TGF- is most apparent under
conditions of oncogenic stress. For example, inactivation of Smad4 or TGF-RII in adenomatosis polyposis
coli (APC)-deficient mice potentiates the progression
of intestinal polyps to carcinoma [83,86]. Similarly,
mammary tumours initiated by polyoma virus middleT oncogene and premalignant lesions initiated by the
KRAS oncoprotein in skin, pancreas, and oral and
oesophagus epithelium progress faster when TGF-
signalling is crippled [8082,84].
Mechanistically, TGF- has several operating arms
to achieve its tumour-suppressive effect which include
regulation of cell proliferation, apoptosis, and indirectly through the tumour stroma.

Regulation of cell proliferation and apoptosis


In epithelial, neuronal, and haematopoietic cells, TGF limits cell proliferation through a coordinated programme of cytostatic gene responses (Figure 2A).
At the core of these events is the induction of
cyclin-dependent kinase (CDK) inhibitors such as

E Meulmeester and P ten Dijke

p15 (INK4B) and p21(WAF1), which are driven by


Smad3/Smad4 complexes in concert with FoxO and
Sp1 transcription factors [8791]. p15 inhibits cell
cycle progression at the late G1 phase by interacting with CDK4/6 and preventing their interaction with
cyclin D [92]. Consequently, the CDK inhibitor p27
is relocated from cyclin DCDK4 complexes to interact with and inhibit cyclin ECDK2 complexes. p21
also inhibits the activity of cyclin ECDK2 complexes
[4,92]. The inactivity of these CDK complexes prevents phosphorylation of pRb, a major switch in cell
cycle progression, and thus progression through G1
into S phase. While the induction of p21 and p15
is important for the cytostatic effect in neuronal and
epithelial cells, in haematopoietic cells TGF- mediates its growth inhibition through the CDK inhibitor
p57 [93].
Simultaneously with the activation of cell cycle progression inhibitors, TGF- represses the c-Myc oncogene that promotes cell proliferation. c-Myc is a transcription factor that activates or represses transcription
depending on its target genes. Since c-Myc inhibits
transcriptional activation of p15 and p21, through Miz1, TGF- has a firm grip on the regulation of these
target genes [94,95]. Furthermore, TGF- represses the
expression of Id1, Id2, and Id3, which are nuclear
factors implicated in differentiation and progression
through the G1S cell cycle transition [96]. More
recently, TGF-s anti-proliferative effect was shown
to be dependent on the eukaryotic translation initiation
factor-4F (eIF4F) [97]. Thus, TGF- orchestrates its

Figure 2. TGF- as a tumour suppressor. (A) TGF- controls cell cycle progression by repressing the oncogenic Myc transcription factor,
which in turn prevents transcriptional activation of the CDK inhibitors p21 and p15 via interaction with MIZ1. TGF- also directly
transcriptionally activates p21 and p15. p15 abolishes the interaction of cyclin D (Cyc D) with the CDK4/6 complex, thereby inactivating
the CDK. p21 (and also p27) inactivates the cyclin E (Cyc E) CDK2 complex, which leads to stalling of the cell cycle at the G1S boundary.
(B) TGF- inhibits apoptosis in a variety of cell types; however, the exact molecular mechanisms remain to be determined. (C) Through
non-cell autonomous signalling, TGF- controls tumour progression via inhibition of paracrine growth factors in the tumour stroma during
the early stages of tumour development. The absence of TGF- signalling in tumour stroma cells relieves the growth inhibitory effect on
epithelial cells, which consequently obtain increased growth, migratory, and invasive properties.
Copyright 2010 Pathological Society of Great Britain and Ireland.
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk

J Pathol 2011; 223: 205218


www.thejournalofpathology.com

TGF- in cancer

growth-suppressive effects via functionally redundant


mechanisms, ie by slowing down cell growth progression and simultaneously de-accelerating proliferative
actions.
In addition to the regulatory role of TGF- in cell
cycle progression, tumour progression is also limited
through control of the apoptotic response. TGF- has
been reported to induce apoptosis in a variety of cell
types under physiological circumstances; however, the
molecular mechanisms remain ill defined (Figure 2B).
Candidates that contribute to the apoptotic functions
of TGF- include the death receptor FAS, growth
arrest and DNA damage inducible 45 (GADD45),
BIM, and death-associated kinase (DAPK) [98101].
Confirmation of the physiological relevance of these
candidates awaits experimental proof using in vivo
model systems. A further thorough understanding of
the mechanisms by which TGF- exerts its cytostatic
and apoptotic responses is important to identify key
points where tumours may hijack the TGF- system to
favour tumour progression.

Indirect regulation of tumour suppression


In addition to cell autonomous regulation of apoptosis
and growth arrest, TGF- further restricts growth and
tumour progression of epithelial cells through blockage
of paracrine factor production in the tumour stroma
(Figure 2C). While the impact of the stromal fibroblast on tumour progression has been known since
early studies on breast cancer, the role of TGF- in
this process emerged from mouse models in which
TGF- signalling was impaired in stromal fibroblasts
[102,103]. Overexpression of a kinase inactive mutant
of TGF-RII in the mammary stromal cells results
in epithelial hyperplasia and increased production of
hepatocyte growth factor (HGF) [103]. These observations were strengthened by a mouse model in which
TGF-RII was conditionally inactivated in fibroblasts,
which results in the development of squamous cell carcinoma of the forestomach and prostatic intraepithelial
neoplasia [104]. The elevated expression of HGF in
the TGF-RII-deficient fibroblast led to the activation
of its receptor (Met) in adjacent epithelial cells. Furthermore, the loss of TGF-RII in fibroblasts results
in increased TGF- and macrophage-stimulating protein (MSP), which all together contribute to augmented
tumour growth, motility, and invasion [105]. Moreover,
in an E-Myc transgenic mouse lymphoma model,
TGF- secreted from non-neoplastic macrophages
prevents tumour development by promoting cellular
senescence [106].

Tumour-promoting roles of TGF- signalling


During the early stages of tumour development, the
TGF- pathway operates as a tumour suppressor,
thereby preventing tumour growth. As seen in large
Copyright 2010 Pathological Society of Great Britain and Ireland.
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk

209

subsets of colon, pancreatic, and gastric cancers, mutation or deletion of TGF- receptors or Smads in
the TGF- pathway leads to its inactivation or perturbation of signalling responses [7476,107110].
Tumours that acquire the ability to bypass the tumoursuppressive arms may exploit certain aspects of TGF-
signalling to actively promote tumour cell progression.
In fact, aggressive tumours resistant to the tumoursuppressive effects of TGF- preserve the core components of TGF- signalling. Different types of tumours
such as gliomas and breast and prostate cancer seem to
acquire mutations preferentially not in the core components of TGF- signalling [109,111114]. Such human
tumours likely obtain resistance to TGF--mediated
growth arrest and importantly retain the ability to
exploit TGF- signalling to induce pathways that promote EMT, tumour invasion, metastatic dissemination,
and evasion of the immune system. Therefore, tumours
with such a signature are highly aggressive.

Epithelial-to-mesenchymal transition (EMT)


and invasion
EMT is a key process during embryonic development
that contributes to the formation of the body plan and
allows differentiation of multiple tissues and organs.
Furthermore, it plays an important role during wound
healing, where keratinocytes recapitulate part of the
EMT process to acquire a more migratory character
to seal the wounded area. Importantly, EMT also
plays a pivotal role in pathological disorders such
as fibrosis and cancer progression. Even though the
EMT process is well accepted in a variety of cancer
cell models, the criticism on its relevance for human
cancer progression has only recently dwindled, due
to convincing morphological evidence of EMT at
the invasive fronts of human tumours [115117]. To
invade normal tissue and spread to distant organs,
carcinoma cells need to lose cell polarity, cellcell
contacts, and acquire fibroblastic-like properties. In
this process of EMT, cells become highly motile
and invasive, which allows survival in an anchorageindependent environment and provides them with stem
cell-like properties (Figure 3). A molecular hallmark
for cells undergoing EMT is decreased expression
of epithelial cellcell junction proteins, such as Ecadherin and zona occludens (ZO-1), while at the same
time acquiring the expression of mesenchymal markers,
such as vimentin, -smooth muscle actin (-SMA), and
fibronectin [1,118121]. Furthermore, up-regulation
of matrix metalloproteases and N-cadherin led to
the degradation of extracellular matrix proteins and
rendered tumour cells more migratory, respectively.
The identification of TGF- as a major inducer of
EMT came from studies in cell culture. Treatment
of normal mouse breast epithelial cells with TGF-
changes the cuboidal shape to an elongated spindle,
accompanied by a decrease in epithelial markers and
increased expression of mesenchymal markers [122].
Canonical TGF-Smad signalling plays a pivotal role
J Pathol 2011; 223: 205218
www.thejournalofpathology.com

210

E Meulmeester and P ten Dijke

Figure 3. TGF- as a tumour promoter. (A) TGF- stimulates the progression of carcinoma in situ towards a more aggressive, motile, and
invasive carcinoma. The tumour cells (yellow) acquire the capacity to invade adjacent tissues and intravasate into the blood stream.
(B) TGF- promotes metastatic dissemination of primary breast carcinoma towards bone by up-regulating PTHrP and IL11, which in turn
activate osteoblasts (green) to secrete RANKL. RANKL leads to the differentiation of precursor cells (blue) into osteoclasts (orange), which
absorb the bone mass and release stored TGF-. As such, a feed-forward loop is created in which TGF- promotes the growth of bone
metastasis. (C) By inducing ANGPTL4, TGF- primes tumour cells for dissemination towards the lungs. ANGPTL4 enhances extravasation
by dissociating vascular endothelial cellcell junctions. TGF-, released by the tumour cells, allows evasion of the immune response by
inactivating CD4+ and CD8+ T-cells that normally inhibit the growth of tumour cells.

in this process, as depletion of Smad3/Smad4 or overexpression of Smad7 completely abolishes induction


of EMT [123125]. Additionally, increased expression of Smad3 and Smad4 in the presence of constitutive active TGF-RI enhances induction of EMT
[126]. Mechanistically, TGF- orchestrates a diverse
transcriptional network that involves transcriptional
activation of Snail, ZEB, and the BHLH family of
proteins [119]. On the other hand, TGF- mediates
down-regulation of the microRNA-200 family that is
required for epithelial cells to undergo EMT [127].
The expression of these factors is regulated by both
Smad-dependent and Smad-independent mechanisms
that lead to the repression of epithelial cell markers
and induction of mesenchymal genes.
Shortly after the identification of TGF- as a regulator of EMT in cell culture conditions, work from
mouse models confirmed TGF- as a critical regulator
of EMT in vivo. During normal craniofacial development, TGF-3 is required for the fusion of the two
Copyright 2010 Pathological Society of Great Britain and Ireland.
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk

palatal shelves, a process that is dependent on EMT.


Consequently, knockout mice for TGF-3 are confronted with a cleft palate [128,129]. Furthermore,
TGF-2 knockouts suffer from a wide variety of developmental defects that can at least in part be explained
by a dysfunctional EMT process [130]. Additionally,
in a mouse model for skin carcinogenesis, increased
expression of TGF-1 in keratinocytes enhances EMT
and increases the rate and aggressiveness of these
tumours, induced by chemical carcinogenesis [131]. In
this model, keratinocyte-specific depletion of Smad2
stimulates EMT by promoting Smad3/Smad4-mediated
Snail transcription, indicating that Smad2 and Smad3
have differential functions [132]. Evidence for TGF as a regulator of EMT in human cancer has been
suggested by Shipitsin et al, who describe an activated TGF- pathway in isolated CD44-positive cancerous cells compared with normal tissue [133]. It
appears that during the TGF--induced EMT process,
epithelial cells also acquire stem cell-like properties
J Pathol 2011; 223: 205218
www.thejournalofpathology.com

TGF- in cancer

besides obtaining a mesenchymal phenotype [134].


Interestingly, Ikushima et al recently reported mechanistic insights into how TGF- maintains stemness in
glioma initiating cells [135]. TGF- directly regulates
the expression of Sox4, which in turn mediates the
induction of Sox2 expression [135].

Metastatic dissemination
After invasion into adjacent tissue, the metastatic
process continues through intravasation, dissemination
to distal capillary beds, extravasation, and growth in a
distal organ [136,137]. Particular tumour types have a
tendency to metastasize to certain organs; for example,
breast tumours tend to metastasize towards the brain,
lungs, bones, and liver. This distribution pattern seems
to be dependent on the expression of a specific set of
genes rather than vasculature, blood flow, and number
of cells delivered to the receiving organ [138140].
The role of TGF- in the metastatic process became
evident with the observation that TGF- immunostaining was much stronger in metastasized breast
tumours than in the primary tumour [11]. In addition,
expression of TGF-RII has a negative correlation with
overall survival in oestrogen receptor (ER)-negative
breast cancer patients [141]. Importantly, chemotherapy or radiation treatment in the MMTV/PyVmT transgenic model of breast metastasis led to increased TGF1 levels as well as increased circulating tumour cells
and lung metastasis [142]. Administration of neutralizing TGF- antibodies prevents this increase in metastasis, thus suggesting an important role for TGF- in the
metastatic process [142]. However, the role of TGF-
in metastasis progression seems to be strongly contextdependent. For example, expression of an activated
TGF-RI transgene increases metastasis to the lungs,
while targeted deletion of TGF-RII resulted in a similar observation [80,143,144]. The picture of TGF-
as a promoter of metastasis is even further obscured
by the observation that short-term stimulation with
TGF- stimulates metastasis formation, while persistent stimulation decreases the metastatic spread to the
lungs [145]. From this study, it has been proposed that
TGF- signalling initially needs to be high in order
to acquire invasive properties for dissemination, while
upon extravasation TGF- signalling is low to allow
proliferation at the secondary site.
Insights into the mechanism by which TGF- stimulates metastatic dissemination came from studies on
bone and lung metastasis. Primary breast and prostate
tumours have in common that they often metastasize
towards bone, a process in which TGF- plays an
important role upon the arrival of tumour cells. The
presence of metastatic tumour cells in the bone microenvironment leads to activation of osteoclasts, which
degrade the bone matrix and consequently release
stored TGF- and other growth factors. TGF- in
turn stimulates the metastatic cells to release osteolytic cytokines such as, amongst others, parathyroid
Copyright 2010 Pathological Society of Great Britain and Ireland.
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk

211

hormone-related protein (PTHrP) [146,147]. The posttranslationally increased levels of PTHrP subsequently
stimulate production of RANK ligand (RANKL) in
osteoblasts, which in turn promotes the differentiation of osteoclast precursors and bone resorption
(Figure 3B) [148,149]. Furthermore, it was recently
demonstrated that TGF-RII also directly phosphorylates the PTH receptor and via this route could also
contribute to bone resorption [72].
Transcriptome comparison, of adrenal with bone
metastasis, originated from a breast cancer epithelial
cell line inoculated into mice, led to the identification
of CTGF and IL-11 within a larger gene signature as
critical factors for bone metastasis [138]. The TGF-mediated induction of IL-11 and CTGF, as well
as the acquired metastatic capacity, requires Smad4dependent signalling [123,150]. Besides Smad4, Smad3
is required for the induction of key target genes
in metastasis [151]. Importantly, depletion of Smad2
enhanced the metastatic process of MDA-MB-231
cells, while Smad3 prevented the metastatic spread
[151].
Besides stimulating metastatic colonization, TGF-
also has the ability to prime tumour cells for metastatic
dissemination. By employing a cell culture-derived
TGF- gene response signature, angiopoietin-like 4
(ANGPTL4) was identified as a key player to prime
breast cancer cells for metastasis towards the lungs
[152]. ANGPTL4 assists the tumour cells to disrupt
the lung capillaries and thereby enables pulmonary
metastasis (Figure 3C). In a previous study ANGPLT4
has also been identified within a TGF--responsive
gene signature that mediates metastasis towards the
lungs [139]. The presence of dominant negative TGFRI or the absence of Smad4 in ER-negative breast
cancer cells prevented their capacity to metastasize
when implanted as mammary tumours.
These studies shed light on a role for TGF- in
the metastatic process, both in the early invasion and
intravasation stage and during the later process involving extravasation and colony formation. However, most
of these studies rely on either inoculation of cell
lines or transplantation of tumours. For future work,
it would be important to study these phenomena in a
more physiological setting employing mouse models in
which metastases arise spontaneously from the primary
tumour. In such a setting, the role of TGF- in the various metastatic processes for different types of cancer
could be investigated employing inducible knockout or
knock-in models.

Immune evasion
In addition to regulation of EMT, invasion, and
metastatic dissemination, TGF- also supports tumour
progression by evading the immune system [2]. The
first evidence for TGF- in immune evasion came
from the observation that TGF- potently inhibits
tumour-induced CD8+ cytolytic T-lymphocyte (CTL)mediated rejection of a murine tumour [153]. Further
J Pathol 2011; 223: 205218
www.thejournalofpathology.com

212

support for TGF- as an immuno-suppressant came


from a transgenic mouse model expressing a dominant negative form of TGF-RII (dn-TGF-RII) in
all T-cells [154]. The absence of TGF- signalling in
T-lymphocytes suppressed growth and metastasis in
this mouse model when challenged with melanoma
or lymphoma cell lines. By employing differential
immuno-depletion in these mice, it was determined that
both T-helper (CD4+ ) and cytolytic T-cells (CD8+ )
were responsible for tumour resistance. In addition,
inactivation of TGF- signalling in CTLs (overexpression of dn-TGF-RII) in a mouse model for spontaneous prostate cancer delays tumour progression [155].
Mechanistically, it was demonstrated that TGF- transcriptionally represses the production of several proapoptotic factors in CTLs such as perforin, granzyme
A, granzyme B, FAS ligand, and interferon- [156].
In addition to its immuno-suppressive effect on CTLs,
TGF- also represses the activity of natural killer (NK)
cells. Inhibition of TGF- using neutralizing TGF-
antibodies increased NK cell activity and consequently
resulted in the suppression of metastasis formation
of an inoculated breast carcinoma cell line [157].
CD4+ CD25+ regulatory T-cells that produce high
amounts of TGF- take part in the inhibition of NK cell
activity [158]. TGF- inhibits NK-mediated cytotoxicity through transcriptional repression of the activating
receptor NKG2D and the type I transmembrane protein NKp30 [159,160]. Neutralizing antibodies towards
TGF- could restore the NKG2D expression, as well
as the anti-tumour reactivity [161]. In glioblastoma
patients, TGF- decreases the expression of NKG2D
in CTLs and NK cells, and represses the expression of
MICA (the ligand of NKG2D) in glioma cells [162].

Therapeutic exploitment of the TGF- pathway


A rationale for designing therapeutic inhibitors that
block TGF- signalling in human cancer was initiated
by the observation that excess TGF- promotes tumour
progression. The current strategies to target TGF-
mainly focus on general inhibition of TGF- signalling,
which can be subdivided into three major approaches.
Firstly, the synthesis of TGF- can be prevented by
administration of antisense molecules. Secondly, ligand
traps (including monoclonal TGF- neutralizing antibodies and soluble TGF-RII or TGF-RIII) prevent
soluble TGF- from activating its signalling cascade.
Thirdly, small molecule inhibitors hinder the kinase
activity of TGF- receptors.

Targeting TGF- signalling using antisense


molecules
During tumour progression, the production of TGF-
is often enhanced either by fibroblasts in the stroma or
by the tumour itself, which correlates with increased
aggressiveness of the tumour. One strategy to diminish
Copyright 2010 Pathological Society of Great Britain and Ireland.
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk

E Meulmeester and P ten Dijke

the production of the ectopic TGF- is by using antisense molecules, which block TGF--mediated gene
expression. Antisense molecules for TGF-1 demonstrate promising results in pre-clinical experiments
by decreasing tumourigenicity [163,164]. A promising
candidate in further development is AP 12 009 (Antisense Pharma), an antisense molecule directed against
TGF-2, which reduces TGF-2-mediated migration
and proliferation, as well as alleviating its immunosuppressive effects [165,166]. The efficacy in phase
I/II clinical trials to treat refractory high-grade glioma
patients exceeded the expected median survival compared with chemotherapy [166]. These results suggest
AP 12 009 as a promising therapeutic target to inhibit
TGF- signalling for malignant tumour therapy. Currently, AP 12 009 is in further clinical studies to test
its efficacy for metastatic melanoma, pancreatic carcinoma, and metastatic colorectal carcinoma.

Ligand traps that sequester TGF-


Another approach to prevent TGF- signalling is
trapping the TGF- ligand employing soluble TGF receptors or neutralizing antibodies. Soluble TGFRII and TGF-RIII have proven their value in
pre-clinical model systems. For example, expression
of soluble TGF-RII prevents pulmonary, pancreatic
or liver metastasis in mouse models [167169], while
intraperitoneal delivery of the soluble ectodomain of
TGF-RIII into athymic nude mice inhibited pulmonary metastasis [170]. Alternatively, TGF- neutralizing antibodies such as 2G7 and 1D11 prevent the
interaction of TGF- with its receptors and thereby
restrain its biological activity [157,171]. Administration of 1D11 following inoculation of 4T1 cells suppressed pulmonary metastasis, while 2G7 transiently
reduced the growth rate of intra-abdominal transplanted
MDA-MB-231 cells and lung metastasis [157,172].
Based on these results, Genzyme developed a monoclonal neutralizing TGF- antibody (GC1008) for
which recently a clinical phase I/II trial was completed to assess the safety and efficacy in patients with
advanced metastatic melanoma or renal cell carcinoma.

Small molecule inhibitors of TGF- signalling


Additionally, small molecule inhibitors that restrain the
kinase activity of TGF- receptors can prevent TGF-
signalling. The compound SB-431 542 prevents TGF-mediated Smad2/3 phosphorylation by restraining the
kinase activity of TGF-RI [173]. Treatment of human
osteosarcoma cells with SB-431 542 inhibits TGF-induced proliferation, while in malignant glioma cells
angiogenesis, proliferation, and motility are inhibited
[174,175]. Another inhibitor of TGF-RI is SD-208,
which upon systemic administration increases median
survival upon implantation of malignant glioma cells
into the mouse brain [176]. In another study, SD-208
was reported to inhibit TGF--induced EMT, migration, and invasion of two murine breast carcinoma cell
lines, while also inhibiting the metastasis formation of
J Pathol 2011; 223: 205218
www.thejournalofpathology.com

TGF- in cancer

these cell lines when inoculated into the mammary fat


pads of mice [177]. The compound Ki26894, also a
TGF-RI kinase inhibitor, prevents the motility and
invasion of breast and gastric cancer cells [178,179].
Furthermore, LY2 109 761 is a small molecule that
inhibits the kinase activity of both TGF-RI and TGFRII, and has been shown to inhibit the metastatic
spread to a variety of organs [180,181]. The small
molecule compound LY2 157 299 (Eli Lilly & Co) is
currently in a clinical phase I trial to address its safety
and pharmacokinetics. LY2 157 299 has been reported
in mouse studies to inhibit the primary tumour growth
of a breast and lung cancer cell line [182].

Conclusion and future perspectives


With the continuously increasing appreciation of the
clinical importance of TGF- as a tumour promoter,
interest in targeting the TGF- pathway for therapeutic
intervention is rising. The use of genetic screens that
delineate the tumour-suppressive versus the tumourpromoting roles of TGF- signalling will provide a
basis for new research that will enable the targeting of its specific oncogenic sub-arms. Several preclinical studies led to clinical trials that underline the
potential to restrain TGF- signalling [183,184]. Furthermore, treatment using soluble TGF-RII prevents
the metastatic spread in a mouse model, while severe
pathological disorders observed as in TGF- knockout
mice remain absent [168]. However, some concerns
regarding the general systemic inhibition of TGF-
remain, due to its tumour-suppressive role [7479].
Furthermore, TGF- also has a profound role in the cardiovascular system, such that general TGF- inhibitors
will most likely have adverse on-target side effects.
Therefore, to determine the treatment regimen, it is of
critical importance to define and select those patients
who are confronted with tumours in which TGF-s
tumour-promoting role prevails. Although individualbased medicine still has a long way to go, the current
state-of-the-art techniques such as microarray, protein microarray or super-SILAC pave the way for a
detailed view of key corrupted pathways in individuals
[185187].

Acknowledgment
We are grateful to all laboratory members for stimulating discussions. We apologize to those whose work
we could not cite due to limited space constraints.
Our work is supported by grants from the LUMC
(vrije beleidsruimte), the Dutch Cancer Society, the
Netherlands Organization for Health and Development,
the Swedish Cancer Foundation, and the Centre for
Biomedical Genetics.

Author contribution statement


EM wrote the manuscript under the supervision of PtD.
Copyright 2010 Pathological Society of Great Britain and Ireland.
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk

213

Teaching materials
PowerPoint slides of the figures from this review
are supplied as supporting information in the online
version of this article.

References
1. Heldin CH, Landstrom M, Moustakas A. Mechanism of TGF-
signaling to growth arrest, apoptosis, and epithelialmesenchymal
transition. Curr Opin Cell Biol 2009; 21: 166176.
2. Yang L, Pang Y, Moses HL. TGF- and immune cells: an important regulatory axis in the tumor microenvironment and progression. Trends Immunol 2010; 31: 220227.
3. Ikushima H, Miyazono K. TGF signalling: a complex web in
cancer progression. Nature Rev Cancer 2010; 10: 415424.
4. Massague J. TGF in Cancer. Cell 2008; 134: 215230.
5. Levy L, Hill CS. Alterations in components of the TGF- superfamily signaling pathways in human cancer. Cytokine Growth
Factor Rev 2006; 17: 4158.
6. Sonis ST, Van Vugt AG, Brien JP, et al . Transforming growth
factor-3 mediated modulation of cell cycling and attenuation
of 5-fluorouracil induced oral mucositis. Oral Oncol 1997; 33:
4754.
7. Goey H, Keller JR, Back T, et al . Inhibition of early murine
hemopoietic progenitor cell proliferation after in vivo locoregional
administration of transforming growth factor-1. J Immunol 1989;
143: 877880.
8. Wikstrom P, Stattin P, Franck-Lissbrant I, et al . Transforming
growth factor 1 is associated with angiogenesis, metastasis,
and poor clinical outcome in prostate cancer. Prostate 1998; 37:
1929.
9. Walker RA, Dearing SJ. Transforming growth factor 1 in ductal
carcinoma in situ and invasive carcinomas of the breast. Eur
J Cancer 1992; 28: 641644.
10. Friedman E, Gold LI, Klimstra D, et al . High levels of transforming growth factor 1 correlate with disease progression in
human colon cancer. Cancer Epidemiol Biomarkers Prev 1995;
4: 549554.
11. Dalal BI, Keown PA, Greenberg AH. Immunocytochemical localization of secreted transforming growth factor-1 to the advancing
edges of primary tumors and to lymph node metastases of human
mammary carcinoma. Am J Pathol 1993; 143: 381389.
12. Picon A, Gold LI, Wang J, et al . A subset of metastatic human
colon cancers expresses elevated levels of transforming growth
factor 1. Cancer Epidemiol Biomarkers Prev 1998; 7: 497504.
13. ten Dijke P, Goumans MJ, Itoh F, et al . Regulation of cell
proliferation by Smad proteins. J Cell Physiol 2002; 191: 116.
14. Ikushima H, Miyazono K. Cellular context-dependent colors of
transforming growth factor- signaling. Cancer Sci 2010; 101:
306312.
15. Flavell RA, Sanjabi S, Wrzesinski SH, et al . The polarization of
immune cells in the tumour environment by TGF. Nature Rev
Immunol 2010; 10: 554567.
16. Roberts AB, Wakefield LM. The two faces of transforming
growth factor in carcinogenesis. Proc Natl Acad Sci U S A 2003;
100: 86218623.
17. Dubois CM, Laprise MH, Blanchette F, et al . Processing of
transforming growth factor 1 precursor by human furin convertase. J Biol Chem 1995; 270: 1061810624.
18. Cheifetz S, Weatherbee JA, Tsang ML, et al . The transforming
growth factor- system, a complex pattern of cross-reactive
ligands and receptors. Cell 1987; 48: 409415.
J Pathol 2011; 223: 205218
www.thejournalofpathology.com

214

19. Ling N, Ying SY, Ueno N, et al . Pituitary FSH is released by


a heterodimer of the -subunits from the two forms of inhibin.
Nature 1986; 321: 779782.
20. Janssens K, ten Dijke P, Janssens S, et al . Transforming growth
factor-1 to the bone. Endocr Rev 2005; 26: 743774.
21. Shi Y, Massague J. Mechanisms of TGF- signaling from cell
membrane to the nucleus. Cell 2003; 113: 685700.
22. Heldin CH, Miyazono K, ten Dijke P. TGF- signalling from cell
membrane to nucleus through SMAD proteins. Nature 1997; 390:
465471.
23. Feng XH, Derynck R. Specificity and versatility in TGF- signaling through Smads. Annu Rev Cell Dev Biol 2005; 21: 659693.
24. Lopez-Casillas F, Wrana JL, Massague J. Betaglycan presents ligand to the TGF signaling receptor. Cell 1993; 73:
14351444.
25. Goumans MJ, Valdimarsdottir G, Itoh S, et al . Activin receptorlike kinase (ALK)1 is an antagonistic mediator of lateral
TGF/ALK5 signaling. Mol Cell 2003; 12: 817828.
26. Daly AC, Randall RA, Hill CS. Transforming growth factor
-induced Smad1/5 phosphorylation in epithelial cells is mediated by novel receptor complexes and is essential for anchorageindependent growth. Mol Cell Biol 2008; 28: 68896902.
27. Hill CS. Nucleocytoplasmic shuttling of Smad proteins. Cell Res
2009; 19: 3646.
28. Shi Y, Wang YF, Jayaraman L, et al . Crystal structure of a Smad
MH1 domain bound to DNA: insights on DNA binding in TGF-
signaling. Cell 1998; 94: 585594.
29. Derynck R, Zhang Y, Feng XH. Smads: transcriptional activators
of TGF- responses. Cell 1998; 95: 737740.
30. Koinuma D, Tsutsumi S, Kamimura N, et al . Chromatin immunoprecipitation on microarray analysis of Smad2/3 binding
sites reveals roles of ETS1 and TFAP2A in transforming growth
factor signaling. Mol Cell Biol 2009; 29: 172186.
31. Koinuma D, Tsutsumi S, Kamimura N, et al . Promoter-wide
analysis of Smad4 binding sites in human epithelial cells. Cancer
Sci 2009; 100: 21332142.
32. Itoh S, Itoh F, Goumans MJ, et al . Signaling of transforming
growth factor- family members through Smad proteins. Eur
J Biochem 2000; 267: 69546967.
33. Itoh S, ten Dijke P. Negative regulation of TGF- receptor/Smad
signal transduction. Curr Opin Cell Biol 2007; 19: 176184.
34. Tsukazaki T, Chiang TA, Davison AF, et al . SARA, a FYVE
domain protein that recruits Smad2 to the TGF receptor. Cell
1998; 95: 779791.
35. Watanabe Y, Itoh S, Goto T, et al . TMEPAI, a transmembrane
TGF--inducible protein, sequesters Smad proteins from active
participation in TGF- signaling. Mol Cell 2010; 37: 123134.
36. Lin X, Duan X, Liang YY, et al . PPM1A functions as a Smad
phosphatase to terminate TGF signaling. Cell 2006; 125:
915928.
37. Sapkota G, Knockaert M, Alarcon C, et al . Dephosphorylation of
the linker regions of Smad1 and Smad2/3 by small C-terminal
domain phosphatases has distinct outcomes for bone morphogenetic protein and transforming growth factor- pathways. J Biol
Chem 2006; 281: 4041240419.
38. Chen HB, Shen J, Ip YT, et al . Identification of phosphatases for
Smad in the BMP/DPP pathway. Genes Dev 2006; 20: 648653.
39. Lo RS, Massague J. Ubiquitin-dependent degradation of TGF-activated smad2. Nature Cell Biol 1999; 1: 472478.
40. Gao S, Alarcon C, Sapkota G, et al . Ubiquitin ligase Nedd4L
targets activated Smad2/3 to limit TGF- signaling. Mol Cell
2009; 36: 457468.
41. Dupont S, Mamidi A, Cordenonsi M, et al . FAM/USP9x, a deubiquitinating enzyme essential for TGF signaling, controls
Smad4 monoubiquitination. Cell 2009; 136: 123135.
Copyright 2010 Pathological Society of Great Britain and Ireland.
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk

E Meulmeester and P ten Dijke

42. Afrakhte M, Moren A, Jossan S, et al . Induction of inhibitory


Smad6 and Smad7 mRNA by TGF- family members. Biochem
Biophys Res Commun 1998; 249: 505511.
43. Nakao A, Afrakhte M, Moren A, et al . Identification of Smad7,
a TGF-inducible antagonist of TGF- signalling. Nature 1997;
389: 631635.
44. Imamura T, Takase M, Nishihara A, et al . Smad6 inhibits signalling by the TGF- superfamily. Nature 1997; 389:
622626.
45. Takase M, Imamura T, Sampath TK, et al . Induction of Smad6
mRNA by bone morphogenetic proteins. Biochem Biophys Res
Commun 1998; 244: 2629.
46. Hayashi H, Abdollah S, Qiu Y, et al . The MAD-related protein
Smad7 associates with the TGF receptor and functions as an
antagonist of TGF signaling. Cell 1997; 89: 11651173.
47. Hata A, Lagna G, Massague J, et al . Smad6 inhibits BMP/Smad1
signaling by specifically competing with the Smad4 tumor suppressor. Genes Dev 1998; 12: 186197.
48. Kavsak P, Rasmussen RK, Causing CG, et al . Smad7 binds to
Smurf2 to form an E3 ubiquitin ligase that targets the TGF
receptor for degradation. Mol Cell 2000; 6: 13651375.
49. Ebisawa T, Fukuchi M, Murakami G, et al . Smurf1 interacts with
transforming growth factor- type I receptor through Smad7
and induces receptor degradation. J Biol Chem 2001; 276:
1247712480.
50. Itoh S, Landstrom M, Hermansson A, et al . Transforming growth
factor 1 induces nuclear export of inhibitory Smad7. J Biol Chem
1998; 273: 2919529201.
51. Dowdy SC, Mariani A, Reinholz MM, et al . Overexpression of
the TGF- antagonist Smad7 in endometrial cancer. Gynecol
Oncol 2005; 96: 368373.
52. Cerutti JM, Ebina KN, Matsuo SE, et al . Expression of Smad4
and Smad7 in human thyroid follicular carcinoma cell lines.
J Endocrinol Invest 2003; 26: 516521.
53. Stroschein SL, Wang W, Zhou S, et al . Negative feedback regulation of TGF- signaling by the SnoN oncoprotein. Science 1999;
286: 771774.
54. Luo K, Stroschein SL, Wang W, et al . The Ski oncoprotein
interacts with the Smad proteins to repress TGF signaling. Genes
Dev 1999; 13: 21962206.
55. Nomura T, Khan MM, Kaul SC, et al . Ski is a component
of the histone deacetylase complex required for transcriptional
repression by Mad and thyroid hormone receptor. Genes Dev
1999; 13: 412423.
56. Akiyoshi S, Inoue H, Hanai J, et al . c-Ski acts as a transcriptional co-repressor in transforming growth factor- signaling
through interaction with smads. J Biol Chem 1999; 274:
3526935277.
57. Subramanian G, Schwarz RE, Higgins L, et al . Targeting endogenous transforming growth factor receptor signaling in SMAD4deficient human pancreatic carcinoma cells inhibits their invasive
phenotype1. Cancer Res 2004; 64: 52005211.
58. Wisotzkey RG, Mehra A, Sutherland DJ, et al . Medea is a
Drosophila Smad4 homolog that is differentially required to
potentiate DPP responses. Development 1998; 125: 14331445.
59. He W, Dorn DC, Erdjument-Bromage H, et al . Hematopoiesis
controlled by distinct TIF1 and Smad4 branches of the TGF
pathway. Cell 2006; 125: 929941.
60. Descargues P, Sil AK, Sano Y, et al . IKK is a critical coregulator of a Smad4-independent TGF-Smad2/3 signaling pathway
that controls keratinocyte differentiation. Proc Natl Acad Sci U S
A 2008; 105: 24872492.
61. Bonni S, Wang HR, Causing CG, et al . TGF- induces assembly
of a Smad2Smurf2 ubiquitin ligase complex that targets SnoN
for degradation. Nature Cell Biol 2001; 3: 587595.
J Pathol 2011; 223: 205218
www.thejournalofpathology.com

TGF- in cancer

62. Conery AR, Cao Y, Thompson EA, et al . Akt interacts directly


with Smad3 to regulate the sensitivity to TGF- induced apoptosis. Nature Cell Biol 2004; 6: 366372.
63. Zhang L, Duan CJ, Binkley C, et al . A transforming growth
factor -induced Smad3/Smad4 complex directly activates protein
kinase A. Mol Cell Biol 2004; 24: 21692180.
64. Hata A, Davis BN. Control of microRNA biogenesis by TGF
signaling pathwaya novel role of Smads in the nucleus.
Cytokine Growth Factor Rev 2009; 20: 517521.
65. Ohkawara B, Shirakabe K, Hyodo-Miura J, et al . Role of the
TAK1NLKSTAT3 pathway in TGF--mediated mesoderm
induction. Genes Dev 2004; 18: 381386.
66. Shim JH, Xiao C, Paschal AE, et al . TAK1, but not TAB 1 or
TAB 2, plays an essential role in multiple signaling pathways
in vivo. Genes Dev 2005; 19: 26682681.
67. Yamashita M, Fatyol K, Jin C, et al . TRAF6 mediates Smadindependent activation of JNK and p38 by TGF-. Mol Cell 2008;
31: 918924.
68. Hutchison N, Hendry BM, Sharpe CC. Rho isoforms have distinct and specific functions in the process of epithelial to mesenchymal transition in renal proximal tubular cells. Cell Signal
2009; 21: 15221531.
69. Bhowmick NA, Ghiassi M, Bakin A, et al . Transforming growth
factor-1 mediates epithelial to mesenchymal transdifferentiation
through a RhoA-dependent mechanism. Mol Biol Cell 2001; 12:
2736.
70. Lee MK, Pardoux C, Hall MC, et al . TGF- activates Erk MAP
kinase signalling through direct phosphorylation of ShcA. EMBO
J 2007; 26: 39573967.
71. Ozdamar B, Bose R, Barrios-Rodiles M, et al . Regulation of the
polarity protein Par6 by TGF receptors controls epithelial cell
plasticity. Science 2005; 307: 16031609.
72. Qiu T, Wu X, Zhang F, et al . TGF- type II receptor phosphorylates PTH receptor to integrate bone remodelling signalling.
Nature Cell Biol 2010; 12: 224234.
73. Derynck R, Zhang YE. Smad-dependent and Smad-independent
pathways in TGF- family signalling. Nature 2003; 425:
577584.
74. Markowitz S, Wang J, Myeroff L, et al . Inactivation of the type
II TGF- receptor in colon cancer cells with microsatellite
instability. Science 1995; 268: 13361338.
75. Grady WM, Myeroff LL, Swinler SE, et al . Mutational inactivation of transforming growth factor receptor type II in microsatellite stable colon cancers. Cancer Res 1999; 59: 320324.
76. Izumoto S, Arita N, Ohnishi T, et al . Microsatellite instability
and mutated type II transforming growth factor- receptor gene
in gliomas. Cancer Lett 1997; 112: 251256.
77. Wang D, Kanuma T, Mizunuma H, et al . Analysis of specific
gene mutations in the transforming growth factor- signal transduction pathway in human ovarian cancer. Cancer Res 2000; 60:
45074512.
78. Goggins M, Shekher M, Turnacioglu K, et al . Genetic alterations
of the transforming growth factor receptor genes in pancreatic
and biliary adenocarcinomas. Cancer Res 1998; 58: 53295332.
79. Chen T, Carter D, Garrigue-Antar L, et al . Transforming growth
factor type I receptor kinase mutant associated with metastatic
breast cancer. Cancer Res 1998; 58: 48054810.
80. Forrester E, Chytil A, Bierie B, et al . Effect of conditional
knockout of the type II TGF- receptor gene in mammary epithelia on mammary gland development and polyomavirus middle
T antigen induced tumor formation and metastasis. Cancer Res
2005; 65: 22962302.
81. Lu SL, Herrington H, Reh D, et al . Loss of transforming growth
factor- type II receptor promotes metastatic head-and-neck
squamous cell carcinoma. Genes Dev 2006; 20: 13311342.
Copyright 2010 Pathological Society of Great Britain and Ireland.
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk

215

82. Ijichi H, Chytil A, Gorska AE, et al . Aggressive pancreatic


ductal adenocarcinoma in mice caused by pancreas-specific
blockade of transforming growth factor- signaling in cooperation with active Kras expression. Genes Dev 2006; 20:
31473160.
83. Munoz NM, Upton M, Rojas A, et al . Transforming growth factor receptor type II inactivation induces the malignant transformation of intestinal neoplasms initiated by Apc mutation. Cancer
Res 2006; 66: 98379844.
84. Guasch G, Schober M, Pasolli HA, et al . Loss of TGF signaling
destabilizes homeostasis and promotes squamous cell carcinomas
in stratified epithelia. Cancer Cell 2007; 12: 313327.
85. Bardeesy N, Cheng KH, Berger JH, et al . Smad4 is dispensable
for normal pancreas development yet critical in progression
and tumor biology of pancreas cancer. Genes Dev 2006; 20:
31303146.
86. Takaku K, Oshima M, Miyoshi H, et al . Intestinal tumorigenesis
in compound mutant mice of both Dpc4 (Smad4) and Apc genes.
Cell 1998; 92: 645656.
87. Gomis RR, Alarcon C, He W, et al . A FoxOSmad synexpression group in human keratinocytes. Proc Natl Acad Sci U S A
2006; 103: 1274712752.
88. Seoane J, Le HV, Shen L, et al . Integration of Smad and forkhead pathways in the control of neuroepithelial and glioblastoma
cell proliferation. Cell 2004; 117: 211223.
89. Pardali K, Kurisaki A, Moren A, et al . Role of Smad proteins and transcription factor Sp1 in p21(Waf1/Cip1) regulation by transforming growth factor-. J Biol Chem 2000; 275:
2924429256.
90. Datto MB, Li Y, Panus JF, et al . Transforming growth factor
induces the cyclin-dependent kinase inhibitor p21 through a p53independent mechanism. Proc Natl Acad Sci U S A 1995; 92:
55455549.
91. Hannon GJ, Beach D. p15INK4B is a potential effector of TGF-induced cell cycle arrest. Nature 1994; 371: 257261.
92. Siegel PM, Massague J. Cytostatic and apoptotic actions of TGF in homeostasis and cancer. Nature Rev Cancer 2003; 3:
807821.
93. Scandura JM, Boccuni P, Massague J, et al . Transforming
growth factor -induced cell cycle arrest of human hematopoietic cells requires p57KIP2 up-regulation. Proc Natl Acad Sci U
S A 2004; 101: 1523115236.
94. Staller P, Peukert K, Kiermaier A, et al . Repression of p15INK4b
expression by Myc through association with Miz-1. Nature Cell
Biol 2001; 3: 392399.
95. Seoane J, Le HV, Massague J. Myc suppression of the p21(Cip1)
Cdk inhibitor influences the outcome of the p53 response to DNA
damage. Nature 2002; 419: 729734.
96. Sikder HA, Devlin MK, Dunlap S, et al . Id proteins in cell
growth and tumorigenesis. Cancer Cell 2003; 3: 525530.
97. Azar R, Alard A, Susini C, et al . 4E-BP1 is a target of Smad4
essential for TGF-mediated inhibition of cell proliferation.
EMBO J 2009; 28: 35143522.
98. Jang CW, Chen CH, Chen CC, et al . TGF- induces apoptosis
through Smad-mediated expression of DAP-kinase. Nature Cell
Biol 2002; 4: 5158.
99. Ohgushi M, Kuroki S, Fukamachi H, et al . Transforming growth
factor -dependent sequential activation of Smad, Bim, and
caspase-9 mediates physiological apoptosis in gastric epithelial
cells. Mol Cell Biol 2005; 25: 1001710028.
100. Yoo J, Ghiassi M, Jirmanova L, et al . Transforming growth
factor--induced apoptosis is mediated by Smad-dependent
expression of GADD45b through p38 activation. J Biol Chem
2003; 278: 4300143007.
J Pathol 2011; 223: 205218
www.thejournalofpathology.com

216

101. Kim SG, Jong HS, Kim TY, et al . Transforming growth factor1 induces apoptosis through Fas ligand-independent activation
of the Fas death pathway in human gastric SNU-620 carcinoma
cells. Mol Biol Cell 2004; 15: 420434.
102. Sakakura T, Sakagami Y, Nishizuka Y. Accelerated mammary
cancer development by fetal salivary mesenchyma isografted to
adult mouse mammary epithelium. J Natl Cancer Inst 1981; 66:
953959.
103. Joseph H, Gorska AE, Sohn P, et al . Overexpression of a kinasedeficient transforming growth factor- type II receptor in mouse
mammary stroma results in increased epithelial branching. Mol
Biol Cell 1999; 10: 12211234.
104. Bhowmick NA, Chytil A, Plieth D, et al . TGF- signaling in
fibroblasts modulates the oncogenic potential of adjacent epithelia.
Science 2004; 303: 848851.
105. Cheng N, Bhowmick NA, Chytil A, et al . Loss of TGF- type
II receptor in fibroblasts promotes mammary carcinoma growth
and invasion through upregulation of TGF--, MSP- and HGFmediated signaling networks. Oncogene 2005; 24: 50535068.
106. Reimann M, Lee S, Loddenkemper C, et al . Tumor stromaderived TGF- limits myc-driven lymphomagenesis via Suv39h1dependent senescence. Cancer Cell 2010; 17: 262272.
107. Myeroff LL, Parsons R, Kim SJ, et al . A transforming growth
factor receptor type II gene mutation common in colon
and gastric but rare in endometrial cancers with microsatellite
instability. Cancer Res 1995; 55: 55455547.
108. Parsons R, Myeroff LL, Liu B, et al . Microsatellite instability
and mutations of the transforming growth factor type II receptor
gene in colorectal cancer. Cancer Res 1995; 55: 55485550.
109. Schutte M, Hruban RH, Hedrick L, et al . DPC4 gene in various
tumor types. Cancer Res 1996; 56: 25272530.
110. Hahn SA, Schutte M, Hoque AT, et al . DPC4, a candidate tumor
suppressor gene at human chromosome 18q21.1. Science 1996;
271: 350353.
111. Vincent F, Hagiwara K, Ke Y, et al . Mutation analysis of the
transforming growth factor type II receptor in sporadic human
cancers of the pancreas, liver, and breast. Biochem Biophys Res
Commun 1996; 223: 561564.
112. Jones E, Pu H, Kyprianou N. Targeting TGF- in prostate cancer:
therapeutic possibilities during tumor progression. Expert Opin
Ther Targets 2009; 13: 227234.
113. Jennings MT, Pietenpol JA. The role of transforming growth
factor in glioma progression. J Neurooncol 1998; 36: 123140.
114. Takenoshita S, Mogi A, Tani M, et al . Absence of mutations in
the analysis of coding sequences of the entire transforming growth
factor- type II receptor gene in sporadic human breast cancers.
Oncol Rep 1998; 5: 367371.
115. Prall F. Tumour budding in colorectal carcinoma. Histopathology
2007; 50: 151162.
116. Kalluri R, Weinberg RA. The basics of epithelialmesenchymal
transition. J Clin Invest 2009; 119: 14201428.
117. Tarin D, Thompson EW, Newgreen DF. The fallacy of epithelial
mesenchymal transition in neoplasia. Cancer Res 2005; 65:
59966000.
118. Moustakas A, Heldin CH. Signaling networks guiding epithelialmesenchymal transitions during embryogenesis and cancer
progression. Cancer Sci 2007; 98: 15121520.
119. Xu J, Lamouille S, Derynck R. TGF--induced epithelial to mesenchymal transition. Cell Res 2009; 19: 156172.
120. Berx G, Raspe E, Christofori G, et al . Pre-EMTing metastasis?
Recapitulation of morphogenetic processes in cancer. Clin Exp
Metastasis 2007; 24: 587597.
121. Huber MA, Kraut N, Beug H. Molecular requirements for epithelialmesenchymal transition during tumor progression. Curr Opin
Cell Biol 2005; 17: 548558.
Copyright 2010 Pathological Society of Great Britain and Ireland.
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk

E Meulmeester and P ten Dijke

122. Miettinen PJ, Ebner R, Lopez AR, et al . TGF- induced transdifferentiation of mammary epithelial cells to mesenchymal
cells: involvement of type I receptors. J Cell Biol 1994; 127:
20212036.
123. Deckers M, van Dinther M, Buijs J, et al . The tumor suppressor
Smad4 is required for transforming growth factor -induced
epithelial to mesenchymal transition and bone metastasis of breast
cancer cells. Cancer Res 2006; 66: 22022209.
124. Valcourt U, Kowanetz M, Niimi H, et al . TGF- and the Smad
signaling pathway support transcriptomic reprogramming during
epithelialmesenchymal cell transition. Mol Biol Cell 2005; 16:
19872002.
125. Dzwonek J, Preobrazhenska O, Cazzola S, et al . Smad3 is a key
nonredundant mediator of transforming growth factor signaling
in Nme mouse mammary epithelial cells. Mol Cancer Res 2009;
7: 13421353.
126. Piek E, Moustakas A, Kurisaki A, et al . TGF-() type I
receptor/ALK-5 and Smad proteins mediate epithelial to mesenchymal transdifferentiation in NMuMG breast epithelial cells.
J Cell Sci 1999; 112: 45574568.
127. Gregory PA, Bracken CP, Bert AG, et al . MicroRNAs as regulators of epithelialmesenchymal transition. Cell Cycle 2008; 7:
31123118.
128. Proetzel G, Pawlowski SA, Wiles MV, et al . Transforming
growth factor-3 is required for secondary palate fusion. Nature
Genet 1995; 11: 409414.
129. Kaartinen V, Voncken JW, Shuler C, et al . Abnormal lung development and cleft palate in mice lacking TGF-3 indicates defects
of epithelialmesenchymal interaction. Nature Genet 1995; 11:
415421.
130. Sanford LP, Ormsby I, Gittenberger-de Groot AC, et al . TGF2
knockout mice have multiple developmental defects that are nonoverlapping with other TGF knockout phenotypes. Development
1997; 124: 26592670.
131. Cui W, Fowlis DJ, Bryson S, et al . TGF1 inhibits the formation
of benign skin tumors, but enhances progression to invasive
spindle carcinomas in transgenic mice. Cell 1996; 86: 531542.
132. Hoot KE, Lighthall J, Han G, et al . Keratinocyte-specific Smad2
ablation results in increased epithelialmesenchymal transition
during skin cancer formation and progression. J Clin Invest 2008;
118: 27222732.
133. Shipitsin M, Campbell LL, Argani P, et al . Molecular definition
of breast tumor heterogeneity. Cancer Cell 2007; 11: 259273.
134. Mani SA, Guo W, Liao MJ, et al . The epithelialmesenchymal
transition generates cells with properties of stem cells. Cell 2008;
133: 704715.
135. Ikushima H, Todo T, Ino Y, et al . Autocrine TGF- signaling
maintains tumorigenicity of glioma-initiating cells through Sryrelated HMG-box factors. Cell Stem Cell 2009; 5: 504514.
136. Nguyen DX, Bos PD, Massague J. Metastasis: from dissemination to organ-specific colonization. Nature Rev Cancer 2009; 9:
274284.
137. Bos PD, Nguyen DX, Massague J. Modeling metastasis in the
mouse. Curr Opin Pharmacol 2010; 10: 571577.
138. Kang Y, Siegel PM, Shu W, et al . A multigenic program mediating breast cancer metastasis to bone. Cancer Cell 2003; 3:
537549.
139. Minn AJ, Gupta GP, Siegel PM, et al . Genes that mediate breast
cancer metastasis to lung. Nature 2005; 436: 518524.
140. Fidler IJ. The pathogenesis of cancer metastasis: the seed and
soil hypothesis revisited. Nature Rev Cancer 2003; 3: 453458.
141. Buck MB, Fritz P, Dippon J, et al . Prognostic significance of
transforming growth factor receptor II in estrogen receptornegative breast cancer patients. Clin Cancer Res 2004; 10:
491498.
J Pathol 2011; 223: 205218
www.thejournalofpathology.com

TGF- in cancer

142. Biswas S, Guix M, Rinehart C, et al . Inhibition of TGF- with


neutralizing antibodies prevents radiation-induced acceleration
of metastatic cancer progression. J Clin Invest 2007; 117:
13051313.
143. Muraoka RS, Koh Y, Roebuck LR, et al . Increased malignancy
of Neu-induced mammary tumors overexpressing active transforming growth factor 1. Mol Cell Biol 2003; 23: 86918703.
144. Siegel PM, Shu W, Cardiff RD, et al . Transforming growth factor signaling impairs Neu-induced mammary tumorigenesis
while promoting pulmonary metastasis. Proc Natl Acad Sci U S
A 2003; 100: 84308435.
145. Giampieri S, Manning C, Hooper S, et al . Localized and
reversible TGF signalling switches breast cancer cells from cohesive to single cell motility. Nature Cell Biol 2009; 11: 12871296.
146. Guise TA, Yin JJ, Taylor SD, et al . Evidence for a causal role
of parathyroid hormone-related protein in the pathogenesis of
human breast cancer-mediated osteolysis. J Clin Invest 1996; 98:
15441549.
147. Yin JJ, Selander K, Chirgwin JM, et al . TGF- signaling blockade inhibits PTHrP secretion by breast cancer cells and bone
metastases development. J Clin Invest 1999; 103: 197206.
148. Kondo H, Guo J, Bringhurst FR. Cyclic adenosine monophosphate/protein kinase A mediates parathyroid hormone/parathyroid
hormone-related protein receptor regulation of osteoclastogenesis and expression of RANKL and osteoprotegerin mRNAs by
marrow stromal cells. J Bone Miner Res 2002; 17: 16671679.
149. Kingsley LA, Fournier PG, Chirgwin JM, et al . Molecular biology of bone metastasis. Mol Cancer Ther 2007; 6: 26092617.
150. Kang Y, He W, Tulley S, et al . Breast cancer bone metastasis
mediated by the Smad tumor suppressor pathway. Proc Natl Acad
Sci U S A 2005; 102: 1390913914.
151. Petersen M, Pardali E, van der Horst G, et al . Smad2 and Smad3
have opposing roles in breast cancer bone metastasis by differentially affecting tumor angiogenesis. Oncogene 2010; 29:
13511361.
152. Padua D, Zhang XH, Wang Q, et al . TGF primes breast tumors
for lung metastasis seeding through angiopoietin-like 4. Cell
2008; 133: 6677.
153. Torre-Amione G, Beauchamp RD, Koeppen H, et al . A highly
immunogenic tumor transfected with a murine transforming
growth factor type 1 cDNA escapes immune surveillance. Proc
Natl Acad Sci U S A 1990; 87: 14861490.
154. Gorelik L, Flavell RA. Immune-mediated eradication of tumors
through the blockade of transforming growth factor- signaling
in T cells. Nature Med 2001; 7: 11181122.
155. Diener KR, Woods AE, Manavis J, et al . Transforming growth
factor--mediated signaling in T lymphocytes impacts on prostatespecific immunity and early prostate tumor progression. Lab
Invest 2009; 89: 142151.
156. Thomas DA, Massague J. TGF- directly targets cytotoxic T cell
functions during tumor evasion of immune surveillance. Cancer
Cell 2005; 8: 369380.
157. Arteaga CL, Hurd SD, Winnier AR, et al . Anti-transforming
growth factor (TGF)- antibodies inhibit breast cancer cell tumorigenicity and increase mouse spleen natural killer cell activity.
Implications for a possible role of tumor cell/host TGF- interactions in human breast cancer progression. J Clin Invest 1993; 92:
25692576.
158. Ghiringhelli F, Menard C, Terme M, et al . CD4+CD25+ regulatory T cells inhibit natural killer cell functions in a transforming growth factor--dependent manner. J Exp Med 2005; 202:
10751085.
159. Lee JC, Lee KM, Kim DW, et al . Elevated TGF-1 secretion and
down-modulation of NKG2D underlies impaired NK cytotoxicity
in cancer patients. J Immunol 2004; 172: 73357340.
Copyright 2010 Pathological Society of Great Britain and Ireland.
Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk

217

160. Castriconi R, Cantoni C, Della CM, et al . Transforming growth


factor 1 inhibits expression of NKp30 and NKG2D receptors:
consequences for the NK-mediated killing of dendritic cells. Proc
Natl Acad Sci U S A 2003; 100: 41204125.
161. Kopp HG, Placke T, Salih HR. Platelet-derived transforming
growth factor- down-regulates NKG2D thereby inhibiting natural killer cell antitumor reactivity. Cancer Res 2009; 69:
77757783.
162. Friese MA, Wischhusen J, Wick W, et al . RNA interference targeting transforming growth factor- enhances NKG2D-mediated
antiglioma immune response, inhibits glioma cell migration and
invasiveness, and abrogates tumorigenicity in vivo. Cancer Res
2004; 64: 75967603.
163. Spearman M, Taylor WR, Greenberg AH, et al . Antisense oligodeoxyribonucleotide inhibition of TGF-1 gene expression and
alterations in the growth and malignant properties of mouse
fibrosarcoma cells. Gene 1994; 149: 2529.
164. Wu RS, Kobie JJ, Besselsen DG, et al . Comparative analysis of
IFN-gamma B7.1 and antisense TGF- gene transfer on the
tumorigenicity of a poorly immunogenic metastatic mammary
carcinoma. Cancer Immunol Immunother 2001; 50: 229240.
165. Hau P, Jachimczak P, Schlingensiepen R, et al . Inhibition of
TGF-2 with AP 12009 in recurrent malignant gliomas: from
preclinical to phase I/II studies. Oligonucleotides 2007; 17:
201212.
166. Schlingensiepen KH, Fischer-Blass B, Schmaus S, et al . Antisense therapeutics for tumor treatment: the TGF-2 inhibitor AP
12009 in clinical development against malignant tumors. Recent
Results Cancer Res 2008; 177: 137150.
167. Muraoka RS, Dumont N, Ritter CA, et al . Blockade of TGF-
inhibits mammary tumor cell viability, migration, and metastases.
J Clin Invest 2002; 109: 15511559.
168. Yang YA, Dukhanina O, Tang B, et al . Lifetime exposure to a
soluble TGF- antagonist protects mice against metastasis without
adverse side effects. J Clin Invest 2002; 109: 16071615.
169. Rowland-Goldsmith MA, Maruyama H, Matsuda K, et al . Soluble type II transforming growth factor- receptor attenuates
expression of metastasis-associated genes and suppresses pancreatic cancer cell metastasis. Mol Cancer Ther 2002; 1:
161167.
170. Bandyopadhyay A, Zhu Y, Malik SN, et al . Extracellular domain
of TGF type III receptor inhibits angiogenesis and tumor growth
in human cancer cells. Oncogene 2002; 21: 35413551.
171. Dasch JR, Pace DR, Waegell W, et al . Monoclonal antibodies
recognizing transforming growth factor-. Bioactivity neutralization and transforming growth factor 2 affinity purification.
J Immunol 1989; 142: 15361541.
172. Nam JS, Suchar AM, Kang MJ, et al . Bone sialoprotein mediates
the tumor cell-targeted prometastatic activity of transforming
growth factor in a mouse model of breast cancer. Cancer Res
2006; 66: 63276335.
173. Inman GJ, Nicolas FJ, Callahan JF, et al . SB-431542 is a potent
and specific inhibitor of transforming growth factor- superfamily
type I activin receptor-like kinase (ALK) receptors ALK4, ALK5,
and ALK7. Mol Pharmacol 2002; 62: 6574.
174. Hjelmeland MD, Hjelmeland AB, Sathornsumetee S, et al . SB431542, a small molecule transforming growth factor--receptor
antagonist, inhibits human glioma cell line proliferation and
motility. Mol Cancer Ther 2004; 3: 737745.
175. Matsuyama S, Iwadate M, Kondo M, et al . SB-431542 and
Gleevec inhibit transforming growth factor--induced proliferation of human osteosarcoma cells. Cancer Res 2003; 63:
77917798.
176. Uhl M, Aulwurm S, Wischhusen J, et al . SD-208, a novel transforming growth factor receptor I kinase inhibitor, inhibits
J Pathol 2011; 223: 205218
www.thejournalofpathology.com

218

177.

178.

179.

180.

181.

E Meulmeester and P ten Dijke

growth and invasiveness and enhances immunogenicity of murine


and human glioma cells in vitro and in vivo. Cancer Res 2004;
64: 79547961.
Ge R, Rajeev V, Ray P, et al . Inhibition of growth and metastasis
of mouse mammary carcinoma by selective inhibitor of transforming growth factor- type I receptor kinase in vivo. Clin Cancer
Res 2006; 12: 43154330.
Ehata S, Hanyu A, Fujime M, et al . Ki26894, a novel transforming growth factor- type I receptor kinase inhibitor, inhibits
in vitro invasion and in vivo bone metastasis of a human breast
cancer cell line. Cancer Sci 2007; 98: 127133.
Shinto O, Yashiro M, Kawajiri H, et al . Inhibitory effect of a
TGF receptor type-I inhibitor, Ki26894, on invasiveness of
scirrhous gastric cancer cells. Br J Cancer 2010; 102: 844851.
Melisi D, Ishiyama S, Sclabas GM, et al . LY2109761, a novel
transforming growth factor receptor type I and type II dual
inhibitor, as a therapeutic approach to suppressing pancreatic
cancer metastasis. Mol Cancer Ther 2008; 7: 829840.
Zhang B, Halder SK, Zhang S, et al . Targeting transforming
growth factor- signaling in liver metastasis of colon cancer.
Cancer Lett 2009; 277: 114120.

Copyright 2010 Pathological Society of Great Britain and Ireland.


Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk

182. Bueno L, de Alwis DP, Pitou C, et al . Semi-mechanistic modelling of the tumour growth inhibitory effects of LY2157299, a
new type I receptor TGF- kinase antagonist, in mice. Eur J Cancer 2008; 44: 142150.
183. Biswas S, Criswell TL, Wang SE, et al . Inhibition of transforming growth factor- signaling in human cancer: targeting a tumor
suppressor network as a therapeutic strategy. Clin Cancer Res
2006; 12: 41424146.
184. Wrzesinski SH, Wan YY, Flavell RA. Transforming growth
factor- and the immune response: implications for anticancer
therapy. Clin Cancer Res 2007; 13: 52625270.
185. Geiger T, Cox J, Ostasiewicz P, et al . Super-SILAC mix for
quantitative proteomics of human tumor tissue. Nature Methods
2010; 7: 383385.
186. Yu X, Schneiderhan-Marra N, Joos TO. Protein microarrays for
personalized medicine. Clin Chem 2010; 56: 376387.
187. Alizadeh AA, Eisen MB, Davis RE, et al . Distinct types of
diffuse large B-cell lymphoma identified by gene expression
profiling. Nature 2000; 403: 503511.

J Pathol 2011; 223: 205218


www.thejournalofpathology.com

Você também pode gostar