ArticlePDF AvailableLiterature Review

Targeted therapies in bladder cancer: An overview of in vivo research

Authors:

Abstract

Survival of patients with muscle-invasive bladder cancer is poor and new therapies are needed. Currently, none of the targeted agents that are approved for cancer therapy have been approved for the treatment of bladder cancer and the few clinical trials that have been performed had limited success, often owing to a lack of efficacy and toxic effects. However, many other novel targeted agents have been investigated in animal models of bladder cancer. EGFR, FGFR-3, VEGF, mTOR, STAT3, the androgen receptor and CD24 are molecular targets that could be efficiently inhibited, resulting in reduced tumour growth, and that have been investigated in multiple independent studies. Several other targets, for example COX-2, IL-12, Bcl-xL, livin and choline kinase α, have also been observed to inhibit tumour growth, but these findings have not been replicated to date. Limitations of several studies include the use of cell lines with mutations downstream of the target, providing resistance to the tested therapy. Furthermore, certain technologies, such as interfering RNAs, although effective in vitro, are not yet ready for clinical applications. Further preclinical research is needed to discover and evaluate other possible targets, but several validated targets are now available to be studied in clinical trials.
NATURE REVIEWS
|
UROLOGY VOLUME 12
|
DECEMBER 2015
|
681
Department of
Pathology (K.E.M.v.K.,
E.C.Z.), Department
ofUrology (T.C.M.Z.,
A.R.A., J.L.B.),
ErasmusMC,
POBox2040, 3000 CA,
Rotterdam,
Netherlands.
Correspondence to:
E.C.Z.
e.zwarthoff@
erasmusmc.nl
Targeted therapies in bladder cancer:
anoverview of invivo research
Kim E.M. van Kessel, Tahlita C.M. Zuiverloon, Arnout R. Alberts, Joost L. Boormans
andEllenC.Zwarthoff
Abstract | Survival of patients with muscle-invasive bladder cancer is poor and new therapies are needed.
Currently, none of the targeted agents that are approved for cancer therapy have been approved for the
treatment of bladder cancer and the few clinical trials that have been performed had limited success, often
owing to a lack of efficacy and toxic effects. However, many other novel targeted agents have been investigated
in animal models of bladder cancer. EGFR, FGFR-3, VEGF, mTOR, STAT3, the androgen receptor and CD24 are
molecular targets that could be efficiently inhibited, resulting in reduced tumour growth, and that have been
investigated in multiple independent studies. Several other targets, for example COX-2, IL-12, Bcl-xL, livin
and choline kinaseα, have also been observed to inhibit tumour growth, but these findings have not been
replicated to date. Limitations of several studies include the use of cell lines with mutations downstream of
the target, providing resistance to the tested therapy. Furthermore, certain technologies, such as interfering
RNAs, although effective invitro, are not yet ready for clinical applications. Further preclinical research is
needed to discover and evaluate other possible targets, but several validated targets are now available to be
studied in clinical trials.
van Kessel, K.E.M. etal. Nat. Rev. Urol. 12, 681–694 (2015); published online 22 September 2015; doi:10.1038/nrurol.2015.231
Introduction
Bladder tumours can be classified as either non-
muscle-invasive bladder cancer (NMIBC) or muscle-
invasive bladder cancer (MIBC). At diagnosis, 75%
of patients present with NMIBC (stageTa, Tis or T1)
and 25%present with MIBC (stage≥T2).1,2 In patients
with NMIBC, although the 5-year survival is >90%, the
recurrence rate is high (>50%), which necessitates costly
long-term surveillance with invasive cystoscopies.1,3–5 By
contrast, MIBC has a poor outcome: the 5-year overall
survival after radical cystectomy and lymph node dissec-
tion ranges from 49% for stage T3–4N0 disease to 74%
for stage T2N0 disease.6 In patients with MIBC, cispla-
tin-based neoadjuvant chemotherapy preceding radical
surgery results in an overall long-term survival benefit of
6%.7 However, overall, no improvement in the survival
of patients with MIBC has been accomplished over the
past 20years.8,9 The high incidence and recurrence rate,
together with the poor survival, make bladder cancer a
serious public health problem.10 Therefore, a clear clini-
cal need for new effective therapies in both NMIBC and
MIBC exists.
In the 1990s, targeted molecular therapy was intro-
duced as a novel treatment strategy in oncology.11 These
therapies aim to interfere with cellular processes that are
essential to cancer cell survival at the molecular level—
for example, by blocking proteins involved in tumour
cell proliferation or tumour cell metabolism, or by
delivering toxic compounds to tumour cells. Anumber
of targeted therapies are now well established for the
treatment of different cancers: imatinib in leukaemia
and gastro intestinal stromal tumours, cetuximab in
colorectal cancer and bevacizumab in kidney cancer.
However, none of the registered targeted therapies have
been approved for the treatment of bladder cancer and
the small number of clinical trials that have been per-
formed showed disappointing effects.12 Patients included
in these phaseI and phaseII clinical trials were mostly
not selected based on their molecular tumour profile,
which might explain the lack of treatment effect.12 In
addition, resistance to therapy occurs quickly, prob-
ably owing to molecular heterogeneity of the tumour.13
This molecular heterogeneity contributes to the clonal
evolution and selection of resistant subclones leading
to therapy resistance.14 Lastly, some trials investigating
therapy with sunitinib or gefitinib were stopped early
because of s ubstantial toxic effects of the treatment.12,15,16
This Review provides a comprehensive overview of
targeted therapies for bladder cancer that have been
investi gated in animal models, some of which have
potential for clinical application. We provide insights
into both the challenges and the promises of the pre-
clinical develop ment of the targeted agents, highlight-
ing the most promis ing studies (Table1), grouped by
the mechanisms and pathways that the investigated
targets are involved in. Summaries of all studies included
in this Review are p rovided in the Supplementary
informationonline.
Competing interests
The authors declare no competing interests.
REVIEWS
© 2015 Macmillan Publishers Limited. All rights reserved
682
|
DECEMBER 2015
|
VOLUME 12 www.nature.com/nrurol
Overview of preclinical animal models
In the development of novel targeted therapies, invitro
and invivo studies precede clinical trials. Only few of the
potential novel therapies will eventually be evaluated in
clinical studies. In bladder cancer research, three types
of animal models are typically used in invivo research:
syngeneic mouse or rat models, and heterotopic or
o rthotopic xenogeneic mouse or rat models (Box1).
Syngeneic tumour models have the advantage of an
intact tumour host environment including an active
immune system. However, obvious drawbacks of these
models exist, as mouse tumours are intrinsically differ-
ent from human tumours. In addition, only few tumour
types are available.17 Xenogeneic tumour models more
closely resemble the tumours of bladder cancer patients;
however, the stroma and vasculature is murine and the
animals are immunodeficient to allow growth of human
cancer cells.17 Furthermore, tumours grownfrominjected
cell lines no longer resemble the tumour from which the
cell line originated, owing to years of cultivation, passag-
ing and expansion. Xenografts that are derived from fresh
patient material represent a possibility to achieve better
resemblance of the true hetero geneous tumour architec-
ture. In 2015, the first study describing a method to create
patient-derived xenografts for bladder cancer studies was
published.18
In studies published since 2009, the heterotopic and
orthotopic animal models were generated using many dif-
ferent bladder cancer cell lines and several of these have
been genotyped, revealing specific oncogenic mutations
(Table2). These mutations are important to consider
when choosing which tumour model to use in invivo
studies, particularly when the target of the tested agent
is upstream of these mutations in the signalling pathway,
as downstream mutations might activate the signal-
ling cascade despite inhibition of the upstream target.
However, for correct interpretation of the treatment
effect, awareness of all potential mutations—downstream
or upstream of the target—that could influence the study
result is required.
Furthermore, in 2013, a study showed that KU-7 bladder
cancer cells are actually HeLa cervical carcinoma cells, as
the original KU-7 cell line had been contaminated with
HeLa cells at the source site at some point before 1984.19
Consequently, KU-7 cells should be considered cervical
cancer cells and not bladder cancer cells. Invivo studies
using KU-7 animal models are listed in Supplementary
Table1 online, but owing to their n onurological origin
they are not being discussed in thisReview.
Receptor tyrosine kinases
Since 2009, many invivo studies have investigated agents
targeted at receptor tyrosine kinases (RTKs). An overview
of these studies is given in Supplementary Table2 online.
RTKs, such as the epidermal growth factor receptor
(EGFR), the fibroblast growth factor receptors (FGFRs)
and the vascular endothelial growth factor receptors
(VEGFRs), are essential for the communication between
cells and their environment. Specific growth factors (for
example, cytokines or hormones) bind to the extracellular
domain of an RTK, thereby activating the intracellular
kinase domain of the receptor. The activation results in
the stimulation of downstream signalling cascades that
affect cell proliferation and cell growth.20 The RTK–
Ras–MAPK pathway and the RTK–PI3K–Akt pathway
are among the most frequently affected signalling path-
ways in cancer (Figure1, Figure2).21,22 In bladder cancer,
RTKs are often overexpressed or mutated, resulting in
o verstimulation of downstream signalling pathways.
EGFR
EGFR is overexpressed in many bladder tumours
andEGFR overexpression correlates with a poor progno-
sis.23,24 Examples of agents targeting EGFR are gefitinib,
a small-molecule inhibitor used for the treatment of
chemotherapy- refractory non-small-cell lung cancer, and
cetuximab, an antibody used for the treatment of head
andneck cancer and advanced colorectal cancer.25,26 The
effect of inhibiting EGFR in bladder cancer was investi-
gated in five mouse studies (Figure2).27–31 Cetuximab was
used in three of these studies and researchers found that
growth of MGH and T24 xenografts was inhibited by this
antibody.27,30,31 In two studies, cetuximab was combined
with photodynamic therapy, which enhanced the inhibi-
tory effect in one study and was synergistic in the other.27,30
Similarly, combining afatinib (an inhibitor of EGFR and
erbB-2) with cetuximab was more effective than cetux-
imab alone.31 Addition of bevacizumab (an antibody
against VEGF-A) to cetuximab and p hotodynamic
therapy resulted in regression of tumourvessels.30
Caution must be taken in interpreting the results of four
of these studies, which used T24 cells28,31 or MGH cells,27,30
as these cell lines carry activating mutations downstream
of the treatment target. Molecular profiling has shown
that T24 cells have a mutation in HRAS, which activates
the Ras–MAPK and PI3K–Akt pathways downstream of
EGFR.32,33 One study in colorectal tumour samples dem-
onstrated that activating mutations in another Ras protein
(GTPase KRas, encoded by KRAS) render the tumour cells
insensitive to cetuximab.34 In one of the two studies using
T24 cells, the authors describe a cetuximab resistance
mechanism but the influence of the HRAS mutation in
this resistance mechanism remains unknown.31 In the
other study using T24 cells, LRIG1 was introduced into
these cells that were then used in a xenograft model.28
Key points
Over the past 20years, survival of patients with muscle-invasive bladder cancer
has not improved
None of the targeted therapies that are approved for other cancers have been
approved for the treatment of bladder cancer
The small number of clinical trials that have been performed in patients with
bladder cancer had limited success owing to several limitations
Specific oncogenic mutations in different cell lines used in invivo
research might render these cell lines insensitive to the therapy that is
beinginvestigated
Inhibition of EGFR, FGFR-3, VEGF, mTOR, STAT3, the androgen receptor and
CD24 resulted in inhibition of tumour growth in multiple invivo studies
Careful patient selection in clinical trials based on the molecular profile of the
tumour will be essential in demonstrating benefit of new targeted therapies
REVIEWS
© 2015 Macmillan Publishers Limited. All rights reserved
NATURE REVIEWS
|
UROLOGY VOLUME 12
|
DECEMBER 2015
|
683
The protein leucine-rich repeats and immunoglobulin-
like domains protein1, which is expressed from LRIG1,
inhibits EGFR through direct interaction;28 however, its
effect on the Ras–MAPK and PI3K–Akt pathways is most
likely nulli fied by the presence of the HRAS mutation in
T24 cells. In addition, the cell line MGH, which was used
in the two studies by Bhuvaneswari and colleagues,27,30 has
a mutation in AKT1; the corresponding protein PKB is
an effector for the PI3K–Akt pathway.32,33 MGH cells also
carry an FGFR3 mutation that might activate the Ras–
MAPK pathway, independently of EGFR.32,33 Although a
beneficial effect of targeting EGFR was observed, the effect
could have been more convincing if a different cell line
had been used to establish the treated mouse xenografts.
Furthermore, the observation that EGFR was down-
regulated does not imply that the downstream signalling
pathways were also downregulated.27,30 By contrast, in the
study of Rebouissou etal.29 in which EGFR was inhibited
by erlotinib, resulting in tumour growth inhibition or
delayed tumour detection by ultrasono graphy, the results
are convincing. The cell lines used (BFTC-905, JMSU1,
KK47, L1207, UMUC6, VMCUB1) were selected using
Table 1 | Promising invivo research*
Target Pathway Tumour type (n)Intervention Study
EGFR Ras–MAPK Orthotopic (40)
Heterotopic (144;
2per mouse)
EGFR inhibitor (erlotinib) 29
FGFR-3 Ras–MAPK Heterotopic (NR) FGFR inhibitor (PD173074) 48
FGFR-3 Ras–MAPK Heterotopic (80) mAb against FGFR-3 (R3Mab) 49
FGFR-3 Ras–MAPK Orthotopic (128) mAb against FGFR-3 (R3Mab) 50
FGFR-3 Ras–MAPK Heterotopic (72) FGFR inhibitor (PD173074) 51
FGFR-3
EGFR
Ras–MAPK Experiment1:
Heterotopic (30)
Experiment2: NR
Experiment1: FGFR inhibitor (PD173074) and/or EGFR inhibitor (getinib)
Experiment2: FGFR inhibitor (PD173074) and/or mAb against EGFR
(cetuximab)
54
VEGFRs Ras–MAPK Metastatic (NR) Tyrosine kinase receptor inhibitor (sunitinib) and/or chemotherapy
(epirubicin)
149
VEGF-A
VEGF-C
VEGFR-3
Ras–MAPK Heterotopic (22) siRNA against VEGF-A or siRNAs against VEGF-C and VEGFR-3 59
VEGF-A
EphB4
Ras–MAPK Heterotopic (32) EphB4 inhibitor (sEphB4-HSA) and/or mAb against VEGF (bevacizumab) 150
mTOR PI3K–Akt Heterotopic (37) Knockdown of p53 and PTEN and mTOR inhibitor (rapamycin) 74
mTOR PI3K–Akt Heterotopic (40) mTOR inhibitor (rapamycin) and/or chemotherapy (cisplatin) 75
mTOR PI3K–Akt Heterotopic (40) mTOR inhibitor (rapamycin) and/or PI3K inhibitor (wortmannin) 76
STAT 3
mTOR
MAPK
JAK
PI3K–Akt
Ras-MAPK
Orthotopic (84) mTOR inhibitor (rapamycin) and/or MAPK inhibitor (UO126) and/or STAT3
inhibitor (S3I-201)
77
STAT 3
Survivin
JAK
Apoptosis
Heterotopic (35) siRNAs against STAT3, survivin or both STAT3 and survivin 151
AR JAK
PI3K–Akt
Ras–MAPK
Heterotopic (18) siRNA against AR 93
AR JAK
PI3K–Akt
Ras–MAPK
Heterotopic (10) Androgen deprivation (surgical castration) 92
AR JAK
PI3K–MAPK
Ras–Akt
Heterotopic (18) siRNA against AR 94
CD24
AR
Metastasis
JAK
PI3K–Akt
Ras–MAPK
Orthotopic (190)
Heterotopic (20)
CD24-knockout mice or androgen deprivation (surgical castration) 98
CD24 Metastasis Heterotopic (16)
Metastatic (44)
mAb against CD24 (ALB9) 95
CD24
HIF-1α
Metastasis
Angiogenesis
Heterotopic (40)
Metastatic (32)
shRNA against CD24 or HIF-1α128
*Additional details can be found in the Supplementary Information online. Abbreviations: AR, androgen receptor; CD24, signal transducer CD24;
EGFR,epidermal growth factor receptor; EphB4, ephrin type-B receptor4; FGFR, fibroblast growth factor receptor; HIF-1α, hypoxia-inducible factor1α;
mAb,monoclonalantibody; MAPK, mitogen-activated protein kinase; mTOR, mammalian target of rapamycin; shRNA, small hairpin RNA; siRNA, small interfering
RNA; STAT3, signal transducer and activator of transcription3; VEGFR, vascular endothelial growth factor receptor.
REVIEWS
© 2015 Macmillan Publishers Limited. All rights reserved
684
|
DECEMBER 2015
|
VOLUME 12 www.nature.com/nrurol
a gene expression classifier, discriminating between high
EGFR expression for a basal-like phenotype or low EGFR
expression for a non-basal-like phenotype, and thoroughly
examined for downstream mutations (Table1).29
Several EGFR-targeted therapies have been evaluated in
clinical trials and a number of clinical trials are ongoing.12
For instance, erlotinib monotherapy was studied in the
neo adjuvant setting in 20patients with MIBC.35 The
authors reported a pathological response rate with down-
staging to <T2 in 35% of patients. Gefitinib and cetux-
imab combined with chemotherapy were studied in the
metastatic setting.36,37 Median overall survival times were
15.1months and 14.3months for gefitinib and cetuximab,
respectively. Gefitinib was also studied as a single-agent,
second-line therapy in the meta static setting: only one of
31 treated patients responded to treatment.38 Nevertheless,
we believe that treatments targeting the EGFR can still
have a role as therapies for patients with bladder cancer in
the future. The disappointing trial results and data from
invivo research clearly highlight the need to carefully
select patients for therapy based onthemolecular profile
of their tumour.
In addition, agents targeting erbB-2 are being investi-
gated in phaseII trials in patients with bladder cancer
but most results are still pending. Examples include a
number of trials in which trastuzumab or lapatinib are
being combined with chemotherapy or given as single
agents.39–44 Promising results are expected, particularly
in the light of new data reported by The Cancer Genome
Atlas (TCGA) Research Network21 and Groenendijk and
co-workers.45 The TCGA study found that ERBB2 was
mutated or amplified in 9% of samples, demonstrating
the importance of ERBB2 mutations in the pathogenesis
of bladder cancer.21 Groenendijk etal.45 found that ERBB2
mutations were associated with a very good response to
neoadjuvant chemotherapy: 24% of complete responders
but none of the nonresponders had ERBB2mutations.
FGFRs
In addition to EGFR overexpression, approximately two-
thirds of all NMIBCs have activating FGFR3 mutations.46
In MIBC, <15% of tumours have FGFR3 m utations,
although >40% of MIBCs overexpress FGFR-3.47 Seven
invivo studies reported the effects of targeting FGFR-3.48–54
In most cases, the cell lines used either carried a mutant
FGFR3 gene or a FGFR3 fusion gene, both leading to acti-
vation of downstream pathways. The FGFR-3 was inhib-
ited by the small-molecule inhibitor PD173074 in four
studies,48,51,52,54 and by the mono clonal antibody R3Mab
in two studies (Figure2).49,50 One study evaluated the effect
of the pan-FGFR inhibitor BGJ-398.53
Five studies reported growth inhibition, reduced cell
proliferation or both.48–51,54 Induction of apoptosis—
which demonstrates cytotoxic effects of the treatment
—was only reported in one study.48 The MGH cell line
used in the study by Lamont etal.52 has an AKT1 muta-
tion, leading to activation of the PI3K pathway regardless
of inhibition of the FGFR; nevertheless, the investiga-
tors did observe a cytostatic response to treatment with
PD173074. The cell line UMUC3 used in the study by
Cheng etal.53 has a KRAS mutation and a deletion in the
PTEN gene, resulting in constitutive activation of both
Ras–MAPK and PI3K–Akt pathways. In this study, block-
age of tumour cell extravasation was reported but not
growth inhibition. In another study, inhibition of FGFR-3
and EGFR was combined in a mouse model using RT112
cells, which have no known downstream mutations in
Ras–MAPK and PI3K–Akt pathways.54 Combination
therapy with the FGFR-3 inhibitor PD173074 and cetux-
imab gave the best result regarding tumour growth
inhibition compared with treatment with either agent
alone, and also led to sustained tumour growth control.54
Overall, these mouse studies provided compelling evi-
dence that targeting FGFR-3 in human bladder tumours
is potentially effective (Table1).
In clinical trials, however, the strategy of targeting
FGFR-3 has not yielded the expected positive results.
The small-molecule inhibitor dovitinib was investigated
as a single-agent, second-line treatment in patients with
metastatic urothelial carcinoma.55 The results were
disappointing and the study was terminated owing
to a lack of beneficial treatment effects. In a phaseIb
study of dovitinib plus gemcitabine combined with cis-
platin or carbop latin in patients with advanced solid
tumours, which included two bladder cancer patients,
no treatment effect was observed.56 Furthermore, com-
bined treatment was poorly tolerated owing to myelo-
suppression. A phaseII trial of dovitinib in patients
with NMIBC who did not respond to BCG tr eatment is
ongoing and reports of the results are pending.57
In summary, studies of the inhibition of RTKs in cell-
line-derived animal models should be performed after
molecular profiling of the chosen cell line to exclude
the possibility of activating mutations downstream
ofthetarget RTK. In human bladder cancer tumours,
mutations in the Ras family of genes are rare (for example,
HRAS mutations exist in 5% of patients)21 and mutations
in the PI3K–Akt pathway occur in approximately 20%
of bladder cancer tumours.21 For these subgroups of
patients, tumour models harbouring mutations in these
pathways will be useful in therapy evaluation. However,
Box 1 | Main types of animal models used in preclinical bladder cancer research
Invivo tumour models can be classified according to the genetic background
oftumour and host and according to tumour location.
Genetic background
Syngeneic model: the tumour arises from tissue of the host animal and has
the same genetic background as the animal; examples include spontaneously
developing cancers, BBN-induced tumours and mice that are genetically
modified to develop tumours
Xenogeneic model: tissue or cells from another species are transplanted into
the host animal
Tumour location
Orthotopic model: the tumour develops in the same anatomical location as the
source cell line or tissue, for example when bladder cancer cells are injected
into the bladder wall
Heterotopic model: the tumour develops in a different anatomical location
than the source cell line or tissue, for example when bladder cancer cells
aresubcutaneously injected into the flank, or results from metastasis
Abbreviation: BBN, N-butyl-N-(4-hydroxybutyl) nitrosamine.
REVIEWS
© 2015 Macmillan Publishers Limited. All rights reserved
NATURE REVIEWS
|
UROLOGY VOLUME 12
|
DECEMBER 2015
|
685
RTK inhibition therapy will probably lack effectivity and
therapies not involving these pathways could be b eneficial
for this patient category.
The data reviewed here suggest that inhibition of RTKs
after stratification of patients based on molecular analy-
sis of the tumour does have potential in bladder cancer
therapy for the majority of patients. In addition, combi-
nation of treatments might enhance treatment effects;
however, some treatment combinations might also result
in increased toxic effects.
VEGFRs and angiogenesis
As tumours become larger, hypoxia will occur unless the
tumour is able to grow new blood vessels. Most tumours
stimulate neoangiogenesis by secreting VEGF. VEGF
binds to its cognate receptor on endothelial cells and
stimulates the formation of new blood vessels.58
In one study, Feng and co-workers59 treated mice
that had heterotopic mouse bladder carcinomas with
intratumoural injections of synthetic small interfering
RNAs (siRNAs) against Veg f a, Ve g fc and Veg f r3 . The
siRNAs specifically inhibited synthesis of the correspond-
ing proteins (VEGF-A, VEGF-C and VEGFR-3, respec-
tively). VEGF-A and VEGF-C are two types of ligands
that can bind to VEGFRs. Interestingly, all mice in the
control group and in the group that received siRNAs
against Veg f c and Ve g f r3 developed lung metastases but
37.5% of mice in the group treated with siRNA against
Ve g fa had no lung metastases. In addition, no animals
in the group treated with siRNA against Veg f a had liver
metastases compared with 50% and 25% of mice in the
control group and the group receiving siRNAs against
Ve g fc and Veg f r3 , respectively.59 These data suggest
that inhibition of VEGF-A expression reduces the
d evelopment of metastases.
Three of the studies on angiogenesis inhibition used
technologies that cannot yet be implemented in clinical
practice (for example, viral vehicles).59–61 Two of these
Table 2 | Bladder cancer cell lines and known oncogenic mutations and gene fusions relevant to this Review
Cell line
[alternative name]
Origin Oncogenic mutations
andgene fusions*
Invivo studies
253J-BV [253JBV, 253J_B-V] Human Unknown 156
5637 Human ERBB2, TP53, RB1 110,150,157
AY-27 Rat Unknown 158–160
BFTC-905 [BFTC905] Human NRAS, TP53 29
BIU-87 Human Unknown 161,162
BTT739 Human Unknown 59,61,90
EJ [MGH-U1] Human Unknown 91,98,130
HT-1197 [HT1197] Human NRAS, PIK3CA 163
HT-1376 [HT1376] Human TP53, RB1 163
J82 Human FGFR3, ERBB2, TP53 51,164,165
JMSU1 Human Unknown 29
KK47 Human Unknown 29
KU-19-19 [KU1919] Human NRAS, AKT1 88
KU-7 Human NA§166–175
L1207 Human Unknown 29
MB49 Mouse Unknown 175–178
MBT-2 [MBT2] Mouse Unknown 60,120,123,124,129,149,179
MGH [MGH-U3, MGHU3, RN] Human FGFR3, AKT1, CDKN2A 27,30,48,52
RT112 [RT-112] Human TP53, FGFR3TACC3 fusion 49,50,52,54,180,181
RT4 [RT-4] Human RHOA 51,74,152,182
SW780 [SW-780] Human FGFR3BIAIAP2L1 fusion 51,52,119,183
T24 [T-24] Human HRAS, TP53 28,31,93,94,99,112,113,127,151,163,178,184–191
TSU-pr1 Human NA|| 143
UMUC1 [UM-UC-1] Human Unknown 50
UMUC3 [UM-UC-3] Human PTEN, KRAS, TP53 53,75,76,87,89,92,95,98,108,111,125,128,152,192–194
UMUC6 [UM-UC-6] Human FGFR3 29
UMUC14 [UM-UC-14] Human FGFR3 48–50,81,119
VMCUB1 [VM-CUB-1] Human TP53, ERBB2, PIK3CA, CDKN2A 29
YTS-1 Human Unknown 144
*Only mutations registered in COSMIC databases are listed.32 Only studies in which the cell lines have been used to create xenografts are listed. §Cervical
carcinoma cell line. ||TSU-pr1 cells have been derived from T24 cells. Abbreviation: NA, not applicable.
REVIEWS
© 2015 Macmillan Publishers Limited. All rights reserved
686
|
DECEMBER 2015
|
VOLUME 12 www.nature.com/nrurol
studies reported tumour growth inhibition,59,61 but one
study did not find an effect on tumour growth.60 All
three studies reported a combination of either reduced
a ngiogenesis, reduction in lymphatic vessels or reduction
of metastasis.59–61
Many clinical trials of agents targeting VEGFs and
VEGFRs in patients with bladder cancer have already
been reported.12 Four studies investigated the use of
bevacizumab (targeting VEGF-A) or sunitinib (targeting
VEGFRs) combined with chemotherapy in the neoadju-
vant setting (Figure2).62–65 Pathological response rates
with downstaging to <T2 ranged from 22% to 53%. One
study of sunitinib was terminated early owing to substan-
tial toxic effects.65 Combining bevacizumab with chemo-
therapy has also been studied as adjuvant treatment in
two studies.64,66 The reported median progression-free
survival was 6.5months in one study64 and 8.2months in
the other study.66 In both studies, the aspired study goal
of 50% improved progression-free survival was not met.
Other studies have analysed monotherapy of sunitinib,
aflibercept (a fusion protein that binds VEGFs) or pazo-
panib (which targets VEGFRs) as second-line treatment
strategies in phaseII trials in patients with advanced
urothelial cancer.67–70 Two studies reported no beneficial
effect of aflibercept and pazopanib,68,69 and the other two
studies reported a partial response in 7% and 17.1% of
patients for sunitinib and pazopanib, respectively.67,70 At
present, three clinical trials investigating anti-VEGF ther-
apies (bevacizumab, pazopanib and sunitinib) in patients
with bladder cancer areongoing.71–73
Overall, VEGF-targeted therapies show promise for
clinical application. However, sunitinib treatment resulted
in high toxicity in clinical trials,12,16 and a more selective
VEGF inhibitor or antibody will be more likely to be well
tolerated. Again, selection of patients based on the down-
stream mutations in their tumour will be n ecessary to
achieve the highest benefits.
Signal transduction
Several studies have identified potential new targets for
bladder cancer therapy in different signal transduction
pathways. Signals coming from cell membrane receptors
must be communicated to downstream effector mol-
ecules. This signal transduction occurs through kinases,
such as MAP kinases and PI3 kinases, or GTPase activity,
for example of Ras proteins (Figure1, Figure2).
The signal transduction cascades offer many possi-
ble new therapeutic targets. As a consequence, inhibi-
tors have not been generated for all potential targets. In
cell culture and animal models, inhibition of a target by
inhibiting translation of its mRNA by siRNAs usually
gives the first insight into whether the target is suit-
able for further study. An overview of invivo studies in
which targeting of signal transduction pathways has been
i nvestigated is provided in Supplementary Table3 online.
The protein mTOR acts in the PI3K–Akt pathway and
is targeted by rapamycin. Several rapamycin analogues,
such as temsirolimus and everolimus, have been devel-
oped and are in use for the treatment of advanced renal
cell carcinoma. In three studies that used hetero topic
xenograft models and one study that used an orthotopic
mouse model, mTOR was targeted via intraperitoneal
injection of rapamycin (Figure2).74–77 All studies showed
that rapamycin reduced tumour growth. In addition,
Makhlin etal.75 showed that the addition of cisplatin
to rapamycin treatment resulted in increased median
survival and a lower proliferation index (decreased
expression of the proliferation marker Ki-67) in com-
parison with rapamycin alone. Similarly, Zhou etal.77
demonstrated that combining rapamycin with inhibi-
tors of MAPK and signal transducer and activator of
transcription3 (STAT3) significantly prolonged sur-
vival in a bladder cancer model in transgenic mice in
comparison with rapamycin alone. Seront etal.76 found
that rapa mycin was less effective in UMUC3 cells in
comparison with UMUC14 cells (UMUC3 cells have a
homozygous deletion in PTEN, whereas UMUC14 cells
carry wild-type PTEN). The researchers also found that
combining rapamycin with a PI3K inhibitor was required
Nature Reviews | Urology
Receptor tyrosine kinases
Signal transduction
Protein synthesis and cell metabolism
Cell cycle proteins and apoptosis
Metastasis, immune system and others
DNA
Transcription
RNA
Protein
synthesis
G2 M
SG1
Oncogene
p53
Apoptosis
Cell
metabolism
Stress
Virus
T cell
CD24
EGFR
FGFRs
VEGFRs
Hormones and
growth factors
Cell
membrane
R
a
s
-
M
A
P
K
p
a
t
h
w
a
y
P
I
3
K
-
A
k
t
p
a
t
h
w
a
y
O
t
h
e
r
p
a
t
h
w
a
y
s
Figure 1 | Cellular processes involved in cancer development. External stimuli,
such as hormones and growth factors, stimulate receptors on the cellular
membrane, leading to activation of downstream pathways. These pathways lead
toactivation of transcription factors and, thus, gene expression. RNA is either
translated to functional proteins or can have regulatory functions on its own.
Incancer cells, many processes eventually lead to increased cellular proliferation
by amplifying cell growth, evading cell cycle arrest and evading apoptosis.
Abbreviations: EGFR, epidermal growth factor receptor; FGFR, fibroblast growth
factor receptor; p53, cellular tumour antigen p53; VEGFR, vascular endothelial
growth factor receptor.
REVIEWS
© 2015 Macmillan Publishers Limited. All rights reserved
NATURE REVIEWS
|
UROLOGY VOLUME 12
|
DECEMBER 2015
|
687
to inhibit UMUC3 cell tumour growth.76 This finding,
however, was not confirmed by Puzio-Kuter etal.74 in
a mouse model of conditional knockout of PTEN and
Makhlin etal.75 in a xenograft model using UMUC3
cells. In summary, invivo studies investigating rapa-
mycin treatment of bladder cancer have demonstrated
that this inhibitor is promising, displaying effective ness
in human xenografted cell lines and models of mouse
bladder cancer (Table1).
Examples of phaseII clinical trials targeting mTOR
and the PI3K–Akt pathway in patients with bladder
cancer include several trials that combined temsiroli-
mus or everolimus with chemotherapy.78–80 The results
of these studies are not yet available. Other trials have
evaluated a single-agent strategy of temsirolimus or
everolimus (Figure2).81–83 One study was terminated
because of futility.82 Although most of the patients in
these trials did not respond to the single-agent treatment,
some patients had stable disease or partial response. For
example, upon treatment with everolimus, eight and
two out of 37 patients in one study81 and 12 and two
out of 45patients in another study83 had stable disease
or partial response, respectively, i ndicating a beneficial
effect in selected groups of patients.
A single-agent strategy with the PI3K inhibitor
BKM120 (also known as buparlisib) is being investigated
in an ongoing trial but study results are still pending
(Figure2).84 BKM120 has previously shown partial
responses in the treatment of patients with breast cancer.85
Furthermore, one trial evaluating the effect of everoli-
mus in patients with NMIBC is currently underway;86
however, the tumours of the patients enrolled in this trial
have not been molecularly characterized. Hence, the
eventual treatment effect might differ between patients
depending on the molecular profile of their tumour,
which could result in an underestimation of the overall
treatment effect. If enough patients are included, strati-
fied analysis based on the molecular tumour profile might
indicate which patients benefit more (or less) than others.
Other targets in signal transduction cascades include
transcription factors (for example, STAT3, NF-κB, zinc
finger protein224 and the androgen receptor [AR])
NF-κB
Nature Reviews | Urology
Growth factors
Cell survival Cell growthApoptosis
Cell cycle Angiogenesis
Cell cycle
p21 (14%) BAX
p53 (49%)
MDM2 (9%)
ARF
CP-31398,
ALT-801 Balomycin A1
Alisertib
Ramucirab,
pazopanib,
bevacizumab,
sunitinib,
aibercept,
siRNA
mTOR
Rapamycin,
temsirolimus,
everolimus
Akt (12%)
VEGF
STAT
JAK-1
E2F
MAPK
MEK
B-raf
Ras (2–5%)*
RTKs (6–15%)* PI3K (17%)
BKM120
Getinib,
cetuximab, afatinib,
erlotinib, trastuzumab,
lapatinib, PD173074,
R3Mab, dovitinib
RPS6
Apoptosis pathway
Ras–MAPK pathway
Activation
Inhibition
Targeting
JAK–STAT pathway
PI3K–Akt pathway
Figure 2 | Cellular pathways affected in bladder cancer and targeting therapies. Multiple small-molecule inhibitors and
antibodies against different receptors and pathways have been tested in preclinical and clinical studies. Agents targeting
protein translation and cell metabolism affect the molecules at the end of signalling pathways and are not shown. Pathways
and targets are located at various sites in and around the tumour cell (Figure1). The percentage of bladder cancer samples
harbouring mutations are shown for targets that have been screened by the TCGA.21 *Percentage depends on the RTK
analysed. Percentage harbouring a copy number variation. Abbreviations: ARF, tumour suppressor ARF; Akt, serine/
threonine-protein kinases; B-raf, serine/threonine-protein kinase B-raf; BAX, apoptosis regulator BAX; E2F, transcription
factors of the E2F family; MAPK, mitogen-activated protein kinases; MDM2, E3 ubiquitin-protein ligase Mdm2; MEK, dual
specificity MAPK kinases1 and2; NF-κB, nuclear factor κB; p21, cyclin-dependent kinase inhibitor1; p53, cellular tumour
antigen p53; RPS6, 40S ribosomal protein S6; RTK, receptor tyrosine kinase; STAT, signal transducer and activator of
transcription1 and3; TCGA, The Cancer Genome Atlas Research Network; VEGF, vascular endothelial growth factor.
REVIEWS
© 2015 Macmillan Publishers Limited. All rights reserved
688
|
DECEMBER 2015
|
VOLUME 12 www.nature.com/nrurol
or proteins that regulate transcription (for example,
histone deacetylases).77,87–96 Inhibition of these transcrip-
tion factors resulted in tumour growth inhibition in all
studies included in this Review.77,87–93,95,96
Testosterone and the AR might have a role in the
increased incidence of bladder cancer in men compared
with women.97 Hence, surgical castration was explored
as therapy in two animal studies: Overdevest etal.98
observed a reduction in tumour growth of orthotopic
and heterotopic xenografts in mice using this approach,
whereas Shiota etal.92 did not find a significant inhibi-
tion of tumour growth in their heterotopic xenograft
model. Inaddition, two studies that investigated siRNAs
targeting the AR in subcutaneous xenografts observed
statistically significant tumour growth inhibition.93,94
Considering these findings, inhibition of the AR might
be a possible therapeutic strategy for men with bladder
cancer (Table1); however, none of the possible tran-
scription factor targets have been evaluated in clinical
studiesyet.
Cell cycle proteins and apoptosis
Cell division occurs after the cell has gone through the
four cell cycle phases G1, S, G2 and M (Figure1). Cell
cycle checkpoint proteins, which regulate progression
through the cell cycle, are often mutated or bypassed in
malignant cells and are therefore considered potential
targets for therapy. Studies that have investigated target-
ing cell cycle control and apoptosis pathways invivo are
listed in Supplementary Table4 online.
In the Mphase, the accurate formation of the mitotic
spindle is controlled during a quality check. Aurora
kinaseA is a protein that is involved in mitotic spindle
formation and function that was evaluated as a poten-
tial therapeutic target for bladder cancer by Zhou etal.99
(Figure2). The researchers found that aurora kinaseA
expression was increased in clinical bladder cancer
tumour samples and that inhibition of aurora kinaseA
with the inhibitor MLN8237 (also known as alisertib) in
T24 bladder cancer xenografts in mice resulted in sup-
pressed tumour growth. A phaseII clinical trial evalu-
ating alisertib combined with paclitaxel in advanced
bladder cancer was registered in 2014.100
The apoptosis pathway can become activated during
the cell cycle, for example when DNA damage occurs.
In addition, apoptosis can be induced by activation
oftheRas–MAPK pathway and subsequent activation of
the tumour suppressor p53, which is encoded by TP53, a
gene that is often mutated in cancer cells. Mutations that
change an amino acid in the DNA-binding domain of p53
inhibit its function as a transcription factor and, as a con-
sequence, expression of the proapoptotic apoptosis regu-
lator BAX. Deletions of the gene likewise lead to loss of
function.101,102 In one study, CP-31398, a compound that
stabilizes the conformation of the DNA-binding domain
of p53 (Figure2), was administered to SV40T-transgenic
mice.103 The SV40 large Tprotein binds and inhibits
both p53 and the retinoblastoma-associated protein
and thereby induces bladder cancer.104 Administration
of CP-31398 resulted in decreased tumour weight and
inhibition of invasive tumour growth with an increase
in apoptosis.73 In clinical trials, p53-targeted therapy is
being evaluated using ALT-801, a fusion protein of IL-2
and a T-cell receptor domain.105 The efficacy of ALT-801
is being investi gated as a combination treatment with
gemcitabine, with or without cisplatin, in patients
with MIBC and NMIBC.106,107 Patient recruitment is
currentlyongoing.
Bcl-xL (Bcl-2-like protein1 expressed from BCL2L1)
is an antiapoptotic protein of the Bcl2 family. One study
investigated the clinicopathological significance of
Bcl-xL expression in patients with upper urinary tract
urothelial carcinoma (Figure2).108 Using a subcutane-
ous UMUC3 xenograft model in mice, the researchers
evaluated the therapeutic effect of the V-ATPase inhibi-
tor bafilomycinA1 against Bcl-xL and observed reduced
tumour growth with signs of apoptosis (confirmed by
terminal deoxynucleotidyl transferase dUTP nick end
labelling [TUNEL] assay, which is used to detect DNA
fragmentation caused by apoptosis). Another target in
this category is livin, which belongs to the inhibitors of
apoptosis proteins (IAP) family and was shown to predict
disease relapse in bladder cancer.109 Inhibiting livin in
an invivo heterotopic xenograft model led to tumour
inhibition and increased apoptosis.110
Targeting and inducing the apoptosis pathway is an
especially interesting strategy in cancer treatment, as the
occurrence of apoptosis shifts the treatment effect from
cytostatic to cytotoxic. This shift could, in theory, lead to
tumour regression rather than only reduced cell prolif-
eration. However, most of the targeted therapies that are
currently used clinically are cytostatic rather than cyto-
toxic. With the exception of the study investigating the
p53 stabilizer CP-31398,103 the number of animals used
in experiments in this category of targets is small and
none of the studies have been replicated to date.
Protein synthesis and cell metabolism
Protein synthesis
During the G1 phase of the cell cycle, many proteins
that are required for subsequent cell division are being
synthesized (translation). As the step of protein trans-
lation is essential to cell proliferation, the participat-
ing proteins are potential targets for anticancer agents.
Supplementary Table4 online provides an overview of
the relevant invivo studies.
Eukaryotic translation initiation factor3B (eIF3b) is a
subunit of the translation initiation factor complex eIF3.
Immunohistochemical assessment of 143bladder cancer
samples showed that eIF3b overexpression was corre-
lated with a high tumour stage and short disease-specific
survival in humans.111 The researchers then used siRNAs
to inhibit eIF3b formation in UMUC3 cells, which
resulted in reduced tumour growth of s ubcutaneous
xenograftsinmice.111
Some bacterial toxins target the function of proteins
during translation. One group used diphtheria toxin
subunitA to inhibit translation of elongation factor2.112,113
Expression of the toxin was controlled by the promoter of
the insulin-like growth factor2 gene (IGF2) and/or the
REVIEWS
© 2015 Macmillan Publishers Limited. All rights reserved
NATURE REVIEWS
|
UROLOGY VOLUME 12
|
DECEMBER 2015
|
689
promoter of the long noncoding RNA gene H19. The
products of IGF2 and H19 are highly expressed in many
tumours but not in normal cells. In two studies using
orthotopic T24 human bladder cancer xenograft models
in mice, treatment inhibited tumour growth, result-
ing in tumours that were 86% and 83% smaller than
incontrols.112,113
Targeting the translational process has also been
investi gated in other types of cancer, for example by tar-
geting eukaryotic translation initiation factor 4E (eIF4e).
After multiple promising preclinical studies, a phaseI
trial114 using an antisense oligonucleotide to block expres-
sion of eIF4e was completed in 2010.115–118 Patients with
stage4 cancer for which no proven therapy existed were
included, regardless of pathological diagnosis. Included
cancer types were mesothelioma, malignant melanoma,
non-small-cell lung carcinoma and several others. The
expression levels of eIF4e in tumour biopsy samples taken
after treatment were lower than those in pre-treatment
samples. Unfortunately, the study did not show any clini-
cal response, which again demonstrates the difficulty of
translating preclinical results to the bedside.118
Cell metabolism
In cancer, cell metabolism is often unbalanced.
Metabolism and catabolism are mediated and controlled
by a multitude of different enzymes, which are there-
fore other potential targets for bladder cancer therapy.
Invivo studies that investigated this strategy are listed
in Supplementary Table4 online. For example, stearoyl-
CoA desaturase1 is a rate-limiting enzyme in fatty acid
desaturation that was targeted by siRNA knockdown or
by an oral small-molecule inhibitor, both combined with
doxycycline therapy. Both strategies led to a significant
reduction in tumour growth in subcutaneous bladder
cancer cell xenografts in mice.119
Choline kinaseα (ChoKα) is a metabolic enzyme
responsible for the generation of phosphorylcholine
(the hydrophilic polar head group of phospholipids).120
Inhibition of this enzyme slowed tumour growth in both
subcutaneous as well as orthotopic xenografts in mice
and prolonged animal survival in the orthotopic mouse
model from a median of 34days to 42days.120 ChoKα is
also overexpressed in lung, breast, prostate, colon, ovary
and haematological malignancies.121 In 2014, a phaseI
trial testing a ChoKα inhibitor in patients with solid
tumours, including bladder cancer, finalized recruitment
but results have not been published yet.122
Two other studies have tested two other drugs that
target metabolic pathways.123,124 Treatment with a com-
bination of α-lipoic acid, which inhibits glycolysis, and
hydroxycitrate, which inhibits lipid synthesis, was able
to improve survival of mice with MBT-2 bladder tumour
xenografts. Addition of octreotide or cisplatin boosted
survival evenfurther.123,124
Overall, targeting protein translation seems to be a
promising strategy for the treatment of bladder cancer,
even though the study targeting eIF4e showed the dif-
ficulty of reproducing favourable preclinical results in a
phaseI clinical trial.
Metastasis, immune system and others
Other processes that involve potential molecular targets
for the treatment of patients with bladder cancer include
the formation of metastases and the immune response.
In addition, novel treatment strategies, such as oncolytic
viruses, are used to exploit tumour-cell-specific molecu-
lar mutations. A summary of studies investigating these
strategies is provided in Supplementary Table5 online.
Potential targets include proteins that interact with the
extracellular matrix, for example urokinase-type plas-
minogen activator (uPA), plasminogen activator inhibi-
tor1 (PAI-1) and versican core protein. These targets
are thought to be important in reducing the metastatic
potential of bladder cancer cells and, hence, might
be employed in preventive treatments.125–127 Another
promising target is the CD antigen CD24, a cell surface
protein and cancer stem cell marker that is involved in
metastatic progression in many cancers. In three pre-
clinical studies, one group showed the relevance of CD24
in bladder cancer development and metastasis.95,98,128
Inhibition of CD24 with an antibody or siRNAs reduced
tumour growth in mice with subcutaneous xenografts and
reduced metastatic load in mice inoculated with tumour
cells via their tail vein.95,98,128 Also, fewer tumours devel-
oped in CD24-knockout animals upon administration of
N-butyl-N-(4-hydroxybutyl)-nitrosamine, which induces
bladder tumours.98 Furthermore, the researchers found
that androgens stimulated CD24 expression in tumour
tissue and surgical castration of male mice inhibited
tumour growth in subcutaneous xenografts.98 As inhibi-
tion of CD24 reduced tumour growth in multiple studies,
we consider this strategy promising as a targeted therapy
in bladder cancer (Table1).
Stimulation of the immune system is another way to
target cancer cells. If the immune system is able to rec-
ognize tumour cells as foreign, the immune response
might destroy the malignant cells.129 One approach
investigated the use of IL-12 in the treatment of bladder
cancer, using transfection of EJ bladder cancer cells with a
plasmid encoding IL-12 to force secretion of IL-12.130 The
hypothesis was that secretion of IL-12 would stimulate
Tlympho cytes and natural killer cells, which should result
in increased tumour cell killing. Mice were injected sub-
cutaneously with nontransfected EJ cells or transfected EJ
cells.130 In addition, some mice were treated with injections
of pirarubicin, a doxorubicin analogue. Combination treat-
ment resulted in significantly smaller tumours compared
with single treatment or no treatment. Although these
results are positive, the need to introduce a gene into a high
percentage of tumour cells in a solid tumour makes this
approach unlikely to be translated into clinical application.
Other bladder cancer treatment strategies targeting the
immune system are currently under investigation in clini-
cal trials.131 These agents have so far not been analysed in
animal models of bladder cancer, but have proven their
efficacy in clinical trials in other types of cancer. Immune
checkpoint inhibitors, such as MPDL3280A (atezoli-
zumab), pembrolizumab, ipilimumab and nivolumab,
are examples of such promising therapies targeting the
immune system. MPDL3280A, an antibody against
REVIEWS
© 2015 Macmillan Publishers Limited. All rights reserved
690
|
DECEMBER 2015
|
VOLUME 12 www.nature.com/nrurol
programmed cell death1 ligand1 (PD-L1), was investi-
gated as a single-agent, second-line therapy in a phaseI
trial in 68 patients with bladder cancer.132 The objective
response rate was 43% in patients with tumours positive for
PD-L1 expression upon immunohistochemical analysis.
Reported overall toxicity was low. In 2014, the FDA granted
MPDL3280A the breakthrough status for bladder cancer,
which will expedite the development and review of the
drug by the FDA.133,134 Many clinical trials of MPDL3280A
are currently ongoing and results arepending.135–142
Finally, three invivo studies employed oncolytic adeno-
viruses.130,143,144 The viruses were genetically modified
in such a way that they specifically attach to a tumour-
associated membrane protein or the extracellular matrix
and then enter and lyse the tumour cells without damag-
ing normal cells. Although this strategy leads to a reduc-
tion of tumour growth, the infection efficiency of the
used viruses is probably not high enough for all tumour
cells to be targeted.145 To achieve extensive effectiveness,
infection should not only result in direct tumour cell
killing, but also antigen release and immune stimula-
tion.146 Thus, as oncolytic viruses have shown suboptimal
potency as monotherapies, combination treatments with,
for example, immune check point inhibitors or immune-
stimulatory molecules should be considered.146,147 For
example, in 2012, a phaseI trial investigating the efficacy
of intravesical administration of an oncolytic adenovirus
expressing granulocyte–macrophage colony-stimulating
factor demonstrated tolerable safety and a complete
response rate of 48.6%.148 However, at the moment, we do
not consider treatment strategies using oncolytic viruses
ready for implementation into the clinic.
Conclusions
Although targeted therapies have proven effective in
other cancer types, none of the registered therapies have
been approved for the treatment of bladder cancer. The
small number of clinical trials that included patients
with this disease had limited success.12 With this Review,
which explores all animal studies investigating targeted
therapeutic interventions for bladder cancer that were
published since 2009, we aimed to provide a current,
comprehensive overview of this exciting field of research.
We found several animal studies that reported tumour
growth inhibition or reduced angiogenesis and were
replicated by different groups or validated by the same
group in separate studies (Table1). Inhibition of FGFR3
seemed very successful using the small-molecule inhibi-
tor PD173074,48,51,52,54 as well as via the monoclonal anti-
body R3Mab49,50 and in combination with inhibition of
EGFR.54 These treatments resulted in tumour growth
inhibition, reduced angiogenesis and increased apoptosis.
Inhibition of EGFR alone using erlotinib was also success-
ful, resulting in tumour growth inhibition in some hetero-
topic xenografts and prolonged survival of mice with
orthotopic xenografts.29 However, the results of clinical
trials investigating therapeutic targeting of FGFR3 and
EGFR have been disappointing,12 most likely owing to the
lack of patient stratification according to the downstream
mutations in the pathways of these receptors.
Inhibition of angiogenesis through blocking VEGF
or its receptor reduced tumour growth in animal
models.30,59–61,149,150 Similarly, inhibition of the cancer stem
cell marker CD24 inhibited tumour growth,95,98,128 which
was also the case for treatments targeting mTOR and the
PI3K–Akt pathway,75–77 S TAT3 77,90,151 and the AR.92–95 Most
experiments showed a cytostatic treatment effect, which
suggests that tumour growth probably resumes when
treatment is stopped. Combining targeted treatment with
chemo therapy usually enhanced the therapeutic effi-
cacy. In clinical trials, combination treatments have been
attempted but have not demonstrated an improvement
in overall survival and sometimes resulted in increased
t oxicity in patients.12
EGFR inhibitors and antibodies have proven to be
effective as targeted therapies in humans with non-small-
cell lung cancer and colorectal cancer.25,26 This work dem-
onstrates that activating mutations in the Ras–MAPK and
PI3K–Akt pathways downstream of EGFR render the
tumour resistant to EGFR inhibition. We have identified
four studies on EGFR inhibition in human bladder cancer
cell line xenografts that had such mutations in one or both
pathways.27,28,30,31 In another study, the researchers care-
fully screened the EGFR downstream pathways of the cell
lines before attempting to target EGFR.29
One limitation of the studies discussed in this Review
is the use of cell lines to create xenografts. Cell lines
might not reflect the entire spectrum of tumour types,
especially because relatively few bladder cancer cell lines
are available. In addition, cell-line-derived tumours
probably lack the tumour heterogeneity observed in
most cancers. Thus, xenografts established directly from
resected tumours might present a better model than
cell lines. Such tumour models also have human stroma
and engage murine stromal cells, which is likely to be
important in cell–cell interactions and cell prolifera-
tion.18 Furthermore, small-molecule inhibitors, inhibi-
tion of cell surface proteins by monoclonal antibodies
and antibodies against ligands of cell surface receptors
have proven their efficacy in the clinic. Inhibition or
knockdown of targets using small, modified RNAs (for
example, siRNAs), transfection of genes and the use of
viral vehicles are efficient invitro but their effectiveness
in delivery is still not good enough for systemic treatment
of tumours and metastases.
Overall, our Review shows that a multitude of differ-
ent targets has been investigated and that most of the
studies have not been replicated to date. Often, relatively
few animals were used to register an effect. In addi-
tion, almost all studies reported inhibition of tumour
growth but not increased apoptosis. Also, only five
studies reported that the investigated treatment had no
effect.53,60,96,125,152 This small number suggests a possible
publication bias. In some papers, the number of treated
animals was not reported. In2010, the ARRIVE (animals
in research: reporting invivo experiments) guidelines—
developed to improve the design, analysis and report-
ing of animal experiments—were published.153,154 We
suggest that s cientists and s cientific journals adhere to
these guidelines in future studies.
REVIEWS
© 2015 Macmillan Publishers Limited. All rights reserved
NATURE REVIEWS
|
UROLOGY VOLUME 12
|
DECEMBER 2015
|
691
Furthermore, bladder cancer research is consistently
underfunded, considering the number of life years lost
and the costs to health-care systems generated by patients
with bladder cancer.155 This lack of funding is the major
reason why research on novel therapies in bladder cancer
lags behind that in other tumour types. The community
of bladder cancer researchers has to invest in reaching out
to the different policy makers and grant organizations to
improve funding.
In conclusion, EGFR, FGFR-3, VEGF, mTOR, STAT3,
the AR and CD24 represent promising and validated
targets for the inhibition of bladder cancer growth in
animal studies. Treatment strategies that combine differ-
ent modalities and/or targets seem to be most promising
and clinical studies should continue to pursue these find-
ings. Careful patient selection in these clinical studies will
be essential in order to demonstrate true benefit. Many of
the other targets that are under preclinical investigation
still need further validation.
Review criteria
The Embase, MEDLINE, OvidSP, Web of Science,
PubMed, Cochrane CENTRAL and Google Scholar
databases were systematically searched (date of
search: December 2014). The full search strategy
conducted in Embase was: (‘molecularly targeted
therapy’/de OR ‘molecular therapy’/exp OR (target*
NEAR/3 (molecul* OR therap* OR antibod* OR inhibit*
OR mirna* OR microrna* OR treat* OR intervent*OR
chemo*)):ab,ti) AND (‘bladder tumor’/exp OR ‘transitional
cell carcinoma’/de OR (‘carcinoma’/de AND ‘bladder’/
de) OR ((bladder OR urothel*) NEAR/3 (tumo* OR
cancer* OR neoplas* OR carcin* OR metasta*
ORtransitional)):ab,ti). After the literature search was
completed, all reviews, expert opinions, editorials,
seminar articles, congress abstracts, articles published
before 2009 and records in languages other than
English or Dutch were excluded. In total, 93 published
animal studies of targeting molecular targets in bladder
cancer were eligible for review.
1. Kompier, L.C. etal. The development of multiple
bladder tumour recurrences in relation to the
FGFR3 mutation status of the primary tumour.
J.Pathol. 218, 104–112 (2009).
2. Babjuk, M. etal. EAU guidelines on non-
muscle-invasive urothelial carcinoma of the
bladder: update 2013. Eur. Urol. 64, 639–653
(2013).
3. Kiemeney, L.A., Witjes, J.A., Heijbroek, R.P.,
Verbeek, A.L. & Debruyne, F.M. Predictability of
recurrent and progressive disease in individual
patients with primary superficial bladder cancer.
J. Urol. 150, 60–64 (1993).
4. Sylvester, R.J. etal. Predicting recurrence and
progression in individual patients with stage Ta
T1 bladder cancer using EORTC risk tables:
acombined analysis of 2596 patients from
seven EORTC trials. Eur. Urol. 49, 466–477
(2006).
5. Surveillance, Epidemiology, and End Results
Program (SEER). seer.cancer.gov [online], http://
seer.cancer.gov/statfacts/html/urinb.html
(2015).
6. Ploussard, G. etal. Conditional survival after
radical cystectomy for bladder cancer:
evidencefor a patient changing risk profile over
time. Eur.Urol. 66, 361–370 (2014).
7. International Collaboration of Trialists etal.
International phaseIII trial assessing
neoadjuvant cisplatin, methotrexate, and
vinblastine chemotherapy for muscle-invasive
bladder cancer: long-term results of the BA06
30894 trial. J. Clin. Oncol. 29, 2171–2177
(2011).
8. Visser, O. & Horenblas, S. Blaascarcinoom:
Kankerzorg in beeld. Intergraal kankercentrum
Nederland [online], https://www.iknl.nl/docs/
default-source/KIB-rapportages/portfolio_kib_
blaascarcinoom.pdf?sfvrsn=2 (2014).
9. Abdollah, F. etal. Incidence, survival and
mortality rates of stage-specific bladder
cancerin United States: a trend analysis.
CancerEpidemiol. 37, 219–225 (2013).
10. Ferlay, J. etal. GLOBOCAN 2012 v1.0, Cancer
Incidence and Mortality Worldwide: IARC
CancerBase No. 11. Lyon, France: International
Agency for Research on Cancer [online], http://
globocan.iarc.fr (2013).
11. DeVita, V.T. Jr, & Chu, E. A history of cancer
chemotherapy. Cancer Res. 68, 8643–8653
(2008).
12. Ghosh, M., Brancato, S.J., Agarwal, P.K.
&Apolo,A.B. Targeted therapies in urothelial
carcinoma. Curr. Opin. Oncol. 26, 305–320
(2014).
13. Gerlinger, M. etal. Intratumour heterogeneity
inurologic cancers: from molecular evidence
toclinical implications. Eur. Urol. 67, 729–737
(2015).
14. Diaz, L.A. Jr. etal. The molecular evolution of
acquired resistance to targeted EGFR blockade in
colorectal cancers. Nature 486, 537–540 (2012).
15. Philips, G.K. etal. A phaseII trial of cisplatin,
fixed dose-rate gemcitabine and gefitinib for
advanced urothelial tract carcinoma: results
ofthe Cancer and Leukaemia Group B 90102.
BJUInt. 101, 20–25 (2007).
16. Meeks, J.J. etal. A systematic review of
neoadjuvant and adjuvant chemotherapy for
muscle-invasive bladder cancer. Eur. Urol. 62,
523–533 (2012).
17. Sausville, E.A. & Burger, A.M. Contributions
ofhuman xenografts to anticancer drug
development. Cancer Res. 66, 3351–3354
(2006).
18. Jäger, W. etal. Patient-derived bladder cancer
xenografts in the preclinical development
ofnovel targeted therapies. Oncotarget 6,
21522–21532 (2015).
19. Jäger, W. etal. Hiding in plain view: genetic
profiling reveals decades old cross
contamination of bladder cancer cell line KU7
with HeLa. J. Urol. 190, 1404–1409 (2013).
20. Schlessinger, J. Cell signaling by receptor
tyrosine kinases. Cell 103, 211–225 (2000).
21. Cancer Genome Atlas Research Network.
Comprehensive molecular characterization
ofurothelial bladder carcinoma. Nature 507,
315–322 (2014).
22. Kim, P.H. etal. Genomic predictors of survival
inpatients with high-grade urothelial carcinoma
of the bladder. Eur. Urol. 67, 198–201 (2015).
23. Black, P.C., Agarwal, P.K. & Dinney, C.P.
Targeted therapies in bladder cancer—an
update. Urol. Oncol. 25, 433–438 (2007).
24. Kassouf, W. etal. Distinctive expression pattern
of ErbB family receptors signifies an aggressive
variant of bladder cancer. J. Urol. 179, 353–358
(2008).
25. Gerber, D.E. Targeted therapies: a new
generation of cancer treatments. Am. Fam.
Physician 77, 311–319 (2008).
26. Krause, D.S. & Van Etten, R.A. Tyrosine kinases
as targets for cancer therapy. N. Engl. J. Med.
353, 172–187 (2005).
27. Bhuvaneswari, R., Gan, Y.Y., Soo, K.C. & Olivo, M.
Targeting EGFR with photodynamic therapy in
combination with erbitux enhances invivo bladder
tumor response. Mol. Cancer 8, 94 (2009).
28. Li, F., Ye, Z.Q., Guo, D.S. & Yang, W.M.
Suppression of bladder cancer cell
tumorigenicity in an athymic mouse model by
adenoviral vector-mediated transfer of LRIG1.
Oncol. Rep. 26, 439–446 (2011).
29. Rebouissou, S. etal. EGFR as a potential
therapeutic target for a subset of muscle-
invasive bladder cancers presenting a basal-like
phenotype. Sci. Transl. Med. 6, 244ra91 (2014).
30. Bhuvaneswari, R., Thong, P., Yuen, G.Y., Olivo, M.
& Chee, S.K. Combined use of anti-VEGF
andanti-EGFR monoclonal antibodies with
photodynamic therapy suppresses tumor growth
in an invivo tumor model. J. Cancer Sci. Ther. 5,
100–105 (2013).
31. Quesnelle, K.M. & Grandis, J.R. Dual kinase
inhibition of EGFR and HER2 overcomes
resistance to cetuximab in a novel invivo model
of acquired cetuximab resistance. Clin. Cancer
Res. 17, 5935–5944 (2011).
32. Catalogue of somatic mutations in cancer
(COSMIC). cancer.sanger.ac.uk [online], http://
cancer.sanger.ac.uk/cell_lines (2015).
33. Halling-Brown, M.D. etal. canSAR: an integrated
cancer public translational research and drug
discovery resource. Nucleic Acids Res. 40,
D947–D956 (2012).
34. Karapetis, C.S. etal. K-ras mutations and
benefit from cetuximab in advanced colorectal
cancer. N. Engl. J. Med. 359, 1757–1765 (2008).
35. Pruthi, R.S. etal. A phaseII trial of neoadjuvant
erlotinib in patients with muscle-invasive bladder
cancer undergoing radical cystectomy: clinical
and pathological results. BJU Int. 106, 349–354
(2010).
36. Philips, G.K. etal. A phaseII trial of cisplatin (C),
gemcitabine (G) and gefitinib for advanced
urothelial tract carcinoma: results of Cancer and
Leukemia Group B (CALGB) 90102. Ann. Oncol.
20, 1074–1079 (2009).
37. Hussain, M. etal. A randomized phase2 trial of
gemcitabine/cisplatin with or without cetuximab
in patients with advanced urothelial carcinoma.
Cancer 120, 2684–2693 (2014).
REVIEWS
© 2015 Macmillan Publishers Limited. All rights reserved
692
|
DECEMBER 2015
|
VOLUME 12 www.nature.com/nrurol
38. Petrylak, D.P. etal. Results of the Southwest
Oncology Group phaseII evaluation (study
S0031) of ZD1839 for advanced transitional
cellcarcinoma of the urothelium. BJU Int. 105,
317–321 (2010).
39. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT00238420 (2015).
40. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT02006667 (2015).
41. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT01828736 (2013).
42. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT01191892 (2012).
43. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT01382706 (2015).
44. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT02013765 (2014).
45. Groenendijk, F.H. etal. ERBB2 mutations
characterize a subgroup of muscle-invasive
bladder cancers with excellent response to
neoadjuvant chemotherapy. Eur. Urol. http://
dx.doi.org/10.1016/j.eururo.2015.01.014.
46. van Rhijn, B.W. etal. Molecular grade (FGFR3/
MIB-1) and EORTC risk scores are predictive in
primary non-muscle-invasive bladder cancer.
Eur.Urol. 58, 433–441 (2010).
47. Tomlinson, D.C., Baldo, O., Harnden, P.
&Knowles, M.A. FGFR3 protein expression
andits relationship to mutation status and
prognostic variables in bladder cancer. J. Pathol.
213, 91–98 (2007).
48. Miyake, M. etal. 1-tert-butyl-3-[6-(3,5-dimethoxy-
phenyl)-2-(4-diethylamino-butylamino)-
pyrido[2,3-d]pyrimidin-7-yl]-urea (PD173074),
aselective tyrosine kinase inhibitor of fibroblast
growth factor receptor-3 (FGFR3), inhibits cell
proliferation of bladder cancer carrying the
FGFR3 gene mutation along with up-regulation
ofp27/Kip1 and G1/G0 arrest. J. Pharmacol. Exp.
Ther. 332, 795–802 (2010).
49. Qing, J. etal. Antibody-based targeting of FGFR3
in bladder carcinoma and t(4;14)-positive
multiple myeloma in mice. J. Clin. Invest. 119,
1216–1229 (2009).
50. Gust, K.M. etal. Fibroblast growth factor
receptor3 is a rational therapeutic target
inbladder cancer. Mol. Cancer Ther. 12,
1245–1254 (2013).
51. Wu, Y.M. etal. Identification of targetable FGFR
gene fusions in diverse cancers. Cancer Discov.
3, 636–647 (2013).
52. Lamont, F.R. etal. Small molecule FGF
receptorinhibitors block FGFR-dependent
urothelial carcinoma growth invitro and invivo.
Br. J. Cancer 104, 75–82 (2011).
53. Cheng, T. etal. Fibroblast growth factor
receptors-1 and -3 play distinct roles in
theregulation of bladder cancer growth
andmetastasis: implications for therapeutic
targeting. PLoS ONE 8, e57284 (2013).
54. Herrera-Abreu, M.T. etal. Parallel RNA
interference screens identify EGFR activation as
an escape mechanism in FGFR3-mutant cancer.
Cancer Discov. 3, 1058–1071 (2013).
55. Milowsky, M.I. etal. Phase2 trial of dovitinib
inpatients with progressive FGFR3-mutated
orFGFR3 wild-type advanced urothelial
carcinoma. Eur. J. Cancer 50, 3145–3152
(2014).
56. Galsky, M.D. etal. PhaseIb study of dovitinib
incombination with gemcitabine plus cisplatin
orgemcitabine plus carboplatin in patients with
advanced solid tumors. Cancer Chemother.
Pharmacol. 74, 465–471 (2014).
57. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT01732107 (2015).
58. Carmeliet, P. VEGF as a key mediator of
angiogenesis in cancer. Oncology 69 (Suppl. 3),
4–10 (2005).
59. Feng, W. etal. siRNA-mediated knockdown
ofVEGF-A, VEGF-C and VEGFR-3 suppresses
thegrowth and metastasis of mouse bladder
carcinoma invivo. Exp. Ther. Med. 1, 899–904
(2010).
60. Yang, H. etal. Soluble vascular endothelial
growth factor receptor-3 suppresses
lymphangiogenesis and lymphatic metastasis
inbladder cancer. Mol. Cancer 10, 36 (2011).
61. Liang, P.H. etal. Construction of a DNA vaccine
encoding Flk-1 extracellular domain and C3d
fusion gene and investigation of its suppressing
effect on tumor growth. Cancer Immunol.
Immunother. 59, 93–101 (2010).
62. Chaudhary, U.B. etal. PhaseII trial of
neoadjuvant cisplatin, gemcitabine, and
bevacizumab followed by radical cystectomy (RC)
in patients with muscle-invasive transitional
cellcarcinoma (TCC) of the bladder. J. Clin. Oncol.
29 (Suppl. 7), abstract 276 (2011).
63. Siefker-Radtke, A.O. etal. Neoadjuvant
chemotherapy with DD-MVAC and bevacizumab in
high-risk urothelial cancer: results from a phaseII
trial at the M.D. Anderson Cancer Center.
J.Clin.Oncol. 30(Suppl.5), abstract 261 (2012).
64. Balar, A.V. etal. PhaseII study of gemcitabine,
carboplatin, and bevacizumab in patients with
advanced unresectable or metastatic urothelial
cancer. J. Clin. Oncol. 31, 724–730 (2013).
65. Galsky, M.D. etal. Gemcitabine, cisplatin, and
sunitinib for metastatic urothelial carcinoma
andas preoperative therapy for muscle-invasive
bladder cancer. Clin. Genitourin. Cancer 11,
175–181 (2013).
66. Hahn, N.M. etal. PhaseII trial of cisplatin,
gemcitabine, and bevacizumab as first-line
therapy for metastatic urothelial carcinoma:
Hoosier Oncology Group GU 04–75. J. Clin.
Oncol. 29, 1525–1530 (2011).
67. Gallagher, D.J. etal. PhaseII study of sunitinib
in patients with metastatic urothelial cancer.
J.Clin. Oncol. 28, 1373–1379 (2010).
68. Twardowski, P. etal. PhaseII study of aflibercept
(VEGF-Trap) in patients with recurrent or
metastatic urothelial cancer, a California
CancerConsortium Trial. Urology 76, 923–926
(2010).
69. Pili, R. etal. A phaseII safety and efficacy study
of the vascular endothelial growth factor
receptor tyrosine kinase inhibitor pazopanib
inpatients with metastatic urothelial cancer.
Clin. Genitourin. Cancer 11, 477–483 (2013).
70. Necchi, A. etal. Pazopanib in advanced and
platinum-resistant urothelial cancer: an open-
label, single group, phase2 trial. Lancet Oncol.
13, 810–816 (2012).
71. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT00942331 (2015).
72. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT01108055 (2014).
73. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT00794950 (2015).
74. Puzio-Kuter, A.M. etal. Inactivation of p53
andPten promotes invasive bladder cancer.
Genes Dev. 23, 675–680 (2009).
75. Makhlin, I. etal. The mTOR pathway affects
proliferation and chemosensitivity of urothelial
carcinoma cells and is upregulated in a subset
of human bladder cancers. BJU Int. 108, E84–E90
(2011).
76. Seront, E. etal. PTEN deficiency is associated
with reduced sensitivity to mTOR inhibitor in
human bladder cancer through the unhampered
feedback loop driving PI3K/Akt activation. Br. J .
Cancer 109, 1586–1592 (2013).
77. Zhou, H. etal. Urothelial tumor initiation requires
deregulation of multiple signaling pathways:
Implications in target-based therapies.
Carcinogenesis 33, 770–780 (2012).
78. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT01090466 (2012).
79. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT01182168 (2015).
80. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT01215136 (2015).
81. Seront, E. etal. PhaseII study of everolimus
inpatients with locally advanced or
metastatictransitional cell carcinoma of the
urothelial tract: clinical activity, molecular
response, and biomarkers. Ann. Oncol. 23,
2663–2670 (2012).
82. Gerullis, H. etal. A phaseII trial of temsirolimus
in second-line metastatic urothelial cancer.
Med.Oncol. 29, 2870–2876 (2012).
83. Milowsky, M.I. etal. PhaseII study of everolimus
in metastatic urothelial cancer. BJU Int. 112,
462–470 (2013).
84. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT01551030 (2014).
85. Rodon, J. etal. PhaseI dose-escalation and
-expansion study of buparlisib (BKM120),
anoralpan-classI PI3K inhibitor, in patients
withadvanced solid tumors. Invest. New Drugs
32, 670–681 (2014).
86. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT01259063 (2015).
87. Harada, Y. etal. Cell-permeable peptide
DEPDC1-ZNF224 interferes with transcriptional
repression and oncogenicity in bladder cancer
cells. Cancer Res. 70, 5829–5839 (2010).
88. Kodaira, K. etal. Potent cytotoxic effect of
anovel nuclear factor-κB inhibitor
dehydroxymethylepoxyquinomicin on human
bladder cancer cells producing various
cytokines. Urology 75, 805–812 (2010).
89. Shanmugam, R. etal. A water soluble
parthenolide analog suppresses invivo tumor
growth of two tobacco-associated cancers, lung
and bladder cancer, by targeting NF-κ B and
generating reactive oxygen species. Int. J. Cancer
128, 2481–2494 (2011).
90. Sen, M. etal. Targeting Stat3 abrogates EGFR
inhibitor resistance in cancer. Clin. Cancer Res.
18, 4986–4996 (2012).
91. Zhu, J. etal. Downregulation of integrin-linked
kinase inhibits epithelial-to-mesenchymal
transition and metastasis in bladder cancer
cells. Cell Signal. 24, 1323–1332 (2012).
92. Shiota, M. etal. Androgen receptor signaling
regulates cell growth and vulnerability to
doxorubicin in bladder cancer. J. Urol. 188,
276–286 (2012).
93. Wu, J.T., Han, B.M., Yu, S.Q., Wang, H.P.
&Xia,S.J. Androgen receptor is a potential
therapeutic target for bladder cancer. Urology
75,820–827 (2010).
94. Jitao, W. etal. Androgen receptor inducing
bladder cancer progression by promoting an
epithelial–mesenchymal transition. Andrologia
46, 1128–1133 (2014).
REVIEWS
© 2015 Macmillan Publishers Limited. All rights reserved
NATURE REVIEWS
|
UROLOGY VOLUME 12
|
DECEMBER 2015
|
693
95. Overdevest, J.B. etal. CD24 offers a
therapeutic target for control of bladder cancer
metastasis based on a requirement for lung
colonization. Cancer Res. 71, 3802–3811
(2011).
96. Ozawa, A. etal. Inhibition of bladder tumour
growth by histone deacetylase inhibitor. BJU Int.
105, 1181–1186 (2010).
97. Miyamoto, H. etal. Promotion of bladder cancer
development and progression by androgen
receptor signals. J. Natl Cancer Inst. 99,
558–568 (2007).
98. Overdevest, J.B. etal. CD24 expression is
important in male urothelial tumorigenesis and
metastasis in mice and is androgen regulated.
Proc. Natl Acad. Sci. USA 109, E3588–E3596
(2012).
99. Zhou, N. etal. The investigational aurora kinase
A inhibitor MLN8237 induces defects in cell
viability and cell-cycle progression in malignant
bladder cancer cells invitro and invivo.
Clin.Cancer Res. 19, 1717–1728 (2013).
100. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT02109328 (2014).
101. Knowles, M. What we could do now: molecular
pathology of bladder cancer. Mol. Pathol. 54,
215–221 (2001).
102. Müllauer, L. etal. Mutations in apoptosis genes:
a pathogenetic factor for human disease.
Mutat.Res. 488, 211–231 (2001).
103. Madka, V. etal. p53-stabilizing agent CP-31398
prevents growth and invasion of urothelial
cancer of the bladder in transgenic UPII-SV40T
mice. Neoplasia 15, 966–974 (2013).
104. Zhang, Z. etal. Urothelium-specific expression
ofan oncogene in transgenic mice induced the
formation of carcinoma insitu and invasive
transitional cell carcinoma. Cancer Res. 59,
3512–3517 (1999).
105. Card, K. etal. A soluble single-chain T-cell
receptor IL-2 fusion protein retains MHC-
restricted peptide specificity and IL-2 bioactivity.
Cancer Immunol. Immunother. 53, 345–357
(2004).
106. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT01326871 (2015).
107. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT01625260 (2015).
108. Yoshimine, S. etal. Prognostic significance of
Bcl-xL expression and efficacy of Bcl-xL targeting
therapy in urothelial carcinoma. Br. J. Cancer 11,
2312–2320 (2013).
109. Gazzaniga, P. etal. Expression and prognostic
significance of LIVIN, SURVIVIN and other
apoptosis-related genes in the progression
ofsuperficial bladder cancer. Ann. Oncol. 14,
85–90 (2003).
110. Liu, C.A. etal. Antisense oligonucleotide
targeting livin induces apoptosis of human
bladder cancer cell via a mechanism involving
caspase3. J. Exp. Clin. Cancer Res. 29, 63
(2010).
111. Wang, H. etal. Translation initiation factor eIF3b
expression in human cancer and its role in tumor
growth and lung colonization. Clin. Cancer Res.
19, 2850–2860 (2013).
112. Amit, D. & Hochberg, A. Development of targeted
therapy for bladder cancer mediated by a double
promoter plasmid expressing diphtheria toxin
under the control of H19 and IGF2-P4 regulatory
sequences. J. Transl. Med. 8, 134 (2010).
113. Amit, D., Tamir, S., Birman, T., Gofrit, O.N.
&Hochberg, A. Development of targeted
therapyfor bladder cancer mediated by a double
promoter plasmid expressing diphtheria toxin
under the control of IGF2-P3 and IGF2-P4
regulatory sequences. Int. J. Clin. Exp. Med. 4,
91–102 (2011).
114. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT00903708 (2012).
115. Graff, J.R. etal. Therapeutic suppression of
translation initiation factor eIF4E expression
reduces tumor growth without toxicity. J. Clin.
Invest. 117, 2638–2648 (2007).
116. Graff, J.R., Konicek, B.W., Carter, J.H.
&Marcusson, E.G. Targeting the eukaryotic
translation initiation factor 4E for cancer therapy.
Cancer Res. 68, 631–634 (2008).
117. Hsieh, T.F. etal. Epidermal growth factor
enhances androgen receptor-mediated bladder
cancer progression and invasion via potentiation
of AR transactivation. Oncol. Rep. 30,
2917–2922 (2013).
118. Hong, D.S. etal. A phase1 dose escalation,
pharmacokinetic, and pharmacodynamic
evaluation of eIF-4E antisense oligonucleotide
LY2275796 in patients with advanced cancer.
Clin. Cancer Res. 17, 6582–6591 (2011).
119. Du, X. etal. FGFR3 stimulates stearoyl CoA
desaturase 1 activity to promote bladder tumor
growth. Cancer Res. 72, 5843–5855 (2012).
120. Hernando, E. etal. A critical role for choline
kinase-α in the aggressiveness of bladder
carcinomas. Oncogene 28, 2425–2435 (2009).
121. de la Cueva, A. etal. Combined 5-FU and ChoKα
inhibitors as a new alternative therapy of
colorectal cancer: evidence in human tumor-
derived cell lines and mouse xenografts.
PLoSONE 8, e64961 (2013).
122. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT01215864 (2014).
123. Abolhassani, M. etal. Screening of well-
established drugs targeting cancer metabolism:
Reproducibility of the efficacy of a highly
effective drug combination in mice. Invest. New
Drugs 30, 1331–1342 (2012).
124. Guais, A. etal. Adding a combination of
hydroxycitrate and lipoic acid (METABLOC™)
tochemotherapy improves effectiveness against
tumor development: experimental results and
case report. Invest. New Drugs 30, 200–211
(2012).
125. Allen, M.D. etal. Prognostic and therapeutic
impact of argininosuccinate synthetase 1 control
in bladder cancer as monitored longitudinally
byPET imaging. Cancer Res. 74, 896–907
(2014).
126. Parviainen, S. etal. CD40 ligand and tdTomato-
armed vaccinia virus for induction of antitumor
immune response and tumor imaging.
GeneTher. 21, 195–204 (2014).
127. Gomes-Giacoia, E., Miyake, M., Goodison, S.
&Rosser, C.J. Targeting plasminogen activator
inhibitor-1 inhibits angiogenesis and tumor
growth in a human cancer xenograft model.
Mol.Cancer Ther. 12, 2697–2708 (2013).
128. Thomas, S. etal. CD24 is an effector of
HIF-1-driven primary tumor growth and
metastasis. Cancer Res. 72, 5600–5612
(2012).
129. Yen, M.C. etal. A novel cancer therapy by skin
delivery of indoleamine 2,3-dioxygenase siRNA.
Clin. Cancer Res. 15, 641–649 (2009).
130. Liu, W., Cao, Y., Fernandez, M.I., Niu, H. & Xiu, Y.
Additive antitumoral effect of interleukin-12 gene
therapy and chemotherapy in the treatment of
urothelial bladder cancer invitro and invivo.
Int.Urol. Nephrol. 43, 721–727 (2011).
131. Kim, J.W., Tomita, Y., Trepel, J. & Apolo, A.B.
Emerging immunotherapies for bladder cancer.
Curr. Opin. Oncol. 27, 191–200 (2015).
132. Powles, T. etal. MPDL3280A (anti-PD-L1)
treatment leads to clinical activity in metastatic
bladder cancer. Nature 515, 3558–62 (2014).
133. Fenner, A. Could MPDL3280A offer a therapeutic
breakthrough in metastatic bladder cancer?
Nat.Rev. Urol. 12, 61 (2015).
134. U.S. Food and Drug Administration. fda.gov
[online], http://www.fda.gov/
RegulatoryInformation/Legislation/
SignificantAmendmentstotheFDCAct/FDASIA/
ucm329491.htm (2015).
135. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT02108652 (2015).
136. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT02302807 (2015).
137. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT02351739 (2015).
138. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT02256436 (2015).
139. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT02335424 (2015).
140. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT02324582 (2015).
141. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT01928394 (2015).
142. US National Library of Medicine. ClinicalTrials.gov
[online], https://clinicaltrials.gov/ct2/show/
NCT02387996 (2015).
143. Fujita, T. etal. Combined therapeutic effects
ofadenoviral vector-mediated GLIPR1 gene
therapy and radiotherapy in prostate and
bladder cancer models. Urol. Oncol. 32, 92–100
(2014).
144. Wang, H. etal. Enhanced antitumor efficacy of
integrin-targeted oncolytic adenovirus
AxdAdB3-F/RGD on bladder cancer. Urology 83,
508.e513–508.e519 (2014).
145. Alemany, R. Viruses in cancer treatment.
Clin.Transl. Oncol. 15, 182–188 (2013).
146. Coffin, R.S. From virotherapy to oncolytic
immunotherapy: where are we now? Curr. Opin.
Virol. 13, 93–100 (2015).
147. Bauzon, M. & Hermiston, T. Armed therapeutic
viruses—a disruptive therapy on the horizon of
cancer immunotherapy. Front. Immunol. 5, 1–10
(2014).
148. Burke, J.M. etal. A first human phase1 study
ofCG0070, a GM-CSF expressing oncolytic
adenovirus, for the treatment of nonmuscle
invasive bladder cancer. J. Urol. 188,
2391–2397 (2012).
149. Wu, C.L., Ping, S.Y., Yu, C.P. & Yu, D.S. Tyrosine
kinase receptor inhibitor-targeted combined
chemotherapy for metastatic bladder cancer.
Kaohsiung J. Med. Sci. 28, 194–203 (2012).
150. Li, X. etal. The differential expression of EphB2
and EphB4 receptor kinases in normal bladder
and in transitional cell carcinoma of the bladder.
PloS One 9, e105326 (2014).
151. Zhang, B., Lu, Z., Hou, Y., Hu, J. & Wang, C.
Theeffects of STAT3 and survivin silencing
onthe growth of human bladder carcinoma cells.
Tumour Biol. 35, 5401–5407 (2014).
152. van der Horst, G. etal. Targeting of α-v integrins
reduces malignancy of bladder carcinoma.
PLoSONE 9, e108464 (2014).
153. Kilkenny, C., Browne, W.J., Cuthill, I.C.,
Emerson, M. & Altman, D.G. Improving
bioscience research reporting: the ARRIVE
guidelines for reporting animal research.
PLoSBiol. 8, e1000412 (2010).
REVIEWS
© 2015 Macmillan Publishers Limited. All rights reserved
694
|
DECEMBER 2015
|
VOLUME 12 www.nature.com/nrurol
154. Baker, D., Lidster, K., Sottomayor, A. & Amor, S.
Two years later: journals are not yet enforcing
the ARRIVE guidelines on reporting standards
forpre-clinical animal studies. PLoS Biol. 12,
e1001756 (2014).
155. Carter, A.J. & Nguyen, C.N. A comparison of
cancer burden and research spending reveals
discrepancies in the distribution of research
funding. BMC Public Health 12, 526 (2012).
156. Joung, Y.H. etal. Combination of AG490,
aJak2inhibitor, and methylsulfonylmethane
synergistically suppresses bladder tumor growth
via the Jak2/STAT3 pathway. Int. J. Oncol. 44,
883–895 (2014).
157. Qin, J. etal. Invitro and invivo inhibitory effect
evaluation of cyclooxygenase-2 inhibitors,
antisense cyclooxygenase-2 cDNA, and their
combination on the growth of human bladder
cancer cells. Biomed. Pharmacother. 63,
241–248 (2009).
158. Chen, S.C., Henry, D.O., Hicks, D.G.,
Reczek,P.R. & Wong, M.K. Intravesical
administration ofplasminogen activator inhibitor
type-1 inhibits invivo bladder tumor invasion and
progression. J. Urol. 181, 336–342 (2009).
159. Bhattacharya, A. etal. Allyl isothiocyanate-rich
mustard seed powder inhibits bladder cancer
growth and muscle invasion. Carcinogenesis 31,
2105–2110 (2010).
160. Bhattacharya, A., Li, Y., Geng, F., Munday, R. &
Zhang, Y. The principal urinary metabolite of allyl
isothiocyanate, N-acetyl-S-(N-allylthiocarbamoyl)
cysteine, inhibits the growth and muscle
invasion of bladder cancer. Carcinogenesis 33,
394–398 (2012).
161. Gao, J. etal. Small interfering RNA targeting
integrin-linked kinase inhibited the growth and
induced apoptosis in human bladder cancer
cells. Int. J. Biochem. Cell Biol. 43, 1294–1304
(2011).
162. Cheng, H. etal. shRNA targeting PLCε inhibits
bladder cancer cell growth invitro and invivo.
Urology 78, e7–e11 (2011).
163. Wu, C.T., Chang, Y.H., Lin, P., Chen, W.C.
&Chen, M.F. Thrombomodulin expression
regulates tumorigenesis in bladder cancer.
BMCCancer 14, 375 (2014).
164. Garg, M. etal. Heat-shock protein 70–72
(HSP70–2) expression in bladder urothelial
carcinoma is associated with tumour
progression and promotes migration and
invasion. Eur. J. Cancer 46, 207–215 (2010).
165. Niedworok, C. etal. The impact of the receptor
ofhyaluronan-mediated motility (RHAMM) on
human urothelial transitional cell cancer of the
bladder. PLoS ONE 8, e75681 (2013).
166. Mansure, J.J. etal. A novel mechanism of
PPARγ induction via EGFR signalling constitutes
rational for combination therapy inbladder
cancer. PLoS ONE 8, e55997 (2013).
167. Shim, J.S. etal. Effect of nitroxoline on
angiogenesis and growth of human bladder
cancer. J. Natl Cancer Inst. 102, 1855–1873
(2010).
168. Shimada, K., Anai, S., Marco, D.A., Fujimoto, K.
& Konishi, N. Cyclooxygenase 2-dependent and
independent activation of Akt through casein
kinase 2 contributes to human bladder cancer
cell survival. BMC Urol. 11, 8 (2011).
169. Matsui, Y. etal. Intravesical combination
treatment with antisense oligonucleotides
targeting heat shock protein-27 and HTI-286 as
a novel strategy for high-grade bladder cancer.
Mol. Cancer Ther. 8, 2402–2411 (2009).
170. Cho, S.D. etal. Activation of nerve growth factor-
induced B α by methylene-substituted
diindolynmethanes in bladder cancer cells
induces apoptosis and inhibits tumor growth.
Mol. Pharmacol. 77, 396–404 (2010).
171. Jeong, K.C. etal. Intravesical instillation of c-MYC
inhibitor KSI-3716 suppresses orthotopic bladder
tumor growth. J. Urol. 191, 510–518 (2014).
172. Ding, S. etal. A potent chemotherapeutic
strategy for bladder cancer: (S)-methoxy-trityl-
L-cystein, a novel Eg5 inhibitor. J. Urol. 184,
1175–1181 (2010).
173. Seo, H.K. etal. Development of replication-
competent adenovirus for bladder cancer
bycontrolling adenovirus E1a and E4 gene
expression with the survivin promoter.
Oncotarget 5, 5615–5623 (2014).
174. Seth, S. etal. RNAi-based therapeutics targeting
survivin and PLK1 for treatment of bladder
cancer. Mol. Ther. 19, 928–935 (2011).
175. Shimada, K. etal. Role of syndecan-1 (CD138)
incell survival of human urothelial carcinoma.
Cancer Sci. 101, 155–160 (2010).
176. Belgorosky, D. etal. Inhibition of nitric oxide is a
good therapeutic target for bladder tumors that
express iNOS. Nitric Oxide 36, 11–18 (2014).
177. Han, C. etal. Target expression of
staphylococcus enterotoxin A from an oncolytic
adenovirus suppresses mouse bladder tumor
growth and recruits CD3+ Tcell. Tumor Biol. 34,
2863–2869 (2013).
178. Fan, Y. etal. TGF-β-induced upregulation of
malat1 promotes bladder cancer metastasis
byassociating with suz12. Clin. Cancer Res. 20,
1531–1541 (2014).
179. Lin, C.C. etal. A novel adjuvant Ling Zhi-8
enhances the efficacy of DNA cancer vaccine
byactivating dendritic cells. Cancer Immunol.
Immunother. 60, 1019–1027 (2011).
180. Acquaviva, J. etal. FGFR3 translocations in
bladder cancer: differential sensitivity to HSP90
inhibition based on drug metabolism.
Mol.Cancer Res. 12, 1042–1054 (2014).
181. Liu, Y. & Kwiatkowski, D.J. Combined CDKN1A/
TP53 mutation in bladder cancer is a
therapeutictarget. Mol. Cancer Ther. 14,
174–182 (2015).
182. Levitt, J.M., Yamashita, H., Jian, W., Lerner, S.P.
& Sonpavde, G. Dasatinib is preclinically active
against Src-overexpressing human transitional
cell carcinoma of the urothelium with activated
Src signaling. Mol. Cancer Ther. 9, 1128–1135
(2010).
183. Marra, E. etal. Growth delay of human bladder
cancer cells by prostate stem cell antigen
downregulation is associated with activation
ofimmune signaling pathways. BMC Cancer 10,
129 (2010).
184. Liu, A.G. etal. RNA interference targeting
adrenomedullin induces apoptosis and reduces
the growth of human bladder urothelial cell
carcinoma. Med. Oncol. 30, 616 (2013).
185. Wei, J. etal. The inhibition of human bladder
cancer growth by calcium carbonate/CaIP6
nanocomposite particles delivering AIB1 siRNA.
Biomaterials 34, 1246–1254 (2013).
186. Chen, H. etal. MicroRNA-449a acts as a tumor
suppressor in human bladder cancer through the
regulation of pocket proteins. Cancer Lett. 320,
40–47 (2012).
187. Zhou, H. etal. Development and characterization
of a potent immunoconjugate targeting the Fn14
receptor on solid tumor cells. Mol. Cancer Ther.
10, 1276–1288 (2011).
188. Liang, P.Y. etal. Overexpression of
immunoglobulin G prompts cell proliferation
andinhibits cell apoptosis in human urothelial
carcinoma. Tumor Biol. 34, 1783–1791
(2013).
189. Zhao, W. etal. Steroid receptor coactivator-3
regulates glucose metabolism in bladder cancer
cells through coactivation of hypoxia inducible
factor 1α. J. Biol. Chem. 289, 11219–11229
(2014).
190. Mao, L. etal. Replication-competent adenovirus
expressing TRAIL synergistically potentiates the
antitumor effect of gemcitabine in bladder
cancer cells. Tumour Biol. 35, 5937–5944
(2014).
191. Li, C. etal. BCMab1, a monoclonal antibody
against aberrantly glycosylated integrin α3β1,
has potent antitumor activity ofbladder cancer
invivo. Clin. Cancer Res. 20, 4001–4013
(2014).
192. Said, N., Sanchez-Carbayo, M., Smith, S.C.
&Theodorescu, D. RhoGDI2 suppresses lung
metastasis in mice by reducing tumor versican
expression and macrophage infiltration. J. Clin.
Invest. 122, 1503–1518 (2012).
193. Foulks, J.M. etal. A small-molecule inhibitor
ofPIM kinases as a potential treatment for
urothelial carcinomas. Neoplasia 16, 403–412
(2014).
194. Zhao, F. etal. Knockdown of a novel lincRNA
AATBC suppresses proliferation and induces
apoptosis in bladder cancer. Oncotarget 6,
1064–1078 (2015).
Acknowledgements
We would like to acknowledge W.M. Bramer for his
contribution to the literature search and database
management. This study was funded by an Erasmus
MC 2012 grant.
Author contributions
E.C.Z., K.E.M.v.K., T.C.M.Z. and A.R.A. researched
data for the article. E.C.Z., K.E.M.v.K., T.C.M.Z. and
J.L.B. substantially contributed to discussion of the
content. E.C.Z., K.E.M.v.K. and J.L.B. wrote article.
Allauthors contributed to review and editing of
themanuscript before submission.
Supplementary information is linked to the online
version of the paper at www.nature.com/nrurol.
REVIEWS
© 2015 Macmillan Publishers Limited. All rights reserved
... Bladder cancer (BC) is the most common malignancy in the urinary system and the 13th most prevalent cause of cancer death worldwide [1]. Approximately 150,000 people die from bladder cancer each year globally [2]. ...
... In the 1990s, targeted molecular therapy was first proposed as a novel strategy in cancer therapy, which focuses on interfering with important cellular processes in cancer cells [5]. Many targeted agents have been investigated in BC, such as EGFR, VEGF, STAT3, and CD24, which are molecular targets that can be efficiently inhibited, resulting in a reduction of tumor growth [1,6,7]. Therefore, the understanding of BC biomarkers has the potential to fundamentally change the diagnosis and treatment of this disease with better clinical effects. ...
Article
Full-text available
Background Bladder cancer is a very common malignancy with a high recurrence rate. The survival of patients with muscle-invasive bladder cancer is poor, and new therapies are needed. Livin has been reported to be upregulated in bladder cancer and influence the proliferation of cancer cells. Materials and methods The Livin gene in human bladder cancer cell line T24 was knocked out, and the differentially expressed genes were identified by RNA-seq and qPCR. Results Livin knockdown affects gene expression and has strong negative effects on some cancer-promoting pathways. Furthermore, combined with bladder cancer clinical sample data downloaded from TCGA and GEO, 2 co-up-regulated genes and 58 co-down-regulated genes were identified and validated, which were associated with cancer proliferation and invasion. Conclusion All these results suggest that Livin plays an important role in bladder cancer and could be a potential anticancer target in clinical therapy.
... BLCA is a prevalent malignant tumor in the genitourinary system (1), and its insidious early symptoms often lead to high mortality rates in patients with advanced stages (19). The limited effectiveness of the current treatment methods, including surgical resection, radiotherapy, and chemotherapy, is mainly attributed to the high metastasis rate of BLCA (20). ...
Preprint
Full-text available
Background Bladder cancer (BLCA) is associated with poor prognosis and the identification of new prognostic markers is needed. This study aimed to investigate the significance of Prolyl 3-hydroxylase 1 (P3H1) as a potential prognostic biomarker for BLCA. Methods The expression levels of P3H1 in BLCA and normal bladder tissues were analyzed using TCGA database, and P3H1's prognostic value was determined using Kaplan-Meier survival analysis. GSEA analysis was performed to investigate its biological functions. Assessment of the tumor microenvironment was conducted using the CIBERSORT algorithm, and the results were verified through RT-PCR and western blotting. Functional analysis of P3H1 was carried out using Colony-formation, CCK-8, wound healing, transwell, and flow cytometry assays. Results The study found that P3H1 was upregulated in BLCA tissues, and elevated expression was linked to an unfavorable prognosis. GSEA analysis revealed putative involvement of P3H1 in the JAK-STAT and WNT signaling pathways. P3H1 expression was also strongly associated with the infiltration of diverse immune cells and immune checkpoints. Conclusion Finally, knocking down P3H1 expression effectively suppressed the ability of BLCA cells to proliferate and migrate, while promoting apoptosis. In conclusion, P3H1 has the potential to serve as a prognostic biomarker for BLCA.
... A recent study of RC48, an anti-HER2 antibodydrug conjugate (payload), has shown promising results in patients with ErbB2 alterations in urothelial cancer [22]. Aberrations in other members of the ERBB family (EGFR amplification [11%], ERBB2 amplification [9%], ERBB2 somatic mutations [9%] and ERBB3 somatic mutations [11%]) have also been investigated as potential drivers in the pathogenesis of urothelial cancer [18,[23][24][25][26][27]. Further, EGFR overexpression in urothelial cancer is correlated with more advanced disease [28,29], tumour recurrence [30,31] and reduced overall survival (OS) [31,32]. ...
Article
Background: Preclinical and early clinical data suggest that the irreversible ErbB family blocker afatinib may be effective in urothelial cancers harbouring ERBB mutations. Methods: This open-label, phase II, single-arm trial (LUX-Bladder 1, NCT02780687) assessed the efficacy and safety of second-line afatinib 40 mg/d in patients with metastatic urothelial carcinoma with ERBB1-3 alterations. The primary endpoint was 6-month progression-free survival rate (PFS6) (cohort A); other endpoints included ORR, PFS, OS, DCR and safety (cohorts A and B). Cohort A was planned to have two stages: stage 2 enrolment was based on observed antitumour activity. Results: Thirty-four patients were enroled into cohort A and eight into cohort B. In cohorts A/B, PFS6 was 11.8%/12.5%, ORR was 5.9%/12.5%, DCR was 50.0%/25.0%, median PFS was 9.8/7.8 weeks and median OS was 30.1/29.6 weeks. Three patients (two ERBB2-amplified [cohort A]; one EGFR-amplified [cohort B]) achieved partial responses. Stage 2 for cohort A did not proceed. All patients experienced adverse events (AEs), most commonly (any/grade 3) diarrhoea (76.2%/9.5%). Two patients (4.8%) discontinued due to AEs and one fatal AE was observed (acute coronary syndrome; not considered treatment-related). Conclusions: An exploratory biomarker analysis suggested that basal-squamous tumours and ERBB2 amplification were associated with superior response to afatinib. Clinical trial registration: NCT02780687.
... Most of patients with superficial bladder carcinoma generally have favorable outcomes. However, 1/3 of these superficial bladder carcinoma patients will progress to muscle invasion [5,6]. Yet the treatment of the muscle-invasive bladder carcinoma patients remains challenging. ...
Article
Full-text available
As a member of PHB (prohibitin1) family, PHB plays important roles in many cancers, but its property in bladder carcinoma aggressiveness is unknown. This research was to explore the function and potential mechanism of PHB in bladder carcinoma in vivo and in vitro. The invasive abilities of cancer cell were determined by transwell and wound-healing assays. The function of PHB was confirmed by gene knockdown and overexpression methods. Further in vivo confirmation was performed in a nude mouse model with lung metastasis. The relationship of PHB and β-catenin was confirmed by immunoprecipitation and immunofluorescence staining assays. The protein expression of epithelial-mescenchymal transition (EMT) and Wnt/β-catenin signaling pathway was tested by immunofluorescence staining and western blotting assay. The depletion of PHB prevented bladder cancer cell invasiveness and inhibited EMT. Contrarily,the abilities of bladder carcinoma cells migration and invasion in vitro as well as metastasis in vivo were enhanced when the PHB overexpressed unnormally. Importantly, the β-catenin was identified to be bound by PHB and β-catenin knockdown reduced the cancer cell migration, invasion and EMT in PHB overexpressing cells. In addition, PHB stabilized β-catenin by inhibiting its ubiqutin-mediated degradation thus leading to increased Wnt/β-catenin signaling. These observations indicate that PHB could promote bladder cancer aggressiveness by binding with β-catenin to prevent the degradation of β-catenin and the localized invasive bladder cancer patients with PHB overexpression should take more aggressive postsurgical adjuvant anticancer therapies.
... Most studies of exosomes in urine have focused on urogenital malignancies, primarily because of their proximity to tissue, which increases the likelihood of isolating exosomes secreted by tumor cells. The search for protein, miRNA, and mRNA biomarkers specific for bladder, prostate, and kidney cancers will be easier with the ability to obtain larger amounts of exosomes from urine samples [122]. The potential application of exosomes in the clinic is an emerging area of research. ...
Chapter
Exosomes are tiny membrane-enveloped vesicles of endosomal origin, typically 40–120nm in diameter, produced by most cells in both normal and pathological situations. These exosomes can be isolated from all biofluids, including urine. In this context, many researchers have focused on the analysis of urinary exosomes because urine can be collected in large quantities, regularly, and with minimal effort. Exosomes contain phospholipids, cholesterol, proteins, glycoconjugates, nucleic acids, and metabolites. Because all organs and tissues produce exosomes, their molecular cargo can provide first-hand information about the physiological and biological state of the site of origin. Many potential disease biomarker candidates have already been identified in urinary exosomes. In this chapter, we performed a bibliometric analysis of the keywords “exosome(s)” and “urine” to identify related terms, diseases and molecular/biological processes, and other related terms. This yielded interesting results suggesting that exosomes in urine may play a role in the pathogenesis of various diseases. Moreover, this chapter discusses exosomes isolation and characterization methodologies and highlight the importance of urinary exosomes and their role in the diagnosis, prognosis and therapy of various diseases. We offer a bibliometric approach and an in-depth analysis on several exosomes' isolation techniques, diagnostic potential for urogenital and specific non-urogenital diseases, as well as an overview of miRNAs significance on urinary exosomes, conferring a more complete status to this review, something that was still lacking in the current literature.
... RTKs control vital biological functions such as proliferation, apoptosis, differentiation, and metabolism. RTKs are frequently overexpressed in BC, resulting in downstream signalling pathways [25]. Previous publications indicate that multiple RTKs families have been found in various cancers, for instance, VEGFR-2 in BC [26] and TACSTD2 in breast cancer [27], non-small-cell lung cancer [28], and thyroid cancer [29]. ...
Article
Full-text available
Bladder cancer (BC) recurrence is one of the primary clinical problems encountered by patients following chemotherapy. However, the mechanisms underlying their resistance to chemotherapy remain unclear. Alteration in the pattern of membrane proteins (MPs) is thought to be associated with this recurrence outcome, often leading to cell dysfunction. Since MPs are found throughout the cell membrane, they have become the focus of attention for cancer diagnosis and treatment. Identifying specific and sensitive biomarkers for BC, therefore, requires a major collaborative effort. This review describes studies on membrane proteins as potential biomarkers to facilitate personalised medicine. It aims to introduce and discuss the types and significant functions of membrane proteins as potential biomarkers for future medicine. Other types of biomarkers such as DNA-, RNA- or metabolite-based biomarkers are not included in this review, but the focus is mainly on cell membrane surface protein-based biomarkers.
Article
Full-text available
Simple Summary Our study used NanoString technology, a high-throughput platform measuring gene expression at the mRNA level to identify a set of genes predictive of clinical outcomes in bladder cancer patients. Twenty-seven differentially expressed genes were correlated with clinicopathological variables including molecular subtypes (luminal, basal, null/double-negative), histological subtypes (conventional urothelial carcinoma or carcinoma with variant histology), clinical subtype (NMIBC and MIBC), tumor stage category (Ta, T1 and T2-4), tumor grade, PD-L1 expression (high vs. low expression), and clinical risk categories (low, intermediate, high, and very high). Then, two risk models integrating the molecular subtypes and the level of expression of TP53, CCND1 and MKI67 were developed. These models provided a score ranging from 0 (best prognosis) to 7 (worst prognosis) that could be used to predict patient’ outcome and guide treatment decisions in bladder cancer. Abstract This study evaluated a panel including the molecular taxonomy subtype and the expression of 27 genes as a diagnostic tool to stratify bladder cancer patients at risk of aggressive behavior, using a well-characterized series of non-muscle invasive bladder cancer (NMIBC) as well as muscle-invasive bladder cancer (MIBC). The study was conducted using the novel NanoString nCounter gene expression analysis. This technology allowed us to identify the molecular subtype and to analyze the gene expression of 27 bladder-cancer-related genes selected through a recent literature search. The differential gene expression was correlated with clinicopathological variables, such as the molecular subtypes (luminal, basal, null/double negative), histological subtype (conventional urothelial carcinoma, or carcinoma with variant histology), clinical subtype (NMIBC and MIBC), tumor stage category (Ta, T1, and T2–4), tumor grade, PD-L1 expression (high vs. low expression), and clinical risk categories (low, intermediate, high and very high). The multivariate analysis of the 19 genes significant for cancer-specific survival in our cohort study series identified TP53 (p = 0.0001), CCND1 (p = 0.0001), MKI67 (p < 0.0001), and molecular subtype (p = 0.005) as independent predictors. A scoring system based on the molecular subtype and the gene expression signature of TP53, CCND1, or MKI67 was used for risk assessment. A score ranging from 0 (best prognosis) to 7 (worst prognosis) was obtained and used to stratify our patients into two (low [score 0–2] vs. high [score 3–7], model A) or three (low [score 0–2] vs. intermediate [score 3–4] vs. high [score 5–7], model B) risk categories with different survival characteristics. Mean cancer-specific survival was longer (122 + 2.7 months) in low-risk than intermediate-risk (79.4 + 9.4 months) or high-risk (6.2 + 0.9 months) categories (p < 0.0001; model A); and was longer (122 + 2.7 months) in low-risk than high-risk (58 + 8.3 months) (p < 0.0001; model B). In conclusion, the molecular risk assessment model, as reported here, might be used better to select the appropriate management for patients with bladder cancer.
Preprint
Full-text available
Background Bladder cancer is a very common malignancy with a high recurrence rate. The survival of patients with muscle-invasive bladder cancer is poor, and new therapies are needed. Livin has been reported to be upregulated in bladder cancer and influence the proliferation of cancer cells. Materials & Methods The Livin gene in human bladder cancer cell line T24 was knocked out, and the differentially expressed genes were identified by RNA-seq and qPCR. Results Livin knockdown affects gene expression and has strong negative effects on some cancer-promoting pathways. Furthermore, combined with bladder cancer clinical sample data downloaded from TCGA and GEO, 2 co-up-regulated genes and 28 co-down-regulated genes were identified and validated, which were associated with cancer proliferation and invasion. Conclusion All these results suggest that Livin plays an important role in bladder cancer and could be a potential anticancer target in clinical therapy.
Article
Full-text available
Photodynamic therapy (PDT) is a promising treatment option for only cancer. It can be used to treat drugresistant and inoperable tumors. However, one of the unsolved problems in PDT is tumor recurrence. The aim of this study is to suppress tumor regrowth by including angiogenesis inhibitors that target vascular endothelial growth factor (VEGF) and epidermal growth factor receptor (EGFR) pathways, to the PDT protocol. Human bladder cancer (MGH cell line) xenografts were induced in Balb/c nude mice. The animals treated with PDT received an intravenous injection of hypericin followed by irradiation after 6 hours, with a broadband light source through a 560-640 NM band-pass filter. A light dosage of 120 J/cm2 and 100 mW/cm2 was administered. Angiogenesis inhibitors, bevacizumab and cetuximab were administered at a dose of 10 mg/kg. Treatment efficacy was assessed by monitoring the tumor growth inhibition of xenograft tumors. Immunohistochemistry was performed to evaluate the expression of VEGF and EGFR. Expression of the major proteins in the VEGF and EGFR pathways was investigated by immunoblotting. One-way ANOVA with Bonferroni correction was performed to analyze the data. Tumors treated with the combination of PDT and inhibitors exhibited significantly greater treatment response compared to control and PDT groups. Downregulation of VEGF and EGFR was observed in tumors treated with PDT + bevacizumab and cetuximab. Our results show that blocking VEGF or/and EGFR pathways along with PDT can effectively suppress tumor growth.
Article
Full-text available
Optimal animal models of muscle invasive bladder cancer (MIBC) are necessary to overcome the current lack of novel targeted therapies for this malignancy. Here we report on the establishment and characterization of patient-derived primary xenografts (PDX). Patient tumors were grafted under the renal capsule of mice and subsequently transplanted over multiple generations. Patient tumor and PDX were processed for analysis of copy number variations by aCGH, gene expression by microarray, and expression of target pathways by immunohistochemistry (IHC). One PDX harbouring an FGFR3 mutation was treated with an inhibitory monoclonal antibody targeting FGFR3. Five PDX were successfully established. Tumor doubling time ranged from 5 to 11 days. Array CGH revealed shared chromosomal aberrations in the patient tumors and PDX. Gene expression microarray and IHC confirmed that PDXs maintain similar patterns to the parental tumors. Tumor growth in the PDX with an FGFR3 mutation was inhibited by the FGFR3 inhibitor. PDXs recapitulate the tumor biology of the patients' primary tumors from which they are derived. Investigations related to tumor biology and drug testing in these models are therefore more likely to be relevant to the disease state in patients. They represent a valuable tool for developing precision therapy in MIBC.
Article
The ultimate goal of prognostic assessment is optimization of individual counseling. Often, however, studies on prognostic factors focus on discriminating between high risk and low risk subgroups without considering the relevance of 1 or more factors for predicting disease outcome in individual patients. We quantified the accuracy of prediction of future recurrences and disease progression in individual patients with primary superficial bladder cancer. The study cohort consisted of 1,674 patients who were followed prospectively between 1983 and 1991 in the Netherlands. By analyzing half of the patients with proportional hazards regression, we computed relative risks of recurrence and progression. A prognostic index score based on these relative risks was then applied to the other half of the patients to determine whether group outcome could be predicted accurately. To assess the accuracy of prediction in individuals we used a method similar to the construction of receiver operating characteristic curves in diagnostic test assessment.
Article
276 Background: Cisplatin-based neoadjuvant chemotherapy has shown survival advantage in patients (pts) with muscle-invasive bladder cancer. This benefit is strongly associated with downstaging of tumor to pT0. It was hypothesized that the addition of antiangiogenic therapy bevacizumab (Avastin [A]) to cisplatin (C) and gemcitabine (G) chemotherapy may offer improved outcomes in these pts. Methods: Patients with clinically localized, muscle-invasive, chemotherapy-naive TCC of bladder (T2-T4a), ECOG PS 0-1 received cisplatin 70 mg/m ² IV day 1, gemcitabine 1000 mg/m ² IV day 1 and 8, bevacizumab 15 mg/kg IV day 1 on a 21- day cycle for up to 4 cycles. After completion of neoadjuvant therapy all pts underwent RC. Those pts not achieving pT0 status received paclitaxel 175 mg/m2 IV day 1 and bevacizumab 15 mg/m ² IV day 1 on a 21-day cycle for up to 3 cycles. The primary objective of the trial was to achieve pT0 status. Results: By October 2010, 15 pts were enrolled, and 13 are evaluable for response. Median age was 64 (range 47-71), 9 (70%) males, 4 (30%) females. 2/13 (15%) developed PD after neo-adjuvant therapy and were taken off protocol.11/13 (85%) underwent RC. 4/13 (31%) pts were downstaged, 3/11 (27%) achieved pTis status and none achieved pT0 status. 2/11 (18%) pts received adjuvant paclitaxel chemotherapy with bevacizumab. Dose reduction or delays were noted in 12/15 (80%) pts. Grade 3 or 4 hematologic toxicity was observed in 5/15 (33%) pts. Grade 3 or 4 nonhematologic toxicity was observed in 3/15 (20%) pts: (renal dysfunction 2 pts, pulmonary embolism 1 pt). Postoperative complications were seen in 5/12 (42%) pts: enterovesicular fistula (1), delayed wound healing (1), prolonged ileus (2), and pelvic abscess (1). With a median follow-up time of 12 months (range 3-45), 8/11 (72%) pts have remained disease free. Conclusions: Neoadjuvant CGA is feasible in pts with clinically localized, muscle-invasive TCC of the bladder. The surgical complications rate appears rather high and close monitoring is needed. Accrual to the trial continuous and updated results will be presented. No significant financial relationships to disclose.
Article
Context: The European Association of Urology (EAU) panel on Non-muscle-invasive Bladder Cancer (NMIBC) released an updated version of the guidelines on Non-muscle-invasive Bladder Cancer. Objective: To present the 2016 EAU guidelines on NMIBC. Evidence acquisition: A broad and comprehensive scoping exercise covering all areas of the NMIBC guidelines published between April 1, 2014, and May 31, 2015, was performed. Databases covered by the search included Medline, Embase, and the Cochrane Libraries. Previous guidelines were updated, and levels of evidence and grades of recommendation were assigned. Evidence synthesis: Tumours staged as TaT1 or carcinoma in situ (CIS) are grouped as NMIBC. Diagnosis depends on cystoscopy and histologic evaluation of the tissue obtained by transurethral resection of the bladder (TURB) in papillary tumours or by multiple bladder biopsies in CIS. In papillary lesions, a complete TURB is essential for the patient's prognosis. If the initial resection is incomplete, there is no muscle in the specimen, or a high-grade or T1 tumour is detected, a second TURB should be performed within 2-6 wk. The risks of both recurrence and progression may be estimated for individual patients using the European Organisation for Research and Treatment of Cancer (EORTC) scoring system and risk tables. The stratification of patients into low-, intermediate-, and high-risk groups is pivotal to recommending adjuvant treatment. For patients with a low-risk tumour and intermediate-risk patients at a lower risk of recurrence, one immediate instillation of chemotherapy is recommended. Patients with an intermediate-risk tumour should receive 1 yr of full-dose bacillus Calmette-Guérin (BCG) intravesical immunotherapy or instillations of chemotherapy for a maximum of 1 yr. In patients with high-risk tumours, full-dose intravesical BCG for 1-3 yr is indicated. In patients at highest risk of tumour progression, immediate radical cystectomy (RC) should be considered. RC is recommended in BCG-refractory tumours. The long version of the guidelines is available at the EAU Web site (www.uroweb.org/guidelines). Conclusions: These abridged EAU guidelines present updated information on the diagnosis and treatment of NMIBC for incorporation into clinical practice. Patient summary: The European Association of Urology has released updated guidelines on Non-muscle-invasive Bladder Cancer (NMIBC). Stratification of patients into low-, intermediate-, and high-risk groups is essential for decisions about adjuvant intravesical instillations. Risk tables can be used to estimate risks of recurrence and progression. Radical cystectomy should be considered only in case of failure of instillations or in NMIBC with the highest risk of progression.
Article
261 Background: DD-MVAC has shown an improved 5-year survival in metastatic urothelial cancer; however, there is no prospective data on its use in the neoadjuvant setting. Previous work suggested that over-expression of VEGF was associated with a high risk of relapse in our neoadjuvant patients, leading to the hypothesis that combining a VEGFR inhibitor with chemotherapy may improve patient outcomes. Methods: Between 8/07 and 12/10, 60 patients with urothelial carcinoma of the bladder or upper tract tumor were enrolled on a prospective phase II clinical trial. Eligibility requirements included at least one of the following: 3-D mass on EUA (cT3b disease), LVI, hydronephrosis, micropapillary features, tumor in a diverticula, and for upper tract tumors a high grade tumor or radiographically measurable sessile mass. The primary endpoint was pathologic down-staging to <=pT1N0M0. Results: Forty-four patients with bladder/urethral tumors and 16 patients with upper tract tumors were enrolled. Pathologic down-staging to <= pT1N0M0 occurred in 53% of patients overall (bladder/urethra 45%, upper tract 75%), and to <= pT0N0M0 in 38% overall (bladder/urethra 39%, upper tract 38%). At a median follow-up of 21 months, the 2-year OS and DSS was 78% and 85%, respectively (bladder 2-yr OS and DSS 75%, 82%; upper tract 93%, 93%). The median OS and DSS have not yet been reached. The most common grade 3 or greater toxicity was neutropenia in 27% of patients, followed by fatigue in 10%. The following grade 3 toxicities were observed in < 10% of patients: mucositis, DVT/PE, hypertension, nausea/vomiting, thrombocytopenia. One patient experienced cardiac ischemia. Conclusions: Neoadjuvant chemotherapy leads to pathologic down-staging in 45% of patients with bladder cancer. DD-MVAC appears an acceptable alternative to traditional M-VAC in the neoadjuvant setting. Although bevacizumab did not impact down-staging based upon historical expectations, determining the effect on recurrence requires longer follow-up.
Article
Viruses have been suggested to be useful as anti-cancer agents since the early 20th century, although following the advent of chemotherapy and radiotherapy work largely stopped until the 1990s when a number of groups began to explore the use of engineered viruses. This overview summarizes the development of the field from the 1990s to the present day, an era when oncolytic viruses have now demonstrated clear clinical benefit to patients. The hurdles and challenges which needed to be overcome are discussed, and in particular the importance of the immune component in achieving a therapeutic effect is highlighted. Today, oncolytic therapy is generally thought of as an immunotherapy, the term 'oncolytic immunotherapy' having been widely adopted. With the advent of immuno-oncology drugs based on immune checkpoint blockade, a clear rationale for synergy between the two approaches, and initial pre-clinical and clinical data suggesting this to be the case, it might be expected that oncolytic immunotherapy combined with checkpoint blockade will provide a cornerstone of future cancer treatment. Copyright © 2015. Published by Elsevier B.V.