ArticlePDF Available

Potential molecular, cellular and microenvironmental mechanism of Sorafenib resistance in hepatocellular carcinoma

Authors:
  • Department of General Surgery

Abstract and Figures

Sorafenib, an orally-available kinase inhibitor, is the only standard clinical treatment against advanced hepatocellular carcinoma. However, development of resistance to sorafenib has raised concern in recent years due to the high-level heterogeneity of individual response to sorafenib treatment. The resistance mechanism underlying the impaired sensitivity to sorafenib is still elusive though some researchers have made great efforts. Here, we provide a systemic insight into the potential molecular, cellular and microenvironmental mechanism of sorafenib resistance in hepatocellular carcinoma depending on abundant previous studies and reports. Copyright © 2015. Published by Elsevier Ireland Ltd.
Content may be subject to copyright.
Mini-review
Potential molecular, cellular and microenvironmental mechanism of
sorafenib resistance in hepatocellular carcinoma
Jiang Chen, Renan Jin, Jie Zhao, Jinghua Liu, Hanning Ying, Han Yan, Senjun Zhou,
Yuelong Liang, Diyu Huang, Xiao Liang, Hong Yu, Hui Lin *, Xiujun Cai **
Department of General Surgery, Sir Run Run Shaw Hospital of Zhejiang University, Hangzhou, Zhejiang, China
ARTICLE INFO
Article history:
Received 4 March 2015
Received in revised form 23 June 2015
Accepted 25 June 2015
Keywords:
Advanced hepatocellular carcinoma (HCC)
Epithelial–mesenchymal transitions (EMT)
and mesenchymal–epithelial transitions
(MET)
Cancer stem cells (CSCs)
Microenvironment
Acquired resistance
ABSTRACT
Sorafenib, an orally-available kinase inhibitor, is the only standard clinical treatment against advanced
hepatocellular carcinoma. However, development of resistance to sorafenib has raised concern in recent
years due to the high-level heterogeneity of individual response to sorafenib treatment. The resistance
mechanism underlying the impaired sensitivity to sorafenib is still elusive though some researchers have
made great efforts. Here, we provide a systemic insight into the potential molecular, cellular and mi-
croenvironmental mechanism of sorafenib resistance in hepatocellular carcinoma depending on abundant
previous studies and reports.
© 2015 Elsevier Ireland Ltd. All rights reserved.
Introduction
Hepatocellular carcinoma (HCC) is one of the most common
primary malignant tumors and the largest cancer-related deaths
ranking only second to lung cancer, with a fairly high and increas-
ing incidence, frequently relapse and dismal prognosis [1,2]. Despite
the remarkable progress in the prevention, detection, and treat-
ment of cancer over the last five decades, no adequate therapy
remains sufficiently effective due to late stage diagnosis and inad-
equate clinical strategies for inhibiting metastasis and promoting
apoptosis [3–6]. Furthermore, multi-drug resistance of cancer cells
has been implicated as a major challenge considering the irreplace-
able role of chemotherapeutic interventions in anti-cancer treatment
[7]. The resistance was postulated to associate with elevated ex-
pression of drug efflux transporters, changes in drug kinetics,
amplification of drug targets or tumor heterogeneity comprising of
genetic variation, the microenvironment, and cell plasticity [8].For
patients with HCC diagnosed at advanced stage, sorafenib is the only
choice of systemic therapy when potentially curative treatment, such
as resection and liver transplantation, may be merely applicable for
patients diagnosed at early stage [9]. Recently, the unstable effica-
cy of sorafenib has raised concern of more and more researchers
and ‘sorafenib resistance’ has become a hot term used to describe
the impaired efficacy of sorafenib, especially for patients with ad-
vanced HCC.
Sorafenib, as a multikinase inhibitor, suppresses tumor angio-
genesis and proliferation by inhibiting serine/threonine kinases, as
well as receptor tyrosine kinases. Intracellular Raf serine/threonine
kinase isoforms inhibited by sorafenib include Raf-1 (or C-Raf), wild-
type B-Raf and mutant B-Raf. Receptor tyrosine kinases inhibited
by sorafenib include vascular endothelial growth factor receptor
(VEGFR)-1, VEGFR-2, VEGFR-3, platelet-derived growth factor re-
ceptor (PDGFR)-b, c-KIT, FMS-like tyrosine kinase 3 (FLT-3) and RET
[10]. Among these kinases, RAF and VEGFR are presumed to be es-
sential for the anti-proliferative effects evoked by sorafenib. RAF
kinases are present at the level of the cancer cells, while the VEGFR
is present on the surface of endothelial cells [11]. Inhibition of the
VEGFR accounts, at least in part, for the antiangiogenic effects of
sorafenib. In majority of patients with advanced HCC, the RAF–
MEK–ERK cascade is often activated by autocrine and paracrine loops,
involving for example the production of amphiregulin, an agonist
of the Epidermal Growth Factor Receptor (EGFR) [12]. The analysis
of vitro experiments reported that there was a strong correlation
between the inhibition of the RAF–MEK–ERK cascade and the anti-
clonogenesis effect of sorafenib [13]. In addition, cytotoxic effect of
sorafenib is also a crucial role in antitumor treatment. Generally,
apoptosis is the major form of cytotoxicity and it is required for tumor
regression and sustained clinical remissions [14]. The pro-apoptotic
effect of sorafenib in HCC cells had also been studied extensively. In
2008, the SHARP (Sorafenib HCC Assessment Randomised Protocol)
* Corresponding author. Tel.: +8613958084556; fax: +86 057186044817.
E-mail address: 369369@zju.edu.cn (H. Lin).
** Corresponding author. Tel.: +8613805738266; fax: +86 057186006605.
E-mail address: cxjzu@hotmail.com (X. Cai).
http://dx.doi.org/10.1016/j.canlet.2015.06.019
0304-3835/© 2015 Elsevier Ireland Ltd. All rights reserved.
Cancer Letters 367 (2015) 1–11
Contents lists available at ScienceDirect
Cancer Letters
journal homepage: www.elsevier.com/locate/canlet
trials showed an improved overall survival in Child–Pugh class A
patients with advanced HCC upon treatment with the antiangiogenic
and antiproliferative agent sorafenib. In addition, Oriental con-
firmed the efficacy and safety of sorafenib again in treatment of
advanced hepatocellular carcinoma by another multicenter, ran-
domized, double-blind, placebo-controlled phase III clinical trials.
However, the promising systemic chemotherapeutic agent
sorafenib only demonstrated relatively limited benefits rather than
eradicated the microscopic residual and cured the patients with ad-
vanced HCC. Sorafenib is beneficial in only around 30% of patients,
and acquired resistance often develops within 6 months [15,16], sug-
gesting the existence of primary and acquired sorafenib resistance
in hepatocellular carcinoma cells. So far, some researchers have in-
vestigated the mechanism underlying resistance to sorafenib from
molecular and clinical points of view. Huang et al. have reported
that overexpression of both αB-Crystallin and 14-3-3ζ decreased
hepatoma cells sensitization to sorafenib in clinical by inducing
epithelial–mesenchymal transition (EMT) in HCC cells [17]. Chiou
et al. presented evidence that glucose-regulated protein 78 (GRP78)
is a positive modifier for sorafenib resistance acquisition in HCC and
GRP78 knockdown enhanced the efficacy of sorafenib-mediated cell
death [18]. Chen et al. also conducted experiments to investigate
sorafenib resistance and the results showed that long-term expo-
sure to sorafenib activated the phosphatidylinositol 3-kinase
(PI3K)/Akt signaling pathway and mediate acquired resistance to
sorafenib in HCC cell lines [19]. All the insights into the molecular
and clinical changes involved in sorafenib treatment provided a
greater understanding of the underlying mechanism of primary and
acquired resistance to sorafenib. However, a systemic and compre-
hensive analysis about the mechanism of acquiring resistance to
sorafenib was still elusive [20]. Here, in order to clarify the drug re-
sistance mechanism clearly, we reviewed studies on sorafenib
resistance in liver cancer from different angles, including molecu-
lar markers, signaling pathways, drug resistance principles, regulation
systems, therapeutic implications and recent approaches. Increas-
ing evidence suggested that deeper insight into the formation of
sorafenib resistance would shed light on the specific drug resis-
tance mechanism and might lead to identification of potential clinical
biomarkers for prognosis evaluation and targets for new therapeu-
tic strategies. Therefore, in the present review, by synthesizing the
retrospective studies and reports, we provided potential molecu-
lar, cellular and microenvironmental mechanism underlying impaired
sensitivity to sorafenib in hepatocellular carcinoma, which had pro-
found effect on inhibiting tumor progression, evaluating patient
survival and predicting sorafenib treatment response.
Epithelial–mesenchymal transitions (EMT)
Epithelial–mesenchymal transitions (EMT) are mainly marked
by the loss of cell–cell interactions and of epithelial apico-basal po-
larity with the concomitant acquisition of mesenchymal markers
and enhanced migratory behavior [21,22]. As is well known, EMT
is of paramount relevance for various developmental processes, in-
cluding gastrulation, neural crest formation and carcinoma
progression. In the last decades, EMT has been extensively studied
and validated in the progression of various carcinomas such as HCC
[23,24]. A large body of researches has claimed the important role
of EMT in facilitating tumor development and progression by driving
metastasis, through the acquisition of enhanced migratory and in-
vasive potential [25–30]. Metastasis, induced by EMT, deconvoluted
into several steps including intravasation, circulation, margin-
ation, extravasation and colonization, is a crucial cause of deaths
in cancer therapy [31,32]. It had been reported that intrahepatic me-
tastases were observed in about 30% of cases after surgical removal
of small HCC nodules and in 80% of HCC autopsy cases [33]. Via EMT,
epithelial cells dissolve intercellular connections and acquire
mesenchymal properties and metastasize to the native or distant
sites. Once cancer cells seed the metastatic site, a mesenchymal to
epithelial transition (MET) occurs, inducing colonization and growth
of the metastatic foci with re-expression of cell adhesion mole-
cule E-cadherin, facilitating tumor cells to seed in the metastatic
sites [34–40]. E-cadherin re-expression accompanied by a partial
MET in the metastatic sites increases post-extravasation survival of
the cancer cells and resistance to multi-drugs [35,37]. Additional-
ly, further study found that a variety of biological molecules such
as HIF, TGF and miRNAs involved in the process of MET through dif-
ferent signaling pathways respectively accelerated the formation of
distant tumor metastasis [41–43]. In epithelial cancers, EMT re-
sulted in the loss of E-cadherin and in turn, tumor cells attain
enhanced migratory and invasive potential. E-cadherin, whose down-
regulated expression is a main biomarker for activation of EMT,
shows reverse relationship with drug resistance [44]. By undergo-
ing EMT, the tumor cells acquired resistance to a number of chemo
and radiotherapies [34,45–48]. The relationship between epithelial–
mesenchymal transition (EMT) and drug resistance was first
described in connection with cancer stem cells by Mani et al., who
inferred that blocking or reversing EMT might cause chemoresistant
cells to revert to chemosensitive cells [49]. In addition, in breast
cancer therapy, it was reported that epithelial cells acquire some
cancer stem cells properties via EMT, such as anti-apoptosis and drug
resistance [50]. In the process of EMT in tumor cells, some biolog-
ical molecules, such as TGF β, Slug and FOXC2, play very crucial role
in explaining the sensitivity of tumor cells to chemotherapy drugs.
Activation of TGF β or FOXC2 and overexpression of Slug impaired
the sensitivity of tumor tissues to drug [49,51,52]. The drug resis-
tance was also reported in correlation with EMT in HCC [17–19].
Various signaling pathways have been involved in regulation of EMT
program [27], indicating that these signaling pathways and mol-
ecules involved are potentially associated with drug resistance. The
Raf/MEK/ERK pathway represents a dominant signaling network pro-
moting proliferation and metastasis and is the main jamming target
pathway of sorafenib [53–56]. The genetic and molecular network
involved in the signal pathway is very complicated. In fact, some
studies have demonstrated that treatment with Raf kinase inhibi-
tors can even paradoxically induce ERK cascade signaling by
promoting dimerization of Raf family members [57,58]. The
biomarkers for sorafenib efficacy, such as Plasma c-KIT, hepato-
cyte growth factor and angiopoietin-2, are likely to be downregulated
in the sorafenib sensitive cell lines while the phospho-Akt and p85
(a regulatory subunit of PI3K) are upregulated in sorafenib resis-
tant cell lines [19,59,60]. Kunnimalaiyaan et al. hypothesized that
Notch signaling pathway is a potential regulator of EMT program
in HCC and NICD3 protein increased E-cadherin and activated tran-
scription factors Snail, Slug, and MMPs and induced EMT [27], which
might lead to development of sorafenib resistance. Acquired resis-
tance to sorafenib in HCC might also be mediated by the activation
of phosphatidylinositol 3-Kinase/Akt signaling pathway [19,61].
Overexpression of both αB-Crystallin (Cryab) and 14-3-3ζ can induce
epithelial–mesenchymal transition (EMT) in HCC cells through ac-
tivation of the extracellular-regulated protein kinase (ERK) cascade,
a crucial up-stream factor of ERK1/2/Fra-1/slug signaling pathway,
which can counteract the effect of sorafenib [62–67]. Therefore, it
is postulated that sorafenib resistance is closely correlated with the
dynamic changes of numerous molecules, especially their up- or
down-regulated expression, involved in the mutual transition
between epithelial and mesenchymal cell phenotypes and various
signaling pathways. Particularly, epithelial cells are more suscep-
tible to a Raf kinase inhibitor, sorafenib, whereas mesenchymal cells
showed significant resistance [68,69]. In addition, PI3K/Akt showed
higher activity and integrin-linked kinase (ILK) exhibit in the mes-
enchymal cells as compared to epithelial cells [70,71]. Many
pathways and molecules involved in these pathways play a crucial
2J. Chen et al./Cancer Letters 367 (2015) 1–11
role in regulation, induction and reversion of EMT, including Wnt,
TGFβ, RTK and other signaling pathways [72–81]. The detailed in-
formation was depicted in Tabl e 1. Therefore, EMT, MET and
concomitant molecules and cell phenotypes are important deter-
minants for the resistance of HCC against sorafenib (Fig. 1).
Cancer stem cells (CSCs)
Cancer stem cells (CSCs) were first identified in the hematopoi-
etic system in 1990s when the cancer stem cells model was
established [85,92]. CSCs are defined as cells within a tumor that
Table 1
Pathways and molecules involved in EMT.
Pathways Molecules
involved
Role Sorafenib
efficacy
References
Wnt Frizzled Promotes tumor cell proliferation and metastasis; induces EMT; increases drug resistance; participates in
crosstalk with other pathways; alters the microenvironment
[38,62,71,82]
GSK-3β
β-catenin
TCF
TGFβ Smad Regulates cell proliferation, differentiation and apoptosis; induces EMT; promotes tumor invasion and
metastasis
[21,26,83]
TIP30
RTK-Ras Raf Regulates cell signaling and the tumor microenvironment; drives tumor development and progression [27,29,42]
MEK
ERK
RTK-PI3K AKT Regulates EMT; promotes motility and metastasis [50,72,84]
mTOR
Hedgehog Gli Regulates cancer cell proliferation and metastasis; induces EMT; increases stromal fibrosis [57,85,86]
PTCH1
Snail/Slug
TNFα IκB Induces EMT; promotes metastasis and invasion; increased Twist expression [28,49,87]
NF-κB
IL-1β IRAK Promotes tumor cell migration and invasion; contributes to higher angiogenesis, tumor growth, tumor
progression; increases inflammation; induces EMT
[28,59,64]
TRAF
NF-κB
IL-6 HSP27 Promotes proliferation; regulates EMT; enhances cancer metastasis; increases inflammation [54,88,89]
STAT3
Notch Snail Upregulates Snail; stimulates Slug; induces EMT; represses E-cadherin expression; promotes tumor cell
migration and invasion; increases chemoresistance
[17,90,91]
Slug
Jagged1
Note: –, specific molecules negatively correlated with Sorafenib efficacy.
Fig. 1. Role of EMT, MET and the molecules involved in the two processes in the development of resistance of HCC to sorafenib. During epithelial–mesenchymal transi-
tions, the expression of E-cadherin and cytokeratins, the cell–cell adhesion molecules, was down-regulated and the levels of Snail, Twist and Zeb, the transcriptional repressors
of E-cadherin, were upregulated. Epithelial phenotype was transformed to mesenchymal phenotype. Epithelial cells are more sensitive to sorafenib treatment while mes-
enchymal cells are resistant to sorafenib treatment. MET, the reverse process of EMT, was accompanied by re-expression of E-cadherin and cytokeratins and down-
regulation of Snail, Twist and Zeb. Metastatic tumor cells could seed in the distant or local sites with higher adhesion via MET. In addition, our team has hypothesized that
the tumor cells acquired sorafenib resistance in a way of an upward spiral through EMT and MTE circulation and the evidence has been collected using the experiments
and detailed data but yet not official report.
3J. Chen et al./Cancer Letters 367 (2015) 1–11
possess the capacity to self-renew and to cause the heteroge-
neous lineages of cancer cells that comprise the tumor [93,94].
According to cancer stem cell hypothesis, in tumors organized into
a hierarchy of heterogeneous cell populations, CSCs are only a minor
part but real driving force behind tumor growth while the remain-
ing cells, though making up the majority of the cells in the tumor,
lack tumorigenic potential [95]. Moreover, CSCs are able to resist
chemotherapy and spring back months or years which may explain
the tragic relapses after surgical or chemotherapeutic treatment [96].
Chemo and radiotherapies can only shrink bulk tumors with quick
efficacy but fail to kill the tiny subset of cells, CSCs, that seed and
resupply the main tumor directly. Perhaps, the expression of drug
efflux pumps and DNA-repair mechanisms are the two main reasons
of resistance of CSCs to various forms of therapy. The preferential
activation of the DNA damage checkpoint response and an in-
crease in DNA-repair capacity have been verified to be associated
with resistance to radiotherapy of brain cancer stem cells [97].Fur-
thermore, virulent multi drug resistant (MDR) phenotypes were
established in the hierarchy of heterogeneous cell populations in
HCC, which may explain the therapeutic refractoriness and dormant
behavior displayed by CSCs [90,93]. Apart from displaying classi-
cal physiological abnormalities and aberrant blood flow behavior,
MDR cancers exhibit several distinctive features such as higher
apoptotic threshold, aerobic glycolysis, regions of hypoxia, and el-
evated activity of drug-efflux transporters, which may function as
safeguards for CSCs from radio- or chemotherapy [90]. Distinc-
tively, Dick et al. suggested that the insensitivity of CSCs to most
treatment such as chemotherapy and radiation perhaps corre-
lated with the relatively more slow growth speed of CSCs as
compared to other rapidly dividing malignant cells [96].
Cancer stem cell (CSC)-like cells could be acquired via EMT with
increased CSC markers on the cell surface. They share common prop-
erties with CSCs, self-renewal ability and drug resistance [98].
Recently, the correlation between EMT and the adoption of cancer
stemness has been extensively studied as a hot research topic. From
the cellular point of view, the involvement of EMT can also be used
to clarify the mechanism under the acquired resistance to sorafenib
[99]. Via EMT, the epithelial cells overexpressed mesenchymal
biomarkers and CSC markers and silenced CSC gain migratory ability
[100]. Moreover, various signaling pathways are bridged between
EMT and CSCs, including Notch, Hedgehog, Wnt, NF-κB and epi-
genetic molecular changes [101,102], which exist in potential
molecular and cellular mechanism underlying the sorafenib resis-
tance but still unclear until now.
In HCC, liver cancer stem cells (LCSCs) display strong prolifer-
ation ability, multi-directional differentiation capacity and high drug
resistance properties, although they merely account for minor pro-
portion in liver cancer tissues [83,103,104]. For advanced HCC treated
with sorafenib, in LRCCs, a novel subpopulation of LCSCs, CSC markers
such as aldehyde dehydrogenase 1 family, wingless-type MMTV-
integration-site family, cell survival and proliferation genes were
observed upregulated, and tumor suppression markers such as apop-
tosis, cell cycle arrest, cell adhesion and stem cells differentiation
genes downregulated, concomitant with non-uniform activation of
specific isoforms of the sorafenib target proteins extracellular-
signal-regulated kinases and v-akt-murine-thymoma-viral-oncogene
homolog (AKT) [105]. It suggested that acquired resistance of HCC
to sorafenib might be explained by LCSCs, which cannot be com-
pletely killed by sorafenib but profile and differentiate into new
tumor tissues, leading to the metastasis and recurrence of HCC.
Therefore, similar to hepatocytes or even progenitor cells of the liver,
LCSCs can be regarded as the critical subset in the pathogenesis of
HCC [70] and provide a new thought line to understand the mech-
anism underlying sorafenib resistance (Fig. 2).
By understanding the cellular mechanisms underlying drug re-
sistance, especially for CSCs, it may help to identify CSC-specific
targets and develop CSC-specific therapies for HCC treatment.
Microenvironment
The liver tumor microenvironment is a complex interaction
network created by non-tumoral cells, molecules and soluble factors
within stroma, supportive and permissive for HCC initiation and pro-
gression. Sorafenib, a multikinase inhibitor with antiangiogenic and
antiproliferative effects for advanced HCC, works on tumor cells and
the stroma. The stroma correlates with diverse critical molecules
Fig. 2. Role of CSCs in the development of resistance of HCC to sorafenib. Tumors were hypothesized to be organized into a hierarchy of heterogeneous cell populations
and CSCs are only a tiny subset of tumor cells. Treated with sorafenib, the majority of tumor cells underwent apoptosis and CSCs were remained, showing their sorafenib
resistance. CSCs reconstructed tumor by causing the heterogeneous lineages of cancer cells, leading to relapse of malignant tumors. In addition, sharing similar character-
istics with normal stem cells, CSCs migrated, invaded into blood vessels, self-renewed, differentiated and led to tumor metastasis with acquired sorafenib resistance. CSCs
possibly contribute to recurrence and metastasis of malignant tumors with acquired drug resistance.
4J. Chen et al./Cancer Letters 367 (2015) 1–11
and signaling pathways responsible for cancer progression, includ-
ing activated Akt, Ras, NF-κB, HIF-1α, myc, hTERT and IRF4 involved
in notch, wnt and hedgehog signaling pathway [39,48,74,106]. The
angiogenic cells, immune cells and cancer associated fibroblastic cells
are three subclasses of stromal components in the liver microen-
vironment. These cells favor tumor progression and promote
resistance to therapy [86,87,106]. Resulting from these cells, an-
giogenesis, inflammation, fibrosis, hypoxia, oxidative stress and
autophagy emerged in the precancerous and neoplastic milieu, all
of which play critical roles in the development, progression, recur-
rence and drug resistance of malignant tumors [88,106–108].In
addition, for HCC patients infected chronically by hepatitis B virus
(HBV) or hepatitis C virus (HCV), viral reactivation is a vital mi-
croenvironmental factor. Role of different microenvironmental factors
in the development of resistance of HCC to sorafenib was de-
picted in Fig. 3.
Angiogenesis
HCC is a highly vascularized tumor and angiogenesis plays an
important role in the growth of liver tumor. Vascular endothelial
growth factor (VEGF) is the most critical proangiogenic factor and
the main target of mutikinases inhibitor, sorafenib [109]. VEGF can
function as a cytokine that directly affects hepatic stellate cells,
Kupffer cells and hepatocytes [110,111], and mediates the dissolu-
tion of the vascular basement membrane and the interstitial matrix
[109]. During the growth of tumor tissues, endothelial cells are ac-
tivated and become proliferative to form new vessel within the
stroma, to provide nutrients and oxygen for tumors. Remarkably,
tumor new vasculature is composed largely of immature vessels with
increased permeability, tortuosity and interstitial fluid pressure than
normal, which facilitates immune cell infiltration, metastasis and
tumor progression but compromises the delivery of therapeutics
and nutrients [91,106]. Platelet-derived endothelial cell growth factor
(PDGF), secreted by cancer cells, is also involved in cell migration
and new vessel maturation. Integrins and cadherins are other two
important mediators involved in tumor neoagiogenesis and they es-
tablish contacts required for new vascular tube formation by
mediating cell–matrix and cell–cell interactions respectively [112].
Inflammation
Inflammation is an essential element of tumor microenviron-
ment that predisposes to cancer initiation. The inflammatory
response in HCC microenvironment is regulated by several growth
factors, in particular, TGF-β, HGF and EGF. TGFβ, a difunctional mol-
ecule, acts as a tumor suppressor in normal and premalignant cells
and paradoxically, an oncogenic growth factor in cancer cells [113].
Epidermal growth factor receptor (EGFR) plays an important role
in tumor-associated angiogenesis by regulating synergistically with
several angiogenic factors [82]. Cytokines and chemokines, se-
creted not only in tumor cells but also in the surrounding tissue,
have been implicated in chronic inflammation, which have pro-
found effect on the development and progression of HCC through
mediating evasion of the immune system, angiogenesis, invasion
and dissemination [106,114].
Fibrosis
Fibrogenesis can entail secretion of several pro-angiogenic factors
by stromal cells, especially MMP, PDGF, TGFβ1, FGF and VEGF, which
provokes resistance to blood flow and reduces metabolic ex-
change of oxygen, favoring hypoxia [106]. It has been reported that
stromal fibroblasts may influence tumor initiation in adjacent epi-
thelia and promote progression [115]. Fibrosis is characterized mainly
by increased stiffness and denser collagen secreted by stromal,
Fig. 3. Role of microenvironment in the development of resistance of HCC to sorafenib. Hepatocellular carcinoma is a highly vascularized tumor. Sorafenib played an antiangiogenic
role by inhibiting VEGF and led to vasculature atrophy, leading to hypoxia. Inadequate oxygen accessibility activated hypoxia-inducible factor 1 (HIF-1α, HIF-2α), whose
expression correlates with angiogenesis, immune evasion, invasion and metastasis, inducing sorafenib resistance. Sorafenib treatment activated mTOR, upregulated expres-
sion of AKT and then caused autophagy, escaping from sorafenib efficacy. In extracellular microenvironment, with the progression of tumors, stromal cells secreted more
and more molecules and the stroma becomes stiffer and denser, affecting delivery and efficacy of sorafenib negatively. TAM are typical inflammatory cells, predisposing to
cancer initiation, and can cause oxidative stress. The relationship between inflammation, oxidative stress and sorafenib resistance has not been investigated directly until
now (—–). Hypoxia could activate fibroblast and HSC to secrete more collagen and resulted in further stiffened stroma, enhancing insensitivity to sorafenib.
5J. Chen et al./Cancer Letters 367 (2015) 1–11
inflammatory and cancer cells, which has a causative impact on the
pathogenesis of liver diseases. In the fibrotic microenvironment, it
is observed that fibrillary collagen types I and II and fibronectin are
deposited excessively in the liver [106]. The situation leads to cel-
lular transformation and metastasis, essential for tumor angiogenesis
and HCC development. The efficacy of sorafenib has been re-
ported to reduce on stiff, collagen-rich microenvironments which
impairs the delivery of chemotherapeutics and promotes aggres-
sive neoplastic cell behavior [84,89].
Hypoxia
Compared to normal liver vasculature, liver tumor vessels have
an abnormal blood flow and are excessively leaky, leading to
hypovascular areas and severe hypoxia. Inadequate oxygen acces-
sibility activates hypoxia-inducible factor 1 (HIF-1), whose expression
correlates with angiogenesis, immune evasion, invasion and me-
tastasis by interacting with several downstream genes [88,107,116].
It has been reported that during exposure to hypoxic condition, HCC
cells tended to experience EMT with increased cell migration and
invasion, and exhibit high resistance to chemotherapy [117]. Chronic
alcohol consumption, sustained antiangiogenic therapy and abnor-
mal metabolic signatures contribute to hypoxia, enhancing
proliferation, angiogenesis, metastasis, chemoresistance, and ra-
dioresistance [33,88,118]. Liang et al. confirmed the hypothesis that
hypoxia caused by the antiangiogenic effects of sustained sorafenib
therapy could induce sorafenib resistance as a cytoprotective adap-
tive response through HIF-1α and NF-κB activation, thereby limiting
sorafenib efficiency [118].
Oxidative stress
Oxidative stress caused by tumor-associated macrophages (TAM),
CSCs and cancer cells leads to an increased mutation, genomic in-
stability, epigenetic changes, and protein dysfunction through
interaction with DNA, RNA, lipid and proteins. Tumor-associated mac-
rophages (TAM) are an important component of the leukocytes in
the tumor microenvironment that can infiltrate most tumors and
promote migration of endothelial cells, neo-angiogenesis and fi-
brosis, associated with poor prognosis [119]. Oxidative stress can
activate fibroblast, the main source of collagen, and produces several
mediators implicated in tumor progression [106], suggesting that
there is a significant correlation between oxidative stress and sorafenib
resistance although no studies have already verify the hypothesis.
Autophagy
Hypoxia in the stroma, together with oxidative stress condi-
tions, can stimulate an evolutionarily conserved self-digestion
pathway, autophagy, an escape phenomenon for cancer cells during
antiangiogenic treatment [108,120]. Sorafenib treatment could lead
to accumulation of autophagosomes as evidenced by conversion from
LC3-I to LC3-II observed by immunoblot in Huh7, HLF and PLC/
PRF/5 cells due to activation of autophagic flux, suggesting that
autophagy is likely to correlate with the underlying sorafenib re-
sistance in HCC [121].
Viral reactivation
Chronic HBV or HCV infection was the major cause of HCC in
either Western or Eastern [122]. Viral reactivation is potentially fatal
and mainly triggers those patients previously exposed to HBV or
HCV [123]. It leads to deterioration of liver function, induces de-
velopment of HCC and significantly affects prognosis of patients with
HCC [124]. Zuo et al. reported that viral reactivation was the most
important correlative risk factor for HCC recurrence [125].As
sorafenib resistance becomes a hot topic, some researchers start to
focus on the correlation between viral reactivation and sorafenib
efficacy, with expectation to explore new treatment against HCC
basing their potential correlation. However, until now, the opin-
ions were still controversial.
Two recent studies reported that sorafenib exhibited a potent
anti-replicative efficacy on HCV [126,127]. An opposite result was
shown in another study, which reported that sorafenib impaired the
antiviral effect of interferon alpha on hepatitis C virus replication
[128].In addition, Zhang et al. showed that sorafenib may contrib-
ute to the reactivation of virus by immunosuppressing the
proliferation of NK cells [129]. Therefore, there is no consistent view
about the correlation about viral reactivation and sorafenib treat-
ment. What’s more, further exploration about the correlation about
viral reactivation and sorafenib resistance just took a small step.
HBV infection in HCC may enhance sorafenib resistance by
downregulating miR-193b and its downstream target protein Mcl-1,
inhibiting sorafenib-induced apoptosis [130]. Compared to HBV-
negative patients, HBV-positive HCC patients showed a less survival
benefit according to a phase III RCT from Asia-Pacific region [131].
On the contrary, HCV-positive patients showed an improvement in
median overall survival and time to progression over non-HCV in-
fected patients by a post-hoc retrospective subgroup analysis of HCV-
positive patients in the Sorafenib Hepatocellular Carcinoma
Assessment Randomized Protocol (SHARP) study [132].
Although no consistent view was obtained about the correla-
tion of viral reactivation, sorafenib treatment and sorafenib resistance,
we believe that the unremitting efforts of researchers will pave a
way for the new clinical treatment of HBV/HCV-associated HCC.
In summary, understanding how biological processes and viral
reactivation in tumor stroma interact with cancer cells and the sig-
naling pathways involved could help to illustrate the mechanism
of sorafenib resistance and identify new therapeutic and
chemopreventive targets.
Current and promising
New agents targeting pathways
Tivantinib (ARQ 197) is an oral, selective small MET tyrosine
kinase inhibitor and demonstrates significant antitumor activity es-
pecially in MET-high patients [133]. Everolimus, as a mTOR inhibitor,
may be a promising therapeutic drug for advanced HCC though no
consistent findings of its antitumor effects were observed [134].
Belinostat may induce apoptosis and tumor regression in HCC pop-
ulation by inhibiting HDAC [135].
Moreover, Tovar et al. reported that brivanib demonstrated an-
titumor activity in xenograft HCC models resistant to sorafenib
through suppressing tumorigenesis and angiogenesis via target-
ing vascular endothelial growth factor(VEGF) and fibroblast growth
factor receptors (FGFR) while Llovet et al. reported a negative result
that brivanib did not significantly improve OS of patients with HCC
who had received sorafenib treatment [136,137]. Another two new
agents, linifanib, a novel ATP-competitive inhibitor, and bevacizumab,
a monoclonal antibody, both could block the activity of VEGF and
show comparative efficacy with sorafenib, as evidenced by the results
of Phase III trial of linifanib and phase II trial of bevacizumab
[138,139]. In addition, PI-88, another new agent inhibiting
heparanase, may confer significant clinical benefits for patients with
HCC by preserving the integrity of the stroma and blocking the ac-
tivity of fibroblastic growth factors (FGF) [140].
Combination of targeted agents
The Raf/(MAPK) signaling pathway is the major targeting source
of sorafenib, whose blockage will inhibit tumorigenesis and
6J. Chen et al./Cancer Letters 367 (2015) 1–11
angiogenesis of HCC. The PI3K/Akt pathway is also involved in the
development and progression of HCC by regulating a large number
of molecules implicated in all aspects of cancer progression. Sorafenib
inhibits the Raf/(MAPK) signaling pathway but activates (PI3K)/
Akt pathway, indicating that there exist cross-talks between the PI3K/
Akt and MAPK/ERK pathways and that the latent compensatory
mechanism of the PI3K/Akt pathway may contribute to sorafenib
resistance in HCC. Therefore, better survival outcomes may be
achieved by using promising combinations of targeted therapies
through inhibiting different pathways involved in HCC. Ras/MAPK
and AKT/mTOR pathways are frequently deregulated in human
hepatocarcinogenesis [141]. The combination of PKI-587, target-
ing PI3K/AKT/mTOR pathway and sorafenib, mainly targeting Ras/
Raf/MAPK signaling pathways, has the advantage over monodrug
therapy on inhibition of HCC cell proliferation by blocking both the
signaling pathways simultaneously [142]. Additionally, results of an
ongoing, randomized, open-label, phase II study are awaited, which
would compare the efficacy of everolimus, targeting PI3K/AKT/
mTOR pathway, combined with sorafenib to sorafenib alone
(ClinicalTrials.gov identifier NCT01005199). However, a recent phase
III, randomized, double-blind, placebo-controlled trial of sorafenib
plus erlotinib (a EGFR inhibitor) in patients with advanced HCC
showed no improved survival compared to sorafenib alone [143].
The horizontal blockades on different pathways by using com-
bination of targeted agents showed a good perspective in HCC
treatment. At the same time, vertical blockades on Ras/MAPK
pathway were also emphasized in exploration of new strategies for
overcoming sorafenib resistance.
The Ras/MAPK pathway is the key pathway in the treatment of
HCC using sorafenib and the inhibition of the RAF kinases is an im-
portant facet of the action of sorafenib in HCC. Combining sorafenib
with other inhibitors active on other kinases or facets of the Ras/
MAPK pathway, such as MEK inhibitors, might achieve a better
efficacy of sorafenib and therefore a better control of tumor growth
[144]. Moreover, refametinib plus sorafenib with reduced doses
showed antitumor activity in patients with HCC. Especially, in pa-
tients with RAS mutations, refametinib/sorafenib combination could
significantly improve their overall survival [145].
In order to improve the impaired efficacy caused by sorafenib
resistance in advanced HCC, drugs combination is a promising di-
rection but the efficacy of sorafenib combined with other molecular
targeted drugs still needs further exploration.
Combination of cytotoxic chemotherapeutic agents
Extensive efforts have been devoted to evaluation of different cy-
totoxic agents by different methods, including Metronomic
capecitabine [146], gemcitabine [147], cisplatin [148] and doxoru-
bicin [149], for treatment of advanced HCC. But the objective
response rate (ORR) to a single cytotoxic regimen was merely 0–10%
with no significant survival benefit [150]. In order to explore new
path to overcome sorafenib resistance and achieve better treat-
ment of advanced HCC, the combination of sorafenib and cytotoxic
agents has attracted focus of researchers. Wang et al. have re-
ported that sorafenib–irinotecan sequential therapy augmented the
anti-tumor efficacy of monotherapy in hepatocellular carcinoma cells
HepG2 [151]. In another study, Liu et al. reported that GEMOX com-
bined with sorafenib as first-line therapy followed by sorafenib as
maintenance therapy was effective with manageable toxicity for
treatment of advanced HCC [150]. Furthermore, the combination
of sorafenib and gemcitabine showed modest clinical efficacy with
good toleration in advanced HCC [152]. In addition, compared with
doxorubicin monotherapy, sorafenib plus doxorubicin resulted in
greater median time to progression, overall survival, and progression-
free survival [153]. The effectiveness and the safety of some cytotoxic
chemotherapeutic agents are still in stage III clinical trial. The clinical
application of cytotoxic chemotherapeutic agents combined with
sorafenib still has a long way to go.
Combination with immunotherapeutic drug
Immune checkpoint blockade has been considered as a prom-
ising therapeutic approach for HCC. Immunotherapeutic drugs are
not metabolized in the liver and HCC is typically immunogenic, so
theoretically predictable pharmacokinetic profiles may be achieved
in HCC with no severe hepatoxicity [154]. However, immunothera-
peutic drug alone sometimes didn’t demonstrate antitumor activity.
A phase I/II trial of CT-011(anti-PD-1 antibodies) in advanced HCC
was initiated but stopped due to slow accrual (ClinicalTrials.gov iden-
tifier NCT00966251). Therefore, combination of immunotherapeutic
drug and sorafenib has recently emerged as a new treatment method
for advanced HCC. Wang et al. reported that this combinatorial ap-
proach resulted in blockade of programmed death-ligand 1 (PDL1)
and sequential effective natural killer cell responses against hepa-
tocellular carcinoma [155]. Chen et al. reported that combination
of anti-PD-1 immunotherapy with sorafenib showed efficacy con-
comitant with targeting of the hypoxic and immunosuppressive
microenvironment in mice model [156]. In a recent study in vivo
and in vitro, the result showed that targeting CD47 in combina-
tion with sorafenib is an attractive therapeutic regimen for the
treatment of HCC patients [9].
Nevertheless, the successful application of immunotherapy in
HCC will have to take into account the liver cancer-specific immune
microenvironment and responses.
Personalized therapy
Indeed, differences in etiology, disease stage and pathological
process make HCC a highly dyshomogeneous malignant tumor. In
terms of quality and quantity, individual differences affect sensi-
tiveness of patients with HCC to sorafenib treatment [6]. Selection
of patient candidates more sensitive to sorafenib treatment is fea-
sible in human malignancies according to their tumor molecular
background. Thus, a molecular identification is crucial to assess the
levels of new targets, and to customize therapies in patients with
HCC.
Low or high level of miR-425-3p has been postulated respon-
sible for stratifying patients with advanced HCC for sorafenib
treatment [157]. VEGF, VEGFR and VEGFA may represent a valu-
able asset to better identify HCC patients more likely to benefit from
sorafenib treatment [158,159]. Mapk14 inhibition has been re-
ported to be associated with HCC sensitization to sorafenib, indicating
that Mapk14 may be used to classify the HCC population sensitive
or resistant to sorafenib [160]. Other potential biomarkers, still in
exploration, may be essential for personalized therapy, including
pERK [161],STAT3[162], EGFR/HER-3 [163], Nanog [164],JNK[165],
HGF/c-KIT [166], FGF3/FGF4 [167], αB-Crystallin [17] and others.
The exploration of biomarkers is a great step moving toward per-
sonalized therapy in patients with HCC. However, the high genetic
heterogeneity of HCC is difficult to refine due to genome-wide ap-
proaches used for genes identification though many progresses have
been made in this regard. With the accelerated progress of clini-
cal management, it is no doubt that the more precise knowledge
of the molecular classes of HCC, better-designed clinical trials, and
more effective blockade of specific pathways with new therapies
will definitely implement personalized therapy for HCC.
Conclusion
Hepatocellular carcinoma (HCC), the most common primary liver
tumor, is notoriously resistant to systemic therapies. In particular,
sorafenib resistance of advanced HCC generated during sustained
7J. Chen et al./Cancer Letters 367 (2015) 1–11
drug treatment has raised great concern throughout the world. Un-
derstanding the underlying mechanism is in urgent need in order
to potentiate the antitumor effect of sorafenib and identify new
promising chemotherapeutic therapies. In this review, we present
potential molecular, cellular and microenvironmental mechanism
to explain sorafenib resistance in HCC by synthesizing the retro-
spective studies and report systematically and comprehensively. EMT
and MET, together with the critical growth factors and signaling
pathway involved in the two transition processes, play a pivotal role
in sorafenib resistance. Furthermore, CSCs and CSC-like cells, as the
clonogenic core of the liver tumor, have an important impact on the
generation of sorafenib resistance. Therapeutic approaches target-
ing CSCs or signaling pathways involved in CSCs are hypothesized
as a breakthrough for treating advanced HCC. In addition, the liver
tumor microenvironment functions as an essential part in tumor
development and progression. The biological processes involved in
tumor microenvironment, angiogenesis, inflammation, fibrosis,
hypoxia, oxidative stress, autophagy and viral reactivation have been
demonstrated significantly or potentially correlated with sorafenib
resistance. This review first provides systemic comment on the mech-
anism of sorafenib resistance in HCC and hopes to motivate new
thoughts about the chemotherapeutic targets or approaches. In order
to overcome the impaired efficacy of sorafenib induced by origi-
nal or acquired resistance, immense amount of studies in the field
of HCC treatment was done with expectation to explore new ther-
apeutic drugs or approaches with significant antitumor activity. These
studies described the state-of-art developments in the field and an-
ticipated the problems to be solved.
Funding
This work was supported by National Natural Science Founda-
tion of China (81201942) and Zhejiang Provincial Natural Science
Foundation of China (LZ14H160002, LY15H160039). The funders had
no role in study design, data collection and analysis, decision to
publish, or preparation of the manuscript.
Conflict of interest
The authors declare that there are no conflicts of interest.
References
[1] G. Yu, Y. Jing, X. Kou, F. Ye, L. Gao, Q. Fan, et al., Hepatic stellate cells secreted
hepatocyte growth factor contributes to the chemoresistance of hepatocellular
carcinoma, PLoS ONE 8 (2013) e73312.
[2] J. Ferlay, I. Soerjomataram, M. Ervik, R. Dikshit, S. Eser, C. Mathers, et al.,
GLOBOCAN 2012 v1. 0, Cancer Incidence and Mortality Worldwide: IARC
CancerBase No. 11 [Internet]. Lyon, Fr Int Agency Res Cancer, 2013. Available
from: <http://globocan.iarc.fr>, (accessed on 12.12.13).
[3] A.J. Page, D.C. Cosgrove, B. Philosophe, T.M. Pawlik, Hepatocellular carcinoma:
diagnosis, management, and prognosis, Surg. Oncol. Clin. N. Am. 23 (2014)
289–311.
[4] A. Pahwa, K. Beckett, S. Channual, N. Tan, D.S. Lu, S.S. Raman, Efficacy of the
American Association for the Study of Liver Disease and Barcelona criteria for
the diagnosis of hepatocellular carcinoma, Abdom. Imaging 39 (2014) 753–760.
[5] L. Wang, M. Yao, Z. Dong, Y. Zhang, D. Yao, Circulating specific biomarkers in
diagnosis of hepatocellular carcinoma and its metastasis monitoring, Tumour
Biol. 35 (2014) 9–20.
[6] G. Giannelli, B. Rani, F. Dituri, Y. Cao, G. Palasciano, Moving towards
personalised therapy in patients with hepatocellular carcinoma: the role of
the microenvironment, Gut 63 (2014) 1668–1676.
[7] Q. Wu, Z. Yang, Y. Nie, Y. Shi, D. Fan, Multi-drug resistance in cancer
chemotherapeutics: mechanisms and lab approaches, Cancer Lett. 347 (2014)
159–166.
[8] N.A. Saunders, F. Simpson, E.W. Thompson, M.M. Hill, L. Endo-Munoz, G.
Leggatt, et al., Role of intratumoural heterogeneity in cancer drug resistance:
molecular and clinical perspectives, EMBO Mol. Med. 4 (2012) 675–684.
[9] J. Lo, E.Y. Lau, R.H. Ching, B.Y. Cheng, M.K. Ma, I.O. Ng, et al., NF-kappaB
mediated CD47 upregulation promotes sorafenib resistance and its blockade
synergizes the effect of sorafenib in hepatocellular carcinoma, Hepatology
(2015).
[10] G.M. Keating, A. Santoro, Sorafenib a review of its use in advanced
hepatocellular carcinoma, Drugs 69 (2009) 223–240.
[11] A. Galmiche, B. Chauffert, J.C. Barbare, New biological perspectives for the
improvement of the efficacy of sorafenib in hepatocellular carcinoma, Cancer
Lett. 346 (2014) 159–162.
[12] J. Castillo, E. Erroba, M.J. Perugorria, M. Santamaria, D.C. Lee, J. Prieto, et al.,
Amphiregulin contributes to the transformed phenotype of human
hepatocellular carcinoma cells, Cancer Res. 66 (2006) 6129–6138.
[13] Z. Ezzoukhry, C. Louandre, E. Trecherel, C. Godin, B. Chauffert, S. Dupont, et al.,
EGFR activation is a potential determinant of primary resistance of
hepatocellular carcinoma cells to sorafenib, Int. J. Cancer 131 (2012) 2961–
2969.
[14] L. Liu, Y. Cao, C. Chen, X. Zhang, A. McNabola, D. Wilkie, et al., Sorafenib blocks
the RAF/MEK/ERK pathway, inhibits tumor angiogenesis, and induces tumor
cell apoptosis in hepatocellular carcinoma model PLC/PRF/5, Cancer Res. 66
(2006) 11851–11858.
[15] A.-L. Cheng, Y.-K. Kang, Z. Chen, C.-J. Tsao, S. Qin, J.S. Kim, et al., Efficacy and
safety of sorafenib in patients in the Asia-Pacific region with advanced
hepatocellular carcinoma: a phase III randomised, double-blind, placebo-
controlled trial, Lancet Oncol. 10 (2009) 25–34.
[16] J.M. Llovet, S. Ricci, V. Mazzaferro, P. Hilgard, E. Gane, J.-F. Blanc, et al., Sorafenib
in advanced hepatocellular carcinoma, NEJM 359 (2008) 378–390.
[17] X.Y. Huang, A.W. Ke, G.M. Shi, X. Zhang, C. Zhang, Y.H. Shi, et al., alphaB-
crystallin complexes with 14-3-3zeta to induce epithelial-mesenchymal
transition and resistance to sorafenib in hepatocellular carcinoma, Hepatology
57 (2013) 2235–2247.
[18] J.-F. Chiou, C.-J. Tai, M.-T. Huang, P.-L. Wei, Y.-H. Wang, J. An, et al., Glucose-
regulated protein 78 is a novel contributor to acquisition of resistance to
sorafenib in hepatocellular carcinoma, Ann. Surg. Oncol. 17 (2010) 603–612.
[19] K.-F. Chen, H.-L. Chen, W.-T. Tai, W.-C. Feng, C.-H. Hsu, P.-J. Chen, et al.,
Activation of phosphatidylinositol 3-kinase/Akt signaling pathway mediates
acquired resistance to sorafenib in hepatocellular carcinoma cells, J. Pharmacol.
Exp. Ther. 337 (2011) 155–161.
[20] Y.Y. Lu, F. Feng, F. Zhang, X. Gao, C. Wang, X. Chang, et al., A novel mechanism
of PXR mediated Sorafenib drug resistance in HCC through MDR and CYP3A4,
Hepatology (2013) 1084A.
[21] M.A. Nieto, A. Cano, The epithelial–mesenchymal transition under control:
global programs to regulate epithelial plasticity, Semin. Cancer Biol. 22 (2012)
361–368.
[22] Y.D. Shaul, E. Freinkman, W.C. Comb, J.R. Cantor, W.L. Tam, P. Thiru, et al.,
Dihydropyrimidine accumulation is required for the epithelial-mesenchymal
transition, Cell 158 (2014) 1094–1109.
[23] J.P. Thiery, H. Acloque, R.Y. Huang, M.A. Nieto, Epithelial-mesenchymal
transitions in development and disease, Cell 139 (2009) 871–890.
[24] K. Uchibori, A. Kasamatsu, M. Sunaga, S. Yokota, T. Sakurada, E. Kobayashi, et al.,
Establishment and characterization of two 5-fluorouracil-resistant
hepatocellular carcinoma cell lines, Int. J. Oncol. 40 (2012) 1005–1010.
[25] F.J. Carmona, V. Davalos, E. Vidal, A. Gomez, H. Heyn, Y. Hashimoto, et al., A
comprehensive DNA methylation profile of epithelial-to-mesenchymal
transition, Cancer Res. 74 (2014) 5608–5619.
[26] A. de Reyniès, M.-C. Jaurand, A. Renier, G. Couchy, I. Hysi, N. Elarouci, et al.,
Molecular classification of malignant pleural mesothelioma: identification of
a poor prognosis subgroup linked to the epithelial-to-mesenchymal transition,
Clin. Cancer Res. 20 (2014) 1323–1334.
[27] S. Kunnimalaiyaan, T.C. Gamblin, M. Kunnimalaiyaan, Role of notch3 in
epithelial-to-mesenchymal transition of HCC, Cancer Res. 74 (2014) 5289.
[28] J.P. Morris, S.C. Wang, M. Hebrok, KRAS, Hedgehog, Wnt and the twisted
developmental biology of pancreatic ductal adenocarcinoma, Nat. Rev. Cancer
10 (2010) 683–695.
[29] S. Lamouille, J. Xu, R. Derynck, Molecular mechanisms of epithelial-
mesenchymal transition, Nat. Rev. Mol. Cell Biol. 15 (2014) 178–196.
[30] Z. Ke, S. Caiping, Z. Qing, W. Xiaojing, Sonic hedgehog-Gli1 signals promote
epithelial-mesenchymal transition in ovarian cancer by mediating PI3K/AKT
pathway, Med. Oncol. 32 (2015) 368.
[31] Z.-J. Zhou, Z. Dai, S.-L. Zhou, Z.-Q. Hu, Q. Chen, Y.-M. Zhao, et al., HNRNPAB
induces epithelial–mesenchymal transition and promotes metastasis of
hepatocellular carcinoma by transcriptionally activating SNAIL, Cancer Res.
74 (2014) 2750–2762.
[32] G. Semenza, Cancer–stromal cell interactions mediated by hypoxia-inducible
factors promote angiogenesis, lymphangiogenesis, and metastasis, Oncogene
32 (2013) 4057–4063.
[33] K. Yuki, S. Hirohashi, M. Sakamoto, T. Kanai, Y. Shimosato, Growth and spread
of hepatocellular carcinoma: a review of 240 consecutive autopsy cases, Cancer
66 (1990) 2174–2179.
[34] J.P. Thiery, Epithelial–mesenchymal transitions in tumour progression, Nat.
Rev. Cancer 2 (2002) 442–454.
[35] F.L. Cousins, A. Murray, A. Esnal, D.A. Gibson, H.O. Critchley, P.T. Saunders,
Evidence from a mouse model that epithelial cell migration and
mesenchymal-epithelial transition contribute to rapid restoration of uterine
tissue integrity during menstruation, PLoS ONE 9 (2014) e86378.
[36] K. Polyak, R.A. Weinberg, Transitions between epithelial and mesenchymal
states: acquisition of malignant and stem cell traits, Nat. Rev. Cancer 9 (2009)
265–273.
[37] Y. Chao, Q. Wu, C. Shepard, A. Wells, Hepatocyte induced re-expression of
E-cadherin in breast and prostate cancer cells increases chemoresistance, Clin.
Exp. Metastasis 29 (2012) 39–50.
8J. Chen et al./Cancer Letters 367 (2015) 1–11
[38] T.S. Gujral, M. Chan, L. Peshkin, P.K. Sorger, M.W. Kirschner, G. MacBeath, A
noncanonical Frizzled2 pathway regulates epithelial-mesenchymal transition
and metastasis, Cell 159 (2014) 844–856.
[39] M. Cojoc, C. Peitzsch, I. Kurth, F. Trautmann, L.A. Kunz-Schughart, G.D. Telegeev,
et al., Aldehyde dehydrogenase is regulated by Beta-catenin/TCF and promotes
radioresistance in prostate cancer progenitor cells, Cancer Res. (2015).
[40] H. Liu, J. Yin, H. Wang, G. Jiang, M. Deng, G. Zhang, et al., FOXO3a modulates
WNT/beta-catenin signaling and suppresses epithelial-to-mesenchymal
transition in prostate cancer cells, Cell. Signal. 27 (2015) 510–518.
[41] B. Bao, A.S. Azmi, S. Ali, A. Ahmad, Y. Li, S. Banerjee, et al., The biological kinship
of hypoxia with CSC and EMT and their relationship with deregulated
expression of miRNAs and tumor aggressiveness, Biochim. Biophys. Acta 2012
(1826) 272–296.
[42] S. Lamouille, R. Derynck, Emergence of the phosphoinositide 3-kinase-Akt-
mammalian target of rapamycin axis in transforming growth factor-β-induced
epithelial-mesenchymal transition, Cells Tissues Organs 193 (2010) 8.
[43] M. Korpal, B.J. Ell, F.M. Buffa, T. Ibrahim, M.A. Blanco, T. Celià-Terrassa, et al.,
Direct targeting of Sec23a by miR-200s influences cancer cell secretome and
promotes metastatic colonization, Nat. Med. 17 (2011) 1101–1108.
[44] J. Gros, C.J. Tabin, Vertebrate limb bud formation is initiated by localized
epithelial-to-mesenchymal transition, Science 343 (2014) 1253–1256.
[45] S. Grünert, M. Jechlinger, H. Beug, Diverse cellular and molecular mechanisms
contribute to epithelial plasticity and metastasis, Nat. Rev. Mol. Cell Biol. 4
(2003) 657–665.
[46] L.A. Timmerman, J. Grego-Bessa, A. Raya, E. Bertran, J.M. Perez-Pomares, J.
Diez, et al., Notch promotes epithelial-mesenchymal transition during
cardiac development and oncogenic transformation, Genes Dev. 18 (2004)
99–115.
[47] K.G. Leong, K. Niessen, I. Kulic, A. Raouf, C. Eaves, I. Pollet, et al., Jagged1-
mediated Notch activation induces epithelial-to-mesenchymal transition
through Slug-induced repression of E-cadherin, J. Exp. Med. 204 (2007)
2935–2948.
[48] C. Sahlgren, M.V. Gustafsson, S. Jin, L. Poellinger, U. Lendahl, Notch signaling
mediates hypoxia-induced tumor cell migration and invasion, Proc. Natl. Acad.
Sci. U.S.A. 105 (2008) 6392–6397.
[49] S.A. Mani, W. Guo, M.-J. Liao, E.N. Eaton, A. Ayyanan, A.Y. Zhou, et al., The
epithelial-mesenchymal transition generates cells with properties of stem cells,
Cell 133 (2008) 704–715.
[50] P. Mallini, T. Lennard, J. Kirby, A. Meeson, Epithelial-to-mesenchymal transition:
what is the impact on breast cancer stem cells and drug resistance, Cancer
Treat. Rev. 40 (2014) 341–348.
[51] P. Bhat-Nakshatri, H. Appaiah, C. Ballas, P. Pick-Franke, R. Goulet, S. Badve, et al.,
SLUG/SNAI2 and tumor necrosis factor generate breast cells with CD44+/CD24-
phenotype, BMC Cancer 10 (2010) 411.
[52] B.G. Hollier, A.A. Tinnirello, S.J. Werden, K.W. Evans, J.H. Taube, T.R. Sarkar,
et al., FOXC2 expression links epithelial–mesenchymal transition and stem
cell properties in breast cancer, Cancer Res. 73 (2013) 1981–1992.
[53] J. Bruix, M. Sherman, Management of hepatocellular carcinoma: an update,
Hepatology 53 (2011) 1020–1022.
[54] D.J. Mulholland, N. Kobayashi, M. Ruscetti, A. Zhi, L.M. Tran, J.T. Huang, et al.,
Pten loss and RAS/MAPK activation cooperate to promote EMT and metastasis
initiated from prostate cancer stem/progenitor cells, Cancer Res. 72 (2012)
1878–1889.
[55] J.M. Buonato, M.J. Lazzara, ERK1/2 blockade prevents epithelial-mesenchymal
transition in lung cancer cells and promotes their sensitivity to EGFR inhibition,
Cancer Res. 74 (2014) 309–319.
[56] C.H. Hou, F.L. Lin, S.M. Hou, J.F. Liu, Cyr61 promotes epithelial-mesenchymal
transition and tumor metastasis of osteosarcoma by Raf-1/MEK/ERK/Elk-1/
TWIST-1 signaling pathway, Mol. Cancer 13 (2014) 236.
[57] M.M. McKay, D.A. Ritt, D.K. Morrison, RAF inhibitor-induced KSR1/B-RAF
binding and its effects on ERK cascade signaling, Curr. Biol. 21 (2011) 563–568.
[58] P.I. Poulikakos, C. Zhang, G. Bollag, K.M. Shokat, N. Rosen, RAF inhibitors
transactivate RAF dimers and ERK signalling in cells with wild-type BRAF,
Nature 464 (2010) 427–430.
[59] K. Miyahara, K. Nouso, K. Yamamoto, Chemotherapy for advanced
hepatocellular carcinoma in the sorafenib age, World J. Gastroenterol. 20 (2014)
4151.
[60] N.C. Schlegel, A. von Planta, D.S. Widmer, R. Dummer, G. Christofori, PI3K
signalling is required for a TGF beta-induced epithelial-mesenchymal-like
transition (EMT-like) in human melanoma cells, Exp. Dermatol. 24 (2015)
22–28.
[61] G.Y. Lin, R.H. Gai, Z.B. Chen, Y.J. Wang, S.D. Liao, R. Dong, et al., The dual
PI3K/mTOR inhibitor NVP-BEZ235 prevents epithelial-mesenchymal transition
induced by hypoxia and TGF-beta 1, Eur. J. Pharmacol. 729 (2014) 45–53.
[62] S.K. Gruvberger-Saal, R. Parsons, Is the small heat shock protein αB-crystallin
an oncogene?, J. Clin. Invest. 116 (2006) 30.
[63] S. Seitz, E. Korsching, J. Weimer, A. Jacobsen, N. Arnold, A. Meindl, et al., Genetic
background of different cancer cell lines influences the gene set involved in
chromosome 8 mediated breast tumor suppression, Genes Chromosomes
Cancer 45 (2006) 612–627.
[64] J.V. Moyano, J.R. Evans, F. Chen, M. Lu, M.E. Werner, F. Yehiely, et al., αB-
crystallin is a novel oncoprotein that predicts poor clinical outcome in breast
cancer, J. Clin. Invest. 116 (2006) 261.
[65] J. Holcakova, L. Hernychová, P. Bouchal, K. Brozkova, J. Zaloudik, D. Valík, et al.,
Identification of alphaB-crystallin, a biomarker of renal cell carcinoma by
SELDI-TOF MS, Int. J. Biol. Markers 23 (2007) 48–53.
[66] Q. Tang, Y.-F. Liu, X.-J. Zhu, Y.-H. Li, J. Zhu, J.-P. Zhang, et al., Expression and
prognostic significance of the αB-crystallin gene in human hepatocellular
carcinoma, Hum. Pathol. 40 (2009) 300–305.
[67] T. Rajakulendran, M. Sahmi, M. Lefrancois, F. Sicheri, M. Therrien, A
dimerization-dependent mechanism drives RAF catalytic activation, Nature
461 (2009) 542–545.
[68] J.S. Bae, S.J. Noh, K.M. Kim, K.Y. Jang, M.J. Chung, D.G. Kim, et al., Serum
response factor induces epithelial to mesenchymal transition with resistance
to sorafenib in hepatocellular carcinoma, Int. J. Oncol. 44 (2014) 129–136.
[69] J. Fernando, A. Malfettone, E.B. Cepeda, R. Vilarrasa-Blasi, E. Bertran, G.
Raimondi, et al., A mesenchymal-like phenotype and expression of CD44
predict lack of apoptotic response to sorafenib in liver tumor cells, Int. J. Cancer
136 (2015) E161–E172.
[70] F. van Zijl, G. Zulehner, M. Petz, D. Schneller, C. Kornauth, M. Hau, et al.,
Epithelial-mesenchymal transition in hepatocellular carcinoma, Future Oncol.
5 (2009) 1169–1179.
[71] P.Gulhati, K.A. Bowen, J.Y. Liu, P.D.Stevens, P.G. Rychahou,M. Chen, et al., mTORC1
and mTORC2 regulate EMT, motility, and metastasis of colorectal cancer via
RhoA and Rac1 signaling pathways, Cancer Res. 71 (2011) 3246–3256.
[72] C.W. Li, W. Xia, L. Huo, S.O. Lim, Y. Wu, J.L. Hsu, et al., Epithelial-mesenchymal
transition induced by TNF-alpha requires NF-kappaB-mediated transcriptional
upregulation of Twist1, Cancer Res. 72 (2012) 1290–1300.
[73] K. Hao, X.D. Tian, C.F. Qin, X.H. Xie, Y.M. Yang, Hedgehog signaling pathway
regulates human pancreatic cancer cell proliferation and metastasis, Oncol.
Rep. 29 (2013) 1124–1132.
[74] Y. Bai, H. Lu, C. Wu, Y. Liang, S. Wang, C. Lin, et al., Resveratrol inhibits
epithelial-mesenchymal transition and renal fibrosis by antagonizing the
hedgehog signaling pathway, Biochem. Pharmacol. 92 (2014) 484–493.
[75] L. Yu, Y. Mu, N. Sa, H. Wang, W. Xu, Tumor necrosis factor alpha induces
epithelial-mesenchymal transition and promotes metastasis via NF-kappaB
signaling pathway-mediated TWIST expression in hypopharyngeal cancer,
Oncol. Rep. 31 (2014) 321–327.
[76] A.G. Pantschenko, I. Pushkar, K.H. Anderson, Y. Wang, L.J. Miller, S.H. Kurtzman,
et al., The interleukin-1 family of cytokines and receptors in human breast
cancer: implications for tumor progression, Int. J. Oncol. 23 (2003) 269–284.
[77] T. Leibovich-Rivkin, Y. Liubomirski, B. Bernstein, T. Meshel, A. Ben-Baruch,
Inflammatory factors of the tumor microenvironment induce plasticity in
nontransformed breast epithelial cells: EMT, invasion, and collapse of normally
organized breast textures, Neoplasia 15 (2013) 1330–1346.
[78] S. Lindsey, S.A. Langhans, Crosstalk of oncogenic signaling pathways during
epithelial-mesenchymal transition, Front. Oncol. 4 (2014) 358.
[79] M. Shiota, J.L. Bishop, K.M. Nip, A. Zardan, A. Takeuchi, T. Cordonnier, et al.,
Hsp27 regulates epithelial mesenchymal transition, metastasis, and circulating
tumor cells in prostate cancer, Cancer Res. 73 (2013) 3109–3119.
[80] P. Li, R. Yang, W.Q. Gao, Contributions of epithelial-mesenchymal transition
and cancer stem cells to the development of castration resistance of prostate
cancer, Mol. Cancer 13 (2014) 55.
[81] N. Sheshadri, J.M. Catanzaro, A.J. Bott, Y. Sun, E. Ullman, E.I. Chen, et al.,
SCCA1/SERPINB3 promotes oncogenesis and epithelial-mesenchymal transition
via the unfolded protein response and IL6 signaling, Cancer Res. 74 (2014)
6318–6329.
[82] A. De Luca, A. Carotenuto, A. Rachiglio, M. Gallo, M.R. Maiello, D. Aldinucci,
et al., The role of the EGFR signaling in tumor microenvironment, J. Cell. Physiol.
214 (2008) 559–567.
[83] N. Oishi, T. Yamashita, S. Kaneko, Molecular biology of liver cancer stem cells,
Liver Cancer 3 (2014) 71.
[84] K.R. Levental, H. Yu, L. Kass, J.N. Lakins, M. Egeblad, J.T. Erler, et al., Matrix
crosslinking forces tumor progression by enhancing integrin signaling, Cell
139 (2009) 891–906.
[85] J.E. Dick, Tumor archaeology: tracking leukemic evolution to its origins, Sci.
Trans. Med. 6 (2014) 238fs223.
[86] D.W. McMillin, J. Delmore, E. Weisberg, J.M. Negri, D.C. Geer, S. Klippel, et al.,
Tumor cell-specific bioluminescence platform to identify stroma-induced
changes to anticancer drug activity, Nat. Med. 16 (2010) 483–489.
[87] C.S. Mitsiades, N. Mitsiades, N.C. Munshi, K.C. Anderson, Focus on multiple
myeloma, Cancer Cell 6 (2004) 439–444.
[88] H.-M. Ni, A. Bhakta, S. Wang, Z. Li, S. Manley, H. Huang, et al., Role of hypoxia
inducing factor-1β in alcohol-induced autophagy, steatosis and liver injury
in mice, PLoS ONE 9 (2014) e115849.
[89] T.V. Nguyen, M. Sleiman, T. Moriarty, W.G. Herrick, S.R. Peyton, Extracellular
matrix stiffness protects carcinoma cells from sorafenib via JNK signaling, in:
Northeast Bioengineering Conference (NEBEC), 2014 40th Annual, IEEE, 2014,
pp. 1–2.
[90] A.K. Iyer, A. Singh, S. Ganta, M.M. Amiji, Role of integrated cancer nanomedicine
in overcoming drug resistance, Adv. Drug Deliv. Rev. 65 (2013) 1784–1802.
[91] R.K. Jain, Normalizing tumor vasculature with anti-angiogenic therapy: a new
paradigm for combination therapy, Nat. Med. 7 (2001) 987–989.
[92] D. Bonnet, J.E. Dick, Human acute myeloid leukemia is organized as a hierarchy
that originates from a primitive hematopoietic cell, Nat. Med. 3 (1997)
730–737.
[93] Z. Zhu, X. Hao, M. Yan, M. Yao, C. Ge, J. Gu, et al., Cancer stem/progenitor cells
are highly enriched in CD133+CD44+population in hepatocellular carcinoma,
Int. J. Cancer 126 (2010) 2067–2078.
[94] M.F. Clarke, J.E. Dick, P.B. Dirks, C.J. Eaves, C.H. Jamieson, D.L. Jones, et al., Cancer
stem cells perspectives on current status and future directions: AACR
Workshop on cancer stem cells, Cancer Res. 66 (2006) 9339–9344.
9J. Chen et al./Cancer Letters 367 (2015) 1–11
[95] L.E. Ailles, I.L. Weissman, Cancer stem cells in solid tumors, Curr. Opin.
Biotechnol. 18 (2007) 460–466.
[96] J.A. Kennedy, F. Barabé, A.G. Poeppl, J.C. Wang, J.E. Dick, Comment on “Tumor
growth need not be driven by rare cancer stem cells”, Science 318 (2007) 1722.
[97] S. Bao, Q. Wu, R.E. McLendon, Y. Hao, Q. Shi, A.B. Hjelmeland, et al., Glioma
stem cells promote radioresistance by preferential activation of the DNA
damage response, Nature 444 (2006) 756–760.
[98] K.-J. Wu, M.-H. Yang, Epithelial-mesenchymal transition and cancer stemness:
the Twist1-Bmi1 connection, Biosci. Rep. 31 (2011) 449–455.
[99] Z.F. Bielecka, A.M. Czarnecka, W. Solarek, A. Kornakiewicz, C. Szczylik,
Mechanisms of acquired resistance to tyrosine kinase inhibitors in clear-cell
renal cell carcinoma (ccRCC), Curr. Signal Transduct. Ther. 8 (2014) 218.
[100] A. Biddle, I.C. Mackenzie, Cancer stem cells and EMT in carcinoma, Cancer
Metastasis Rev. 31 (2012) 285–293.
[101] C. Chang, M. Hung, The role of EZH2 in tumour progression, Br. J. Cancer 106
(2012) 243–247.
[102] T. Kiesslich, D. Neureiter, Advances in targeting the Hedgehog signaling
pathway in cancer therapy, Expert Opin. Ther. Targets 16 (2012) 151–156.
[103] J.-J. Zhou, X.-G. Deng, X.-Y. He, Y. Zhou, M. Yu, W.-C. Gao, et al., Knockdown
of NANOG enhances chemosensitivity of liver cancer cells to doxorubicin by
reducing MDR1 expression, Int. J. Oncol. 44 (2014) 2034–2040.
[104] R. Gedaly, R. Galuppo, Y. Musgrave, P. Angulo, J. Hundley, M. Shah, et al.,
PKI-587 and sorafenib alone and in combination on inhibition of liver cancer
stem cell proliferation, J. Surg. Res. 185 (2013) 225–230.
[105] H.-W. Xin, C.M. Ambe, D.M. Hari, G.W. Wiegand, T.C. Miller, J.-Q. Chen, et al.,
Label-retaining liver cancer cells are relatively resistant to sorafenib, Gut 62
(2013) 1777–1786.
[106] V. Hernandez-Gea, S. Toffanin, S.L. Friedman, J.M. Llovet, Role of the
microenvironment in the pathogenesis and treatment of hepatocellular
carcinoma, Gastroenterology 144 (2013) 512–527.
[107] Z. Von Marschall, T. Cramer, M. Höcker, G. Finkenzeller, B. Wiedenmann, S.
Rosewicz, Dual mechanism of vascular endothelial growth factor upregulation
by hypoxia in human hepatocellular carcinoma, Gut 48 (2001) 87–96.
[108] Y.-H. Shi, Z.-B. Ding, J. Zhou, B. Hui, G.-M. Shi, A.-W. Ke, et al., Targeting
autophagy enhances sorafenib lethality for hepatocellular carcinoma via ER
stress-related apoptosis, Autophagy 7 (2011) 1159–1172.
[109] A.X. Zhu, D.G. Duda, D.V. Sahani, R.K. Jain, HCC and angiogenesis: possible
targets and future directions, Nat. Rev. Clin. Oncol. 8 (2011) 292–301.
[110] J. LeCouter, D.R. Moritz, B. Li, G.L. Phillips, X.H. Liang, H.-P. Gerber, et al.,
Angiogenesis-independent endothelial protection of liver: role of VEGFR-1,
Science 299 (2003) 890–893.
[111] B.M. Lichtenberger, P.K. Tan, H. Niederleithner, N. Ferrara, P. Petzelbauer, M.
Sibilia, Autocrine VEGF signaling synergizes with EGFR in tumor cells to
promote epithelial cancer development, Cell 140 (2010) 268–279.
[112] S. Coulon, F. Heindryckx, A. Geerts, C. Van Steenkiste, I. Colle, H.
Van Vlierberghe, Angiogenesis in chronic liver disease and its complications,
Liver Int. 31 (2011) 146–162.
[113] S.J. Faivre, A. Santoro, R.K. Kelley, P. Merle, E. Gane, J.-Y. Douillard, et al., A phase
2 study of a novel transforming growth factor-beta (TGF-{beta} 1) receptor I
kinase inhibitor, LY2157299 monohydrate (LY), in patients with advanced
hepatocellular carcinoma (HCC), in: ASCO Annual Meeting Proceedings, 2014,
p. LBA173.
[114] E.T. Roussos, J.S. Condeelis, A. Patsialou, Chemotaxis in cancer, Nat. Rev. Cancer
11 (2011) 573–587.
[115] A.J. Trimboli, C.Z. Cantemir-Stone, F. Li, J.A. Wallace, A. Merchant, N. Creasap,
et al., Pten in stromal fibroblasts suppresses mammary epithelial tumours,
Nature 461 (2009) 1084–1091.
[116] X.Z. Wu, G.R. Xie, D. Chen, Hypoxia and hepatocellular carcinoma: the
therapeutic target for hepatocellular carcinoma, J. Gastroenterol. Hepatol. 22
(2007) 1178–1182.
[117] M. Jiao, K.-J. Nan, Activation of PI3 kinase/Akt/HIF-1α pathway contributes
to hypoxia-induced epithelial-mesenchymal transition and chemoresistance
in hepatocellular carcinoma, Int. J. Oncol. 40 (2012) 461–468.
[118] Y. Liang, T. Zheng, R. Song, J. Wang, D. Yin, L. Wang, et al., Hypoxia-mediated
sorafenib resistance can be overcome by EF24 through Von Hippel-Lindau
tumor suppressor-dependent HIF-1α inhibition in hepatocellular carcinoma,
Hepatology 57 (2013) 1847–1857.
[119] K. Shirabe, Y. Mano, J. Muto, R. Matono, T. Motomura, T. Toshima, et al., Role
of tumor-associated macrophages in the progression of hepatocellular
carcinoma, Surg. Today 42 (2012) 1–7.
[120] M.D. Bareford, M.A. Park, A. Yacoub, H.A. Hamed, Y. Tang, N. Cruickshanks,
et al., Sorafenib enhances pemetrexed cytotoxicity through an autophagy-
dependent mechanism in cancer cells, Cancer Res. 71 (2011) 4955–4967.
[121] S. Shimizu, T. Takehara, H. Hikita, T. Kodama, H. Tsunematsu, T. Miyagi, et al.,
Inhibition of autophagy potentiates the antitumor effect of the multikinase
inhibitor sorafenib in hepatocellular carcinoma, Int. J. Cancer 131 (2012)
548–557.
[122] C. Berasain, J. Castillo, M.J. Perugorria, M.U. Latasa, J. Prieto, M.A. Avila,
Inflammation and liver cancer: new molecular links, Ann. N. Y. Acad. Sci. 1155
(2009) 206–221.
[123] V. Pattullo, Hepatitis B reactivation in the setting of chemotherapy and
immunosuppression prevention is better than cure, World J. Hepatol. 7
(2015) 954–967.
[124] S. D’Angelo, M. Secondulfo, R. De Cristofano, P. Sorrentino, Sorafenib and
entecavir: the dioscuri of treatment for advanced hepatocellular carcinoma?,
World J. Gastroenterol. 19 (2013) 2141–2143.
[125] C. Zuo, M. Xia, Q. Wu, H. Zhu, J. Liu, C. Liu, Role of antiviral therapy in reducing
recurrence and improving survival in hepatitis B virus-associated
hepatocellular carcinoma following curative resection (Review), Oncol. Lett.
9 (2015) 527–534.
[126] V. Descamps, F. Helle, C. Louandre, E. Martin, E. Brochot, L. Izquierdo, et al.,
The kinase-inhibitor sorafenib inhibits multiple steps of the Hepatitis C Virus
infectious cycle in vitro, Antiviral Res. 118 (2015) 93–102.
[127] R. Cabrera, A.R. Limaye, P. Horne, R. Mills, C. Soldevila-Pico, V. Clark, et al.,
The anti-viral effect of sorafenib in hepatitis C-related hepatocellular
carcinoma, Aliment. Pharmacol. Ther. 37 (2013) 91–97.
[128] K. Himmelsbach, E. Hildt, The kinase inhibitor Sorafenib impairs the antiviral
effect of interferon alpha on hepatitis C virus replication, Eur. J. Cell Biol. 92
(2013) 12–20.
[129] Q.B. Zhang, H.C. Sun, K.Z. Zhang, Q.A. Jia, Y. Bu, M. Wang, et al., Suppression
of natural killer cells by sorafenib contributes to prometastatic effects in
hepatocellular carcinoma, PLoS ONE 8 (2013) e55945.
[130] K. Mao, J. Zhang, C. He, K. Xu, J. Liu, J. Sun, et al., Restoration of miR-193b
sensitizes Hepatitis B virus-associated hepatocellular carcinoma to sorafenib,
Cancer Lett. 352 (2014) 245–252.
[131] A.L. Cheng, Z. Guan, Z. Chen, C.J. Tsao, S. Qin, J.S. Kim, et al., Efficacy and safety
of sorafenib in patients with advanced hepatocellular carcinoma according
to baseline status: subset analyses of the phase III Sorafenib Asia-Pacific trial,
Eur. J. Cancer 48 (2012) 1452–1465.
[132] L. Bolondi, W. Caspary, J. Bennouna, W. Thomson, F. Van Steenbergen, M. Degos,
et al., Clinical benefit of sorafenib in hepatitis C patients with hepatocellular
carcinoma (HCC): subgroup analysis of the SHARP trial.[Abstract]. Presented
at the 2008 Gastrointestinal Cancers Symposium; Orlando, FL, 2008, pp. 25–
27.
[133] E. Rota Caremoli, R. Labianca, Tivantinib: critical review with a focus on
hepatocellular carcinoma, Expert Opin. Investig. Drugs 23 (2014) 1563–1574.
[134] K. Yamanaka, M. Petrulionis, S. Lin, C. Gao, U. Galli, S. Richter, et al., Therapeutic
potential and adverse events of everolimus for treatment of hepatocellular
carcinoma systematic review and meta-analysis, Cancer Med. 2 (2013)
862–871.
[135] W. Yeo, H.C. Chung, S.L. Chan, L.Z. Wang, R. Lim, J. Picus, et al., Epigenetic
therapy using belinostat for patients with unresectable hepatocellular
carcinoma: a multicenter phase I/II study with biomarker and pharmacokinetic
analysis of tumors from patients in the Mayo Phase II Consortium and the
Cancer Therapeutics Research Group, J. Clin. Oncol. 30 (2012) 3361–3367.
[136] V. Tovar, Cornella, A. Villanueva, FGF signaling is involved in acquired resistance
to sorafenib in an in vivo model of hepatocellular carcinoma, Presented at the
International Liver Cancer Association Conference, Hong Kong, China,
September 2–4, 2011 (abstr 0–006).
[137] J.M. Llovet, T. Decaens, J.L. Raoul, E. Boucher, M. Kudo, C. Chang, et al., Brivanib
in patients with advanced hepatocellular carcinoma who were intolerant to
sorafenib or for whom sorafenib failed: results from the randomized phase
III BRISK-PS study, J. Clin. Oncol. 31 (2013) 3509–3516.
[138] C. Cainap, S.K. Qin, W.T. Huang, I.J. Chung, H.M. Pan, Y. Cheng, et al., Phase III
trial of linifanib versus sorafenib in patients with advanced hepatocellular
carcinoma (HCC), J. Clin. Oncol. 31 (2013).
[139] P. Fang, J.H. Hu, Z.G. Cheng, Z.F. Liu, J.L. Wang, S.C. Jiao, Efficacy and safety of
bevacizumab for the treatment of advanced hepatocellular carcinoma: a
systematic review of phase II trials, PLoS ONE 7 (2012).
[140] C.J. Liu, J. Chang, P.H. Lee, D.Y. Lin, C.C. Wu, L.B. Jeng, et al., Adjuvant heparanase
inhibitor PI-88 therapy for hepatocellular carcinoma recurrence, World J.
Gastroenterol. 20 (2014) 11384–11393.
[141] C.M. Wang, A. Cigliano, S. Delogu, J. Armbruster, F. Dombrowski, M. Evert, et al.,
Functional crosstalk between AKT/mTOR and Ras/MAPK pathways in
hepatocarcinogenesis Implications for the treatment of human liver cancer,
Cell Cycle 12 (2013) 1999–2010.
[142] R. Gedaly, P. Angulo, J. Hundley, M.F. Daily, C. Chen, B.M. Evers, PKI-587 and
sorafenib targeting PI3K/AKT/mTOR and Ras/Raf/MAPK pathways
synergistically inhibit HCC cell proliferation, J. Surg. Res. 176 (2012) 542–548.
[143] A.X. Zhu, O. Rosmorduc, T.R. Evans, P.J. Ross, A. Santoro, F.J. Carrilho, et al.,
SEARCH: a phase III, randomized, double-blind, placebo-controlled trial of
sorafenib plus erlotinib in patients with advanced hepatocellular carcinoma,
J. Clin. Oncol. 33 (2015) 559–566.
[144] H. Huynh, V.C. Ngo, H.N. Koong, D. Poon, S.P. Choo, H.C. Toh, et al., AZD6244
enhances the anti-tumor activity of sorafenib in ectopic and orthotopic models
of human hepatocellular carcinoma (HCC), J. Hepatol. 52 (2010) 79–87.
[145] H.Y. Lim, J. Heo, H.J. Choi, C.Y. Lin, J.H. Yoon, C. Hsu, et al., A phase II study of
the efficacy and safety of the combination therapy of the MEK inhibitor
refametinib (BAY 86–9766) plus sorafenib for Asian patients with unresectable
hepatocellular carcinoma, Clin. Cancer Res. 20 (2014) 5976–5985.
[146] G. Brandi, F. de Rosa, V. Agostini, S. di Girolamo, P. Andreone, L. Bolondi, et al.,
Metronomic capecitabine in advanced hepatocellular carcinoma patients: a
phase II study, Oncologist 18 (2013) 1256–1257.
[147] J.S. Hammond, J. Franko, S.E. Holloway, J.T. Heckman, P.D. Orons, T.C. Gamblin,
Gemcitabine transcatheter arterial chemoembolization for unresectable
hepatocellular carcinoma, Hepatogastroenterology 61 (2014) 1339–1343.
[148] Y. Wang, Y. Liu, Y. Liu, W. Zhou, H. Wang, G. Wan, et al., A polymeric prodrug
of cisplatin based on pullulan for the targeted therapy against hepatocellular
carcinoma, Int. J. Pharm. 483 (2015) 89–100.
[149] W.W. Qi, H.Y. Yu, H. Guo, J. Lou, Z.M. Wang, P. Liu, et al., Doxorubicin-loaded
glycyrrhetinic acid modified recombinant human serum albumin nanoparticles
for targeting liver tumor chemotherapy, Mol. Pharm. 12 (2015) 675–683.
10 J. Chen et al./Cancer Letters 367 (2015) 1–11
[150] Y. Liu, H. Yue, S. Xu, F. Wang, N. Ma, K. Li, et al., First-line gemcitabine and
oxaliplatin (GEMOX) plus sorafenib, followed by sorafenib as maintenance
therapy, for patients with advanced hepatocellular carcinoma: a preliminary
study, Int. J. Clin. Oncol. (2015).
[151] Z. Wang, Z. Zhao, T. Wu, L. Song, Y. Zhang, Sorafenib-irinotecan sequential
therapy augmented the anti-tumor efficacy of monotherapy in hepatocellular
carcinoma cells HepG2, Neoplasma 62 (2015) 172–179.
[152] V. Srimuninnimit, V. Sriuranpong, S. Suwanvecho, Efficacy and safety of
sorafenib in combination with gemcitabine in patients with advanced
hepatocellular carcinoma: a multicenter, open-label, single-arm phase II study,
Asia Pac. J. Clin. Oncol. 10 (2014) 255–260.
[153] G.K. Abou-Alfa, P. Johnson, J.J. Knox, M. Capanu, I. Davidenko, J. Lacava, et al.,
Doxorubicin plus sorafenib vs doxorubicin alone in patients with advanced
hepatocellular carcinoma: a randomized trial, JAMA 304 (2010) 2154–2160.
[154] T. Hato, L. Goyal, T.F. Greten, D.G. Duda, A.X. Zhu, Immune checkpoint blockade
in hepatocellular carcinoma: current progress and future directions,
Hepatology 60 (2014) 1776–1782.
[155] Y. Wang, H. Li, Q. Liang, B. Liu, X. Mei, Y. Ma, Combinatorial immunotherapy
of sorafenib and blockade of programmed death-ligand 1 induces effective
natural killer cell responses against hepatocellular carcinoma, Tumour Biol.
(2014).
[156] Y. Chen, R.R. Ramjiawan, T. Reiberger, M.R. Ng, T. Hato, Y. Huang, et al., CXCR4
inhibition in tumor microenvironment facilitates anti-programmed death
receptor-1 immunotherapy in sorafenib-treated hepatocellular carcinoma in
mice, Hepatology 61 (2015) 1591–1602.
[157] V. Vaira, M. Roncalli, C. Carnaghi, A. Faversani, M. Maggioni, C. Augello, et al.,
MicroRNA-425-3p predicts response to sorafenib therapy in patients with
hepatocellular carcinoma, Liver Int. 35 (2015) 1077–1086.
[158] M. Scartozzi, L. Faloppi, G. Svegliati Baroni, C. Loretelli, F. Piscaglia, M. Iavarone,
et al., VEGF and VEGFR genotyping in the prediction of clinical outcome for
HCC patients receiving sorafenib: the ALICE-1 study, Int. J. Cancer 135 (2014)
1247–1256.
[159] X. Luo, G.S. Feng, VEGFA genomic amplification tailors treatment of HCCs with
sorafenib, Cancer Discov. 4 (2014) 640–641.
[160] R. Rudalska, D. Dauch, T. Longerich, K. McJunkin, T. Wuestefeld, T.W. Kang, et al.,
In vivo RNAi screening identifies a mechanism of sorafenib resistance in liver
cancer, Nat. Med. 20 (2014) 1138–1146.
[161] D. Chen, P. Zhao, S.Q. Li, W.K. Xiao, X.Y. Yin, B.G. Peng, et al., Prognostic impact
of pERK in advanced hepatocellular carcinoma patients treated with sorafenib,
Eur. J. Surg. Oncol. 39 (2013) 974–980.
[162] W.T. Tai, A.L. Cheng, C.W. Shiau, H.P. Huang, J.W. Huang, P.J. Chen, et al., Signal
transducer and activator of transcription 3 is a major kinase-independent
target of sorafenib in hepatocellular carcinoma, J. Hepatol. 55 (2011) 1041–
1048.
[163] M.J. Blivet-Van Eggelpoel, H. Chettouh, L. Fartoux, L. Aoudjehane, V. Barbu,
C. Rey, et al., Epidermal growth factor receptor and HER-3 restrict cell response
to sorafenib in hepatocellular carcinoma cells, J. Hepatol. 57 (2012) 108–
115.
[164] J.J. Shan, J.J. Shen, L.M. Liu, F. Xia, C. Xu, G.J. Duan, et al., Nanog regulates self-
renewal of cancer stem cells through the insulin-like growth factor pathway
in human hepatocellular carcinoma, Hepatology 56 (2012) 1004–1014.
[165] S. Hagiwara, M. Kudo, T. Nagai, T. Inoue, K. Ueshima, N. Nishida, et al., Activation
of JNK and high expression level of CD133 predict a poor response to sorafenib
in hepatocellular carcinoma, Br. J. Cancer 106 (2012) 1997–2003.
[166] J.M. Llovet, C.E.A. Pena, C.D. Lathia, M. Shan, G. Meinhardt, J. Bruix, et al., Plasma
biomarkers as predictors of outcome in patients with advanced hepatocellular
carcinoma, Clin. Cancer Res. 18 (2012) 2290–2300.
[167] T. Arao, K. Ueshima, K. Matsumoto, T. Nagai, H. Kimura, S. Hagiwara, et al.,
FGF3/FGF4 amplification and multiple lung metastases in responders to
sorafenib in hepatocellular carcinoma, Hepatology 57 (2013) 1407–1415.
11J. Chen et al./Cancer Letters 367 (2015) 1–11
... Nevertheless, around 30% of individuals exhibit an inadequate response to sorafenib or acquire resistance following several months of treatment [6]. Multiple mechanisms, such as the tumor microenvironment, autophagy, epithelial-mesenchymal transition, and cancer stem cells, have been indicated as factors contributing to sorafenib resistance in HCC [7]. Moreover, given that sorafenib is among the limited number of tyrosine kinase inhibitors capable of triggering ferroptosis, there is a growing emphasis on understanding the resistance mechanism to sorafenib-induced ferroptosis in HCC [8,9]. ...
Article
Full-text available
Background Advanced hepatocellular carcinoma (HCC) can be treated with sorafenib, which is the primary choice for targeted therapy. Nevertheless, the effectiveness of sorafenib is greatly restricted due to resistance. Research has shown that exosomes and circular RNAs play a vital role in the cancer’s malignant advancement. However, the significance of exosomal circular RNAs in the development of resistance to sorafenib in HCC remains uncertain. Methods Ultracentrifugation was utilized to isolate exosomes (Exo-SR) from the sorafenib-resistant HCC cells’ culture medium. Transcriptome sequencing and differential expression gene analysis were used to identify the targets of Exo-SR action in HCC cells. To identify the targets of Exo-SR action in HCC cells, transcriptome sequencing and analysis of differential expression genes were employed. To evaluate the impact of exosomal circUPF2 on resistance to sorafenib in HCC, experiments involving gain-of-function and loss-of-function were conducted. RNA pull-down assays and mass spectrometry analysis were performed to identify the RNA-binding proteins interacting with circUPF2. RNA immunoprecipitation (RIP), RNA pull-down, electrophoretic mobility shift assay (EMSA), immunofluorescence (IF) -fluorescence in situ hybridization (FISH), and rescue assays were used to validate the interactions among circUPF2, IGF2BP2 and SLC7A11. Finally, a tumor xenograft assay was used to examine the biological functions and underlying mechanisms of Exo-SR and circUPF2 in vivo. Results A novel exosomal circRNA, circUPF2, was identified and revealed to be significantly enriched in Exo-SR. Exosomes with enriched circUPF2 enhanced sorafenib resistance by promoting SLC7A11 expression and suppressing ferroptosis in HCC cells. Mechanistically, circUPF2 acts as a framework to enhance the creation of the circUPF2-IGF2BP2-SLC7A11 ternary complex contributing to the stabilization of SLC7A11 mRNA. Consequently, exosomal circUPF2 promotes SLC7A11 expression and enhances the function of system Xc- in HCC cells, leading to decreased sensitivity to ferroptosis and resistance to sorafenib. Conclusions The resistance to sorafenib in HCC is facilitated by the exosomal circUPF2, which promotes the formation of the circUPF2-IGF2BP2-SLC7A11 ternary complex and increases the stability of SLC7A11 mRNA. Focusing on exosomal circUPF2 could potentially be an innovative approach for HCC treatment. Graphical Abstract
... In the early stage of liver cancer, specific clinical symptoms are often absent, leading to missed opportunities for early surgical intervention upon diagnosis [3,4]. Sorafenib, approved by the U.S. Food and Drug Administration (FDA) in 2007, was the first oral multikinase inhibitor for treating advanced HCC and remains the most commonly prescribed first-line chemotherapy for advanced liver cancer patients [5,6]. However, in clinical practice, many patients receiving sorafenib have quickly developed resistance [7]. ...
... This signaling also promotes the invasiveness and motility of cancer cells (Neuzillet et al. 2014). Moreover, B-RAF triggers angiogenesis via HIF-1α and VEGF, while C-RAF (RAF-1) ISSN: 2708-7182 (Print) Chen et al. 2015). ...
Article
Full-text available
Hepatocellular carcinoma (HCC) is a severe and increasingly prevalent health issue affecting individuals globally. Recent research endeavors in the clinical domains have lately focused more on the MAPK signaling pathway in HCC. Activating mutations in the RAS and RAF genes, which greatly activate the MAPK pathway in malignancies, are rare in HCC patients, yet over 50% of them have activated the pathway. This suggests that other factors may be responsible for the activation of the signaling pathway in HCC. MAPK signaling is important to carcinogenesis, and it is often altered in human cancers. The drug resistance in targeted therapy against RTKs in HCC may arise from mutations in downstream components (RAS, RAF, MEK, ERK), resistant mutations within RTKs, and additional alternative pathways like PI3K and YAP may also develop the resistance. Epigenetic processes and chromatin remodeling are crucial to pharmacological tolerance to MAPK regulation. This review will focus on the latest developments in our knowledge of the cellular and molecular processes to activate the MAPK signaling pathway, as well as possible treatment approaches that specifically target this pathway in relation to HCC. The study also investigates the clinical efficacy of molecular-targeted treatments, including tyrosine kinase inhibitors and immunological checkpoint inhibitors and highlights the use of combination therapy for HCC.
Article
Liver cancer remains one of the most prevalent malignancies worldwide with high incidence and mortality rates. Due to its subtle onset, liver cancer is commonly diagnosed at a late stage when surgical interventions are no longer feasible. This situation highlights the critical role of systemic treatments, including targeted therapies, in bettering patient outcomes. Despite numerous studies on the mechanisms underlying liver cancer, tyrosine kinase inhibitors (TKIs) are the only widely used clinical inhibitors, represented by sorafenib, whose clinical application is greatly limited by the phenomenon of drug resistance. Here we show an in-depth discussion of the signaling pathways frequently implicated in liver cancer pathogenesis and the inhibitors targeting these pathways under investigation or already in use in the management of advanced liver cancer. We elucidate the oncogenic roles of these pathways in liver cancer especially hepatocellular carcinoma (HCC), as well as the current state of research on inhibitors respectively. Given that TKIs represent the sole class of targeted therapeutics for liver cancer employed in clinical practice, we have particularly focused on TKIs and the mechanisms of the commonly encountered phenomena of its resistance during HCC treatment. This necessitates the imperative development of innovative targeted strategies and the urgency of overcoming the existing limitations. This review endeavors to shed light on the utilization of targeted therapy in advanced liver cancer, with a vision to improve the unsatisfactory prognostic outlook for those patients.
Article
Hepatocellular carcinoma (HCC), one of the most prevalent and destructive causes of cancer‐related deaths worldwide, approximately 70% of patients with HCC exhibit advanced disease at diagnosis, limiting the potential for radical treatment. For such patients, lenvatinib, a long‐awaited alternative to sorafenib for first‐line targeted therapy, has become a key treatment. Unfortunately, despite some progress, the prognosis for advanced HCC remains poor because of drug resistance development. However, the molecular mechanisms underlying lenvatinib resistance and ways to relief drug resistance in HCC are largely unknown and lack of systematic summary; thus, this review not only aims to explore factors contributing to lenvatinib resistance in HCC, but more importantly, summary potential methods to conquer or mitigate the resistance. The results suggest that abnormal activation of pathways, drug transport, epigenetics, tumour microenvironment, cancer stem cells, regulated cell death, epithelial–mesenchymal transition, and other mechanisms are involved in the development of lenvatinib resistance in HCC and subsequent HCC progression. To improve the therapeutic outcomes of lenvatinib, inhibiting acquired resistance, combined therapies, and nano‐delivery carriers may be possible approaches.
Article
Signal transducer and activator of transcription 3 (STAT3) plays an important role in the occurrence and progression of tumors, leading to resistance and poor prognosis. Activation of STAT3 signaling is frequently detected in hepatocellular carcinoma (HCC), but potent and less toxic STAT3 inhibitors have not been discovered. Here, based on antisense technology, we designed a series of stabilized modified antisense oligonucleotides targeting STAT3 mRNA (STAT3 ASOs). Treatment with STAT3 ASOs decreased the STAT3 mRNA and protein levels in HCC cells. STAT3 ASOs significantly inhibited the proliferation, survival, migration, and invasion of cancer cells by specifically perturbing STAT3 signaling. Treatment with STAT3 ASOs decreased the tumor burden in an HCC xenograft model. Moreover, aberrant STAT3 signaling activation is one of multiple signaling pathways involved in sorafenib resistance in HCC. STAT3 ASOs effectively sensitized resistant HCC cell lines to sorafenib in vitro and improved the inhibitory potency of sorafenib in a resistant HCC xenograft model. The developed STAT3 ASOs enrich the tools capable of targeting STAT3 and modulating STAT3 activity, serve as a promising strategy for treating HCC and other STAT3-addicted tumors, and alleviate the acquired resistance to sorafenib in HCC patients. A series of novel STAT3 antisense oligonucleotide were designed and showed potent anti-cancer efficacy in hepatocellular carcinoma in vitro and in vivo by targeting STAT3 signaling. Moreover, the selected STAT3 ASOs enhance sorafenib sensitivity in resistant cell model and xenograft model.
Article
Aim: Investigation of the pharmacokinetics of sorafenib (SRF) in rats with hepatocellular carcinoma (HCC). Methods: A reproducible ultra-HPLC–MS method for simultaneous determination of serum SRF, N-hydroxymethyl sorafenib and N-demethylation sorafenib. Results: Both the maximum serum concentrations (2.5-times) and the area under the serum concentration–time curve from 0 h to infinity (4.5-times) of SRF were observed to be significantly higher, with a greater than 3.0-fold decrease in the clearance rate in the HCC-bearing rats compared with these values in healthy animals. Further study revealed approximately 3.8- and 3.2-times increases in the apparent Michaelis constant for N-hydroxymethyl sorafenib and N-demethylation sorafenib conversions in the HCC-bearing rats. Conclusion: The low efficiency for the SRF conversions was a key contributor to the increased serum concentrations of SRF.
Article
No effective systemic therapy exists for patients with advanced hepatocellular carcinoma. A preliminary study suggested that sorafenib, an oral multikinase inhibitor of the vascular endothelial growth factor receptor, the platelet-derived growth factor receptor, and Raf may be effective in hepatocellular carcinoma. Methods In this multicenter, phase 3, double-blind, placebo-controlled trial, we randomly assigned 602 patients with advanced hepatocellular carcinoma who had not received previous systemic treatment to receive either sorafenib (at a dose of 400 mg twice daily) or placebo. Primary outcomes were overall survival and the time to symptomatic progression. Secondary outcomes included the time to radiologic progression and safety. Results At the second planned interim analysis, 321 deaths had occurred, and the study was stopped. Median overall survival was 10.7 months in the sorafenib group and 7.9 months in the placebo group (hazard ratio in the sorafenib group, 0.69; 95% confidence interval, 0.55 to 0.87; P
Article
249 Background: Linifanib (ABT-869; Lin) is a potent and selective inhibitor of the vascular endothelial growth factor (VEGF) and platelet-derived growth factor (PDGF) receptor tyrosine kinase families. In a phase II trial in patients (pts) with advanced HCC, Lin showed clinical activity (objective response rate [ORR] 10.5% in Child-Pugh A [CPA] pts). This open-label, global phase 3 trial evaluated Lin versus sorafenib (Sor) as first-line therapy in pts with advanced CPA HCC (NCT01009593). Methods: Pts were randomized 1:1 to Lin 17.5 mg QD or Sor 400 mg BID and stratified by region (non-Asia/Japan/rest of Asia), ECOG performance status (0/1), vascular invasion or extrahepatic spread (yes/no) and HBV infection (yes/no). The primary efficacy endpoint was overall survival (OS); both non-inferiority (margin 1.0491) and superiority hypotheses were to be tested. Secondary efficacy endpoints included time to progression (TTP) and ORR, using RECIST v1.1. AE severity was graded using NCI-CTCAE v4.0. Results: 1035 ...
Conference Paper
Tumor progression is coincident with mechanochemical changes in the extracellular matrix (ECM). We hypothesized that tumor stroma stiffening, alongside a shift in the ECM composition from a basement membrane-like microenvironment toward a dense network of collagen-rich fibers during tumorigenesis, confers resistance to otherwise powerful chemotherapeutics. We created a high-throughput drug screening platform based on our poly(ethylene glycol)-phosphorylcholine (PEG-PC) hydrogel system, and customized it to capture the stiffness and integrin-binding profile of in vivo tumors. We report that the efficacy of a Raf kinase inhibitor, sorafenib, is reduced on stiff, collagen-rich microenvironments, independent of ROCK activity. Instead, sustained activation of JNK mediated this resistance, and combining a JNK inhibitor with sorafenib eliminated stiffness-mediated resistance in triple negative breast cancer cells. Overall, we discovered that β1 integrin and its downstream effector JNK mediate sorafenib resistance during tumor stiffening. These results also highlight the need for more advanced cell culture platforms, such as our high-throughput PEG-PC system, with which to screen chemotherapeutics.