Next Article in Journal
Wnt Signaling in Neural Crest Ontogenesis and Oncogenesis
Next Article in Special Issue
ATM Regulated PTEN Degradation Is XIAP E3 Ubiquitin Ligase Mediated in p85α Deficient Cancer Cells and Influence Platinum Sensitivity
Previous Article in Journal
Inhibition of TOR in Chlamydomonas reinhardtii Leads to Rapid Cysteine Oxidation Reflecting Sustained Physiological Changes
Previous Article in Special Issue
Skeletal Muscle Response to Deflazacort, Dexamethasone and Methylprednisolone
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mitogen-Activated Protein Kinases (MAPKs) and Cholangiocarcinoma: The Missing Link

by
Chaobo Chen
1,2,3,
Leonard J. Nelson
4,
Matías A. Ávila
5,6 and
Francisco Javier Cubero
1,2,*
1
Department of Immunology, Ophthalmology & ENT, Complutense University School of Medicine, 28040 Madrid, Spain
212
de Octubre Health Research Institute (imas12), 28040 Madrid, Spain
3
Department of General Surgery, Wuxi Xishan People’s Hospital, Wuxi 214000, China
4
Institute for Bioengineering (IBioE), School of Engineering, Faraday Building, The University of Edinburgh, Edinburgh EH9 3 JL, Scotland, UK
5
Hepatology Program, Center for Applied Medical Research (CIMA), University of Navarra, 31008 Pamplona, Spain
6
Centro de Investigacion Biomedica en Red, Enfermedades Hepáticas y Digestivas (CIBERehd), 28029 Madrid, Spain
*
Author to whom correspondence should be addressed.
Cells 2019, 8(10), 1172; https://doi.org/10.3390/cells8101172
Submission received: 28 August 2019 / Revised: 18 September 2019 / Accepted: 25 September 2019 / Published: 28 September 2019
(This article belongs to the Special Issue Protein Kinases: Signalling and Disease)

Abstract

:
In recent years, the incidence of both liver and biliary tract cancer has increased. Hepatocellular carcinoma (HCC) and cholangiocarcinoma (CCA) are the two most common types of hepatic malignancies. Whereas HCC is the fifth most common malignant tumor in Western countries, the prevalence of CCA has taken an alarming increase from 0.3 to 2.1 cases per 100,000 people. The lack of specific biomarkers makes diagnosis very difficult in the early stages of this fatal cancer. Thus, the prognosis of CCA is dismal and surgery is the only effective treatment, whilst recurrence after resection is common. Even though chemotherapy and radiotherapy may prolong survival in patients with CCA, the 5-year survival rate is still very low—a significant global problem in clinical diagnosis and therapy. The mitogen-activated protein kinase (MAPK) pathway plays an important role in signal transduction by converting extracellular stimuli into a wide range of cellular responses including inflammatory response, stress response, differentiation, survival, and tumorigenesis. Dysregulation of the MAPK cascade involves key signaling components and phosphorylation events that play an important role in tumorigenesis. In this review, we discuss the pathophysiological role of MAPK, current therapeutic options, and the current situation of MAPK-targeted therapies in CCA.

1. Introduction

Cholangiocarcinoma (CCA)

Cholangiocarcinoma (CCA) is a malignant adenocarcinoma characterized by biliary tract differentiation and is the second most common primary liver tumor. Although CCA in most countries is considered a rare form of cancer with an incidence rate of <6 cases per 100,000 of the population [1,2,3], the global incidence of CCA (1/100,000) is steadily on the rise, reaching up to 2.1/100,000 in Western countries, and having increased almost 165% in the United States over the last 30 years period [4,5,6,7]. Risk factors for CCA include chronic viral hepatitis (types B and C), liver cirrhosis, biliary stone disease, chronic infections (infection with Clonorchis sinensis, Opisthorchis viverrini, and Opisthorchis felineus species in Southeast Asia), metabolic syndrome, obesity, and toxic substances (e.g., tobacco and alcohol). In addition, other factors such as parasitic infections, hepatolithiasis, Caroli’s disease, and primary sclerosing cholangitis (PSC) are identified risk factors [8]. Although hepatolithiasis is more commonly found in Asia, it has also been associated with CCA in Western populations [9,10]. In addition, CCA development has also been linked to genetic variants such as DNA repair, inflammation, and carcinogen metabolism [3].
Depending on the anatomical location, CCA can be basically divided into three categories: intrahepatic (iCCA), perihilar (pCCA), and distal (dCCA) [5,11]. The above-mentioned elevation in global CCA incidence is mainly due to the increase in iCCA cases [12]. It is noteworthy that although these three subtypes share common features, they also have important inter- and intra-tumoral differences that may affect the pathogenesis and outcome of the disease [13,14,15,16]. According to a recent study, by using integrated molecular analysis research methods, CCA can be divided into two categories: inflammatory or proliferative. The former is characterized by the activation of inflammatory signaling pathways, overexpression of cytokines, and the activation of the signal transducer and activator of the transcription 3 (STAT3) pathway, which is characterized by the activation of oncogenic signaling pathways including RAS, mitogen-activated protein kinases (MAPKs), and epithelial to mesenchymal transition (EMT) [17]. Consequently, a number of important signaling pathways involved in CCA have been discovered including RAF/mitogen-activated protein kinase (MEK)/MAPK, transforming growth factor-β (TGF-β)/SMADs, phosphatidylinositol 3-kinase (PI3K)/AKT, WNT, interleukin-6 (IL-6)/STAT-3, and NOTCH [18,19]. The MAPKs not only play an important role in the initiation and formation of CCA, but are also closely linked to many CCA-related signaling pathways, as above-mentioned.

2. The MAPK Signaling Pathways: ERK1/2, JNK-1/2/3, and p38

There are at least three different MAPK signal transduction pathways that modulate and transduce extracellular signals into the nucleus to induce response genes in mammalian cells including ERK1/2, c-Jun NH2-terminal kinase(JNK)1/2/3, and p38 [20,21]. The ERK kinase family is composed of ERK1 (p44) and ERK2 (p42). The JNK kinase family is composed of JNK1, JNK2, and JNK3. Finally, the p38 MAPKs family can be further divided into two subgroups, p38α and p38β, and p38γ and p38δ, respectively [22,23]. In addition, there are other atypical MAPKs that have unique regulation and function including ERK3/4, ERK7/8, and Nemo-like kinase (NLK) [20,21,24,25].
Growth factor-initiated signaling is associated with the ERK pathway, whereas the JNK and p38 pathways are activated by cytokines, growth factors, environmental stress, and other stimuli (Figure 1). In the p38 pathway, MKK3/6 acts as mitogen-activated protein kinase kinases (MAP2Ks) and is activated by mitogen-activated protein kinase kinase kinases (MAP3Ks) such as apoptosis signal-regulating kinase 1 (ASK1), transforming growth factor-β-activated kinase 1 (TAK1), mitogen-activated protein kinase kinase kinase 1 (MEKK1), and mitogen-activated protein kinase kinase kinase 11 (MLK3). These MAP3Ks also play a role in the JNK pathway, where they target MAP2Ks such as MAP kinase 4 (MKK4)/MAP kinase 7 (MKK7), which activate JNK1/2. In the ERK signaling pathway, ERK1/2 is activated by MEK1/2 and activated by RAF, which is activated by RAS, and RAS is related to the plasma membrane by activation of receptor tyrosine kinases (RTKs), epidermal growth factor receptor (EGFR), vascular endothelial growth factor receptor (VEGFR), and platelet-derived growth factor receptor (PDGFR) [20,26].

2.1. JNK/MAPKs

In normal liver, the JNK family is minimally or transiently activated, the latter usually physiological, whereas sustained activation is pathological. In addition to their central role in hepatic physiological responses (such as cell death, apoptosis, and cell cycle regulation) [27,28], JNK plays a carcinogenic role by promoting inflammation, proliferation, invasion, and angiogenesis, depending on the specific context and duration of the activation of the JNK signaling pathway [29,30]. Earlier studies have reported that the JNK signaling pathway plays a significant role in the development of HCC [31,32]. In particular, Jnk1 seems to be more important in malignant transformation and hepatocellular carcinoma (HCC) development [33,34]. However, during the progression of liver disease, the deletion of Jnk1 significantly exacerbated apoptosis, compensatory proliferation, and carcinogenesis in experimental chronic liver disease. In turn, Jnk2 was reported to modulate necrosis and inflammation [35]. Using Puma KO mice in diethylnitrosamine (DEN)-induced HCC, HCC formation was suppressed, a phenomenon associated with reduced cell death and compensatory proliferation of hepatocytes. However, such effects were inhibited by the administration of the JNK inhibitor SP600125 [36]. In human HCC cell lines, Jnk1 activation is associated with tumor size. By knocking down Jnk1 but not Jnk2, proliferation of human HCC occurs via upregulation of cmyc and downregulation of p21 [37].
The JNK signaling pathway has been recently implicated in CCA development. JNK plays an important role in regulating the interaction between different pro-apoptotic and anti-apoptotic proteins in response to both external and internal apoptotic stimuli [38]. In human CCA, a high expression of TNFA in cells near CCA lesions was found and phosphorylation of JNK in cholangiocytes (80% in CCA patients) as well as the accumulation of reactive oxygen species (ROS) around the peripheral hepatocytes occurred. Accumulation of ROS leads to caspase-3-dependent apoptosis via JNK. At the same time, Yuan et al. [39], showed that mitochondrial dysfunction and oxidative stress promoted cholangiocellular overgrowth and tumorigenesis. In this study, cholangiocellular proliferation and oncogenic transformation were promoted through the JNK signaling pathway, which was activated by Kupffer cells (KC), subsequently leading to liver damage, ROS, and the paracrine release of TNF. Therefore, KC-derived TNF might promote cholangiocellular cell proliferation, differentiation, and carcinogenesis, suggesting that the ROS/TNF/JNK axis plays a significant role in the development of CCA. Another study that tested the antioxidant activity of using guggulsterone, a steroid found in the resin of the guggul plant (Commiphora wightii) in both HuCC-T1 and RBE CCA cell lines, showed that guggulsterone could induce the apoptosis of human CCA cells via ROS-mediated activation of the JNK signaling pathway [40]. Feng et al. [41] reported that JNK exerted its carcinogenic effect in human CCA cells, partially because the mammalian target of the rapamycin (mTOR) pathway was regulated by the induction of glucose-regulated protein 78 (GRP78). They showed that eukaryotic initiation factor-α (eIF2a)/activated transcription factor 4 (ATF4) signaling promoted the accumulation of GRP78; whilst JNK was both promoted by the activation of mTOR and characterized by a high expression of GRP78. These results indicated that GRP78 contributed to the pro-tumorigenic function of JNK in human CCA cells, partly through promoting the eIF2α/ATF4/GRP78 pathway, and through JNK/mTOR signaling. As above-mentioned, the JNK/MAPKs signaling pathway may be a key target in the treatment of CCA; however, more studies are needed to test this possibility.

2.2. p38/MAPKs

The p38 kinase plays a key role by promoting proliferation, invasion, inflammation, and angiogenesis in the occurrence and development of CCA [42,43,44,45], also affecting the growth of malignant human CCA cells, and maintaining the phenotype of transformed cells [45,46]. p38δ can be used for differential diagnosis of CCA, since it is not expressed in HCC cells [47]. IL-6 receptors and tyrosine kinases receptors such as Met (c-MET) and the EGFR family members ERBB2 and ERBB1 (EGFR) are key signaling pathways in CCA. Members of the EGFR family, particularly EGFR and ERBB2 (HER-2/neu), are involved in the pathogenesis of CCA [15,48,49]. Dai et al. [50] reported that both p38 and C-MET, the tyrosine kinase receptor for hepatocyte growth factor (HGF), promote the proliferation and invasion of human CCA cells, and p38 promotes CCA formation via sustained activation of C-MET. In addition, using ATO (arsenic trioxide) alone or in combination with metformin to treat CCA cell lines, Ling et al. [42] found that inactivation of p38 by the inhibitor SB203580 or specific siRNA could enhance the anticancer efficacy of single drug or combination of metformin and ATO, especially using ATO alone. Furthermore, the expression of P38δ was upregulated in CCA when compared to normal biliary tissue. Inhibition of P38δ expression by siRNA transfection significantly reduced CCA cell proliferation and invasion. Conversely, overexpression of P38δ in CCA cells can induce an increase in tumor invasiveness.

2.3. ERK/MAPKs

Previous studies have reported an upregulation of the RAS-ERK1/2 signaling pathway in CCA [51]. TGF-β transmits signals through SMAD2/3 phosphorylation, but can also transduce signals via RAS-ERK1/2, PI3K, p38, and Rho pathways [52,53,54]. Rac1, a member of the Rho family of small GTPases, plays a pivotal role in the development of CCA [55,56]. Li et al. [57] demonstrated that integrin β6 promoted tumor cell malignant behavior by activating Rac1. In biliary dysplasia, hyperplasia and CCA, TGF-β is overexpressed, a mechanism associated with CCA initiation and formation [51,58,59]. Moreover, ERK1/2 activation has also been linked with TGF-β-induced epithelial to mesenchymal transition (EMT) and cell invasiveness. One study using CCA cell lines indicated that TGF-β enhanced cell invasiveness and mesenchymal features in CCA cell lines [58]. Recently, a study using CCA cell lines (KKU-M213 and HuCCA-1) showed that TGF-β activates ERK1/2 via the SMAD2/3 pathway, enhancing TGF-β activity to promote tumor growth [60]. As above-mentioned, TGF-β can also transduce signals via the RAS-ERK/PI3K pathways [52,53,54]. The PI3K/AKT/PTEN signaling pathway is significantly overexpressed in human CCA tissue [61], therefore it is another important pathway involved in the development of CCA associated with RAS/ERK/MAPKs. PI3K is one of the most important factors in RAS activation and can be involved in the regulation of various functions including cell growth, cell cycle entry, cell survival, cytoskeleton reorganization, and apoptosis [15,62,63] (Figure 1). Loss of phosphatase function of PTEN will result in the constitutive activation of (PI3K/protein kinase B) AKT signaling pathway in CCA. Inhibition of PI3K/AKT signaling effectively blocked the proliferation and invasive behavior of CCA [61]. At the same time, Hyunho et al. [64] using CCA cell lines (SCK and Choi-CK) with inactivation of AKT, showed decreased expression of BCL2, and enhanced expression of BAX, thereby inducing the apoptosis of resistant cells; whilst the inhibition of ERK1/2 activation did not induce apoptosis, but decreased tumor cell growth. These results indicated that the AKT/ERK1/2 signaling transduction pathway might mediate apoptosis in CCA cell lines. Furthermore, in CCA, the expression of the HGF receptor (HGFR) encoded by the MET gene, also known as cMET, is increased (12–58%). Activation of MET promotes cell invasion and triggers metastasis by directly participating in tumor angiogenesis [65]. Experiments with CCA cell lines indicate that HGFR-dependent CCA cell invasiveness occurs via the AKT/ERK signaling pathway [66,67,68]. In summary, abundant evidence suggests that the ERK signaling pathway is intimately related to cholangiocarcinogenesis. In the future, the ERK signaling pathway may be an effective target for the diagnosis and treatment of CCA; however, more research is needed to confirm this.

2.4. ‘Other’ Kinases

Kinases such as transforming growth factor-β-activated kinase 1 (TAK1) belong to the MAP3Ks family. TAK1 is activated in response to cytokines such as TNF, LPS, and TGF-β in multiple cellular systems [69]. Activation of the TGF-β receptor triggers downstream signaling mediated by SMAD family of proteins, which promotes EMT progression in CCA (discussed below). However, phosphorylated TAK1 activates IKK (IκB kinase) and MKK4/7, causing the activation of NF-κB and JNK. Deletion of TAK1 in liver parenchymal cells (both hepatocytes and cholangiocytes) promotes hepatocyte death, inflammation, fibrosis, and hepatocarcinogenesis, coinciding with biliary ductopenia and cholestasis [70,71,72]. ASK1 is another member of the MAPK family, known as MAP3K5 (mitogen-activated protein kinase kinase kinase 5), which can activate the P38 and JNK pathways in response to both oxidative and ER (endoplasmic reticulum) stress as well as stress-induced via inflammatory cytokines (such as TNFα) [73,74,75].

3. EMT Transition

Previously, it was generally believed that HCC originated from hepatocytes and that CCA originates from bile duct cells [62]. Moreover, biliary epithelial cells (BECs) may be involved in liver fibrosis [76]. During acute and chronic cholestasis, liver fibrosis involving an increasing number of ductules (the finest ramifications of the biliary tree), is accompanied by polymorphonuclear leukocyte infiltration and increased extracellular matrix deposition, leading to periportal fibrosis and finally PBC (primary biliary cholangitis) [77]. Diaz et al. [78] presented histological data suggesting that EMT occurs in human liver fibrosis, particularly in diseases associated with prominent bile ductular proliferation including biliary atresia and PBC. Other studies indicated that proliferating cholangiocytes play a key role in the induction of fibrosis via EMT [76,79,80]. In CCA, developmental changes seen in cholangiocytes may be due to the activation of hepatic stem/progenitor cells, the proliferation of bile duct cells, or through ductal metaplasia of mature hepatocytes [81,82]. However, alongside the progress seen in histopathological studies, there is now a better understanding of the origins of BECs. Both hepatic progenitor cells (HPCs) [83] and CCA stem cells [62] seem to be capable of differentiating into BECs (Figure 2), whilst CCA stem cells can also originate from HPCs [62,84,85]. Micrographs of in vivo fate-tracking experiments showed the contribution of hepatocytes and HPCs to bile duct repair or hepatocyte injury. In order to cope with biliary tract injury, signals derived from myofibroblasts mediate the activation and proliferation of the HPCs niche. This process can be triggered by NOTCH signal transduction, via JAG1, a NOTCH ligand provided by the stroma of the portal vein, which induces the differentiation of hepatocytes into BECs [86,87,88]. These results indicate that CCA can be derived from common HPCs, which can differentiate into both hepatocytes and cholangiocytes [89,90].
In addition, chronic inflammatory processes can induce the production of a variety of cytokines including TNF, IL-6, TGF-β, and PDGF, promoting the development of CCA by affecting BEC function and proliferation [18,19]. In other words, EMT may also play an important role in CCA progression.
In 1968, Hay [91] noted that during chick embryo development, epithelial cells underwent through several differentiation and dedifferentiation stages and migrated a considerable distance in the body. All of these processes require an interconversion between the epithelial and mesenchymal cell phenotypes (including ‘multipotent’, mesenchymal stem cells) called EMT, and the opposite process, mesenchymal–epithelial transformation (MET). EMT is a special physiological process in which epithelial cells lose polarity and exhibit a physiological process of mesenchymal phenotype. Subsequently, as research progresses, multiple in vivo studies have also been reported. EMT is critical for embryonic development, wound healing, stem cell biology, and plays a key role in the tumorigenic process [92,93]. The first change occurs when epithelial cells lose cell–cell adhesions (at specialized adherent junctions) with the disruption of two major components E-cadherin and β-catenin. Tight junction fractures also occur and result in the loss of apical-basal polarity including cell-extracellular matrix (ECM) contacts. EMT is now known to be mediated by the so-called EMT-activated transcription factors (EMT-ATFs) including SNAIL, TWIST, and ZEB families [92]. EMT-ATFs are key mediators of cell plasticity, playing an important role in all stages of cancer progression from initiation, primary tumor growth, invasion, dissemination, and metastasis to colonization including resistance to therapy [92,94]. The process from primary tumor to metastasis is very complex, and cancer cells need to adapt to multiple and changing environments, which are usually harsh. This plasticity of tumor cells is reflected by the so called back-and-forth transition from differentiation to undifferentiation or partial EMT-associated cancer cell phenotypes, which promote the formation of tumor cells and facilitate their metastasis [95]. In line with CCA development, at the biochemical level, various morphogenetic and environmental signals including TGF-β, WNT, EGFR, and PDGF, inflammatory cytokines and integrin receptor ligands promote EMT [96,97] (Figure 3). Furthermore, TGF-β is also involved in several steps of cancer from tumorigenesis to metastasis, which promotes cancer cell invasion and metastasis through an EMT process, although at later stages [98,99,100,101,102].
The canonical TGF-β pathway relies on the downstream SMAD family [103], which are subsequently phosphorylated after receiving signals from membrane receptors (e.g., EGFR, TNFR, and ILR). The two phosphorylated R-SMADs are then brought together, recruiting SMAD (co-SMAD) and SMAD4 to form a trimer that can be transported from the cytosol to the nucleus [104]. In the nucleus, the heterotrimeric complex binds to DNA in a sequence-specific manner alone or in combination with other coactivators to trigger the transcription of target genes such as SNAIL1, SNAIL2, and other oncogenes [105,106]. The formation of protein complexes between phosphorylated R-SMADs and SMAD4 is a central event in the TGF-β signaling pathway including apoptosis, cell growth and differentiation, and extracellular matrix production [107,108]. Activation of the TGF-β-induced SMADs independent pathway (i.e., the non-canonical pathway) is achieved via TAK1 and JNK/P38 signaling. Interestingly, JNK induced phosphorylation of R-SMAD can activate the EMT programming process. Both JNK and p38 can synergize with SMADs to promote non-canonical TGF-β-induced apoptosis [96,109,110,111]. In SMAD-independent pathways TGF-β can also induce AKT phosphorylation and rapidly activate PI3K, which is possibly involved in SMAD2-induced EMT [112]. Studies on keratinocytes indicate that the PI3K/AKT pathway contributes to completion of TGF-β-induced SMADs-dependent EMT. Similarly, the TGF-β-induced ERK/MAPKs pathway contributes to EMT induction, as ERK is required to remove cell adhesion junctions, which leads to increased cell migration [96,110,113] (Figure 3).

4. Tumor-Associated Macrophages (TAMs) and Cancer Associated Fibroblasts (CAFs)

Macrophages and fibroblasts are the main components of infiltrating stromal cells, called tumor-associated macrophages (TAMs) and cancer-associated fibroblasts (CAFs), respectively. CAFs are a heterogeneous cell population; every tumor type consists of tumor- and tissue-specific CAF subpopulations [114,115,116]. Different markers have been used to identify these subgroups including α-smooth muscle actin (α-SMA), fibroblast-specific protein 1 (FSP1), fibroblast activation protein (FAP), and neuron glial antigen-2 (NG2) [116,117]. In human CCA, hepatic stellate cells (HSCs) are thought to be the dominant cellular source of CAFs [118], whilst other studies suggest that CCA-associated CAFs might be derived from portal fibroblasts (PFs) and from bone marrow derived fibrocytes (BMDFs) [119,120,121]. Activation of CAFs is associated with an invasive phenotype of the tumor in CCA [122]. TAMs and CAFs have also been reported to be involved in tumor progression [56,123,124] as well as in the RAS, MAPK, and MET signaling cascades. On the other hand, TAMs can promote the secretion of tumor growth-promoting factors such as TNF-α, IL6, TGF-β, PDGFR, and VEGFR, among others. These factors can usually act on EMT, promoting tumor growth, and metastasis in CCA [56,125,126,127]. Cadamuro et al. [56] suggest that CCA cells recruit CAFs by secreting platelet-derived growth factor D (PDGF-D), promoting CCA cell migration through PDGFR-β, RHO-GTPase, and JNK activation. The recruitment of macrophages is often supported by CAFs and epithelial tumor cells, and is also stimulated by regulatory signaling pathways (e.g., NOTCH and IL6/STAT3) [128,129,130]. In addition, neutrophil migration can be induced by the P13K/AKT and MAPKs signaling pathways, which are manifested by the recruitment of neutrophils in CCA [131]. According to recent studies, CAFs are able to promote CCA progression through interacting autocrine and paracrine signaling pathways, especially in promoting EMT [97,132].
As above-mentioned, the NOTCH signaling pathway (including the NOTCH ligand JAG1) seems to be another particularly important reason for CCA formation. In a murine model of liver cancer driven by transfection of activated forms of AKT and RAS oncogenes, inhibiting NOTCH and its ligand JAG1 markedly impairs CCA development. According to cell-fate tracing experiments, combined activation of NOTCH and AKT could induce mature hepatocytes to develop into CCA in mice [133,134] (Figure 2). Moreover, a recent study from Thongsom’s group [135] demonstrated that piperlongumine (PL), a naturally-occurring small molecule, could induce CCA cellular apoptosis through ROS-mediated JNK/ERK activation and proteasome inhibition, while the susceptibility of CCA to PL therapy may depend on the activation of AKT. In fact, overwhelming evidence indicates that it is possible to convert hepatocytes into CCA cells when exposed to specific stimuli, and that the key driving factor in this process is NOTCH signal transduction [133,134]. Indeed, activation of the NOTCH signaling pathway induces EMT and increases CCA cell line migration [136,137,138].
Several reports have linked the p38 signaling pathway to the regulation of EMT [139,140,141,142]. As noted previously, there is crosstalk between the p38 and the JNK signaling pathways. The JNK and p38 MAPKs pathways share several upstream regulatory factors, so multiple stimuli may activate both pathways. As previously shown, p38 responds to ROS accumulation to prevent tumorigenesis by promoting growth stagnation and apoptosis [143,144]. ROS can promote TNF-α to activate the JNK signaling pathway and lead to apoptosis. However, TNF-α can also activate NF-κB to inhibit ROS through related genes such as MnSOD (manganese-superoxide dismutase) and FHC (ferritin heavy chain), thereby preventing JNK activation and inhibiting apoptosis [145,146]. While activation of p38 promotes MSK1/2 (mitogen- and stress-activated protein kinase 1/2), and subsequently, NF-κB [147,148]. Furthermore, the JNK signaling pathway has been shown to regulate tricellulin-mediated tight junctions in human pancreatic ductal epithelial cells, indicating that it may play an important role in the differentiation of ductular cells [149,150].
As demonstrated in previous studies, TNF-α may alter the differentiation of hepatocytes in liver disease [151]. In TNF-α-stimulated fibroblast death, JNK activation plays a dual role, inhibiting not only TNF-α-stimulated apoptosis, but also TNF-α-induced necrosis [152,153]. Earlier studies by Nishikawa’s group [151] demonstrated that, within the collagen-rich matrix, TNF specifically stimulated branching morphogenesis associated with the expression of bile duct/ductular markers including cytokeratin 19 (CK-19). Meanwhile, the phosphorylation of JNK and c-Jun increased during collagen gel culture, and TNF-α strongly enhanced bile duct transdifferentiation of hepatocytes, principally by inhibiting hepatocyte differentiation and enhancing ductular morphogenesis. Interestingly, in human liver-fluke-associated CCA, overexpression of TNF-α and NF-κB are characteristic [154].
Several studies have shown that CAFs from human CCA are capable of synthesizing HB-EGF (heparin binding EGF like growth factor) with the activation of EGFR in CCA cells, promoting disruption of cell–cell junctions and increased invasiveness [97,155,156]. EGFR expression and signaling are closely related to the development and progression of CCA [17,157]. Abnormal phosphorylation of EGFR family receptors activates p38, increases COX-2, inhibits apoptosis, and promotes tumor growth [158,159]. Simultaneously, epigenetic regulation of IL-6 promotes the progression of CCA by affecting promoter methylation and gene expression in growth regulatory pathways including EGFR [160]. Another study using sulfated galactans (SG) isolated from the red seaweed Gracilaria fisheri inhibited CCA cell proliferation by inhibiting EGFR and ERK phosphorylation, and blocking EGF-induced cell proliferation. These results suggest that SG may interact with EGFR and mediate cell proliferation inhibition, partly by blocking the endogenous activation of the EGFR/ERK/MAPKs signaling pathway [161].

5. Biomarkers and Diagnosis

The development of CCA involves genetic alterations of related oncogenes in humans [12,130,162,163,164,165,166,167] (Table 1). Mutations common to tumors all along the chromosome include tumor suppressor genes (TP53, and PTEN), chromatin-remodeling genes (ARID1A, ARID1B, BAP1, PBRM1), and gain of function of oncogenes (KRAS, BRAF, and PIK3CA). However, mutations in the ERK/MAPKs pathway components are relatively common in CCA. In particular, KRAS mutations are associated with a decrease in both progression-free, and overall survival in CCA patients [168]. KRAS activating mutations are frequently detected (22%, range 5–57%) [17,169,170], especially in codon 12 hotspots, and have recently been identified as an independent predictor of poor survival after surgery [170,171].
Oncogenes in humans CCA include KMT2C (lysine methyltransferase 2C), roundabout guidance receptor 2 (ROBO2), ring finger protein 43 (RNF43), paternally expressed 3 (PEG3), guanine nucleotide binding protein (G protein) alpha stimulating activity polypeptide 1 (GNAS), (BRCA-associated protein 1 (BAP1), V-Ki-ras2 Kirsten rat sarcoma viral oncogene homolog (KRAS), neuroblastoma RAS viral (v-ras) oncogene homolog (NRAS), v-raf murine sarcoma viral oncogene homolog B1 (BRAF), polybromo 1 (PBRM1), AT rich interactive domain 1A (SWI-like) (ARID1A), AT-rich interaction domain 1B (ARID1B); phosphoinositide-3-kinase, catalytic, alpha polypeptide (PIK3CA), phosphatase and tensin homolog deleted on chromosome ten (PTEN), protein tyrosine phosphatase non-receptor type 3 (PTPN3), cyclin dependent kinase 2a/p16 (CDKN2A), mothers against decapentaplegic homolog 4 (Drosophila) (SMAD4), tumor protein p53 (TP53), isocitrate dehydrogenase 1 (IDH1), and isocitrate dehydrogenase 2 (IDH2).
As previously mentioned, carcinogenesis in CCA is characterized by recruitment of fibroblasts, ECM remodeling, changes in immune cell migration patterns, and promotion of angiogenesis [172]. Matrix metalloproteinases that degrade and remodel the ECM including MMP1, MMP2, MMP3, and MMP9 are strongly expressed in CCA and are associated with invasive tumors. CCA does not always exhibit a tubular contour, which is easily recognized in histopathological studies; instead, cancer cells often appear nodular or exhibit a form of HCC resembling a neuroendocrine or mixed solid shape [173]. CCA cells usually grow along the wall of the portal vein in connective tissue and directly intrude into the nearby bile ducts.
In practice, immunohistochemical detection of CK-7, CK-19, CK-20, CDX-2, and CA199 biomarkers is often used to identify CCA [12,174]; whilst CK-7 and CAM5.2 biomarkers are strongly expressed in CCA cells. Immunohistochemical detection of human EpCAM on the cell membrane is positive in 90% of cases, while the presence of CK-19 is found in approximately 70% to 80% of cases [164,175,176]. In addition, within the NOTCH signaling pathway, SOX9 plays an important role in the pathophysiology of the biliary tract, and SOX9 has been shown to be associated with the progression of EMT in CCA [177,178]. Unfortunately, none of these immunophenotypes are entirely specific for CCA, since they are often also seen in pancreatic ducts and bile duct primordial cells. Accurate diagnosis of CCA depends mainly on clinical symptoms, histopathology, and imaging findings [173]. Even though recent studies have found that modified branched-chain DNA probes for albumin mRNA (used for in situ detection of albumin expression) have a detection rate of 99% for CCA and 100% for liver cancer [179], these are not detected exclusively in CCA, not even in early stages. As such, most patients are diagnosed with CCA at advanced stages and have limited treatment options, resulting in a poor prognosis [180].
The combination of clinical and biochemical findings including abdominal pain/distension, pruritus, jaundice, and weight loss, imaging techniques like ultrasound (US), computed tomography (CT), magnetic resonance imaging (MRI), 18F-fluorodeoxyglucose positron emission tomography (18FDG-PET), and serum analysis of non-specific tumor biomarkers such as CA199, CEA (carcinoembryonic antigen), and CD133 [180], usually help to diagnose CCA, but ultimately liver biopsy and pathological diagnosis are used as the gold standard [181].

6. Therapy

6.1. Surgical Treatment

Currently, surgical resection is still the preferred treatment for CCA. However, only a small number of patients (about 35%) are diagnosed sufficiently early to undergo surgery [182]. In 2012, the 5-year survival rates of surgically treated CCA, iCCA, pCCA, and dCCA were 22–44%, 11–41%, and 27–37%, respectively [4]. However, a recent study including 574 patients with pCCA that underwent right hepatic artery resection and reconstruction, with a perioperative mortality rate of less than 5%, showed a 5-year survival rate of 30% [183]. Contraindications to CCA surgical resection include bilateral, multifocal disease, distant metastases, and comorbidities associated with surgical risk that exceed the patient’s expected surgical benefits. Regional lymph node metastasis is not considered an absolute contraindication to resection, although N1 disease (CCA with regional lymph node metastases including nodes along the cystic duct, common bile duct, hepatic artery, and portal vein) is one of the prognostic factors representing poor prognosis [184,185,186]. Although Bismuth-Corlette type IV pCCA is not considered to be an absolute contraindication to surgical resection, subsequent orthotopic liver transplantation (OLT) is a valid treatment after neoadjuvant radiotherapy [187]. The recurrence rate was 20% and the 5-year survival rate was 68% [187]. However, the selection criteria were strict and 25% to 31% of patients developed disease progression while waiting for OLT and were excluded from the protocol [186,187]. Therefore, single-OLT is not recommended as a CCA monotherapy because of high recurrence rates and long-term survival rates of less than 20% [5,188]. Current guidelines for CCA are still controversial, but due to regional differences between Eastern and Western centers, different surgical resection criteria and surgical strategies are based on applied areas, especially between the Western (US and Europe) and the Eastern centers, with aggressive surgical procedures (including extended hepatectomy and combination with vascular resection in early pCCA). Thus, the actual resection rate of CCA patients has increased, and the prognosis improved in East Asia [183,189,190].

6.2. Non-Surgical Treatment: Targeted Therapies

The complexity of the pathogenesis and the apparent heterogeneity of CCA have hindered use of efficacious clinical therapeutics in their management [39,191,192]. The combination of gemcitabine and cisplatin (Table 2) is the first-line chemotherapy for patients with advanced CCA (not suitable for local treatment and surgery) and is also suitable for various anatomical disease subtypes. The median survival of the combination was 11.7 months, and 8.1 months with gemcitabine alone [193,194]. Activation of EGFR leads to downstream activation of MAPKs, however, the efficiency of the EGFR inhibitor erlotinib in clinical trials of CCA is limited according to a Multicenter Phase II clinical trial testing [165,195]. A recent phase III randomized controlled trial showed that by combining erlotinib with gemcitabine/cisplatin, patients had a higher response rate (31% vs. 14%, p = 0.004) and longer survival (5.9 months vs. 3.0 months, p = 0.049) [196].
Locoregional treatments including radiofrequency ablation, transcatheter arterial chemoembolization (TACE), drug eluting beads-TACE, selective 90Y internal artery radiotherapy, and microsphere or external radiation therapy are also applied in CCA [186]. Since technology and equipment have improved, the safety and efficacy of radiotherapy for CCA are better. Advances in these techniques and devices can also increase the radiation therapy dose of biliary tumors and/or improve the protection of non-malignant tissues, thereby improving the therapeutic benefit of radiation therapy for CCA [197,198,199,200].

Molecularly Targeted Therapy

Recently, there are several new methods available for CCA including molecularly targeted therapy and immunotherapy. The FGFR inhibitor NVP-BGJ398 exhibits the potential to treat CCA [201,202]. Some clinical studies are currently testing patients with advanced CCA who have FGFR changes (NCT02150967), (NCT01703481), (NCT02699606), and (NCT02265341) [12,203,204]. BGJ398 and another small molecule, FGFR inhibitor PD173072, inhibit the downstream carcinogenic activity of MAPK signaling and FGFR fusion kinase. The clinical efficacy of BGJ398 is currently under assessment in a phase I clinical study for advanced solid tumors, with FGFR1/FGFR2 amplification or FGFR3 mutation (NCT01004224) [165,202,205]. ALK (EML4(echinoderm microtubule-associated protein-like 4)–ALK(anaplastic lymphoma kinase)) and the ROS1 inhibitor Ceritinib are currently being evaluated in a phase II clinical trial of patients with ROS1-positive or ALK-positive advanced-stage pCCA or iCCA (NCT02374489), (NCT02638909), and (NCT02568267) [12]. Anti-MEK therapies in CCA have been tested [206,207,208,209,210]. Binimetinib (MEK162), a potent, selective oral MEK1/2 inhibitor, showed promising evidence of activity in patients with biliary tract cancer [207]. PD901, a MEK inhibitor, has been shown to be effective against CCA harboring KRAS oncogenic mutations via inhibition of cell proliferation and modulation of the tumor microenvironment [208]. Cabozantinib is a multi-kinase inhibitor with anti-MET and anti-VEGFR2 activity. Unfortunately, its activity is limited, median progression-free survival (PFS) is 1.8 months, and has significant toxicity to unselected patients with CCA [209]. Similarly, trametinib, in combination with pazopanib, an oral VEGFR tyrosine kinase inhibitor, in patients with advanced CCA, did not achieve a statistically significant improvement in 4-month PFS over the prespecified null hypothesized 4-month PFS [210]. Sorafenib, a multi-kinase inhibitor with anti-VEGFR and RAF family kinase activity (inhibitor of the RAF/MEK/ERK pathway) [211], combined with the EGFR inhibitor Erlotinib in the treatment of advanced CCA, has been disappointing [212]. Due to chemoresistance, gene therapy capable of selectively inducing human organic cation transporter type 1 (hOCT1) in tumor cells was suggested as a potentially useful chemosensitization strategy to improve the response of CCA to sorafenib [213,214]. Again, it has been reported that hot-spot mutations in the IDH1 and IDH2 genes are frequent in CCA and can promote epigenetic alterations through the regulation of DNA demethylases activity [215]. Small molecules that inhibit the neomorphic activity of mutant IDH can reverse DNA methylation, and, in fact, may promote cancer cell differentiation. Such inhibitors may however also be candidates for the treatment of CCA [216,217,218,219]. Therefore, small molecule inhibitors of mutant IDH1 or IDH2 have demonstrated efficacy, and, in fact, orally available compounds such as AG-120 have entered clinical trials (inhibitor of mutant IDH1) [220]. Preliminary studies treating CCA patients with mutated IDH1 with AG-120 in Phase I clinical trials (dose escalation and dose amplification cohort studies) showed that this inhibitor was safe and effective (NCT02073994) [221]. Enasidenib, a selective inhibitor of mutant IDH2, is currently being studied in multi-phase I/II trials in patients with IDH2-mutant advanced-stage solid tumors including CCA(NCT02273739) [12]. Anetumab ravtansine, an anti-mesothelin antibody-drug conjugate, is undergoing Phase I trials to recruit patients with advanced-stage CCA with abnormal expression of mesothelin [12]. CDKN2A is a gene encoding the tumor suppressor proteins p16INK4A and p14ARF and is an important negative regulator of cell cycle progression. Repetitive focal loss of cdkna2 highlights the potential of CDK4/6 inhibitors in the treatment of CCA including ribociclib and palbociclib [16,166,222]. Several studies have found a subset of CCA patients with high expression of PD-1/PD-L1, suggesting that these patients may be suitable for treatment with PD-1 or PD-L1 inhibitors (NCT02703714, NCT02628067) [223,224,225] (Table 1).
Inhibitors of the JNK signaling pathway may also play an important role in therapy against CCA. To date, a number of small molecule inhibitors have been developed that may regulate specific JNK signaling pathways [31]. One study using the RBE human CCA cell line treated with SP600125, an inhibitor of JNK, demonstrated enhanced TGF-β-induced cell apoptosis in a SMAD4-dependent manner, suggesting that JNK inhibition may be an ideal therapeutic candidate for the treatment of human CCA [233]. However, as JNKs play a pivotal role in normal biological functions, the direct inhibition of JNK itself may also have unexpected side effects. Therefore, the development of highly selective JNK1 or JNK2 inhibitors is essential for the safe treatment of cancer [234]. In this regard, it is important to note that selective inhibition between the individual and their associated JNK proteins may be involved in normal cell functions (such as adaptive immunity and brain development), and thus may lead to adverse cellular effects and unwanted toxicity [235,236,237,238]. Unfortunately, MAPKs-directed therapy is not showing the same clinical efficacy as the experimental models. Thus, further research should also focus on developing substrate specific inhibitors that can make important contributions to mitigate different types of cancer [24] including CCA.

7. Conclusions

Recent advances have led to a more profound understanding of the role of the MAPKs signaling pathway in liver neoplastic diseases. Clinical symptoms, histopathology, and gathering of detailed molecular information are of great significance for the diagnosis and treatment of CCA. Utilizing improved experimental models, the important roles of the MAPKs signaling pathway in the pathogenesis of CCA are emerging. Although the exact potential risk factors for CCA have not yet been accurately identified, perhaps the mutation profiles identified are closely related to the underlying causes of CCA. Moreover, some of these mutated proteins can become therapeutic targets for the treatment of CCA. Exciting new research is revealing that liver cancer progression can be attenuated by inhibiting or blocking dysregulated cell transduction components of the MAPKs signaling pathway. If we can elucidate the signal transduction pattern of liver cancer development and carry out specific blockade in certain aspects, it will herald significant breakthroughs and progress in the diagnosis and treatment of CCA.

Author Contributions

C.C.: manuscript draft and preparation; L.J.N. and M.A.Á.: critically reviewed the manuscript; F.J.C.; supervised, edited, added information, and reviewed manuscript content.

Funding

This work was supported by the MINECO Retos SAF2016-78711, EXOHEP-CM S2017/BMD-3727, NanoLiver-CM Y2018/NMT-4949, UCM-25-2019, ERAB EA/18-14, AMMF 2018/117, and COST Action CA17112. We also acknowledge the funding from the Instituto de Salud Carlos III grant FIS16/01126, AECC grant for rare tumors 2017, Hepacare Project “la Caixa”, and Gobierno de Navarra Salud Grant 58/17. F.J.C. is a Ramón y Cajal Researcher (RYC-2014-15242) affiliated with the UCM group “Lymphocyte Immunobiology”, No. 920631 (imas12-associated, Ref. IBL-6) and a Gilead Liver Research Scholar 2018.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations (in Alphabetical Order)

ALKanaplastic lymphoma kinase
ARID1AAT rich interactive domain 1A (SWI-like)
ASK1apoptosis signal-regulating kinase 1
ATF4activated transcription factor 4
ATOarsenic trioxide
BAP1BRCA-associated protein 1
BECsbiliary epithelial cells
BMDFsbone marrow derived fibrocytes
BRAFv-raf murine sarcoma viral oncogene homolog B1;
CA199carbohydrate antigen 19-9
CAFscancer-associated fibroblasts
CCAcholangiocarcinoma
CDKN2Acyclin dependent kinase 2a/p16
CDX2caudal type homeobox 2
CEAcarcinoembryonic antigen
c-Mettyrosine-protein kinase Met
CK-7cytokeratin 7
CK-19cytokeratin 19
CK-20cytokeratin 20
CTcomputed tomography
dCCAdistal cholangiocarcinoma
ECMextracellular matrix
EGFepidermal growth factor
EGFRepidermal growth factor receptor
eIF2aeukaryotic initiation factor-α
EML4echinoderm microtubule-associated protein-like 4
EMTepithelial to mesenchymal transition
EMT-ATFsEMT-activated transcription factors
ERKextracellular signal-regulated kinase
ERSendoplasmic reticulum stress
FAPFibroblast activation protein
18FDG-PET18F-fluorodeoxyglucose positron emission tomography
FHCferritin heavy chain
FSP1fibroblast-specific protein 1
GEFguanine nucleotide exchange factor
GNASguanine nucleotide binding protein (G protein) alpha stimulating activity polypeptide 1
GPCRsG protein-coupled receptors
GRP78glucose-regulated protein 78
HB-EGFheparin binding EGF like growth factor
HCChepatocellular carcinoma
HNFhepatocyte nuclear factor
hOCT1human organic cation transporter type 1
HPCshepatic progenitor cells
HSCshepatic stellate cells
iCCAintrahepatic cholangiocarcinoma
IDH1isocitrate dehydrogenase 1
IDH2isocitrate dehydrogenase 2
IKKinhibitor of NF-κB kinase
IL6interleukin 6
ILRinterleukin receptor
JNKc-Jun NH2-terminal kinase
KCKupffer cell
KMT2Clysine methyltransferase 2C
KRASV-Ki-ras2 Kirsten rat sarcoma viral oncogene homolog
LPCliver parenchymal cells
LPSlipopolysaccharide
MAPKsmitogen-activated protein kinases
MAP2Ksmitogen-activated protein kinase kinases
MAP3Ksmitogen-activated protein kinase kinase kinases
MAP3K5mitogen-activated protein kinase kinase kinase 5
MEKmitogen-activated protein kinase
MEKK1mitogen-activated protein kinase kinase kinase 1
METmesenchymal to epithelial transition
MKK3MAP kinase 3
MKK4MAP kinase 4
MKK6MAP kinase 6
MKK7MAP kinase 7
MLK3mitogen-activated protein kinase kinase kinase 11
MMPsMatrix metalloproteinases
MnSODmanganese-superoxide dismutase
MRImagnetic resonance imaging
mTORmammalian target of rapamycin
NF-κBnuclear factor (NF)-κB
NLKNemo-like kinase
NRASneuroblastoma RAS viral (v-ras) oncogene homolog
OLTorthotopic liver transplantation
PBCprimary biliary cholangitis
PBRM1polybromo 1
pCCAperihilar cholangiocarcinoma
PDGFplatelet-derived growth factor
PDGF-Dplatelet derived growth factor D
PDGFRplatelet-derived growth factor receptor
PDGFR-βplatelet-derived growth factor receptor-β
PEG3paternally expressed 3
PTENphosphatase and tensin homolog deleted on chromosome ten
PFSprogression-free survival
PFsportal fibroblasts
PI3Kphosphatidylinositol 3-kinase
PIK3CAphosphoinositide-3-kinase, catalytic, alpha polypeptide
PLpiperlongumine
PSCprimary sclerosing cholangitis
RNF43ring finger protein 43
ROBO2roundabout guidance receptor 2
ROSreactive oxygen species
RTKsreceptor tyrosine kinases
SGsulfated galactans
SMAD4sulfated galactans; mothers against decapentaplegic homolog 4 (Drosophila)
SOSson of sevenless
α-SMAα-smooth muscle actin
STAT3signal transducer and activator of transcription 3
TACEtranscatheter arterial chemoembolization
TAK1transforming growth factor-β-activated kinase 1
TAMstumor-associated macrophages
TNFtumor necrosis factor
TNF-αtumor necrosis factor alpha
TNFRtumor necrosis factor receptor
TGF-βtransforming growth factor-β
TP53Tumor protein p53
USultrasound
VEGFvascular endothelial growth factor
VEGFRvascular endothelial growth factor receptor

References

  1. Shin, H.R.; Oh, J.K.; Lim, M.K.; Shin, A.; Kong, H.J.; Jung, K.W.; Won, Y.J.; Park, S.; Park, S.J.; Hong, S.T. Descriptive epidemiology of cholangiocarcinoma and clonorchiasis in Korea. J. Korean Med. Sci. 2010, 25, 1011–1016. [Google Scholar] [CrossRef] [PubMed]
  2. McLean, L.; Patel, T. Racial and ethnic variations in the epidemiology of intrahepatic cholangiocarcinoma in the United States. Liver Int. 2006, 26, 1047–1053. [Google Scholar] [CrossRef] [PubMed]
  3. Tyson, G.L.; El-Serag, H.B. Risk factors for cholangiocarcinoma. Hepatology 2011, 54, 173–184. [Google Scholar] [CrossRef] [PubMed]
  4. Khan, S.A.; Davidson, B.R.; Goldin, R.D.; Heaton, N.; Karani, J.; Pereira, S.P.; Rosenberg, W.M.; Tait, P.; Taylor-Robinson, S.D.; Thillainayagam, A.V.; et al. Guidelines for the diagnosis and treatment of cholangiocarcinoma: An update. Gut 2012, 61, 1657–1669. [Google Scholar] [CrossRef] [PubMed]
  5. Bridgewater, J.; Galle, P.R.; Khan, S.A.; Llovet, J.M.; Park, J.W.; Patel, T.; Pawlik, T.M.; Gores, G.J. Guidelines for the diagnosis and management of intrahepatic cholangiocarcinoma. J. Hepatol. 2014, 60, 1268–1289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Shaib, Y.H.; Davila, J.A.; McGlynn, K.; El-Serag, H.B. Rising incidence of intrahepatic cholangiocarcinoma in the United States: A true increase? J. Hepatol. 2004, 40, 472–477. [Google Scholar] [CrossRef] [PubMed]
  7. Dodson, R.M.; Weiss, M.J.; Cosgrove, D.; Herman, J.M.; Kamel, I.; Anders, R.; Geschwind, J.F.; Pawlik, T.M. Intrahepatic cholangiocarcinoma: Management options and emerging therapies. J. Am. Coll. Surg. 2013, 217, 736–750 e734. [Google Scholar] [CrossRef]
  8. Labib, P.L.; Goodchild, G.; Pereira, S.P. Molecular pathogenesis of cholangiocarcinoma. BMC Cancer 2019, 19, 185. [Google Scholar] [CrossRef] [PubMed]
  9. Welzel, T.M.; Mellemkjaer, L.; Gloria, G.; Sakoda, L.C.; Hsing, A.W.; El Ghormli, L.; Olsen, J.H.; McGlynn, K.A. Risk factors for intrahepatic cholangiocarcinoma in a low-risk population: A nationwide case-control study. Int. J. Cancer 2007, 120, 638–641. [Google Scholar] [CrossRef]
  10. Welzel, T.M.; Graubard, B.I.; El-Serag, H.B.; Shaib, Y.H.; Hsing, A.W.; Davila, J.A.; McGlynn, K.A. Risk factors for intrahepatic and extrahepatic cholangiocarcinoma in the United States: A population-based case-control study. Clin. Gastroenterol. Hepatol. 2007, 5, 1221–1228. [Google Scholar] [CrossRef]
  11. Banales, J.M.; Cardinale, V.; Carpino, G.; Marzioni, M.; Andersen, J.B.; Invernizzi, P.; Lind, G.E.; Folseraas, T.; Forbes, S.J.; Fouassier, L.; et al. Expert consensus document: Cholangiocarcinoma: Current knowledge and future perspectives consensus statement from the European network for the study of cholangiocarcinoma (ens-cca). Nat. Rev. Gastroenterol. Hepatol. 2016, 13, 261–280. [Google Scholar] [CrossRef] [PubMed]
  12. Rizvi, S.; Khan, S.A.; Hallemeier, C.L.; Kelley, R.K.; Gores, G.J. Cholangiocarcinoma—Evolving concepts and therapeutic strategies. Nat. Rev. Clin. Oncol. 2018, 15, 95–111. [Google Scholar] [CrossRef] [PubMed]
  13. Komuta, M.; Govaere, O.; Vandecaveye, V.; Akiba, J.; Van Steenbergen, W.; Verslype, C.; Laleman, W.; Pirenne, J.; Aerts, R.; Yano, H.; et al. Histological diversity in cholangiocellular carcinoma reflects the different cholangiocyte phenotypes. Hepatology 2012, 55, 1876–1888. [Google Scholar] [CrossRef] [PubMed]
  14. Aishima, S.; Oda, Y. Pathogenesis and classification of intrahepatic cholangiocarcinoma: Different characters of perihilar large duct type versus peripheral small duct type. J. Hepatobil. Pancreat. Sci. 2015, 22, 94–100. [Google Scholar] [CrossRef] [PubMed]
  15. Rizvi, S.; Gores, G.J. Pathogenesis, diagnosis, and management of cholangiocarcinoma. Gastroenterology 2013, 145, 1215–1229. [Google Scholar] [CrossRef] [PubMed]
  16. Nakamura, H.; Arai, Y.; Totoki, Y.; Shirota, T.; Elzawahry, A.; Kato, M.; Hama, N.; Hosoda, F.; Urushidate, T.; Ohashi, S.; et al. Genomic spectra of biliary tract cancer. Nat. Genet. 2015, 47, 1003–1010. [Google Scholar] [CrossRef] [PubMed]
  17. Sia, D.; Hoshida, Y.; Villanueva, A.; Roayaie, S.; Ferrer, J.; Tabak, B.; Peix, J.; Sole, M.; Tovar, V.; Alsinet, C.; et al. Integrative molecular analysis of intrahepatic cholangiocarcinoma reveals 2 classes that have different outcomes. Gastroenterology 2013, 144, 829–840. [Google Scholar] [CrossRef] [PubMed]
  18. Al-Bahrani, R.; Abuetabh, Y.; Zeitouni, N.; Sergi, C. Cholangiocarcinoma: Risk factors, environmental influences and oncogenesis. Ann. Clin. Lab. Sci. 2013, 43, 195–210. [Google Scholar]
  19. Isomoto, H.; Kobayashi, S.; Werneburg, N.W.; Bronk, S.F.; Guicciardi, M.E.; Frank, D.A.; Gores, G.J. Interleukin 6 upregulates myeloid cell leukemia-1 expression through a stat3 pathway in cholangiocarcinoma cells. Hepatology 2005, 42, 1329–1338. [Google Scholar] [CrossRef]
  20. Cargnello, M.; Roux, P.P. Activation and function of the mapks and their substrates, the mapk-activated protein kinases. Microbiol. Mol. Biol. Rev. 2011, 75, 50–83. [Google Scholar] [CrossRef] [PubMed]
  21. Johnson, G.L.; Lapadat, R. Mitogen-activated protein kinase pathways mediated by erk, jnk, and p38 protein kinases. Science 2002, 298, 1911–1912. [Google Scholar] [CrossRef] [PubMed]
  22. Brancho, D.; Tanaka, N.; Jaeschke, A.; Ventura, J.J.; Kelkar, N.; Tanaka, Y.; Kyuuma, M.; Takeshita, T.; Flavell, R.A.; Davis, R.J. Mechanism of p38 map kinase activation in vivo. Genes Dev. 2003, 17, 1969–1978. [Google Scholar] [CrossRef] [PubMed]
  23. Remy, G.; Risco, A.M.; Inesta-Vaquera, F.A.; Gonzalez-Teran, B.; Sabio, G.; Davis, R.J.; Cuenda, A. Differential activation of p38mapk isoforms by mkk6 and mkk3. Cell. Signal. 2010, 22, 660–667. [Google Scholar] [CrossRef] [PubMed]
  24. Bubici, C.; Papa, S. Jnk signalling in cancer: In need of new, smarter therapeutic targets. Br. J. Pharmacol. 2014, 171, 24–37. [Google Scholar] [CrossRef] [PubMed]
  25. Mao, X.; Bravo, I.G.; Cheng, H.; Alonso, A. Multiple independent kinase cascades are targeted by hyperosmotic stress but only one activates stress kinase p38. Exp. Cell. Res. 2004, 292, 304–311. [Google Scholar] [CrossRef] [PubMed]
  26. Roberts, P.J.; Der, C.J. Targeting the raf-mek-erk mitogen-activated protein kinase cascade for the treatment of cancer. Oncogene 2007, 26, 3291–3310. [Google Scholar] [CrossRef] [PubMed]
  27. Win, S.; Than, T.A.; Zhang, J.; Oo, C.; Min, R.W.M.; Kaplowitz, N. New insights into the role and mechanism of c-jun-n-terminal kinase signaling in the pathobiology of liver diseases. Hepatology 2018, 67, 2013–2024. [Google Scholar] [CrossRef] [PubMed]
  28. Win, S.; Than, T.A.; Min, R.W.; Aghajan, M.; Kaplowitz, N. C-jun n-terminal kinase mediates mouse liver injury through a novel sab (sh3bp5)-dependent pathway leading to inactivation of intramitochondrial src. Hepatology 2016, 63, 1987–2003. [Google Scholar] [CrossRef] [PubMed]
  29. Mingo-Sion, A.M.; Marietta, P.M.; Koller, E.; Wolf, D.M.; Van Den Berg, C.L. Inhibition of jnk reduces g2/m transit independent of p53, leading to endoreduplication, decreased proliferation, and apoptosis in breast cancer cells. Oncogene 2004, 23, 596–604. [Google Scholar] [CrossRef]
  30. Vivanco, I.; Palaskas, N.; Tran, C.; Finn, S.P.; Getz, G.; Kennedy, N.J.; Jiao, J.; Rose, J.; Xie, W.; Loda, M.; et al. Identification of the jnk signaling pathway as a functional target of the tumor suppressor pten. Cancer Cell 2007, 11, 555–569. [Google Scholar] [CrossRef]
  31. Wang, J.; Tai, G. Role of c-jun n-terminal kinase in hepatocellular carcinoma development. Target. Oncol. 2016, 11, 723–738. [Google Scholar] [CrossRef] [PubMed]
  32. Cubero, F.J.; Zhao, G.; Nevzorova, Y.A.; Hatting, M.; Al Masaoudi, M.; Verdier, J.; Peng, J.; Schaefer, F.M.; Hermanns, N.; Boekschoten, M.V.; et al. Haematopoietic cell-derived jnk1 is crucial for chronic inflammation and carcinogenesis in an experimental model of liver injury. J. Hepatol. 2015, 62, 140–149. [Google Scholar] [CrossRef] [PubMed]
  33. Chen, F. Jnk-induced apoptosis, compensatory growth, and cancer stem cells. Cancer Res. 2012, 72, 379–386. [Google Scholar] [CrossRef] [PubMed]
  34. Seki, E.; Brenner, D.A.; Karin, M. A liver full of jnk: Signaling in regulation of cell function and disease pathogenesis, and clinical approaches. Gastroenterology 2012, 143, 307–320. [Google Scholar] [CrossRef] [PubMed]
  35. Cazanave, S.C.; Mott, J.L.; Elmi, N.A.; Bronk, S.F.; Werneburg, N.W.; Akazawa, Y.; Kahraman, A.; Garrison, S.P.; Zambetti, G.P.; Charlton, M.R.; et al. Jnk1-dependent puma expression contributes to hepatocyte lipoapoptosis. J. Biol. Chem. 2009, 284, 26591–26602. [Google Scholar] [CrossRef]
  36. Qiu, W.; Wang, X.; Leibowitz, B.; Yang, W.; Zhang, L.; Yu, J. Puma-mediated apoptosis drives chemical hepatocarcinogenesis in mice. Hepatology 2011, 54, 1249–1258. [Google Scholar] [CrossRef] [PubMed]
  37. Chang, Q.; Zhang, Y.; Beezhold, K.J.; Bhatia, D.; Zhao, H.; Chen, J.; Castranova, V.; Shi, X.; Chen, F. Sustained jnk1 activation is associated with altered histone h3 methylations in human liver cancer. J. Hepatol. 2009, 50, 323–333. [Google Scholar] [CrossRef]
  38. Dhanasekaran, D.N.; Reddy, E.P. Jnk signaling in apoptosis. Oncogene 2008, 27, 6245–6251. [Google Scholar] [CrossRef]
  39. Yuan, D.; Huang, S.; Berger, E.; Liu, L.; Gross, N.; Heinzmann, F.; Ringelhan, M.; Connor, T.O.; Stadler, M.; Meister, M.; et al. Kupffer cell-derived tnf triggers cholangiocellular tumorigenesis through jnk due to chronic mitochondrial dysfunction and ros. Cancer Cell 2017, 31, 771–789 e776. [Google Scholar] [CrossRef]
  40. Zhong, F.; Tong, Z.T.; Fan, L.L.; Zha, L.X.; Wang, F.; Yao, M.Q.; Gu, K.S.; Cao, Y.X. Guggulsterone-induced apoptosis in cholangiocarcinoma cells through ros/jnk signaling pathway. Am. J. Cancer Res. 2016, 6, 226–237. [Google Scholar]
  41. Feng, C.; He, K.; Zhang, C.; Su, S.; Li, B.; Li, Y.; Duan, C.Y.; Chen, S.; Chen, R.; Liu, Y.; et al. Jnk contributes to the tumorigenic potential of human cholangiocarcinoma cells through the mtor pathway regulated grp78 induction. PLoS ONE 2014, 9, e90388. [Google Scholar] [CrossRef] [PubMed]
  42. Ling, S.; Xie, H.; Yang, F.; Shan, Q.; Dai, H.; Zhuo, J.; Wei, X.; Song, P.; Zhou, L.; Xu, X.; et al. Metformin potentiates the effect of arsenic trioxide suppressing intrahepatic cholangiocarcinoma: Roles of p38 mapk, erk3, and mtorc1. J. Hematol. Oncol. 2017, 10, 59. [Google Scholar] [CrossRef] [PubMed]
  43. Del Barco Barrantes, I.; Nebreda, A.R. Roles of p38 mapks in invasion and metastasis. Biochem. Soc. Trans. 2012, 40, 79–84. [Google Scholar] [CrossRef]
  44. Rousseau, S.; Dolado, I.; Beardmore, V.; Shpiro, N.; Marquez, R.; Nebreda, A.R.; Arthur, J.S.; Case, L.M.; Tessier-Lavigne, M.; Gaestel, M.; et al. Cxcl12 and c5a trigger cell migration via a pak1/2-p38alpha mapk-mapkap-k2-hsp27 pathway. Cell. Signal. 2006, 18, 1897–1905. [Google Scholar] [CrossRef] [PubMed]
  45. Platanias, L.C. Map kinase signaling pathways and hematologic malignancies. Blood 2003, 101, 4667–4679. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Yamagiwa, Y.; Marienfeld, C.; Tadlock, L.; Patel, T. Translational regulation by p38 mitogen-activated protein kinase signaling during human cholangiocarcinoma growth. Hepatology 2003, 38, 158–166. [Google Scholar] [CrossRef]
  47. Tan, F.L.; Ooi, A.; Huang, D.; Wong, J.C.; Qian, C.N.; Chao, C.; Ooi, L.; Tan, Y.M.; Chung, A.; Cheow, P.C.; et al. P38delta/mapk13 as a diagnostic marker for cholangiocarcinoma and its involvement in cell motility and invasion. Int. J. Cancer 2010, 126, 2353–2361. [Google Scholar] [CrossRef]
  48. Yoshikawa, D.; Ojima, H.; Iwasaki, M.; Hiraoka, N.; Kosuge, T.; Kasai, S.; Hirohashi, S.; Shibata, T. Clinicopathological and prognostic significance of egfr, vegf, and her2 expression in cholangiocarcinoma. Br. J. Cancer 2008, 98, 418–425. [Google Scholar] [CrossRef]
  49. Nakazawa, K.; Dobashi, Y.; Suzuki, S.; Fujii, H.; Takeda, Y.; Ooi, A. Amplification and overexpression of c-erbb-2, epidermal growth factor receptor, and c-met in biliary tract cancers. J. Pathol. 2005, 206, 356–365. [Google Scholar] [CrossRef]
  50. Dai, R.; Li, J.; Fu, J.; Chen, Y.; Wang, R.; Zhao, X.; Luo, T.; Zhu, J.; Ren, Y.; Cao, J.; et al. The tyrosine kinase c-met contributes to the pro-tumorigenic function of the p38 kinase in human bile duct cholangiocarcinoma cells. J. Biol. Chem. 2012, 287, 39812–39823. [Google Scholar] [CrossRef]
  51. Sia, D.; Tovar, V.; Moeini, A.; Llovet, J.M. Intrahepatic cholangiocarcinoma: Pathogenesis and rationale for molecular therapies. Oncogene 2013, 32, 4861–4870. [Google Scholar] [CrossRef] [PubMed]
  52. Papadimitriou, E.; Vasilaki, E.; Vorvis, C.; Iliopoulos, D.; Moustakas, A.; Kardassis, D.; Stournaras, C. Differential regulation of the two rhoa-specific gef isoforms net1/net1a by tgf-beta and mir-24: Role in epithelial-to-mesenchymal transition. Oncogene 2012, 31, 2862–2875. [Google Scholar] [CrossRef] [PubMed]
  53. Olieslagers, S.; Pardali, E.; Tchaikovski, V.; ten Dijke, P.; Waltenberger, J. Tgf-beta1/alk5-induced monocyte migration involves pi3k and p38 pathways and is not negatively affected by diabetes mellitus. Cardiovasc. Res. 2011, 91, 510–518. [Google Scholar] [CrossRef] [PubMed]
  54. Chapnick, D.A.; Warner, L.; Bernet, J.; Rao, T.; Liu, X. Partners in crime: The tgfbeta and mapk pathways in cancer progression. Cell Biosci. 2011, 1, 42. [Google Scholar] [CrossRef] [PubMed]
  55. Miller, T.; Yang, F.; Wise, C.E.; Meng, F.; Priester, S.; Munshi, M.K.; Guerrier, M.; Dostal, D.E.; Glaser, S.S. Simvastatin stimulates apoptosis in cholangiocarcinoma by inhibition of rac1 activity. Dig. Liver Dis. 2011, 43, 395–403. [Google Scholar] [CrossRef] [PubMed]
  56. Cadamuro, M.; Nardo, G.; Indraccolo, S.; Dall’olmo, L.; Sambado, L.; Moserle, L.; Franceschet, I.; Colledan, M.; Massani, M.; Stecca, T.; et al. Platelet-derived growth factor-d and rho gtpases regulate recruitment of cancer-associated fibroblasts in cholangiocarcinoma. Hepatology 2013, 58, 1042–1053. [Google Scholar] [CrossRef] [PubMed]
  57. Li, Z.; Biswas, S.; Liang, B.; Zou, X.; Shan, L.; Li, Y.; Fang, R.; Niu, J. Integrin beta6 serves as an immunohistochemical marker for lymph node metastasis and promotes cell invasiveness in cholangiocarcinoma. Sci. Rep. 2016, 6, 30081. [Google Scholar] [CrossRef] [PubMed]
  58. Sato, Y.; Harada, K.; Itatsu, K.; Ikeda, H.; Kakuda, Y.; Shimomura, S.; Shan Ren, X.; Yoneda, N.; Sasaki, M.; Nakanuma, Y. Epithelial-mesenchymal transition induced by transforming growth factor-{beta}1/snail activation aggravates invasive growth of cholangiocarcinoma. Am. J. Pathol. 2010, 177, 141–152. [Google Scholar] [CrossRef]
  59. Zen, Y.; Harada, K.; Sasaki, M.; Chen, T.C.; Chen, M.F.; Yeh, T.S.; Jan, Y.Y.; Huang, S.F.; Nimura, Y.; Nakanuma, Y. Intrahepatic cholangiocarcinoma escapes from growth inhibitory effect of transforming growth factor-beta1 by overexpression of cyclin d1. Lab. Investig. 2005, 85, 572–581. [Google Scholar] [CrossRef]
  60. Sritananuwat, P.; Sueangoen, N.; Thummarati, P.; Islam, K.; Suthiphongchai, T. Blocking erk1/2 signaling impairs tgf-beta1 tumor promoting function but enhances its tumor suppressing role in intrahepatic cholangiocarcinoma cells. Cancer Cell Int. 2017, 17, 85. [Google Scholar] [CrossRef]
  61. Yothaisong, S.; Dokduang, H.; Techasen, A.; Namwat, N.; Yongvanit, P.; Bhudhisawasdi, V.; Puapairoj, A.; Riggins, G.J.; Loilome, W. Increased activation of pi3k/akt signaling pathway is associated with cholangiocarcinoma metastasis and pi3k/mtor inhibition presents a possible therapeutic strategy. Tumour Biol. 2013, 34, 3637–3648. [Google Scholar] [CrossRef] [PubMed]
  62. Kokuryo, T.; Yokoyama, Y.; Nagino, M. Recent advances in cancer stem cell research for cholangiocarcinoma. J. Hepatobil. Pancreat. Sci. 2012, 19, 606–613. [Google Scholar] [CrossRef] [PubMed]
  63. Utispan, K.; Sonongbua, J.; Thuwajit, P.; Chau-In, S.; Pairojkul, C.; Wongkham, S.; Thuwajit, C. Periostin activates integrin alpha5beta1 through a pi3k/aktdependent pathway in invasion of cholangiocarcinoma. Int. J. Oncol. 2012, 41, 1110–1118. [Google Scholar] [CrossRef] [PubMed]
  64. Yoon, H.; Min, J.K.; Lee, J.W.; Kim, D.G.; Hong, H.J. Acquisition of chemoresistance in intrahepatic cholangiocarcinoma cells by activation of akt and extracellular signal-regulated kinase (erk)1/2. Biochem. Biophys. Res. Commun. 2011, 405, 333–337. [Google Scholar] [CrossRef]
  65. Zhang, Y.W.; Su, Y.; Volpert, O.V.; Vande Woude, G.F. Hepatocyte growth factor/scatter factor mediates angiogenesis through positive vegf and negative thrombospondin 1 regulation. Proc. Natl. Acad. Sci. USA 2003, 100, 12718–12723. [Google Scholar] [CrossRef] [PubMed]
  66. Miyamoto, M.; Ojima, H.; Iwasaki, M.; Shimizu, H.; Kokubu, A.; Hiraoka, N.; Kosuge, T.; Yoshikawa, D.; Kono, T.; Furukawa, H.; et al. Prognostic significance of overexpression of c-met oncoprotein in cholangiocarcinoma. Br. J. Cancer 2011, 105, 131–138. [Google Scholar] [CrossRef]
  67. Terada, T.; Nakanuma, Y.; Sirica, A.E. Immunohistochemical demonstration of met overexpression in human intrahepatic cholangiocarcinoma and in hepatolithiasis. Hum. Pathol. 1998, 29, 175–180. [Google Scholar]
  68. Menakongka, A.; Suthiphongchai, T. Involvement of pi3k and erk1/2 pathways in hepatocyte growth factor-induced cholangiocarcinoma cell invasion. World J. Gastroenterol. 2010, 16, 713–722. [Google Scholar] [CrossRef]
  69. Rincon, M.; Davis, R.J. Regulation of the immune response by stress-activated protein kinases. Immunol. Rev. 2009, 228, 212–224. [Google Scholar] [CrossRef]
  70. Bettermann, K.; Vucur, M.; Haybaeck, J.; Koppe, C.; Janssen, J.; Heymann, F.; Weber, A.; Weiskirchen, R.; Liedtke, C.; Gassler, N.; et al. Tak1 suppresses a nemo-dependent but nf-kappab-independent pathway to liver cancer. Cancer Cell 2010, 17, 481–496. [Google Scholar] [CrossRef]
  71. Malato, Y.; Willenbring, H. The map3k tak1: A chock block to liver cancer formation. Hepatology 2010, 52, 1506–1509. [Google Scholar] [CrossRef] [PubMed]
  72. Inokuchi, S.; Aoyama, T.; Miura, K.; Osterreicher, C.H.; Kodama, Y.; Miyai, K.; Akira, S.; Brenner, D.A.; Seki, E. Disruption of tak1 in hepatocytes causes hepatic injury, inflammation, fibrosis, and carcinogenesis. Proc. Natl. Acad. Sci. USA 2010, 107, 844–849. [Google Scholar] [CrossRef] [PubMed]
  73. Nygaard, G.; Di Paolo, J.A.; Hammaker, D.; Boyle, D.L.; Budas, G.; Notte, G.T.; Mikaelian, I.; Barry, V.; Firestein, G.S. Regulation and function of apoptosis signal-regulating kinase 1 in rheumatoid arthritis. Biochem. Pharmacol. 2018, 151, 282–290. [Google Scholar] [CrossRef] [PubMed]
  74. Fujino, G.; Noguchi, T.; Matsuzawa, A.; Yamauchi, S.; Saitoh, M.; Takeda, K.; Ichijo, H. Thioredoxin and traf family proteins regulate reactive oxygen species-dependent activation of ask1 through reciprocal modulation of the n-terminal homophilic interaction of ask1. Mol. Cell. Biol. 2007, 27, 8152–8163. [Google Scholar] [CrossRef] [PubMed]
  75. Hattori, K.; Naguro, I.; Runchel, C.; Ichijo, H. The roles of ask family proteins in stress responses and diseases. Cell Commun. Signal. 2009, 7, 9. [Google Scholar] [CrossRef]
  76. Glaser, S.S.; Gaudio, E.; Miller, T.; Alvaro, D.; Alpini, G. Cholangiocyte proliferation and liver fibrosis. Expert Rev. Mol. Med. 2009, 11, e7. [Google Scholar] [CrossRef] [PubMed]
  77. Roskams, T.; Desmet, V. Ductular reaction and its diagnostic significance. Semin. Diagn. Pathol. 1998, 15, 259–269. [Google Scholar] [PubMed]
  78. Diaz, R.; Kim, J.W.; Hui, J.J.; Li, Z.; Swain, G.P.; Fong, K.S.; Csiszar, K.; Russo, P.A.; Rand, E.B.; Furth, E.E.; et al. Evidence for the epithelial to mesenchymal transition in biliary atresia fibrosis. Hum. Pathol. 2008, 39, 102–115. [Google Scholar] [CrossRef] [PubMed]
  79. Burns, W.C.; Kantharidis, P.; Thomas, M.C. The role of tubular epithelial-mesenchymal transition in progressive kidney disease. Cells Tissues Organs 2007, 185, 222–231. [Google Scholar] [CrossRef] [PubMed]
  80. Robertson, H.; Ali, S.; McDonnell, B.J.; Burt, A.D.; Kirby, J.A. Chronic renal allograft dysfunction: The role of t cell-mediated tubular epithelial to mesenchymal cell transition. J. Am. Soc. Nephrol. 2004, 15, 390–397. [Google Scholar] [CrossRef]
  81. Roskams, T.A.; Theise, N.D.; Balabaud, C.; Bhagat, G.; Bhathal, P.S.; Bioulac-Sage, P.; Brunt, E.M.; Crawford, J.M.; Crosby, H.A.; Desmet, V.; et al. Nomenclature of the finer branches of the biliary tree: Canals, ductules, and ductular reactions in human livers. Hepatology 2004, 39, 1739–1745. [Google Scholar] [CrossRef]
  82. Alvaro, D.; Mancino, M.G.; Glaser, S.; Gaudio, E.; Marzioni, M.; Francis, H.; Alpini, G. Proliferating cholangiocytes: A neuroendocrine compartment in the diseased liver. Gastroenterology 2007, 132, 415–431. [Google Scholar] [CrossRef] [PubMed]
  83. Lee, J.S.; Heo, J.; Libbrecht, L.; Chu, I.S.; Kaposi-Novak, P.; Calvisi, D.F.; Mikaelyan, A.; Roberts, L.R.; Demetris, A.J.; Sun, Z.; et al. A novel prognostic subtype of human hepatocellular carcinoma derived from hepatic progenitor cells. Nat. Med. 2006, 12, 410–416. [Google Scholar] [CrossRef] [PubMed]
  84. Zhang, F.; Chen, X.P.; Zhang, W.; Dong, H.H.; Xiang, S.; Zhang, W.G.; Zhang, B.X. Combined hepatocellular cholangiocarcinoma originating from hepatic progenitor cells: Immunohistochemical and double-fluorescence immunostaining evidence. Histopathology 2008, 52, 224–232. [Google Scholar] [CrossRef] [PubMed]
  85. Kumar, M.; Zhao, X.; Wang, X.W. Molecular carcinogenesis of hepatocellular carcinoma and intrahepatic cholangiocarcinoma: One step closer to personalized medicine? Cell Biosci. 2011, 1, 5. [Google Scholar] [CrossRef]
  86. Planas-Paz, L.; Orsini, V.; Boulter, L.; Calabrese, D.; Pikiolek, M.; Nigsch, F.; Xie, Y.; Roma, G.; Donovan, A.; Marti, P.; et al. The rspo-lgr4/5-znrf3/rnf43 module controls liver zonation and size. Nat. Cell Biol. 2016, 18, 467–479. [Google Scholar] [CrossRef]
  87. Guest, R.V.; Boulter, L.; Dwyer, B.J.; Forbes, S.J. Understanding liver regeneration to bring new insights to the mechanisms driving cholangiocarcinoma. NPJ Regener. Med. 2017, 2, 13. [Google Scholar] [CrossRef] [Green Version]
  88. Zong, Y.; Panikkar, A.; Xu, J.; Antoniou, A.; Raynaud, P.; Lemaigre, F.; Stanger, B.Z. Notch signaling controls liver development by regulating biliary differentiation. Development 2009, 136, 1727–1739. [Google Scholar] [CrossRef] [Green Version]
  89. Theise, N.D.; Yao, J.L.; Harada, K.; Hytiroglou, P.; Portmann, B.; Thung, S.N.; Tsui, W.; Ohta, H.; Nakanuma, Y. Hepatic ‘stem cell’ malignancies in adults: Four cases. Histopathology 2003, 43, 263–271. [Google Scholar] [CrossRef]
  90. Chiba, T.; Kita, K.; Zheng, Y.W.; Yokosuka, O.; Saisho, H.; Iwama, A.; Nakauchi, H.; Taniguchi, H. Side population purified from hepatocellular carcinoma cells harbors cancer stem cell-like properties. Hepatology 2006, 44, 240–251. [Google Scholar] [CrossRef]
  91. Hay, E.D. Organization and fine structure of epithelium and mesenchyme in the developing chick embryo. Epithel. Mesenchymal Interact. 1968, 2, 31–35. [Google Scholar]
  92. Brabletz, T.; Kalluri, R.; Nieto, M.A.; Weinberg, R.A. Emt in cancer. Nat. Rev. Cancer 2018, 18, 128–134. [Google Scholar] [CrossRef] [PubMed]
  93. Zhang, Q.; Agoston, A.T.; Pham, T.H.; Zhang, W.; Zhang, X.; Huo, X.; Peng, S.; Bajpai, M.; Das, K.; Odze, R.D.; et al. Acidic bile salts induce epithelial to mesenchymal transition via vegf signaling in non-neoplastic barrett’s cells. Gastroenterology 2019, 156, 130–144.e110. [Google Scholar] [CrossRef] [PubMed]
  94. Nieto, M.A.; Huang, R.Y.; Jackson, R.A.; Thiery, J.P. Emt: 2016. Cell 2016, 166, 21–45. [Google Scholar] [CrossRef] [PubMed]
  95. Chaffer, C.L.; San Juan, B.P.; Lim, E.; Weinberg, R.A. Emt, cell plasticity and metastasis. Cancer Metastasis Rev. 2016, 35, 645–654. [Google Scholar] [CrossRef] [PubMed]
  96. Zhang, J.; Tian, X.J.; Xing, J. Signal transduction pathways of emt induced by tgf-beta, shh, and wnt and their crosstalks. J. Clin. Med. 2016, 5, 41. [Google Scholar] [CrossRef] [PubMed]
  97. Vaquero, J.; Guedj, N.; Claperon, A.; Nguyen Ho-Bouldoires, T.H.; Paradis, V.; Fouassier, L. Epithelial-mesenchymal transition in cholangiocarcinoma: From clinical evidence to regulatory networks. J. Hepatol. 2017, 66, 424–441. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Jakowlew, S.B. Transforming growth factor-beta in cancer and metastasis. Cancer Metastasis Rev. 2006, 25, 435–457. [Google Scholar] [CrossRef] [PubMed]
  99. Heldin, C.H.; Landstrom, M.; Moustakas, A. Mechanism of tgf-beta signaling to growth arrest, apoptosis, and epithelial-mesenchymal transition. Curr. Opin. Cell Biol. 2009, 21, 166–176. [Google Scholar] [CrossRef]
  100. Seoane, J. Escaping from the tgfbeta anti-proliferative control. Carcinogenesis 2006, 27, 2148–2156. [Google Scholar] [CrossRef]
  101. Sheahan, S.; Bellamy, C.O.; Dunbar, D.R.; Harrison, D.J.; Prost, S. Deficiency of g1 regulators p53, p21cip1 and/or prb decreases hepatocyte sensitivity to tgfbeta cell cycle arrest. BMC Cancer 2007, 7, 215. [Google Scholar] [CrossRef] [PubMed]
  102. Connolly, E.C.; Freimuth, J.; Akhurst, R.J. Complexities of tgf-beta targeted cancer therapy. Int. J. Biol. Sci. 2012, 8, 964–978. [Google Scholar] [CrossRef] [PubMed]
  103. Heldin, C.H.; Miyazono, K.; ten Dijke, P. Tgf-beta signalling from cell membrane to nucleus through smad proteins. Nature 1997, 390, 465–471. [Google Scholar] [CrossRef] [PubMed]
  104. Wu, J.W.; Fairman, R.; Penry, J.; Shi, Y. Formation of a stable heterodimer between smad2 and smad4. J. Biol. Chem. 2001, 276, 20688–20694. [Google Scholar] [CrossRef] [PubMed]
  105. Thuault, S.; Tan, E.J.; Peinado, H.; Cano, A.; Heldin, C.H.; Moustakas, A. Hmga2 and smads co-regulate snail1 expression during induction of epithelial-to-mesenchymal transition. J. Biol. Chem. 2008, 283, 33437–33446. [Google Scholar] [CrossRef] [PubMed]
  106. Brandl, M.; Seidler, B.; Haller, F.; Adamski, J.; Schmid, R.M.; Saur, D.; Schneider, G. Ikk(alpha) controls canonical tgf(ss)-smad signaling to regulate genes expressing snail and slug during emt in panc1 cells. J. Cell Sci. 2010, 123, 4231–4239. [Google Scholar] [CrossRef] [PubMed]
  107. Chacko, B.M.; Qin, B.Y.; Tiwari, A.; Shi, G.; Lam, S.; Hayward, L.J.; De Caestecker, M.; Lin, K. Structural basis of heteromeric smad protein assembly in tgf-beta signaling. Mol. Cell 2004, 15, 813–823. [Google Scholar] [CrossRef]
  108. Lagna, G.; Hata, A.; Hemmati-Brivanlou, A.; Massague, J. Partnership between dpc4 and smad proteins in tgf-beta signalling pathways. Nature 1996, 383, 832–836. [Google Scholar] [CrossRef] [PubMed]
  109. Zhang, Y.E. Non-smad pathways in tgf-beta signaling. Cell Res. 2009, 19, 128–139. [Google Scholar] [CrossRef]
  110. Yamamura, Y.; Hua, X.; Bergelson, S.; Lodish, H.F. Critical role of smads and ap-1 complex in transforming growth factor-beta-dependent apoptosis. J. Biol. Chem. 2000, 275, 36295–36302. [Google Scholar] [CrossRef]
  111. Yamashita, M.; Fatyol, K.; Jin, C.; Wang, X.; Liu, Z.; Zhang, Y.E. Traf6 mediates smad-independent activation of jnk and p38 by tgf-beta. Mol. Cell 2008, 31, 918–924. [Google Scholar] [CrossRef] [PubMed]
  112. Bakin, A.V.; Tomlinson, A.K.; Bhowmick, N.A.; Moses, H.L.; Arteaga, C.L. Phosphatidylinositol 3-kinase function is required for transforming growth factor beta-mediated epithelial to mesenchymal transition and cell migration. J. Biol. Chem. 2000, 275, 36803–36810. [Google Scholar] [CrossRef] [PubMed]
  113. Giannelli, G.; Koudelkova, P.; Dituri, F.; Mikulits, W. Role of epithelial to mesenchymal transition in hepatocellular carcinoma. J. Hepatol. 2016, 65, 798–808. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Tremblay, G. Stromal aspects of breast carcinoma. Exp. Mol. Pathol. 1979, 31, 248–260. [Google Scholar] [CrossRef]
  115. Kwa, M.Q.; Herum, K.M.; Brakebusch, C. Cancer-associated fibroblasts: How do they contribute to metastasis? Clin. Exp. Metastasis 2019, 36, 71–86. [Google Scholar] [CrossRef] [PubMed]
  116. Zeltz, C.; Primac, I.; Erusappan, P.; Alam, J.; Noel, A.; Gullberg, D. Cancer-associated fibroblasts in desmoplastic tumors: Emerging role of integrins. Semin. Cancer Biol. 2019. [Google Scholar] [CrossRef] [PubMed]
  117. Santi, A.; Kugeratski, F.G.; Zanivan, S. Cancer associated fibroblasts: The architects of stroma remodeling. Proteomics 2018, 18, e1700167. [Google Scholar] [CrossRef]
  118. Okabe, H.; Beppu, T.; Hayashi, H.; Ishiko, T.; Masuda, T.; Otao, R.; Horlad, H.; Jono, H.; Ueda, M.; Shinriki, S.; et al. Hepatic stellate cells accelerate the malignant behavior of cholangiocarcinoma cells. Ann. Surg. Oncol. 2011, 18, 1175–1184. [Google Scholar] [CrossRef]
  119. Itou, R.A.; Uyama, N.; Hirota, S.; Kawada, N.; Wu, S.; Miyashita, S.; Nakamura, I.; Suzumura, K.; Sueoka, H.; Okada, T.; et al. Immunohistochemical characterization of cancer-associated fibroblasts at the primary sites and in the metastatic lymph nodes of human intrahepatic cholangiocarcinoma. Hum. Pathol. 2019, 83, 77–89. [Google Scholar] [CrossRef]
  120. Sirica, A.E.; Campbell, D.J.; Dumur, C.I. Cancer-associated fibroblasts in intrahepatic cholangiocarcinoma. Curr. Opin. Gastroenterol. 2011, 27, 276–284. [Google Scholar] [CrossRef]
  121. Razumilava, N.; Gores, G.J. Cholangiocarcinoma. Lancet 2014, 383, 2168–2179. [Google Scholar] [CrossRef] [Green Version]
  122. Cohen, S.J.; Alpaugh, R.K.; Palazzo, I.; Meropol, N.J.; Rogatko, A.; Xu, Z.; Hoffman, J.P.; Weiner, L.M.; Cheng, J.D. Fibroblast activation protein and its relationship to clinical outcome in pancreatic adenocarcinoma. Pancreas 2008, 37, 154–158. [Google Scholar] [CrossRef] [PubMed]
  123. Komohara, Y.; Takeya, M. Cafs and tams: Maestros of the tumour microenvironment. J. Pathol. 2017, 241, 313–315. [Google Scholar] [CrossRef] [PubMed]
  124. Hashimoto, O.; Yoshida, M.; Koma, Y.; Yanai, T.; Hasegawa, D.; Kosaka, Y.; Nishimura, N.; Yokozaki, H. Collaboration of cancer-associated fibroblasts and tumour-associated macrophages for neuroblastoma development. J. Pathol. 2016, 240, 211–223. [Google Scholar] [CrossRef] [PubMed]
  125. Techasen, A.; Loilome, W.; Namwat, N.; Dokduang, H.; Jongthawin, J.; Yongvanit, P. Cytokines released from activated human macrophages induce epithelial mesenchymal transition markers of cholangiocarcinoma cells. Asian Pac. J. Cancer Prev. 2012, 13, 115–118. [Google Scholar] [PubMed]
  126. Berasain, C.; Avila, M.A. Platelet-derived growth factor d: A new player in the complex cross-talk between cholangiocarcinoma cells and cancer-associated fibroblasts. Hepatology 2013, 58, 853–855. [Google Scholar] [CrossRef] [PubMed]
  127. Raggi, C.; Invernizzi, P.; Andersen, J.B. Impact of microenvironment and stem-like plasticity in cholangiocarcinoma: Molecular networks and biological concepts. J. Hepatol. 2015, 62, 198–207. [Google Scholar] [CrossRef] [Green Version]
  128. Bansal, R.; van Baarlen, J.; Storm, G.; Prakash, J. The interplay of the notch signaling in hepatic stellate cells and macrophages determines the fate of liver fibrogenesis. Sci. Rep. 2015, 5, 18272. [Google Scholar] [CrossRef] [PubMed]
  129. Mauer, J.; Chaurasia, B.; Goldau, J.; Vogt, M.C.; Ruud, J.; Nguyen, K.D.; Theurich, S.; Hausen, A.C.; Schmitz, J.; Bronneke, H.S.; et al. Signaling by il-6 promotes alternative activation of macrophages to limit endotoxemia and obesity-associated resistance to insulin. Nat. Immunol. 2014, 15, 423–430. [Google Scholar] [CrossRef]
  130. Hogdall, D.; Lewinska, M.; Andersen, J.B. Desmoplastic tumor microenvironment and immunotherapy in cholangiocarcinoma. Trends Cancer 2018, 4, 239–255. [Google Scholar] [CrossRef]
  131. Zhou, S.L.; Dai, Z.; Zhou, Z.J.; Chen, Q.; Wang, Z.; Xiao, Y.S.; Hu, Z.Q.; Huang, X.Y.; Yang, G.H.; Shi, Y.H.; et al. Cxcl5 contributes to tumor metastasis and recurrence of intrahepatic cholangiocarcinoma by recruiting infiltrative intratumoral neutrophils. Carcinogenesis 2014, 35, 597–605. [Google Scholar] [CrossRef] [PubMed]
  132. Cadamuro, M.; Brivio, S.; Spirli, C.; Joplin, R.E.; Strazzabosco, M.; Fabris, L. Autocrine and paracrine mechanisms promoting chemoresistance in cholangiocarcinoma. Int. J. Mol. Sci. 2017, 18, 149. [Google Scholar] [CrossRef] [PubMed]
  133. Fan, B.; Malato, Y.; Calvisi, D.F.; Naqvi, S.; Razumilava, N.; Ribback, S.; Gores, G.J.; Dombrowski, F.; Evert, M.; Chen, X.; et al. Cholangiocarcinomas can originate from hepatocytes in mice. J. Clin. Investig. 2012, 122, 2911–2915. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Sekiya, S.; Suzuki, A. Intrahepatic cholangiocarcinoma can arise from notch-mediated conversion of hepatocytes. J. Clin. Investig. 2012, 122, 3914–3918. [Google Scholar] [CrossRef] [PubMed]
  135. Thongsom, S.; Suginta, W.; Lee, K.J.; Choe, H.; Talabnin, C. Piperlongumine induces g2/m phase arrest and apoptosis in cholangiocarcinoma cells through the ros-jnk-erk signaling pathway. Apoptosis 2017, 22, 1473–1484. [Google Scholar] [CrossRef] [PubMed]
  136. Huntzicker, E.G.; Hotzel, K.; Choy, L.; Che, L.; Ross, J.; Pau, G.; Sharma, N.; Siebel, C.W.; Chen, X.; French, D.M. Differential effects of targeting notch receptors in a mouse model of liver cancer. Hepatology 2015, 61, 942–952. [Google Scholar] [CrossRef] [PubMed]
  137. Greenhill, C. Liver cancer: Different effects of the notch receptors in liver cancer revealed. Nat. Rev. Gastroenterol. Hepatol. 2014, 11, 703. [Google Scholar] [PubMed]
  138. Wu, W.R.; Zhang, R.; Shi, X.D.; Zhu, M.S.; Xu, L.B.; Zeng, H.; Liu, C. Notch1 is overexpressed in human intrahepatic cholangiocarcinoma and is associated with its proliferation, invasiveness and sensitivity to 5-fluorouracil in vitro. Oncol. Rep. 2014, 31, 2515–2524. [Google Scholar] [CrossRef]
  139. Thiery, J.P.; Acloque, H.; Huang, R.Y.; Nieto, M.A. Epithelial-mesenchymal transitions in development and disease. Cell 2009, 139, 871–890. [Google Scholar] [CrossRef]
  140. Aguirre-Ghiso, J.A. Models, mechanisms and clinical evidence for cancer dormancy. Nat. Rev. Cancer 2007, 7, 834–846. [Google Scholar] [CrossRef] [Green Version]
  141. Hong, J.; Zhou, J.; Fu, J.; He, T.; Qin, J.; Wang, L.; Liao, L.; Xu, J. Phosphorylation of serine 68 of twist1 by mapks stabilizes twist1 protein and promotes breast cancer cell invasiveness. Cancer Res. 2011, 71, 3980–3990. [Google Scholar] [CrossRef] [PubMed]
  142. Radisky, D.C.; Levy, D.D.; Littlepage, L.E.; Liu, H.; Nelson, C.M.; Fata, J.E.; Leake, D.; Godden, E.L.; Albertson, D.G.; Nieto, M.A.; et al. Rac1b and reactive oxygen species mediate mmp-3-induced emt and genomic instability. Nature 2005, 436, 123–127. [Google Scholar] [CrossRef] [PubMed]
  143. Dolado, I.; Swat, A.; Ajenjo, N.; De Vita, G.; Cuadrado, A.; Nebreda, A.R. P38alpha map kinase as a sensor of reactive oxygen species in tumorigenesis. Cancer Cell 2007, 11, 191–205. [Google Scholar] [CrossRef] [PubMed]
  144. Cuenda, A.; Rousseau, S. P38 map-kinases pathway regulation, function and role in human diseases. Biochim. Biophys. Acta 2007, 1773, 1358–1375. [Google Scholar] [CrossRef]
  145. Papa, S.; Bubici, C.; Zazzeroni, F.; Pham, C.G.; Kuntzen, C.; Knabb, J.R.; Dean, K.; Franzoso, G. The nf-kappab-mediated control of the jnk cascade in the antagonism of programmed cell death in health and disease. Cell Death Differ. 2006, 13, 712–729. [Google Scholar] [CrossRef] [PubMed]
  146. Kamata, H.; Honda, S.; Maeda, S.; Chang, L.; Hirata, H.; Karin, M. Reactive oxygen species promote tnfalpha-induced death and sustained jnk activation by inhibiting map kinase phosphatases. Cell 2005, 120, 649–661. [Google Scholar] [CrossRef] [PubMed]
  147. Perkins, N.D. Integrating cell-signalling pathways with nf-kappab and ikk function. Nat. Rev. Mol. Cell Biol. 2007, 8, 49–62. [Google Scholar] [CrossRef]
  148. Saccani, S.; Pantano, S.; Natoli, G. P38-dependent marking of inflammatory genes for increased nf-kappa b recruitment. Nat. Immunol. 2002, 3, 69–75. [Google Scholar] [CrossRef]
  149. Kojima, T.; Fuchimoto, J.; Yamaguchi, H.; Ito, T.; Takasawa, A.; Ninomiya, T.; Kikuchi, S.; Ogasawara, N.; Ohkuni, T.; Masaki, T.; et al. C-jun n-terminal kinase is largely involved in the regulation of tricellular tight junctions via tricellulin in human pancreatic duct epithelial cells. J. Cell Physiol. 2010, 225, 720–733. [Google Scholar] [CrossRef]
  150. Xia, Y.; Karin, M. The control of cell motility and epithelial morphogenesis by jun kinases. Trends Cell Biol. 2004, 14, 94–101. [Google Scholar] [CrossRef]
  151. Kamiya, A.; Gonzalez, F.J. Tnf-alpha regulates mouse fetal hepatic maturation induced by oncostatin m and extracellular matrices. Hepatology 2004, 40, 527–536. [Google Scholar] [CrossRef] [PubMed]
  152. Varfolomeev, E.E.; Ashkenazi, A. Tumor necrosis factor: An apoptosis junkie? Cell 2004, 116, 491–497. [Google Scholar] [CrossRef]
  153. Morgan, M.J.; Liu, Z.G. Reactive oxygen species in tnfalpha-induced signaling and cell death. Mol. Cells 2010, 30, 1–12. [Google Scholar] [CrossRef] [PubMed]
  154. Seubwai, W.; Wongkham, C.; Puapairoj, A.; Khuntikeo, N.; Pugkhem, A.; Hahnvajanawong, C.; Chaiyagool, J.; Umezawa, K.; Okada, S.; Wongkham, S. Aberrant expression of nf-kappab in liver fluke associated cholangiocarcinoma: Implications for targeted therapy. PLoS ONE 2014, 9, e106056. [Google Scholar]
  155. Sirica, A.E. The role of cancer-associated myofibroblasts in intrahepatic cholangiocarcinoma. Nat. Rev. Gastroenterol. Hepatol. 2011, 9, 44–54. [Google Scholar] [CrossRef]
  156. Claperon, A.; Mergey, M.; Aoudjehane, L.; Ho-Bouldoires, T.H.; Wendum, D.; Prignon, A.; Merabtene, F.; Firrincieli, D.; Desbois-Mouthon, C.; Scatton, O.; et al. Hepatic myofibroblasts promote the progression of human cholangiocarcinoma through activation of epidermal growth factor receptor. Hepatology 2013, 58, 2001–2011. [Google Scholar] [CrossRef]
  157. Claperon, A.; Guedj, N.; Mergey, M.; Vignjevic, D.; Desbois-Mouthon, C.; Boissan, M.; Saubamea, B.; Paradis, V.; Housset, C.; Fouassier, L. Loss of ebp50 stimulates egfr activity to induce emt phenotypic features in biliary cancer cells. Oncogene 2012, 31, 1376–1388. [Google Scholar] [CrossRef]
  158. Endo, K.; Yoon, B.I.; Pairojkul, C.; Demetris, A.J.; Sirica, A.E. Erbb-2 overexpression and cyclooxygenase-2 up-regulation in human cholangiocarcinoma and risk conditions. Hepatology 2002, 36, 439–450. [Google Scholar] [CrossRef]
  159. Han, C.; Leng, J.; Demetris, A.J.; Wu, T. Cyclooxygenase-2 promotes human cholangiocarcinoma growth: Evidence for cyclooxygenase-2-independent mechanism in celecoxib-mediated induction of p21waf1/cip1 and p27kip1 and cell cycle arrest. Cancer Res. 2004, 64, 1369–1376. [Google Scholar] [CrossRef]
  160. Wehbe, H.; Henson, R.; Meng, F.; Mize-Berge, J.; Patel, T. Interleukin-6 contributes to growth in cholangiocarcinoma cells by aberrant promoter methylation and gene expression. Cancer Res. 2006, 66, 10517–10524. [Google Scholar] [CrossRef]
  161. Sae-Lao, T.; Tohtong, R.; Bates, D.O.; Wongprasert, K. Sulfated galactans from red seaweed gracilaria fisheri target egfr and inhibit cholangiocarcinoma cell proliferation. Am. J. Chin. Med. 2017, 45, 615–633. [Google Scholar] [CrossRef] [PubMed]
  162. Ong, C.K.; Subimerb, C.; Pairojkul, C.; Wongkham, S.; Cutcutache, I.; Yu, W.; McPherson, J.R.; Allen, G.E.; Ng, C.C.; Wong, B.H.; et al. Exome sequencing of liver fluke-associated cholangiocarcinoma. Nat. Genet. 2012, 44, 690–693. [Google Scholar] [CrossRef]
  163. Jiao, Y.; Pawlik, T.M.; Anders, R.A.; Selaru, F.M.; Streppel, M.M.; Lucas, D.J.; Niknafs, N.; Guthrie, V.B.; Maitra, A.; Argani, P.; et al. Exome sequencing identifies frequent inactivating mutations in bap1, arid1a and pbrm1 in intrahepatic cholangiocarcinomas. Nat. Genet. 2013, 45, 1470–1473. [Google Scholar] [CrossRef] [PubMed]
  164. Macias, R.I.R.; Banales, J.M.; Sangro, B.; Muntane, J.; Avila, M.A.; Lozano, E.; Perugorria, M.J.; Padillo, F.J.; Bujanda, L.; Marin, J.J.G. The search for novel diagnostic and prognostic biomarkers in cholangiocarcinoma. Biochim. Biophys. Acta Mol. Basis Dis. 2018, 1864, 1468–1477. [Google Scholar] [CrossRef] [PubMed]
  165. Rizvi, S.; Borad, M.J.; Patel, T.; Gores, G.J. Cholangiocarcinoma: Molecular pathways and therapeutic opportunities. Semin. Liver Dis. 2014, 34, 456–464. [Google Scholar] [CrossRef] [PubMed]
  166. Ross, J.S.; Wang, K.; Gay, L.; Al-Rohil, R.; Rand, J.V.; Jones, D.M.; Lee, H.J.; Sheehan, C.E.; Otto, G.A.; Palmer, G.; et al. New routes to targeted therapy of intrahepatic cholangiocarcinomas revealed by next-generation sequencing. Oncologist 2014, 19, 235–242. [Google Scholar] [CrossRef] [PubMed]
  167. Farshidfar, F.; Zheng, S.; Gingras, M.C.; Newton, Y.; Shih, J.; Robertson, A.G.; Hinoue, T.; Hoadley, K.A.; Gibb, E.A.; Roszik, J.; et al. Integrative genomic analysis of cholangiocarcinoma identifies distinct idh-mutant molecular profiles. Cell Rep. 2017, 18, 2780–2794. [Google Scholar] [CrossRef]
  168. Churi, C.R.; Shroff, R.; Wang, Y.; Rashid, A.; Kang, H.C.; Weatherly, J.; Zuo, M.; Zinner, R.; Hong, D.; Meric-Bernstam, F.; et al. Mutation profiling in cholangiocarcinoma: Prognostic and therapeutic implications. PLoS ONE 2014, 9, e115383. [Google Scholar] [CrossRef]
  169. Andersen, J.B.; Spee, B.; Blechacz, B.R.; Avital, I.; Komuta, M.; Barbour, A.; Conner, E.A.; Gillen, M.C.; Roskams, T.; Roberts, L.R.; et al. Genomic and genetic characterization of cholangiocarcinoma identifies therapeutic targets for tyrosine kinase inhibitors. Gastroenterology 2012, 142, 1021–1031 e1015. [Google Scholar] [CrossRef]
  170. Chen, T.C.; Jan, Y.Y.; Yeh, T.S. K-ras mutation is strongly associated with perineural invasion and represents an independent prognostic factor of intrahepatic cholangiocarcinoma after hepatectomy. Ann. Surg. Oncol. 2012, 19 (Suppl. 3), S675–S681. [Google Scholar] [CrossRef]
  171. Isa, T.; Tomita, S.; Nakachi, A.; Miyazato, H.; Shimoji, H.; Kusano, T.; Muto, Y.; Furukawa, M. Analysis of microsatellite instability, k-ras gene mutation and p53 protein overexpression in intrahepatic cholangiocarcinoma. Hepatogastroenterology 2002, 49, 604–608. [Google Scholar] [PubMed]
  172. Junttila, M.R.; de Sauvage, F.J. Influence of tumour micro-environment heterogeneity on therapeutic response. Nature 2013, 501, 346–354. [Google Scholar] [CrossRef] [PubMed]
  173. Jiang, K.; Al-Diffhala, S.; Centeno, B.A. Primary liver cancers-part 1: Histopathology, differential diagnoses, and risk stratification. Cancer Control 2018, 25, 1073274817744625. [Google Scholar] [CrossRef] [PubMed]
  174. Chan, E.S.; Yeh, M.M. The use of immunohistochemistry in liver tumors. Clin. Liver Dis. 2010, 14, 687–703. [Google Scholar] [CrossRef] [PubMed]
  175. Bioulac-Sage, P.; Cubel, G.; Taouji, S.; Scoazec, J.Y.; Leteurtre, E.; Paradis, V.; Sturm, N.; Nhieu, J.T.; Wendum, D.; Bancel, B.; et al. Immunohistochemical markers on needle biopsies are helpful for the diagnosis of focal nodular hyperplasia and hepatocellular adenoma subtypes. Am. J. Surg. Pathol. 2012, 36, 1691–1699. [Google Scholar] [CrossRef] [PubMed]
  176. Han, C.P.; Hsu, J.D.; Koo, C.L.; Yang, S.F. Antibody to cytokeratin (ck8/ck18) is not derived from cam5.2 clone, and anticytokeratin cam5.2 (becton dickinson) is not synonymous with the antibody (ck8/ck18). Hum. Pathol. 2010, 41, 616–617, author reply 617. [Google Scholar] [CrossRef]
  177. Razumilava, N.; Gores, G.J. Notch-driven carcinogenesis: The merging of hepatocellular cancer and cholangiocarcinoma into a common molecular liver cancer subtype. J. Hepatol. 2013, 58, 1244–1245. [Google Scholar] [CrossRef] [Green Version]
  178. Matsushima, H.; Kuroki, T.; Kitasato, A.; Adachi, T.; Tanaka, T.; Hirabaru, M.; Hirayama, T.; Kuroshima, N.; Hidaka, M.; Soyama, A.; et al. Sox9 expression in carcinogenesis and its clinical significance in intrahepatic cholangiocarcinoma. Dig. Liver Dis. 2015, 47, 1067–1075. [Google Scholar] [CrossRef]
  179. Ferrone, C.R.; Ting, D.T.; Shahid, M.; Konstantinidis, I.T.; Sabbatino, F.; Goyal, L.; Rice-Stitt, T.; Mubeen, A.; Arora, K.; Bardeesey, N.; et al. The ability to diagnose intrahepatic cholangiocarcinoma definitively using novel branched DNA-enhanced albumin rna in situ hybridization technology. Ann. Surg. Oncol. 2016, 23, 290–296. [Google Scholar] [CrossRef]
  180. Macias, R.I.R.; Kornek, M.; Rodrigues, P.M.; Paiva, N.A.; Castro, R.E.; Urban, S.; Pereira, S.P.; Cadamuro, M.; Rupp, C.; Loosen, S.H.; et al. Diagnostic and prognostic biomarkers in cholangiocarcinoma. Liver Int. 2019, 39, 108–122. [Google Scholar] [CrossRef] [Green Version]
  181. Forner, A.; Vidili, G.; Rengo, M.; Bujanda, L.; Ponz-Sarvise, M.; Lamarca, A. Clinical presentation, diagnosis and staging of cholangiocarcinoma. Liver Int. 2019, 39, 98–107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Jarnagin, W.R.; Fong, Y.; DeMatteo, R.P.; Gonen, M.; Burke, E.C.; Bodniewicz, B.J.; Youssef, B.M.; Klimstra, D.; Blumgart, L.H. Staging, resectability, and outcome in 225 patients with hilar cholangiocarcinoma. Ann. Surg. 2001, 234, 507–517. [Google Scholar] [CrossRef] [PubMed]
  183. Nagino, M.; Ebata, T.; Yokoyama, Y.; Igami, T.; Sugawara, G.; Takahashi, Y.; Nimura, Y. Evolution of surgical treatment for perihilar cholangiocarcinoma: A single-center 34-year review of 574 consecutive resections. Ann. Surg. 2013, 258, 129–140. [Google Scholar] [CrossRef] [PubMed]
  184. Endo, I.; Gonen, M.; Yopp, A.C.; Dalal, K.M.; Zhou, Q.; Klimstra, D.; D’Angelica, M.; DeMatteo, R.P.; Fong, Y.; Schwartz, L.; et al. Intrahepatic cholangiocarcinoma: Rising frequency, improved survival, and determinants of outcome after resection. Ann. Surg. 2008, 248, 84–96. [Google Scholar] [CrossRef] [PubMed]
  185. Weber, S.M.; Jarnagin, W.R.; Klimstra, D.; DeMatteo, R.P.; Fong, Y.; Blumgart, L.H. Intrahepatic cholangiocarcinoma: Resectability, recurrence pattern, and outcomes. J. Am. Coll. Surg. 2001, 193, 384–391. [Google Scholar] [CrossRef]
  186. Blechacz, B. Cholangiocarcinoma: Current knowledge and new developments. Gut Liver 2017, 11, 13–26. [Google Scholar] [CrossRef] [PubMed]
  187. Darwish Murad, S.; Kim, W.R.; Therneau, T.; Gores, G.J.; Rosen, C.B.; Martenson, J.A.; Alberts, S.R.; Heimbach, J.K. Predictors of pretransplant dropout and posttransplant recurrence in patients with perihilar cholangiocarcinoma. Hepatology 2012, 56, 972–981. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Hong, J.C.; Jones, C.M.; Duffy, J.P.; Petrowsky, H.; Farmer, D.G.; French, S.; Finn, R.; Durazo, F.A.; Saab, S.; Tong, M.J.; et al. Comparative analysis of resection and liver transplantation for intrahepatic and hilar cholangiocarcinoma: A 24-year experience in a single center. Arch. Surg. 2011, 146, 683–689. [Google Scholar] [CrossRef]
  189. Mansour, J.C.; Aloia, T.A.; Crane, C.H.; Heimbach, J.K.; Nagino, M.; Vauthey, J.N. Hilar cholangiocarcinoma: Expert consensus statement. HPB 2015, 17, 691–699. [Google Scholar]
  190. Song, S.C.; Choi, D.W.; Kow, A.W.; Choi, S.H.; Heo, J.S.; Kim, W.S.; Kim, M.J. Surgical outcomes of 230 resected hilar cholangiocarcinoma in a single centre. ANZ J. Surg. 2013, 83, 268–274. [Google Scholar] [CrossRef]
  191. Cardinale, V.; Carpino, G.; Reid, L.; Gaudio, E.; Alvaro, D. Multiple cells of origin in cholangiocarcinoma underlie biological, epidemiological and clinical heterogeneity. World J. Gastrointest. Oncol. 2012, 4, 94–102. [Google Scholar] [CrossRef] [PubMed]
  192. Nakanuma, Y.; Sato, Y.; Harada, K.; Sasaki, M.; Xu, J.; Ikeda, H. Pathological classification of intrahepatic cholangiocarcinoma based on a new concept. World J. Hepatol. 2010, 2, 419–427. [Google Scholar] [CrossRef] [PubMed]
  193. Valle, J.; Wasan, H.; Palmer, D.H.; Cunningham, D.; Anthoney, A.; Maraveyas, A.; Madhusudan, S.; Iveson, T.; Hughes, S.; Pereira, S.P.; et al. Cisplatin plus gemcitabine versus gemcitabine for biliary tract cancer. N. Engl. J. Med. 2010, 362, 1273–1281. [Google Scholar] [CrossRef] [PubMed]
  194. Park, K.; Kim, K.P.; Park, S.; Chang, H.M. Comparison of gemcitabine plus cisplatin versus capecitabine plus cisplatin as first-line chemotherapy for advanced biliary tract cancer. Asia Pac. J. Clin. Oncol. 2017, 13, 13–20. [Google Scholar] [CrossRef] [PubMed]
  195. Lubner, S.J.; Mahoney, M.R.; Kolesar, J.L.; Loconte, N.K.; Kim, G.P.; Pitot, H.C.; Philip, P.A.; Picus, J.; Yong, W.P.; Horvath, L.; et al. Report of a multicenter phase ii trial testing a combination of biweekly bevacizumab and daily erlotinib in patients with unresectable biliary cancer: A phase ii consortium study. J. Clin. Oncol. 2010, 28, 3491–3497. [Google Scholar] [CrossRef] [PubMed]
  196. Lee, J.; Park, S.H.; Chang, H.M.; Kim, J.S.; Choi, H.J.; Lee, M.A.; Jang, J.S.; Jeung, H.C.; Kang, J.H.; Lee, H.W.; et al. Gemcitabine and oxaliplatin with or without erlotinib in advanced biliary-tract cancer: A multicentre, open-label, randomised, phase 3 study. Lancet Oncol. 2012, 13, 181–188. [Google Scholar] [CrossRef]
  197. Crane, C.H.; Koay, E.J. Solutions that enable ablative radiotherapy for large liver tumors: Fractionated dose painting, simultaneous integrated protection, motion management, and computed tomography image guidance. Cancer 2016, 122, 1974–1986. [Google Scholar] [CrossRef] [PubMed]
  198. Hong, T.S.; Wo, J.Y.; Yeap, B.Y.; Ben-Josef, E.; McDonnell, E.I.; Blaszkowsky, L.S.; Kwak, E.L.; Allen, J.N.; Clark, J.W.; Goyal, L.; et al. Multi-institutional phase ii study of high-dose hypofractionated proton beam therapy in patients with localized, unresectable hepatocellular carcinoma and intrahepatic cholangiocarcinoma. J. Clin. Oncol. 2016, 34, 460–468. [Google Scholar] [CrossRef] [PubMed]
  199. Patel, S.; Ragab, O.; Kamrava, M. Another solution that enables ablative radiotherapy for large liver tumors: Percutaneous interstitial high-dose rate brachytherapy. Cancer 2016, 122, 2766. [Google Scholar] [CrossRef]
  200. Mukewar, S.; Gupta, A.; Baron, T.H.; Gores, G.; Furutani, K.; Haddock, M.G.; Hallemeier, C.L. Endoscopically inserted nasobiliary catheters for high dose-rate brachytherapy as part of neoadjuvant therapy for perihilar cholangiocarcinoma. Endoscopy 2015, 47, 878–883. [Google Scholar] [CrossRef] [PubMed]
  201. Javle, M.; Lowery, M.; Shroff, R.T.; Weiss, K.H.; Springfeld, C.; Borad, M.J.; Ramanathan, R.K.; Goyal, L.; Sadeghi, S.; Macarulla, T.; et al. Phase ii study of bgj398 in patients with fgfr-altered advanced cholangiocarcinoma. J. Clin. Oncol. 2018, 36, 276–282. [Google Scholar] [CrossRef] [PubMed]
  202. Guagnano, V.; Kauffmann, A.; Wohrle, S.; Stamm, C.; Ito, M.; Barys, L.; Pornon, A.; Yao, Y.; Li, F.; Zhang, Y.; et al. Fgfr genetic alterations predict for sensitivity to nvp-bgj398, a selective pan-fgfr inhibitor. Cancer Discov. 2012, 2, 1118–1133. [Google Scholar] [CrossRef] [PubMed]
  203. Rizvi, S.; Yamada, D.; Hirsova, P.; Bronk, S.F.; Werneburg, N.W.; Krishnan, A.; Salim, W.; Zhang, L.; Trushina, E.; Truty, M.J.; et al. A hippo and fibroblast growth factor receptor autocrine pathway in cholangiocarcinoma. J. Biol. Chem. 2016, 291, 8031–8047. [Google Scholar] [CrossRef] [PubMed]
  204. Tabernero, J.; Bahleda, R.; Dienstmann, R.; Infante, J.R.; Mita, A.; Italiano, A.; Calvo, E.; Moreno, V.; Adamo, B.; Gazzah, A.; et al. Phase i dose-escalation study of jnj-42756493, an oral pan-fibroblast growth factor receptor inhibitor, in patients with advanced solid tumors. J. Clin. Oncol. 2015, 33, 3401–3408. [Google Scholar] [CrossRef]
  205. Arai, Y.; Totoki, Y.; Hosoda, F.; Shirota, T.; Hama, N.; Nakamura, H.; Ojima, H.; Furuta, K.; Shimada, K.; Okusaka, T.; et al. Fibroblast growth factor receptor 2 tyrosine kinase fusions define a unique molecular subtype of cholangiocarcinoma. Hepatology 2014, 59, 1427–1434. [Google Scholar] [CrossRef] [PubMed]
  206. Ewald, F.; Norz, D.; Grottke, A.; Hofmann, B.T.; Nashan, B.; Jucker, M. Dual inhibition of pi3k-akt-mtor- and raf-mek-erk-signaling is synergistic in cholangiocarcinoma and reverses acquired resistance to mek-inhibitors. Investig. New Drugs 2014, 32, 1144–1154. [Google Scholar] [CrossRef]
  207. Finn, R.S.; Ahn, D.H.; Javle, M.M.; Tan, B.R., Jr.; Weekes, C.D.; Bendell, J.C.; Patnaik, A.; Khan, G.N.; Laheru, D.; Chavira, R.; et al. Phase 1b investigation of the mek inhibitor binimetinib in patients with advanced or metastatic biliary tract cancer. Investig. New Drugs 2018, 36, 1037–1043. [Google Scholar] [CrossRef] [PubMed]
  208. Wang, P.; Song, X.; Utpatel, K.; Shang, R.; Yang, Y.M.; Xu, M.; Zhang, J.; Che, L.; Gordan, J.; Cigliano, A.; et al. Mek inhibition suppresses k-ras wild-type cholangiocarcinoma in vitro and in vivo via inhibiting cell proliferation and modulating tumor microenvironment. Cell Death Dis. 2019, 10, 120. [Google Scholar] [CrossRef]
  209. Goyal, L.; Zheng, H.; Yurgelun, M.B.; Abrams, T.A.; Allen, J.N.; Cleary, J.M.; Knowles, M.; Regan, E.; Reardon, A.; Khachatryan, A.; et al. A phase 2 and biomarker study of cabozantinib in patients with advanced cholangiocarcinoma. Cancer 2017, 123, 1979–1988. [Google Scholar] [CrossRef] [Green Version]
  210. Shroff, R.T.; Yarchoan, M.; O’Connor, A.; Gallagher, D.; Zahurak, M.L.; Rosner, G.; Ohaji, C.; Sartorius-Mergenthaler, S.; Parkinson, R.; Subbiah, V.; et al. The oral vegf receptor tyrosine kinase inhibitor pazopanib in combination with the mek inhibitor trametinib in advanced cholangiocarcinoma. Br. J. Cancer 2017, 116, 1402–1407. [Google Scholar] [CrossRef]
  211. Yokoi, K.; Kobayashi, A.; Motoyama, H.; Kitazawa, M.; Shimizu, A.; Notake, T.; Yokoyama, T.; Matsumura, T.; Takeoka, M.; Miyagawa, S.I. Survival pathway of cholangiocarcinoma via akt/mtor signaling to escape raf/mek/erk pathway inhibition by sorafenib. Oncol. Rep. 2018, 39, 843–850. [Google Scholar] [CrossRef] [PubMed]
  212. El-Khoueiry, A.B.; Rankin, C.; Siegel, A.B.; Iqbal, S.; Gong, I.Y.; Micetich, K.C.; Kayaleh, O.R.; Lenz, H.J.; Blanke, C.D. S0941: A phase 2 swog study of sorafenib and erlotinib in patients with advanced gallbladder carcinoma or cholangiocarcinoma. Br. J. Cancer 2014, 110, 882–887. [Google Scholar] [CrossRef] [PubMed]
  213. Lozano, E.; Macias, R.I.R.; Monte, M.J.; Asensio, M.; Del Carmen, S.; Sanchez-Vicente, L.; Alonso-Pena, M.; Al-Abdulla, R.; Munoz-Garrido, P.; Satriano, L.; et al. Causes of hoct1-dependent cholangiocarcinoma resistance to sorafenib and sensitization by tumor-selective gene therapy. Hepatology 2019. [Google Scholar] [CrossRef] [PubMed]
  214. Marin, J.J.G.; Lozano, E.; Herraez, E.; Asensio, M.; Di Giacomo, S.; Romero, M.R.; Briz, O.; Serrano, M.A.; Efferth, T.; Macias, R.I.R. Chemoresistance and chemosensitization in cholangiocarcinoma. Biochim. Biophys. Acta Mol. Basis Dis. 2018, 1864, 1444–1453. [Google Scholar] [CrossRef] [PubMed]
  215. O’Rourke, C.J.; Lafuente-Barquero, J.; Andersen, J.B. Epigenome remodeling in cholangiocarcinoma. Trends Cancer 2019, 5, 335–350. [Google Scholar] [CrossRef] [PubMed]
  216. Borger, D.R.; Tanabe, K.K.; Fan, K.C.; Lopez, H.U.; Fantin, V.R.; Straley, K.S.; Schenkein, D.P.; Hezel, A.F.; Ancukiewicz, M.; Liebman, H.M.; et al. Frequent mutation of isocitrate dehydrogenase (idh)1 and idh2 in cholangiocarcinoma identified through broad-based tumor genotyping. Oncologist 2012, 17, 72–79. [Google Scholar] [CrossRef] [PubMed]
  217. Wang, F.; Travins, J.; DeLaBarre, B.; Penard-Lacronique, V.; Schalm, S.; Hansen, E.; Straley, K.; Kernytsky, A.; Liu, W.; Gliser, C.; et al. Targeted inhibition of mutant idh2 in leukemia cells induces cellular differentiation. Science 2013, 340, 622–626. [Google Scholar] [CrossRef] [PubMed]
  218. Wang, P.; Dong, Q.; Zhang, C.; Kuan, P.F.; Liu, Y.; Jeck, W.R.; Andersen, J.B.; Jiang, W.; Savich, G.L.; Tan, T.X.; et al. Mutations in isocitrate dehydrogenase 1 and 2 occur frequently in intrahepatic cholangiocarcinomas and share hypermethylation targets with glioblastomas. Oncogene 2013, 32, 3091–3100. [Google Scholar] [CrossRef] [PubMed]
  219. Rohle, D.; Popovici-Muller, J.; Palaskas, N.; Turcan, S.; Grommes, C.; Campos, C.; Tsoi, J.; Clark, O.; Oldrini, B.; Komisopoulou, E.; et al. An inhibitor of mutant idh1 delays growth and promotes differentiation of glioma cells. Science 2013, 340, 626–630. [Google Scholar] [CrossRef]
  220. Burris, H.; Mellinghoff, I.; Maher, E.; Wen, P.; Beeram, M.; Touat, M.; Faris, J.; Azad, N.; Cloughesy, T.; Gore, L.; et al. The first reported results of ag.120, a first.In.Class, potent inhibitor of the idh1 mutant protein, in a phase 1 study of patients with advanced idh1.Mutant solid tumors, including gliomas. Mol. Cancer. Ther. 2015, 14, PL04–05. [Google Scholar]
  221. Popovici-Muller, J.; Lemieux, R.M.; Artin, E.; Saunders, J.O.; Salituro, F.G.; Travins, J.; Cianchetta, G.; Cai, Z.; Zhou, D.; Cui, D.; et al. Discovery of ag-120 (ivosidenib): A first-in-class mutant idh1 inhibitor for the treatment of idh1 mutant cancers. ACS Med. Chem. Lett. 2018, 9, 300–305. [Google Scholar] [CrossRef] [PubMed]
  222. Farshidfar, F.; Zheng, S.; Gingras, M.C.; Newton, Y.; Shih, J.; Robertson, A.G.; Hinoue, T.; Hoadley, K.A.; Gibb, E.A.; Roszik, J.; et al. Integrative genomic analysis of cholangiocarcinoma identifies distinct idh-mutant molecular profiles. Cell Rep. 2017, 19, 2878–2880. [Google Scholar] [CrossRef] [PubMed]
  223. Gani, F.; Nagarajan, N.; Kim, Y.; Zhu, Q.; Luan, L.; Bhaijjee, F.; Anders, R.A.; Pawlik, T.M. Program death 1 immune checkpoint and tumor microenvironment: Implications for patients with intrahepatic cholangiocarcinoma. Ann. Surg. Oncol. 2016, 23, 2610–2617. [Google Scholar] [CrossRef] [PubMed]
  224. Fontugne, J.; Augustin, J.; Pujals, A.; Compagnon, P.; Rousseau, B.; Luciani, A.; Tournigand, C.; Cherqui, D.; Azoulay, D.; Pawlotsky, J.M.; et al. Pd-l1 expression in perihilar and intrahepatic cholangiocarcinoma. Oncotarget 2017, 8, 24644–24651. [Google Scholar] [CrossRef] [PubMed]
  225. Bang, Y.J.; Doi, T.; De Braud, F.; Piha-Paul, S.; Hollebecque, A.; Razak, A.A.; Lin, C.C.; Ott, P.A.; He, A.R.; Yuan, S.S.; et al. Safety and efficacy of pembrolizumab (mk.3475) in patients (pts) with advanced biliary tract cancer: Interim results of keynote.028 [abstract]. Eur. J. Cancer 2015, 51 (Suppl. 3), S122. [Google Scholar] [CrossRef]
  226. Fan, B.; Mellinghoff, I.K.; Wen, P.Y.; Lowery, M.A.; Goyal, L.; Tap, W.D.; Pandya, S.S.; Manyak, E.; Jiang, L.; Liu, G.; et al. Clinical pharmacokinetics and pharmacodynamics of ivosidenib, an oral, targeted inhibitor of mutant idh1, in patients with advanced solid tumors. Investig. New Drugs 2019, 1–12. [Google Scholar] [CrossRef] [PubMed]
  227. Valade, E.; Dosne, A.G.; Xie, H.; Kleiman, R.; Li, L.Y.; Perez-Ruixo, J.J.; Ouellet, D. Assessment of the effect of erdafitinib on cardiac safety: Analysis of ecgs and exposure-qtc in patients with advanced or refractory solid tumors. Cancer Chemother. Pharmacol. 2019, 84, 621–633. [Google Scholar] [CrossRef]
  228. Borad, M.J.; Gores, G.J.; Roberts, L.R. Fibroblast growth factor receptor 2 fusions as a target for treating cholangiocarcinoma. Curr. Opin. Gastroenterol. 2015, 31, 264–268. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  229. Roskoski, R., Jr. Ros1 protein-tyrosine kinase inhibitors in the treatment of ros1 fusion protein-driven non-small cell lung cancers. Pharmacol. Res. 2017, 121, 202–212. [Google Scholar] [CrossRef]
  230. Sigal, D.; Tartar, M.; Xavier, M.; Bao, F.; Foley, P.; Luo, D.; Christiansen, J.; Hornby, Z.; Maneval, E.C.; Multani, P. Activity of entrectinib in a patient with the first reported ntrk fusion in neuroendocrine cancer. J. Natl. Compr. Cancer Netw. 2017, 15, 1317–1322. [Google Scholar] [CrossRef]
  231. Huang, C.Y.; Hsieh, F.S.; Wang, C.Y.; Chen, L.J.; Chang, S.S.; Tsai, M.H.; Hung, M.H.; Kuo, C.W.; Shih, C.T.; Chao, T.I.; et al. Palbociclib enhances radiosensitivity of hepatocellular carcinoma and cholangiocarcinoma via inhibiting ataxia telangiectasia-mutated kinase-mediated DNA damage response. Eur. J. Cancer 2018, 102, 10–22. [Google Scholar] [CrossRef] [PubMed]
  232. Chung, H.C.; Ros, W.; Delord, J.P.; Perets, R.; Italiano, A.; Shapira-Frommer, R.; Manzuk, L.; Piha-Paul, S.A.; Xu, L.; Zeigenfuss, S.; et al. Efficacy and safety of pembrolizumab in previously treated advanced cervical cancer: Results from the phase ii keynote-158 study. J. Clin. Oncol. 2019, 37, 1470–1478. [Google Scholar] [CrossRef] [PubMed]
  233. Lin, Y.; Zhang, B.; Liang, H.; Lu, Y.; Ai, X.; Zhang, B.; Chen, X. Jnk inhibitor sp600125 enhances tgf-beta-induced apoptosis of rbe human cholangiocarcinoma cells in a smad-dependent manner. Mol. Med. Rep. 2013, 8, 1623–1629. [Google Scholar] [CrossRef] [PubMed]
  234. Waetzig, V.; Herdegen, T. Context-specific inhibition of jnks: Overcoming the dilemma of protection and damage. Trends Pharmacol. Sci. 2005, 26, 455–461. [Google Scholar] [CrossRef] [PubMed]
  235. Yang, Y.M.; Bost, F.; Charbono, W.; Dean, N.; McKay, R.; Rhim, J.S.; Depatie, C.; Mercola, D. C-jun nh(2)-terminal kinase mediates proliferation and tumor growth of human prostate carcinoma. Clin. Cancer Res. 2003, 9, 391–401. [Google Scholar]
  236. Chromik, A.M.; Muller, A.M.; Korner, J.; Belyaev, O.; Holland-Letz, T.; Schmitz, F.; Herdegen, T.; Uhl, W.; Mittelkotter, U. Genetic deletion of jnk1 and jnk2 aggravates the dss-induced colitis in mice. J. Investig. Surg. 2007, 20, 23–33. [Google Scholar] [CrossRef] [PubMed]
  237. Ke, H.; Harris, R.; Coloff, J.L.; Jin, J.Y.; Leshin, B.; Miliani de Marval, P.; Tao, S.; Rathmell, J.C.; Hall, R.P.; Zhang, J.Y. The c-jun nh2-terminal kinase 2 plays a dominant role in human epidermal neoplasia. Cancer Res. 2010, 70, 3080–3088. [Google Scholar] [CrossRef]
  238. Yoon, C.H.; Kim, M.J.; Kim, R.K.; Lim, E.J.; Choi, K.S.; An, S.; Hwang, S.G.; Kang, S.G.; Suh, Y.; Park, M.J.; et al. C-jun n-terminal kinase has a pivotal role in the maintenance of self-renewal and tumorigenicity in glioma stem-like cells. Oncogene 2012, 31, 4655–4666. [Google Scholar] [CrossRef]
Figure 1. MAPK signaling pathway: ERK1/2, JNK1/2, and p38. Activation of ERK begins with the phosphorylation of MEK1/2, followed by activation of tyrosine and threonine residues. Activated RAF binds to and phosphorylates the kinases MEK1/2 as well as activated RAS. The RAS activation occurs at the plasma membrane and is mediated by son of sevenless (SOS), a guanine nucleotide exchange factor (GEF). Signals from cell surface receptors are passed through RAS-GTP to the RAF(A/B/C) and/or PI3K, the latter then activates the PI3K/AKT signaling pathway. RAF also receives signals from the activated ligands’ G protein-coupled receptors (GPCRs). Activated RAF is capable of phosphorylating MEK, and subsequently, the ERK/MAPK signaling pathway. After activation of ERK, ERK1/2 moves to the cytoplasm and nucleus to phosphorylate other proteins. These proteins are responsible for cell regulation, growth, differentiation, and mitosis. JNK is activated in response to cytokines, growth factors, pathogens, stress, etc., and is associated with the transformation of oncogenes and growth factor pathways. Activation of JNK requires dual phosphorylation tyrosine and threonine, the MAP2Ks that catalyze this reaction are known as MKK4 (also known as SEK1) and MKK7. MKK4/7 are phosphorylated and activated by MAP3Ks, TAK1, and ASK1. Tumor necrosis factor (TNF) receptor signaling and ROS might be the major upstream mediators of JNK activation. Abnormal activation of the JNK signaling pathway is linked to the development of cancer, diabetes, inflammatory diseases, and neurodegenerative diseases. The activation of p38 is mediated by upstream kinases, MAP kinase 3 (MKK3), and MAP kinase 6 (MKK6). MKK3/6 are activated by MAP3Ks such as ASK1 and TAK1, which respond to various extracellular stimuli including osmotic shock, inflammatory cytokines, lipopolysaccharides (LPS), UV light, and growth factors.
Figure 1. MAPK signaling pathway: ERK1/2, JNK1/2, and p38. Activation of ERK begins with the phosphorylation of MEK1/2, followed by activation of tyrosine and threonine residues. Activated RAF binds to and phosphorylates the kinases MEK1/2 as well as activated RAS. The RAS activation occurs at the plasma membrane and is mediated by son of sevenless (SOS), a guanine nucleotide exchange factor (GEF). Signals from cell surface receptors are passed through RAS-GTP to the RAF(A/B/C) and/or PI3K, the latter then activates the PI3K/AKT signaling pathway. RAF also receives signals from the activated ligands’ G protein-coupled receptors (GPCRs). Activated RAF is capable of phosphorylating MEK, and subsequently, the ERK/MAPK signaling pathway. After activation of ERK, ERK1/2 moves to the cytoplasm and nucleus to phosphorylate other proteins. These proteins are responsible for cell regulation, growth, differentiation, and mitosis. JNK is activated in response to cytokines, growth factors, pathogens, stress, etc., and is associated with the transformation of oncogenes and growth factor pathways. Activation of JNK requires dual phosphorylation tyrosine and threonine, the MAP2Ks that catalyze this reaction are known as MKK4 (also known as SEK1) and MKK7. MKK4/7 are phosphorylated and activated by MAP3Ks, TAK1, and ASK1. Tumor necrosis factor (TNF) receptor signaling and ROS might be the major upstream mediators of JNK activation. Abnormal activation of the JNK signaling pathway is linked to the development of cancer, diabetes, inflammatory diseases, and neurodegenerative diseases. The activation of p38 is mediated by upstream kinases, MAP kinase 3 (MKK3), and MAP kinase 6 (MKK6). MKK3/6 are activated by MAP3Ks such as ASK1 and TAK1, which respond to various extracellular stimuli including osmotic shock, inflammatory cytokines, lipopolysaccharides (LPS), UV light, and growth factors.
Cells 08 01172 g001
Figure 2. The hierarchy of cholangiocarcinoma stem cells. Hepatic stem cells grow and differentiate into hepatic progenitor cells and mature hepatocytes (hepatocytes and cholangiocytes). CCA stem cells are produced by genetic abnormalities of hepatic stem cells, hepatic progenitor cells, or CCA cells, which then grow and differentiate to CCA cells. The NOTCH signaling pathway might induce mature hepatocytes to develop into CCA (in mice). Cholangiocytes can also transform into CCA cells through ROS and inflammation-induced genetic and epigenetic variation.
Figure 2. The hierarchy of cholangiocarcinoma stem cells. Hepatic stem cells grow and differentiate into hepatic progenitor cells and mature hepatocytes (hepatocytes and cholangiocytes). CCA stem cells are produced by genetic abnormalities of hepatic stem cells, hepatic progenitor cells, or CCA cells, which then grow and differentiate to CCA cells. The NOTCH signaling pathway might induce mature hepatocytes to develop into CCA (in mice). Cholangiocytes can also transform into CCA cells through ROS and inflammation-induced genetic and epigenetic variation.
Cells 08 01172 g002
Figure 3. Schematic diagram of the EMT and MET progress in the CCA program. Healthy epithelial cells have a tight junction between the apical-basal polarity and interaction with the basement membrane. These cells express markers of epithelial cell lines such as the adhesive connexin E-cadherin. After stimulation, epithelial cells undergo EMT progression (EMT-induced EMT-TFs results in downregulation), lose cell to cell and cell–ECM connections, lose epithelial markers, express interstitial markers (N-CADHERIN, β-CATENIN), obtain front–back polarity, and enhance migration. During tissue repair or inflammation regression, MET is initiated, and the cells reacquire their epithelial phenotype. Cytokines (TGF-β, TNF-α, and IL-6), tyrosine kinase receptors (EGF, HB-EGF), and their receptors (TGFβR, EGFR, TNFR, and ILR) involved in developmental processes play a key role in the induction of the EMT program by activating intracellular signaling pathways (including MAPKs, IL-6/STAT3, NOTCH/JAG1, and WNT), involved in CCA progression and metastasis.
Figure 3. Schematic diagram of the EMT and MET progress in the CCA program. Healthy epithelial cells have a tight junction between the apical-basal polarity and interaction with the basement membrane. These cells express markers of epithelial cell lines such as the adhesive connexin E-cadherin. After stimulation, epithelial cells undergo EMT progression (EMT-induced EMT-TFs results in downregulation), lose cell to cell and cell–ECM connections, lose epithelial markers, express interstitial markers (N-CADHERIN, β-CATENIN), obtain front–back polarity, and enhance migration. During tissue repair or inflammation regression, MET is initiated, and the cells reacquire their epithelial phenotype. Cytokines (TGF-β, TNF-α, and IL-6), tyrosine kinase receptors (EGF, HB-EGF), and their receptors (TGFβR, EGFR, TNFR, and ILR) involved in developmental processes play a key role in the induction of the EMT program by activating intracellular signaling pathways (including MAPKs, IL-6/STAT3, NOTCH/JAG1, and WNT), involved in CCA progression and metastasis.
Cells 08 01172 g003
Table 1. Genetic alterations of related oncogenes in human CCA.
Table 1. Genetic alterations of related oncogenes in human CCA.
OncogenesTumor Suppressor GenesChromatin-Remodeling GenesGain of Function of Oncogenes
MLL3TP53ARID1AKRAS
ROBO2PTENARID1BBRAF
RNF43 BAP1PIK3CA
PEG3 PBRM1
GNAS
NRAS
PTPN3
CDKN2A
SMAD4
IDH1/2
Table 2. Targeted therapies against cholangiocarcinoma.
Table 2. Targeted therapies against cholangiocarcinoma.
DRUGTargetPhaseIdentifier
AG-120 [226]IDH1INCT02073994
Enasidenib [12]IDH2I/IINCT02273739
JNJ-42756493 [227]FGFRINCT01703481
BGJ398+ PD173072 [165,202,205]FGFRINCT01004224
NVP-BGJ398 [201,202]FGFRIINCT02150967
Ponatinib [228]FGFRIINCT02265341
Ceritinib [229]ROS1-ALKIINCT02638909
LDK378 (Ceritinib) [229]ROS2-ALKIINCT02374489
Entrectinib [230]ROS3-ALKIINCT02568267
Binimetinib (MEK162) [207]MEK1/2INCT00959127
Trametinib and Pazopanib [210]MEK/VEGFRIINCT01438554
Cabozantinib [209]c-MET-VEGFR2IINCT01954745
Cisplatin + gemcitabine [193]EGFRIIINCT00262769
Gemcitabine/Oxaliplatin + Erlotinib [196]EGFRIIINCT01149122
Sorafenib [212]VEGFR-PDGFR-BRAFIINCT00238212
Anetumab ravtansine [12]anti-mesothelin antibody-drug conjugateINCT03102320
Palbociclib [231]CDK4/6 InhibitorI/IINCT03065062
Pembrolizumab + GM-CSF [224]PD1IINCT02703714
Pembrolizumab [232]PD1IINCT02628067

Share and Cite

MDPI and ACS Style

Chen, C.; Nelson, L.J.; Ávila, M.A.; Cubero, F.J. Mitogen-Activated Protein Kinases (MAPKs) and Cholangiocarcinoma: The Missing Link. Cells 2019, 8, 1172. https://doi.org/10.3390/cells8101172

AMA Style

Chen C, Nelson LJ, Ávila MA, Cubero FJ. Mitogen-Activated Protein Kinases (MAPKs) and Cholangiocarcinoma: The Missing Link. Cells. 2019; 8(10):1172. https://doi.org/10.3390/cells8101172

Chicago/Turabian Style

Chen, Chaobo, Leonard J. Nelson, Matías A. Ávila, and Francisco Javier Cubero. 2019. "Mitogen-Activated Protein Kinases (MAPKs) and Cholangiocarcinoma: The Missing Link" Cells 8, no. 10: 1172. https://doi.org/10.3390/cells8101172

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop