Next Article in Journal
Effective Antiviral Application of Antisense in Plants by Exploiting Accessible Sites in the Target RNA
Next Article in Special Issue
Establishment and Validation of Novel Prognostic Subtypes in Hepatocellular Carcinoma Based on Bile Acid Metabolism Gene Signatures Using Bulk and Single-Cell RNA-Seq Data
Previous Article in Journal
Epstein–Barr Virus and Human Endogenous Retrovirus in Japanese Patients with Autoimmune Demyelinating Disorders
Previous Article in Special Issue
FARSB Facilitates Hepatocellular Carcinoma Progression by Activating the mTORC1 Signaling Pathway
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

New and Old Key Players in Liver Cancer

by
Ángel M. Cuesta
1,2,†,
Nerea Palao
1,2,†,
Paloma Bragado
1,2,
Alvaro Gutierrez-Uzquiza
1,2,
Blanca Herrera
1,2,3,
Aránzazu Sánchez
1,2,3 and
Almudena Porras
1,2,*
1
Departamento de Bioquímica y Biología Molecular, Facultad de Farmacia, Universidad Complutense de Madrid (UCM), 28040 Madrid, Spain
2
Instituto de Investigación Sanitaria del Hospital Clínico San Carlos (IdISSC), 28040 Madrid, Spain
3
Centro de Investigación Biomédica en Red de Enfermedades Hepáticas y Digestivas (CIBEREHD-ISCIII), 28040 Madrid, Spain
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(24), 17152; https://doi.org/10.3390/ijms242417152
Submission received: 2 October 2023 / Revised: 27 November 2023 / Accepted: 28 November 2023 / Published: 5 December 2023
(This article belongs to the Special Issue Molecular Mechanisms of Liver Cancer)

Abstract

:
Liver cancer represents a major health problem worldwide with growing incidence and high mortality, hepatocellular carcinoma (HCC) being the most frequent. Hepatocytes are likely the cellular origin of most HCCs through the accumulation of genetic alterations, although hepatic progenitor cells (HPCs) might also be candidates in specific cases, as discussed here. HCC usually develops in a context of chronic inflammation, fibrosis, and cirrhosis, although the role of fibrosis is controversial. The interplay between hepatocytes, immune cells and hepatic stellate cells is a key issue. This review summarizes critical aspects of the liver tumor microenvironment paying special attention to platelets as new key players, which exert both pro- and anti-tumor effects, determined by specific contexts and a tight regulation of platelet signaling. Additionally, the relevance of specific signaling pathways, mainly HGF/MET, EGFR and TGF-β is discussed. HGF and TGF-β are produced by different liver cells and platelets and regulate not only tumor cell fate but also HPCs, inflammation and fibrosis, these being key players in these processes. The role of C3G/RAPGEF1, required for the proper function of HGF/MET signaling in HCC and HPCs, is highlighted, due to its ability to promote HCC growth and, regulate HPC fate and platelet-mediated actions on liver cancer.

1. Introduction

Liver cancer (including hepatocarcinoma (HCC), cholangiocarcinoma (CCA), and combined hepatocellular-cholangiocarcinoma (cHCC-CCA)) represents a global health challenge, and its incidence is growing worldwide. HCC is the most common form of primary liver cancer accounting for around 90% of cases [1]. According to the World Health Organization (WHO), in 2020 liver cancer was the sixth most common cancer worldwide in incidence, accounting for 4.7% of cases, and the third most common in the number of cancer-related deaths, accounting for 8.3% of cases. However, there is a striking variation in HCC incidence rates across geographic regions. The highest incidence and mortality of HCC are observed in East Asia and Africa, but they are also increasing in Europe and the USA. Furthermore, HCC has a strong male predominance likely related to a clustering of risk factors and differences in sex hormones [2,3,4].
The etiology of HCC and the implicated signaling mechanisms need to be fully characterized. Liver cancer, specifically HCC, is usually associated with chronic inflammation, fibrosis and cirrhosis [1]. However, a new HCC subclass, not associated with inflammation, and enriched in CTNNB1 gene mutations and PTK2/FAK1 overexpression has been described [5]. Fibrosis is a common manifestation of chronic liver disease (CLD) that can progress to cirrhosis, eventually leading to HCC development. Several risk factors, frequently associated with chronic liver inflammation, are involved in HCC development, including hepatitis viruses, carcinogens, alcohol intake, hereditary diseases, metabolic syndrome, and fatty liver disease [4,6,7]. Hepatitis B virus (HBV) infection is the most prominent risk factor for HCC, while the risk attributed to hepatitis C virus (HCV) infection has decreased thanks to antiviral treatments and the extended use of vaccines. In contrast, non-alcoholic fatty liver disease (NAFLD) and non-alcoholic steatohepatitis (NASH), a more severe form of NAFLD, are becoming the fastest growing causes of HCC, which are not associated with cirrhosis in most cases [8]. NAFLD and NASH are strongly connected with metabolic syndrome (a condition associated with type 2 diabetes, dyslipidemia, and hypertension) [9]. Therefore, in 2020, a group of international experts in the field proposed a change of the term from NAFLD to MAFLD (metabolic associated fatty liver disease) [10]. Moreover, in recent years, it has been considered that HCCs generated by MAFLD present specific characteristics [8], being more differentiated and having a higher inflammatory component.
HCC is the result of a complex long-term multistep process regulated by the interplay of several risk factors, and the severity of the underlying CLD. HCC results from the accumulation of somatic genetic and epigenomic alterations in the cells of origin. There is an average of 40–60 somatic mutations in coding regions, most of them in ‘passenger’ genes, but a few in cancer ‘driver’ genes, involved in activating key signaling pathways for hepatocarcinogenesis such as those implicated in telomere maintenance, cell cycle control, chromatin modification, and oxidative stress, with WNT–β-catenin, transforming growth factor-β (TGF-β) signaling and receptor tyrosine kinase (RTK) signaling cascades being relevant, among others [11,12,13,14].
Early-stage HCC lesions are small and can normally be cured by minimally invasive methods. Nevertheless, around 50% of the cases are diagnosed after developing symptomatic advanced-stage HCC. This significantly minimizes the possibility of treatment, and death is ensured after a few months [1,15]. Different options for first-, second- and subsequent line treatments are currently available, such as combinations of immunotherapy with tyrosine kinase inhibitors (TKIs) or antiangiogenic agents (e.g., vascular endothelial growth factor (VEGF) inhibitors) [16,17].

2. Liver Cancer Microenvironment

The cross-talk between hepatocytes and other cell populations, constituents of the liver microenvironment, such as hepatic stellate cells (HSCs), liver sinusoidal endothelial cells (LSECs), liver-resident macrophages Kupffer cells, KCs), other immune cells, and hepatic progenitor cells (HPCs), is involved in hepatocarcinogenesis in the context of CLD, and in metastasis generation in later stages [1,18]. In addition to this, the recruitment of immune cells and platelets to the liver during CLD also contributes to the hepatocarcinogenic process. The release of several growth factors, cytokines, chemokines, and extracellular matrix (ECM) proteins by liver cells and/or newly recruited cells, plays a major role in this process. Hence, this makes the pre-cancerous microenvironment of CLD to evolve towards a tumor microenvironment (TME) with immunosuppressive properties and altered angiogenesis, favoring the tumorigenic process. However, the TME is highly heterogeneous, which hinders its study within the tumor tissue in patients. Hence, recent works have also paid attention to the peritumor microenvironment (PME), a non-tumor tissue also involved in the occurrence (PME-O) and progression (PME-P) of HCC. Different proteome profiles have been identified in PME-O and PME-P, which are immune proteins’ key players in PME-O, while proteins involved in inflammation, angiogenesis and metabolism are relevant in PME-P [19].
It is also noteworthy to mention that despite having common characteristics, TME features are likely dependent on liver cancer etiology. The role and relevance of some major components of the liver microenvironment in HCC development and progression are further discussed below.

2.1. Fibrosis: Hepatic Stellate Cells and Cancer Associated Fibroblasts

As mentioned before, the majority of HCCs arise in a fibrotic or cirrhotic liver [20]. Fibrosis results from a chronic liver injury. The activation of HSCs and their further conversion into myofibroblasts (MFBs) are considered key steps in fibrosis development. However, MFBs can have different origins [21]. HSCs are the major source, but other cells such as portal fibroblasts, hepatocytes undergoing epithelial-mesenchymal-transition (EMT), fibrocytes (bone marrow-derived cells), or HPCs can also lead to MFBs [22].
In a normal liver, HSCs are non-proliferative cells that express glial fibrillary acidic protein (GFAP), desmin, vimentin, and platelet-derived growth factor receptor (PDGFR), among other markers [21]. However, upon chronic liver injury they are activated by growth factors mainly released by KCs, such as TGF-β1, considered their major activator, and cytokines such as tumor necrosis factor-α (TNF-α) and interleukin-1β (IL-1 β) [23,24]. After its activation and transdifferentiation into MFBs [25], they express α-smooth muscle actin (α-SMA) and ECM proteins, mainly collagen I and II [22]. In addition, activated HSCs also secrete TGF-β1 and chemokines such as chemokine (C-C motif) ligand 2 and ligand 5 (CCL2, CCL5) and chemokine (C-X-C motif) ligand 8-10 (CXCL8-10) [26] that regulate immune cells, fibrosis and hepatocarcinogenesis (Figure 1). Hence, during fibrosis development HSCs undergo dramatic transcriptomic changes with the sequential activation of inflammatory, migratory, and ECM–producing programs, as determined by single cell RNA sequencing (sc-RNAseq) [27]. Continuous activation of HSCs leads to a non-resolutive repair response evolving to fibrosis and cirrhosis.
Although fibrosis, cirrhosis and HCC are strongly associated, the casual relationship between fibrosis and HCC development is still unclear [28]. The use of different mouse models has led to controversial results. Some studies have shown an association between reduced fibrosis and less HCC risk [29,30]. Nevertheless, other studies did not find a correlation between fibrosis and HCC [31]. These discrepancies might be due to the existence of different subtypes of HSCs with specific functions. The analysis of human and mouse data revealed the existence of two main subtypes of HSCs in the fibrotic liver with opposed functions in hepatocarcinogenesis [32]. On one side, there are quiescent and cytokine-producing HSCs enriched in HGF (hepatocyte growth factor) that protect against hepatocyte cell death and HCC development and on the other side, activated myofibroblastic HSCs enriched in collagen I, which promote HCC development and increase ECM stiffness. An imbalance between these HSC types in favor of myofibroblastic HSCs during CLD favors HCC promotion. In relation to this, matrix stiffness could also play a key role. This is supported by recent studies indicating that a higher matrix stiffness promotes a cancer stem-like phenotype [33] and HCC progression [34]. More recent data derived from sc-RNAseq analyses support a more complex situation due to the existence of several subpopulations of HSCs in CLD [35]. Hence, activated HSCs are considered to be highly heterogenous and have been classified into distinct subpopulations: proliferative HSCs (pHSCs), inflammatory HSCs (iHSCs), intermediate activated vascular HSCs (vHSCs), contractile and migratory HSCs (cmHSCs), and fibrogenic myofibroblasts (myHSCs). However, the specific functions of each subpopulation in liver cancer as well as the regulatory signals remain unclear.
In the classic model, once liver cancer is generated, HSC activation in the peritumoral tissue contributes to HCC progression [20,36] through the secretion of growth factors, cytokines and chemokines such as VEGF, HGF, TGF-β1, PDGF, IL-6 and CX3CL1. These signals can act either directly on cancer cells inducing survival, proliferation and/or migration, or indirectly by enhancing angiogenesis and modifying the immune response to favor an immunosuppressive context. For example, KCs are stimulated by HSC-derived cytokines leading to the exhaustion of activated T cells [37,38], and the expansion of regulatory T cells (Tregs) or M2 macrophage polarization. An anti-inflammatory phenotype can also be induced by HSC-derived CX3CL1 through regulation of KCs and infiltrated monocytes [39,40]. Recently, through a sc-RNAseq analysis of human liver samples, new subpopulations of pro-fibrotic macrophages, endothelial cells, and PDGFRα+ collagen-producing mesenchymal cells which collaborate to promote fibrosis have been identified [41].
During HCC progression, HSCs and HSC-derived MFBs are believed to transform into cancer-associated fibroblasts (CAFs) [28], although CAFs could also be originated from cancer cells, endothelial cells, portal cells or mesenchymal bone marrow stem cells [42]. CAFs play a key role inhibiting immune cell response, promoting angiogenesis and HCC progression by secreting growth factors, cytokines and chemokines such as HGF, IL-6, VEGF, CCL2, or CCL7 [18,43,44,45]. However, the existence of different types of CAFs with distinct gene expression profiles and functions in hepatocarcinogenesis has recently been uncovered [35]. Among them, myofibroblastic CAFs, which produce collagen I, suppress tumor growth by mechanically restraining tumor spread, while inflammatory CAFs which secrete HGF, promote tumor growth [46]. Nevertheless, CAFs are heterogeneous and suffer dynamic changes during hepatocarcinogenesis that are likely dependent on the specific context.
Based on the above information, we can state that the classic association between fibrosis and liver cancer still stands. In this respect, there is a great interest in blocking transdifferentiation of HSCs into MFBs or, at least, to reduce it, as a potential therapy for cancer. However, recent published data support the idea of a more complex activity of HSCs and CAFs as they have different phenotypes and functions. This likely determines matrix stiffness and the secretion of specific growth factors, cytokines and chemokines, which leads to the activation or inhibition of tumor progression. In addition, HSCs and CAFs play an important role in the immune response as they can exert an immunosuppressive effect. Therefore, more efforts are required to understand all the complexity of fibrosis and its relationship with liver cancer.

2.2. Liver Endothelial Sinusoidal Cells

Liver sinusoidal endothelial cells (LSECs), mainly localized in the liver sinusoidal wall, have open fenestrations and lack a basement membrane to facilitate substance exchange from the sinusoidal space to the parenchyma and the space of Disse. Upon the alteration of KCs function, LSECs uptake foreign substances from the blood stream [42].
In response to liver injury, LSECs also contribute to induce HSCs proliferation by releasing VEGF [24]. In addition, they play a relevant role regulating the immune system and tumorigenesis [42], contributing to leukocyte recruitment into the injured liver [47].

2.3. Liver Immune Cells

The liver is an organ rich in immunocompetent cells [42,48,49]. Different immune cell populations are present in the liver under normal conditions, including KCs, natural killer (NK) cells, lymphocytes (from innate (NKT cells) and adaptive systems), and dendritic cells (DCs) [42], which exert a net immunosuppressive function [50]. This prevents excessive inflammation in response to bacterial components and dietary antigens coming from the gastrointestinal tract through the portal vein. In addition, liver immune cells and inflammatory signals, such as IL-1β, IL-6 and TNF-α are necessary for liver regeneration in response to liver damage (Figure 1).
In CLD, the immune system is dysregulated by an excessive production of pro-inflammatory cytokines (IL-1β, IL-6, TNF-α), induced by the release of damage-associated molecular patterns (DAMPs) by dying hepatocytes [50]. Initially, KCs kill damaged hepatocytes and tumor cells via phagocytosis and secrete the aforementioned pro-inflammatory cytokines and chemokines (e.g., CCL2) to recruit and activate other immune cells. All this leads to the recruitment of monocytes (that differentiate into macrophages), neutrophils and other immune cells (DCs and CD8+ T cells), which amplify the response [51]. The interaction between platelets and KCs also induces the secretion of KC-derived cytokines and chemokines that promote NK and CD8+ T cell recruitment, at least, in NASH [52,53]. KCs also release TGF-β1 and PDGF, activating HSCs. NKs would kill activated HSCs in the early stages, preventing fibrosis progression [54,55]. However, differences in the response of KCs and other immune cells can take place during CLD and its progression to HCC depending on the etiology of the disease. For example, during NAFLD/MAFLD progression to HCC, resident KCs are quickly depleted and replaced by monocyte-derived KCs and hepatic lipid-associated macrophages [56], which can also derive from adipose tissue multipotent adult stem cells in obesity. Hence, serum levels of factors, like stem cell growth factor-beta (SCGF-β), that regulate macrophage differentiation might be useful in predicting liver disease severity and insulin resistance in obese patients, although further studies are required [57].
DCs play an important role both during CLD progression and once HCC is generated. They infiltrate the tumor, presenting antigens from tumor cells to T cells to prime and co-stimulate CD8+ T cells [43]. CD4+ T cells are also involved in the immune response in CLD-HCC with different populations of these cells exerting specific, or even opposed actions [52].
CD8+ cytotoxic T cells (CTLs) are the main killers of cancer cells upon recognition of antibodies on DCs [43]. CD4+ T cells can also activate CD8+ T cells by CD40L and enhance their proliferation and maturation into memory CD8+ T cells in the initial tumor stages. Hence, increased infiltration of CD8+ T cells in HCC patients is associated with a better prognosis [50]. However, a high expression of regulatory factors such as VEGF, CXCL17 or IL-10 can restrict CTLs response. On the other hand, different subpopulations of CD8+ T cells can play specific roles depending on the context. Thus, using different NASH mouse models and experimental approaches, a pro-inflammatory and pro-tumorigenic role has been demonstrated for CD8+ T cells, facilitated by platelet-mediated recruitment [53,58]. Particularly, CD8+CXCR6+PD1+ T cells mediate the auto-aggressive killing of liver cells and are associated with chronic liver damage and HCC development in NASH [59,60]. Therefore, further understanding of their role in HCC is required, as it likely depends on the specific context and/or stage.
NK cells are essential to kill tumor cells through perforin and granulin release. In addition, they produce pro-inflammatory cytokines and chemokines that contribute to their anti-tumor function [43]. However, liver specific NK cells are maintained in a hyporesponsive state, even after their activation by pro-inflammatory signals [61]. Moreover, the NK cell population decreases in blood and tumor tissues from HCC patients [50].
In response to liver damage, and in HCC, bone marrow derived monocytes are also recruited to the liver, where they differentiate into macrophages [62,63,64]. CCL2 secreted by hepatocytes, KCs, HSCs, and platelets [51] and acting through CCR2, is essential for this recruitment [65]. These recruited macrophages together with KCs constitute the tumor-associated macrophages (TAM) that accumulate within the tumor [42].
TAMs can undergo M1 (classic) or M2 (alternative) polarization, although an overlap in M1/M2 signatures also occurs. M1 macrophages are pro-inflammatory, hence they secrete TNF-α, IL-6 or IL-1β and produce reactive oxygen species (ROS) and nitric oxygen (NO) to eliminate tumor cells [66]. In contrast, M2 macrophages secrete anti-inflammatory cytokines (e.g., IL-10, TGF-β1), chemokines (e.g., CCL17, CCL22), and growth factors (e.g., VEGF, fibroblast growth factor (FGF)), which promote cancer progression, EMT, immune evasion and angiogenesis [42,67]. M2 TAMs replace M1 TAMs in the advanced HCC stages, Wnt ligands produced by hepatic tumor cells being important for M2 polarization [68].
Tregs are a subtype of CD4+ T cells that express CD25 and FOXP3 (forkhead box P3) [69]. Their main function is to maintain autoimmune tolerance and homeostasis by reducing the immune response. Consequently, once Tregs are recruited by CCL20/CCR6 [70] and activated by IL-10 and TGF-β1, they play an anti-inflammatory and immunosuppressive role in HCC through impairing CD8+ T cell actions [71]. Accordingly, high levels of Tregs are present in the TME of HCC patients with a poor prognosis [72,73]. However, it remains unclear how Tregs operate in certain scenarios, for instance, how they regulate NASH development and progression to HCC [52].
Myeloid-derived suppressor cells (MDSCs) are a heterogeneous population of bone marrow-derived immature myeloid cells, classified into granulocytes (also called polymorphonuclear (PMN-MDSCs)) and mononuclear (M-MDSCs) [74]. Their major role is to suppress anti-tumor immunity through the inhibition of DCs, CTLs, NK and B cells and to activate Tregs and TAMs, as well as to promote angiogenesis [43]. Their recruitment to the liver during HCC development and progression is mainly promoted by cytokines (e.g., granulocyte-macrophage colony-stimulating factor (GM-CSF), granulocyte colony-stimulating factor (G-CSF), and IL-6) and chemokines, such as CCL2, according to data derived from different murine models of HCC [75]. However, other signals like CCL26 could also contribute to it [42].
In HCC patients, MDSCs are increased in the blood and within the tumor, inducing programmed death-ligand 1 (PD-L1) expression in advanced stages through interaction with KCs [76] to suppress the adaptive immune response.
Tumor associated neutrophils (TANs) have a similar origin to that of PMN-MDSCs and share some phenotypic traits [77]. However, they are more heterogeneous and present different functional properties. Interferon β (IFN-β) induces an antitumor N1 polarization of TANs, which can kill tumor cells and activate CTLs. In contrast, the N2 phenotype is induced by TGF-β, cytokines and chemokines, promoting tumor development and immunosuppression through the inhibition of T-cell response, also enhancing angiogenesis [78,79]. N2 TANs produce CCL2 and CCL17, which promote macrophage and Tregs’ recruitment [80].
The recruitment of neutrophils into HCC tumor tissue, after G-CSF and IL-17-induced mobilization, is mediated by chemokines such as CXCL1–3 and CXCL5–8, which bind to the CXCR2 receptor in neutrophils [79,81]. Hence, CXCL5 secreted by tumor cells increases TANs infiltration, leading to a positive feedback loop [82].
TANs induce tumor cell migration and invasion and promote metastasis generation [79]. Neutrophils also protect tumor cells from being killed by NK cells [83]. Moreover, the formation of platelet-neutrophil clusters induced by platelet-derived CXCL5 and CXCL7 leads to the generation of premetastatic niches [84,85].
On the other hand, neutrophil extracellular traps (NETs) derived from TANs have emerged as important regulators of HCC [86]. NETs are released into the tumor tissue and circulation, and are able to trap tumor cells, which favors its spreading through the blood to generate metastasis [87].
More recently, B cells have been considered to play a role in HCC [1]. Studies derived from mouse models found both tumor-promoting and suppressing activities for B cells in HCC [88]. The release of selective cytokines by different populations of B cells could account for distinct effects in specific contexts [43].

2.4. Platelets

Platelets are also a component of the liver microenvironment and are key players in fibrosis, CLD and HCC. Under physiological conditions, platelets promote hemostasis and vascular integrity, also regulating the immune system, which allows for the protection of the liver from pathogens [89]. In CLD and HCC their function seems to be dual (Figure 1). Hence, both protective and deleterious effects of platelets have been described, playing specific roles in different pathologies and pathological stages [89].
Different clinical studies have shown that the presence of thrombocytosis is associated with worse prognosis in patients with HCC [90], unlike thrombocytopenia, which promotes patients’ survival [91]. Thus, a high platelet count directly correlates with greater tumor aggressiveness and poorer survival of HCC patients [91]. In addition, clinical trials have shown that antiplatelet therapy using clopidogrel or aspirin reduces liver fibrosis risk [92] and HCC, and decreases liver-related mortality in patients with chronic hepatitis B and C. This suggests that platelets may contribute to HCC progression [93,94,95]. Moreover, patients with advanced HCV-induced fibrosis or NASH had increased serum and intrahepatic levels of platelet-derived CXCL4, which promotes HSC proliferation, chemotaxis, and chemokine expression [96]. Therefore, platelet count has been incorporated as a prognostic marker, and a criteria for the selection of treatment in patients at risk of developing HCC and in patients with viral liver cirrhosis [97,98,99].
In HCC, platelets interact with different liver cells, including HSCs, LSECs and immune cells, in addition to having a potential direct effect on tumor cells [100,101]. In mouse models of acute viral hepatitis, platelet depletion decreases intrahepatic accumulation of virus-specific CTLs and delays organ damage, suggesting that platelets can enhance adaptive immunity [102]. In agreement with this, in an HBV transgenic mouse model, antiplatelet therapy inhibits liver fibrosis and HCC development by reducing the number of intrahepatic HBV-specific CD8+ T and virus-non-specific inflammatory cells [103].
In other types of liver injury, platelets also promote damage, fibrosis and HCC. For instance, platelet number, activation and aggregation are increased in NASH [53]. In the early and advanced stages of NAFLD, platelet recruitment to the liver is mediated by KCs through hyaluronan-CD44 binding, and platelet activation by Glycoprotein GPIbα, contributing to NASH and HCC development [53]. Antiplatelet therapy with aspirin or clopidogrel blocks intrahepatic immune cell infiltration, reduces inflammatory cytokines and hepatocyte damage, preventing NASH-induced HCC [53]. Furthermore, in MDR2(Abcb4)-null mice, and control mice fed with a 3,5-diethoxycarbonyl-1,4-dihydrocollidine (DDC)-supplemented diet that leads to cholestasis, platelets activate HSCs through PDGF-β secretion and promote fibrosis [104]. In contrast, our unpublished results indicate that livers from transgenic mice overexpressing C3G in platelets are partially protected from carbon tetrachloride (CCl4)-induced fibrosis, while the deletion of platelet C3G increases liver fibrosis. This suggests that platelet C3G, which is involved in platelet activation [105,106], has a protective effect on fibrosis. However, diethylnitrosamine (DEN)+CCl4-induced liver cancer is exacerbated in wild type (wt) mice as compared to mice lacking platelet C3G, which suggests that platelet C3G promotes HCC progression. Therefore, these studies support the idea that platelets regulate liver immune response and can either promote or reduce fibrosis. However, in all these contexts platelet activation contributes to HCC development and/or progression in different scenarios of liver chronic damage.
Nevertheless, other studies have shown that platelets can also have a protective role in liver regeneration, CLDs and fibrosis, also preventing HCC. For instance, platelet-derived serotonin induces hepatocyte proliferation in a model of liver regeneration, and the release of growth factors and cytokines (e.g., IL-6), at the site of liver injury [107]. Thrombopoietin mediated platelet generation after liver hepatectomy also reduces liver fibrosis and promotes liver regeneration [108]. In a model of CCl4-induced fibrosis, platelets also reduce liver fibrosis progression through MMP-9 induction and TGF-β downregulation [109]. Human platelets also attenuate liver fibrosis in severe combined immunodeficiency (SCID) mice by secreting HGF that inhibits HSC activation, upregulates MMPs, and restrains hepatocyte apoptosis [110]. Furthermore, in NAFLD mouse models, platelets restrain HCC growth by enhancing CD8+ T cell-dependent anti-tumor immunity through P2Y12/leukotrienes-dependent CD40L release [111]. Moreover, in a mouse model of cholestasis-induced liver fibrosis, platelets are activated by HSCs and play an antifibrotic role by reducing collagen I deposition by HSCs through enhancing the HGF-MET signaling pathway [112].
Altogether, all these data support that platelets are relevant regulators of CLD and liver cancer through modulation of immune cells, HSCs and other liver cells. However, their effects are likely dependent on the context.

2.5. Hepatic Stem/Progenitor Cells in Liver Cancer

The adult hepatic stem/progenitor cell (HPC) population constitutes another cellular component of the liver microenvironment during chronic liver injury. Evidence in rodents and humans locates the liver stem cell niche at the Canals of Hering—the adult remnants of fetal/neonatal liver ductal plates—in the periportal area. From there, a transit amplifying cell population of bipotent cells, capable of differentiating into both hepatocytes and cholangiocytes arises and expands in a context of CLD to repair the liver [113,114,115]. This makes HPCs potential candidates as cells of origin in liver cancer, since chronic liver injury constitutes a predominant risk factor for all types of liver tumors (HCC, CCA, and cHCC-CCA) that precedes the big majority of them [116,117,118]. Nonetheless, the potential of HPCs to generate liver tumors has been a subject of intense debate. There is compelling evidence for HPCs’ tumorigenic potential (thoroughly discussed in other recent reviews) [119,120,121]. The strongest evidence is the fact that HPCs/oval cells are susceptible to neoplastic transformation, giving rise to HCC upon genetic alterations implying activation of oncogenes (such as Ras) or silencing of tumor suppressors (including ARF/INK4a and p53) [122,123,124,125]. In the same line, HCC tumors spontaneously developed in mice with a deletion in embryonic liver fodrin (ELF), an adaptor protein required for TGF-β signaling. These tumors were proposed to derive from transformed stem cells with the inactivation of TGF-β signaling and subsequent activation of IL-6 signaling [126]. In agreement with this, chronic exposure of HPCs to TGF-β can confer tumor initiating cell properties and promote hepatocarcinogenesis through a miR216a/Phosphatase and Tensin Homolog (PTEN)/Akt-dependent pathway, under specific conditions [127], but not all, based on our own data [128]. Furthermore, conditional deletion of mammalian orthologs of Hippo kinase, the serine-threonine kinases Mst1 and Mst2, leads to excessive proliferation and appearance of liver tumors at 5–6 months of age, preceded by accumulation of small proliferating periductal cells expressing HPC markers. This, together with the observation of tumors of both biliary and hepatocyte lineages, made authors link HPC activation by suppression of Hippo pathway with tumor formation [129]. These findings are in accordance with gene expression profiling data that identified a subset of HCCs consistent with a an HPC origin, which is among those with the worst prognosis [130,131]. Altogether, evidence supports that, at least, a fraction of HCC could derive from transformed HPCs as a consequence of the accumulation of genetic alterations and a cell maturation arrest process. Nevertheless, the key issue is which cell is more prone to neoplastic transformation, mature hepatocyte or HPC, and which of them more likely becomes a tumor-initiating cell. The use of state-of-the-art strategies is helping on this. One recently reported experimental model in this field is a double knockout mouse combining the loss of autophagy (ATG5 or 7) and PTEN genes that results in inflammation and fibrosis, concomitant with the development of an extensive ductular reaction, which leads to HCC formation at 4–5 months of age. Lineage tracing-based studies revealed that HCCs generated from ductular liver progenitor cells are derived from dedifferentiated hepatocytes. These findings prompt us to be more careful when interpreting the cellular origin of HCC based on the stem/progenitor-like properties of tumor cells, and support that hepatocytes are the cells of origin of HCC, at least, upon injury induced by autophagy deficiency [132]. Earlier lineage tracing studies in Rosa-YFP (yellow fluorescent protein) mouse reached similar conclusions. Using adeno-associated virus (AAV)-Thyroxine binding globulin (TBG)-Cre or Foxl1-Cre mice to label hepatocytes or stem/progenitor cells, respectively, and two models of chemical hepatocarcinogenesis (induced by DEN/CCl4 or DEN/TCPOBOP (1,4-Bis[2-(3,5-Dichloropyridyloxyl)] benzene), the authors also demonstrated that tumor formation was driven by transformed hepatocytes [133]. These studies highlight that oncogenic reprogramming and acquisition of stemness in hepatocytes can be important steps in HCC development. On the other hand, Tummala et al. [134] using a mouse model expressing hepatocyte-specific human unconventional prefoldin RPB5 interactor (hURI-tetOFFhep) that causes DNA damage and results in liver tumor development, demonstrated that both HPCs and adult hepatocytes can contribute to liver tumorigenesis. However, about 70% of HCCs originate from hepatocytes, regardless of HPC expansion during early stages of hepatocarcinogenesis. Of note, studies with Mdr2-KOFoxl1-Cre; RosaYFP mice that develop both HCC and cHCC-CCA tumors showed that HPCs are the source of cHCC-CCA tumors, but not of HCCs [133,135]. Hence, no universal conclusions should be made from specific experimental models. Clearly, different cancer models lead to different tumorigenic processes, making it clear that the cell of origin of HCC is context specific. Nonetheless, the idea of HPCs as a major source of liver tumors, specifically HCC, should be dismissed, since there is sufficient evidence supporting a relatively low risk of their malignant transformation. This is an interesting issue to keep in mind when evaluating therapeutic strategies to push regeneration in CLD.

3. Signaling Pathways in HCC

Among all signaling pathways regulating liver cancer development and progression, we have focused on those elicited by the receptor tyrosine kinases (RTKs) MET and Epidermal growth factor receptor (EGFR), and TGF-β.

3.1. Tyrosine Kinase Receptors: MET and EGFR

HGF acting through MET represents a key signaling pathway that controls epithelial morphogenesis through its mitogenic, motogenic and pro-survival activities, being essential for liver development [136,137,138]. HGF/MET signaling is also necessary for reparative responses, activated upon liver injury, both during hepatocyte-mediated regeneration and liver regeneration associated with HPC expansion [139,140,141,142] and exerts hepatoprotective effects against liver fibrosis through different mechanisms [143].
A number of in vitro and in vivo analyses have been performed to clarify the role of HGF/MET in HCC. In vitro data support the notion that HGF/MET has a pro-invasive effect and relate it to its ability to induce EMT [144,145]. Thus, HCC cell lines with high MET levels display a mesenchymal phenotype [146,147]. Different laboratories have analyzed the effects of inactivating or overexpressing MET and/or HGF in in vivo experimental models of HCC with discordant results. As an example of the complexity of MET signaling effects in HCC, the loss of MET in hepatocytes led to bigger tumors with shorter latency compared to controls in the DEN-induced HCC model. On the contrary, transgenic mouse models of MET overexpression in hepatocytes spontaneously generate liver tumors [148,149,150]. On the other hand, MET knockout accelerates and enhances chemically-mediated HCC initiation, but does not affect phenobarbital-induced HCC promotion [151]. Therefore, the scenario is intricate and suggests that context might be important, and a fine balance of HGF/MET signaling is necessary to maintain liver homeostasis.
It is well established that the HGF/MET signaling pathway is overactivated in around 50% of HCC patients and nearly all liver metastases. Furthermore, comparative functional genomics identified a MET-regulated signature in a subgroup of HCC with aggressive phenotype and poor prognosis [152]. Aberrant activation of MET signaling can be achieved by alterations at different levels, including gene amplification, activating mutations, downregulation of microRNAs targeting MET, and autocrine signaling due to HGF overexpression or activation by other ligands [153]. However, neither MET expression nor HGF plasma levels can be used as diagnostic or prognostic factors in HCC, as studies led to controversial results. Therefore, it deserves further analysis. One possibility is their potential utility in specific patient subgroups, particularly those with the MET specific signature, or pathological status. Another possibility would be to use them together with other HCC risk parameters [154,155].
Importantly, the HGF/MET axis has emerged as a therapeutic target in HCC. Different types of HGF inhibitors have been designed, including HGF neutralizing antibodies, HGF antagonists, MET-selective TKIs or multitarget TKIs [156]. Cabozantinib belongs to this last group, and it has recently been included as a second line targeted therapy. Cabozantinib targets MET and receptors associated with angiogenesis (such as VEGFR or AXL). How the effect on these different targets contribute to reaching the final antitumoral effect is not known, nor the specific contribution of MET inhibition to the overall effect. Data obtained with selective TKIs shed light on this question. Positive results obtained in clinical trials with Tepotinib and Capmatinib in MET positive-advanced HCC [157] support a key contribution of MET inhibition to the final effects of TKIs. However, HGF/MET targeting in HCC presents some key problems still unresolved. First, understanding why other TKI drugs, such as (ARQ 197), a TKI selective for MET [158] or the multitarget TKI, golvatinib, in combination with sorafenib [159], failed to be effective for advanced HCC treatment. Second, due to its important regenerative functions, the inhibition of the HGF/MET axis may be detrimental for the liver disease that frequently accompanies HCC development [156]. Targeting HGF/MET downstream effectors could solve some of these problems that current therapy, focused on the receptor level, presents. Thus, the identification of specific and druggable targets in the HGF/MET signaling pathway, rather than focusing on proteins shared with other pathways, is desirable. Third, sorafenib, an oral multikinase inhibitor, is one of the options for standard first-line systemic therapy for HCC [160]. However, patients acquire resistance within 6 months, which is one of the major challenges of HCC management [161]. As aberrant activation of HGF/MET is one of the mechanisms involved in sorafenib resistance in HCC [162], inhibition of this pathway could aid to obtain better results with the current therapy. However, it is also necessary to consider (as described in other sections) that HGF is produced by different components of the liver TME. In addition, HGF/MET signaling plays specific roles in different liver cells that change during liver tumor generation and progression. Therefore, this needs to be taken into consideration.
EGFR is another member of the RTK family, activated by several ligands including TGF-α, AREG and EGF [163,164]. Together with its well-established role in liver regeneration [165], EGFR signaling is also involved in different stages of CLD. Hence, numerous studies point to a pro-fibrogenic role of EGFR in the liver [166,167,168,169]. Various EGFR ligands are overexpressed in response to chronic liver injury as shown in experimental models and human cirrhotic tissue. This may favor the hepatocarcinogenic process, as observed in different models of hepatic injury ending in HCC, in which pharmacological or genetic inhibition of EGFR prevents tumor development [170,171,172]. It is interesting to point out that EGFR is expressed in liver macrophages both in human HCC and mouse HCC models, being critical for HCC development as demonstrated by specific deletion in KCs/macrophages [173]. The activation of EGFR has also been detected in liver macrophages in experimental models of chronic injury [168,174]. These findings, together with a delay in the DEN-induced inflammatory response and tumor formation in mice expressing a dominant negative mutant EGFR lacking catalytic activity in hepatocytes [172], evidence an interesting regulatory role for EGFR in the HCC inflammatory microenvironment, which involves both hepatocytes and immune cells, underneath its pro-tumorigenic effect. Additionally, EGFR-dependent activation of HSCs has also been reported [175], which provides a potential additional mechanism by which EGFR would contribute to fibrogenesis, the activation of inflammation, and therefore, liver tumor formation.
Although aberrant EGFR activation through EGFR gene amplification and/or mutation has been detected in various types of cancer, this is not frequently found in HCC. In fact, EGFR overexpression is often observed in HCC, but it is not usually associated with an increase in gene copy number [176,177,178] or somatic mutations in the exon encoding the catalytic domain [179]. Interestingly, regulators of EGFR signaling have been described as altered, e.g., ERBB receptor feedback inhibitor 1 (ERRFI1), a negative regulator of EGFR that inhibits its catalytic activity and mediates its lysosomal degradation, is frequently deleted in HCC (Cancer Genome Atlas Research Network, 2017) [180]. EGFR ligands are also overexpressed in HCC [178,181,182]. Importantly, EGF is part of several gene signatures of different nature that are associated with HCC (angiogenic-, immune- related, and others) [154,183]. This suggests that EGFR signaling contributes to HCC development at different mechanistic levels.
Other evidence supports a key role for EGFR signaling in different types of liver cancer. Thus, EGFR expression correlates with a high proliferative activity, the presence of intrahepatic metastasis, poor differentiation, and bad prognosis in HCC [177,184]. EGFR activation is also associated with poor prognosis in carcinomas of the biliary tract [185].
All this, together with encouraging in vitro data using different types of EGFR signaling inhibitors (TKIs and EGFR targeted antibodies), paved the way for clinical trials aiming to impact EGFR signaling in HCC. However, to date, EGFR inhibition therapy cannot be used for HCC treatment [182], as clinical trials were not successful (Table 1), even the combination of inhibitors of different signaling molecules. For example, the TKI erlotinib—a potent EGFR inhibitor—alone (trial 1) or in combination with bevacizumab (anti VEGF) (trial 2) were administered in three different phase II clinical trials (NCT00365391, NCT02273362 and NCT00356889, respectively) [186,187,188]. The two formers induce a poor partial response (PR) (1 patient out of 27), a median time to disease progression of 3.0 months and a median survival time of 9.5 months in the combined therapy [189]. The latter, evaluated in patients with CCA, showed a PR of 12% after 6 months, a survival time of 9.9 months and a progression free survival (PFS) of 4.4 months.
The combination of erlotinib plus sorafenib phase II (NCT01093222) [190] also failed due to a very low PFS (of 2 months) and an overall survival (OS) of 6 months [190]. The safety and efficacy of the TKI lapatinib were also studied in a phase II trial (NCT00107536). The results indicate it is safe, but its OS and PFS are 12.6 and 1.9 months, respectively. Hence, efficacy needs further investigations [191].
Panitumumab (a monoclonal anti-EGFR antibody) plus gemcitabine and irinotecan [192] have also been tested in a phase II, showing a response of 31.4% and good tolerability, but in a short period of 5 months.
Based on all this, further research is required to define the resistance mechanisms to EGFR inhibition that operates in HCC. In this sense, EGFR signaling as a driver or modulator of liver inflammation in CLD and cancer [193,194] should be taken into consideration when thinking about EGFR inhibition to combat HCC. A better comprehension of this specific aspect of the EGFR signaling role in liver disease deserves further research. In this regard, it would be important to better understand the specific actions of EGFR signaling in the liver TME.

3.2. TGF-β

It is widely recognized that TGF-β plays critical roles in tumor initiation, development, and the generation of metastasis in several cancer types. Strikingly, TGF-β switches from a potent cytostatic and pro-apoptotic effect in normal epithelial cells to a tumor promoter activity at the late stages of the disease, a phenomenon known as the “TGF-β paradox” [195]. In a revealing work by Coulouarn et al. (2018) [196] two different TGF-β gene signatures were proposed in HCC. An “early signature”, characterized by the expression of genes related to tumor suppressor activities, and associated with the longer survival of HCC patients, and a “late signature” that includes a set of genes related to TGF-β tumor promoter effects such as positive regulators of cell cycle and EMT.
TGF-β is involved in all stages of CLD, from steatosis and inflammation to fibrosis, cirrhosis and HCC [195,196,197,198]. Therefore, many clinical studies point at TGF-β as a strong candidate for being a diagnostic and prognostic marker, as well as a target of therapeutic strategies. In fact, circulating TGF-β1 levels are elevated in HCC patients [199], which likely makes it a suitable marker for HCC diagnosis. The expression of TGF-β1 is also significantly higher in HCC tissues than in normal liver, and a correlation between TGF-β levels and poor prognosis, extrahepatic metastasis, lower survival rates and lower post-operative disease-free survival has been established [200].
The idea of targeting TGF-β in HCC is certainly promising but not that simple. Unfortunately, how HCC cells surpass the suppressive effect of TGF-β is not completely clear, although a few mechanisms have been described. Inactivating somatic mutations is one of them. A recent genomic and transcriptomic study [201] reported that almost 40% of HCC samples contain mutations in genes of the TGF-β signaling pathway. Some of them correlate with the inactivation of the pathway, due to the loss of TGF-β suppressor activity, while others result in an overactivation of the pathway, again making evident the TGF-β paradox. Other mechanisms include alterations in TGF-β receptor internalization [202,203]; epigenetic silencing events [204]; aberrant epitranscriptomic RNA modifications [205] or the expression of specific miRNAs that allow cells to escape from TGF-β-induced apoptosis [206]. Additionally, the overactivation of survival pathways that interfere with the Smad canonical pathway and/or crosstalk with non-canonical TGF-β signaling impacting on transcriptional regulation and the suppressive effects of TGF-β can also occur [207,208,209].
Likewise, the pro-tumorigenic effect of the TGF-β pathway is mediated by complex and diverse mechanisms that combine actions on the cancer cell itself, but also on the tumor stroma and microenvironment [210]. In this respect, since TGF-β is already present during the inflammatory response activated in CLD, it could build up a microenvironment optimal for HCC growth. TGF-β is an important player in the dialogue between cancer and stroma cells, as mentioned before, favoring the secretion of cytokines, like CLCF1 (CAF-derived cardiotrophin-like cytokine factor 1) [211]. This effect, together with the fact that TGF-β is a potent inducer of EMT, result in a positive feedback loop that sustains HCC cells´ proliferation and invasion through the release of growth factors, chemokines and cytokines. TGF-β is also a pro-angiogenic factor in HCC, acting both directly on the endothelial cells or indirectly, modulating the production of VEGF [212,213]. Finally, TGF-β is a key immunoregulator with immunosuppressive effects on innate and adaptive immune cells that result in tumor immune escape in HCC [214].
Encouraged by the promising results of clinical trials of galunisertib (a TGF-βRI kinase inhibitor factor), alone or in combination, for the treatment of advanced HCC [215], three different phase II trials (Table 1) investigated galunisertib, alone or in combination with sorafenib (NCT02178358) [216], the anti-PD-1 antibody nivolumab (NCT02423343) [217] or the anti-VEGFR antibody ramucirumab (NCT01246986) [218]. In the first one, neither the single nor the combination approach showed any benefit in OS nor in PFS. The second one had only one patient, and therefore, no conclusions can be drawn. For the latter, the combination with ramucirumab lacked results, but the combination with nivolumab showed better PFS. Both approaches showed similar reduction of AFP (alpha-fetoprotein) and TGF-β levels after the treatments [219]. Therefore, anti TGF-β therapy needs to be reevaluated considering the complexity of TGF-β effects. On one hand, further research is required to clarify the mechanisms behind the “TGF-β paradox” in HCC and to better understand the diverse functions of TGF-β in the different hepatic cell populations, and its secretion by several liver cells and platelets. On the other hand, a careful stratification of patients according to their TGF-β profile may help to guide clinical decision to improve patient outcomes.
Table 1. Completed clinical trials for HCC or CCA with published results.
Table 1. Completed clinical trials for HCC or CCA with published results.
Trial IDStudy TitleTargetInterventionResultsPhaseReferences
NCT01271504E7050 in combination with Sorafenib versus Sorafenib alone as first line therapy in participants with hepatocellular carcinoma (HCC)METGolvatinib
Sorafenib
-
TEAEs: 100%
-
SAEs: 47.6%
-
TTP: 10.29 weeks
-
PFS: 10.29 weeks
-
OS: 27.86 weeks
-
OR: 4.8%
1–2[159]
NCT00107536Lapatinib ditosylate in treating patients with unresectable liver or biliary tract cancerEGFRLapatinib
-
PFS: 1.9 months
-
OS: 12.6 months
-
Target-EGFR/EGFR-P protein expression: 0%
-
Expression profile and mutations of genes critical for EGFR and ERBB2 Signaling: 0%
2[191,220]
NCT00365391Bevacizumab and Erlotinib in treating patients with advanced liver cancerEGFRErlotinib
Bevacizumab
-
OR: 4.3%
-
TTP: 3 months
-
OS: 9.5 months
-
TTF: 2 months
2[187]
NCT00356889Bevacizumab and Erlotinib hydrochloride in treating patients with metastatic or unresectable biliary tumorsEGFRErlotinib
Bevacizumab
-
OR: 12%
-
OS: 9.9 months
-
PFS: 4.4 months
-
DOR: 8.4 months
2[186]
NCT00753675Vandetanib, Gemcitabine or placebo plus Gemcitabine or Vandetanib monotherapy in advanced biliary tract cancerEGFRVandetanib
Gemcitabine
-
PFS: 105–148 days
-
OR: 3.5–15.5%
-
DOR: 127–277 days
-
OS: 228–307 days
2[221]
NCT00948935Study of Gemcitabine, Irinotecan and Panitumumab in patients with advanced and metastatic biliary tract adenocarcinomaEGFRPanitumumab
Gemcitabine Irinotecan
-
PFS: 69%
-
OR: 31.4%
2[192,222]
NCT01093222Sorafenib Tosylate and Erlotinib hydrochloride in treating patients with locally advanced, unresectable, or metastatic gallbladder cancer or cholangiocarcinoma (CCA)EGFRErlotinib
Sorafenib
-
PFS: 2 months
-
OS: 6 months
-
OR: 6%
-
TEAEs: 100%
2[190]
NCT02273362Erlotinib hydrochloride in preventing liver cancer in patients with cirrhosis of the liverEGFRErlotinib
-
Response (at Least a 50% Reduction in Liver Phospho-EGFR Staining): 100, 60, 60%
-
TEAEs: 0–16.7%
1–2[188,189]
NCT01246986A Study of LY2157299 in participants with HCCTGF-βGalunisertib
Sorafenib
Ramucirumab
-
Change from baseline in relationship of AFP to OS: 17.9–24.2%
-
Change from baseline in relationship of TGF-β to OS: 10.1–22.88%
-
TTP: 7.1–36 weeks
-
OS: 29.6–89.6 weeks
-
PFS: 6.6–28.4 weeks
-
OR: 0.0–3.7%
-
DOR: 37.6–47.2%
-
TTF: 9.9–49.3 weeks
-
TTW: 113–113 days
2[218,219,223]
NCT02178358A Study of LY2157299 in participants with advanced HCCTGF-βGalunisertib
Sorafenib
-
OS: 75–81.6%
-
TTP: 1.4–4.1 months
-
PFS: 1.4–4.1 months
-
OR: 0.03–0.16%
2[216]
NCT02423343A Study of Galunisertib (LY2157299) in combination with Nivolumab in advanced refractory solid tumors and in recurrent or refractory NSCLC, or HCCTGF-βGalunisertib
Nivolumab
-
PFS: 5.26, 5.39,
-
OR: 24, 0%
-
DOR: 9.03 months
-
TTR: 4.2 months
-
OS: 11.99–14.52 months
1–2[217]
Compilation of the interventional clinical trials targeting MET, EGFR, or TGF-β, registered at the EU Clinical Trials Register (https://www.clinicaltrialsregister.eu, accessed on 28 September 2023) and the U.S. National Library of Medicine (https://clinicaltrials.gov, accessed on 28 September 2023). Only completed interventional trials with results posted are listed. Abbreviations: (TEAEs) Treatment-emergent Adverse Events; (SAEs) Serious Adverse Events; (TTP) Time to Progression; (PFS) Progression Free Survival; (OS) Overall Survival; (OR) Overall/Objective response (Complete Response or Partial Response); (DOR) Duration of Tumor Response; (TTF) Time to Treatment Failure; (TTW) Time to Worsening of Symptoms. Glossary: (TEAEs) Treatment-emergent Adverse Events. Undesirable events not existing prior to medical treatment, or an already existing event that worsens either in intensity or frequency following the treatment. (SAEs) Serious Adverse Events. A life-threatening adverse event (related with the medical treatment). (TTP) Time to Progression. The interval of time from the start of treatment to disease progression. Similar to (TTW) Time to Worsening of Symptoms. (PFS) Progression Free Survival. In a clinical trial, the interval of time from random assignment (placebo or treatment) to disease progression or death from any cause. (OS) Overall Survival. In a clinical trial, the length of time from randomization to death (OR) Overall/Objective response (Complete Response or Partial Response). The percentage of patients who achieve a complete response or partial response. Complete response: total disappearance of lesions. Partial response: reduction in the sum of maximal tumor diameters by at least 30% or more (following Response Evaluation Criteria in Solid Tumors (RECIST)). (DOR) Duration of Tumor Response (also known as Duration of clinical benefit (DoCB)). In a clinical trial, the period from randomization to disease progression or death in patients who achieve complete or partial response. (TTF) Time to Treatment Failure. In a clinical trial, the interval from chemotherapy (treatment) initiation to premature discontinuation (loss of efficacy, SAEs, death, patient voluntary termination, etc.).

3.3. C3G, a New Signaling Player in HCC

C3G (Crk SH3-domain-binding guanine-nucleotide-releasing factor), encoded by RapGEF1 gene is a guanine nucleotide exchange factor (GEF) for Rap and other Ras proteins [224], which can also act through mechanisms not dependent on its GEF activity [225]. C3G plays a dual role in cancer acting as either a tumor suppressor or promoter depending on tumor type and stage [225,226,227,228,229].
C3G is expressed at low levels in the adult liver and hepatocytes, while in HPCs and neonatal hepatocytes is highly expressed [227]. Importantly, C3G expression is upregulated in HCC patient samples, liver cancer mouse models and HCC cell lines, promoting tumor growth (Figure 2). Hence, high RapGEF1 mRNA levels correlate with tumor progression and a lower patient survival rate. In addition to C3G upregulation, genomic databases analyses show that RapGEF1 is commonly altered in liver cancer (HCC and CCA). An analysis performed on 1829 samples/1710 patients showed genetic alterations in 2% of the samples, amplification or missense events being the most predominant aberrations [230,231,232,233]. Interestingly, most missense mutations are located in conserved regions (encoding catalytic, N-terminal inhibitory, or proline-rich domains) and might have functional consequences dysregulating C3G activity. It is also important to highlight that C3G is necessary for a proper activation of HGF/MET signaling in HCC cells. Additionally, as mentioned before, our own unpublished data also suggest that platelet C3G favors DEN+CCl4-induced liver cancer.
On the other hand, recent published data suggest that the Crk-C3G-Rap1 pathway could improve endothelial cell integrity in HCC upon treatment with the anti-tumor compound lenvatinib, facilitating tumor suppression when combined with anti-PD-1 [234].
C3G also regulates HPC biology [235]. C3G down-regulation favors a partial EMT associated with stemness and enhanced migration. C3G is required for HGF/MET signaling and its pro-invasive activity in HPCs, while TGF-β signaling is enhanced when C3G is down-regulated (Figure 2). Therefore, changes in C3G levels alter HPC signaling, which may impact liver repair in CLD. Thus, C3G down-regulation is detected in livers from DDC-treated mice, an experimental model of cholestatic liver injury in which HPC expansion occurs. Whether this might become relevant for HPC malignant transformation and HCC development in specific contexts deserves further analysis.

4. Liver Cancer Mouse Models

Several mouse models have been used to study liver cancer development, including genetically engineered mouse models, exposure to chemical agents, intrahepatic or intrasplenic injection of tumor cells and xenograft approaches [236]. The choice of the appropriate model depends on the scope of the research since these models cover many conditions that go from the chronic administration of chemicals to induce tumor development to acute injection of tumor cells (HCC or CCA). Additionally, these methods can be combined with extra agents to mimic the liver disease environment using specific diets, injection of chemotoxic agents or the expression of inflammation promoting genes. We will briefly review the available animal models on liver cancer with special focus on the pathways mentioned previously.
Several genetically engineered mouse models mimic HCC/CCA based on the activation of oncogenes or the inactivation of tumor suppressor genes through different strategies that go from genome editing to infection with HBV or HCV. Most of these strategies target tumor-related genes/pathways that have proven to be relevant in liver cancer like the MET/HGF axis, the EGFR, ErbB-2A, Wnt, P53, PTEN, Akt, TGF-α, Myc, E2F, or KRAS pathways (Table 2) [149,150,237,238,239,240,241,242,243,244,245,246,247,248].
Induced models include the use of agents (specific diets, carcinogens, etc.) to generate liver tumors [236,249,250]. Several chemotoxic agents have been used to promote liver cancer including agents able to alkylate DNA and/or promote oxidative stress (DEN, N-Nitrosomorpholine (NMOR), Dimethylnitrosoamine (DMN), 2-acetylaminofluorene (2-AAF), thioacetamide (TAA) [251,252], agents that promote RAS mutations (DMBA (dimethylbenz(a)anthracene) [253], or the cholangiocarcinogen furan [254,255,256,257]. Furthermore, dietary-based models have been developed to induce NAFLD/NASH or other pre-cancer conditions relying on either high content of nutrients (carbohydrate, fat and cholesterol) or nutrient deficient diets (Table 2) [258]. Among the second group the most frequently used are the methionine and choline-deficient (MCD), and choline-deficient, L-amino acid-defined (CDAA) diets [259,260,261,262,263,264]. These models of NASH may progress to HCC after 20 months. However, the natural resistance of mice to developing HCC makes them long-term models and sometimes unsuitable for liver cancer studies unless combined with other strategies. Thus, combined models have been used to promote HCC like the choline-deficient high-fat diet (CD-HFD) that combines obesity and NASH, promoting HCC development within 12–24 months. Similarly, choline-deficient, L-amino acid-defined, high-fat diet (CDA-HFD) induces NASH/NAFLD with a quick progression to HCC [258,259,263,264].
In addition to the previous models, the most common method to establish HCC or CCA is through the injection of tumor cell lines (implantation models) through heterotopic/orthotopic injections in immunodeficient mice models [236,265]. Additionally, the syngeneic models allow for the use of a recipient with a functional immune system granting the study of immune cells on tumor formation and progression. Lately, the implantation models have incorporated human immune cells that invade the tumor and mimic a more realistic TME (humanized mouse models). To fully recapitulate human disease HCC or CCA, liver cancer-induced models are usually combined with agents that are able to induce liver diseases (HBV, HCV, alcohol, CCl4 or special diet) [236,265].
Lastly, treatment to induce chronic cholestasis in combination with DMN or DEN promotes CCA and induces genetic aberrations that serve as the basis for several animal models of CCA [250,266].
Despite the great variety of models, most of them share common genetic alterations that are able to induce human liver cancer, some of the most relevant: oncogenes (MYC, MET, RAS, NICD1, ERB2, NOTCH, AKT), tumor suppressors (P53, PTEN, FBXW7), and the WNT pathway (CTNNB1) [236]. Several of these pathways converge on the activation of intracellular pathways that rely on small GTPases (Ras, Rac, Rho) to transduce the oncogenic signaling responsible for the cellular transformation highlighting the role of their modulators like GEFs and GAPs.
Table 2. Liver cancer mouse models.
Table 2. Liver cancer mouse models.
Genetically engineered mouse models
Model (Oncogene/TSG)Altered PathwayType of CancerReferences
HBV, HCVViral modelHCC 13–24 months[267,268]
WNT1, CTNNB1Wnt pathwayHCC[240,248]
NOTCH1Notch pathwayCCA[247]
P53, myc, E2FCell cycleHCC[237,245,246]
PTEN, PTEN/SMAD4PI3K/Akt pathwayHCC, CCA (SMAD4/PTEN)[243,244]
IGF2Insulin growth factor pathwayHCC[242]
EGFR, ERBB2EGF pathwayHCC, CCA[172,241]
HGFR (met), HGFHGF signalingHCC (combination with DEN, b-catenin)[149,150,239,240,269,270]
TGF-α (+Myc), TGF-α/TGF-βEGFR signalingHCC[271,272,273,274]
KRAS/HRASRas signalingHCC, CCA (combination with PTEN)[237]
Chemotoxic agents
Model (agent)Mechanism of actionType of cancerReferences
DEN/DEN-CCL4Genotoxic hepatocarcinogen50–90 weeks 100% HCC[252,257,275,276,277,278]
NMORGenotoxic12 weeks HCC with lung metastasis[279]
DMNAlkylate DNA and/or promote oxidative stressPromote HCC[280,281]
2-AAFAlkylate DNA and/or promote oxidative stressPromote HCC[282]
DMBAInduces Ras mutationPromote HCC[253]
TAAGenotoxicPromote HCC and CCA[251,252]
FuranGenotoxicPromote CCA[256]
Dietary models
Model (diet)Mechanism of actionType of cancerReferences
High nutrientNASH/NAFLDmore than 80 weeks/20% HCC[258]
MCDOxidative DNA damage and chromosomal instability30–35 weeks 25–100% HCC[261,262]
CDEOxidative DNA damage and chromosomal instability30–35 weeks 25–100% HCC[260]
CDAAOxidative DNA damage and chromosomal instability84 weeks 100% HCC[259]
CDHFDOxidative DNA damage and chromosomal instability30–35 weeks 100% HCC[258,264]
CDAHFDOxidative DNA damage and chromosomal instability30–35 weeks 100% HCC[258,263]
Implantation models [264,265,266]
ModelCommentsAdvantagesDisadvantages
HeterotopicSubcutaneous inoculation of human cultured cellsQuick evaluation of tumor growthNo immune response
OrthotopicLiver implantation of human cultured cellsReproduce TME in immunodeficient mouseUnable to trigger an immune response
SyngeneicHeterotopic or orthotopic implantation of mouse tumor cellsReproduce TME and mimic the metastatic behavior of HC in immunocompetent mouseDifferences among human and mouse disease
Humanized mouse modelsTransplantation of cancer patient tissue directly into immunodeficient miceGenetic and histological similarities.
Identification of treatments
Reflect human disease and allow pharmacological testing
Summary of the most relevant methods to induce liver cancer. Abbreviations: HCC, hepatocellular carcinoma; CCA, cholangiocarcinoma; DEN, diethylnitrosamine; CCL4, Carbon Tetrachloride; NMOR, N-Nitrosomorpholine; DMN, dimethylnitrosamine; 2-AAF, 2-acetylaminofluorene; DMBA, 9,10-dimethyl-1,2-benzanthracene; TAA, Thioacetamide; MCD, methionine and choline- deficient; CDE, choline-deficient and methionine-supplemented; CDAA, choline-deficient l-amino-defined diet; CDHFD, choline-deficient high-fat diet; CDAHFD, choline-deficient, L-amino acid-defined, high-fat diet.

5. Conclusions and Future Directions

Most recent studies reinforce the idea that the TME plays a key role in liver cancer, with platelets emerging as important players with a potential value for diagnosis, prognosis, and treatment. On the other hand, although fibrosis has been classically associated with HCC development and progression, several exceptions came out, as well as new mechanisms controlling fibrosis. Therefore, the complex interplay between fibrosis and liver cancer deserves a deep analysis to find out the mechanisms determining its positive or negative role. For this, it could be helpful to understand particular features (etiology- dependent and-independent) of the fibrotic process that define its specific impact on liver cancer.
It is also remarkable that HGF/MET/C3G, EGFR and TGF-β signaling regulates not only hepatocyte cell fate and function, but also the function of most liver cells including HSCs, CAFs, and immune cells. Moreover, HGF, EGF and TGF-β can be secreted by platelets and different cells from the liver microenvironment. This, together with the generation of resistance to treatments, explain the limited effects of HCC therapy with inhibitors of these pathways. Hence, new approaches should be considered to improve treatment based on patient classification into subgroups according to gene signatures that reflect not only tumor cell gene profile, but also tumor microenvironment and tumor stage.

Author Contributions

Conceptualization, A.P. and A.S.; writing—original draft preparation, Á.M.C., N.P., P.B., A.G.-U., B.H., A.S. and A.P.; writing—review and editing, A.P., A.S., Á.M.C.; B.H. and N.P.; funding acquisition, A.P., A.S., B.H., P.B., A.G.-U. and Á.M.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by grants from the Spanish Ministry of Economy and Competitiveness [PID2022-137717OB-C21 to AP/AC funded by MCIN/AEI; PID2019-104143RB-C22 to AP funded by MCIN/AEI/10.13039/501100011033; PID2019-104991RB-I00 to PB, PID2020-117650RA-I00 to A G-U and PID2021-124830OB-I00 to AS/BH]. All funding was cosponsored by the European FEDER Program. NP was a recipient of FPU fellowship from Spanish Ministry of Education. A G-U.

Acknowledgments

We apologize to the authors whose work is not included in the references owing to space limitations.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Llovet, J.M.; Kelley, R.K.; Villanueva, A.; Singal, A.G.; Pikarsky, E.; Roayaie, S.; Lencioni, R.; Koike, K.; Zucman-Rossi, J.; Finn, R.S. Hepatocellular Carcinoma. Nat. Rev. Dis. Primers 2021, 7, 6. [Google Scholar] [CrossRef] [PubMed]
  2. Zheng, B.; Zhu, Y.J.; Wang, H.Y.; Chen, L. Gender Disparity in Hepatocellular Carcinoma (HCC): Multiple Underlying Mechanisms. Sci. China Life Sci. 2017, 60, 575–584. [Google Scholar] [CrossRef] [PubMed]
  3. Llovet, J.M.; Zucman-Rossi, J.; Pikarsky, E.; Sangro, B.; Schwartz, M.; Sherman, M.; Gores, G. Hepatocellular Carcinoma. Nat. Rev. Dis. Primers 2016, 2, 16018. [Google Scholar] [CrossRef] [PubMed]
  4. Philips, C.A.; Rajesh, S.; Nair, D.C.; Ahamed, R.; Abduljaleel, J.K.; Augustine, P. Hepatocellular Carcinoma in 2021: An Exhaustive Update. Cureus 2021, 13, e19274. [Google Scholar] [CrossRef] [PubMed]
  5. Montironi, C.; Castet, F.; Haber, P.K.; Pinyol, R.; Torres-Martin, M.; Torrens, L.; Mesropian, A.; Wang, H.; Puigvehi, M.; Maeda, M.; et al. Inflamed and Non-Inflamed Classes of HCC: A Revised Immunogenomic Classification. Gut 2023, 72, 129–140. [Google Scholar] [CrossRef] [PubMed]
  6. Ogunwobi, O.O.; Harricharran, T.; Huaman, J.; Galuza, A.; Odumuwagun, O.; Tan, Y.; Ma, G.X.; Nguyen, M.T. Mechanisms of Hepatocellular Carcinoma Progression. World J. Gastroenterol. 2019, 25, 2279–2293. [Google Scholar] [CrossRef]
  7. Suresh, D.; Srinivas, A.N.; Kumar, D.P. Etiology of Hepatocellular Carcinoma: Special Focus on Fatty Liver Disease. Front. Oncol. 2020, 10, 601710. [Google Scholar] [CrossRef]
  8. Vetrano, E.; Rinaldi, L.; Mormone, A.; Giorgione, C.; Galiero, R.; Caturano, A.; Nevola, R.; Marfella, R.; Sasso, F.C. Non-Alcoholic Fatty Liver Disease (NAFLD), Type 2 Diabetes, and Non-Viral Hepatocarcinoma: Pathophysiological Mechanisms and New Therapeutic Strategies. Biomedicines 2023, 11, 468. [Google Scholar] [CrossRef]
  9. Anstee, Q.M.; Reeves, H.L.; Kotsiliti, E.; Govaere, O.; Heikenwalder, M. From NASH to HCC: Current Concepts and Future Challenges. Nat. Rev. Gastroenterol. Hepatol. 2019, 16, 411–428. [Google Scholar] [CrossRef]
  10. Eslam, M.; Sanyal, A.J.; George, J.; Sanyal, A.; Neuschwander-Tetri, B.; Tiribelli, C.; Kleiner, D.E.; Brunt, E.; Bugianesi, E.; Yki-Järvinen, H.; et al. MAFLD: A Consensus-Driven Proposed Nomenclature for Metabolic Associated Fatty Liver Disease. Gastroenterology 2020, 158, 1999–2014.e1. [Google Scholar] [CrossRef]
  11. Craig, A.J.; von Felden, J.; Garcia-Lezana, T.; Sarcognato, S.; Villanueva, A. Tumour Evolution in Hepatocellular Carcinoma. Nat. Rev. Gastroenterol. Hepatol. 2020, 17, 139–152. [Google Scholar] [CrossRef] [PubMed]
  12. Llovet, J.M.; Montal, R.; Sia, D.; Finn, R.S. Molecular Therapies and Precision Medicine for Hepatocellular Carcinoma. Nat. Rev. Clin. Oncol. 2018, 15, 599–616. [Google Scholar] [CrossRef] [PubMed]
  13. Schulze, K.; Nault, J.C.; Villanueva, A. Genetic Profiling of Hepatocellular Carcinoma Using Next-Generation Sequencing. J. Hepatol. 2016, 65, 1031–1042. [Google Scholar] [CrossRef] [PubMed]
  14. Zucman-Rossi, J.; Villanueva, A.; Nault, J.C.; Llovet, J.M. Genetic Landscape and Biomarkers of Hepatocellular Carcinoma. Gastroenterology 2015, 149, 1226–1239.e4. [Google Scholar] [CrossRef]
  15. Dimitroulis, D.; Damaskos, C.; Valsami, S.; Davakis, S.; Garmpis, N.; Spartalis, E.; Athanasiou, A.; Moris, D.; Sakellariou, S.; Kykalos, S.; et al. From Diagnosis to Treatment of Hepatocellular Carcinoma: An Epidemic Problem for Both Developed and Developing World. World J. Gastroenterol. 2017, 23, 5282–5294. [Google Scholar] [CrossRef] [PubMed]
  16. Reig, M.; Forner, A.; Rimola, J.; Ferrer-Fàbrega, J.; Burrel, M.; Garcia-Criado, Á.; Kelley, R.K.; Galle, P.R.; Mazzaferro, V.; Salem, R.; et al. BCLC Strategy for Prognosis Prediction and Treatment Recommendation: The 2022 Update. J. Hepatol. 2022, 76, 681–693. [Google Scholar] [CrossRef]
  17. Wang, Y.; Deng, B. Hepatocellular Carcinoma: Molecular Mechanism, Targeted Therapy, and Biomarkers. Cancer Metastasis Rev. 2023, 42, 629–652. [Google Scholar] [CrossRef]
  18. Santhakumar, C.; Gane, E.J.; Liu, K.; McCaughan, G.W. Current Perspectives on the Tumor Microenvironment in Hepatocellular Carcinoma. Hepatol. Int. 2020, 14, 947–957. [Google Scholar] [CrossRef]
  19. Gu, Y.; Guo, Y.; Gao, N.; Fang, Y.; Xu, C.; Hu, G.; Guo, M.; Ma, Y.; Zhang, Y.; Zhou, J.; et al. The Proteomic Characterization of the Peritumor Microenvironment in Human Hepatocellular Carcinoma. Oncogene 2022, 41, 2480–2491. [Google Scholar] [CrossRef] [PubMed]
  20. Dhar, D.; Baglieri, J.; Kisseleva, T.; Brenner, D.A. Mechanisms of Liver Fibrosis and Its Role in Liver Cancer. Exp. Biol. Med. 2020, 245, 96–108. [Google Scholar] [CrossRef]
  21. Yavuz, B.G.; Pestana, R.C.; Abugabal, Y.I.; Krishnan, S.; Chen, J.; Hassan, M.M.; Wolff, R.A.; Rashid, A.; Amin, H.M.; Kaseb, A.O. Origin and Role of Hepatic Myofibroblasts in Hepatocellular Carcinoma. Oncotarget 2020, 11, 1186–1201. [Google Scholar] [CrossRef] [PubMed]
  22. Acharya, P.; Chouhan, K.; Weiskirchen, S.; Weiskirchen, R. Cellular Mechanisms of Liver Fibrosis. Front. Pharmacol. 2021, 12, 671640. [Google Scholar] [CrossRef] [PubMed]
  23. Caja, L.; Dituri, F.; Mancarella, S.; Caballero-Diaz, D.; Moustakas, A.; Giannelli, G.; Fabregat, I. TGF-β and the Tissue Microenvironment: Relevance in Fibrosis and Cancer. Int. J. Mol. Sci. 2018, 19, 1294. [Google Scholar] [CrossRef] [PubMed]
  24. Sánchez, P.S.; Rigual, M.d.M.; Djouder, N. Inflammatory and Non-Inflammatory Mechanisms Controlling Cirrhosis Development. Cancers 2021, 13, 5045. [Google Scholar] [CrossRef] [PubMed]
  25. Mederacke, I.; Hsu, C.C.; Troeger, J.S.; Huebener, P.; Mu, X.; Dapito, D.H.; Pradere, J.-P.; Schwabe, R.F. Fate Tracing Reveals Hepatic Stellate Cells as Dominant Contributors to Liver Fibrosis Independent of Its Aetiology. Nat. Commun. 2013, 4, 2823. [Google Scholar] [CrossRef] [PubMed]
  26. Marra, F.; Tacke, F. Roles for Chemokines in Liver Disease. Gastroenterology 2014, 147, 577–594.e1. [Google Scholar] [CrossRef]
  27. Yang, W.; He, H.; Wang, T.; Su, N.; Zhang, F.; Jiang, K.; Zhu, J.; Zhang, C.; Niu, K.; Wang, L.; et al. Single-Cell Transcriptomic Analysis Reveals a Hepatic Stellate Cell–Activation Roadmap and Myofibroblast Origin During Liver Fibrosis in Mice. Hepatology 2021, 74, 2774–2790. [Google Scholar] [CrossRef]
  28. Affo, S.; Yu, L.-X.; Schwabe, R.F. The Role of Cancer-Associated Fibroblasts and Fibrosis in Liver Cancer. Annu. Rev. Pathol. Mech. Dis. 2017, 12, 153–186. [Google Scholar] [CrossRef]
  29. Lujambio, A.; Akkari, L.; Simon, J.; Grace, D.; Tschaharganeh, D.F.; Bolden, J.E.; Zhao, Z.; Thapar, V.; Joyce, J.A.; Krizhanovsky, V.; et al. Non-Cell-Autonomous Tumor Suppression by P53. Cell 2013, 153, 449–460. [Google Scholar] [CrossRef]
  30. Duran, A.; Hernandez, E.D.; Reina-Campos, M.; Castilla, E.A.; Subramaniam, S.; Raghunandan, S.; Roberts, L.R.; Kisseleva, T.; Karin, M.; Diaz-Meco, M.T.; et al. P62/SQSTM1 by Binding to Vitamin D Receptor Inhibits Hepatic Stellate Cell Activity, Fibrosis, and Liver Cancer. Cancer Cell 2016, 30, 595–609. [Google Scholar] [CrossRef]
  31. Kocabayoglu, P.; Lade, A.; Lee, Y.A.; Dragomir, A.-C.; Sun, X.; Fiel, M.I.; Thung, S.; Aloman, C.; Soriano, P.; Hoshida, Y.; et al. β-PDGF Receptor Expressed by Hepatic Stellate Cells Regulates Fibrosis in Murine Liver Injury, but Not Carcinogenesis. J. Hepatol. 2015, 63, 141–147. [Google Scholar] [CrossRef]
  32. Filliol, A.; Saito, Y.; Nair, A.; Dapito, D.H.; Yu, L.X.; Ravichandra, A.; Bhattacharjee, S.; Affo, S.; Fujiwara, N.; Su, H.; et al. Opposing Roles of Hepatic Stellate Cell Subpopulations in Hepatocarcinogenesis. Nature 2022, 610, 356–365. [Google Scholar] [CrossRef]
  33. Wei, J.; Yao, J.; Yang, C.; Mao, Y.; Zhu, D.; Xie, Y.; Liu, P.; Yan, M.; Ren, L.; Lin, Y.; et al. Heterogeneous Matrix Stiffness Regulates the Cancer Stem-like Cell Phenotype in Hepatocellular Carcinoma. J. Transl. Med. 2022, 20, 555. [Google Scholar] [CrossRef] [PubMed]
  34. Zhao, W.; Mo, H.; Liu, R.; Chen, T.; Yang, N.; Liu, Z. Matrix Stiffness-Induced Upregulation of Histone Acetyltransferase KAT6A Promotes Hepatocellular Carcinoma Progression through Regulating SOX2 Expression. Br. J. Cancer 2022, 127, 202–210. [Google Scholar] [CrossRef] [PubMed]
  35. Cogliati, B.; Yashaswini, C.N.; Wang, S.; Sia, D.; Friedman, S.L. Friend or Foe? The Elusive Role of Hepatic Stellate Cells in Liver Cancer. Nat. Rev. Gastroenterol. Hepatol. 2023, 20, 647–661. [Google Scholar] [CrossRef] [PubMed]
  36. Ezhilarasan, D. Hepatic Stellate Cells in the Injured Liver: Perspectives beyond Hepatic Fibrosis. J. Cell Physiol. 2022, 237, 436–449. [Google Scholar] [CrossRef] [PubMed]
  37. Yu, M.-C.; Chen, C.-H.; Liang, X.; Wang, L.; Gandhi, C.R.; Fung, J.J.; Lu, L.; Qian, S. Inhibition of T-Cell Responses by Hepatic Stellate Cells via B7-H1-Mediated T-Cell Apoptosis in Mice. Hepatology 2004, 40, 1312–1321. [Google Scholar] [CrossRef] [PubMed]
  38. Chinnadurai, R.; Grakoui, A. B7-H4 Mediates Inhibition of T Cell Responses by Activated Murine Hepatic Stellate Cells. Hepatology 2010, 52, 2177–2185. [Google Scholar] [CrossRef]
  39. Karlmark, K.R.; Zimmermann, H.W.; Roderburg, C.; Gassler, N.; Wasmuth, H.E.; Luedde, T.; Trautwein, C.; Tacke, F. The Fractalkine Receptor CX3CR1 Protects against Liver Fibrosis by Controlling Differentiation and Survival of Infiltrating Hepatic Monocytes. Hepatology 2010, 52, 1769–1782. [Google Scholar] [CrossRef]
  40. Aoyama, T.; Inokuchi, S.; Brenner, D.A.; Seki, E. CX3CL1-CX3CR1 Interaction Prevents Carbon Tetrachloride-Induced Liver Inflammation and Fibrosis in Mice. Hepatology 2010, 52, 1390–1400. [Google Scholar] [CrossRef]
  41. Ramachandran, P.; Dobie, R.; Wilson-Kanamori, J.R.; Dora, E.F.; Henderson, B.E.P.; Luu, N.T.; Portman, J.R.; Matchett, K.P.; Brice, M.; Marwick, J.A.; et al. Resolving the Fibrotic Niche of Human Liver Cirrhosis at Single-Cell Level. Nature 2019, 575, 512–518. [Google Scholar] [CrossRef] [PubMed]
  42. Oura, K.; Morishita, A.; Tani, J.; Masaki, T. Tumor Immune Microenvironment and Immunosuppressive Therapy in Hepatocellular Carcinoma: A Review. Int. J. Mol. Sci. 2021, 22, 5801. [Google Scholar] [CrossRef] [PubMed]
  43. Hao, X.; Sun, G.; Zhang, Y.; Kong, X.; Rong, D.; Song, J.; Tang, W.; Wang, X. Targeting Immune Cells in the Tumor Microenvironment of HCC: New Opportunities and Challenges. Front. Cell Dev. Biol. 2021, 9, 775462. [Google Scholar] [CrossRef]
  44. Li, Y.; Wang, R.; Xiong, S.; Wang, X.; Zhao, Z.; Bai, S.; Wang, Y.; Zhao, Y.; Cheng, B. Cancer-Associated Fibroblasts Promote the Stemness of CD24+ Liver Cells via Paracrine Signaling. J. Mol. Med. 2019, 97, 243–255. [Google Scholar] [CrossRef] [PubMed]
  45. Xiong, S.; Wang, R.; Chen, Q.; Luo, J.; Wang, J.; Zhao, Z.; Li, Y.; Wang, Y.; Wang, X.; Cheng, B. Cancer-Associated Fibroblasts Promote Stem Cell-like Properties of Hepatocellular Carcinoma Cells through IL-6/STAT3/Notch Signaling. Am. J. Cancer Res. 2018, 8, 302–316. [Google Scholar] [PubMed]
  46. Bhattacharjee, S.; Hamberger, F.; Ravichandra, A.; Miller, M.; Nair, A.; Affo, S.; Filliol, A.; Chin, L.; Savage, T.M.; Yin, D.; et al. Tumor Restriction by Type I Collagen Opposes Tumor-Promoting Effects of Cancer-Associated Fibroblasts. J. Clin. Investig. 2021, 131, e146987. [Google Scholar] [CrossRef] [PubMed]
  47. Wohlleber, D.; Knolle, P.A. The Role of Liver Sinusoidal Cells in Local Hepatic Immune Surveillance. Clin. Transl. Immunol. 2016, 5, e117. [Google Scholar] [CrossRef]
  48. Crispe, I.N. The Liver as a Lymphoid Organ. Annu. Rev. Immunol. 2009, 27, 147–163. [Google Scholar] [CrossRef]
  49. Ringelhan, M.; Pfister, D.; O’Connor, T.; Pikarsky, E.; Heikenwalder, M. The Immunology of Hepatocellular Carcinoma. Nat. Immunol. 2018, 19, 222–232. [Google Scholar] [CrossRef]
  50. Li, X.; Ramadori, P.; Pfister, D.; Seehawer, M.; Zender, L.; Heikenwalder, M. The Immunological and Metabolic Landscape in Primary and Metastatic Liver Cancer. Nat. Rev. Cancer 2021, 21, 541–557. [Google Scholar] [CrossRef]
  51. Triantafyllou, E.; Woollard, K.J.; McPhail, M.J.W.; Antoniades, C.G.; Possamai, L.A. The Role of Monocytes and Macrophages in Acute and Acute-on-Chronic Liver Failure. Front. Immunol. 2018, 9, 2948. [Google Scholar] [CrossRef] [PubMed]
  52. Ramadori, P.; Kam, S.; Heikenwalder, M. T Cells: Friends and Foes in NASH Pathogenesis and Hepatocarcinogenesis. Hepatology 2022, 75, 1038–1049. [Google Scholar] [CrossRef] [PubMed]
  53. Malehmir, M.; Pfister, D.; Gallage, S.; Szydlowska, M.; Inverso, D.; Kotsiliti, E.; Leone, V.; Peiseler, M.; Surewaard, B.G.J.; Rath, D.; et al. Platelet GPIbα Is a Mediator and Potential Interventional Target for NASH and Subsequent Liver Cancer. Nat. Med. 2019, 25, 641–655. [Google Scholar] [CrossRef] [PubMed]
  54. Tian, Z.; Chen, Y.; Gao, B. Natural Killer Cells in Liver Disease. Hepatology 2013, 57, 1654–1662. [Google Scholar] [CrossRef] [PubMed]
  55. Park, O.; Jeong, W.-I.; Wang, L.; Wang, H.; Lian, Z.-X.; Gershwin, M.E.; Gao, B. Diverse Roles of Invariant Natural Killer T Cells in Liver Injury and Fibrosis Induced by Carbon Tetrachloride. Hepatology 2009, 49, 1683–1694. [Google Scholar] [CrossRef]
  56. Talamantes, S.; Lisjak, M.; Gilglioni, E.H.; Llamoza-Torres, C.J.; Ramos-Molina, B.; Gurzov, E.N. Non-Alcoholic Fatty Liver Disease and Diabetes Mellitus as Growing Aetiologies of Hepatocellular Carcinoma. JHEP Rep. 2023, 5, 100811. [Google Scholar] [CrossRef]
  57. Tarantino, G.; Citro, V.; Balsano, C.; Capone, D. Could SCGF-Beta Levels Be Associated with Inflammation Markers and Insulin Resistance in Male Patients Suffering from Obesity-Related NAFLD? Diagnostics 2020, 10, 395. [Google Scholar] [CrossRef]
  58. Wolf, M.J.; Adili, A.; Piotrowitz, K.; Abdullah, Z.; Boege, Y.; Stemmer, K.; Ringelhan, M.; Simonavicius, N.; Egger, M.; Wohlleber, D.; et al. Metabolic Activation of Intrahepatic CD8+ T Cells and NKT Cells Causes Nonalcoholic Steatohepatitis and Liver Cancer via Cross-Talk with Hepatocytes. Cancer Cell 2014, 26, 549–564. [Google Scholar] [CrossRef]
  59. Yahoo, N.; Dudek, M.; Knolle, P.; Heikenwälder, M. Role of Immune Responses in the Development of NAFLD-Associated Liver Cancer and Prospects for Therapeutic Modulation. J Hepatol 2023, 79, 538–551. [Google Scholar] [CrossRef]
  60. Dudek, M.; Pfister, D.; Donakonda, S.; Filpe, P.; Schneider, A.; Laschinger, M.; Hartmann, D.; Hüser, N.; Meiser, P.; Bayerl, F.; et al. Auto-Aggressive CXCR6+ CD8 T Cells Cause Liver Immune Pathology in NASH. Nature 2021, 592, 444–449. [Google Scholar] [CrossRef]
  61. Lassen, M.G.; Lukens, J.R.; Dolina, J.S.; Brown, M.G.; Hahn, Y.S. Intrahepatic IL-10 Maintains NKG2A+Ly49 Liver NK Cells in a Functionally Hyporesponsive State. J. Immunol. 2010, 184, 2693–2701. [Google Scholar] [CrossRef]
  62. Varol, C.; Yona, S.; Jung, S. Origins and Tissue-context-dependent Fates of Blood Monocytes. Immunol. Cell Biol. 2009, 87, 30–38. [Google Scholar] [CrossRef] [PubMed]
  63. Dal-Secco, D.; Wang, J.; Zeng, Z.; Kolaczkowska, E.; Wong, C.H.Y.; Petri, B.; Ransohoff, R.M.; Charo, I.F.; Jenne, C.N.; Kubes, P. A Dynamic Spectrum of Monocytes Arising from the In Situ Reprogramming of CCR2+ Monocytes at a Site of Sterile Injury. J. Exp. Med. 2015, 212, 447–456. [Google Scholar] [CrossRef] [PubMed]
  64. Karlmark, K.R.; Weiskirchen, R.; Zimmermann, H.W.; Gassler, N.; Ginhoux, F.; Weber, C.; Merad, M.; Luedde, T.; Trautwein, C.; Tacke, F. Hepatic Recruitment of the Inflammatory Gr1+ Monocyte Subset upon Liver Injury Promotes Hepatic Fibrosis. Hepatology 2009, 50, 261–274. [Google Scholar] [CrossRef] [PubMed]
  65. Li, X.; Yao, W.; Yuan, Y.; Chen, P.; Li, B.; Li, J.; Chu, R.; Song, H.; Xie, D.; Jiang, X.; et al. Targeting of Tumour-Infiltrating Macrophages via CCL2/CCR2 Signalling as a Therapeutic Strategy against Hepatocellular Carcinoma. Gut 2017, 66, 157–167. [Google Scholar] [CrossRef] [PubMed]
  66. Kuang, D.-M.; Peng, C.; Zhao, Q.; Wu, Y.; Chen, M.-S.; Zheng, L. Activated Monocytes in Peritumoral Stroma of Hepatocellular Carcinoma Promote Expansion of Memory T Helper 17 Cells. Hepatology 2010, 51, 154–164. [Google Scholar] [CrossRef] [PubMed]
  67. Noy, R.; Pollard, J.W. Tumor-Associated Macrophages: From Mechanisms to Therapy. Immunity 2014, 41, 49–61. [Google Scholar] [CrossRef]
  68. Yang, Y.; Ye, Y.-C.; Chen, Y.; Zhao, J.-L.; Gao, C.-C.; Han, H.; Liu, W.-C.; Qin, H.-Y. Crosstalk between Hepatic Tumor Cells and Macrophages via Wnt/β-Catenin Signaling Promotes M2-like Macrophage Polarization and Reinforces Tumor Malignant Behaviors. Cell Death Dis. 2018, 9, 793. [Google Scholar] [CrossRef]
  69. Samstein, R.M.; Arvey, A.; Josefowicz, S.Z.; Peng, X.; Reynolds, A.; Sandstrom, R.; Neph, S.; Sabo, P.; Kim, J.M.; Liao, W.; et al. Foxp3 Exploits a Pre-Existent Enhancer Landscape for Regulatory T Cell Lineage Specification. Cell 2012, 151, 153–166. [Google Scholar] [CrossRef]
  70. Chen, K.-J.; Lin, S.-Z.; Zhou, L.; Xie, H.-Y.; Zhou, W.-H.; Taki-Eldin, A.; Zheng, S.-S. Selective Recruitment of Regulatory T Cell through CCR6-CCL20 in Hepatocellular Carcinoma Fosters Tumor Progression and Predicts Poor Prognosis. PLoS ONE 2011, 6, e24671. [Google Scholar] [CrossRef]
  71. Garnelo, M.; Tan, A.; Her, Z.; Yeong, J.; Lim, C.J.; Chen, J.; Lim, K.H.; Weber, A.; Chow, P.; Chung, A.; et al. Interaction between Tumour-Infiltrating B Cells and T Cells Controls the Progression of Hepatocellular Carcinoma. Gut 2017, 66, 342–351. [Google Scholar] [CrossRef]
  72. Fu, J.; Xu, D.; Liu, Z.; Shi, M.; Zhao, P.; Fu, B.; Zhang, Z.; Yang, H.; Zhang, H.; Zhou, C.; et al. Increased Regulatory T Cells Correlate with CD8 T-Cell Impairment and Poor Survival in Hepatocellular Carcinoma Patients. Gastroenterology 2007, 132, 2328–2339. [Google Scholar] [CrossRef]
  73. Langhans, B.; Nischalke, H.D.; Krämer, B.; Dold, L.; Lutz, P.; Mohr, R.; Vogt, A.; Toma, M.; Eis-Hübinger, A.M.; Nattermann, J.; et al. Role of Regulatory T Cells and Checkpoint Inhibition in Hepatocellular Carcinoma. Cancer Immunol. Immunother. 2019, 68, 2055–2066. [Google Scholar] [CrossRef]
  74. Veglia, F.; Perego, M.; Gabrilovich, D. Myeloid-Derived Suppressor Cells Coming of Age. Nat. Immunol. 2018, 19, 108–119. [Google Scholar] [CrossRef]
  75. Kapanadze, T.; Gamrekelashvili, J.; Ma, C.; Chan, C.; Zhao, F.; Hewitt, S.; Zender, L.; Kapoor, V.; Felsher, D.W.; Manns, M.P.; et al. Regulation of Accumulation and Function of Myeloid Derived Suppressor Cells in Different Murine Models of Hepatocellular Carcinoma. J. Hepatol. 2013, 59, 1007–1013. [Google Scholar] [CrossRef] [PubMed]
  76. Hoechst, B.; Voigtlaender, T.; Ormandy, L.; Gamrekelashvili, J.; Zhao, F.; Wedemeyer, H.; Lehner, F.; Manns, M.P.; Greten, T.F.; Korangy, F. Myeloid Derived Suppressor Cells Inhibit Natural Killer Cells in Patients with Hepatocellular Carcinoma via the NKp30 Receptor. Hepatology 2009, 50, 799–807. [Google Scholar] [CrossRef] [PubMed]
  77. Coffelt, S.B.; Wellenstein, M.D.; de Visser, K.E. Neutrophils in Cancer: Neutral No More. Nat. Rev. Cancer 2016, 16, 431–446. [Google Scholar] [CrossRef]
  78. Liu, K.; Wang, F.-S.; Xu, R. Neutrophils in Liver Diseases: Pathogenesis and Therapeutic Targets. Cell Mol. Immunol. 2021, 18, 38–44. [Google Scholar] [CrossRef] [PubMed]
  79. Chen, L.; Gu, J.; Qian, Y.; Li, M.; Qian, Y.; Xu, M.; Li, J.; Wen, Y.; Xia, L.; Li, J.; et al. Deletion of C-C Motif Chemokine Ligand 5 Worsens Invariant Natural Killer T-Cell–Mediated Hepatitis via Compensatory Up-Regulation of CXCR2–Related Chemokine Activity. Cell Mol. Gastroenterol. Hepatol. 2019, 7, 623–639. [Google Scholar] [CrossRef] [PubMed]
  80. Zhou, S.-L.; Zhou, Z.-J.; Hu, Z.-Q.; Huang, X.-W.; Wang, Z.; Chen, E.-B.; Fan, J.; Cao, Y.; Dai, Z.; Zhou, J. Tumor-Associated Neutrophils Recruit Macrophages and T-Regulatory Cells to Promote Progression of Hepatocellular Carcinoma and Resistance to Sorafenib. Gastroenterology 2016, 150, 1646–1658.e17. [Google Scholar] [CrossRef]
  81. Li, L.; Xu, L.; Yan, J.; Zhen, Z.-J.; Ji, Y.; Liu, C.-Q.; Lau, W.Y.; Zheng, L.; Xu, J. CXCR2–CXCL1 Axis Is Correlated with Neutrophil Infiltration and Predicts a Poor Prognosis in Hepatocellular Carcinoma. J. Exp. Clin. Cancer Res. 2015, 34, 129. [Google Scholar] [CrossRef]
  82. Zhou, S.; Yin, D.; Hu, Z.; Luo, C.; Zhou, Z.; Xin, H.; Yang, X.; Shi, Y.; Wang, Z.; Huang, X.; et al. A Positive Feedback Loop Between Cancer Stem-Like Cells and Tumor-Associated Neutrophils Controls Hepatocellular Carcinoma Progression. Hepatology 2019, 70, 1214–1230. [Google Scholar] [CrossRef] [PubMed]
  83. Spiegel, A.; Brooks, M.W.; Houshyar, S.; Reinhardt, F.; Ardolino, M.; Fessler, E.; Chen, M.B.; Krall, J.A.; DeCock, J.; Zervantonakis, I.K.; et al. Neutrophils Suppress Intraluminal NK Cell–Mediated Tumor Cell Clearance and Enhance Extravasation of Disseminated Carcinoma Cells. Cancer Discov. 2016, 6, 630–649. [Google Scholar] [CrossRef] [PubMed]
  84. Labelle, M.; Begum, S.; Hynes, R.O. Platelets Guide the Formation of Early Metastatic Niches. Proc. Natl. Acad. Sci. USA 2014, 111, E3053–E3061. [Google Scholar] [CrossRef] [PubMed]
  85. Szczerba, B.M.; Castro-Giner, F.; Vetter, M.; Krol, I.; Gkountela, S.; Landin, J.; Scheidmann, M.C.; Donato, C.; Scherrer, R.; Singer, J.; et al. Neutrophils Escort Circulating Tumour Cells to Enable Cell Cycle Progression. Nature 2019, 566, 553–557. [Google Scholar] [CrossRef] [PubMed]
  86. Honda, M.; Kubes, P. Neutrophils and Neutrophil Extracellular Traps in the Liver and Gastrointestinal System. Nat. Rev. Gastroenterol. Hepatol. 2018, 15, 206–221. [Google Scholar] [CrossRef]
  87. Yang, L.; Liu, Q.; Zhang, X.; Liu, X.; Zhou, B.; Chen, J.; Huang, D.; Li, J.; Li, H.; Chen, F.; et al. DNA of Neutrophil Extracellular Traps Promotes Cancer Metastasis via CCDC25. Nature 2020, 583, 133–138. [Google Scholar] [CrossRef]
  88. Schneider, C.; Teufel, A.; Yevsa, T.; Staib, F.; Hohmeyer, A.; Walenda, G.; Zimmermann, H.W.; Vucur, M.; Huss, S.; Gassler, N.; et al. Adaptive Immunity Suppresses Formation and Progression of Diethylnitrosamine-Induced Liver Cancer. Gut 2012, 61, 1733–1743. [Google Scholar] [CrossRef]
  89. Mussbacher, M.; Brunnthaler, L.; Panhuber, A.; Starlinger, P.; Assinger, A. Till Death Do Us Part—The Multifaceted Role of Platelets in Liver Diseases. Int. J. Mol. Sci. 2021, 22, 3113. [Google Scholar] [CrossRef]
  90. Liu, P.; Hsu, C.; Su, C.; Huang, Y.; Hou, M.; Rich, N.E.; Fujiwara, N.; Hoshida, Y.; Singal, A.G.; Huo, T. Thrombocytosis Is Associated with Worse Survival in Patients with Hepatocellular Carcinoma. Liver Int. 2020, 40, 2522–2534. [Google Scholar] [CrossRef]
  91. Scheiner, B.; Kirstein, M.; Popp, S.; Hucke, F.; Bota, S.; Rohr-Udilova, N.; Reiberger, T.; Müller, C.; Trauner, M.; Peck-Radosavljevic, M.; et al. Association of Platelet Count and Mean Platelet Volume with Overall Survival in Patients with Cirrhosis and Unresectable Hepatocellular Carcinoma. Liver Cancer 2019, 8, 203–217. [Google Scholar] [CrossRef] [PubMed]
  92. Poujol-Robert, A.; Boëlle, P.-Y.; Conti, F.; Durand, F.; Duvoux, C.; Wendum, D.; Paradis, V.; Mackiewicz, V.; Chazouillères, O.; Corpechot, C.; et al. Aspirin May Reduce Liver Fibrosis Progression: Evidence from a Multicenter Retrospective Study of Recurrent Hepatitis C after Liver Transplantation. Clin. Res. Hepatol. Gastroenterol. 2014, 38, 570–576. [Google Scholar] [CrossRef]
  93. Lee, M.; Chung, G.E.; Lee, J.; Oh, S.; Nam, J.Y.; Chang, Y.; Cho, H.; Ahn, H.; Cho, Y.Y.; Yoo, J.; et al. Antiplatelet Therapy and the Risk of Hepatocellular Carcinoma in Chronic Hepatitis B Patients on Antiviral Treatment. Hepatology 2017, 66, 1556–1569. [Google Scholar] [CrossRef] [PubMed]
  94. Lee, S.K.; Song, M.J.; Kim, S.H.; Park, M. Comparing Various Scoring System for Predicting Overall Survival According to Treatment Modalities in Hepatocellular Carcinoma Focused on Platelet-Albumin-Bilirubin (PALBI) and Albumin-Bilirubin (ALBI) Grade: A Nationwide Cohort Study. PLoS ONE 2019, 14, e0216173. [Google Scholar] [CrossRef] [PubMed]
  95. Simon, T.G.; Duberg, A.-S.; Aleman, S.; Chung, R.T.; Chan, A.T.; Ludvigsson, J.F. Association of Aspirin with Hepatocellular Carcinoma and Liver-Related Mortality. N. Engl. J. Med. 2020, 382, 1018–1028. [Google Scholar] [CrossRef]
  96. Zaldivar, M.M.; Pauels, K.; von Hundelshausen, P.; Berres, M.-L.; Schmitz, P.; Bornemann, J.; Kowalska, M.A.; Gassler, N.; Streetz, K.L.; Weiskirchen, R.; et al. CXC Chemokine Ligand 4 (Cxcl4) Is a Platelet-Derived Mediator of Experimental Liver Fibrosis. Hepatology 2010, 51, 1345–1353. [Google Scholar] [CrossRef] [PubMed]
  97. Carr, B.I.; Guerra, V. Thrombocytosis and Hepatocellular Carcinoma. Dig. Dis. Sci. 2013, 58, 1790–1796. [Google Scholar] [CrossRef]
  98. Carr, B.I.; Pancoska, P.; Giannini, E.G.; Farinati, F.; Ciccarese, F.; Ludovico Rapaccini, G.; Di Marco, M.; Benvegnù, L.; Zoli, M.; Borzio, F.; et al. Identification of Two Clinical Hepatocellular Carcinoma Patient Phenotypes from Results of Standard Screening Parameters. Semin. Oncol. 2014, 41, 406–414. [Google Scholar] [CrossRef]
  99. Pang, Q.; Qu, K.; Zhang, J.-Y.; Song, S.-D.; Liu, S.-S.; Tai, M.-H.; Liu, H.-C.; Liu, C. The Prognostic Value of Platelet Count in Patients with Hepatocellular Carcinoma. Medicine 2015, 94, e1431. [Google Scholar] [CrossRef]
  100. Maini, M.K.; Schurich, A. Platelets Harness the Immune Response to Drive Liver Cancer. Proc. Natl. Acad. Sci. USA 2012, 109, 12840–12841. [Google Scholar] [CrossRef]
  101. Pavlovic, N.; Rani, B.; Gerwins, P.; Heindryckx, F. Platelets as Key Factors in Hepatocellular Carcinoma. Cancers 2019, 11, 1022. [Google Scholar] [CrossRef] [PubMed]
  102. Iannacone, M.; Sitia, G.; Isogawa, M.; Marchese, P.; Castro, M.G.; Lowenstein, P.R.; Chisari, F.V.; Ruggeri, Z.M.; Guidotti, L.G. Platelets Mediate Cytotoxic T Lymphocyte–Induced Liver Damage. Nat. Med. 2005, 11, 1167–1169. [Google Scholar] [CrossRef] [PubMed]
  103. Sitia, G.; Aiolfi, R.; Di Lucia, P.; Mainetti, M.; Fiocchi, A.; Mingozzi, F.; Esposito, A.; Ruggeri, Z.M.; Chisari, F.V.; Iannacone, M.; et al. Antiplatelet Therapy Prevents Hepatocellular Carcinoma and Improves Survival in a Mouse Model of Chronic Hepatitis B. Proc. Natl. Acad. Sci. USA 2012, 109, E2165–E2172. [Google Scholar] [CrossRef]
  104. Yoshida, S.; Ikenaga, N.; Liu, S.B.; Peng, Z.-W.; Chung, J.; Sverdlov, D.Y.; Miyamoto, M.; Kim, Y.O.; Ogawa, S.; Arch, R.H.; et al. Extrahepatic Platelet-Derived Growth Factor-β, Delivered by Platelets, Promotes Activation of Hepatic Stellate Cells and Biliary Fibrosis in Mice. Gastroenterology 2014, 147, 1378–1392. [Google Scholar] [CrossRef]
  105. Gutiérrez-Herrero, S.; Maia, V.; Gutiérrez-Berzal, J.; Calzada, N.; Sanz, M.; González-Manchón, C.; Pericacho, M.; Ortiz-Rivero, S.; González-Porras, J.R.; Arechederra, M.; et al. C3G Transgenic Mouse Models with Specific Expression in Platelets Reveal a New Role for C3G in Platelet Clotting through Its GEF Activity. Biochim. Biophys. Acta Mol. Cell Res. 2012, 1823, 1366–1377. [Google Scholar] [CrossRef]
  106. Gutiérrez-Herrero, S.; Fernández-Infante, C.; Hernández-Cano, L.; Ortiz-Rivero, S.; Guijas, C.; Martín-Granado, V.; Ramón González-Porras, J.; Balsinde, J.; Porras, A.; Guerrero, C. C3G Contributes to Platelet Activation and Aggregation by Regulating Major Signaling Pathways. Signal Transduct. Target. Ther. 2020, 5, 29. [Google Scholar] [CrossRef] [PubMed]
  107. Lesurtel, M.; Graf, R.; Aleil, B.; Walther, D.J.; Tian, Y.; Jochum, W.; Gachet, C.; Bader, M.; Clavien, P.-A. Platelet-Derived Serotonin Mediates Liver Regeneration. Science 2006, 312, 104–107. [Google Scholar] [CrossRef] [PubMed]
  108. Murata, S.; Hashimoto, I.; Nakano, Y.; Myronovych, A.; Watanabe, M.; Ohkohchi, N. Single Administration of Thrombopoietin Prevents Progression of Liver Fibrosis and Promotes Liver Regeneration After Partial Hepatectomy in Cirrhotic Rats. Ann. Surg. 2008, 248, 821–828. [Google Scholar] [CrossRef]
  109. Watanabe, M.; Murata, S.; Hashimoto, I.; Nakano, Y.; Ikeda, O.; Aoyagi, Y.; Matsuo, R.; Fukunaga, K.; Yasue, H.; Ohkohchi, N. Platelets Contribute to the Reduction of Liver Fibrosis in Mice. J. Gastroenterol. Hepatol. 2009, 24, 78–89. [Google Scholar] [CrossRef]
  110. Takahashi, K. Human Platelets Inhibit Liver Fibrosis in Severe Combined Immunodeficiency Mice. World J. Gastroenterol. 2013, 19, 5250. [Google Scholar] [CrossRef]
  111. Ma, C.; Fu, Q.; Diggs, L.P.; McVey, J.C.; McCallen, J.; Wabitsch, S.; Ruf, B.; Brown, Z.; Heinrich, B.; Zhang, Q.; et al. Platelets Control Liver Tumor Growth through P2Y12-Dependent CD40L Release in NAFLD. Cancer Cell 2022, 40, 986–998.e5. [Google Scholar] [CrossRef]
  112. Kodama, T.; Takehara, T.; Hikita, H.; Shimizu, S.; Li, W.; Miyagi, T.; Hosui, A.; Tatsumi, T.; Ishida, H.; Tadokoro, S.; et al. Thrombocytopenia Exacerbates Cholestasis-Induced Liver Fibrosis in Mice. Gastroenterology 2010, 138, 2487–2498.e7. [Google Scholar] [CrossRef]
  113. Kuwahara, R.; Kofman, A.V.; Landis, C.S.; Swenson, E.S.; Barendswaard, E.; Theise, N.D. The Hepatic Stem Cell Niche: Identification by Label-Retaining Cell Assay. Hepatology 2008, 47, 1994–2002. [Google Scholar] [CrossRef]
  114. Wright, N.; Samuelson, L.; Walkup, M.H.; Chandrasekaran, P.; Gerber, D.A. Enrichment of a Bipotent Hepatic Progenitor Cell from Naïve Adult Liver Tissue. Biochem. Biophys. Res. Commun. 2008, 366, 367–372. [Google Scholar] [CrossRef]
  115. Zhang, L.; Theise, N.; Chua, M.; Reid, L.M. The Stem Cell Niche of Human Livers: Symmetry between Development and Regeneration. Hepatology 2008, 48, 1598–1607. [Google Scholar] [CrossRef]
  116. Rumgay, H.; Arnold, M.; Ferlay, J.; Lesi, O.; Cabasag, C.J.; Vignat, J.; Laversanne, M.; McGlynn, K.A.; Soerjomataram, I. Global Burden of Primary Liver Cancer in 2020 and Predictions to 2040. J. Hepatol. 2022, 77, 1598–1606. [Google Scholar] [CrossRef]
  117. Dasgupta, P.; Henshaw, C.; Youlden, D.R.; Clark, P.J.; Aitken, J.F.; Baade, P.D. Global Trends in Incidence Rates of Primary Adult Liver Cancers: A Systematic Review and Meta-Analysis. Front. Oncol. 2020, 10, 171. [Google Scholar] [CrossRef] [PubMed]
  118. Farazi, P.A.; De Pinho, R.A. Hepatocellular Carcinoma Pathogenesis: From Genes to Environment. Nat. Rev. Cancer 2006, 6, 674–687. [Google Scholar] [CrossRef] [PubMed]
  119. Schneller, D.; Angel, P. Cellular Origin of Hepatocellular Carcinoma. In Hepatocellular Carcinoma; Codon Publications: Brisbane, Australia, 2019; pp. 1–28. [Google Scholar] [CrossRef]
  120. Holczbauer, Á.; Wangensteen, K.J.; Shin, S. Cellular Origins of Regenerating Liver and Hepatocellular Carcinoma. JHEP Rep. 2022, 4, 100416. [Google Scholar] [CrossRef] [PubMed]
  121. Gromowski, T.; Lukacs-Kornek, V.; Cisowski, J. Current View of Liver Cancer Cell-of-Origin and Proposed Mechanisms Precluding Its Proper Determination. Cancer Cell Int. 2023, 23, 3. [Google Scholar] [CrossRef] [PubMed]
  122. Braun, L.; Goyette, M.; Yaswen, P.; Thompson, N.L.; Fausto4, N. Growth in Culture and Tumorigenicity after Transfection with the Ras Oncogene of Liver Epithelial Cells from Carcinogen-Treated Rats. Cancer Res. 1987, 47, 4116–4124. Available online: http://aacrjournals.org/cancerres/article-pdf/47/15/4116/2429364/cr0470154116.pdf (accessed on 29 September 2023). [PubMed]
  123. Strauss, R.P.; Audsley, K.M.; Passman, A.M.; van Vuuren, J.H.; Finch-Edmondson, M.L.; Callus, B.A.; Yeoh, G.C. Loss of ARF/INK4A Promotes Liver Progenitor Cell Transformation Toward Tumorigenicity Supporting Their Role in Hepatocarcinogenesis. Gene Expr. 2020, 20, 39–52. [Google Scholar] [CrossRef] [PubMed]
  124. Olivera-Salazar, R.; García-Arranz, M.; Sánchez, A.; Olmedillas-López, S.; Vega-Clemente, L.; Serrano, L.J.; Herrera, B.; García-Olmo, D. Oncological Transformation in Vitro of Hepatic Progenitor Cell Lines Isolated from Adult Mice. Sci. Rep. 2022, 12, 3149. [Google Scholar] [CrossRef]
  125. Zender, L.; Spector, M.S.; Xue, W.; Flemming, P.; Cordon-Cardo, C.; Silke, J.; Fan, S.-T.; Luk, J.M.; Wigler, M.; Hannon, G.J.; et al. Identification and Validation of Oncogenes in Liver Cancer Using an Integrative Oncogenomic Approach. Cell 2006, 125, 1253–1267. [Google Scholar] [CrossRef]
  126. Tang, D.; Chen, Y.; Fu, G.-B.; Yuan, T.-J.; Huang, W.-J.; Wang, Z.-Y.; Li, W.-J.; Jiao, Y.-F.; Yu, W.-F.; Yan, H.-X. EpCAM Inhibits Differentiation of Human Liver Progenitor Cells into Hepatocytes in Vitro by Activating Notch1 Signaling. Biochem. Biophys. Res. Commun. 2020, 525, 238–243. [Google Scholar] [CrossRef]
  127. Wu, K.; Ding, J.; Chen, C.; Sun, W.; Ning, B.-F.; Wen, W.; Huang, L.; Han, T.; Yang, W.; Wang, C.; et al. Hepatic Transforming Growth Factor Beta Gives Rise to Tumor-Initiating Cells and Promotes Liver Cancer Development. Hepatology 2012, 56, 2255–2267. [Google Scholar] [CrossRef] [PubMed]
  128. Almalé, L.; García-Álvaro, M.; Martínez-Palacián, A.; García-Bravo, M.; Lazcanoiturburu, N.; Addante, A.; Roncero, C.; Sanz, J.; López, M.; Bragado, P.; et al. C-Met Signaling Is Essential for Mouse Adult Liver Progenitor Cells Expansion After Transforming Growth Factor-β-Induced Epithelial–Mesenchymal Transition and Regulates Cell Phenotypic Switch. Stem Cells 2019, 37, 1108–1118. [Google Scholar] [CrossRef]
  129. Lu, L.; Li, Y.; Kim, S.M.; Bossuyt, W.; Liu, P.; Qiu, Q.; Wang, Y.; Halder, G.; Finegold, M.J.; Lee, J.-S.; et al. Hippo Signaling Is a Potent in Vivo Growth and Tumor Suppressor Pathway in the Mammalian Liver. Proc. Natl. Acad. Sci. USA 2010, 107, 1437–1442. [Google Scholar] [CrossRef]
  130. 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]
  131. Yamashita, T.; Forgues, M.; Wang, W.; Jin, W.K.; Ye, Q.; Jia, H.; Budhu, A.; Zanetti, K.A.; Chen, Y.; Qin, L.X.; et al. EpCAM and α-Fetoprotein Expression Defines Novel Prognostic Subtypes of Hepatocellular Carcinoma. Cancer Res. 2008, 68, 1451–1461. [Google Scholar] [CrossRef]
  132. Barthet, V.J.A.; Brucoli, M.; Ladds, M.J.G.W.; Nössing, C.; Kiourtis, C.; Baudot, A.D.; O’Prey, J.; Zunino, B.; Müller, M.; May, S.; et al. Autophagy Suppresses the Formation of Hepatocyte-Derived Cancer-Initiating Ductular Progenitor Cells in the Liver. Sci. Adv. 2021, 7, eabf9141. [Google Scholar] [CrossRef]
  133. Shin, S.; Wangensteen, K.J.; Teta-Bissett, M.; Wang, Y.J.; Mosleh-Shirazi, E.; Buza, E.L.; Greenbaum, L.E.; Kaestner, K.H. Genetic Lineage Tracing Analysis of the Cell of Origin of Hepatotoxin-induced Liver Tumors in Mice. Hepatology 2016, 64, 1163–1177. [Google Scholar] [CrossRef] [PubMed]
  134. Tummala, K.S.; Brandt, M.; Teijeiro, A.; Graña, O.; Schwabe, R.F.; Perna, C.; Djouder, N. Hepatocellular Carcinomas Originate Predominantly from Hepatocytes and Benign Lesions from Hepatic Progenitor Cells. Cell Rep. 2017, 19, 584–600. [Google Scholar] [CrossRef] [PubMed]
  135. Rosenberg, N.; Van Haele, M.; Lanton, T.; Brashi, N.; Bromberg, Z.; Adler, H.; Giladi, H.; Peled, A.; Goldenberg, D.S.; Axelrod, J.H.; et al. Combined Hepatocellular-Cholangiocarcinoma Derives from Liver Progenitor Cells and Depends on Senescence and IL-6 Trans-Signaling. J. Hepatol. 2022, 77, 1631–1641. [Google Scholar] [CrossRef]
  136. Nakamura, T.; Sakai, K.; Nakamura, T.; Matsumoto, K. Hepatocyte Growth Factor Twenty Years on: Much More than a Growth Factor. J. Gastroenterol. Hepatol. 2011, 26, 188–202. [Google Scholar] [CrossRef]
  137. Bladt, F.; Riethmacher, D.; Isenmann, S.; Aguzzi, A.; Birchmeier, C. Essential Role for the C-Met Receptor in the Migration of Myogenic Precursor Cells into the Limb Bud. Nature 1995, 376, 768–771. [Google Scholar] [CrossRef] [PubMed]
  138. Zhao, Y.; Ye, W.; Wang, Y.D.; Chen, W.D. HGF/c-Met: A Key Promoter in Liver Regeneration. Front. Pharmacol. 2022, 13, 808855. [Google Scholar] [CrossRef]
  139. Factor, V.M.; Seo, D.; Ishikawa, T.; Kaposi-Novak, P.; Marquardt, J.U.; Andersen, J.B.; Conner, E.A.; Thorgeirsson, S.S. Loss of C-Met Disrupts Gene Expression Program Required for G2/M Progression during Liver Regeneration in Mice. PLoS ONE 2010, 5, e12739. [Google Scholar] [CrossRef]
  140. Huh, C.-G.; Factor, V.M.; Sánchez, A.; Uchida, K.; Conner, E.A.; Thorgeirsson, S.S. Hepatocyte Growth Factorc-Met Signaling Pathway Is Required for Efficient Liver Regeneration and Repair. Proc. Natl. Acad. Sci. USA 2004, 101, 4477–4482. [Google Scholar] [CrossRef]
  141. Shiota, G.; Kunisada, T.; Oyama, K.; Udagawa, A.; Nomi, T.; Tanaka, K.; Tsutsumi, A.; Isono, M.; Nakamura, T.; Hamada, H.; et al. In Vivo Transfer of Hepatocyte Growth Factor Gene Accelerates Proliferation of Hepatic Oval Cells in a 2-acetylaminofluorene/Partial Hepatectomy Model in Rats. FEBS Lett. 2000, 470, 325–330. [Google Scholar] [CrossRef]
  142. Ishikawa, T.; Factor, V.M.; Marquardt, J.U.; Raggi, C.; Seo, D.; Kitade, M.; Conner, E.A.; Thorgeirsson, S.S. Hepatocyte Growth Factor/c-Met Signaling Is Required for Stem-Cell-Mediated Liver Regeneration in Mice. Hepatology 2012, 55, 1215–1226. [Google Scholar] [CrossRef]
  143. Giordano, S.; Columbano, A. Met as a Therapeutic Target in HCC: Facts and Hopes. J. Hepatol. 2014, 60, 442–452. [Google Scholar] [CrossRef]
  144. Salvi, A.; Arici, B.; Portolani, N.; Giulini, S.M.; De Petro, G.; Barlati, S. In Vitro C-Met Inhibition by Antisense RNA and Plasmid-Based RNAi down-Modulates Migration and Invasion of Hepatocellular Carcinoma Cells. Int. J. Oncol. 2007, 31, 451–460. [Google Scholar] [CrossRef]
  145. Xie, B.; Xing, R.; Chen, P.; Gou, Y.; Li, S.; Xiao, J.; Dong, J. Down-Regulation of c-Met Expression Inhibits Human HCC Cells Growth and Invasion by RNA Interference. J. Surg. Res. 2010, 162, 231–238. [Google Scholar] [CrossRef] [PubMed]
  146. Ogunwobi, O.O.; Liu, C. Hepatocyte Growth Factor Upregulation Promotes Carcinogenesis and Epithelial-Mesenchymal Transition in Hepatocellular Carcinoma via Akt and COX-2 Pathways. Clin. Exp. Metastasis 2011, 28, 721–731. [Google Scholar] [CrossRef] [PubMed]
  147. You, H.; Ding, W.; Dang, H.; Jiang, Y.; Rountree, C.B. C-Met Represents a Potential Therapeutic Target for Personalized Treatment in Hepatocellular Carcinoma. Hepatology 2011, 54, 879–889. [Google Scholar] [CrossRef]
  148. Fan, Y.; Arechederra, M.; Richelme, S.; Daian, F.; Novello, C.; Calderaro, J.; Di Tommaso, L.; Morcrette, G.; Rebouissou, S.; Donadon, M.; et al. A Phosphokinome-based Screen Uncovers New Drug Synergies for Cancer Driven by Liver-specific Gain of Nononcogenic Receptor Tyrosine Kinases. Hepatology 2017, 66, 1644–1661. [Google Scholar] [CrossRef] [PubMed]
  149. Wang, R.; Ferrell, L.D.; Faouzi, S.; Maher, J.J.; Bishop, J.M. Activation of the Met Receptor by Cell Attachment Induces and Sustains Hepatocellular Carcinomas in Transgenic Mice. J. Cell Biol. 2001, 153, 1023–1034. [Google Scholar] [CrossRef] [PubMed]
  150. Takami, T.; Kaposi-Novak, P.; Uchida, K.; Gomez-Quiroz, L.E.; Conner, E.A.; Factor, V.M.; Thorgeirsson, S.S. Loss of Hepatocyte Growth Factor/c-Met Signaling Pathway Accelerates Early Stages of N-nitrosodiethylamine–Induced Hepatocarcinogenesis. Cancer Res. 2007, 67, 9844–9851. [Google Scholar] [CrossRef]
  151. Marx-Stoelting, P.; Borowiak, M.; Knorpp, T.; Birchmeier, C.; Buchmann, A.; Schwarz, M. Hepatocarcinogenesis in Mice with a Conditional Knockout of the Hepatocyte Growth Factor Receptor C-Met. Int. J. Cancer 2009, 124, 1767–1772. [Google Scholar] [CrossRef]
  152. Kaposi-Novak, P. Met-Regulated Expression Signature Defines a Subset of Human Hepatocellular Carcinomas with Poor Prognosis and Aggressive Phenotype. J. Clin. Investig. 2006, 116, 1582–1595. [Google Scholar] [CrossRef]
  153. Yu, J.; Chen, G.G.; Lai, P.B.S. Targeting Hepatocyte Growth Factor/C-mesenchymal–Epithelial Transition Factor Axis in Hepatocellular Carcinoma: Rationale and Therapeutic Strategies. Med. Res. Rev. 2021, 41, 507–524. [Google Scholar] [CrossRef]
  154. Wang, Q.; Xu, L.; Wu, Q.; Zhang, M.; Zhang, J. Association between the Risk of Hepatitis Virus-Related Hepatocellular Carcinoma and EGF Polymorphism: A PRISMA-Compliant Updated Meta-Analysis. Medicine 2022, 101, e31280. [Google Scholar] [CrossRef] [PubMed]
  155. Arechederra, M.; Daian, F.; Yim, A.; Bazai, S.K.; Richelme, S.; Dono, R.; Saurin, A.J.; Habermann, B.H.; Maina, F. Hypermethylation of Gene Body CpG Islands Predicts High Dosage of Functional Oncogenes in Liver Cancer. Nat. Commun. 2018, 9, 3164. [Google Scholar] [CrossRef] [PubMed]
  156. Bouattour, M.; Raymond, E.; Qin, S.; Cheng, A.; Stammberger, U.; Locatelli, G.; Faivre, S. Recent Developments of C-Met as a Therapeutic Target in Hepatocellular Carcinoma. Hepatology 2018, 67, 1132–1149. [Google Scholar] [CrossRef] [PubMed]
  157. Qin, S.; Chan, S.L.; Sukeepaisarnjaroen, W.; Han, G.; Choo, S.P.; Sriuranpong, V.; Pan, H.; Yau, T.; Guo, Y.; Chen, M.; et al. A Phase II Study of the Efficacy and Safety of the MET Inhibitor Capmatinib (INC280) in Patients with Advanced Hepatocellular Carcinoma. Ther. Adv. Med. Oncol. 2019, 11, 175883591988900. [Google Scholar] [CrossRef] [PubMed]
  158. Zhao, S.; Wu, W.; Jiang, H.; Ma, L.; Pan, C.; Jin, C.; Mo, J.; Wang, L.; Wang, K. Selective Inhibitor of the C-Met Receptor Tyrosine Kinase in Advanced Hepatocellular Carcinoma: No Beneficial Effect with the Use of Tivantinib? Front. Immunol. 2021, 12, 731527. [Google Scholar] [CrossRef] [PubMed]
  159. Eisai Inc. E7050 in Combination with Sorafenib Versus Sorafenib Alone as First Line Therapy in Participants with Hepatocellular Carcinoma. NIH, US National library of Medicine. Available online: https://clinicaltrials.gov/study/NCT01271504 (accessed on 28 September 2023).
  160. Vogel, A.; Martinelli, E.; Vogel, A.; Cervantes, A.; Chau, I.; Daniele, B.; Llovet, J.M.; Meyer, T.; Nault, J.-C.; Neumann, U.; et al. Updated Treatment Recommendations for Hepatocellular Carcinoma (HCC) from the ESMO Clinical Practice Guidelines. Ann. Oncol. 2021, 32, 801–805. [Google Scholar] [CrossRef] [PubMed]
  161. Llovet, J.M.; Ricci, S.; Mazzaferro, V.; Hilgard, P.; Gane, E.; Blanc, J.-F.; de Oliveira, A.C.; Santoro, A.; Raoul, J.-L.; Forner, A.; et al. Sorafenib in Advanced Hepatocellular Carcinoma. N. Engl. J. Med. 2008, 359, 378–390. [Google Scholar] [CrossRef]
  162. Chen, J.; Jin, R.; Zhao, J.; Liu, J.; Ying, H.; Yan, H.; Zhou, S.; Liang, Y.; Huang, D.; Liang, X.; et al. Potential Molecular, Cellular and Microenvironmental Mechanism of Sorafenib Resistance in Hepatocellular Carcinoma. Cancer Lett. 2015, 367, 1–11. [Google Scholar] [CrossRef]
  163. Sabbah, D.A.; Hajjo, R.; Sweidan, K. Review on Epidermal Growth Factor Receptor (EGFR) Structure, Signaling Pathways, Interactions, and Recent Updates of EGFR Inhibitors. Curr. Top. Med. Chem. 2020, 20, 815–834. [Google Scholar] [CrossRef]
  164. Burgess, A.W. Regulation of Signaling from the Epidermal Growth Factor Family. J. Phys. Chem. B 2022, 126, 7475–7485. [Google Scholar] [CrossRef]
  165. Bhushan, B.; Michalopoulos, G.K. Role of Epidermal Growth Factor Receptor in Liver Injury and Lipid Metabolism: Emerging New Roles for an Old Receptor. Chem. Biol. Interact. 2020, 324, 109090. [Google Scholar] [CrossRef]
  166. Xu, H.; Liu, L.; Cong, M.; Liu, T.; Sun, S.; Ma, H.; You, H.; Jia, J.; Wang, P. EGF Neutralization Antibodies Attenuate Liver Fibrosis by Inhibiting Myofibroblast Proliferation in Bile Duct Ligation Mice. Histochem. Cell Biol. 2020, 154, 107–116. [Google Scholar] [CrossRef]
  167. Perugorria, M.J.; Latasa, M.U.; Nicou, A.; Cartagena-Lirola, H.; Castillo, J.; Goñi, S.; Vespasiani-Gentilucci, U.; Zagami, M.G.; Lotersztajn, S.; Prieto, J.; et al. The Epidermal Growth Factor Receptor Ligand Amphiregulin Participates in the Development of Mouse Liver Fibrosis. Hepatology 2008, 48, 1251–1261. [Google Scholar] [CrossRef] [PubMed]
  168. Scheving, L.A.; Zhang, X.; Threadgill, D.W.; Russell, W.E. Hepatocyte ERBB3 and EGFR Are Required for Maximal CCl4-Induced Liver Fibrosis. Am. J. Physiol.-Gastrointest. Liver Physiol. 2016, 311, G807–G816. [Google Scholar] [CrossRef] [PubMed]
  169. McKee, C.; Sigala, B.; Soeda, J.; Mouralidarane, A.; Morgan, M.; Mazzoccoli, G.; Rappa, F.; Cappello, F.; Cabibi, D.; Pazienza, V.; et al. Amphiregulin Activates Human Hepatic Stellate Cells and Is Upregulated in Non Alcoholic Steatohepatitis. Sci. Rep. 2015, 5, 8812. [Google Scholar] [CrossRef] [PubMed]
  170. Schiffer, E.; Housset, C.; Cacheux, W.; Wendum, D.; Desbois-Mouthon, C.; Rey, C.; Clergue, F.; Poupon, R.; Barbu, V.; Rosmorduc, O. Gefitinib, an EGFR Inhibitor, Prevents Hepatocellular Carcinoma Development in the Rat Liver with Cirrhosis. Hepatology 2005, 41, 307–314. [Google Scholar] [CrossRef] [PubMed]
  171. Fuchs, B.C.; Hoshida, Y.; Fujii, T.; Wei, L.; Yamada, S.; Lauwers, G.Y.; McGinn, C.M.; DePeralta, D.K.; Chen, X.; Kuroda, T.; et al. Epidermal Growth Factor Receptor Inhibition Attenuates Liver Fibrosis and Development of Hepatocellular Carcinoma. Hepatology 2014, 59, 1577–1590. [Google Scholar] [CrossRef] [PubMed]
  172. López-Luque, J.; Caballero-Díaz, D.; Martinez-Palacián, A.; Roncero, C.; Moreno-Càceres, J.; García-Bravo, M.; Grueso, E.; Fernández, A.; Crosas-Molist, E.; García-Álvaro, M.; et al. Dissecting the Role of Epidermal Growth Factor Receptor Catalytic Activity during Liver Regeneration and Hepatocarcinogenesis. Hepatology 2016, 63, 604–619. [Google Scholar] [CrossRef] [PubMed]
  173. Lanaya, H.; Natarajan, A.; Komposch, K.; Li, L.; Amberg, N.; Chen, L.; Wculek, S.K.; Hammer, M.; Zenz, R.; Peck-Radosavljevic, M.; et al. EGFR Has a Tumour-Promoting Role in Liver Macrophages during Hepatocellular Carcinoma Formation. Nat. Cell Biol. 2014, 16, 972–981. [Google Scholar] [CrossRef]
  174. Choung, S.; Kim, J.M.; Joung, K.H.; Lee, E.S.; Kim, H.J.; Ku, B.J. Epidermal Growth Factor Receptor Inhibition Attenuates Non-Alcoholic Fatty Liver Disease in Diet-Induced Obese Mice. PLoS ONE 2019, 14, e0210828. [Google Scholar] [CrossRef] [PubMed]
  175. Qian, Y.; Han, J.; Zhou, L.; Yu, Q.; Xu, J.; Jin, Z.; Yang, Y.; Jiang, L.; Lou, D. Inhibition of Epidermal Growth Factor Receptor (EGFR) Reduces Lipopolysaccharide (LPS)-Induced Activation and Inflammatory Cytokines in Hepatic Stellate Cells In Vitro. Med. Sci. Monit. 2018, 24, 5533–5541. [Google Scholar] [CrossRef] [PubMed]
  176. Villanueva, A.; Toffanin, S.; Llovet, J.M. Linking Molecular Classification of Hepatocellular Carcinoma and Personalized Medicine: Preliminary Steps. Curr. Opin. Oncol. 2008, 20, 444–453. [Google Scholar] [CrossRef] [PubMed]
  177. Ito, Y.; Takeda, T.; Sakon, M.; Tsujimoto, M.; Higashiyama, S.; Noda, K.; Miyoshi, E.; Monden, M.; Matsuura, N. Expression and Clinical Significance of Erb-B Receptor Family in Hepatocellular Carcinoma. Br. J. Cancer 2001, 84, 1377–1383. [Google Scholar] [CrossRef] [PubMed]
  178. Furuse, J. Growth Factors as Therapeutic Targets in HCC. Crit. Rev. Oncol. Hematol. 2008, 67, 8–15. [Google Scholar] [CrossRef] [PubMed]
  179. Su, M.-C.; Lien, H.-C.; Jeng, Y.-M. Absence of Epidermal Growth Factor Receptor Exon 18–21 Mutation in Hepatocellular Carcinoma. Cancer Lett. 2005, 224, 117–121. [Google Scholar] [CrossRef]
  180. Ally, A.; Balasundaram, M.; Carlsen, R.; Chuah, E.; Clarke, A.; Dhalla, N.; Holt, R.A.; Jones, S.J.M.; Lee, D.; Ma, Y.; et al. Comprehensive and Integrative Genomic Characterization of Hepatocellular Carcinoma. Cell 2017, 169, 1327–1341.e23. [Google Scholar] [CrossRef]
  181. Liu, Z.; Chen, D.; Ning, F.; Du, J.; Wang, H. EGF Is Highly Expressed in Hepatocellular Carcinoma (HCC) and Promotes Motility of HCC Cells via Fibronectin. J. Cell Biochem. 2018, 119, 4170–4183. [Google Scholar] [CrossRef]
  182. Berasain, C.; Avila, M.A. The EGFR Signalling System in the Liver: From Hepatoprotection to Hepatocarcinogenesis. J. Gastroenterol. 2014, 49, 9–23. [Google Scholar] [CrossRef]
  183. Qu, X.; Zhang, L.; Li, S.; Li, T.; Zhao, X.; Wang, N.; Shi, Y. M6A-Related Angiogenic Genes to Construct Prognostic Signature, Reveal Immune and Oxidative Stress Landscape, and Screen Drugs in Hepatocellular Carcinoma. Oxid. Med. Cell Longev. 2022, 2022, 8301888. [Google Scholar] [CrossRef]
  184. Breuhahn, K.; Longerich, T.; Schirmacher, P. Dysregulation of Growth Factor Signaling in Human Hepatocellular Carcinoma. Oncogene 2006, 25, 3787–3800. [Google Scholar] [CrossRef] [PubMed]
  185. 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]
  186. National Cancer Institute (NCI). Bevacizumab and Erlotinib Hydrochloride in Treating Patients with Metastatic or Unresectable Biliary Tumors. Available online: https://clinicaltrials.gov/study/NCT00356889 (accessed on 28 September 2023).
  187. National Cancer Institute (NCI). Bevacizumab and Erlotinib in Treating Patients with Advanced Liver Cancer. Available online: https://clinicaltrials.gov/study/NCT00365391 (accessed on 28 September 2023).
  188. National Cancer Institute (NCI). Erlotinib Hydrochloride in Preventing Liver Cancer in Patients with Cirrhosis of the Liver. Available online: https://clinicaltrials.gov/study/NCT02273362 (accessed on 28 September 2023).
  189. Philip, P.A.; Mahoney, M.R.; Holen, K.D.; Northfelt, D.W.; Pitot, H.C.; Picus, J.; Flynn, P.J.; Erlichman, C. Phase 2 Study of Bevacizumab plus Erlotinib in Patients with Advanced Hepatocellular Cancer. Cancer 2012, 118, 2424–2430. [Google Scholar] [CrossRef] [PubMed]
  190. National Cancer Institute (NCI). Sorafenib Tosylate and Erlotinib Hydrochloride in Treating Patients with Locally Advanced, Unresectable, or Metastatic Gallbladder Cancer or Cholangiocarcinoma. Available online: https://clinicaltrials.gov/study/NCT01093222 (accessed on 28 September 2023).
  191. Bekaii-Saab, T.; Markowitz, J.; Prescott, N.; Sadee, W.; Heerema, N.; Wei, L.; Dai, Z.; Papp, A.; Campbell, A.; Culler, K.; et al. A Multi-Institutional Phase II Study of the Efficacy and Tolerability of Lapatinib in Patients with Advanced Hepatocellular Carcinomas. Clin. Cancer Res. 2009, 15, 5895–5901. [Google Scholar] [CrossRef] [PubMed]
  192. Sohal, D.P.S.; Mykulowycz, K.; Uehara, T.; Teitelbaum, U.R.; Damjanov, N.; Giantonio, B.J.; Carberry, M.; Wissel, P.; Jacobs-Small, M.; O’Dwyer, P.J.; et al. A Phase II Trial of Gemcitabine, Irinotecan and Panitumumab in Advanced Cholangiocarcinoma. Ann. Oncol. 2013, 24, 3061–3065. [Google Scholar] [CrossRef] [PubMed]
  193. Berasain, C.; Perugorria, M.J.; Latasa, M.U.; Castillo, J.; Goñi, S.; Santamaría, M.; Prieto, J.; Avila, M.A. The Epidermal Growth Factor Receptor: A Link Between Inflammation and Liver Cancer. Exp. Biol. Med. 2009, 234, 713–725. [Google Scholar] [CrossRef]
  194. Shaker, M.E.; Gomaa, H.A.M.; Abdelgawad, M.A.; El-Mesery, M.; Shaaban, A.A.; Hazem, S.H. Emerging Roles of Tyrosine Kinases in Hepatic Inflammatory Diseases and Therapeutic Opportunities. Int. Immunopharmacol. 2023, 120, 110373. [Google Scholar] [CrossRef] [PubMed]
  195. David, C.J.; Massagué, J. Contextual Determinants of TGFβ Action in Development, Immunity and Cancer. Nat. Rev. Mol. Cell Biol. 2018, 19, 419–435. [Google Scholar] [CrossRef]
  196. Coulouarn, C.; Factor, V.M.; Thorgeirsson, S.S. Transforming Growth Factor-β Gene Expression Signature in Mouse Hepatocytes Predicts Clinical Outcome in Human Cancer. Hepatology 2008, 47, 2059–2067. [Google Scholar] [CrossRef]
  197. Dooley, S.; ten Dijke, P. TGF-β in Progression of Liver Disease. Cell Tissue Res. 2012, 347, 245–256. [Google Scholar] [CrossRef]
  198. Fabregat, I.; Moreno-Càceres, J.; Sánchez, A.; Dooley, S.; Dewidar, B.; Giannelli, G.; ten Dijke, P. TGF-β Signalling and Liver Disease. FEBS J. 2016, 283, 2219–2232. [Google Scholar] [CrossRef] [PubMed]
  199. Dong, Z.-Z.; Yao, D.-F.; Yao, M.; Qiu, L.-W.; Zong, L.; Wu, W.; Wu, X.-H.; Yao, D.-B.; Meng, X.-Y. Clinical Impact of Plasma TGF-Beta1 and Circulating TGF-Beta1 MRNA in Diagnosis of Hepatocellular Carcinoma. Hepatobiliary Pancreat. Dis. Int. 2008, 7, 288–295. [Google Scholar]
  200. Devan, A.R.; Pavithran, K.; Nair, B.; Murali, M.; Nath, L.R. Deciphering the Role of Transforming Growth Factor-Beta 1 as a Diagnostic-Prognostic-Therapeutic Candidate against Hepatocellular Carcinoma. World J. Gastroenterol. 2022, 28, 5250–5264. [Google Scholar] [CrossRef]
  201. Chen, J.; Zaidi, S.; Rao, S.; Chen, J.-S.; Phan, L.; Farci, P.; Su, X.; Shetty, K.; White, J.; Zamboni, F.; et al. Analysis of Genomes and Transcriptomes of Hepatocellular Carcinomas Identifies Mutations and Gene Expression Changes in the Transforming Growth Factor-β Pathway. Gastroenterology 2018, 154, 195–210. [Google Scholar] [CrossRef]
  202. Meyer, C.; Liu, Y.; Dooley, S. Caveolin and TGF-β Entanglements. J. Cell Physiol. 2013, 228, 2097–2102. [Google Scholar] [CrossRef] [PubMed]
  203. Moreno-Càceres, J.; Caballero-Díaz, D.; Nwosu, Z.C.; Meyer, C.; López-Luque, J.; Malfettone, A.; Lastra, R.; Serrano, T.; Ramos, E.; Dooley, S.; et al. The Level of Caveolin-1 Expression Determines Response to TGF-β as a Tumour Suppressor in Hepatocellular Carcinoma Cells. Cell Death Dis. 2017, 8, e3098. [Google Scholar] [CrossRef] [PubMed]
  204. Gungor, M.Z.; Uysal, M.; Senturk, S. The Bright and the Dark Side of TGF-β Signaling in Hepatocellular Carcinoma: Mechanisms, Dysregulation, and Therapeutic Implications. Cancers 2022, 14, 940. [Google Scholar] [CrossRef]
  205. Liu, A.M.; Xu, Z.; Shek, F.H.; Wong, K.F.; Lee, N.P.; Poon, R.T.; Chen, J.; Luk, J.M. MiR-122 Targets Pyruvate Kinase M2 and Affects Metabolism of Hepatocellular Carcinoma. PLoS ONE 2014, 9, e86872. [Google Scholar] [CrossRef]
  206. Huang, S.; He, X.; Ding, J.; Liang, L.; Zhao, Y.; Zhang, Z.; Yao, X.; Pan, Z.; Zhang, P.; Li, J.; et al. Upregulation of MiR-23a∼27a∼24 Decreases Transforming Growth Factor-Beta-Induced Tumor-Suppressive Activities in Human Hepatocellular Carcinoma Cells. Int. J. Cancer 2008, 123, 972–978. [Google Scholar] [CrossRef]
  207. Murillo, M.M.; del Castillo, G.; Sánchez, A.; Fernández, M.; Fabregat, I. Involvement of EGF Receptor and C-Src in the Survival Signals Induced by TGF-Β1 in Hepatocytes. Oncogene 2005, 24, 4580–4587. [Google Scholar] [CrossRef]
  208. Caja, L.; Bertran, E.; Campbell, J.; Fausto, N.; Fabregat, I. The Transforming Growth Factor-Beta (TGF-β) Mediates Acquisition of a Mesenchymal Stem Cell-like Phenotype in Human Liver Cells. J. Cell Physiol. 2011, 226, 1214–1223. [Google Scholar] [CrossRef]
  209. Fischer, A.N.M.; Fuchs, E.; Mikula, M.; Huber, H.; Beug, H.; Mikulits, W. PDGF Essentially Links TGF-β Signaling to Nuclear β-Catenin Accumulation in Hepatocellular Carcinoma Progression. Oncogene 2007, 26, 3395–3405. [Google Scholar] [CrossRef] [PubMed]
  210. Gonzalez-Sanchez, E.; Vaquero, J.; Férnandez-Barrena, M.G.; Lasarte, J.J.; Avila, M.A.; Sarobe, P.; Reig, M.; Calvo, M.; Fabregat, I. The TGF-β Pathway: A Pharmacological Target in Hepatocellular Carcinoma? Cancers 2021, 13, 3248. [Google Scholar] [CrossRef] [PubMed]
  211. Song, M.; He, J.; Pan, Q.; Yang, J.; Zhao, J.; Zhang, Y.; Huang, Y.; Tang, Y.; Wang, Q.; He, J.; et al. Cancer-Associated Fibroblast-Mediated Cellular Crosstalk Supports Hepatocellular Carcinoma Progression. Hepatology 2021, 73, 1717–1735. [Google Scholar] [CrossRef] [PubMed]
  212. Benetti, A.; Berenzi, A.; Gambarotti, M.; Garrafa, E.; Gelati, M.; Dessy, E.; Portolani, N.; Piardi, T.; Giulini, S.M.; Caruso, A.; et al. Transforming Growth Factor-Β1 and CD105 Promote the Migration of Hepatocellular Carcinoma–Derived Endothelium. Cancer Res. 2008, 68, 8626–8634. [Google Scholar] [CrossRef] [PubMed]
  213. Jiang, F.; Wang, X.; Liu, Q.; Shen, J.; Li, Z.; Li, Y.; Zhang, J. Inhibition of TGF-β/SMAD3/NF-ΚB Signaling by MicroRNA-491 Is Involved in Arsenic Trioxide-Induced Anti-Angiogenesis in Hepatocellular Carcinoma Cells. Toxicol. Lett. 2014, 231, 55–61. [Google Scholar] [CrossRef]
  214. Batlle, E.; Massagué, J. Transforming Growth Factor-β Signaling in Immunity and Cancer. Immunity 2019, 50, 924–940. [Google Scholar] [CrossRef] [PubMed]
  215. Holmgaard, R.B.; Schaer, D.A.; Li, Y.; Castaneda, S.P.; Murphy, M.Y.; Xu, X.; Inigo, I.; Dobkin, J.; Manro, J.R.; Iversen, P.W.; et al. Targeting the TGFβ Pathway with Galunisertib, a TGFβRI Small Molecule Inhibitor, Promotes Anti-Tumor Immunity Leading to Durable, Complete Responses, as Monotherapy and in Combination with Checkpoint Blockade. J. Immunother. Cancer 2018, 6, 47. [Google Scholar] [CrossRef] [PubMed]
  216. Eli Lilly and Company. A Study of LY2157299 in Participants with Advanced Hepatocellular Carcinoma. Available online: https://clinicaltrials.gov/study/NCT02178358 (accessed on 28 September 2023).
  217. Eli Lilly and Company. A Study of Galunisertib (LY2157299) in Combination with Nivolumab in Advanced Refractory Solid Tumors and in Recurrent or Refractory NSCLC, or Hepatocellular Carcinoma. Available online: https://clinicaltrials.gov/study/NCT02423343 (accessed on 28 September 2023).
  218. Eli Lilly and Company. A Study of LY2157299 in Participants with Hepatocellular Carcinoma. Available online: https://clinicaltrials.gov/study/NCT01246986 (accessed on 28 September 2023).
  219. Giannelli, G.; Santoro, A.; Kelley, R.K.; Gane, E.; Paradis, V.; Cleverly, A.; Smith, C.; Estrem, S.T.; Man, M.; Wang, S.; et al. Biomarkers and Overall Survival in Patients with Advanced Hepatocellular Carcinoma Treated with TGF-ΒRI Inhibitor Galunisertib. PLoS ONE 2020, 15, e0222259. [Google Scholar] [CrossRef]
  220. National Cancer Institute (NCI). Lapatinib Ditosylate in Treating Patients with Unresectable Liver or Biliary Tract Cancer. Available online: https://clinicaltrials.gov/study/NCT00107536 (accessed on 30 September 2023).
  221. Genzyme, a Sanofi Company. Vandetanib Gemcitabine or Placebo Plus Gemcitabine or Vandetanib Monotherapy in Advanced Biliary Tract Cancer (VANGOGH). Available online: https://clinicaltrials.gov/study/NCT00753675 (accessed on 29 September 2023).
  222. Abramson Cancer Center at Penn Medicine. Study of Gemcitabine, Irinotecan and Panitumumab in Patients with Advanced and Metastatic Biliary Tract Adenocarcinoma. Available online: https://clinicaltrials.gov/study/NCT00948935 (accessed on 29 September 2023).
  223. Faivre, S.; Santoro, A.; Kelley, R.K.; Gane, E.; Costentin, C.E.; Gueorguieva, I.; Smith, C.; Cleverly, A.; Lahn, M.M.; Raymond, E.; et al. Novel Transforming Growth Factor Beta Receptor I Kinase Inhibitor Galunisertib (LY2157299) in Advanced Hepatocellular Carcinoma. Liver Int. 2019, 39, 1468–1477. [Google Scholar] [CrossRef]
  224. Gotoh, T.; Hattori, S.; Nakamura, S.; Kitayama, H.; Noda, M.; Takai, Y.; Kaibuchi, K.; Matsui, H.; Hatase, O.; Takahashi, H.; et al. Identification of Rap1 as a Target for the Crk SH3 Domain-Binding Guanine Nucleotide-Releasing Factor C3G. Mol. Cell. Biol. 1995, 15, 6746–6753. Available online: http://mcb.asm.org/ (accessed on 29 September 2023). [CrossRef] [PubMed]
  225. Guerrero, C.; Fernandez-Medarde, A.; Rojas, J.; Font de Mora, J.; Esteban, L.; Santos, E. Transformation Suppressor Activity of C3G Is Independent of Its CDC25-Homology Domain. Oncogene 1998, 16, 613–624. [Google Scholar] [CrossRef]
  226. Priego, N.; Arechederra, M.; Sequera, C.; Bragado, P.; Vázquez-Carballo, A.; Gutiérrez-Uzquiza, Á.; Martín-Granado, V.; Ventura, J.J.; Kazanietz, M.G.; Guerrero, C.; et al. C3G Knock-down Enhances Migration and Invasion by Increasing Rap1-Mediated P38α Activation, While It Impairs Tumor Growth through P38α-Independent Mechanisms. Oncotarget 2016, 7, 45060–45078. [Google Scholar] [CrossRef] [PubMed]
  227. Sequera, C.; Bragado, P.; Manzano, S.; Arechederra, M.; Richelme, S.; Gutiérrez-Uzquiza, A.; Sánchez, A.; Maina, F.; Guerrero, C.; Porras, A. C3G Is Upregulated in Hepatocarcinoma, Contributing to Tumor Growth and Progression and to HGF/MET Pathway Activation. Cancers 2020, 12, 2282. [Google Scholar] [CrossRef] [PubMed]
  228. Manzano, S.; Gutierrez-uzquiza, A.; Bragado, P.; Cuesta, A.M.; Guerrero, C.; Porras, A. C3G Protein, a New Player in Glioblastoma. Int. J. Mol. Sci. 2021, 22, 10018. [Google Scholar] [CrossRef]
  229. Manzano, S.; Gutierrez-Uzquiza, A.; Bragado, P.; Sequera, C.; Herranz, Ó.; Rodrigo-Faus, M.; Jauregui, P.; Morgner, S.; Rubio, I.; Guerrero, C.; et al. C3G Downregulation Induces the Acquisition of a Mesenchymal Phenotype That Enhances Aggressiveness of Glioblastoma Cells. Cell Death Dis. 2021, 12, 348. [Google Scholar] [CrossRef]
  230. Gao, J.; Aksoy, B.A.; Dogrusoz, U.; Dresdner, G.; Gross, B.; Sumer, S.O.; Sun, Y.; Jacobsen, A.; Sinha, R.; Larsson, E.; et al. Integrative Analysis of Complex Cancer Genomics and Clinical Profiles Using the CBioPortal. Sci. Signal. 2013, 6, pl1. [Google Scholar] [CrossRef]
  231. Cerami, E.; Gao, J.; Dogrusoz, U.; Gross, B.E.; Sumer, S.O.; Aksoy, B.A.; Jacobsen, A.; Byrne, C.J.; Heuer, M.L.; Larsson, E.; et al. The CBio Cancer Genomics Portal: An Open Platform for Exploring Multidimensional Cancer Genomics Data. Cancer Discov. 2012, 2, 401–404. [Google Scholar] [CrossRef]
  232. Sequera, C.; Manzano, S.; Guerrero, C.; Porras, A. How Rap and Its GEFs Control Liver Physiology and Cancer Development. C3G Alterations in Human Hepatocarcinoma. Hepatic Oncol. 2018, 5, HEP05. [Google Scholar] [CrossRef]
  233. The Results Published Here Are in Whole or Part Based Upon Data Generated by the TCGA Research Network. Available online: https://www.cancer.gov/tcga (accessed on 25 October 2023).
  234. Yang, J.; Guo, Z.; Song, M.; Pan, Q.; Zhao, J.; Huang, Y.; Han, Y.; Ouyang, D.; Yang, C.; Chen, H.; et al. Lenvatinib Improves Anti-PD-1 Therapeutic Efficacy by Promoting Vascular Normalization via the NRP-1-PDGFRβ Complex in Hepatocellular Carcinoma. Front. Immunol. 2023, 14, 1212577. [Google Scholar] [CrossRef] [PubMed]
  235. Palao, N.; Sequera, C.; Cuesta, Á.M.; Baquero, C.; Bragado, P.; Gutierrez-Uzquiza, A.; Sánchez, A.; Guerrero, C.; Porras, A. C3G Down-Regulation Enhances pro-Migratory and Stemness Properties of Oval Cells by Promoting an Epithelial-Mesenchymal-like Process. Int. J. Biol. Sci. 2022, 18, 5873–5884. [Google Scholar] [CrossRef] [PubMed]
  236. Brown, Z.J.; Heinrich, B.; Greten, T.F. Mouse Models of Hepatocellular Carcinoma: An Overview and Highlights for Immunotherapy Research. Nat. Rev. Gastroenterol. Hepatol. 2018, 15, 536–554. [Google Scholar] [CrossRef] [PubMed]
  237. Sandgren, E.P.; Quaife, C.J.; Pinkert, C.A.; Palmiter, R.D.; Brinster, R.L. Oncogene-Induced Liver Neoplasia in Transgenic Mice. Oncogene 1989, 4, 715–724. [Google Scholar]
  238. Arechederra, M.; Bazai, S.K.; Abdouni, A.; Sequera, C.; Mead, T.J.; Richelme, S.; Daian, F.; Audebert, S.; Dono, R.; Lozano, A.; et al. ADAMTSL5 Is an Epigenetically Activated Gene Underlying Tumorigenesis and Drug Resistance in Hepatocellular Carcinoma. J. Hepatol. 2021, 74, 893–906. [Google Scholar] [CrossRef]
  239. Sakata, H.; Takayama, H.; Sharp, R.; Rubin, J.S.; Merlino, G.; LaRochelle, W.J. Hepatocyte Growth Factor/Scatter Factor Overexpression Induces Growth, Abnormal Development, and Tumor Formation in Transgenic Mouse Livers. Cell Growth Differ. 1996, 7, 1513–1523. [Google Scholar] [PubMed]
  240. Tward, A.D.; Jones, K.D.; Yant, S.; Cheung, S.T.; Fan, S.T.; Chen, X.; Kay, M.A.; Wang, R.; Bishop, J.M. Distinct Pathways of Genomic Progression to Benign and Malignant Tumors of the Liver. Proc. Natl. Acad. Sci. USA 2007, 104, 14771–14776. [Google Scholar] [CrossRef]
  241. Kiguchi, K.; Carbajal, S.; Chan, K.; Beltrán, L.; Ruffino, L.; Shen, J.; Matsumoto, T.; Yoshimi, N.; DiGiovanni, J. Constitutive Expression of ErbB-2 in Gallbladder Epithelium Results in Development of Adenocarcinoma. Cancer Res. 2001, 61, 6971–6976. [Google Scholar]
  242. Rogler, C.E.; Yang, D.; Rossetti, L.; Donohoe, J.; Alt, E.; Chang, C.J.; Rosenfeld, R.; Neely, K.; Hintz, R. Altered Body Composition and Increased Frequency of Diverse Malignancies in Insulin-like Growth Factor-II Transgenic Mice. J. Biol. Chem. 1994, 269, 13779–13784. [Google Scholar] [CrossRef]
  243. Xu, X.; Kobayashi, S.; Qiao, W.; Li, C.; Xiao, C.; Radaeva, S.; Stiles, B.; Wang, R.-H.; Ohara, N.; Yoshino, T.; et al. Induction of Intrahepatic Cholangiocellular Carcinoma by Liver-Specific Disruption of Smad4 and Pten in Mice. J. Clin. Investig. 2006, 116, 1843–1852. [Google Scholar] [CrossRef]
  244. Horie, Y.; Suzuki, A.; Kataoka, E.; Sasaki, T.; Hamada, K.; Sasaki, J.; Mizuno, K.; Hasegawa, G.; Kishimoto, H.; Iizuka, M.; et al. Hepatocyte-Specific Pten Deficiency Results in Steatohepatitis and Hepatocellular Carcinomas. J. Clin. Investig. 2004, 113, 1774–1783. [Google Scholar] [CrossRef]
  245. Conner, E.A.; Lemmer, E.R.; Sánchez, A.; Factor, V.M.; Thorgeirsson, S.S. E2F1 Blocks and C-Myc Accelerates Hepatic Ploidy in Transgenic Mouse Models. Biochem. Biophys. Res. Commun. 2003, 302, 114–120. [Google Scholar] [CrossRef]
  246. Lewis, B.C.; Klimstra, D.S.; Socci, N.D.; Xu, S.; Koutcher, J.A.; Varmus, H.E. The Absence of P53 Promotes Metastasis in a Novel Somatic Mouse Model for Hepatocellular Carcinoma. Mol. Cell. Biol. 2005, 25, 1228–1237. [Google Scholar] [CrossRef]
  247. Zender, S.; Nickeleit, I.; Wuestefeld, T.; Sörensen, I.; Dauch, D.; Bozko, P.; El-Khatib, M.; Geffers, R.; Bektas, H.; Manns, M.P.; et al. A Critical Role for Notch Signaling in the Formation of Cholangiocellular Carcinomas. Cancer Cell 2013, 23, 784–795. [Google Scholar] [CrossRef]
  248. Cadoret, A.; Ovejero, C.; Saadi-Kheddouci, S.; Souil, E.; Fabre, M.; Romagnolo, B.; Kahn, A.; Perret, C. Hepatomegaly in Transgenic Mice Expressing an Oncogenic Form of Beta-Catenin. Cancer Res. 2001, 61, 3245–3249. [Google Scholar]
  249. Cho, K.; Ro, S.W.; Seo, S.H.; Jeon, Y.; Moon, H.; Kim, D.Y.; Kim, S.U. Genetically Engineered Mouse Models for Liver Cancer. Cancers 2020, 12, 14. [Google Scholar] [CrossRef] [PubMed]
  250. Loeuillard, E.; Fischbach, S.R.; Gores, G.J.; Rizvi, S. Animal Models of Cholangiocarcinoma. Biochim. Biophys. Acta-Mol. Basis Dis. 2019, 1865, 982–992. [Google Scholar] [CrossRef] [PubMed]
  251. Li, X.; Benjamin, I.S.; Alexander, B. Reproducible Production of Thioacetamide-Induced Macronodular Cirrhosis in the Rat with No Mortality. J. Hepatol. 2002, 36, 488–493. [Google Scholar] [CrossRef] [PubMed]
  252. Praet, M.M.; Roels, H.J. Histogenesis of Cholangiomas and Cholangiocarcinomas in Thioacetamide Fed Rats. Exp. Pathol. 1984, 26, 3–14. [Google Scholar] [CrossRef]
  253. Yoshimoto, S.; Loo, T.M.; Atarashi, K.; Kanda, H.; Sato, S.; Oyadomari, S.; Iwakura, Y.; Oshima, K.; Morita, H.; Hattori, M.; et al. Obesity-Induced Gut Microbial Metabolite Promotes Liver Cancer through Senescence Secretome. Nature 2013, 499, 97–101. [Google Scholar] [CrossRef] [PubMed]
  254. Koen, H.; Pugh, T.D.; Goldfarb, S. Centrilobular Distribution of Diethylnitrosamine-Induced Hepatocellular Foci in the Mouse. Lab. Investig. 1983, 49, 78–81. [Google Scholar]
  255. Newell, P.; Villanueva, A.; Friedman, S.L.; Koike, K.; Llovet, J.M. Experimental Models of Hepatocellular Carcinoma. J. Hepatol. 2008, 48, 858–879. [Google Scholar] [CrossRef]
  256. Groos, J.; Bannasch, P.; Schwarz, M.; Kopp-Schneider, A. Comparison of Mode of Action of Four Hepatocarcinogens: A Model-Based Approach. Toxicol. Sci. 2007, 99, 446–454. [Google Scholar] [CrossRef] [PubMed]
  257. National Toxicology Program. Toxicology and Carcinogenesis Studies of Furan (CAS No. 110-00-9) in F344 Rats and B6C3F1 Mice(Gavage Studies). Natl. Toxicol. Program Tech. Rep. Ser. 1993, 402, 1–286. [Google Scholar]
  258. Gallage, S.; Avila, J.E.B.; Ramadori, P.; Focaccia, E.; Rahbari, M.; Ali, A.; Malek, N.P.; Anstee, Q.M.; Heikenwalder, M. A Researcher’s Guide to Preclinical Mouse NASH Models. Nat. Metab. 2022, 4, 1632–1649. [Google Scholar] [CrossRef] [PubMed]
  259. Denda, A.; Kitayama, W.; Kishida, H.; Murata, N.; Tsutsumi, M.; Tsujiuchi, T.; Nakae, D.; Konishi, Y. Development of Hepatocellular Adenomas and Carcinomas Associated with Fibrosis in C57BL/6J Male Mice given a Choline-Deficient, L-Amino Acid-Defined Diet. Jpn. J. Cancer Res. 2002, 93, 125–132. [Google Scholar] [CrossRef] [PubMed]
  260. Denda, A.; Tang, Q.; Endoh, T.; Tsujiuchi, T.; Horiguchi, K.; Noguchi, O.; Mizumoto, Y.; Nakae, D.; Konishi, Y. Prevention by Acetylsalicylic Acid of Liver Cirrhosis and Carcinogenesis as Well as Generations of 8-Hydroxydeoxyguanosine and Thiobarbituric Acid-Reactive Substances Caused by a Choline-Deficient, L-Amino Acid-Defined Diet in Rats. Carcinogenesis 1994, 15, 1279–1283. [Google Scholar] [CrossRef] [PubMed]
  261. Itagaki, H.; Shimizu, K.; Morikawa, S.; Ogawa, K.; Ezaki, T. Morphological and Functional Characterization of Non-Alcoholic Fatty Liver Disease Induced by a Methionine-Choline-Deficient Diet in C57BL/6 Mice. Int. J. Clin. Exp. Pathol. 2013, 6, 2683–2696. [Google Scholar] [PubMed]
  262. Rizki, G.; Arnaboldi, L.; Gabrielli, B.; Yan, J.; Lee, G.S.; Ng, R.K.; Turner, S.M.; Badger, T.M.; Pitas, R.E.; Maher, J.J. Mice Fed a Lipogenic Methionine-Choline-Deficient Diet Develop Hypermetabolism Coincident with Hepatic Suppression of SCD-1. J. Lipid Res. 2006, 47, 2280–2290. [Google Scholar] [CrossRef] [PubMed]
  263. Ikawa-Yoshida, A.; Matsuo, S.; Kato, A.; Ohmori, Y.; Higashida, A.; Kaneko, E.; Matsumoto, M. Hepatocellular Carcinoma in a Mouse Model Fed a Choline-Deficient, L-Amino Acid-Defined, High-Fat Diet. Int. J. Exp. Pathol. 2017, 98, 221–233. [Google Scholar] [CrossRef]
  264. Kishida, N.; Matsuda, S.; Itano, O.; Shinoda, M.; Kitago, M.; Yagi, H.; Abe, Y.; Hibi, T.; Masugi, Y.; Aiura, K.; et al. Development of a Novel Mouse Model of Hepatocellular Carcinoma with Nonalcoholic Steatohepatitis Using a High-Fat, Choline-Deficient Diet and Intraperitoneal Injection of Diethylnitrosamine. BMC Gastroenterol. 2016, 16, 61. [Google Scholar] [CrossRef]
  265. He, L.; Tian, D.-A.; Li, P.-Y.; He, X.-X. Mouse Models of Liver Cancer: Progress and Recommendations. Oncotarget 2015, 6, 23306–23322. [Google Scholar] [CrossRef]
  266. Yang, H.; Li, T.W.H.; Peng, J.; Tang, X.; Ko, K.S.; Xia, M.; Aller, M.-A. A Mouse Model of Cholestasis-Associated Cholangiocarcinoma and Transcription Factors Involved in Progression. Gastroenterology 2011, 141, 378–388. [Google Scholar] [CrossRef]
  267. Kamegaya, Y.; Hiasa, Y.; Zukerberg, L.; Fowler, N.; Blackard, J.T.; Lin, W.; Choe, W.H.; Schmidt, E.V.; Chung, R.T. Hepatitis C Virus Acts as a Tumor Accelerator by Blocking Apoptosis in a Mouse Model of Hepatocarcinogenesis. Hepatology 2005, 41, 660–667. [Google Scholar] [CrossRef] [PubMed]
  268. Koike, K.; Moriya, K.; Iino, S.; Yotsuyanagi, H.; Endo, Y.; Miyamura, T.; Kurokawa, K. High-Level Expression of Hepatitis B Virus HBx Gene and Hepatocarcinogenesis in Transgenic Mice. Hepatology 1994, 19, 810–819. [Google Scholar] [CrossRef] [PubMed]
  269. Horiguchi, N.; Takayama, H.; Toyoda, M.; Otsuka, T.; Fukusato, T.; Merlino, G.; Takagi, H.; Mori, M. Hepatocyte Growth Factor Promotes Hepatocarcinogenesis through C-Met Autocrine Activation and Enhanced Angiogenesis in Transgenic Mice Treated with Diethylnitrosamine. Oncogene 2002, 21, 1791–1799. [Google Scholar] [CrossRef] [PubMed]
  270. Arechederra, M.; Recalde, M.; Gárate-rascón, M.; Fernández-barrena, M.G.; Ávila, M.A.; Berasain, C. Epigenetic Biomarkers for the Diagnosis and Treatment of Liver Disease. Cancers 2021, 13, 1265. [Google Scholar] [CrossRef] [PubMed]
  271. Baek, J.Y.; Morris, S.M.; Campbell, J.; Fausto, N.; Yeh, M.M.; Grady, W.M. TGF-Beta Inactivation and TGF-Alpha Overexpression Cooperate in an in Vivo Mouse Model to Induce Hepatocellular Carcinoma That Recapitulates Molecular Features of Human Liver Cancer. Int. J. Cancer 2010, 127, 1060–1071. [Google Scholar] [CrossRef] [PubMed]
  272. Jhappan, C.; Stahle, C.; Harkins, R.N.; Fausto, N.; Smith, G.H.; Merlino, G.T. TGF Alpha Overexpression in Transgenic Mice Induces Liver Neoplasia and Abnormal Development of the Mammary Gland and Pancreas. Cell 1990, 61, 1137–1146. [Google Scholar] [CrossRef]
  273. Santoni-Rugiu, E.; Nagy, P.; Jensen, M.R.; Factor, V.M.; Thorgeirsson, S.S. Evolution of Neoplastic Development in the Liver of Transgenic Mice Co-Expressing c-Myc and Transforming Growth Factor-Alpha. Am. J. Pathol. 1996, 149, 407–428. [Google Scholar]
  274. Murakami, H.; Sanderson, N.D.; Nagy, P.; Marino, P.A.; Merlino, G.; Thorgeirsson, S.S. Transgenic Mouse Model for Synergistic Effects of Nuclear Oncogenes and Growth Factors in Tumorigenesis: Interaction of c-Myc and Transforming Growth Factor Alpha in Hepatic Oncogenesis. Cancer Res. 1993, 53, 1719–1723. [Google Scholar]
  275. Uehara, T.; Pogribny, I.P.; Rusyn, I. The DEN and CCl4 -Induced Mouse Model of Fibrosis and Inflammation-Associated Hepatocellular Carcinoma. Curr. Protoc. Pharmacol. 2014, 66, 14.30.1–14.30.10. [Google Scholar] [CrossRef]
  276. Poirier, L.A. Hepatocarcinogenesis by Diethylnitrosamine in Rats Fed High Dietary Levels of Lipotropes. J. Natl. Cancer Inst. 1975, 54, 137–140. [Google Scholar] [CrossRef]
  277. Pascale, R.M.; Simile, M.M.; Feo, F. Genomic Abnormalities in Hepatocarcinogenesis. Implications for a Chemopreventive Strategy. Anticancer Res. 1993, 13, 1341–1356. [Google Scholar] [PubMed]
  278. Solt, D.B.; Medline, A.; Farber, E. Rapid Emergence of Carcinogen-Induced Hyperplastic Lesions in a New Model for the Sequential Analysis of Liver Carcinogenesis. Am. J. Pathol. 1977, 88, 595–618. [Google Scholar] [PubMed]
  279. Shariff, M.I.F.; Tognarelli, J.M.; Lewis, M.R.; Want, E.J.; Mohamed, F.E.Z.; Ladep, N.G.; Crossey, M.M.E.; Khan, S.A.; Jalan, R.; Holmes, E.; et al. Plasma Lipid Profiling in a Rat Model of Hepatocellular Carcinoma: Potential Modulation Through Quinolone Administration. J. Clin. Exp. Hepatol. 2015, 5, 286–294. [Google Scholar] [CrossRef] [PubMed]
  280. Barnes, J.M.; Magee, P.N. Some Toxic Properties of Dimethylnitrosamine. Br. J. Ind. Med. 1954, 11, 167–174. [Google Scholar] [CrossRef] [PubMed]
  281. Encell, L.; Foiles, P.G.; Gold, B. The Relationship between N-nitrosodimethylamine Metabolism and DNA Methylation in Isolated Rat Hepatocytes. Carcinogenesis 1996, 17, 1127–1134. [Google Scholar] [CrossRef] [PubMed]
  282. Weisburger, E.K.; Ulland, B.M.; Nam, J.; Gart, J.J.; Weisburger, J.H. Carcinogenicity Tests of Certain Environmental and Industrial Chemicals. J. Natl. Cancer Inst. 1981, 67, 75–88. [Google Scholar] [PubMed]
Figure 1. Regulation of liver tumor development and progression by the tumor microenvironment. The role played by different constituents of liver tumor microenvironment is shown as well as the mediators of their actions (growth factors, cytokines, chemokines, and other molecules). Cell populations include: myofibroblasts (MFBs), cancer associated fibroblasts (CAFs), platelets and different populations of the immune system: cytotoxic T lymphocytes (CTLs), regulatory T lymphocytes (Tregs), M1 and M2 tumor associated macrophages (TAMs), myeloid derived suppressor cells (MDSCs), N1 and N2 tumor associated neutrophils (TANs) and natural killer cells (NKs). Created in BioRender.com.
Figure 1. Regulation of liver tumor development and progression by the tumor microenvironment. The role played by different constituents of liver tumor microenvironment is shown as well as the mediators of their actions (growth factors, cytokines, chemokines, and other molecules). Cell populations include: myofibroblasts (MFBs), cancer associated fibroblasts (CAFs), platelets and different populations of the immune system: cytotoxic T lymphocytes (CTLs), regulatory T lymphocytes (Tregs), M1 and M2 tumor associated macrophages (TAMs), myeloid derived suppressor cells (MDSCs), N1 and N2 tumor associated neutrophils (TANs) and natural killer cells (NKs). Created in BioRender.com.
Ijms 24 17152 g001
Figure 2. Functions of C3G in liver cancer. The scheme shows C3G functions in liver cancer at different levels. C3G protein expression is low in adult hepatocytes but increases in HCC cells promoting tumor growth and HGF/MET signaling. High C3G levels in patient tumor samples are associated with poor survival in patients. C3G present in platelets also contributes to enhance liver tumor growth in a mouse model of HCC associated with fibrosis (our unpublished data). On the other hand, hepatic progenitor cells (HPCs) express high C3G levels, which facilitate adhesion and allow a proper HGF/MET signaling. In contrast, low levels of C3G increase cell migration, invasion and stemness in HPCs. Arrows indicate upregulation ( ) or downregulation ( ). Created in BioRender.com.
Figure 2. Functions of C3G in liver cancer. The scheme shows C3G functions in liver cancer at different levels. C3G protein expression is low in adult hepatocytes but increases in HCC cells promoting tumor growth and HGF/MET signaling. High C3G levels in patient tumor samples are associated with poor survival in patients. C3G present in platelets also contributes to enhance liver tumor growth in a mouse model of HCC associated with fibrosis (our unpublished data). On the other hand, hepatic progenitor cells (HPCs) express high C3G levels, which facilitate adhesion and allow a proper HGF/MET signaling. In contrast, low levels of C3G increase cell migration, invasion and stemness in HPCs. Arrows indicate upregulation ( ) or downregulation ( ). Created in BioRender.com.
Ijms 24 17152 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Cuesta, Á.M.; Palao, N.; Bragado, P.; Gutierrez-Uzquiza, A.; Herrera, B.; Sánchez, A.; Porras, A. New and Old Key Players in Liver Cancer. Int. J. Mol. Sci. 2023, 24, 17152. https://doi.org/10.3390/ijms242417152

AMA Style

Cuesta ÁM, Palao N, Bragado P, Gutierrez-Uzquiza A, Herrera B, Sánchez A, Porras A. New and Old Key Players in Liver Cancer. International Journal of Molecular Sciences. 2023; 24(24):17152. https://doi.org/10.3390/ijms242417152

Chicago/Turabian Style

Cuesta, Ángel M., Nerea Palao, Paloma Bragado, Alvaro Gutierrez-Uzquiza, Blanca Herrera, Aránzazu Sánchez, and Almudena Porras. 2023. "New and Old Key Players in Liver Cancer" International Journal of Molecular Sciences 24, no. 24: 17152. https://doi.org/10.3390/ijms242417152

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