Next Article in Journal
Pathogenesis and Clinical Management of Uterine Serous Carcinoma
Next Article in Special Issue
Identification of the Novel Oncogenic Role of SAAL1 and Its Therapeutic Potential in Hepatocellular Carcinoma
Previous Article in Journal
Adoptive Cell Therapy—Harnessing Antigen-Specific T Cells to Target Solid Tumours
Previous Article in Special Issue
Inflammatory Mechanisms of HCC Development
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Cancer Stem Cell in Hepatocellular Carcinoma

by
Lucas-Alexander Schulte
1,
Juan Carlos López-Gil
2,3,
Bruno Sainz, Jr.
2,3,4,* and
Patrick C. Hermann
1,*
1
Department of Internal Medicine I, Ulm University, 89081 Ulm, Germany
2
Department of Biochemistry, Universidad Autonoma de Madrid (UAM), 28029 Madrid, Spain
3
Department of Cancer Biology, Instituto de Investigaciones Biomedicas “Alberto Sols” (IIBM), CSIC-UAM, 28029 Madrid, Spain
4
Chronic Diseases and Cancer, Area 3-Instituto Ramon y Cajal de Investigacion Sanitaria (IRYCIS), 28029 Madrid, Spain
*
Authors to whom correspondence should be addressed.
Cancers 2020, 12(3), 684; https://doi.org/10.3390/cancers12030684
Submission received: 25 February 2020 / Revised: 11 March 2020 / Accepted: 12 March 2020 / Published: 14 March 2020
(This article belongs to the Special Issue Liver Cancer and Potential Therapeutic Targets)

Abstract

:
The recognition of intra-tumoral cellular heterogeneity has given way to the concept of the cancer stem cell (CSC). According to this concept, CSCs are able to self-renew and differentiate into all of the cancer cell lineages present within the tumor, placing the CSC at the top of a hierarchical tree. The observation that these cells—in contrast to bulk tumor cells—are able to exclusively initiate new tumors, initiate metastatic spread and resist chemotherapy implies that CSCs are solely responsible for tumor recurrence and should be therapeutically targeted. Toward this end, dissecting and understanding the biology of CSCs should translate into new clinical therapeutic approaches. In this article, we review the CSC concept in cancer, with a special focus on hepatocellular carcinoma.

1. Introduction

In solid tumor oncology, there are only few examples of medical treatments that are curative or that achieve sustained and long-term remission. In general, most patients retain at least some tumor burden and/or suffer relapse after initial treatment response. Several hypotheses have tried to provide an explanation for this phenomenon. The clonal evolution model refers to a model in which a therapy is administered at doses that cause tolerable, non-permanent damage to healthy tissues, causing only moderate side effects, whereas tumor cells, which are more susceptible to therapy, are targeted and eliminated [1]. Based on pure probability, however, some cancer cell clones will possess resistance to therapy due to additional acquired mutations or microenvironmental cues that will facilitate their survival, giving rise to a new clonal tumor that can grow despite continued treatment. In the clonal evolution model, heterogeneity within each tumor is a result of randomly acquired (and subsequently selected) alterations, and every tumor cell equally contributes to the fate of the tumor. Furthermore, all cancer cells have tumor-initiating capacity, and those that survive therapy will drive relapse.
In a way, the clonal evolution model fits the Darwinian concept of “survival of the fittest.” Those cells that adapt to new challenges confer a selective advantage to their progeny. This can lead to therapy-induced selection of a dominant clone with new biological features, and to the loss of those cells that were “addicted” to the previous molecular target, as has been observed with anti-EGFR (epithelial growth factor receptor) therapy in colorectal carcinoma [2]. While clinical observations such as acquired drug resistance or tumor relapse can be explained by the clonal evolution model, evidence to the contrary exists. Most compelling are studies showing that tumorigenic activity is restricted to certain cancer cell populations. That is, the majority of “bulk” tumor cells either fail altogether to generate tumors in xenotransplantation models, or exponentially higher numbers of cells have to be transplanted in order to generate a tumor [3]. At the same time, distinct subpopulations of cancer cells with specific surface marker expression profiles are exceptionally tumorigenic and give rise to fully-fledged tumors after xenotransplantation of extremely low numbers of cells. This phenomenon was first observed in acute myeloid leukemia (AML) in 1994 by John Dick and colleagues [4]. In this study, the authors demonstrated that CD34+/CD38− AML cells had significantly higher tumorigenic potential in SCID mice than the CD34+/CD38+ or CD34− fractions. This discovery of intra-tumoral heterogeneity and tumor hierarchy led to the advent of the cancer stem cell (CSC) concept, with a CSC being a tumor cell that can self-renew, differentiate into all tumor cell lineages, exhibit exclusive tumor-initiating capacity and possess high chemotherapeutic resistance [5]. This model places the CSC at the apex of the tumor hierarchy, undergoing asymmetric and symmetric divisions, the latter ensuring perpetual self-renewal and maintenance of the CSC sub-population. Likewise, the model proposes that CSCs are exclusively able to intravasate into the circulation, survive and successfully metastasize to distant locations. Since metastasis and tumor relapse dictate overall survival, targeting CSCs provides new hope for improving long-term survival.
Hepatocellular carcinoma (HCC) is a frequent and prognostically poor malignancy of the liver. Currently, HCC represents the fourth leading cause of cancer-related death worldwide and the sixth most commonly diagnosed cancer [6]. HCC is typically divided into two molecular subgroups: (1) proliferation class, which is poorly differentiated, expresses high alpha-fetoprotein (AFP) levels and is commonly associated with hepatitis B virus (HBV) infection; and (2) non-proliferation class, which is more commonly associated with alcohol-related HCC or hepatitis C virus (HCV) infection [7]. Treatment of HCC is hampered by several factors: (1) the tumor is highly resistant to classical chemotherapy and is usually detected at late stages with no hope for complete surgical resection, and perhaps most importantly, (2) HCC typically develops in critically damaged cirrhotic livers that have suffered years of active hepatitis or toxin intake, and therefore, patient tolerance to therapy is severely limited. In the following article, we will summarize the role of CSCs in cancer, and provide an overview of the current knowledge of CSCs in HCC.

2. Cancer Stem Cells and Their Role in Cancer

2.1. Cancer Stem Cells

2.1.1. CSCs and Their Clinical Relevance

The first evidence for the existence of CSCs came from studies in AML in 1994 by Dick and colleagues [4], who provided the first proof of a hierarchical architecture in a human cancer. By xenotransplanting AML cells into severe combined immunodeficient (SCID) mice, the authors found profound inter-patient differences in the cell numbers needed for leukemia induction. Specifically, after using FACS (sorting) on AML cells, the leukemia-inducing cells were exclusively found in a subset of CD34+/CD38− cells. These cells have been discovered in many other malignancies, such as pancreatic, breast, glioma, prostate, lung and liver carcinomas [8,9,10,11,12,13,14,15], and are referred to as CSCs, tumor-initiating cells (TICs) or cancer stem-like cells [16]. Importantly, throughout the last two decades, many groups have demonstrated that these sub-populations of cells exhibit inherent plasticity and dormancy, mimicking many of the characteristics of somatic stem cells [5]. Importantly, similarly to tissue-resident stem cells, the differentiation of CSCs into other tumor cells is not unidirectional. Meaning, a tumor’s CSC pool can be replenished by de-differentiation of “bulk cells” via a process referred to as transdifferentiation. This concept was experimentally shown in two studies from 2017. First, de Sousa e Melo et al. used mouse-derived colon cancer organoid cultures (Apcmin; KrasG12V; p53CRIPSR; SMAD4CRIPSR) that expressed EGFP (enhanced green fluorescent protein) and the diphtheria-toxin (DT) receptor (DTR) under the control of the leucine-rich repeat-containing G-protein coupled receptor 5 (LGR5) promoter to initiate tumors in recipient mice [17]. The authors showed that DT treatment inhibited tumor growth concomitant with loss of the EGFP+/LGR5+ cells. Shimokawa et al. targeted the LGR5+ CSC population in organoid cultures of human colorectal cancer using an inducible suicide-gene caspase 9 (iCasp9) approach [18], and demonstrated that upon iCasp9 induction, organoid-derived xenograft growth in vivo was significantly reduced. Interestingly, upon removal of DT or the inducer, tumors regrew and the LGR5+ CSC population reemerged, indicating replenishment of the CSC pool from LGR5-negative cells. However, such transdifferentiation was not observed by Chen et al. in a mouse model of glioma, where the authors labeled glioblastoma cells using a nestin-ΔTK-IRES-GFP transgene. Upon treatment with ganciclovir (which specifically eliminates herpes simplex virus type-1 thymidine kinase (TK)-expressing cells) and temozolomide, tumor development was abrogated [19]; however, no replenishment of the CSC pool from other glioblastoma cells was observed.
The clinical relevance of the CSC model is based on the observation that the resistance of CSCs to chemotherapy and radiation makes them responsible for the relapse of malignant diseases [20]. Likewise, CSCs can undergo epithelial-to-mesenchymal transition (EMT), invade, circulate in the blood stream and extravasate at distant sites to form metastatic lesions [9,21,22,23,24,25]. Thus, from a therapeutic perspective, long-term effective anti-cancer therapies must also eliminate these CSCs, and much effort has been invested in defining and understanding the mechanisms underlying the aforementioned key properties of CSC biology. CSCs seem to enter a more dormant state with reduced proliferative activity. In this state of G0-arrest, cells resist chemotherapy and persist, sometimes for years, eventually causing relapse [26]. Furthermore, CSCs exhibit increased DNA repair and reduced apoptosis compared to bulk tumor cells [27,28,29,30], and strongly express drug efflux transporters, such as multi-drug resistant (MDR) proteins, a feature that can also be used to identify and isolate these cells [31]. Perhaps equally important with respect to the maintenance and protection of CSCs is the niche, a location within the tumor microenvironment (TME) that “nurses” CSCs by providing anti-apoptotic, stemness-maintaining factors and matrix components. This special TME is believed to not only help CSCs resist therapeutic efforts but also to play an important role in the trans-differentiation of non-CSCs into CSCs.

2.1.2. CSC Identification and Markers

Two hallmarks of CSCs are (1) their ability to re-establish tumors with the same molecular and phenotypic heterogeneity (and perhaps subtype) as the original tumor (differentiation), and (2) the maintenance of the CSC pool via unlimited self-renewal, ensuring that the CSC as the source of tumor replenishment is never lost. These two properties are typically validated in (xeno-)transplantation models where putative CSCs are implanted in mice and tumor formation is assessed, and in subsequent serial transplantation models to asses unlimited self-renewal. A large number of cellular markers have been used to identify and experimentally validate CSCs, including CD133, LGR5, CD44, Aldehyde dehydrogenase 1 (ALDH1), Mucin 1, ATP-binding cassette super-family G member 2 (ABCG2) and many more. However, it is important to note that to date no universal CSC marker or combination of markers have been identified/established. While the function of some of these markers in CSC biology is unknown, CSC-specific functions for many of these markers have been described. ALDH1 is one of the aldehyde-dehydrogenase isoforms and is overexpressed in many CSCs, clearing toxic aldehydes and lowering oxidative stress. Additionally, ALDH1 has been shown to confer a survival advantage to CSCs of many tumor entities, and its expression is relevant for prognosis [32,33]. CD133 (prominin-1) is a commonly used CSC marker. It is a lipid-raft bound transmembrane protein, which is involved in the activation of EMT-promoting signals, such as autocrine interleukin (IL) IL-1 and NFκB signaling [34,35,36]. It has been used successfully to identify CSCs in pancreatic [9] and breast cancer [37], hepatocellular carcinoma [38] and glioblastomas [15]. Several ATP-binding-cassette (ABC) proteins have also been identified as CSC markers. These MDR proteins are themselves upregulated by stemness-associated signaling pathways, such as Wnt, and act as active efflux pumps for several substances, including anti-cancer drugs, and are therefore functionally important [39]. Furthermore, ABC proteins such as ABCG2 are the basis for functional assays to enrich for CSCs. Staining with the DNA binding dye Hoechst-33342 identifies CSCs with high expression and activity of ABCG2, as only these cells are capable of depleting this intracellular stain, while non-CSCs will retain the dye. These non-Hoechst-retaining cells can be identified in cytometry plots as a “side population” [26]. Interestingly, ABCG2 was also shown by Miranda-Lorenzo et al. to promote the accumulation of riboflavin in cytoplasmic vesicles. This accumulation of riboflavin results in autofluorescence, which has been used successfully to enrich for CSCs in pancreatic, colorectal and liver cancers [40].

2.1.3. The CSC Niche, the Tumor Microenvironment and Metastasis

While little is known about the CSC “niche,” like somatic stem cells, CSCs are likely to reside within a supportive environment that regulates their stemness [41]. The CSC niche is formed by the interplay of different cells and factors, which ensures the generation of a proper environment to maintain the CSC population. While the niche is a changing entity that undergoes significant variations depending on the location of the CSC(s) within the TME and the exact moment in time during tumor evolution, an assortment of common factors predominates throughout the life of the CSC niche. Fibroblasts, mesenchymal stem cells and immune cells are the most important support cells found in the CSC niche, although other cell types (e.g., endothelial cells) also contribute to the niche. These cells secrete factors that can modulate the fate of the CSCs, and CSCs will produce factors to adapt the niche to their needs [42,43]. In the end, a bidirectional and dynamic environment is formed between CSCs and the cells of the niche, and disrupting this environment could be detrimental for CSCs. Fibroblasts, a key structural cell type in many tissues, exert essential repair functions following tissue injury. In cancer, however, fibroblasts can become cancer-associated fibroblasts (CAFs), which are characterized by rapid proliferation and production of tumor-promoting factors (e.g., VEGF (vascular endothelial growth factor), SDF1 (stroma-derived factor 1), HGF (hepatocyte growth factor) or CXCLs (Chemokine [C-X-C Motif] Ligands)) that enhance tumor growth and progression [44]. The communication circuit between CSCs and CAFs is essential for maintaining stemness and is mediated by the activation of signaling pathways such as STAT-3-NF-κB [45], NOTCH [46] or Wnt [22,43]. This interplay can also promote migration, expansion and de-differentiation of bulk tumor cells into CSCs. The roles of CAFs and CSCs were recently reviewed in Alguacil-Núñez et al. in 2018 [47].
Mesenchymal stem cells (MSCs) are cells with the ability to differentiate into structural lineages (osteoblasts, adipocytes and chondrocytes) and have been isolated from a wide variety of organs and tissues [48]. Bone marrow-derived MSCs are recruited to the TME, providing the CSC niche with a combination of secreted cytokines that modulate the stroma [49], the immune system [50] and the MSCs themselves [51]. This facilitates CSC maintenance, but also chemoresistance within the TME [52].
Finally, the niche is also characterized by the infiltration of anti-inflammatory/pro-tumoral immune cells. Anti-inflammatory cells will be attracted to deal with the effect of pro-inflammatory innate immune cells (e.g., M1 macrophages and natural killer cells) and cytotoxic T cells. Tumor-associated macrophages (TAMs) and myeloid-derived suppressor cells (MDSCs) [53] enhance stemness and favor a pro-tumoral and pro-metastatic state by releasing cytokines (e.g., IL-6) that activate STAT3, ultimately inducing several stem-related pathways, EMT, and the expression of CSC markers, such as CD133 [54]. Underscoring the significance of the niche for the maintenance of CSCs, TAM depletion leads to a reduction in CSC numbers, tumor size and metastasis [55], while MDSC depletion increases the effect of cytotoxic T cells and reduces tumor growth [56].
Metastasis depends on several steps. First, a tumor cell has to lose its intercellular adhesions (via EMT) and invade the surrounding tissue, migrate to nearby lymphatic or blood vessels, enter the circulation, survive splenic and intravasal immune surveillance, stop its travel in a suitable target tissue, extravasate and form metastatic lesions (via reversed EMT; i.e., mesenchymal-to-epithelial transition (MET)). Increasing evidence suggests that the TME activates pro-metastatic signals in CSCs via factors secreted by the niche, or by other factors that affect the CSC state, such as hypoxia, acidic pH or glucose deprivation. VEGF is a pro-angiogenic transmitter with multiple functions in tumor biology. Intra-tumoral hypoxia is a key driver of VEGF production, and CSCs can produce VEGF [57]. Together, hypoxia and VEGF lead to the formation of atypical vessels with increased permeability. Furthermore, VEGF can induce EMT in CSCs, activating the circuitry needed for CSCs to lose their intercellular adhesions and enter nearby vessels [58,59,60,61,62,63]. Indeed, anti-VEGF treatments have been successfully used in AFPhigh HCC patients [64], but their specific efficacy against CSCs is unclear [65,66]. The latter is likely due to CSCs being able to “hijack” angiogenesis in vasculogenic mimicry [67]; that is, CSCs forming the endothelium in tumor-associated vessels [68,69,70,71].
Once in the blood stream, CSCs have to overcome new obstacles. Only about 1 out of 500 CSCs is able to survive in circulation [72], and elaborate mechanisms are in place to protect CSCs from elimination by the immune system [73]. For example, CSCs decrease major histocompatibility complex I (MHC-I) expression and antigen presentation, evading immune cell attack via the expression of PD-group ligands and instigating bystander cells to create an immunosuppressive milieu. Furthermore, CSCs can shield themselves by forming adducts with transforming growth factor-beta (TGF-β)-producing platelets and fibroblasts, and they can protect their micro-milieu by forming spheres in the bloodstream [73,74,75,76,77,78,79,80,81,82]. To successfully form distant metastases, the metastatic site must provide the appropriate milieu necessary to support all essential CSC properties. Exosomes or exosomal-like extracellular vesicles (ECV) are small membrane-protected bodies that can carry bioactive compounds, such as micro-RNAs and proteins, to cells at distant sites. In cancer, it has been shown that these ECVs are responsible for the formation of pre-metastatic CSC niches in distant organs, where they adhere by site-specific integrin interactions. These pre-metastatic niches are required for the successful seeding of CSCs and are built with the collaboration and support of ECV-educated bone marrow-derived cells and macrophages that produce vascular structures and promote the activation of stemness-related pathways, such as Wnt, Notch or Src signaling [83,84,85,86,87,88,89,90]. Once at the metastatic site, CSCs are thought to attach to the nearby endothelium by integrin- and lectin-mediated cell adhesion molecules and start proliferating in the capillaries to which they adhere, followed by invasion [91,92].

3. CSCs and Their Role in Hepatocellular Carcinoma

3.1. Stem Cells in the Normal Liver

Compared to other organs, the liver has the unique ability of extensive regeneration after damage-induced tissue loss. In fact, less than 50% of the original parenchyma is sufficient to restore the original liver mass [93]. Experiments have proven the liver to be capable of multiple cycles of self-renewal, raising hopes for regenerative therapeutic approaches [94,95,96,97]. Depending on the severity of the liver damage and possible replication restrictions of the remaining hepatocytes, several patterns of repair have been observed [98,99,100]. Partial hepatectomy prompts a strong growth signal to the remaining liver tissue, leading to the proliferation of hepatocytes and subsequent regeneration of the liver mass. In experimental settings, hepatocytes are able to divide more than 60 times without reaching replicative senescence or undergoing obvious functional loss [94]. However, other models of extensive liver damage show a reduced proliferative capacity of hepatocytes. In this case a distinct cell population residing in the periportal area near the smallest bile ducts starts to proliferate and to differentiate into hepatocytes and cholangiocytes in a “ductular reaction”. Many observations support this scenario, but there is also contradictory evidence from lineage-tracing experiments [101]. While the primordial hepatic stem cell responsible for cellular repopulation has not been definitively discovered to date, a distinct liver-resident “oval cell” most likely represents the bipotent progeny of these stem cells and proliferates to rebuild the lost parenchyma in mice [102]. Marker profiling shows heterogeneity amongst oval cells suggesting different lineages, but a common marker profile combines markers of biliary, hepatoblastic and hematopoetic progenitor cells [103,104,105]. This implicitly strict model of the potency of single cells, however, is contradicted by other studies showing that hepatocytes can transdifferentiate into other lineages without genetic manipulation, proving the bipotency of hepatocytes [106,107]. Although Sox9 lineage tracing of oval cells showed that these cells are bipotent in organoid cultures, they only marginally contribute to liver regeneration, even in the typical “ductular reaction” models [108]. Therefore, it could be hypothesized that the repopulative relevance of oval cells is not so much due to bipotency, but rather due to their resistance to hepatotoxic agents as a consequence of their less differentiated state/relative quiescence that prevents them from metabolizing hepatotoxins [109].
The gold standard to demonstrate cellular inheritance or cell fate is lineage tracing. In most cases, a cell type-specific promotor driving the expression of the Cre-recombinase and a loxP-STOP-loxP flanked reporter construct are inserted into the germline DNA. Thus, the (fluorescent) reporter is expressed when the cell type-specific promotor becomes transcriptionally active and Cre is expressed, leading to excision of the stop cassette and the permanent expression of the reporter construct in the cell and all its progeny. Models in which a hepatocyte has been shown to be the cell responsible for liver regeneration are supported by many transplantation and lineage tracing experiments [94,95,96,97,110], but there is also evidence that other cell populations inside and outside the liver are capable of liver regeneration [111,112,113,114,115]. The bone marrow is linked to hepatic stem cell pools in several ways. Oval cells exhibit markers suggestive of a bone marrow descendance [116], and hematopoiesis moves from the liver and spleen to the bone marrow during ageing [117]. Accordingly, it was shown in allogenic bone marrow transplant models that liver repopulation could ultimately stem from the grafted marrow [118]. While this seems not to have major relevance in liver regeneration following injury [119], the role of the bone marrow in HCC carcinogenesis has yet to be determined. MSCs originate from multiple tissues, including bone marrow or adipose tissue, and represent another possible source of hepatic progenitor cells [120], since studies have shown MSCs to be capable of engrafting in liver repopulation models [121,122].
The cellular compartment giving rise to the tumor-initiating cell in HCC is under intense debate, and either progenitor/stem cells or differentiated liver cells could be responsible. Experimental evidence indicates a model where every cell can be a key player [123]. There is profound inter-tumoral heterogeneity in HCC with respect to genetic profiles, allowing for the discrimination of at least two genetic/molecular subgroups, as described above [7,124], and transcriptional profiling can distinguish different prognostic subgroups. Interestingly, a subtype with a progenitor cell-like expression pattern has been identified and seems to be associated with activation of AP-1 transcription factors and poor prognosis [125]. Likewise, RNA and protein expression levels of Epithelial cell adhesion molecule (EpCAM) and AFP could define prognostically distinct subgroups of HCC, and EpCAM expression was associated with a progenitor gene expression signature, including tyrosine-protein kinase KIT (c-Kit), Wnt activity and CK19 expression [126]. In turn, a CK19 expression signature correlates with prognosis [127].
Several lines of evidence indicate that hepatic progenitor cells (HPCs) could represent the ancestors of the TIC in HCC. Embryonic liver fodrin (ELF) is a downstream target of TGF-β/SMAD signaling that is expressed in a stem cell-like subpopulation of the liver, indicating a possible “stem” cell origin. Interestingly, heterozygous ELF knock-out mice spontaneously develop HCC, regulated by hampered TGF-β signaling [128]. In another model, Wu et al. subjected an HPC-like cell line (WB-F344) to long-term in vitro treatment with TGF-β, resulting in an AKT-dependent enhanced tumorigenic potential in NOD-SCID mice [129], and suggesting HPCs to be the source of TICs in HCC. Genetically engineered fetal murine hepatoblasts (i.e., progenitor cells) harboring alterations inactivating p53 and activating several oncogenes (c-myc, activated Akt (Akt1) or oncogenic Ras (H-RasV12)) formed tumors in preconditioned wild-type recipient livers [130]. Furthermore, attenuation of the developmental pathways Hippo and Neurofibromatosis type 2 (which have been associated with HCC) led to HPC expansion and tumor formation [131,132]. In β-catenin-stabilized mouse models, only HPCs can generate tumors, while hepatocytes need further genetic alterations to form malignant liver tumors [133,134]. Finally, restricting liver cell survival by epigenetic induction of G2-arrest combined with STAT3-activation leads to HCC formation with HPC-like features [135].
While there is significant evidence to support HPCs as the “cell of origin” in HCC, hepatocytes have also been shown to be responsible for HCC development. Lineage-tracing models revealed that in certain HCC models, tumors are derived from hepatocytes and not from HPCs. Using Hepatocyte nuclear factor -1beta (HNF-1β) as an HPC marker, no contribution to genetically or chemically-induced HCC could be attributed to HPCs [101]. In another hepatocyte tracing model, nearly all chemically or genetically induced HCCs were the progeny of mature hepatocytes [136,137,138]. Recently, a self-maintaining pericentral group of LGR5+ hepatocytes was shown to be highly susceptible to hepatocarcinogenesis, and was determined to be primarily responsible for tumor development in diethylnitrosamin (DEN)-induced HCC [110]. LGR5 regulates chemoresistance via Wnt potentiation, p53 suppression and EMT induction in HCC, all of which are typical characteristics of CSCs [139,140]. Furthermore, LGR5 is an established CSC marker in colorectal cancer [18,141]. These observations indicate that in HCC, the mechanism of CSC/TIC generation may be the induction of stem cell traits rather than cellular inheritance. This scenario is further supported by the observation that Nestin expression following p53 loss is associated with the dedifferentiation of mature hepatocytes into progenitor-like cells in hepatocarcinogenesis, a process that is mediated by lineage-specific mutations that target Wnt signaling [142].

3.2. Identification of CSCs in HCC

CSCs have been characterized in HCC by different approaches. Figure 1 and Table 1 provide an overview of the most well-known HCC CSC markers and their physiological functions. Since every method to isolate CSCs relies on specific (and sometimes few) properties or individual methodological approaches, one should not consider the identified cell populations as pure, but rather as subpopulations enriched in CSCs. It is likely that the different approaches also identify varying CSC subpopulations, so comparing the results of different approaches has to be done with great caution.
A side population (SP) of cells can be isolated by flow cytometry based on their ability to efflux Hoechst dyes. This indicates their ABC-transporter activity, which is mediated by ABCG2, ABCG5 and MDR1 [150]. This side population was first identified in two out of four tested HCC cell lines [151], and sorting for these cells revealed that in xenotransplantation models, 1000 SP cells generated tumors, while 1 × 106 non-SP cells were unable to do so. Furthermore, tumors derived from SP cells differentiated into SP and non-SP cells and showed increased expression of stemness-associated genes. Similar to the results in cell lines, a corresponding SP was identified in primary HCCs [152], establishing the side population as a putative CSC population in HCC and linking a CSC phenotype to drug resistance (Table 2).
CD24 was identified as a CSC-associated cell surface marker in HCC by gene expression profiling. CD24+ cells were enriched by cisplatin treatment in xenotransplanted mice and increased in tumorigenicity upon serial transplantation. Interestingly, CD24 expression was associated with worse clinical outcomes and is linked to stemness characteristics. Furthermore, CD24 expression is essential for metastasis, differentiation and self-renewal in HCC cell lines. Via phosphorylation of STAT3, CD24 induces NANOG, a key factor for self-renewal in embryonic stem cells and during tumor development [143,144], thus enabling HCC cell self-renewal [145]. In addition to increased cisplatin resistance, CD24 expression correlates with sorafenib-resistance and clinical outcome in HCC patients. This was shown to be due to the activation of autophagy, and was partially reversible upon autophagy inhibition [159]. CD24 expression in HCC has also been associated with iNOS-induced TACE/ADAM17-dependent activation of Notch1, which is linked to tumor progression and stemness [160].
CD24 expression in HCC largely overlaps with the expression of several other important CSC cell surface markers; namely, CD133, CD44 and EpCAM. CD133 has been detected in HCC cell lines and primary HCC tissue, and isolated CD133+ cells show high tumorigenicity and clonogenicity compared to CD133− cells [161]. In the human hepatoma cell line Huh-7, CD133+ cells formed tumors in SCID mice more efficiently than their CD133− counterparts, and the CD133+ population expressed fewer markers of mature hepatocytes such as glutamine synthetase and cytochrome P450 3A4 [162]. Others have found that CD133 expression is linked to a broader differentiation capacity; enhanced self-renewal and colony forming ability; and increased expression of stemness genes [38,146]. On a functional level, CD133 has been found to be associated with neurotensin-induced IL-8 and CXCL1 pathway activation, leading to enhanced MAPK pathway activity, which in turn results in increased tumorigenicity, angiogenesis and self-renewal. Similar to most CSC markers, CD133 expression has been associated with worse outcomes in HCC [163]. In an attempt to target CD133 therapeutically, Smith et al. used an anti-CD133-antibody (AC133) conjugated to the cytotoxic drug monomethyl auristatin F (MMAF) to effectively target CD133+ HCC cells. Importantly, this functionalized antibody effectively delayed the growth of Hep3B-derived tumors in SCID mice [164]. Interestingly, the NIH has listed a CD133-CART (chimeric antigen receptor T-cell) phase-I study for several malignant tumors, including liver cancer (NCT02541370), underscoring the potential usefulness of strategies targeting CSCs.
The CSC-associated surface marker CD44 has many physiological functions, including binding to hyaluronic acid, osteopontin, fibronectin and collagen [165], and it is a co-stimulatory protein of the EGF receptor and c-Met. CD44 is upregulated on TICs and proliferating cells in HCC [166], and has been shown to be essential for EMT in breast cancer [167]. CD44−/− mice were resistant to HCC induction using the hepatic procarcinogen diethylnitrosamine, and CD44 on murine hepatocytes is induced by carcinogen exposure and suppresses p53-mediated genomic surveillance via AKT and Mdm2 [148]. The upregulation of CD44 was due to IL-6/STAT3 signaling, the latter of which has already been described as a mechanism of HCC induction [168] and CSC maintenance [169]. In Huh7 cells, the CD44 standard isoform (CD44s) was shown to be required for tumorigenicity and drug resistance. Its loss also downregulated the expression of other CSC markers, including CD133 and EpCAM; impaired the antioxidative capacity of Huh7 cells via downregulation of glutathione peroxidase 1; and reduced NOTCH3 and its target genes. Again, like other CSC markers, CD44 expression is associated with poor patient prognosis [170]. While CD44 has been thoroughly evaluated as a therapeutic target, its translation to the clinic has been minimal, likely due to several detectable CD44 isoforms that are generated through alternative splicing [171]. Interestingly, since an activation of xCT (a part of the glutathione redox cycle responsible for L-glutamate transport) was shown to be mediated by CD44 on tumor cells [172], an F18-labeled L-glutamate analogue is currently in clinical evaluation as a tracer for PET imaging (NCT02379377).
EpCAM, a transmembrane protein with homotypic cell–cell adhesion function, is widely expressed on many epithelial cancer cells. HCCs can be subdivided into EpCAM+ tumors with a progenitor-like gene expression pattern and EpCAM- tumors with a more mature phenotype. This was demonstrated in microarray analyses in a large cohort of HCC tumors, where stemness-associated genes, such as c-Kit and Wnt/β-catenin, were upregulated in EpCAM+ HCCs [126]. EpCAM expression in HCC depends directly on Wnt/β-catenin activity, and the downstream transcription factor Tcf-4 was shown to directly bind the EpCAM promotor [173]. Additional efforts have led to the discovery that ZFX, a zinc-finger transcription factor known to promote stemness in human stem cells, augments β-catenin nuclear translocation and Wnt activation in HCC CSCs, enhancing their stemness [174]. Furthermore, EpCAM represents an important therapeutic target, with EpCAM-targeting antibodies being assessed in several clinical trials. EpCAM is a tempting target due to its strong expression on tumor cells and on CSCs. Catumaxomab, an FDA-approved chimeric antibody that binds to antigens CD3 and EpCAM, is used in the treatment of malignant ascites and has shown promising effects in malignant pleural effusion [175]. Furthermore, CART-cells with specificity against EpCAM are used in clinical trials (e.g., NCT03013712). Likewise, activated cytokine induced killer (CIK) cells with EpCAM specificity are currently in development (NCT03146637).
CD90 is a glycoprotein of the immunoglobulin family of surface proteins that was linked to CSCs in HCC due to their ability to recreate tumors in serial transplantation experiments after isolation from the blood of HCC patients [30,176]. Interestingly, compared to EpCAM, CD90 seems to be expressed on different CSC populations. Yamashita T et al. demonstrated that while EpCAM appears to enrich for CSCs with a more epithelial gene expression profile, CD90+ HCC cells were found to be more mesenchymal. Accordingly, EpCAM+ cells revealed local tumorigenicity, while CD90+ cells efficiently formed lung metastases upon transplantation in mice. However, in co-cultures, these subpopulations positively interacted with CD90+ cells enhancing the motility of EpCAM+ cells in vitro through TGF-β signaling. The authors also showed high c-Kit expression in metastatic CD90+ cells, suggesting therapeutic potential for imatinib against CD90+ CSCs [177]. CyclinD1-dependent SMAD activation also appears to be essential for the maintenance and chemosensitivity of CD90+ CSCs in HCC. Specifically, Xia et al. demonstrated that >50% of xenotransplanted HCCs could be eliminated by SMAD inhibition [178]. It is important to note, however, that the used small molecule inhibitor SB431542 is a potent inhibitor of the TGF-β/Activin/NODAL pathway that inhibits ALK5, ALK4, and ALK7 and not SMADs directly. The link between CD90 and a more metastatic CSC phenotype was further explored in a study aimed at the identification of markers for circulating tumor stem cells. The authors showed that CXCR4 is an essential contributor to CD90+ mediated distant metastasis, as only CD90+CXCR4+ HCC cells developed metastases in NOD/SCID mice, whereas CD90−CXCR4−, CD90−CXCR4+ and CD90+CXCR4− cells failed to do so [179]. The use of CXCR4 as a marker to identify metastatic CSCs is in line with previous studies in pancreatic cancer, in which we showed that CD133+/CXCR4+ CSCs have exclusive tumorigenic and metastasis-initiating capacity [9]. CD90 expression negatively impacts prognosis [180], but targeting CD90 as a means of treating HCC is still at the experimental stage and no drugs targeting CD90 are approved currently. A very interesting study using magneto-liposomes (to induce local hyperthermia via magnetic fields) loaded with Fe3O4 and anti-CD90 antibodies to target Huh7-induced HCCs in mice resulted in tumor size reduction and showed specificity against CD90+ CSCs [181]. Similarly, using the “energy restriction mimetic” OSU-CG5, Chen WC et al. were able to show reduction of the CD90+ population in fresh liver tumor samples and repression of the tumor growth established with HCC cell lines with ectopic CD90 expression [147].
CD13 can serve as a CSC marker in HCC [182]; however, it is only expressed in a semi-quiescent CSC subpopulation. CD13+ cells are predominantly in G0-phase and were found to be enriched in the tumor periphery after treatment with doxorubicin or 5-FU. Functionally, CD13 protects cells from genotoxic damage by activating reactive oxygen species (ROS) scavenging systems and by supporting CSC self-renewal. Importantly, inhibition of CD13 using ubenimex led to a strong increase in therapeutic efficacy of 5-FU treatment of HCCs in xenograft models and hampered the ability of CSC to self-renew and initiate tumors [154], indicating sensitization of CSCs to therapy. The same authors linked TGF-β-induced EMT to CD13 upregulation, promoting CSC survival by abrogating ROS production [183]. Several experimental approaches have been designed to target CD13 in HCC, most of them showing promising efficacy, especially in combination with cytotoxic agents [184,185,186,187,188,189]. The above-mentioned ubenimex is a protease inhibitor blocking CD13 activity and has been evaluated in other diseases; i.e., squamous cell lung carcinoma, lymphedema and pulmonary hypertension. It is usually well-tolerated, but while effective against CD13+ CSCs in mouse models of HCC [154], clinical trials investigating ubenimex in HCC are still lacking.
OV-6 is the epitope for anti-OV-6 antibodies derived from hybridomas generated from BALB/c mice immunized with the oval cell antigen [190]. OV-6 was shown to be expressed in ductular and hepatocytic regenerates after liver injury in humans [191] and in HCC cell lines [192], and in primary HCCs [193]. OV-6 expression has been associated with CSC traits [194], and OV-6+ HCC correlated with poorer prognosis in tissue microarrays containing samples from 208 HCC patients [195]. While studies in esophageal carcinomas have found a link between OV-6 expression and β-catenin stabilization via ATG7, a similar link still needs to be demonstrated in HCC. However, OV6+ CSCs have already been described to have highly active Wnt/β-catenin signaling in HCC [196].
Zhao et al. identified isoform 5 of the cell surface calcium channel α2δ1 as a putative CSC marker in HCC in 2013 [149]. When α2δ1+ cells were subcutaneously injected in NOD/SCID mice, they showed increased tumorigenic potential in comparison with α2δ1− cells. Indeed, α2δ1+ cells expressed other CSCs markers, such as CD133, CD13 or EpCAM, further supporting this subpopulation of cells to be CSCs. Tissue sample analysis showed that α2δ1 could be found in most HCC samples and in surrounding tissue, the latter correlating with pathological factors, such as cirrhosis and shorter survival. Regarding the biological role of α2δ1 overexpression in CSCs, it was demonstrated that α2δ1 regulates calcium signaling, and intracellular calcium concentration and oscillation were shown to be higher in α2δ1+ cells. α2δ1 silencing leads to suppressed ERK1/2 phosphorylation, which suggests a role for calcium flux in CSC biology. This hypothesis was supported by ERK1/2 inhibition experiments [149]. Other types of cancer also contain a subset of α2δ1+ cells (small-cell lung cancer, laryngeal squamous cell carcinoma and gastric cancer), showing not only the same stem cell-like properties [197], but also the same chemoresistant phenotype [198] and overexpression of ERK1/2 [199]. In all cases, inhibition of α2δ1 has been proposed as a new promising therapeutic target for CSCs, including HCC CSCs [197,198,199,200].

3.3. Hepatitis B and C Viral Infections in HCC Initiation and CSC Maintenance

While the genetic origins of HCC are still not completely known, with no one set of driver mutations predominating as with other cancers, it is clear that HCC development is tightly linked to chronic liver damage and inflammation [201]. Thus, it is not surprising that infection with HBV or HCV has been determined as one of the leading risk factors for HCC development. As with alcohol-steatohepatitis-related HCC, cirrhosis due to chronic inflammation is likely to also occur before viral hepatitis-related HCC. However, other risk factors, such as alcohol consumption, obesity or diabetes, can synergize with chronic viral hepatic infections, promoting a multifactorial evolution towards HCC [202].
The link between HBV or HCV infection and HCC raises many questions with respect to the genesis of CSCs, such as whether hepatitis viruses promote the evolution of CSCs, whether these viruses preferentially replicate in CSCs, or whether viral products/replication can regulate stemness in cancer cells. It has been reported that cells with CSC markers—such as EpCAM—expressing stem-like transcriptional factors, such as Nanog or Oct4, also express the Hepatitis-B X-protein (HBx) of HBV, in either normal or truncated form [203,204]. HBx is a 16.5 KDa viral protein that apart from playing a role in the HBV life cycle, can activate mitogenic signaling cascades, altering expression of proliferation genes via transcription factors such as NF-κB, AP-1, AP-2, c-EBP and ATF/CREB [205]. HBx, however, is not the only HBV protein related to a stem-like phenotype. The viral PreS1 protein induces CD133, CD117 and CD90 expression in cancer cells, leading to an increase in sphere formation, migration, tumorigenesis and tumor growth in nude mice [206]. Other studies have suggested that HBV-positive cells have increased Wnt pathway activity, mediated by upregulation of EpCAM after loss of PCR2 function, leading to the expression of CSCs markers and resulting in poor prognosis [207]. The chronic viral-related inflammatory environment has also been proposed to be an important factor that could increase CSC proliferation due to cytokine-induced enhancement of Oct4/Nanog [208].
HCV infection/replication has been directly related to the acquisition of a stem-like phenotype in HCC. HCV subgenomic replicon replication has been linked to the enhancement of stem-like properties. When expressed in HCC cells, HCV subgenomic RNA replication enhanced the expression of DCAMKL-1, LGR5, CD133, AFP, CK-19, Lin28 and c-Myc, all of which have been associated with CSCs in HCC. Importantly, inhibition of the replicon reversed this phenotype, indicating that HCV replication and the production of non-structural HCV proteins can promote a stem-like state [209]. The latter are also involved in the acquisition of a CSC phenotype (as in HBV). Toll-like receptor 4, for example, is induced by the expression of the NS5A HCV protein in hepatocytes, and this induction correlates with the upregulation of Nanog and CD133 [210]. Finally, sphere formation assays with HCV-infected and non-infected HCC cell lines showed the increased sphere formation capacity of infected cells with an upregulation of stemness markers such as CD133, c-Kit, CD105, Sox2 and CD45 [211]. Moreover, microenvironmental factors and signaling cascades can influence HCV replication, and consequently, its link to promoting HCC CSCs. Shirasaki et al. showed that the osteopontin-CD44 axis in EpCAM+/CD44+ CSCs is important for mediating higher HCV replication in these cells through STAT1 reduction. The authors observed a significant difference compared to the EpCAM/CD44 non-CSC population, suggesting that HCV can replicate to higher levels in CSCs. As detailed above, this higher replication and/or increased production of viral proteins can induce or maintain the CSC state [212].
While the precise mechanism(s) behind how HBV and HCV promote the genesis and maintenance of CSCs is still not yet entirely understood, insights gained from the above-mentioned and future studies could ultimately be used to control hepatitis virus replication and simultaneously eliminate CSCs in HCC.

3.4. The CSC Niche in HCC: Ectopic Lymphoid Structures and Hepatic Stellate Cells

As detailed above, CSCs need a supportive microenvironment for their maintenance and survival. This niche generates a protective and growth-promoting milieu via cellular players, the extracellular matrix (ECM) and soluble factors. The niche is of special interest, since therapeutic interventions could potentially affect the CSC pool either directly or indirectly by elimination of the protective environment. The niche is thought to protect CSCs from chemotherapy, prevent apoptosis and maintain stemness. In the case of HCC, a structure called the “ectopic lymphoid structure (ELS)” or “tertiary lymphoid tissue” has been recognized to represent a putative microenvironment that promotes CSC growth and even the “education” of TICs into fully self-sufficient CSCs [213]. Because ELSs are associated with a more favorable prognosis in some malignancies, including carcinomas of the lung, skin, colon and breast [214,215,216], they were thought to be the anatomic correlate of an anti-cancer immune response. In breast cancer, ELS mainly contains follicular helper T cells; Th1, Th2 and Th17 effector memory cells; and regulatory T cells [216]. On the contrary, in HCC, anatomical detection of these structures was associated with an increased risk for late recurrence, and using a molecular detection method based on gene expression profiling, ELSs were associated with poor overall survival. Using liver specific expression of a constitutively active IKK-β variant in mice, activated hepatic NFκB signaling induced the formation of ELS after 7 months and of HCC after 20 months in 100% of the mice analyzed in the study. Interestingly, lymphocytes were essential for tumor formation, since a cross with a RAG−/− mouse with constitutive IKK-β activation resulted in a reversal of the HCC phenotype [213]. This observation is in line with the consensus that HCCs emerge in livers suffering from (non-)alcoholic steatohepatitis or viral hepatitis. Importantly, Svinarenko et al. were recently able to reproduce this signaling link in vivo using a conditional liver-specific IKK-β activation model with p53 loss [217].
Another characteristic of the CSC niche in HCC is the presence of a stromal cell population known as hepatic stellate cells (HSCs). These cells exhibit a myofibroblast-like phenotype in conditions of chronic liver injury. In cancer, these cells share many of the same functions as CAFs at the level of the CSC niche. It has been reported that HSCs secrete HGF or IL-6 that induce the expression of stemness markers in CSCs, and this increased expression is mediated through STAT3 signaling [218,219]. The investigation of the crosstalk between HSCs and CSCs resulted in the identification of the transcription factor Forkhead Box M1 (FOXM1) as a main player in HSC activation and maintenance of stemness in CSCs in vitro. FOXM1 inhibition in co-cultures of these two populations disrupted the crosstalk, whereas the overexpression of FOXM1 reversed the effects of the inhibition [220]. These lines of research highlight additional potential targets to interrupt the communication between HSCs and cancer (stem) cells, with the ultimate goal of targeting the CSC via multiple approaches.

4. Discussion

In the sections above, we review the current knowledge about CSCs in general and in the more specific context of HCC. Based on the CSC model, only by targeting and eliminating CSCs along with the more differentiated cells of the bulk tumor will we be able to achieve long-term curative responses and prevent metastatic spread. At the same time, CSCs display highly effective mechanisms of therapy resistance, such as the expression of drug efflux pumps, (semi-)quiescence and high ROS scavenging. Thus, it is unlikely that CSCs will be effectively targeted with classical cytostatic drugs. Furthermore, most HCC patients suffer from side effects associated with therapy, since HCC typically arises in severely damaged, cirrhotic livers. These patients suffer from chronic anemia and thrombocytopenia; are prone to infections; and often present with highly reduced (and therefore therapy-limiting) liver function and subsequently reduced dosage tolerance. One possible approach to reduce toxicity would be to utilize the inherent properties of CSCs. For example, inhibiting efflux pumps in CSCs could enhance their susceptibility to cytostatic drugs, reducing the necessary dose and thus increasing drug tolerance in the patient. For example, the calcium channel blocker verapamil showed efficacy in vitro, sensitizing SP cells to gemcitabine in pancreatic cancer [221], and has shown promising results in HCC [222,223]. ABC transporter inhibitors such as β-caryophyllene oxide have been shown to sensitize HCC cells to sorafenib by favoring its intracellular accumulation [224], but these compounds are still in development and their clinical utility is still unknown. Indeed, even third-generation inhibitors of p-glycoprotein drug transporters have not shown clinical efficacy [225]. Targeting CSC-specific surface markers or stemness-associated pathways for cancer therapy has led to several clinical trials that are currently ongoing to test highly specific molecules or antibodies that target CSCs. However, a key obstacle with these strategies is the expression of these markers or pathways on/in other stem cells, such as a hematopoietic stem cells or adult stem cells; another is the subsequent adverse effects of a targeted CSC therapy. In summary, the recognition of CSCs as a target sheds new light on potential therapeutic approaches in HCC, even though effective treatments against these highly aggressive cells still need to be developed.

5. Conclusions

Several lines of evidence point to the CSC as a key therapeutic target in HCC, as in many other tumors. Thus, this population must be successfully eliminated in order to prevent metastatic spread and tumor relapse, but due to their high therapy resistance and high cellular plasticity, effective means still have to be found to specifically identify these cells and to adequately target them, especially in HCC. To conclude on a promising note, the recent IMbrave150 trial showed impressive therapeutic effects of a combined anti-VEGF/anti-PD-L1 targeted therapy in HCC. In this trial, overall survival and progression free survival of HCC patients improved significantly and relevantly compared to sorafenib in first-line treatment, making this trial the first successful first-line trial in HCC since more than a decade (NCT03434379). Targeting CSCs could in part account for this positive outcome, since in many tumor types, a selective upregulation of PD-L1 has been observed on CSCs [226,227,228,229,230], enabling them to even induce apoptosis of T-cells in vitro. Furthermore, PD-L1 expression seems to be linked to EMT via ZEB-1 [231], which is highly expressed in HCC [232], closing the circle of CSC, EMT, therapy resistance and immune evasion. Several other mechanisms for immune evasion have been demonstrated in CSC, but available data are still sparse and in part conflicting [233]. MHC-I expression was lower in CSCs of colorectal cancer and glioblastoma patients [226,234,235]. Likewise, defects in the antigen processing machinery of CSC were observed [235]. In conclusion, these findings point to mechanisms of immune evasion that are highly active on CSC, and therefore, CSC depletion could have contributed to the success of the IMbrave150 trial. On the other hand, the single agent anti-PD-1 trial CheckMate 459 missed its primary endpoint of overall survival, which points to additional effects of the anti-VEGF agent and its possible interference with CSCs by mechanisms described elsewhere in this article.

Author Contributions

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

Funding

Work in the laboratory of P.C.H. is supported by a Max Eder Fellowship of the German Cancer Aid (111746), by a Collaborative Research Centre grant of the German Research Foundation (316249678–SFB 1279) and by a Hector Foundation Cancer Research grant (M65.1). Work in the laboratory of B.S.J. is supported by a Rámon y Cajal Merit Award (RYC-2012-12104) from the Ministerio de Economía y Competitividad, Spain. L.-A.S. is supported by the clinician scientist program of Ulm University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Greaves, M.; Maley, C.C. Clonal evolution in cancer. Nature 2012, 481, 306–313. [Google Scholar] [CrossRef]
  2. Zhao, B.; Wang, L.; Qiu, H.; Zhang, M.; Sun, L.; Peng, P.; Yu, Q.; Yuan, X. Mechanisms of resistance to anti-EGFR therapy in colorectal cancer. Oncotarget 2017, 8, 3980–4000. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Rycaj, K.; Tang, D.G. Cell-of-Origin of Cancer versus Cancer Stem Cells: Assays and Interpretations. Cancer Res. 2015, 75, 4003–4011. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Lapidot, T.; Sirard, C.; Vormoor, J.; Murdoch, B.; Hoang, T.; Caceres-Cortes, J.; Minden, M.; Paterson, B.; Caligiuri, M.A.; Dick, J.E. A cell initiating human acute myeloid leukaemia after transplantation into SCID mice. Nature 1994, 367, 645–648. [Google Scholar] [CrossRef] [PubMed]
  5. Batlle, E.; Clevers, H. Cancer stem cells revisited. Nat. Med. 2017, 23, 1124–1134. [Google Scholar] [CrossRef]
  6. GBD 2015 Disease and Injury Incidence and Prevalence Collaborators. Global, regional, and national life expectancy, all-cause mortality, and cause-specific mortality for 249 causes of death, 1980–2015: A systematic analysis for the Global Burden of Disease Study 2015. Lancet 2016, 388, 1459–1544. [Google Scholar] [CrossRef] [Green Version]
  7. 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]
  8. Li, C.; Heidt, D.G.; Dalerba, P.; Burant, C.F.; Zhang, L.; Adsay, V.; Wicha, M.; Clarke, M.F.; Simeone, D.M. Identification of pancreatic cancer stem cells. Cancer Res. 2007, 67, 1030–1037. [Google Scholar] [CrossRef] [Green Version]
  9. Hermann, P.C.; Huber, S.L.; Herrler, T.; Aicher, A.; Ellwart, J.W.; Guba, M.; Bruns, C.J.; Heeschen, C. Distinct populations of cancer stem cells determine tumor growth and metastatic activity in human pancreatic cancer. Cell Stem Cell 2007, 1, 313–323. [Google Scholar] [CrossRef] [Green Version]
  10. Kim, C.F.; Jackson, E.L.; Woolfenden, A.E.; Lawrence, S.; Babar, I.; Vogel, S.; Crowley, D.; Bronson, R.T.; Jacks, T. Identification of bronchioalveolar stem cells in normal lung and lung cancer. Cell 2005, 121, 823–835. [Google Scholar] [CrossRef] [Green Version]
  11. O’Brien, C.A.; Pollett, A.; Gallinger, S.; Dick, J.E. A human colon cancer cell capable of initiating tumour growth in immunodeficient mice. Nature 2007, 445, 106–110. [Google Scholar] [CrossRef] [PubMed]
  12. Collins, A.T.; Berry, P.A.; Hyde, C.; Stower, M.J.; Maitland, N.J. Prospective identification of tumorigenic prostate cancer stem cells. Cancer Res. 2005, 65, 10946–10951. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Szotek, P.P.; Pieretti-Vanmarcke, R.; Masiakos, P.T.; Dinulescu, D.M.; Connolly, D.; Foster, R.; Dombkowski, D.; Preffer, F.; Maclaughlin, D.T.; Donahoe, P.K. Ovarian cancer side population defines cells with stem cell-like characteristics and Mullerian Inhibiting Substance responsiveness. Proc. Natl. Acad. Sci. USA 2006, 103, 11154–11159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Ginestier, C.; Hur, M.H.; Charafe-Jauffret, E.; Monville, F.; Dutcher, J.; Brown, M.; Jacquemier, J.; Viens, P.; Kleer, C.G.; Liu, S.; et al. ALDH1 is a marker of normal and malignant human mammary stem cells and a predictor of poor clinical outcome. Cell Stem Cell 2007, 1, 555–567. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Singh, S.K.; Clarke, I.D.; Terasaki, M.; Bonn, V.E.; Hawkins, C.; Squire, J.; Dirks, P.B. Identification of a cancer stem cell in human brain tumors. Cancer Res. 2003, 63, 5821–5828. [Google Scholar] [PubMed]
  16. Clarke, M.; Dick, J.; Dirks, P.; Eaves, C.; Jamieson, C.; Jones, D.; Visvader, J.; Weissman, I.; Wahl, G. Cancer Stem Cells--Perspectives on Current Status and Future Directions: AACR Workshop on Cancer Stem Cells. Cancer Res. 2006, 66, 9339–9344. [Google Scholar] [CrossRef] [Green Version]
  17. de Sousa e Melo, F.; Kurtova, A.V.; Harnoss, J.M.; Kljavin, N.; Hoeck, J.D.; Hung, J.; Anderson, J.E.; Storm, E.E.; Modrusan, Z.; Koeppen, H.; et al. A distinct role for Lgr5(+) stem cells in primary and metastatic colon cancer. Nature 2017, 543, 676–680. [Google Scholar] [CrossRef]
  18. Shimokawa, M.; Ohta, Y.; Nishikori, S.; Matano, M.; Takano, A.; Fujii, M.; Date, S.; Sugimoto, S.; Kanai, T.; Sato, T. Visualization and targeting of LGR5(+) human colon cancer stem cells. Nature 2017, 545, 187–192. [Google Scholar] [CrossRef]
  19. Chen, J.; Li, Y.; Yu, T.S.; McKay, R.M.; Burns, D.K.; Kernie, S.G.; Parada, L.F. A restricted cell population propagates glioblastoma growth after chemotherapy. Nature 2012, 488, 522–526. [Google Scholar] [CrossRef] [Green Version]
  20. Steinbichler, T.B.; Savic, D.; Dudas, J.; Kvitsaridze, I.; Skvortsov, S.; Riechelmann, H.; Skvortsova, I.I. Cancer stem cells and their unique role in metastatic spread. Semin. Cancer Biol. 2019. [Google Scholar] [CrossRef]
  21. Chaffer, C.L.; Weinberg, R.A. A perspective on cancer cell metastasis. Science 2011, 331, 1559–1564. [Google Scholar] [CrossRef] [PubMed]
  22. Todaro, M.; Gaggianesi, M.; Catalano, V.; Benfante, A.; Iovino, F.; Biffoni, M.; Apuzzo, T.; Sperduti, I.; Volpe, S.; Cocorullo, G.; et al. CD44v6 is a marker of constitutive and reprogrammed cancer stem cells driving colon cancer metastasis. Cell Stem Cell 2014, 14, 342–356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Beck, B.; Lapouge, G.; Rorive, S.; Drogat, B.; Desaedelaere, K.; Delafaille, S.; Dubois, C.; Salmon, I.; Willekens, K.; Marine, J.C.; et al. Different levels of Twist1 regulate skin tumor initiation, stemness, and progression. Cell Stem Cell 2015, 16, 67–79. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Pang, R.; Law, W.L.; Chu, A.C.; Poon, J.T.; Lam, C.S.; Chow, A.K.; Ng, L.; Cheung, L.W.; Lan, X.R.; Lan, H.Y.; et al. A subpopulation of CD26+ cancer stem cells with metastatic capacity in human colorectal cancer. Cell Stem Cell 2010, 6, 603–615. [Google Scholar] [CrossRef] [Green Version]
  25. Dalerba, P.; Clarke, M.F. Cancer stem cells and tumor metastasis: First steps into uncharted territory. Cell Stem Cell 2007, 1, 241–242. [Google Scholar] [CrossRef] [Green Version]
  26. Goodell, M.A.; Brose, K.; Paradis, G.; Conner, A.S.; Mulligan, R.C. Isolation and functional properties of murine hematopoietic stem cells that are replicating in vivo. J. Exp. Med. 1996, 183, 1797–1806. [Google Scholar] [CrossRef] [Green Version]
  27. Bao, S.; Wu, Q.; Mclendon, R.; Hao, Y.; Shi, Q.; Hjelmeland, A.; Dewhirst, M.; Bigner, D.; Rich, J. Glioma stem cells promote radioresistance by preferential activation of the DNA damage response. Nature 2006, 444, 756–760. [Google Scholar] [CrossRef]
  28. Blanpain, C.; Mohrin, M.; Sotiropoulou, P.A.; Passegue, E. DNA-damage response in tissue-specific and cancer stem cells. Cell Stem Cell 2011, 8, 16–29. [Google Scholar] [CrossRef]
  29. Todaro, M.; Alea, M.; Distefano, A.; Cammareri, P.; Vermeulen, L.; Iovino, F.; Tripodo, C.; Russo, A.; Gulotta, G.; Medema, J. Colon Cancer Stem Cells Dictate Tumor Growth and Resist Cell Death by Production of Interleukin-4. Cell Stem Cell 2007, 1, 389–402. [Google Scholar] [CrossRef] [Green Version]
  30. Yang, Z.F.; Ho, D.W.; Ng, M.N.; Lau, C.K.; Yu, W.C.; Ngai, P.; Chu, P.W.; Lam, C.T.; Poon, R.T.; Fan, S.T. Significance of CD90+ cancer stem cells in human liver cancer. Cancer Cell 2008, 13, 153–166. [Google Scholar] [CrossRef] [Green Version]
  31. Hirschmann-Jax, C.; Foster, A.E.; Wulf, G.G.; Nuchtern, J.G.; Jax, T.W.; Gobel, U.; Goodell, M.A.; Brenner, M.K. A distinct “side population” of cells with high drug efflux capacity in human tumor cells. Proc. Natl. Acad. Sci. USA 2004, 101, 14228–14233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Marcato, P.; Dean, C.A.; Giacomantonio, C.A.; Lee, P.W. Aldehyde dehydrogenase: Its role as a cancer stem cell marker comes down to the specific isoform. Cell Cycle 2011, 10, 1378–1384. [Google Scholar] [CrossRef] [PubMed]
  33. Mizuno, T.; Suzuki, N.; Makino, H.; Furui, T.; Morii, E.; Aoki, H.; Kunisada, T.; Yano, M.; Kuji, S.; Hirashima, Y.; et al. Cancer stem-like cells of ovarian clear cell carcinoma are enriched in the ALDH-high population associated with an accelerated scavenging system in reactive oxygen species. Gynecol. Oncol. 2015, 137, 299–305. [Google Scholar] [CrossRef] [PubMed]
  34. Huber, M.A.; Azoitei, N.; Baumann, B.; Grunert, S.; Sommer, A.; Pehamberger, H.; Kraut, N.; Beug, H.; Wirth, T. NF-kappaB is essential for epithelial-mesenchymal transition and metastasis in a model of breast cancer progression. J. Clin. Investig. 2004, 114, 569–581. [Google Scholar] [CrossRef] [Green Version]
  35. Nomura, A.; Banerjee, S.; Chugh, R.; Dudeja, V.; Yamamoto, M.; Vickers, S.M.; Saluja, A.K. CD133 initiates tumors, induces epithelial-mesenchymal transition and increases metastasis in pancreatic cancer. Oncotarget 2015, 6, 8313–8322. [Google Scholar] [CrossRef] [Green Version]
  36. Nomura, A.; Gupta, V.K.; Dauer, P.; Sharma, N.S.; Dudeja, V.; Merchant, N.; Saluja, A.K.; Banerjee, S. NFkappaB-Mediated Invasiveness in CD133(+) Pancreatic TICs Is Regulated by Autocrine and Paracrine Activation of IL1 Signaling. Mol. Cancer Res. 2018, 16, 162–172. [Google Scholar] [CrossRef] [Green Version]
  37. Wright, M.H.; Calcagno, A.M.; Salcido, C.D.; Carlson, M.D.; Ambudkar, S.V.; Varticovski, L. Brca1 breast tumors contain distinct CD44+/CD24− and CD133+ cells with cancer stem cell characteristics. Breast Cancer Res. 2008, 10, R10. [Google Scholar] [CrossRef] [Green Version]
  38. Ma, S.; Chan, K.; Hu, L.; Lee, T.; Wo, J.; Ng, I.; Zheng, B.; Guan, X. Identification and Characterization of Tumorigenic Liver Cancer Stem/Progenitor Cells. Gastroenterology 2007, 132, 2542–2556. [Google Scholar] [CrossRef]
  39. Harpstrite, S.E.; Gu, H.; Natarajan, R.; Sharma, V. Interrogation of multidrug resistance (MDR1) P-glycoprotein (ABCB1) expression in human pancreatic carcinoma cells: Correlation of 99mTc-Sestamibi uptake with western blot analysis. Nucl. Med. Commun. 2014, 35, 1067–1070. [Google Scholar] [CrossRef] [Green Version]
  40. Miranda-Lorenzo, I.; Dorado, J.; Lonardo, E.; Alcala, S.; Serrano, A.G.; Clausell-Tormos, J.; Cioffi, M.; Megias, D.; Zagorac, S.; Balic, A.; et al. Intracellular autofluorescence: A biomarker for epithelial cancer stem cells. Nat. Methods 2014, 11, 1161–1169. [Google Scholar] [CrossRef]
  41. Plaks, V.; Kong, N.; Werb, Z. The cancer stem cell niche: How essential is the niche in regulating stemness of tumor cells? Cell Stem Cell 2015, 16, 225–238. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Chan, L.-H.; Luk, S.T.; Ma, S. Turning Hepatic Cancer Stem Cells Inside Out—A Deeper Understanding through Multiple Perspectives. Mol. Cells 2015, 38, 202–209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Vermeulen, L.; Felipe De Sousa, E.M.; van der Heijden, M.; Cameron, K.; de Jong, J.H.; Borovski, T.; Tuynman, J.B.; Todaro, M.; Merz, C.; Rodermond, H.; et al. Wnt activity defines colon cancer stem cells and is regulated by the microenvironment. Nat. Cell Biol. 2010, 12, 468–476. [Google Scholar] [CrossRef] [PubMed]
  44. Junttila, M.R.; de Sauvage, F.J. Influence of tumour micro-environment heterogeneity on therapeutic response. Nature 2013, 501, 346–354. [Google Scholar] [CrossRef] [PubMed]
  45. Iliopoulos, D.; Hirsch, H.A.; Wang, G.; Struhl, K. Inducible formation of breast cancer stem cells and their dynamic equilibrium with non-stem cancer cells via IL6 secretion. Proc. Natl. Acad. Sci. USA 2011, 108, 1397. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Tsuyada, A.; Chow, A.; Wu, J.; Somlo, G.; Chu, P.; Loera, S.; Luu, T.; Li, A.X.; Wu, X.; Ye, W.; et al. CCL2 Mediates Cross-talk between Cancer Cells and Stromal Fibroblasts That Regulates Breast Cancer Stem Cells. Cancer Res. 2012, 72, 2768. [Google Scholar] [CrossRef] [Green Version]
  47. Alguacil-Nunez, C.; Ferrer-Ortiz, I.; Garcia-Verdu, E.; Lopez-Pirez, P.; Llorente-Cortijo, I.M.; Sainz, B., Jr. Current perspectives on the crosstalk between lung cancer stem cells and cancer-associated fibroblasts. Crit. Rev. Oncol. Hematol. 2018, 125, 102–110. [Google Scholar] [CrossRef] [Green Version]
  48. Nombela-Arrieta, C.; Ritz, J.; Silberstein, L.E. The elusive nature and function of mesenchymal stem cells. Nat. Rev. Mol. Cell Biol. 2011, 12, 126–131. [Google Scholar] [CrossRef] [Green Version]
  49. Mishra, P.J.; Mishra, P.J.; Humeniuk, R.; Medina, D.J.; Alexe, G.; Mesirov, J.P.; Ganesan, S.; Glod, J.W.; Banerjee, D. Carcinoma-Associated Fibroblast–Like Differentiation of Human Mesenchymal Stem Cells. Cancer Res. 2008, 68, 4331. [Google Scholar] [CrossRef] [Green Version]
  50. Petty, A.J.; Yang, Y. Tumor-Associated Macrophages in Hematologic Malignancies: New Insights and Targeted Therapies. Cells 2019, 8, 1526. [Google Scholar] [CrossRef] [Green Version]
  51. Timaner, M.; Letko-Khait, N.; Kotsofruk, R.; Benguigui, M.; Beyar-Katz, O.; Rachman-Tzemah, C.; Raviv, Z.; Bronshtein, T.; Machluf, M.; Shaked, Y. Therapy-Educated Mesenchymal Stem Cells Enrich for Tumor-Initiating Cells. Cancer Res. 2018, 78, 1253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Roodhart, J.M.; Daenen, L.G.; Stigter, E.C.; Prins, H.J.; Gerrits, J.; Houthuijzen, J.M.; Gerritsen, M.G.; Schipper, H.S.; Backer, M.J.; van Amersfoort, M.; et al. Mesenchymal Stem Cells Induce Resistance to Chemotherapy through the Release of Platinum-Induced Fatty Acids. Cancer Cell 2011, 20, 370–383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Sica, A.; Porta, C.; Amadori, A.; Pastò, A. Tumor-associated myeloid cells as guiding forces of cancer cell stemness. Cancer Immunol. Immunother. 2017, 66, 1025–1036. [Google Scholar] [CrossRef] [PubMed]
  54. Won, C.; Kim, B.-H.; Yi, E.H.; Choi, K.-J.; Kim, E.-K.; Jeong, J.-M.; Lee, J.-H.; Jang, J.-J.; Yoon, J.-H.; Jeong, W.-I.; et al. Signal transducer and activator of transcription 3-mediated CD133 up-regulation contributes to promotion of hepatocellular carcinoma. Hepatology 2015, 62, 1160–1173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Mitchem, J.B.; Brennan, D.J.; Knolhoff, B.L.; Belt, B.A.; Zhu, Y.; Sanford, D.E.; Belaygorod, L.; Carpenter, D.; Collins, L.; Piwnica-Worms, D.; et al. Targeting Tumor-Infiltrating Macrophages Decreases Tumor-Initiating Cells, Relieves Immunosuppression, and Improves Chemotherapeutic Responses. Cancer Res. 2013, 73, 1128. [Google Scholar] [CrossRef] [PubMed]
  56. Otvos, B.; Silver, D.J.; Mulkearns-Hubert, E.E.; Alvarado, A.G.; Turaga, S.M.; Sorensen, M.D.; Rayman, P.; Flavahan, W.A.; Hale, J.S.; Stoltz, K.; et al. Cancer Stem Cell-Secreted Macrophage Migration Inhibitory Factor Stimulates Myeloid Derived Suppressor Cell Function and Facilitates Glioblastoma Immune Evasion. Stem Cells 2016, 34, 2026–2039. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Ping, Y.F.; Bian, X.W. Consice review: Contribution of cancer stem cells to neovascularization. Stem Cells 2011, 29, 888–894. [Google Scholar] [CrossRef]
  58. Steinbichler, T.B.; Dudas, J.; Skvortsov, S.; Ganswindt, U.; Riechelmann, H.; Skvortsova, I.I. Therapy resistance mediated by cancer stem cells. Semin. Cancer Biol. 2018, 53, 156–167. [Google Scholar] [CrossRef]
  59. Li, S.; Li, Q. Cancer stem cells and tumor metastasis (Review). Int. J. Oncol. 2014, 44, 1806–1812. [Google Scholar] [CrossRef] [Green Version]
  60. Gilkes, D.M.; Semenza, G.L. Role of hypoxia-inducible factors in breast cancer metastasis. Future Oncol. 2013, 9, 1623–1636. [Google Scholar] [CrossRef] [Green Version]
  61. Lee, S.Y.; Ju, M.K.; Jeon, H.M.; Lee, Y.J.; Kim, C.H.; Park, H.G.; Han, S.I.; Kang, H.S. Oncogenic Metabolism Acts as a Prerequisite Step for Induction of Cancer Metastasis and Cancer Stem Cell Phenotype. Oxid. Med. Cell Longev. 2018, 2018, 1027453. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Bao, S.; Wu, Q.; Sathornsumetee, S.; Hao, Y.; Li, Z.; Hjelmeland, A.B.; Shi, Q.; McLendon, R.E.; Bigner, D.D.; Rich, J.N. Stem cell-like glioma cells promote tumor angiogenesis through vascular endothelial growth factor. Cancer Res. 2006, 66, 7843–7848. [Google Scholar] [CrossRef] [Green Version]
  63. Folkins, C.; Shaked, Y.; Man, S.; Tang, T.; Lee, C.R.; Zhu, Z.; Hoffman, R.M.; Kerbel, R.S. Glioma tumor stem-like cells promote tumor angiogenesis and vasculogenesis via vascular endothelial growth factor and stromal-derived factor 1. Cancer Res. 2009, 69, 7243–7251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Zhu, A.X.; Kang, Y.K.; Yen, C.J.; Finn, R.S.; Galle, P.R.; Llovet, J.M.; Assenat, E.; Brandi, G.; Pracht, M.; Lim, H.Y.; et al. Ramucirumab after sorafenib in patients with advanced hepatocellular carcinoma and increased alpha-fetoprotein concentrations (REACH-2): A randomised, double-blind, placebo-controlled, phase 3 trial. Lancet Oncol. 2019, 20, 282–296. [Google Scholar] [CrossRef]
  65. Brossa, A.; Grange, C.; Mancuso, L.; Annaratone, L.; Satolli, M.A.; Mazzone, M.; Camussi, G.; Bussolati, B. Sunitinib but not VEGF blockade inhibits cancer stem cell endothelial differentiation. Oncotarget 2015, 6, 11295–11309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Ramezani, S.; Vousooghi, N.; Kapourchali, F.R.; Hadjighasem, M.; Hayat, P.; Amini, N.; Joghataei, M.T. Rolipram potentiates bevacizumab-induced cell death in human glioblastoma stem-like cells. Life Sci. 2017, 173, 11–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Fernandez-Cortes, M.; Delgado-Bellido, D.; Oliver, F.J. Vasculogenic Mimicry: Become an Endothelial Cell “But Not So Much”. Front. Oncol. 2019, 9, 803. [Google Scholar] [CrossRef] [Green Version]
  68. Wang, R.; Chadalavada, K.; Wilshire, J.; Kowalik, U.; Hovinga, K.E.; Geber, A.; Fligelman, B.; Leversha, M.; Brennan, C.; Tabar, V. Glioblastoma stem-like cells give rise to tumour endothelium. Nature 2010, 468, 829–833. [Google Scholar] [CrossRef]
  69. Shen, R.; Ye, Y.; Chen, L.; Yan, Q.; Barsky, S.H.; Gao, J.X. Precancerous stem cells can serve as tumor vasculogenic progenitors. PLoS ONE 2008, 3, e1652. [Google Scholar] [CrossRef] [Green Version]
  70. Bussolati, B.; Grange, C.; Sapino, A.; Camussi, G. Endothelial cell differentiation of human breast tumour stem/progenitor cells. J. Cell Mol. Med. 2009, 13, 309–319. [Google Scholar] [CrossRef] [Green Version]
  71. Ricci-Vitiani, L.; Pallini, R.; Biffoni, M.; Todaro, M.; Invernici, G.; Cenci, T.; Maira, G.; Parati, E.A.; Stassi, G.; Larocca, L.M.; et al. Tumour vascularization via endothelial differentiation of glioblastoma stem-like cells. Nature 2010, 468, 824–828. [Google Scholar] [CrossRef] [PubMed]
  72. Baccelli, I.; Schneeweiss, A.; Riethdorf, S.; Stenzinger, A.; Schillert, A.; Vogel, V.; Klein, C.; Saini, M.; Bauerle, T.; Wallwiener, M.; et al. Identification of a population of blood circulating tumor cells from breast cancer patients that initiates metastasis in a xenograft assay. Nat. Biotechnol. 2013, 31, 539–544. [Google Scholar] [CrossRef] [PubMed]
  73. Mohme, M.; Riethdorf, S.; Pantel, K. Circulating and disseminated tumour cells—Mechanisms of immune surveillance and escape. Nat. Rev. Clin. Oncol. 2017, 14, 155–167. [Google Scholar] [CrossRef] [PubMed]
  74. Chambers, A.F.; Groom, A.C.; MacDonald, I.C. Dissemination and growth of cancer cells in metastatic sites. Nat. Rev. Cancer 2002, 2, 563–572. [Google Scholar] [CrossRef] [PubMed]
  75. Malanchi, I.; Santamaria-Martinez, A.; Susanto, E.; Peng, H.; Lehr, H.A.; Delaloye, J.F.; Huelsken, J. Interactions between cancer stem cells and their niche govern metastatic colonization. Nature 2011, 481, 85–89. [Google Scholar] [CrossRef]
  76. Aguirre-Ghiso, J.A. Models, mechanisms and clinical evidence for cancer dormancy. Nat. Rev. Cancer 2007, 7, 834–846. [Google Scholar] [CrossRef] [Green Version]
  77. Labelle, M.; Begum, S.; Hynes, R.O. Direct signaling between platelets and cancer cells induces an epithelial-mesenchymal-like transition and promotes metastasis. Cancer Cell 2011, 20, 576–590. [Google Scholar] [CrossRef] [Green Version]
  78. Palumbo, J.S.; Talmage, K.E.; Massari, J.V.; La Jeunesse, C.M.; Flick, M.J.; Kombrinck, K.W.; Jirouskova, M.; Degen, J.L. Platelets and fibrin(ogen) increase metastatic potential by impeding natural killer cell-mediated elimination of tumor cells. Blood 2005, 105, 178–185. [Google Scholar] [CrossRef] [Green Version]
  79. Wang, S.; Zhang, Y.; Cong, W.; Liu, J.; Zhang, Y.; Fan, H.; Xu, Y.; Lin, H. Breast cancer stem-like cells can promote metastasis by activating platelets and down-regulating antitumor activity of natural killer cells. J. Tradit. Chin. Med. 2016, 36, 530–537. [Google Scholar] [CrossRef] [Green Version]
  80. Duda, D.G.; Duyverman, A.M.; Kohno, M.; Snuderl, M.; Steller, E.J.; Fukumura, D.; Jain, R.K. Malignant cells facilitate lung metastasis by bringing their own soil. Proc. Natl. Acad. Sci. USA 2010, 107, 21677–21682. [Google Scholar] [CrossRef] [Green Version]
  81. Ao, Z.; Shah, S.H.; Machlin, L.M.; Parajuli, R.; Miller, P.C.; Rawal, S.; Williams, A.J.; Cote, R.J.; Lippman, M.E.; Datar, R.H.; et al. Identification of Cancer-Associated Fibroblasts in Circulating Blood from Patients with Metastatic Breast Cancer. Cancer Res. 2015, 75, 4681–4687. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Denes, V.; Lakk, M.; Makarovskiy, A.; Jakso, P.; Szappanos, S.; Graf, L.; Mandel, L.; Karadi, I.; Geck, P. Metastasis blood test by flow cytometry: In vivo cancer spheroids and the role of hypoxia. Int. J. Cancer 2015, 136, 1528–1536. [Google Scholar] [CrossRef] [PubMed]
  83. Fidler, I.J. Selection of successive tumour lines for metastasis. Nat. New. Biol. 1973, 242, 148–149. [Google Scholar] [CrossRef] [PubMed]
  84. Peinado, H.; Aleckovic, M.; Lavotshkin, S.; Matei, I.; Costa-Silva, B.; Moreno-Bueno, G.; Hergueta-Redondo, M.; Williams, C.; Garcia-Santos, G.; Ghajar, C.; et al. Melanoma exosomes educate bone marrow progenitor cells toward a pro-metastatic phenotype through MET. Nat. Med. 2012, 18, 883–891. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Fong, M.Y.; Zhou, W.; Liu, L.; Alontaga, A.Y.; Chandra, M.; Ashby, J.; Chow, A.; O’Connor, S.T.; Li, S.; Chin, A.R.; et al. Breast-cancer-secreted miR-122 reprograms glucose metabolism in premetastatic niche to promote metastasis. Nat. Cell. Biol. 2015, 17, 183–194. [Google Scholar] [CrossRef] [Green Version]
  86. Hoshino, A.; Costa-Silva, B.; Shen, T.L.; Rodrigues, G.; Hashimoto, A.; Tesic Mark, M.; Molina, H.; Kohsaka, S.; Di Giannatale, A.; Ceder, S.; et al. Tumour exosome integrins determine organotropic metastasis. Nature 2015, 527, 329–335. [Google Scholar] [CrossRef] [Green Version]
  87. Bos, P.D.; Zhang, X.H.; Nadal, C.; Shu, W.; Gomis, R.R.; Nguyen, D.X.; Minn, A.J.; van de Vijver, M.J.; Gerald, W.L.; Foekens, J.A.; et al. Genes that mediate breast cancer metastasis to the brain. Nature 2009, 459, 1005–1009. [Google Scholar] [CrossRef]
  88. Minn, A.J.; Gupta, G.P.; Siegel, P.M.; Bos, P.D.; Shu, W.; Giri, D.D.; Viale, A.; Olshen, A.B.; Gerald, W.L.; Massague, J. Genes that mediate breast cancer metastasis to lung. Nature 2005, 436, 518–524. [Google Scholar] [CrossRef]
  89. Iqbal, W.; Alkarim, S.; AlHejin, A.; Mukhtar, H.; Saini, K.S. Targeting signal transduction pathways of cancer stem cells for therapeutic opportunities of metastasis. Oncotarget 2016, 7, 76337–76353. [Google Scholar] [CrossRef] [Green Version]
  90. Zhang, X.H.; Wang, Q.; Gerald, W.; Hudis, C.A.; Norton, L.; Smid, M.; Foekens, J.A.; Massague, J. Latent bone metastasis in breast cancer tied to Src-dependent survival signals. Cancer Cell 2009, 16, 67–78. [Google Scholar] [CrossRef] [Green Version]
  91. Deneve, E.; Riethdorf, S.; Ramos, J.; Nocca, D.; Coffy, A.; Daures, J.P.; Maudelonde, T.; Fabre, J.M.; Pantel, K.; Alix-Panabieres, C. Capture of viable circulating tumor cells in the liver of colorectal cancer patients. Clin. Chem. 2013, 59, 1384–1392. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Al-Mehdi, A.B.; Tozawa, K.; Fisher, A.B.; Shientag, L.; Lee, A.; Muschel, R.J. Intravascular origin of metastasis from the proliferation of endothelium-attached tumor cells: A new model for metastasis. Nat. Med. 2000, 6, 100–102. [Google Scholar] [CrossRef] [PubMed]
  93. Fausto, N.; Campbell, J.S.; Riehle, K.J. Liver regeneration. Hepatology 2006, 43, S45–S53. [Google Scholar] [CrossRef] [PubMed]
  94. Overturf, K.; al-Dhalimy, M.; Ou, C.N.; Finegold, M.; Grompe, M. Serial transplantation reveals the stem-cell-like regenerative potential of adult mouse hepatocytes. Am. J. Pathol. 1997, 151, 1273–1280. [Google Scholar]
  95. Overturf, K.; Al-Dhalimy, M.; Finegold, M.; Grompe, M. The repopulation potential of hepatocyte populations differing in size and prior mitotic expansion. Am. J. Pathol. 1999, 155, 2135–2143. [Google Scholar] [CrossRef] [Green Version]
  96. Rhim, J.A.; Sandgren, E.P.; Degen, J.L.; Palmiter, R.D.; Brinster, R.L. Replacement of diseased mouse liver by hepatic cell transplantation. Science 1994, 263, 1149–1152. [Google Scholar] [CrossRef]
  97. Sandgren, E.P.; Palmiter, R.D.; Heckel, J.L.; Daugherty, C.C.; Brinster, R.L.; Degen, J.L. Complete hepatic regeneration after somatic deletion of an albumin-plasminogen activator transgene. Cell 1991, 66, 245–256. [Google Scholar] [CrossRef]
  98. Yao, Z.; Mishra, L. Cancer stem cells and hepatocellular carcinoma. Cancer Biol. Ther. 2009, 8, 1691–1698. [Google Scholar] [CrossRef] [Green Version]
  99. Mishra, L.; Banker, T.; Murray, J.; Byers, S.; Thenappan, A.; He, A.R.; Shetty, K.; Johnson, L.; Reddy, E.P. Liver stem cells and hepatocellular carcinoma. Hepatology 2009, 49, 318–329. [Google Scholar] [CrossRef] [Green Version]
  100. Christ, B.; Pelz, S. Implication of hepatic stem cells in functional liver repopulation. Cytom. A 2013, 83, 90–102. [Google Scholar] [CrossRef]
  101. Jors, S.; Jeliazkova, P.; Ringelhan, M.; Thalhammer, J.; Durl, S.; Ferrer, J.; Sander, M.; Heikenwalder, M.; Schmid, R.M.; Siveke, J.T.; et al. Lineage fate of ductular reactions in liver injury and carcinogenesis. J. Clin. Investig. 2015, 125, 2445–2457. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Sell, S. Heterogeneity and plasticity of hepatocyte lineage cells. Hepatology 2001, 33, 738–750. [Google Scholar] [CrossRef] [PubMed]
  103. Alison, M.R.; Vig, P.; Russo, F.; Bigger, B.W.; Amofah, E.; Themis, M.; Forbes, S. Hepatic stem cells: From inside and outside the liver? Cell Prolif. 2004, 37, 1–21. [Google Scholar] [CrossRef] [PubMed]
  104. Fausto, N. Liver regeneration and repair: Hepatocytes, progenitor cells, and stem cells. Hepatology 2004, 39, 1477–1487. [Google Scholar] [CrossRef]
  105. Shafritz, D.A.; Oertel, M.; Menthena, A.; Nierhoff, D.; Dabeva, M.D. Liver stem cells and prospects for liver reconstitution by transplanted cells. Hepatology 2006, 43, S89–S98. [Google Scholar] [CrossRef]
  106. Michalopoulos, G.K.; Barua, L.; Bowen, W.C. Transdifferentiation of rat hepatocytes into biliary cells after bile duct ligation and toxic biliary injury. Hepatology 2005, 41, 535–544. [Google Scholar] [CrossRef]
  107. Li, W.C.; Horb, M.E.; Tosh, D.; Slack, J.M. In vitro transdifferentiation of hepatoma cells into functional pancreatic cells. Mech. Dev. 2005, 122, 835–847. [Google Scholar] [CrossRef]
  108. Tarlow, B.D.; Finegold, M.J.; Grompe, M. Clonal tracing of Sox9+ liver progenitors in mouse oval cell injury. Hepatology 2014, 60, 278–289. [Google Scholar] [CrossRef] [Green Version]
  109. Schmelzer, E.; Wauthier, E.; Reid, L.M. The phenotypes of pluripotent human hepatic progenitors. Stem. Cells 2006, 24, 1852–1858. [Google Scholar] [CrossRef]
  110. Ang, C.H.; Hsu, S.H.; Guo, F.; Tan, C.T.; Yu, V.C.; Visvader, J.E.; Chow, P.K.H.; Fu, N.Y. Lgr5(+) pericentral hepatocytes are self-maintained in normal liver regeneration and susceptible to hepatocarcinogenesis. Proc. Natl. Acad. Sci. USA 2019, 116, 19530–19540. [Google Scholar] [CrossRef] [Green Version]
  111. Gouw, A.S.; Clouston, A.D.; Theise, N.D. Ductular reactions in human liver: Diversity at the interface. Hepatology 2011, 54, 1853–1863. [Google Scholar] [CrossRef] [PubMed]
  112. Riehle, K.J.; Dan, Y.Y.; Campbell, J.S.; Fausto, N. New concepts in liver regeneration. J. Gastroenterol. Hepatol. 2011, 26 (Suppl. 1), 203–212. [Google Scholar] [CrossRef] [Green Version]
  113. 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] [PubMed]
  114. Dorrell, C.; Erker, L.; Schug, J.; Kopp, J.L.; Canaday, P.S.; Fox, A.J.; Smirnova, O.; Duncan, A.W.; Finegold, M.J.; Sander, M.; et al. Prospective isolation of a bipotential clonogenic liver progenitor cell in adult mice. Genes. Dev. 2011, 25, 1193–1203. [Google Scholar] [CrossRef] [Green Version]
  115. Furuyama, K.; Kawaguchi, Y.; Akiyama, H.; Horiguchi, M.; Kodama, S.; Kuhara, T.; Hosokawa, S.; Elbahrawy, A.; Soeda, T.; Koizumi, M.; et al. Continuous cell supply from a Sox9-expressing progenitor zone in adult liver, exocrine pancreas and intestine. Nat. Genet. 2011, 43, 34–41. [Google Scholar] [CrossRef] [PubMed]
  116. Petersen, B.E.; Goff, J.P.; Greenberger, J.S.; Michalopoulos, G.K. Hepatic oval cells express the hematopoietic stem cell marker Thy-1 in the rat. Hepatology 1998, 27, 433–445. [Google Scholar] [CrossRef] [PubMed]
  117. Keller, G.; Lacaud, G.; Robertson, S. Development of the hematopoietic system in the mouse. Exp. Hematol. 1999, 27, 777–787. [Google Scholar] [CrossRef]
  118. Petersen, B.E.; Bowen, W.C.; Patrene, K.D.; Mars, W.M.; Sullivan, A.K.; Murase, N.; Boggs, S.S.; Greenberger, J.S.; Goff, J.P. Bone marrow as a potential source of hepatic oval cells. Science 1999, 284, 1168–1170. [Google Scholar] [CrossRef]
  119. Menthena, A.; Deb, N.; Oertel, M.; Grozdanov, P.N.; Sandhu, J.; Shah, S.; Guha, C.; Shafritz, D.A.; Dabeva, M.D. Bone marrow progenitors are not the source of expanding oval cells in injured liver. Stem Cells 2004, 22, 1049–1061. [Google Scholar] [CrossRef]
  120. Aurich, I.; Mueller, L.P.; Aurich, H.; Luetzkendorf, J.; Tisljar, K.; Dollinger, M.M.; Schormann, W.; Walldorf, J.; Hengstler, J.G.; Fleig, W.E.; et al. Functional integration of hepatocytes derived from human mesenchymal stem cells into mouse livers. Gut 2007, 56, 405–415. [Google Scholar] [CrossRef] [Green Version]
  121. Sgodda, M.; Aurich, H.; Kleist, S.; Aurich, I.; Konig, S.; Dollinger, M.M.; Fleig, W.E.; Christ, B. Hepatocyte differentiation of mesenchymal stem cells from rat peritoneal adipose tissue in vitro and in vivo. Exp. Cell Res. 2007, 313, 2875–2886. [Google Scholar] [CrossRef] [PubMed]
  122. Aurich, H.; Sgodda, M.; Kaltwasser, P.; Vetter, M.; Weise, A.; Liehr, T.; Brulport, M.; Hengstler, J.G.; Dollinger, M.M.; Fleig, W.E.; et al. Hepatocyte differentiation of mesenchymal stem cells from human adipose tissue in vitro promotes hepatic integration in vivo. Gut 2009, 58, 570–581. [Google Scholar] [CrossRef] [PubMed]
  123. Schneller, D.; Angel, P. Cellular Origin of Hepatocellular Carcinoma. In Hepatocellular Carcinoma; Tirnitz-Parker, J.E.E., Ed.; Codon Publications: Brisbane, AU, Australia, 24 October 2019. [Google Scholar] [CrossRef] [Green Version]
  124. 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] [PubMed] [Green Version]
  125. 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]
  126. Yamashita, T.; Forgues, M.; Wang, W.; Kim, J.W.; Ye, Q.; Jia, H.; Budhu, A.; Zanetti, K.A.; Chen, Y.; Qin, L.X.; et al. EpCAM and alpha-fetoprotein expression defines novel prognostic subtypes of hepatocellular carcinoma. Cancer Res. 2008, 68, 1451–1461. [Google Scholar] [CrossRef] [Green Version]
  127. Andersen, J.B.; Loi, R.; Perra, A.; Factor, V.M.; Ledda-Columbano, G.M.; Columbano, A.; Thorgeirsson, S.S. Progenitor-derived hepatocellular carcinoma model in the rat. Hepatology 2010, 51, 1401–1409. [Google Scholar] [CrossRef] [Green Version]
  128. Tang, Y.; Kitisin, K.; Jogunoori, W.; Li, C.; Deng, C.X.; Mueller, S.C.; Ressom, H.W.; Rashid, A.; He, A.R.; Mendelson, J.S.; et al. Progenitor/stem cells give rise to liver cancer due to aberrant TGF-beta and IL-6 signaling. Proc. Natl. Acad. Sci. USA 2008, 105, 2445–2450. [Google Scholar] [CrossRef] [Green Version]
  129. 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]
  130. 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] [Green Version]
  131. Xu, M.Z.; Yao, T.J.; Lee, N.P.; Ng, I.O.; Chan, Y.T.; Zender, L.; Lowe, S.W.; Poon, R.T.; Luk, J.M. Yes-associated protein is an independent prognostic marker in hepatocellular carcinoma. Cancer 2009, 115, 4576–4585. [Google Scholar] [CrossRef] [Green Version]
  132. Benhamouche, S.; Curto, M.; Saotome, I.; Gladden, A.B.; Liu, C.H.; Giovannini, M.; McClatchey, A.I. Nf2/Merlin controls progenitor homeostasis and tumorigenesis in the liver. Genes Dev. 2010, 24, 1718–1730. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Harada, N.; Oshima, H.; Katoh, M.; Tamai, Y.; Oshima, M.; Taketo, M.M. Hepatocarcinogenesis in mice with beta-catenin and Ha-ras gene mutations. Cancer Res. 2004, 64, 48–54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Mokkapati, S.; Niopek, K.; Huang, L.; Cunniff, K.J.; Ruteshouser, E.C.; deCaestecker, M.; Finegold, M.J.; Huff, V. beta-catenin activation in a novel liver progenitor cell type is sufficient to cause hepatocellular carcinoma and hepatoblastoma. Cancer Res. 2014, 74, 4515–4525. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Nikolaou, K.C.; Moulos, P.; Chalepakis, G.; Hatzis, P.; Oda, H.; Reinberg, D.; Talianidis, I. Spontaneous development of hepatocellular carcinoma with cancer stem cell properties in PR-SET7-deficient livers. EMBO J. 2015, 34, 430–447. [Google Scholar] [CrossRef]
  136. Mu, X.; Espanol-Suner, R.; Mederacke, I.; Affo, S.; Manco, R.; Sempoux, C.; Lemaigre, F.P.; Adili, A.; Yuan, D.; Weber, A.; et al. Hepatocellular carcinoma originates from hepatocytes and not from the progenitor/biliary compartment. J. Clin. Investig. 2015, 125, 3891–3903. [Google Scholar] [CrossRef] [Green Version]
  137. 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] [Green Version]
  138. Tummala, K.S.; Brandt, M.; Teijeiro, A.; Grana, 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] [Green Version]
  139. Ma, Z.; Guo, D.; Wang, Q.; Liu, P.; Xiao, Y.; Wu, P.; Wang, Y.; Chen, B.; Liu, Z.; Liu, Q. Lgr5-mediated p53 Repression through PDCD5 leads to doxorubicin resistance in Hepatocellular Carcinoma. Theranostics 2019, 9, 2967–2983. [Google Scholar] [CrossRef]
  140. Liu, J.; Yu, G.Z.; Cheng, X.K.; Li, X.D.; Zeng, X.T.; Ren, X.Q. LGR5 promotes hepatocellular carcinoma metastasis through inducting epithelial-mesenchymal transition. Oncotarget 2017, 8, 50896–50903. [Google Scholar] [CrossRef] [Green Version]
  141. Cao, H.Z.; Liu, X.F.; Yang, W.T.; Chen, Q.; Zheng, P.S. LGR5 promotes cancer stem cell traits and chemoresistance in cervical cancer. Cell Death Dis. 2017, 8, e3039. [Google Scholar] [CrossRef] [Green Version]
  142. Tschaharganeh, D.F.; Xue, W.; Calvisi, D.F.; Evert, M.; Michurina, T.V.; Dow, L.E.; Banito, A.; Katz, S.F.; Kastenhuber, E.R.; Weissmueller, S.; et al. p53-dependent Nestin regulation links tumor suppression to cellular plasticity in liver cancer. Cell 2014, 158, 579–592. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Jeter, C.R.; Badeaux, M.; Choy, G.; Chandra, D.; Patrawala, L.; Liu, C.; Calhoun-Davis, T.; Zaehres, H.; Daley, G.Q.; Tang, D.G. Functional evidence that the self-renewal gene NANOG regulates human tumor development. Stem Cells 2009, 27, 993–1005. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Silva, J.; Nichols, J.; Theunissen, T.W.; Guo, G.; van Oosten, A.L.; Barrandon, O.; Wray, J.; Yamanaka, S.; Chambers, I.; Smith, A. Nanog is the gateway to the pluripotent ground state. Cell 2009, 138, 722–737. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Lee, T.K.; Castilho, A.; Cheung, V.C.; Tang, K.H.; Ma, S.; Ng, I.O. CD24(+) liver tumor-initiating cells drive self-renewal and tumor initiation through STAT3-mediated NANOG regulation. Cell Stem Cell 2011, 9, 50–63. [Google Scholar] [CrossRef] [Green Version]
  146. Ma, S.; Tang, K.H.; Chan, Y.P.; Lee, T.K.; Kwan, P.S.; Castilho, A.; Ng, I.; Man, K.; Wong, N.; To, K.F.; et al. miR-130b Promotes CD133(+) liver tumor-initiating cell growth and self-renewal via tumor protein 53-induced nuclear protein 1. Cell Stem Cell 2010, 7, 694–707. [Google Scholar] [CrossRef] [Green Version]
  147. Chen, W.C.; Chang, Y.S.; Hsu, H.P.; Yen, M.C.; Huang, H.L.; Cho, C.Y.; Wang, C.Y.; Weng, T.Y.; Lai, P.T.; Chen, C.S.; et al. Therapeutics targeting CD90-integrin-AMPK-CD133 signal axis in liver cancer. Oncotarget 2015, 6, 42923–42937. [Google Scholar] [CrossRef]
  148. Dhar, D.; Antonucci, L.; Nakagawa, H.; Kim, J.Y.; Glitzner, E.; Caruso, S.; Shalapour, S.; Yang, L.; Valasek, M.A.; Lee, S.; et al. Liver Cancer Initiation Requires p53 Inhibition by CD44-Enhanced Growth Factor Signaling. Cancer Cell 2018, 33, 1061–1077.e6. [Google Scholar] [CrossRef] [Green Version]
  149. Zhao, W.; Wang, L.; Han, H.; Jin, K.; Lin, N.; Guo, T.; Chen, Y.; Cheng, H.; Lu, F.; Fang, W.; et al. 1B50-1, a mAb Raised against Recurrent Tumor Cells, Targets Liver Tumor-Initiating Cells by Binding to the Calcium Channel α2δ1 Subunit. Cancer Cell 2013, 23, 541–556. [Google Scholar] [CrossRef] [Green Version]
  150. Chen, W.; Zhang, Y.W.; Li, Y.; Zhang, J.W.; Zhang, T.; Fu, B.S.; Zhang, Q.; Jiang, N. Constitutive expression of Wnt/betacatenin target genes promotes proliferation and invasion of liver cancer stem cells. Mol. Med. Rep. 2016, 13, 3466–3474. [Google Scholar] [CrossRef] [Green Version]
  151. Chiba, T.; Kita, K.; Zheng, Y.W.; Yokosuka, O.; Saisho, H.; Iwama, A.; Nakauchi, H.; Taniguchi, H. Side population purified from hepatocellular carcinoma cells harbors cancer stem cell-like properties. Hepatology 2006, 44, 240–251. [Google Scholar] [CrossRef]
  152. Xia, H.; Cao, J.; Li, Q.; Lv, Y.; Jia, W.; Ren, W.; Cheng, Q.; Song, X.; Xu, G. Hepatocellular Carcinoma-propagating Cells are Detectable by Side Population Analysis and Possess an Expression Profile Reflective of a Primitive Origin. Sci. Rep. 2016, 6, 34856. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Talukdar, S.; Bhoopathi, P.; Emdad, L.; Das, S.; Sarkar, D.; Fisher, P.B. Dormancy and cancer stem cells: An enigma for cancer therapeutic targeting. Adv. Cancer Res. 2019, 141, 43–84. [Google Scholar] [CrossRef] [PubMed]
  154. Haraguchi, N.; Ishii, H.; Mimori, K.; Tanaka, F.; Ohkuma, M.; Kim, H.M.; Akita, H.; Takiuchi, D.; Hatano, H.; Nagano, H.; et al. CD13 is a therapeutic target in human liver cancer stem cells. J. Clin. Investig. 2010, 120, 3326–3339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Jiao, X.; Velasco-Velazquez, M.A.; Wang, M.; Li, Z.; Rui, H.; Peck, A.R.; Korkola, J.E.; Chen, X.; Xu, S.; DuHadaway, J.B.; et al. CCR5 Governs DNA Damage Repair and Breast Cancer Stem Cell Expansion. Cancer Res. 2018, 78, 1657–1671. [Google Scholar] [CrossRef] [Green Version]
  156. Gallmeier, E.; Hermann, P.C.; Mueller, M.T.; Machado, J.G.; Ziesch, A.; De Toni, E.N.; Palagyi, A.; Eisen, C.; Ellwart, J.W.; Rivera, J.; et al. Inhibition of ataxia telangiectasia- and Rad3-related function abrogates the in vitro and in vivo tumorigenicity of human colon cancer cells through depletion of the CD133(+) tumor-initiating cell fraction. Stem Cells 2011, 29, 418–429. [Google Scholar] [CrossRef]
  157. Zhou, J.J.; Deng, X.G.; He, X.Y.; Zhou, Y.; Yu, M.; Gao, W.C.; Zeng, B.; Zhou, Q.B.; Li, Z.H.; Chen, R.F. Knockdown of NANOG enhances chemosensitivity of liver cancer cells to doxorubicin by reducing MDR1 expression. Int. J. Oncol. 2014, 44, 2034–2040. [Google Scholar] [CrossRef] [Green Version]
  158. Prieto-Vila, M.; Takahashi, R.U.; Usuba, W.; Kohama, I.; Ochiya, T. Drug Resistance Driven by Cancer Stem Cells and Their Niche. Int. J. Mol. Sci. 2017, 18, 2574. [Google Scholar] [CrossRef] [Green Version]
  159. Lu, S.; Yao, Y.; Xu, G.; Zhou, C.; Zhang, Y.; Sun, J.; Jiang, R.; Shao, Q.; Chen, Y. CD24 regulates sorafenib resistance via activating autophagy in hepatocellular carcinoma. Cell Death Dis. 2018, 9, 646. [Google Scholar] [CrossRef]
  160. Wang, R.; Li, Y.; Tsung, A.; Huang, H.; Du, Q.; Yang, M.; Deng, M.; Xiong, S.; Wang, X.; Zhang, L.; et al. iNOS promotes CD24(+)CD133(+) liver cancer stem cell phenotype through a TACE/ADAM17-dependent Notch signaling pathway. Proc. Natl. Acad. Sci. USA 2018, 115, E10127–E10136. [Google Scholar] [CrossRef] [Green Version]
  161. Yin, S.; Li, J.; Hu, C.; Chen, X.; Yao, M.; Yan, M.; Jiang, G.; Ge, C.; Xie, H.; Wan, D.; et al. CD133 positive hepatocellular carcinoma cells possess high capacity for tumorigenicity. Int. J. Cancer 2007, 120, 1444–1450. [Google Scholar] [CrossRef]
  162. Suetsugu, A.; Nagaki, M.; Aoki, H.; Motohashi, T.; Kunisada, T.; Moriwaki, H. Characterization of CD133+ hepatocellular carcinoma cells as cancer stem/progenitor cells. Biochem. Biophys. Res. Commun. 2006, 351, 820–824. [Google Scholar] [CrossRef] [PubMed]
  163. Ma, Y.C.; Yang, J.Y.; Yan, L.N. Relevant markers of cancer stem cells indicate a poor prognosis in hepatocellular carcinoma patients: A meta-analysis. Eur. J. Gastroenterol. Hepatol. 2013, 25, 1007–1016. [Google Scholar] [CrossRef] [PubMed]
  164. Smith, L.M.; Nesterova, A.; Ryan, M.C.; Duniho, S.; Jonas, M.; Anderson, M.; Zabinski, R.F.; Sutherland, M.K.; Gerber, H.P.; Van Orden, K.L.; et al. CD133/prominin-1 is a potential therapeutic target for antibody-drug conjugates in hepatocellular and gastric cancers. Br. J. Cancer 2008, 99, 100–109. [Google Scholar] [CrossRef] [PubMed]
  165. Ponta, H.; Sherman, L.; Herrlich, P.A. CD44: From adhesion molecules to signalling regulators. Nat. Rev. Mol. Cell. Biol. 2003, 4, 33–45. [Google Scholar] [CrossRef] [PubMed]
  166. Zoller, M. CD44: Can a cancer-initiating cell profit from an abundantly expressed molecule? Nat. Rev. Cancer 2011, 11, 254–267. [Google Scholar] [CrossRef] [PubMed]
  167. Brown, R.L.; Reinke, L.M.; Damerow, M.S.; Perez, D.; Chodosh, L.A.; Yang, J.; Cheng, C. CD44 splice isoform switching in human and mouse epithelium is essential for epithelial-mesenchymal transition and breast cancer progression. J. Clin. Investig. 2011, 121, 1064–1074. [Google Scholar] [CrossRef] [Green Version]
  168. Zhou, M.; Yang, H.; Learned, R.M.; Tian, H.; Ling, L. Non-cell-autonomous activation of IL-6/STAT3 signaling mediates FGF19-driven hepatocarcinogenesis. Nat. Commun. 2017, 8, 15433. [Google Scholar] [CrossRef]
  169. Wan, S.; Zhao, E.; Kryczek, I.; Vatan, L.; Sadovskaya, A.; Ludema, G.; Simeone, D.M.; Zou, W.; Welling, T.H. Tumor-associated macrophages produce interleukin 6 and signal via STAT3 to promote expansion of human hepatocellular carcinoma stem cells. Gastroenterology 2014, 147, 1393–1404. [Google Scholar] [CrossRef] [Green Version]
  170. Mima, K.; Okabe, H.; Ishimoto, T.; Hayashi, H.; Nakagawa, S.; Kuroki, H.; Watanabe, M.; Beppu, T.; Tamada, M.; Nagano, O.; et al. CD44s regulates the TGF-beta-mediated mesenchymal phenotype and is associated with poor prognosis in patients with hepatocellular carcinoma. Cancer Res. 2012, 72, 3414–3423. [Google Scholar] [CrossRef] [Green Version]
  171. Jordan, A.R.; Racine, R.R.; Hennig, M.J.; Lokeshwar, V.B. The Role of CD44 in Disease Pathophysiology and Targeted Treatment. Front. Immunol. 2015, 6, 182. [Google Scholar] [CrossRef]
  172. Ishimoto, T.; Nagano, O.; Yae, T.; Tamada, M.; Motohara, T.; Oshima, H.; Oshima, M.; Ikeda, T.; Asaba, R.; Yagi, H.; et al. CD44 variant regulates redox status in cancer cells by stabilizing the xCT subunit of system xc(-) and thereby promotes tumor growth. Cancer Cell 2011, 19, 387–400. [Google Scholar] [CrossRef] [Green Version]
  173. Yamashita, T.; Budhu, A.; Forgues, M.; Wang, X.W. Activation of hepatic stem cell marker EpCAM by Wnt-beta-catenin signaling in hepatocellular carcinoma. Cancer Res. 2007, 67, 10831–10839. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Wang, C.; Fu, S.Y.; Wang, M.D.; Yu, W.B.; Cui, Q.S.; Wang, H.R.; Huang, H.; Dong, W.; Zhang, W.W.; Li, P.P.; et al. Zinc finger protein X-linked promotes expansion of EpCAM(+) cancer stem-like cells in hepatocellular carcinoma. Mol. Oncol. 2017, 11, 455–469. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Eyvazi, S.; Farajnia, S.; Dastmalchi, S.; Kanipour, F.; Zarredar, H.; Bandehpour, M. Antibody Based EpCAM Targeted Therapy of Cancer, Review and Update. Curr. Cancer Drug Targets 2018, 18, 857–868. [Google Scholar] [CrossRef] [PubMed]
  176. Yang, Z.F.; Ngai, P.; Ho, D.W.; Yu, W.C.; Ng, M.N.; Lau, C.K.; Li, M.L.; Tam, K.H.; Lam, C.T.; Poon, R.T.; et al. Identification of local and circulating cancer stem cells in human liver cancer. Hepatology 2008, 47, 919–928. [Google Scholar] [CrossRef] [PubMed]
  177. Yamashita, T.; Honda, M.; Nakamoto, Y.; Baba, M.; Nio, K.; Hara, Y.; Zeng, S.S.; Hayashi, T.; Kondo, M.; Takatori, H.; et al. Discrete nature of EpCAM+ and CD90+ cancer stem cells in human hepatocellular carcinoma. Hepatology 2013, 57, 1484–1497. [Google Scholar] [CrossRef]
  178. Xia, W.; Lo, C.M.; Poon, R.Y.C.; Cheung, T.T.; Chan, A.C.Y.; Chen, L.; Yang, S.; Tsao, G.S.W.; Wang, X.Q. Smad inhibitor induces CSC differentiation for effective chemosensitization in cyclin D1- and TGF-beta/Smad-regulated liver cancer stem cell-like cells. Oncotarget 2017, 8, 38811–38824. [Google Scholar] [CrossRef]
  179. Zhu, L.; Zhang, W.; Wang, J.; Liu, R. Evidence of CD90+CXCR4+ cells as circulating tumor stem cells in hepatocellular carcinoma. Tumour Biol. J. Int. Soc. Oncodevelopmental Biol. Med. 2015, 36, 5353–5360. [Google Scholar] [CrossRef]
  180. Kim, B.H.; Park, J.W.; Kim, J.S.; Lee, S.K.; Hong, E.K. Stem Cell Markers Predict the Response to Sorafenib in Patients with Hepatocellular Carcinoma. Gut Liver 2019, 13, 342–348. [Google Scholar] [CrossRef]
  181. Yang, R.; An, L.Y.; Miao, Q.F.; Li, F.M.; Han, Y.; Wang, H.X.; Liu, D.P.; Chen, R.; Tang, S.Q. Effective elimination of liver cancer stem-like cells by CD90 antibody targeted thermosensitive magnetoliposomes. Oncotarget 2016, 7, 35894–35916. [Google Scholar] [CrossRef] [Green Version]
  182. Yamanaka, C.; Wada, H.; Eguchi, H.; Hatano, H.; Gotoh, K.; Noda, T.; Yamada, D.; Asaoka, T.; Kawamoto, K.; Nagano, H.; et al. Clinical significance of CD13 and epithelial mesenchymal transition (EMT) markers in hepatocellular carcinoma. Jpn. J. Clin. Oncol. 2018, 48, 52–60. [Google Scholar] [CrossRef] [PubMed]
  183. Kim, H.M.; Haraguchi, N.; Ishii, H.; Ohkuma, M.; Okano, M.; Mimori, K.; Eguchi, H.; Yamamoto, H.; Nagano, H.; Sekimoto, M.; et al. Increased CD13 expression reduces reactive oxygen species, promoting survival of liver cancer stem cells via an epithelial-mesenchymal transition-like phenomenon. Ann. Surg. Oncol. 2012, 19 (Suppl. 3), S539–S548. [Google Scholar] [CrossRef]
  184. Toshiyama, R.; Konno, M.; Eguchi, H.; Takemoto, H.; Noda, T.; Asai, A.; Koseki, J.; Haraguchi, N.; Ueda, Y.; Matsushita, K.; et al. Poly(ethylene glycol)-poly(lysine) block copolymer-ubenimex conjugate targets aminopeptidase N and exerts an antitumor effect in hepatocellular carcinoma stem cells. Oncogene 2019, 38, 244–260. [Google Scholar] [CrossRef] [PubMed]
  185. Guo, Q.; Sui, Z.G.; Xu, W.; Quan, X.H.; Sun, J.L.; Li, X.; Ji, H.Y.; Jing, F.B. Ubenimex suppresses Pim-3 kinase expression by targeting CD13 to reverse MDR in HCC cells. Oncotarget 2017, 8, 72652–72665. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Yamashita, M.; Wada, H.; Eguchi, H.; Ogawa, H.; Yamada, D.; Noda, T.; Asaoka, T.; Kawamoto, K.; Gotoh, K.; Umeshita, K.; et al. A CD13 inhibitor, ubenimex, synergistically enhances the effects of anticancer drugs in hepatocellular carcinoma. Int. J. Oncol. 2016, 49, 89–98. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Sun, Z.P.; Zhang, J.; Shi, L.H.; Zhang, X.R.; Duan, Y.; Xu, W.F.; Dai, G.; Wang, X.J. Aminopeptidase N inhibitor 4cc synergizes antitumor effects of 5-fluorouracil on human liver cancer cells through ROS-dependent CD13 inhibition. Biomed. Pharmacother. Biomed. Pharmacother. 2015, 76, 65–72. [Google Scholar] [CrossRef] [PubMed]
  188. Dou, C.; Fang, C.; Zhao, Y.; Fu, X.; Zhang, Y.; Zhu, D.; Wu, H.; Liu, H.; Zhang, J.; Xu, W.; et al. BC-02 eradicates liver cancer stem cells by upregulating the ROS-dependent DNA damage. Int. J. Oncol. 2017, 51, 1775–1784. [Google Scholar] [CrossRef] [Green Version]
  189. Inagaki, Y.; Tang, W.; Zhang, L.; Du, G.; Xu, W.; Kokudo, N. Novel aminopeptidase N (APN/CD13) inhibitor 24F can suppress invasion of hepatocellular carcinoma cells as well as angiogenesis. Biosci. Trends 2010, 4, 56–60. [Google Scholar]
  190. Dunsford, H.A.; Sell, S. Production of monoclonal antibodies to preneoplastic liver cell populations induced by chemical carcinogens in rats and to transplantable Morris hepatomas. Cancer Res. 1989, 49, 4887–4893. [Google Scholar]
  191. Roskams, T.; De Vos, R.; Van Eyken, P.; Myazaki, H.; Van Damme, B.; Desmet, V. Hepatic OV-6 expression in human liver disease and rat experiments: Evidence for hepatic progenitor cells in man. J. Hepatol. 1998, 29, 455–463. [Google Scholar] [CrossRef]
  192. Parent, R.; Marion, M.J.; Furio, L.; Trepo, C.; Petit, M.A. Origin and characterization of a human bipotent liver progenitor cell line. Gastroenterology 2004, 126, 1147–1156. [Google Scholar] [CrossRef] [PubMed]
  193. Jia, S.Q.; Ren, J.J.; Dong, P.D.; Meng, X.K. Probing the hepatic progenitor cell in human hepatocellular carcinoma. Gastroenterol. Res. Pract. 2013, 2013, 145253. [Google Scholar] [CrossRef] [PubMed]
  194. Yang, W.; Wang, C.; Lin, Y.; Liu, Q.; Yu, L.X.; Tang, L.; Yan, H.X.; Fu, J.; Chen, Y.; Zhang, H.L.; et al. OV6(+) tumor-initiating cells contribute to tumor progression and invasion in human hepatocellular carcinoma. J. Hepatol. 2012, 57, 613–620. [Google Scholar] [CrossRef] [PubMed]
  195. Zhu, J.; Yu, H.; Chen, S.; Yang, P.; Dong, Z.; Ling, Y.; Tang, H.; Bai, S.; Yang, W.; Tang, L.; et al. Prognostic significance of combining high mobility group Box-1 and OV-6 expression in hepatocellular carcinoma. Sci. China Life Sci. 2018, 61, 912–923. [Google Scholar] [CrossRef]
  196. Yang, W.; Yan, H.X.; Chen, L.; Liu, Q.; He, Y.Q.; Yu, L.X.; Zhang, S.H.; Huang, D.D.; Tang, L.; Kong, X.N.; et al. Wnt/beta-catenin signaling contributes to activation of normal and tumorigenic liver progenitor cells. Cancer Res. 2008, 68, 4287–4295. [Google Scholar] [CrossRef] [Green Version]
  197. Zhang, Z.; Zhao, W.; Lin, X.; Gao, J.; Zhang, Z.; Shen, L. Voltage-dependent calcium channel α2δ1 subunit is a specific candidate marker for identifying gastric cancer stem cells. CMAR 2019, 11, 4707–4718. [Google Scholar] [CrossRef] [Green Version]
  198. Huang, C.; Li, Y.; Zhao, W.; Zhang, A.; Lu, C.; Wang, Z.; Liu, L. α2δ1 may be a potential marker for cancer stem cell in laryngeal squamous cell carcinoma. CBM 2019, 24, 97–107. [Google Scholar] [CrossRef] [Green Version]
  199. Yu, J.; Wang, S.; Zhao, W.; Duan, J.; Wang, Z.; Chen, H.; Tian, Y.; Wang, D.; Zhao, J.; An, T.; et al. Mechanistic Exploration of Cancer Stem Cell Marker Voltage-Dependent Calcium Channel α2δ1 Subunit-mediated Chemotherapy Resistance in Small-Cell Lung Cancer. Clin. Cancer Res. 2018, 24, 2148–2158. [Google Scholar] [CrossRef] [Green Version]
  200. Sainz, B., Jr.; Heeschen, C. Standing out from the crowd: Cancer stem cells in hepatocellular carcinoma. Cancer Cell 2013, 23, 431–433. [Google Scholar] [CrossRef] [Green Version]
  201. El-Serag, H.B.; Rudolph, K.L. Hepatocellular carcinoma: Epidemiology and molecular carcinogenesis. Gastroenterology 2007, 132, 2557–2576. [Google Scholar] [CrossRef]
  202. Sukowati, C.H.C. Significance of hepatitis virus infection in the oncogenic initiation of hepatocellular carcinoma. WJG 2016, 22, 1497. [Google Scholar] [CrossRef] [PubMed]
  203. Ng, K.-Y.; Chai, S.; Tong, M.; Guan, X.-Y.; Lin, C.-H.; Ching, Y.-P.; Xie, D.; Cheng, A.S.-L.; Ma, S. C-terminal truncated hepatitis B virus X protein promotes hepatocellular carcinogenesis through induction of cancer and stem cell-like properties. Oncotarget 2016, 7. [Google Scholar] [CrossRef] [PubMed]
  204. Wang, C.; Yang, W.; Yan, H.-X.; Luo, T.; Zhang, J.; Tang, L.; Wu, F.-Q.; Zhang, H.-L.; Yu, L.-X.; Zheng, L.-Y.; et al. Hepatitis B virus X (HBx) induces tumorigenicity of hepatic progenitor cells in 3,5-diethoxycarbonyl-1,4-dihydrocollidine-treated HBx transgenic mice. Hepatology 2012, 55, 108–120. [Google Scholar] [CrossRef] [PubMed]
  205. Mani, S.K.K.; Andrisani, O. Hepatitis B Virus-Associated Hepatocellular Carcinoma and Hepatic Cancer Stem Cells. Genes 2018, 9, 137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Liu, Z.; Dai, X.; Wang, T.; Zhang, C.; Zhang, W.; Zhang, W.; Zhang, Q.; Wu, K.; Liu, F.; Liu, Y.; et al. Hepatitis B virus PreS1 facilitates hepatocellular carcinoma development by promoting appearance and self-renewal of liver cancer stem cells. Cancer Lett. 2017, 400, 149–160. [Google Scholar] [CrossRef] [PubMed]
  207. Mani, S.K.K.; Zhang, H.; Diab, A.; Pascuzzi, P.E.; Lefrançois, L.; Fares, N.; Bancel, B.; Merle, P.; Andrisani, O. EpCAM-regulated intramembrane proteolysis induces a cancer stem cell-like gene signature in hepatitis B virus-infected hepatocytes. J. Hepatol. 2016, 65, 888–898. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Chang, T.-S.; Chen, C.-L.; Wu, Y.-C.; Liu, J.-J.; Kuo, Y.C.; Lee, K.-F.; Lin, S.-Y.; Lin, S.-E.; Tung, S.-Y.; Kuo, L.-M.; et al. Inflammation Promotes Expression of Stemness-Related Properties in HBV-Related Hepatocellular Carcinoma. PLoS ONE 2016, 11, e0149897. [Google Scholar] [CrossRef] [Green Version]
  209. Ali, N.; Allam, H.; May, R.; Sureban, S.M.; Bronze, M.S.; Bader, T.; Umar, S.; Anant, S.; Houchen, C.W. Hepatitis C Virus-Induced Cancer Stem Cell-Like Signatures in Cell Culture and Murine Tumor Xenografts. J. Virol. 2011, 85, 12292–12303. [Google Scholar] [CrossRef] [Green Version]
  210. Machida, K.; Chen, C.-L.; Liu, J.-C.; Kashiwabara, C.; Feldman, D.; French, S.W.; Sher, L.; Hyeongnam, J.J.; Tsukamoto, H. Cancer stem cells generated by alcohol, diabetes, and hepatitis C virus: Toll-like receptor signaling in liver diseases. J. Gastroenterol. Hepatol. 2012, 27, 19–22. [Google Scholar] [CrossRef] [Green Version]
  211. Kwon, Y.-C.; Bose, S.K.; Steele, R.; Meyer, K.; Di Bisceglie, A.M.; Ray, R.B.; Ray, R. Promotion of Cancer Stem-Like Cell Properties in Hepatitis C Virus-Infected Hepatocytes. J. Virol. 2015, 89, 11549–11556. [Google Scholar] [CrossRef] [Green Version]
  212. Shirasaki, T.; Honda, M.; Yamashita, T.; Nio, K.; Shimakami, T.; Shimizu, R.; Nakasyo, S.; Murai, K.; Shirasaki, N.; Okada, H.; et al. The osteopontin-CD44 axis in hepatic cancer stem cells regulates IFN signaling and HCV replication. Sci. Rep. 2018, 8, 13143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Finkin, S.; Yuan, D.; Stein, I.; Taniguchi, K.; Weber, A.; Unger, K.; Browning, J.L.; Goossens, N.; Nakagawa, S.; Gunasekaran, G.; et al. Ectopic lymphoid structures function as microniches for tumor progenitor cells in hepatocellular carcinoma. Nat. Immunol. 2015, 16, 1235–1244. [Google Scholar] [CrossRef] [PubMed]
  214. Di Caro, G.; Bergomas, F.; Grizzi, F.; Doni, A.; Bianchi, P.; Malesci, A.; Laghi, L.; Allavena, P.; Mantovani, A.; Marchesi, F. Occurrence of tertiary lymphoid tissue is associated with T-cell infiltration and predicts better prognosis in early-stage colorectal cancers. Clin. Cancer Res. 2014, 20, 2147–2158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Dieu-Nosjean, M.C.; Antoine, M.; Danel, C.; Heudes, D.; Wislez, M.; Poulot, V.; Rabbe, N.; Laurans, L.; Tartour, E.; de Chaisemartin, L.; et al. Long-term survival for patients with non-small-cell lung cancer with intratumoral lymphoid structures. J. Clin. Oncol. 2008, 26, 4410–4417. [Google Scholar] [CrossRef] [Green Version]
  216. Gu-Trantien, C.; Loi, S.; Garaud, S.; Equeter, C.; Libin, M.; de Wind, A.; Ravoet, M.; Le Buanec, H.; Sibille, C.; Manfouo-Foutsop, G.; et al. CD4(+) follicular helper T cell infiltration predicts breast cancer survival. J. Clin. Investig. 2013, 123, 2873–2892. [Google Scholar] [CrossRef]
  217. Svinarenko, M.; Katz, S.F.; Tharehalli, U.; Mulaw, M.A.; Maier, H.J.; Sunami, Y.; Fischer, S.K.; Chen, Y.; Heurich, S.; Erkert, L.; et al. An IKK/NF-kappaB Activation/p53 Deletion Sequence Drives Liver Carcinogenesis and Tumor Differentiation. Cancers 2019, 11, 1410. [Google Scholar] [CrossRef] [Green Version]
  218. Cui, Y.; Sun, S.; Ren, K.; Quan, M.; Song, Z.; Zou, H.; Li, D.; Cao, J. Reversal of liver cancer-associated stellate cell-induced stem-like characteristics in SMMC-7721 cells by 8-bromo-7-methoxychrysin via inhibiting STAT3 activation. Oncol. Rep. 2016, 35, 2952–2962. [Google Scholar] [CrossRef] [Green Version]
  219. Yu, G.; Jing, Y.; Kou, X.; Ye, F.; Gao, L.; Fan, Q.; Yang, Y.; Zhao, Q.; Li, R.; Wu, M.; et al. Hepatic Stellate Cells Secreted Hepatocyte Growth Factor Contributes to the Chemoresistance of Hepatocellular Carcinoma. PLoS ONE 2013, 8, e73312. [Google Scholar] [CrossRef]
  220. Chen, A.; Xu, C.; Luo, Y.; Liu, L.; Song, K.; Deng, G.; Yang, M.; Cao, J.; Yuan, L.; Li, X. Disruption of crosstalk between LX-2 and liver cancer stem-like cells from MHCC97H cells by DFOG via inhibiting FOXM1. Acta Biochim. Et Biophys. Sin. 2019, 51, 1267–1275. [Google Scholar] [CrossRef]
  221. Zhao, L.; Zhao, Y.; Schwarz, B.; Mysliwietz, J.; Hartig, R.; Camaj, P.; Bao, Q.; Jauch, K.W.; Guba, M.; Ellwart, J.W.; et al. Verapamil inhibits tumor progression of chemotherapy-resistant pancreatic cancer side population cells. Int. J. Oncol. 2016, 49, 99–110. [Google Scholar] [CrossRef] [Green Version]
  222. Zhang, T.; Ma, K.; Huang, J.; Wang, S.; Liu, Y.; Fan, G.; Liu, M.; Yang, G.; Wang, C.; Fan, P. CDKN2B is critical for verapamil-mediated reversal of doxorubicin resistance in hepatocellular carcinoma. Oncotarget 2017, 8, 110052–110063. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Jalali, A.; Ghasemian, S.; Najafzadeh, H.; Galehdari, H.; Seifi, M.R.; Zangene, F.; Dehdardargahi, S. Verapamil and rifampin effect on p-glycoprotein expression in hepatocellular carcinoma. Jundishapur. J. Nat. Pharm. Prod. 2014, 9, e17741. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Di Giacomo, S.; Briz, O.; Monte, M.J.; Sanchez-Vicente, L.; Abete, L.; Lozano, E.; Mazzanti, G.; Di Sotto, A.; Marin, J.J.G. Chemosensitization of hepatocellular carcinoma cells to sorafenib by beta-caryophyllene oxide-induced inhibition of ABC export pumps. Arch. Toxicol. 2019, 93, 623–634. [Google Scholar] [CrossRef] [PubMed]
  225. Chung, F.S.; Santiago, J.S.; Jesus, M.F.; Trinidad, C.V.; See, M.F. Disrupting P-glycoprotein function in clinical settings: What can we learn from the fundamental aspects of this transporter? Am. J. Cancer Res. 2016, 6, 1583–1598. [Google Scholar]
  226. Di Tomaso, T.; Mazzoleni, S.; Wang, E.; Sovena, G.; Clavenna, D.; Franzin, A.; Mortini, P.; Ferrone, S.; Doglioni, C.; Marincola, F.M.; et al. Immunobiological characterization of cancer stem cells isolated from glioblastoma patients. Clin. Cancer Res. 2010, 16, 800–813. [Google Scholar] [CrossRef] [Green Version]
  227. Wei, J.; Barr, J.; Kong, L.Y.; Wang, Y.; Wu, A.; Sharma, A.K.; Gumin, J.; Henry, V.; Colman, H.; Priebe, W.; et al. Glioblastoma cancer-initiating cells inhibit T-cell proliferation and effector responses by the signal transducers and activators of transcription 3 pathway. Mol. Cancer Ther. 2010, 9, 67–78. [Google Scholar] [CrossRef] [Green Version]
  228. Lee, Y.; Shin, J.H.; Longmire, M.; Wang, H.; Kohrt, H.E.; Chang, H.Y.; Sunwoo, J.B. CD44+ Cells in Head and Neck Squamous Cell Carcinoma Suppress T-Cell-Mediated Immunity by Selective Constitutive and Inducible Expression of PD-L1. Clin. Cancer Res. 2016, 22, 3571–3581. [Google Scholar] [CrossRef] [Green Version]
  229. Wu, Y.; Chen, M.; Wu, P.; Chen, C.; Xu, Z.P.; Gu, W. Increased PD-L1 expression in breast and colon cancer stem cells. Clin. Exp. Pharmacol. Physiol. 2017, 44, 602–604. [Google Scholar] [CrossRef]
  230. Dianzani, C.; Minelli, R.; Gigliotti, C.L.; Occhipinti, S.; Giovarelli, M.; Conti, L.; Boggio, E.; Shivakumar, Y.; Baldanzi, G.; Malacarne, V.; et al. B7h triggering inhibits the migration of tumor cell lines. J. Immunol. 2014, 192, 4921–4931. [Google Scholar] [CrossRef] [Green Version]
  231. Chen, L.; Gibbons, D.L.; Goswami, S.; Cortez, M.A.; Ahn, Y.H.; Byers, L.A.; Zhang, X.; Yi, X.; Dwyer, D.; Lin, W.; et al. Metastasis is regulated via microRNA-200/ZEB1 axis control of tumour cell PD-L1 expression and intratumoral immunosuppression. Nat. Commun. 2014, 5, 5241. [Google Scholar] [CrossRef]
  232. Qin, Y.; Yu, J.; Zhang, M.; Qin, F.; Lan, X. ZEB1 promotes tumorigenesis and metastasis in hepatocellular carcinoma by regulating the expression of vimentin. Mol. Med. Rep. 2019, 19, 2297–2306. [Google Scholar] [CrossRef] [Green Version]
  233. Ruiu, R.; Tarone, L.; Rolih, V.; Barutello, G.; Bolli, E.; Riccardo, F.; Cavallo, F.; Conti, L. Cancer stem cell immunology and immunotherapy: Harnessing the immune system against cancer’s source. Prog. Mol. Biol. Transl. Sci. 2019, 164, 119–188. [Google Scholar] [CrossRef] [PubMed]
  234. Tallerico, R.; Todaro, M.; Di Franco, S.; Maccalli, C.; Garofalo, C.; Sottile, R.; Palmieri, C.; Tirinato, L.; Pangigadde, P.N.; La Rocca, R.; et al. Human NK cells selective targeting of colon cancer-initiating cells: A role for natural cytotoxicity receptors and MHC class I molecules. J. Immunol. 2013, 190, 2381–2390. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Volonte, A.; Di Tomaso, T.; Spinelli, M.; Todaro, M.; Sanvito, F.; Albarello, L.; Bissolati, M.; Ghirardelli, L.; Orsenigo, E.; Ferrone, S.; et al. Cancer-initiating cells from colorectal cancer patients escape from T cell-mediated immunosurveillance in vitro through membrane-bound IL-4. J. Immunol. 2014, 192, 523–532. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Established markers for cancer stem cells in hepatocellular carcinoma (HCC) and possible functions. MDR: multidrug resistance protein, ATP-dependent substrate export; α2δ1: calcium voltage-gated channel auxiliary subunit Alpha2Delta1, calcium channel; EpCAM: epithelial cell adhesion molecule, single-trans-membrane cell surface adhesion molecule; CD133: prominin 1, pentaspan transmembrane molecule; CD24, CD90: GPI-anchored cell surface molecules; CD44: single-trans-membrane cell surface molecule with multiple functions, including cell–matrix and cell–cell interactions. mTOR: mammalian target of rapamycin. Mdm2: murine double minute 2. MAPK: mitogen activated protein kinases. ERK: extracellular signal-regulated kinases.
Figure 1. Established markers for cancer stem cells in hepatocellular carcinoma (HCC) and possible functions. MDR: multidrug resistance protein, ATP-dependent substrate export; α2δ1: calcium voltage-gated channel auxiliary subunit Alpha2Delta1, calcium channel; EpCAM: epithelial cell adhesion molecule, single-trans-membrane cell surface adhesion molecule; CD133: prominin 1, pentaspan transmembrane molecule; CD24, CD90: GPI-anchored cell surface molecules; CD44: single-trans-membrane cell surface molecule with multiple functions, including cell–matrix and cell–cell interactions. mTOR: mammalian target of rapamycin. Mdm2: murine double minute 2. MAPK: mitogen activated protein kinases. ERK: extracellular signal-regulated kinases.
Cancers 12 00684 g001
Table 1. Surface molecules linked to cancer stem cell (CSC) traits in HCC and their putative oncogenic and stemness supporting functions (Figure 1).
Table 1. Surface molecules linked to cancer stem cell (CSC) traits in HCC and their putative oncogenic and stemness supporting functions (Figure 1).
MDR ProteinsUpregulation in HCC-CSC and contribute to drug resistance by active outward transport of drugs [31]
CD24Upregulation in HCC CSC leads to Nanog-upregulation and therefore stemness-conservation [143,144,145]
CD133Activates autocrine signals ultimately leading to pro-oncogenic MAPK signaling [38,146]
CD90Activates AMPK and its downstream target mTOR [147]
CD44Mdm2 Activation [148]
EpCAMInduced by β-catenin signaling [126]
α2δ1Subunit of voltage-gated calcium channel complex, ERK1/2 activation [149]
Table 2. Major mechanisms of resistance of CSCs to therapy.
Table 2. Major mechanisms of resistance of CSCs to therapy.
Mechanisms of Resistance to THERAPY of CSCs
Dormancy: G0-cell cycle arrest with high resistance to genotoxic agents [153]
Increased DNA-Repair and reduced apoptotic response to DNA damage [154,155,156]
Active efflux-pumping of antineoplastic agents via MDR proteins [157]
Anti-apoptotic tumor microenvironment: The Cancer Stem Cell Niche [158]

Share and Cite

MDPI and ACS Style

Schulte, L.-A.; López-Gil, J.C.; Sainz, B., Jr.; Hermann, P.C. The Cancer Stem Cell in Hepatocellular Carcinoma. Cancers 2020, 12, 684. https://doi.org/10.3390/cancers12030684

AMA Style

Schulte L-A, López-Gil JC, Sainz B Jr., Hermann PC. The Cancer Stem Cell in Hepatocellular Carcinoma. Cancers. 2020; 12(3):684. https://doi.org/10.3390/cancers12030684

Chicago/Turabian Style

Schulte, Lucas-Alexander, Juan Carlos López-Gil, Bruno Sainz, Jr., and Patrick C. Hermann. 2020. "The Cancer Stem Cell in Hepatocellular Carcinoma" Cancers 12, no. 3: 684. https://doi.org/10.3390/cancers12030684

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