Download PDF
Review  |  Open Access  |  4 Nov 2018

Stemness features in liver cancer

Views: 3831 |  Downloads: 989 |  Cited:   5
Hepatoma Res 2018;4:69.
10.20517/2394-5079.2018.96 |  © The Author(s) 2018.
Author Information
Article Notes
Cite This Article

Abstract

Heterogeneity is a cardinal hallmark of cancer, including primary liver cancer (PLC), and occurs at different layers including putative cell-of-origin. Current evidence suggests that within cellular subpopulations in PLC there are stem-like cells, the cancer stem cells (CSCs). The CSC concept has been recently proposed as an explanation of such intra-tumor heterogeneity. According to this model, CSCs are responsible for tumor initiation, recurrence, metastasis as well as drug-resistance. However, although the CSC hypothesis is intriguing and supported by a large number of experimental studies, there are still open questions regarding the origin of putative CSCs. Since chemo-resistance and recurrence represent major issues in PLC treatment, the development of new therapeutic strategies is needed, for which a good understanding of tumor behavior and in particular of CSCs biology is an imperative prerequisite. In this review we summarize the regulatory pathways that support CSC features in PLC. Moreover, we highlight the key features of hepatic CSC, in terms of enhanced drug-resistance, increased metastatic potential and metabolic rearrangement. Knowledge of the molecular mechanisms underlying CSC biology may provide novel options for PLC combination therapies.

Keywords

Hepatocellular carcinoma, cholangiocarcinoma, cancer stem cells, tumor heterogeneity, drug-resistance

Multiple cells-of-origin of primary liver cancer

Primary liver cancer (PLC) is one of the most common cancers worldwide and the second leading cause of cancer-related mortality[1,2]. The major forms of PLC comprise hepatocellular carcinoma (HCC) and cholangiocarcinoma (CCA)[1,3-5]. HCC accounts for approximately 90% of all PLCs[1,3], while CCA is the second most common form and accounts for about 5% of all PLCs[3-5]. HCC causes over 600,000 deaths worldwide annually, and its incidence and mortality are increasing at a fast rate[6-10]. On the other hand, CCA is characterized by a very poor prognosis, with a 5-years survival lower than 20%, and its incidence and worldwide mortality are also increasing[5,11-13]. The high mortality rate of CCA may depend on its non-specific or silent clinical features and the lack of specific markers that make it difficult to diagnose[14-16].

Many studies carried out in these last years have attempted to define which type of epithelial cell [hepatocytes, cholangiocytes, hepatic progenitor cells (HPCs) or all three] should be considered as the PLC cell of origin[17]. For a long time, HCC and CCA have been commonly accepted to derive from hepatocytes and cholangiocytes, respectively. Since mature hepatocytes and cholangiocytes have an enormous self-renewal capacity and longevity, they meet the requirements to be targets for oncogenesis[17-23]. Detailed analyses of a wide range of PLC tumor types have reported that a rare form of combined HCC-CCA (cHCC-CCA) has intermediate characteristics between HCC and intrahepatic CCA (iCCA), suggesting that they could share the same stem/progenitor cell origin[18-24]. In this regard, since most PLCs arise on the background of chronic liver disease in the presence of an extensive activation of the HPC compartment (the so-called ductular reaction), several studies suggested that PLCs can be derived from HPCs rather than from mature cell types[25]. HPCs situated in the canal of Hering physiologically act as a reserve cell compartment activated in case of liver damage or when mature hepatocytes and/or cholangiocytes replication is compromised. These cells are bipotential, and may differentiate into either hepatocytes or cholangiocytes[26-28]. During the differentiation in malignant cells, bipotential HPCs undergo maturation arrest and give rise to a spectrum of tumor phenotypes with both admixed hepatocellular and cholangiocellular features, such as cholangiolocellular carcinoma and cHCC-CCA[29-31]. Additionally, a new subtype of CCA-like HCC (CLHCC) has been discovered and characterized as HCC expressing CCA-like traits[32]. CLHCC co-express embryonic stem cell (ESC) traits and hepatoblast-like genomic signatures, suggesting a HPC origin. These lines of evidence provided important insight into the heterogeneous progression of PLCs, which imply a common evolutionary origin from cells at different developmental stages[31-33]. The hypothesis of a progenitor cell origin has been supported by new advancement in genome wide analysis. Indeed, it has been suggested that iCCA and HCC are closely related at molecular level[19,29,34,35], since both tumor types share common copy number variations[11,36].

Such phenotypic variability and presence of progenitor cell features in PLC can be explained in two ways: either the cell of origin is a progenitor cell with acquired genetic alterations or, alternatively, mature tumor cells de-differentiate acquiring progenitor cell features during carcinogenesis (de-differentiation theory[37-40]). Interestingly, new findings provide direct evidence that any cell in the hepatic lineage can be the cell of origin of PLC[41]. In this regard, it has been recently suggested the development of iCCA by lineage conversion of malignant hepatocytes, through a co-activation of both Notch and protein kinase B (AKT) signaling, contributes to the acquisition of stem/progenitor cell features[42,43]. In spite of the marked plasticity in the underlying cells of origin, current evidence suggests that most PLCs are derived from undifferentiated cells with stem-like capabilities[40].

Understanding the concept of cancer stem cell

Extensive clinical and pathobiological heterogeneity at the level of cellular morphologies, genetic fingerprints and responses to therapies is a cardinal hallmark of cancer, including PLC. Such tumor complexity may reflect the presence of different cell subtypes with distinct self-renewal and differentiation potentials[40,44-46]. The traditional view of cancer development is based on a stochastic model, which states that every malignant cell may undergo genetic and/or epigenetic alterations and clonally expand to initiate tumor growth. Thus, every cell within the tumor may be equally responsible for tumor initiation and progression[47-51]. Unlike the stochastic model, the hierchical or cancer stem cell (CSC) model may explain intra-tumor heterogeneity representing tumor as a hierchically organized tissue with CSCs at the apex in the pyramid and more committed and differentiated tumor cell types progressively down[47-50]. According to this model, CSCs represent a fraction of cells resident in the tumor endowed with stem-like features like the ability to self-renew and differentiate into heterogeneous tumor cell progeny as well as with the unresponsiveness to treatments[52,53], and represent the unit of selection within the tumor, while any other bulk tumor cells lead to clonal exhaustion[50]. More importantly, CSCs are thought to be a unique cellular subset responsible not only for tumor initiation but also for tumor growth maintenance, tumor recurrence and metastasis, showing intrinsic resistance to chemotherapeutic drugs compared to bulk tumor cells[52,54-56]. In this view, the existence of CSCs represent an entirely distinct dimension of intra-tumoral heterogeneity[57].

Interestingly, a third model has been recently proposed to explain the intra-tumor heterogeneity, the so-called “CSC plasticity model”. According with this theory, tumor cells represent a very plastic and dynamic population, with the ability to continuously shift between non-CSC and CSC states, in response to intrinsic and extrinsic stimuli. In this view, the stochastic and the CSC model not only are not mutually exclusive, but can be integrated with each other, adding a new level of tumor complexity[58].

The idea that tumor initiation and progression are driven by stem-like cells is still a subject of debate, since the first time it was proposed[59] until today. While CSC existence has been confirmed in a growing range of hematologic and solid tumors (e.g., acute myeloid leukemia, pancreatic cancer, breast cancer, lung cancer, hepatocellular carcinoma, head and neck cancer, colon cancer, prostate cancer, melanoma, and glioblastoma), no agreement has yet been reached regarding the origin of putative CSCs[60]. Some reports have indicated that CSCs can originate from normal resident stem cells, due to their inherent self-renewal capacity and long life span that can allow them to accumulate oncogenic and epigenetic modifications, resulting in malignant transformation. Alternatively, CSCs may originate from more committed progenitor cells[47], or even from differentiated non-CSCs that re-acquire stem cell properties by de-differentiation or reprogramming processes[61,62]. Thus, tumor hierarchical organization does not imply that CSCs originated from normal stem cells, and the CSC model does not address the cell-of-origin, that represents the normal cell that acquires the first cancer-promoting mutation(s) and is not necessarily related to the CSC concept[63,64]. These considerations interconnect with the debate on the true nature of the cell-of-origin of PLC. While it has already been accepted that HCC progression is driven by CSCs[22,65-69], very few studies have indicated the presence of CSCs in CCA[70] (reviewed in[71]).

Regulatory pathways involved in PLC-associated stemness

Many of the identified CSC regulatory pathways are also known to be involved in normal stem-cell maintenance as well as in self-renewal potential and pluripotency of embryonic stem cells[72-77]. Here, we will briefly review the key regulatory pathways that support stemness features in the context of PLC [Figure 1].

Stemness features in liver cancer

Figure 1. Regulatory pathways of liver cancer stem cells (CSCs). Primary liver cancers are heterogeneously composed by bulk tumor cells and CSCs. Liver CSCs are characterized by the activation of several molecular regulatory pathways that contribute to support the maintenance of CSC stemness features, including Wnt/β-catenin, Notch, Hedgehog, Hippo/Yes-associated protein (YAP), transforming growth factor-β (TGF-β), mitogen-activated protein kinase (MAPK)/extracellular signal-regulated kinase (ERK) and Janus kinase (JAK)/signal transducers and activators of transcription (STAT) signaling pathways

Wingless-type MMTV integration site family member (Wnt)/β-catenin pathway

Disruption of Wnt/β-catenin signaling results from both genetic and epigenetic changes in many tumors, including PLC. Wnt/β-catenin canonical signaling pathway appears to be involved in stemness maintenance in both embryonic and cancer stem cells[78,79]. Extracellular Wnt ligand binds to Frizzled cell surface receptors leading to increased cytoplasmic β-catenin levels, with the following induction of Wnt key target genes[31,55,80]. Notably, β-catenin is expressed in 58% of CCA, mutated in 8% of cases and it is considered an early determinant in CCA-progression[71]. In up to 90% of HCCs, the Wnt receptor FZD-7 is overexpressed, and 20%-40% of HCCs have unusual cytoplasmic and nuclear accumulation of β-catenin[81]. Moreover, in 25% of HCCs, β-catenin and Axin1 mutations are observed[69,81].

Notch signaling pathway

The Notch canonical signaling plays an important role in cell differentiation, proliferation and apoptosis, as well as in stem cell and HPCs maintenance[31,71,82,83]. Moreover, Notch signaling is implicated in bile duct morphogenesis (reviewed in[84]), and dysfunction in this pathway may result in reduced detoxification, ultimately leading to liver damage and iCCA development. Interestingly, the expression of Notch receptors 1 and 3 correlates with CCA progression and poor survival[71], whereas overexpression of Notch receptors 1 and 4 in HCC exerts tumorigenic effect[85]. Moreover, in up to 30% of HCCs, nuclear expression of Notch 1 and 3 is associated with the presence of stem cell signatures, supporting the role of Notch in promoting the expansion of the CSC niche[81,86]. Since Notch signaling can contribute to either CCA or HCC, it has been suggested that this pathway could be deregulated in bipotential HPCs[82].

Hedgehog signaling pathway

The Hedgehog (Hh) pathway regulates embryonic development, cell differentiation, regeneration and stem cell biology. The aberrant activation of the Hh pathway has been reported in different malignancies[87], and its correlation with prognosis is well known[88]. In addition to HCC carcinogenesis and HPC proliferation, activation of Hh pathway promotes CCA proliferation[71,79]. Notably, Sonic Hh (Shh) is the predominant ligand in the liver and is overexpressed in over 60% of HCCs[31,69,81,89].

Hippo signaling pathway

The Hippo signaling cascade is an evolutionarily conserved pathway involved in organ development[90-92]. This pathway has been implicated in multiple events during tumor onset. Strong evidence indicates a significant role of Hippo signaling in regulating stem cells, including HPCs[93-95]. Yes-associated protein 1 (YAP1) is a primary effector of the Hippo cascade and is frequently expressed in HCC and cHCC-CCA mixed tumor types, which retain stemness-related features[94]. Furthermore, constitutive activation of YAP in bile ducts, in association with AKT, seems to be essential in inducing CCA in a murine biliary injury model[31,96].

Phosphatidyl inositol 3-kinase/AKT signaling

AKT plays a critical role in many human cancers, including HCC and CCA[3,97]. AKT signaling can be triggered downstream of tyrosine kinase receptors activation, phosphatidyl inositol 3-kinase (PI3K) constitutive activation or loss of phosphatase and tensin homolog (PTEN)[3]. PTEN deletion results in the proliferation of a CD133+ population[71,98]. PI3K signaling promotes stem-like properties of HCC cells and it is implicated in HCC chemo- and radio-resistance as well as in epithelial-to-mesenchymal transition (EMT) and metastasis[99-104]. Notably, the co-activation of AKT and neuroblastoma rat sarcoma viral oncogene homolog (N-RAS) oncogenes leads to development of cHCC-CCA-like liver tumors, through the expansion of HPCs or malignant conversion of hepatocyte into progenitor-like cells[42].

Mitogen-activated protein kinase/extracellular signal-regulated kinases signaling pathway

The mitogen-activated protein kinase (MAPK) cascade regulates many important cell function, such as proliferation, invasion and survival and is critical for HPCs proliferation[71]. Gain-of-function mutations of KRAS are some of the most frequent mutations observed in iCCA, defining a class of patients characterized by poor outcome and enriched in CCA stem like-cells and tumor recurrence predicting signatures. Moreover, these mutations are also detected in patients with primary sclerosing cholangitis, suggesting that this could be an early event that contributes to the malignant transformation of cholangiocytes[36]. It is known that the MAPK pathway is directly associated with HCC cell growth and tumor-initiating capability[105-107]. Moreover, the long non-coding RNA H19 is highly expressed in HCC cells, where it activates the MAPK/extracellular signal-regulated kinases signaling pathway, regulating oxidative stress and chemotherapy resistance of CD133+ HCC CSC[108].

Transforming growth factor-β signaling

The transforming growth factor-β (TGF-β) pathway plays a key role in self-renewal and maintenance of an undifferentiated stem cell state. Its disruption is implicated in CCA development through impairment of stem cell differentiation and deregulated proliferation of HPCs[98]. Nonetheless, the role of TGF-β in PLC development is still controversial. Indeed, TGF-β acts as a tumor suppressor early in tumor initiation, whereas at late stages it promotes tumor growth, metastasis and EMT. It has been demonstrated that TGF-β1/Snail activation induces EMT in CCA both in vitro and in vivo, and this is associated with a higher CCA aggressiveness[109]. Moreover, TGF-β is upregulated in 40% of HCCs[69,81,89], and it may promote HCC progression via regulatory T cells recruitment and subsequent creation of a tumor suitable microenvironment[110,111].

Janus kinase/signal transducers and activators of transcription signaling

Several lines of evidences highlight the central role of interleukin (IL)-6/signal transducers and activators of transcription 3 (STAT3) signaling in CCA. Binding of IL-6 to the gp130 receptor leads to Janus kinases (JAKs) (JAK1, JAK2 and TYK2) and STAT3 activation, inducing the transcription of target genes essential for cell growth, differentiation and proliferation (reviewed in[34,112]). STAT3 signaling is also involved in maintenance of CSC population[113-115] and EMT-triggering in diverse tumors, including PLC[116,117]. Increased IL-6 expression has been reported to drive CSCs expansion through STAT3 activation in HCC[118]. Moreover, a recent study has demonstrated that EMT+ metastatic CSCs can be generated in a β2SP+/- mouse model of HCC, mainly due to overexpression of IL-6 in addition to the partial disruption of TGF-β signaling[119].

Key features of liver CSCs

Drug-resistance

A fundamental aspect contributing to poor PLC survival rate is the unresponsiveness to conventional therapies[11,12]. Currently, effective treatment is limited to surgical resection for both HCC and CCA, as well as liver transplantation for HCC. Unfortunately, 80% of HCC patients are diagnosed at an advanced tumor stage, which is not amenable to curative treatment[8-10]. Although other treatment procedures (e.g., cryosurgery, radiofrequency ablation and embolization) are also available, they are mostly palliative approaches and the treatment regime is shifting towards systemic chemotherapy[9]. Moreover, more than 70% of patients with early-stage HCC develop post-surgery recurrence[9,110]. Likewise, CCAs are generally asymptomatic in early stages and are usually diagnosed at an advanced unresectable stage. Moreover, although chemotherapy improves the patients’ quality of life, it still remains only a palliative treatment[5,13,120]. Therefore, the majority of patients with unresectable CCA undergoes a rapid decline in clinical conditions and dies within 12 months of the onset of symptoms. Thus, PLC still remains a fatal disease, mainly due to frequent tumor recurrence and chemoresistance.

CSCs represent a peculiar sub-compartment of tumor cell population crucially involved in recurrence, metastasis as well as drug resistance[86,121][Figure 2]. CSCs can escape drug-induced cell death through different intrinsic and external mechanisms. The intrinsic mechanisms consist of enhancement of DNA damage repair pathways, self-renewal ability of CSCs, high expression of drug efflux-related proteins and over activation of growth- and other stem-related pathway. The external mechanisms refer to the role of the tumor microenvironment (TME) on CSC drug resistance. This includes TME-derived EMT signals, hypoxia stimulation and angiogenesis trigger[122]. Consistently, increasing evidence suggests that sorafenib resistance in HCC correlates with the activation of EMT and enrichment of CSC traits[123-125].

Stemness features in liver cancer

Figure 2. Key features of liver cancer stem cells (CSCs). Liver CSCs present some common key functional features. CSCs represent a peculiar sub-compartment of tumor cells crucially involved in drug resistance. Current therapeutic strategies for the treatment of hepatic cancer mostly focus on the inhibition of tumor growth, resulting in the death of only bulk tumor cells. CSCs are able to survive thanks to intrinsic and extrinsic mechanisms of chemo-resistance and subsequently they can give rise to primary liver cancer recurrence. Moreover, CSCs are the only cells endowed with metastatic potential. While mature tumor cells mostly die through natural apoptosis into blood circulation, CSCs, however, have a significantly higher viability, enhanced homing ability into the bloodstream as circulating CSCs as well as higher distant metastasis initiation capability. Next, CSCs are characterized by metabolic changes, but, no consensus has been reached on the metabolic features of CSCs, which are very plastic and so able to modify their metabolic features according to specific needs. Then, CSCs can exhibit enhanced glycolytic activity as well as increased mitochondrial oxidative phosphorylation (OXPHOS) with subsequent increased fatty acid oxidation, depending on tumor context. All these features make CSCs particularly hard to eradicate. FAO: fatty acid oxidation

Several CSC markers seem to be implicated in drug resistance, such as CD13, that protects PLC CSCs from apoptosis and ROS-dependent DNA damage induced by different chemotherapeutic drugs (e.g., 5-FU)[86]. The HCC epithelial cell adhesion molecule (EpCAM)+ CSCs also show chemo-resistance against genotoxic agents like 5-FU[9]. Next, CD133+ HCC CSCs exhibited chemo-resistance to fluorouracil and doxorubicin through AKT and Bcl-2 pathway activation. Furthermore, CD133+ CSC and CSC spheres isolated from HCC cell lines display enhanced resistance to a panel of chemotherapeutic drugs (e.g., paclitaxel, methotrexate, vinblastine, cisplatin, carboplatin, docetaxel, irinotecan, etc.)[126,127]. According to these data, we have recently demonstrated that CCA CSCs isolated by tumor sphere assay possess higher resistance to common chemotherapeutic agents[70]. Additionally, laminin-332 expression is fundamental for maintaining self-renewal abilities of hepatic CSCs and for inducing mTOR-associated resistance to doxorubicin and sorafenib. Laminin-332 not only protects hepatic cancer cells against chemotherapy but also stimulates simultaneously cell proliferation upon sorafenib exposure, and it has been hypothesized that while laminin-332 may induce quiescence in PLC in “normal” circumstances, under cellular stress (e.g., sorafenib treatment) it could stimulate PLC cells to react by enhancing their proliferation[17,86].

Metastatic activity

The spread of circulating tumor cells (CTCs) in the blood plays a major role in tumor recurrence and metastasis initiation. Nevertheless, only a subset of CTCs can survive in the bloodstream, migrate to distant sites and establish secondary tumors. Consistent with CSC-hypothesis, stem-like CTCs might represent a potential source for cancer relapse and metastasis[121,128][Figure 2]. In fact, mature tumor cells have only a short blood circulation time and mostly die through natural apoptosis. CSCs, however, have shown to have significantly higher viability, enhanced homing ability into the bloodstream as well as higher distant metastasis initiation capability compared to other tumor cells[121,122,128]. According to the CSC-hypothesis, circulating CSCs (cCSCs) are particularly difficult to eradicate, with a consequent permanence of minimal residual disease and tumor recurrence[121,128].

Some putative markers has been proposed for identification of liver cCSCs. It has been demonstrated that CD90+ cCSCs express key stem-like genes (e.g., BMI1, CD44, OCT4, WNT3A, STAT3 and HIF-1α) at very high levels, also when compared to tissue CD90+ CSCs[121,129,130]. Moreover, CD90+ CXCR4+ cCSCs are able to initiate tumor metastasis formation in transplanted mice, enhancing the metastasis initiating ability of CSCs[121,131]. Considering that intercellular adhesion molecule 1 (ICAM1) inhibition by shRNA results in reduced metastasis in mice, ICAM1 has been proposed as another cCSC marker in PLC patients[121,132]. An explanation for the different metastatic activity observed between CSCs and other tumor cells might be the EMT status of CSCs, which enables them to have a prominent role in the metastasis and invasion[122]. Malignant cells undergo molecular changes typical of EMT, which represents a key stage of the metastatic multistep process, and eventually undergo a mesenchymal-to-epithelial transition (MET) to generate secondary tumors in target organs. Hence, CSCs mediate tumor metastasis by maintaining plasticity to transition between epithelial or mesenchymal states, and the EMT process represents the potential link between CSCs and circulating metastasis-initiating cells[121,133]. For example, in the CCA cell line TFK-1, TGF-β1 is able to induce not only EMT, but also CSC generation with a consequent decreased sensitivity to the chemotherapeutic agent 5-FU. Furthermore, the EMT-related overexpression of hepatic transmembrane 4L six family member 5 (TM4SF5) has a potential role in generating HCC cCSCs with metastatic properties through interaction with CD44[121,134]. In addition, HCC CSCs isolated by sphere assay are associated with an enhanced expression of the variant isoforms of CD44, which are related to CSC chemo-resistance, as well as with an increased frequency of intrahepatic metastasis when injected in the spleen of NOD-Rag1null IL2rγnull double mutant mice (NRG mice). Also in this case, enhancement of the EMT correlates to the metastatic potential and CSC state[135]. Another study has revealed that CD44 is associated with a mesenchymal phenotype in HCC cell lines, and knockdown of CD44 reverses EMT and inhibits lung metastasis of HCC cells in a murine model[136]. Another gene expression analysis of microarray data from 238 HCC cases has revealed an enriched EMT signature in CD90+ stem-like cells[137]. Finally, a recent study has found that CD44 protein levels are enhanced after TGF-β1 treatment and that interaction between CD44 and TGF-β1 induces EMT and CSC phenotypes through β-catenin signaling in HCC[138]. All these findings strengthen the hypothesis of an existing link between EMT and CSC cellular states in relation with the metastatic process.

Metabolic reprogramming

Starting from the pioneering work of Otto Warburg, several observations have indicated that tumor genetic alterations imply also cell metabolism reorganization[139,140]. In particular, it has been shown that tumor cells produce ATP via glycolysis and accumulate extracellular lactate even under normoxic conditions[140,141], and often present a limited or absent mitochondrial oxidative phosphorylation (OXPHOS)[140]. Although metabolic reprogramming is currently considered a hallmark of cancer, no consensus has been reached on the metabolic features of CSCs. which are very plastic and capable of either reside in a dormant state, or rapidly proliferate to replenish the tumor mass. A number of studies suggest that CSCs more strongly favor the glycolytic pathway compared to bulk tumor cells, while other studies report that mitochondrial oxidative metabolism is the prevalent source of energy for CSC (reviewed in[141]) [Figure 2]. However, even if investigation of PLC metabolism is still at its very beginning in comparison with other tumor systems, recent evidence has revealed the importance of the metabolic rearrangement in PLC CSCs. CD44+ CCA CSCs adapt their redox status regulation according to their needs and contribute to reactive oxygen species (ROS) defense promoting glutathione synthesis by way of xCT (a cysteine-glutamate transporter), resulting in evasion of cell death[142]. Moreover, CD133+ HCC CSCs are characterized by high glycolytic metabolism with concomitant overexpression of glycolytic genes and enhanced extracellular acidification rate, demonstrating that CD133+ cells are more glycolytic compared to CD133- cells. Further, CD133+ cells stemness features are significantly reduced when glycolysis is inhibited[143]. Extensive trascriptome and metabolome analysis of CD133+ HCC cells revealed the key role of MYC in the regulation of glycolytic metabolism in HCC CSCs[144].

There is also an increasing interest in lipid metabolism and specifically in alterations in lipid and cholesterol-associated pathways. It is well known that proliferating tumor cells require lipids and cholesterol, and they may increase the uptake of exogenous lipids and lipoproteins or hyper-activate metabolic pathways deputed to produce lipids and cholesterol. When specifically looking at the stem cell compartment, it has been demonstrated that stem-like cells rely on fatty acid oxidation (FAO) for the generation of ATP and NADH[145][Figure 2]. Metabolism analysis has revealed that NAD+ concentrations are increased in CD133+ cells, and this is directly correlated with SIRT1-dependent enhanced FAO[144]. In HCC, genome-wide transcriptional profiling and Ingenuity pathway analysis have suggested NANOG to be the connecting point between FAO and stem-like features, because of its simultaneous OXPHOS repression and FAO activation actions[145]. Moreover, it has been observed that stearoyl-CoA desaturase 1 (SCD1), a central enzyme involved in the conversion of saturated fatty acids into monounsaturated fatty acids (MUFAs), regulates liver CSCs[146]. In addition, enhanced activation of SCD1 and the consequent production of MUFAs appear to be a potential hallmark of CSCs[141].

All these findings prompt metabolic plasticity as a central force that enables CSCs to modify their replicative capabilities according to specific needs [Figure 2]. Further, emerging evidence suggests that CSCs may adopt specific metabolic phenotypes based on their location within tumor mass[147].

Conclusions and clinical implications

Unresponsiveness to current conventional therapies remains one of the major challenges in PLC. Current therapeutic strategies for the treatment of hepatic cancer mostly focus on the inhibition of tumor growth, with unsatisfactory results. Future treatments are likely to target CSCs and their specialized niche. In this view, it is imperative to decipher the molecular mechanism behind chemoresistance of PLC cells and especially of CSCs, with the objective to develop novel therapeutic strategies targeting features, markers or signaling pathways essentials for CSC biology.

Since CSCs are characterized by metabolic changes, drugs that inhibit OXPHOS have been studied as potential anticancer agents. Metformin, which interfere with OXPHOS by inhibiting NADH-coenzyme Q oxidoreductase (complex I), is a key example and has been shown to be particularly cytotoxic for CSCs, as well as for cells with mutations in OXPHOS complex I[148,149]. Despite metabolic studies in the field of liver CSC are still at an early stage, the dual inhibition of glycolytic and mitochondrial energy pathways may represent a promising superior therapeutic approach to effective eradicate heterogeneous liver CSCs and to overcome therapeutic resistance.

Moreover, since EMT pathway and CSC features seem to be intimately linked, improving our understanding of these cellular states may help to develop novel therapies. The plasticity of CSCs further suggests that simultaneously targeting CSCs existing in both epithelial and mesenchymal states rather than either state alone is needed to achieve complete tumor eradication[150]. Hence, future investigations in this direction are imperative.

It is important to underline that the development of CSC-specific therapeutic strategies imply the presence of a common recognized method for isolation and subsequent characterization of liver CSCs. During the last decade a large number of studies have aimed to identify liver CSCs and several attempts have been made to enrich liver CSCs. Common strategies for PLC CSC enrichment, varied from the widely used classical antigenic approach that relies on surface CSC markers detection (e.g., CD133[46,56,151-154], CD44[71,153-155], OV6[156], CD90[129,130,157], EpCAM[22,68,71,158], CD13[159], CD24[153,154,160], CD47[161]) to functional techniques including side population (SP) analysis[65,162-165], Aldefluor assay[166-169] and tumor-sphere formation[65,67,70,170,171]. In all different published studies, enriched PLC CSC subsets have been then tested in immune-deficient mice for the in vivo tumorigenic potential[22,56,65,67,68,129,130,151,152,155,156,159-162,166,170].

One important challenge in developing new therapeutic strategies is the dynamic and plastic behavior of tumor cells, especially of CSC. As it’s well known, a central role in the regulation of cancer cell plasticity is played not only by genetic alterations, but also by epigenetic changes, including DNA methylation, histone modifications and non-coding RNA (ncRNA) activity[58]. By acting at transcriptional, post-transcriptional and translational level, ncRNAs represent key regulators of CSCs by modulating several biological processes including asymmetric division, unresponsiveness to treatments and EMT, thus affecting tumor progression and recurrence[58]. In addition, recent studies also suggest that similar to normal stem cells, CSCs seem to reside in specialized microenvironment (“CSC-niche”)[46,50,70,172], whose signals can support self-renewal and drug-resistance features and, thereby, may influence the plasticity of CSCs[173-177]. Therefore, targeting only CSCs may not be enough, and continued development of therapies targeting CSCs and their microenvironment in combination with chemotherapy may be essential to improve the outcomes of PLC patients.

Declarations

Authors’ contributions

Analysis of publications and drafting of the manuscript: Correnti M, Booijink R, Di Maira G

Critically revised the manuscript: Raggi C, Marra F

Read and approved the final manuscript: All authors

Availability of data and materials

Not applicable.

Financial support and sponsorship

Raggi C and Correnti M were supported by the Italian Foundation of Cancer Research (AIRC). Work on liver cancer in Dr. Marra’s laboratory is supported by AIRC, Istituto Toscano Tumori (ITT), Fondazione Umberto Veronesi, University of Florence.

Conflicts of interest

All authors declared that there are no conflicts of interest.

Ethical approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Copyright

© The Author(s) 2018.

REFERENCES

1. El-Serag HB. Hepatocellular carcinoma. N Engl J Med 2011;365:1118-27.

2. Oikawa T. Cancer stem cells and their cellular origins in primary liver and biliary tract cancers. Hepatology 2016;64:645-51.

3. Kumar M, Zhao X, Wang XW. Molecular carcinogenesis of hepatocellular carcinoma and intrahepatic cholangiocarcinoma: one step closer to personalized medicine? Cell Biosci 2011;1:5.

4. Shaib Y, El-Serag HB. The epidemiology of cholangiocarcinoma. Semin Liver Dis 2004;24:115-25.

5. Banales JM, Cardinale V, Carpino G, Marzioni M, Andersen JB, et al. Expert consensus document: cholangiocarcinoma: current knowledge and future perspectives consensus statement from the European Network for the Study of Cholangiocarcinoma (ENS-CCA). Nat Rev Gastroenterol Hepatol 2016;13:261-80.

6. Machida K. Existence of cancer stem cells in hepatocellular carcinoma: myth or reality? Hepatol Int 2017;11:143-7.

7. Sia D, Llovet JM. Liver cancer: translating ‘-omics’ results into precision medicine for hepatocellular carcinoma. Nat Rev Gastroenterol Hepatol 2017;14:571-2.

8. Bruix J, Han KH, Gores G, Llovet JM, Mazzaferro V. Liver cancer: approaching a personalized care. J Hepatol 2015;62:S144-56.

9. Lohitesh K, Chowdhury R, Mukherjee S. Resistance a major hindrance to chemotherapy in hepatocellular carcinoma: an insight. Cancer Cell Int 2018;18:44.

10. Lu LC, Hsu CH, Hsu C, Cheng AL. Tumor Heterogeneity in hepatocellular carcinoma: facing the challenges. Liver Cancer 2016;5:128-38.

11. Andersen JB, Spee B, Blechacz BR, Avital I, Komuta M, et al. Genomic and genetic characterization of cholangiocarcinoma identifies therapeutic targets for tyrosine kinase inhibitors. Gastroenterology 2012;142:1021-31.e15.

12. Blechacz B, Gores GJ. Cholangiocarcinoma: advances in pathogenesis, diagnosis, and treatment. Hepatology 2008;48:308-21.

13. Rizvi S, Gores GJ. Pathogenesis, diagnosis, and management of cholangiocarcinoma. Gastroenterology 2013;145:1215-29.

14. Zabron A, Edwards RJ, Khan SA. The challenge of cholangiocarcinoma: dissecting the molecular mechanisms of an insidious cancer. Dis Model Mech 2013;6:281-92.

15. Khan SA, Thomas HC, Davidson BR, Taylor-Robinson SD. Cholangiocarcinoma. Lancet 2005;366:1303-14.

16. Blechacz B, Komuta M, Roskams T, Gores GJ. Clinical diagnosis and staging of cholangiocarcinoma. Nat Rev Gastroenterol Hepatol 2011;8:512-22.

17. Govaere O, Wouters J, Petz M, Vandewynckel YP, Van den Eynde K, et al. Laminin-332 sustains chemoresistance and quiescence as part of the human hepatic cancer stem cell niche. J Hepatol 2016;64:609-17.

18. Berthiaume EP, Wands J. The molecular pathogenesis of cholangiocarcinoma. Semin Liver Dis 2004;24:127-37.

19. Komuta M, Spee B, Vander Borght S, De Vos R, Verslype C, et al. Clinicopathological study on cholangiolocellular carcinoma suggesting hepatic progenitor cell origin. Hepatology 2008;47:1544-56.

20. Lee JS, Heo J, Libbrecht L, Chu IS, Kaposi-Novak P, et al. A novel prognostic subtype of human hepatocellular carcinoma derived from hepatic progenitor cells. Nat Med 2006;12:410-6.

21. Nomoto K, Tsuneyama K, Cheng C, Takahashi H, Hori R, et al. Intrahepatic cholangiocarcinoma arising in cirrhotic liver frequently expressed p63-positive basal/stem-cell phenotype. Pathol Res Pract 2006;202:71-6.

22. Yamashita T, Ji J, Budhu A, Forgues M, Yang W, et al. EpCAM-positive hepatocellular carcinoma cells are tumor-initiating cells with stem/progenitor cell features. Gastroenterology 2009;136:1012-24.

23. Zhou H, Wang H, Zhou D, Wang Q, Zou S, et al. Hepatitis B virus-associated intrahepatic cholangiocarcinoma and hepatocellular carcinoma may hold common disease process for carcinogenesis. Eur J Cancer 2010;46:1056-61.

24. Brunt E, Aishima S, Clavien PA, Fowler K, Goodman Z, et al. cHCC-CCA: consensus terminology for primary liver carcinomas with both hepatocytic and cholangiocytic differentation. Hepatology 2018;68:113-26.

25. Sato K, Marzioni M, Meng F, Francis H, Glaser S, et al. Ductular reaction in liver diseases: pathological mechanisms and translational significances. Hepatology 2018; doi: 10.1002/hep.30150.

26. Gouw AS, Clouston AD, Theise ND. Ductular reactions in human liver: diversity at the interface. Hepatology 2011;54:1853-63.

27. Kuwahara R, Kofman AV, Landis CS, Swenson ES, Barendswaard E, et al. The hepatic stem cell niche: identification by label-retaining cell assay. Hepatology 2008;47:1994-2002.

28. Theise ND, Saxena R, Portmann BC, Thung SN, Yee H, et al. The canals of Hering and hepatic stem cells in humans. Hepatology 1999;30:1425-33.

29. Roskams T. Liver stem cells and their implication in hepatocellular and cholangiocarcinoma. Oncogene 2006;25:3818-22.

30. Kitade M, Factor VM, Andersen JB, Tomokuni A, Kaji K, et al. Specific fate decisions in adult hepatic progenitor cells driven by MET and EGFR signaling. Genes Dev 2013;27:1706-17.

31. Raggi C, Invernizzi P, Andersen JB. Impact of microenvironment and stem-like plasticity in cholangiocarcinoma: molecular networks and biological concepts. J Hepatol 2015;62:198-207.

32. Woo HG, Lee JH, Yoon JH, Kim CY, Lee HS, et al. Identification of a cholangiocarcinoma-like gene expression trait in hepatocellular carcinoma. Cancer Res 2010;70:3034-41.

33. Zhang F, Chen XP, Zhang W, Dong HH, Xiang S, et al. Combined hepatocellular cholangiocarcinoma originating from hepatic progenitor cells: immunohistochemical and double-fluorescence immunostaining evidence. Histopathology 2008;52:224-32.

34. Sia D, Tovar V, Moeini A, Llovet JM. Intrahepatic cholangiocarcinoma: pathogenesis and rationale for molecular therapies. Oncogene 2013;32:4861-70.

35. Cardinale V, Carpino G, Reid L, Gaudio E, Alvaro D. Multiple cells of origin in cholangiocarcinoma underlie biological, epidemiological and clinical heterogeneity. World J Gastrointest Oncol 2012;4:94-102.

36. Andersen JB, Thorgeirsson SS. Genetic profiling of intrahepatic cholangiocarcinoma. Curr Opin Gastroenterol 2012;28:266-72.

37. Carpino G, Cardinale V, Onori P, Franchitto A, Berloco PB, et al. Biliary tree stem/progenitor cells in glands of extrahepatic and intraheptic bile ducts: an anatomical in situ study yielding evidence of maturational lineages. J Anat 2012;220:186-99.

38. Razumilava N, Gores GJ. Notch-driven carcinogenesis: the merging of hepatocellular cancer and cholangiocarcinoma into a common molecular liver cancer subtype. J Hepatol 2013;58:1244-5.

39. Zucman-Rossi J, Nault JC, Zender L. Primary liver carcinomas can originate from different cell types: a new level of complexity in hepatocarcinogenesis. Gastroenterology 2013;145:53-5.

40. Raggi C, Mousa HS, Correnti M, Sica A, Invernizzi P. Cancer stem cells and tumor-associated macrophages: a roadmap for multitargeting strategies. Oncogene 2016;35:671-82.

41. Holczbauer A, Factor VM, Andersen JB, Marquardt JU, Kleiner DE, et al. Modeling pathogenesis of primary liver cancer in lineage-specific mouse cell types. Gastroenterology 2013;145:221-31.

42. Fan B, Malato Y, Calvisi DF, Naqvi S, Razumilava N, et al. Cholangiocarcinomas can originate from hepatocytes in mice. J Clin Invest 2012;122:2911-5.

43. Sekiya S, Suzuki A. Intrahepatic cholangiocarcinoma can arise from Notch-mediated conversion of hepatocytes. J Clin Invest 2012;122:3914-8.

44. Fidler IJ, Hart IR. Biological diversity in metastatic neoplasms: origins and implications. Science 1982;217:998-1003.

45. Heppner GH, Miller BE. Tumor heterogeneity: biological implications and therapeutic consequences. Cancer Metastasis Rev 1983;2:5-23.

46. Kreso A, Dick JE. Evolution of the cancer stem cell model. Cell Stem Cell 2014;14:275-91.

47. Beachy PA, Karhadkar SS, Berman DM. Tissue repair and stem cell renewal in carcinogenesis. Nature 2004;432:324-31.

48. Clarke MF, Dick JE, Dirks PB, Eaves CJ, Jamieson CH, et al. Cancer stem cells--perspectives on current status and future directions: AACR Workshop on cancer stem cells. Cancer Res 2006;66:9339-44.

49. Morrison SJ, Kimble J. Asymmetric and symmetric stem-cell divisions in development and cancer. Nature 2006;441:1068-74.

50. 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;16:225-38.

51. Boesch M, Sopper S, Zeimet AG, Reimer D, Gastl G, et al. Heterogeneity of cancer stem cells: rationale for targeting the stem cell niche. Biochim Biophys Acta 2016;1866:276-89.

52. Yamashita T, Wang XW. Cancer stem cells in the development of liver cancer. J Clin Invest 2013;123:1911-8.

53. Mani SA, Guo W, Liao MJ, Eaton EN, Ayyanan A, et al. The epithelial-mesenchymal transition generates cells with properties of stem cells. Cell 2008;133:704-15.

54. O’Brien CA, Pollett A, Gallinger S, Dick JE. A human colon cancer cell capable of initiating tumour growth in immunodeficient mice. Nature 2007;445:106-10.

55. Borovski T, De Sousa EMF, Vermeulen L, Medema JP. Cancer stem cell niche: the place to be. Cancer Res 2011;71:634-9.

56. Ma S, Lee TK, Zheng BJ, Chan KW, Guan XY. CD133+ HCC cancer stem cells confer chemoresistance by preferential expression of the Akt/PKB survival pathway. Oncogene 2008;27:1749-58.

57. Pattabiraman DR, Weinberg RA. Tackling the cancer stem cells - what challenges do they pose? Nat Rev Drug Discov 2014;13:497-512.

58. da Silva-Diz V, Lorenzo-Sanz L, Bernat-Peguera A, Lopez-Cerda M, Munoz P. Cancer cell plasticity: impact on tumor progression and therapy response. Semin Cancer Biol 2018; doi: 10.1016/j.semcancer.2018.08.009.

59. Virchow R. An address on the value of pathological experiments. Br Med J 1881;2:198-203.

60. Sainz B, Jr., Carron E, Vallespinos M, Machado HL. Cancer stem cells and macrophages: implications in tumor biology and therapeutic strategies. Mediators Inflamm 2016;2016:9012369.

61. Rais Y, Zviran A, Geula S, Gafni O, Chomsky E, et al. Deterministic direct reprogramming of somatic cells to pluripotency. Nature 2013;502:65-70.

62. Chaffer CL, Brueckmann I, Scheel C, Kaestli AJ, Wiggins PA, et al. Normal and neoplastic nonstem cells can spontaneously convert to a stem-like state. Proc Natl Acad Sci U S A 2011;108:7950-5.

63. Wang JC, Dick JE. Cancer stem cells: lessons from leukemia. Trends Cell Biol 2005;15:494-501.

64. Cheng L, Ramesh AV, Flesken-Nikitin A, Choi J, Nikitin AY. Mouse models for cancer stem cell research. Toxicol Pathol 2010;38:62-71.

65. Marquardt JU, Raggi C, Andersen JB, Seo D, Avital I, et al. Human hepatic cancer stem cells are characterized by common stemness traits and diverse oncogenic pathways. Hepatology 2011;54:1031-42.

66. Ma S, Chan KW, Hu L, Lee TK, Wo JY, et al. Identification and characterization of tumorigenic liver cancer stem/progenitor cells. Gastroenterology 2007;132:2542-56.

67. Raggi C, Factor VM, Seo D, Holczbauer A, Gillen MC, et al. Epigenetic reprogramming modulates malignant properties of human liver cancer. Hepatology 2014;59:2251-62.

68. Yamashita T, Forgues M, Wang W, Kim JW, Ye Q, et al. EpCAM and alpha-fetoprotein expression defines novel prognostic subtypes of hepatocellular carcinoma. Cancer Res 2008;68:1451-61.

69. Hsieh A, Kim HS, Lim SO, Yu DY, Jung G. Hepatitis B viral X protein interacts with tumor suppressor adenomatous polyposis coli to activate Wnt/beta-catenin signaling. Cancer Lett 2011;300:162-72.

70. Raggi C, Correnti M, Sica A, Andersen JB, Cardinale V, et al. Cholangiocarcinoma stem-like subset shapes tumor-initiating niche by educating associated macrophages. J Hepatol 2017;66:102-15.

71. Kokuryo T, Yokoyama Y, Nagino M. Recent advances in cancer stem cell research for cholangiocarcinoma. J Hepatobiliary Pancreat Sci 2012;19:606-13.

72. Takahashi K, Yamanaka S. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 2006;126:663-76.

73. Schoenhals M, Kassambara A, De Vos J, Hose D, Moreaux J, et al. Embryonic stem cell markers expression in cancers. Biochem Biophys Res Commun 2009;383:157-62.

74. Ben-Porath I, Thomson MW, Carey VJ, Ge R, Bell GW, et al. An embryonic stem cell-like gene expression signature in poorly differentiated aggressive human tumors. Nat Genet 2008;40:499-507.

75. Jeter CR, Badeaux M, Choy G, Chandra D, Patrawala L, et al. Functional evidence that the self-renewal gene NANOG regulates human tumor development. Stem Cells 2009;27:993-1005.

76. Segal E, Friedman N, Koller D, Regev A. A module map showing conditional activity of expression modules in cancer. Nat Genet 2004;36:1090-8.

77. Wong DJ, Liu H, Ridky TW, Cassarino D, Segal E, et al. Module map of stem cell genes guides creation of epithelial cancer stem cells. Cell Stem Cell 2008;2:333-44.

78. Nusse R, van Ooyen A, Cox D, Fung YK, Varmus H. Mode of proviral activation of a putative mammary oncogene (int-1) on mouse chromosome 15. Nature 1984;307:131-6.

79. Mikhail S, He AR. Liver cancer stem cells. Int J Hepatol 2011;2011:486954.

80. Salomon DS, Kim N, Saeki T, Ciardiello F. Transforming growth factor-alpha: an oncodevelopmental growth factor. Cancer Cells 1990;2:389-97.

81. Oishi N, Wang XW. Novel therapeutic strategies for targeting liver cancer stem cells. Int J Biol Sci 2011;7:517-35.

82. Villanueva A, Llovet JM. Second-line therapies in hepatocellular carcinoma: emergence of resistance to sorafenib. Clin Cancer Res 2012;18:1824-6.

83. Morell CM, Strazzabosco M. Notch signaling and new therapeutic options in liver disease. J Hepatol 2014;60:885-90.

84. Zong Y, Stanger BZ. Molecular mechanisms of liver and bile duct development. Wiley Interdiscip Rev Dev Biol 2012;1:643-55.

85. Gao J, Song Z, Chen Y, Xia L, Wang J, et al. Deregulated expression of Notch receptors in human hepatocellular carcinoma. Dig Liver Dis 2008;40:114-21.

86. Cadamuro M, Brivio S, Spirli C, Joplin RE, Strazzabosco M, et al. Autocrine and paracrine mechanisms promoting chemoresistance in cholangiocarcinoma. Int J Mol Sci 2017;18.

87. Thayer SP, di Magliano MP, Heiser PW, Nielsen CM, Roberts DJ, et al. Hedgehog is an early and late mediator of pancreatic cancer tumorigenesis. Nature 2003;425:851-6.

88. Liao X, Siu MK, Au CW, Wong ES, Chan HY, et al. Aberrant activation of hedgehog signaling pathway in ovarian cancers: effect on prognosis, cell invasion and differentiation. Carcinogenesis 2009;30:131-40.

89. Oishi N, Yamashita T, Kaneko S. Molecular biology of liver cancer stem cells. Liver Cancer 2014;3:71-84.

90. Kango-Singh M, Singh A. Regulation of organ size: insights from the Drosophila Hippo signaling pathway. Dev Dyn 2009;238:1627-37.

91. Huang J, Wu S, Barrera J, Matthews K, Pan D. The Hippo signaling pathway coordinately regulates cell proliferation and apoptosis by inactivating Yorkie, the Drosophila Homolog of YAP. Cell 2005;122:421-34.

92. Lian I, Kim J, Okazawa H, Zhao J, Zhao B, et al. The role of YAP transcription coactivator in regulating stem cell self-renewal and differentiation. Genes Dev 2010;24:1106-18.

93. Lu L, Li Y, Kim SM, Bossuyt W, Liu P, et al. Hippo signaling is a potent in vivo growth and tumor suppressor pathway in the mammalian liver. Proc Natl Acad Sci U S A 2010;107:1437-42.

94. Kim GJ, Kim H, Park YN. Increased expression of Yes-associated protein 1 in hepatocellular carcinoma with stemness and combined hepatocellular-cholangiocarcinoma. PLoS One 2013;8:e75449.

95. Lee KP, Lee JH, Kim TS, Kim TH, Park HD, et al. The Hippo-Salvador pathway restrains hepatic oval cell proliferation, liver size, and liver tumorigenesis. Proc Natl Acad Sci U S A 2010;107:8248-53.

96. Li J, Razumilava N, Gores GJ, Walters S, Mizuochi T, et al. Biliary repair and carcinogenesis are mediated by IL-33-dependent cholangiocyte proliferation. J Clin Invest 2014;124:3241-51.

97. Stauffer JK, Scarzello AJ, Andersen JB, De Kluyver RL, Back TC, et al. Coactivation of AKT and beta-catenin in mice rapidly induces formation of lipogenic liver tumors. Cancer Res 2011;71:2718-27.

98. Lau CK, Yang ZF, Fan ST. Role of stem cells in normal liver and cancer. Anticancer Agents Med Chem 2011;11:522-8.

99. Zhang XL, Jia Q, Lv L, Deng T, Gao J. Tumorspheres derived from HCC cells are enriched with cancer stem cell-like cells and present high chemoresistance dependent on the Akt pathway. Anticancer Agents Med Chem 2015;15:755-63.

100. Feng X, Jiang J, Shi S, Xie H, Zhou L, et al. Knockdown of miR-25 increases the sensitivity of liver cancer stem cells to TRAIL-induced apoptosis via PTEN/PI3K/Akt/Bad signaling pathway. Int J Oncol 2016;49:2600-10.

101. Shi DM, Bian XY, Qin CD, Wu WZ. miR-106b-5p promotes stem cell-like properties of hepatocellular carcinoma cells by targeting PTEN via PI3K/Akt pathway. Onco Targets Ther 2018;11:571-85.

102. Zhu M, Li W, Lu Y, Dong X, Lin B, et al. HBx drives alpha fetoprotein expression to promote initiation of liver cancer stem cells through activating PI3K/AKT signal pathway. Int J Cancer 2017;140:1346-55.

103. Xia H, Ooi LL, Hui KM. MicroRNA-216a/217-induced epithelial-mesenchymal transition targets PTEN and SMAD7 to promote drug resistance and recurrence of liver cancer. Hepatology 2013;58:629-41.

104. Zhang Y, Zheng L, Ding Y, Li Q, Wang R, et al. MiR-20a induces cell radioresistance by activating the PTEN/PI3K/Akt signaling pathway in hepatocellular carcinoma. Int J Radiat Oncol Biol Phys 2015;92:1132-40.

105. Chiba T, Suzuki E, Yuki K, Zen Y, Oshima M, et al. Disulfiram eradicates tumor-initiating hepatocellular carcinoma cells in ROS-p38 MAPK pathway-dependent and -independent manners. PLoS One 2014;9:e84807.

106. Galuppo R, Maynard E, Shah M, Daily MF, Chen C, et al. Synergistic inhibition of HCC and liver cancer stem cell proliferation by targeting RAS/RAF/MAPK and WNT/beta-catenin pathways. Anticancer Res 2014;34:1709-13.

107. Zhang L, Zhang L, Li H, Ge C, Zhao F, et al. CXCL3 contributes to CD133(+) CSCs maintenance and forms a positive feedback regulation loop with CD133 in HCC via Erk1/2 phosphorylation. Sci Rep 2016;6:27426.

108. Ding K, Liao Y, Gong D, Zhao X, Ji W. Effect of long non-coding RNA H19 on oxidative stress and chemotherapy resistance of CD133+ cancer stem cells via the MAPK/ERK signaling pathway in hepatocellular carcinoma. Biochem Biophys Res Commun 2018;502:194-201.

109. Sato Y, Harada K, Itatsu K, Ikeda H, Kakuda Y, et al. Epithelial-mesenchymal transition induced by transforming growth factor-{beta}1/Snail activation aggravates invasive growth of cholangiocarcinoma. Am J Pathol 2010;177:141-52.

110. Chan SL, Chan AW, Yeo W. Novel therapeutic targets and predictive markers for hepatocellular carcinoma. Expert Opin Ther Targets 2015;19:973-83.

111. Chan LH, Luk ST, Ma S. Turning hepatic cancer stem cells inside out--a deeper understanding through multiple perspectives. Mol Cells 2015;38:202-9.

112. Andersen JB, Thorgeirsson SS. A perspective on molecular therapy in cholangiocarcinoma: present status and future directions. Future Medicine (Hepatic Oncology) 2014;1:143-57.

113. Zhou B, Damrauer JS, Bailey ST, Hadzic T, Jeong Y, et al. Erythropoietin promotes breast tumorigenesis through tumor-initiating cell self-renewal. J Clin Invest 2014;124:553-63.

114. Cook AM, Li L, Ho Y, Lin A, Stein A, et al. Role of altered growth factor receptor-mediated JAK2 signaling in growth and maintenance of human acute myeloid leukemia stem cells. Blood 2014;123:2826-37.

115. Sherry MM, Reeves A, Wu JK, Cochran BH. STAT3 is required for proliferation and maintenance of multipotency in glioblastoma stem cells. Stem Cells 2009;27:2383-92.

116. Liu RY, Zeng Y, Lei Z, Wang L, Yang H, et al. JAK/STAT3 signaling is required for TGF-beta-induced epithelial-mesenchymal transition in lung cancer cells. Int J Oncol 2014;44:1643-51.

117. Rokavec M, Oner MG, Li H, Jackstadt R, Jiang L, et al. IL-6R/STAT3/miR-34a feedback loop promotes EMT-mediated colorectal cancer invasion and metastasis. J Clin Invest 2014;124:1853-67.

118. Wang X, Sun W, Shen W, Xia M, Chen C, et al. Long non-coding RNA DILC regulates liver cancer stem cells via IL-6/STAT3 axis. J Hepatol 2016;64:1283-94.

119. Mitra A, Yan J, Xia X, Zhou S, Chen J, et al. IL6-mediated inflammatory loop reprograms normal to epithelial-mesenchymal transition(+) metastatic cancer stem cells in preneoplastic liver of transforming growth factor beta-deficient beta2-spectrin(+/-) mice. Hepatology 2017;65:1222-36.

120. Aljiffry M, Walsh MJ, Molinari M. Advances in diagnosis, treatment and palliation of cholangiocarcinoma: 1990-2009. World J Gastroenterol 2009;15:4240-62.

121. Correnti M, Raggi C. Stem-like plasticity and heterogeneity of circulating tumor cells: current status and prospect challenges in liver cancer. Oncotarget 2017;8:7094-115.

122. Qiu L, Li H, Fu S, Chen X, Lu L. Surface markers of liver cancer stem cells and innovative targeted-therapy strategies for HCC. Oncol Lett 2018;15:2039-48.

123. Chen J, Jin R, Zhao J, Liu J, Ying H, et al. Potential molecular, cellular and microenvironmental mechanism of sorafenib resistance in hepatocellular carcinoma. Cancer Lett 2015;367:1-11.

124. Chow AK, Ng L, Lam CS, Wong SK, Wan TM, et al. The Enhanced metastatic potential of hepatocellular carcinoma (HCC) cells with sorafenib resistance. PLoS One 2013;8:e78675.

125. Zhang W, Sun HC, Wang WQ, Zhang QB, Zhuang PY, et al. Sorafenib down-regulates expression of HTATIP2 to promote invasiveness and metastasis of orthotopic hepatocellular carcinoma tumors in mice. Gastroenterology 2012;143:1641-9 e5.

126. Hashimoto N, Tsunedomi R, Yoshimura K, Watanabe Y, Hazama S, et al. Cancer stem-like sphere cells induced from de-differentiated hepatocellular carcinoma-derived cell lines possess the resistance to anti-cancer drugs. BMC Cancer 2014;14:722.

127. Jang JW, Song Y, Kim SH, Kim JS, Kim KM, et al. CD133 confers cancer stem-like cell properties by stabilizing EGFR-AKT signaling in hepatocellular carcinoma. Cancer Lett 2017;389:1-10.

128. Oskarsson T, Batlle E, Massague J. Metastatic stem cells: sources, niches, and vital pathways. Cell Stem Cell 2014;14:306-21.

129. Yang ZF, Ho DW, Ng MN, Lau CK, Yu WC, et al. Significance of CD90(+) cancer stem cells in human liver cancer. Cancer Cell 2008;13:153-66.

130. Yang ZF, Ngai P, Ho DW, Yu WC, Ng MNP, et al. Identification of local and circulating cancer stem cells in human liver cancer. Hepatology 2008;47:919-28.

131. Zhu L, Zhang W, Wang J, Liu R. Evidence of CD90+CXCR4+ cells as circulating tumor stem cells in hepatocellular carcinoma. Tumour Biol 2015;36:5353-60.

132. Liu S, Li N, Yu X, Xiao X, Cheng K, et al. Expression of intercellular adhesion molecule 1 by hepatocellular carcinoma stem cells and circulating tumor cells. Gastroenterology 2013;144:1031-41 e10.

133. Shuang ZY, Wu WC, Xu J, Lin G, Liu YC, et al. Transforming growth factor-beta1-induced epithelial-mesenchymal transition generates ALDH-positive cells with stem cell properties in cholangiocarcinoma. Cancer Lett 2014;354:320-8.

134. Lee D, Na J, Ryu J, Kim HJ, Nam SH, et al. Interaction of tetraspan(in) TM4SF5 with CD44 promotes self-renewal and circulating capacities of hepatocarcinoma cells. Hepatology 2015;61:1978-97.

135. Nishiyama M, Tsunedomi R, Yoshimura K, Hashimoto N, Matsukuma S, et al. Metastatic ability and the epithelial-mesenchymal transition in induced cancer stem-like hepatoma cells. Cancer Sci 2018;109:1101-9.

136. Gao Y, Ruan B, Liu W, Wang J, Yang X, et al. Knockdown of CD44 inhibits the invasion and metastasis of hepatocellular carcinoma both in vitro and in vivo by reversing epithelial-mesenchymal transition. Oncotarget 2015;6:7828-37.

137. Yamashita T, Honda M, Nakamoto Y, Baba M, Nio K, et al. Discrete nature of EpCAM+ and CD90+ cancer stem cells in human hepatocellular carcinoma. Hepatology 2013;57:1484-97.

138. Park NR, Cha JH, Jang JW, Bae SH, Jang B, et al. Synergistic effects of CD44 and TGF-beta1 through AKT/GSK-3beta/beta-catenin signaling during epithelial-mesenchymal transition in liver cancer cells. Biochem Biophys Res Commun 2016;477:568-74.

139. Warburg O. On the origin of cancer cells. Science 1956;123:309-14.

140. Palorini R, Votta G, Balestrieri C, Monestiroli A, Olivieri S, et al. Energy metabolism characterization of a novel cancer stem cell-like line 3AB-OS. J Cell Biochem 2014;115:368-79.

141. Mancini R, Noto A, Pisanu ME, De Vitis C, Maugeri-Sacca M, et al. Metabolic features of cancer stem cells: the emerging role of lipid metabolism. Oncogene 2018;37:2367-78.

142. Thanee M, Loilome W, Techasen A, Sugihara E, Okazaki S, et al. CD44 variant-dependent redox status regulation in liver fluke-associated cholangiocarcinoma: a target for cholangiocarcinoma treatment. Cancer Sci 2016;107:991-1000.

143. Song K, Kwon H, Han C, Zhang J, Dash S, et al. Active glycolytic metabolism in CD133(+) hepatocellular cancer stem cells: regulation by MIR-122. Oncotarget 2015;6:40822-35.

144. Hur W, Ryu JY, Kim HU, Hong SW, Lee EB, et al. Systems approach to characterize the metabolism of liver cancer stem cells expressing CD133. Sci Rep 2017;7:45557.

145. Chen CL, Uthaya Kumar DB, Punj V, Xu J, Sher L, et al. NANOG metabolically reprograms tumor-initiating stem-like cells through tumorigenic changes in oxidative phosphorylation and fatty acid metabolism. Cell Metab 2016;23:206-19.

146. Ma MKF, Lau EYT, Leung DHW, Lo J, Ho NPY, et al. Stearoyl-CoA desaturase regulates sorafenib resistance via modulation of ER stress-induced differentiation. J Hepatol 2017;67:979-90.

147. Snyder V, Reed-Newman TC, Arnold L, Thomas SM, Anant S. Cancer stem cell metabolism and potential therapeutic targets. Front Oncol 2018;8:203.

148. Janzer A, German NJ, Gonzalez-Herrera KN, Asara JM, Haigis MC, et al. Metformin and phenformin deplete tricarboxylic acid cycle and glycolytic intermediates during cell transformation and NTPs in cancer stem cells. Proc Natl Acad Sci U S A 2014;111:10574-9.

149. Martinez-Outschoorn UE, Peiris-Pages M, Pestell RG, Sotgia F, Lisanti MP. Cancer metabolism: a therapeutic perspective. Nat Rev Clin Oncol 2017;14:11-31.

150. Jayachandran A, Dhungel B, Steel JC. Epithelial-to-mesenchymal plasticity of cancer stem cells: therapeutic targets in hepatocellular carcinoma. J Hematol Oncol 2016;9:74.

151. Suetsugu A, Nagaki M, Aoki H, Motohashi T, Kunisada T, et al. Characterization of CD133+ hepatocellular carcinoma cells as cancer stem/progenitor cells. Biochem Biophys Res Commun 2006;351:820-4.

152. Rountree CB, Ding W, He L, Stiles B. Expansion of CD133-expressing liver cancer stem cells in liver-specific phosphatase and tensin homolog deleted on chromosome 10-deleted mice. Stem Cells 2009;27:290-9.

153. Agrawal S, Kuvshinoff BW, Khoury T, Yu J, Javle MM, et al. CD24 expression is an independent prognostic marker in cholangiocarcinoma. J Gastrointest Surg 2007;11:445-51.

154. Riener MO, Vogetseder A, Pestalozzi BC, Clavien PA, Probst-Hensch N, et al. Cell adhesion molecules P-cadherin and CD24 are markers for carcinoma and dysplasia in the biliary tract. Hum Pathol 2010;41:1558-65.

155. Zhu Z, Hao X, Yan M, Yao M, Ge C, et al. Cancer stem/progenitor cells are highly enriched in CD133+CD44+ population in hepatocellular carcinoma. Int J Cancer 2010;126:2067-78.

156. Yang W, Yan HX, Chen L, Liu Q, He YQ, et al. Wnt/beta-catenin signaling contributes to activation of normal and tumorigenic liver progenitor cells. Cancer Res 2008;68:4287-95.

157. Ho DW, Yang ZF, Yi K, Lam CT, Ng MN, et al. Gene expression profiling of liver cancer stem cells by RNA-sequencing. PLoS One 2012;7:e37159.

158. de Boer CJ, van Krieken JH, Janssen-van Rhijn CM, Litvinov SV. Expression of Ep-CAM in normal, regenerating, metaplastic, and neoplastic liver. J Pathol 1999;188:201-6.

159. Haraguchi N, Ishii H, Mimori K, Tanaka F, Ohkuma M, et al. CD13 is a therapeutic target in human liver cancer stem cells. J Clin Invest 2010;120:3326-39.

160. Lee TK, Castilho A, Cheung VC, Tang KH, Ma S, et al. CD24(+) liver tumor-initiating cells drive self-renewal and tumor initiation through STAT3-mediated NANOG regulation. Cell Stem Cell 2011;9:50-63.

161. Lee TK, Cheung VC, Lu P, Lau EY, Ma S, et al. Blockade of CD47-mediated cathepsin S/protease-activated receptor 2 signaling provides a therapeutic target for hepatocellular carcinoma. Hepatology 2014;60:179-91.

162. Chiba T, Kita K, Zheng YW, Yokosuka O, Saisho H, et al. Side population purified from hepatocellular carcinoma cells harbors cancer stem cell-like properties. Hepatology 2006;44:240-51.

163. Cao L, Fan X, Jing W, Liang Y, Chen R, et al. Osteopontin promotes a cancer stem cell-like phenotype in hepatocellular carcinoma cells via an integrin-NF-kappaB-HIF-1alpha pathway. Oncotarget 2015;6:6627-40.

164. Guo Z, Jiang JH, Zhang J, Yang HJ, Yang FQ, et al. COX-2 promotes migration and invasion by the side population of cancer stem cell-like hepatocellular carcinoma cells. Medicine (Baltimore) 2015;94:e1806.

165. Yang X, Wang J, Qu S, Zhang H, Ruan B, et al. MicroRNA-200a suppresses metastatic potential of side population cells in human hepatocellular carcinoma by decreasing ZEB2. Oncotarget 2015;6:7918-29.

166. Ma S, Chan KW, Lee TK, Tang KH, Wo JY, et al. Aldehyde dehydrogenase discriminates the CD133 liver cancer stem cell populations. Mol Cancer Res 2008;6:1146-53.

167. Carpentino JE, Hynes MJ, Appelman HD, Zheng T, Steindler DA, et al. Aldehyde dehydrogenase-expressing colon stem cells contribute to tumorigenesis in the transition from colitis to cancer. Cancer Res 2009;69:8208-15.

168. Ginestier C, Hur MH, Charafe-Jauffret E, Monville F, Dutcher J, 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;1:555-67.

169. Li T, Su Y, Mei Y, Leng Q, Leng B, et al. ALDH1A1 is a marker for malignant prostate stem cells and predictor of prostate cancer patients’ outcome. Lab Invest 2010;90:234-44.

170. Cardinale V, Renzi A, Carpino G, Torrice A, Bragazzi MC, et al. Profiles of cancer stem cell subpopulations in cholangiocarcinomas. Am J Pathol 2015;185:1724-39.

171. Cao L, Zhou Y, Zhai B, Liao J, Xu W, et al. Sphere-forming cell subpopulations with cancer stem cell properties in human hepatoma cell lines. BMC Gastroenterol 2011;11:71.

172. Hanahan D, Coussens LM. Accessories to the crime: functions of cells recruited to the tumor microenvironment. Cancer Cell 2012;21:309-22.

173. Jinushi M, Chiba S, Yoshiyama H, Masutomi K, Kinoshita I, et al. Tumor-associated macrophages regulate tumorigenicity and anticancer drug responses of cancer stem/initiating cells. Proc Natl Acad Sci U S A 2011;108:12425-30.

174. Lu H, Clauser KR, Tam WL, Frose J, Ye X, et al. A breast cancer stem cell niche supported by juxtacrine signalling from monocytes and macrophages. Nat Cell Biol 2014;16:1105-17.

175. Su S, Liu Q, Chen J, Chen F, He C, et al. A positive feedback loop between mesenchymal-like cancer cells and macrophages is essential to breast cancer metastasis. Cancer Cell 2014;25:605-20.

176. Wan S, Zhao E, Kryczek I, Vatan L, Sadovskaya A, et al. Tumor-associated macrophages produce interleukin 6 and signal via STAT3 to promote expansion of human hepatocellular carcinoma stem cells. Gastroenterology 2014;147:1393-404.

177. Zhou W, Ke SQ, Huang Z, Flavahan W, Fang X, et al. Periostin secreted by glioblastoma stem cells recruits M2 tumour-associated macrophages and promotes malignant growth. Nat Cell Biol 2015;17:170-82.

Cite This Article

Export citation file: BibTeX | RIS

OAE Style

Correnti M, Booijink R, Di Maira G, Raggi C, Marra F. Stemness features in liver cancer. Hepatoma Res 2018;4:69. http://dx.doi.org/10.20517/2394-5079.2018.96

AMA Style

Correnti M, Booijink R, Di Maira G, Raggi C, Marra F. Stemness features in liver cancer. Hepatoma Research. 2018; 4: 69. http://dx.doi.org/10.20517/2394-5079.2018.96

Chicago/Turabian Style

Correnti, Margherita, Richell Booijink, Giovanni Di Maira, Chiara Raggi, Fabio Marra. 2018. "Stemness features in liver cancer" Hepatoma Research. 4: 69. http://dx.doi.org/10.20517/2394-5079.2018.96

ACS Style

Correnti, M.; Booijink R.; Di Maira G.; Raggi C.; Marra F. Stemness features in liver cancer. Hepatoma. Res. 2018, 4, 69. http://dx.doi.org/10.20517/2394-5079.2018.96

About This Article

Special Issue

This article belongs to the Special Issue New Insights in Liver Cancer
© The Author(s) 2018. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, sharing, adaptation, distribution and reproduction in any medium or format, for any purpose, even commercially, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Data & Comments

Data

Views
3831
Downloads
989
Citations
5
Comments
0
16

Comments

Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at support@oaepublish.com.

0
Download PDF
Cite This Article 6 clicks
Like This Article 16 likes
Share This Article
Scan the QR code for reading!
See Updates
Contents
Figures
Related
Hepatoma Research
ISSN 2454-2520 (Online) 2394-5079 (Print)

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/