Open access peer-reviewed chapter

Therapeutic Approaches Targeting Cancer Stem Cells

Written By

Shin Mukai

Submitted: 06 October 2022 Reviewed: 09 November 2022 Published: 24 December 2022

DOI: 10.5772/intechopen.108963

From the Edited Volume

Possibilities and Limitations in Current Translational Stem Cell Research

Edited by Diana Kitala

Chapter metrics overview

200 Chapter Downloads

View Full Metrics

Abstract

Cancer stem cells (CSCs) have been identified in many types of cancer since their discovery in leukemia in the 1990s. CSCs have self-renewal and differentiation capacity, and are thought to be a key driver for the establishment and growth of tumours. Several intracellular signalling pathways are reported to play a significant role in the regulation of the biological activities of CSCs. Thus, many researchers have considered CSCs to be a compelling therapeutic target for cancer, and blockade of CSC-related signalling pathways can be efficacious for the treatment of multiple cancer types. This chapter succinctly summarises the recent progress in the development of treatments targeting signalling pathways related to the functions of CSCs.

Keywords

  • therapeutic modalities
  • drug development
  • signalling pathways
  • self-renewal
  • cancer stem cells

1. Introduction

Cancer is a life-threatening disease in which abnormal cells grow and divide, resulting in the destruction of normal body tissues. The last several decades have seen advancements in cancer treatments [1]. However, the conventional therapeutic methods often fail due to cancer recurrence and metastasis [2]. This can be explained by the existence of cancer stem cells (CSCs), which are a minor population in tumours and can survive most traditional cancer therapies killing cancer cells with proliferative properties [3]. Cancer relapse, metastasis, multidrug resistance, and radiation resistance can be induced by the transient arrest of CSCs at the G0 phase, leading to the production of new malignant tumours [4]. The ability of CSCs to self-renew and differentiate into multiple cellular subtypes allows them to generate tumours [4]. Thus, researchers have regarded CSCs as a promising target for the treatment of cancer since they were discovered in leukemia in the 1990s [5, 6]. CSCs have also been found to be a subpopulation of many types of tumours, and tissue-specific expression of CSC markers has been reported [7]. It is known that the biological functions of CSCs are controlled by several signalling pathways [8]. This chapter focuses on the research status in cancer therapies targeting the signalling pathways that are believed to control the properties of CSCs.

Advertisement

2. General biology of CSCs

Since CSCs were found in leukemia in the 1990s, they have been studied intensively [5, 6]. However, the origin of CSCs remains to be elucidated [9]. It has been reported that (a) CSCs possess self-renewal capacity and high proliferation rate, (b) CSCs are able to generate and maintain tumours, and (c) cancer recurrence may be induced by the unlimited self-renewal capacity of CSCs [10]. CSCs are shown to be a distinct subpopulation in haematologic malignancies and solid tumours, and cell surface markers of CSCs in various types of cancer have been reported and can be used for the identification and isolation of CSCs (Figure 1) [11, 12, 13, 14, 15]. Signalling pathways such as Wnt/β-catenin, Notch and Hedgehog signalling pathways are thought to regulate the properties of CSCs [8]. Emerging evidence supports the clinical relevance of CSCs [16]. In particular, CSCs are shown to be resistant to conventional chemotherapy and radiation therapy, and they are believed to be a key player for cancer recurrence and metastasis [16]. Thus, understanding of signalling pathways that control the functions of CSCs can be useful for the creation of novel therapeutic interventions for cancer.

Figure 1.

Cell surface markers for CSCs in various organs. This figure was created with BioRender.

Advertisement

3. Development of therapeutic modalities targeting CSCs

3.1 Notch signalling

The Notch signalling pathway is a highly conserved pathway in mammals and controls a variety of cellular functions [17]. In mammals, the Notch pathway comprises four Notch receptors (Notch 1–4) and five Notch ligands (Jagged 1, Jagged 2, Delta-like ligand (DLL)-1, DLL-3, and DLL-4), and its primary effector is the transcription factor CSL (CBF1/RBP, Su(H)/Lag-1), which is crucial for the activation of the genes downstream of the Notch signalling pathway [18]. The Notch signalling pathway is divided into canonical and noncanonical pathways: the CSL-dependent pathway (canonical Notch signalling pathway) and non-CSL-dependent pathway (non-canonical Notch signalling pathway) [19]. Evidence suggests that these pathways play an instrumental role in preserving the existence of stem cells and initiating embryonic or foetal cell differentiation [20].

In the canonical Notch signalling pathway (Figure 2), the binding of receptor and ligand is induced through metalloproteinase- and γ-secretase-mediated proteolytic activation of the Notch receptor, leading to the release of the intracellular NOTCH domain (NICD) [21]. Subsequently, NICD migrates to the nucleus and forms a complex with CSL. As a result, the transcription and expression of the downstream target genes are triggered, leading to self-renewal, differentiation and proliferation [21].

Figure 2.

Brief diagram of the canonical Notch signalling pathway. This figure was created with BioRender.

Recent studies suggest that there is non-canonical Notch signalling, which can be activated either with or without ligand interaction [22]. In addition, the activation of non-canonical Notch signalling can occur in a γ-secretase-dependent or -independent manner [22]. In the case where non-canonical Notch signalling occurs in a γ-secretase-independent way, Notch remains bound to membrane [22]. Non-canonical Notch signalling does not require CSL [23, 24]. Instead, NICD or membrane bound Notch interacts with (a) Wnt, PI3K, mTORC2 and/or AKT pathways in the cytoplasm, and (b) NF-κB, YY1 and/or HIF-1α pathways in the nucleus [23, 24]. It is suggested that non-canonical Notch signalling plays a role in cell survival, metabolism and differentiation [23, 24]. Compared with the canonical Notch signalling pathway, there is less information on the non-canonical Notch signalling pathway [25]. Thus, more work will be needed for the identification of potential therapeutic targets in the non-canonical Notch signalling pathway [25].

Notch inhibition is believed to be a promising therapeutic approach to target CSCs, which are resistant to conventional methods such as chemotherapy and radiation [26]. As γ-secretase is a key player in Notch signalling, a great deal of effort has been invested in the development of γ-secretase inhibitors (GSIs) (Figure 3) [27]. It should be noted that GSIs show anti-CSC effects and that they were the first Notch inhibitors to reach clinical development [27]. However, one of the major concerns is the toxicity of GSIs [28]. In particular, serious toxicity in the gastrointestinal tract can be caused by GSIs [28]. Several GSIs have entered clinical trials thus far. Data from a Phase II clinical trial suggest that RO4929097 did not show sufficient efficacy for the treatment of metastatic melanoma and platinum-resistant ovarian cancer as monotherapy [29, 30]. Nirogacestat (PF-3084014) is another GSI undergoing clinical trials for desmoid tumours [31]. The result of Phase II clinical trials indicates that (a) treatment of desmoid tumour fibromatosis patients with Nirogacestat could be a promising approach, (b) the objective response rate was 71.4% and (c) relatively low doses and high tolerability were achieved, resulting in prolonged disease control [31, 32]. Another study shows the antitumour and antimetastatic effects of Nirogacestat in hepatocellular carcinoma [33]. The potent pan-Notch Inhibitor BMS-906024 has advanced into clinical trials, and the data suggest that it could be effective for the treatment of leukemia and solid tumours [34]. The clinical development status of other small molecule inhibitors is as follows: (a) The GSI MK-0752 is in a Phase I clinical trial for the treatment of pancreatic ductal adenocarcinoma [27], (b) the GSI LY900009 is in a Phase I clinical trial for ovarian cancer [35], (c) CB-103, which inhibits the interaction between NICD and CSL, is in a Phase I clinical trial for adenoid cystic carcinoma (ACC), colorectal cancer, breast cancer and prostate cancer [36], and (d) Crenigacestat (LY3039478), which is an oral Notch and GSI inhibitor, is in a Phase I clinical trial for solid tumours [37]. Recently, cryoelectron microscopy (cryo-EM) structures of γ-secretase in complex with each of the two GSI clinical candidates Semagacestat and Avagacestat have been reported, and these pieces of information might be useful for the design of novel GSIs [38].

Figure 3.

Structures of Notch signalling inhibitors in clinical development.

A study suggests that the expression of the Notch ligand DLL4 is increased in gastric cancer, enhancing self-renewal ability of CSCs [39]. Therefore, inhibition of DLL4 can be an effective approach to treating cancer [39]. Currently, Enoticumab (REGN421), a fully human IgG1 monoclonal antibody against DLL4, is in a phase I clinical trail for the treatment of solid tumours [40].

3.2 Wnt signalling

The Wnt signalling pathway is known to play a pivotal role in embryogenesis and tissue repair by controlling proliferation, differentiation, apoptosis and cell-to-cell interactions [41, 42]. The Wnt signalling pathway is classified into the canonical Wnt pathway (β-catenin dependent) and the noncanonical Wnt pathway (β-catenin independent) [43]. Wnt ligands are required for the activation of Wnt signalling, and the acyltransferase Porcupine is known to be essential for the production of Wnt ligands [44]. In canonical Wnt signalling (Figure 4), the absence of Wnt ligands leads to the degradation of β-catenin due to phosphorylation by glycogen synthase kinase 3β (GSK3β), and thereby translocation of β-catenin from the cytoplasm to the nucleus does not occur [45]. In the presence of Wnt ligands, their ligation to Frizzled proteins and LRP5/6 receptors induces the activation of the cytoplasmic protein DVL and the subsequent suppression of GSK3β [46]. It enables β-catenin to migrate to the nucleus and trigger target gene transcription by binding to TCF/LEF transcription factors [46]. Noncanonical Wnt signalling does not require the cytoplasmic stabilisation of β-catenin or its translocation into the nucleus [47]. The noncanonical Wnt signalling is subdivided into the following two well-characterised pathways: the planar cell polarity (Wnt/PCP) pathway and the Wnt-Calcium (Wnt/Ca2+) pathway [48]. In the Wnt/PCP pathway, Wnt ligands bind to the Frizzled receptor, leading to the activation of Dishevelled (DVL) protein [49]. Activated DVL forms the DVL-Rac complex and DVL-Rho complex [49]. The former stimulates the Rho kinase (ROCK), and the latter stimulates the c-Jun N-terminal Kinase (JNK) [50]. JNK is known to translocate into the nucleus and trigger the transcription of target genes [51]. The Wnt/Ca2+ pathway involves the activation of PLC and PKC, and an increase in intracellular Ca2+ [52]. The phosphatase calcineurin is stimulated by Ca2+ and dephosphorylates the transcriptions factor NF-AT, resulting in the migration of NF-AT to the nucleus [52].

Figure 4.

Brief diagram of the canonical Wnt signalling pathway. This figure was created with BioRender.

Dysregulation of Wnt signalling is observed in many types of cancer [53]. The Wnt signalling cascade is reported to play an important role in controlling the properties of CSCs [53]. Thus, many scholars have been striving to create therapeutic modalities to target Wnt signalling for safe and effective elimination of CSCs, and several small molecule inhibitors and monoclonal antibodies have entered clinical trials (Figure 5) [53]. Evidence suggests that pharmacological inhibition of Porcupine can selectively block Wnt signalling and suppress tumour growth [54]. Thus, Porcupine is considered to be a promising therapeutic target for cancer [54]. LGK-974, a small molecule disrupting the enzymatic activity of Porcupine, is currently in Phase I clinical trials for pancreatic cancer, melanoma and triple-negative breast cancer [55]. Cryo-electron microscopy (cryo-EM) structures of Porcupine in complex with LGK-974 are available in the PDB database (PDB ID: 7URD) [55], and this information could be useful for creation of novel Porcupine inhibitors. Another Porcupine inhibitor ETC-159 is in a Phase I clinical trial for the treatment of solid tumours [56]. The small molecule inhibitor PRI-724 is reported to block canonical Wnt signalling by preventing the interaction between β-catenin and its coactivator CREB binding protein (CBP) [57]. PRI-724 is in a Phase II clinical trial for advanced myeloid malignancies [58]. The therapeutic monoclonal antibody OMP-18R5 (vantictumab), which can target the Frizzled receptors, is now in a Phase I clinical trial for the treatment of non-small-cell lung cancer (NSCLC), pancreatic cancer and breast cancer [59, 60, 61, 62]. The recombinant fusion protein ipafricept (OMP-54F28) can inhibit Wnt signalling by binding to Wnt ligands, and its safety and effectiveness are being evaluated in clinical trials [63, 64, 65].

Figure 5.

Structures of Wnt signalling inhibitors in clinical development.

3.3 Hedgehog signalling

The Hedgehog (Hh) signalling pathway contributes to the control of cell proliferation, cell survival, cell differentiation, and stem cell maintenance and development [66]. In Hh signalling (Figure 6), the autoproteolytic cleavage of Hh ligand precursor proteins leads to the production of an N-terminal protein, followed by dual lipid modification [67]. Subsequently, active Hh ligands are released through mediation of Dispatched and Scube2 [67]. In the absence of Hh ligand, Patched (PTCH) prevents the activation and ciliary localisation of Smoothened (SMO) [68, 69]. As a result, the Glioma-Associated Oncogene Homolog (GLI) forms a complex with Suppressor of Fused (SUFU), which precludes GLI from translocating into the nucleus [68, 69]. In the presence of Hh ligand, the binding of Hh to PTCH allows SMO to interact with β-arrestin (Arrb2) and kinesin family member 3A (KIF3A), leading to the ciliary localisation of SMO [68, 69]. As a result, GLI is released from SUFU and subsequently migrates to the nucleus, triggering the transcription of Hh target genes [68, 69].

Figure 6.

Brief diagram of the canonical Hedgehog signalling pathway. This figure was created with BioRender.

Evidence suggests that the dysregulation of Hh signalling is associated with detrimental events such as the self-renewal and metastasis of cancer stem cells [67]. Thus, therapeutic targeting of Hh signalling has accorded a great deal of attention from many researchers, and SMO has been regarded as the most promising pharmacological target [68]. Indeed, several SMO inhibitors have been approved by the Food and Drug Administration (FDA) or are undergoing clinical trials (Figure 7) [70]. Vismodegib (GDC-0449) was the first SMO inhibitor that was granted FDA approval to treat basal cell carcinoma in 2011 [71]. Mounting evidence suggests the inhibitory activity of Vismodegib against self-renewal and mammosphere formation of breast CSCs [72]. Data from Phase II clinical trials demonstrate that (a) Vismodegib could be used as a neoadjuvant chemotherapy agent for patients with triple-negative breast cancer (NCT02694224) [73], (b) Vismodegib could be efficacious for the treatment of pancreatic cancer by suppressing self-renewal, proliferation and survival of pancreatic CSCs [74] and (c) Vismodegib could be effective for untreated metastatic colorectal cancer by reducing the stem cell markers of colon CSCs [75, 76]. These results support the notion that Vismodegib can inhibit the activities of CSCs by blocking the Hh signalling pathway. The SMO inhibitor Sonidegib (LDE225) was approved by FDA for the treatment of advanced basal cell carcinoma in 2015 [77]. A recent study indicates that Sonidegib can make triple-negative breast cancer more sensitive to Paclitaxel and improve clinical outcomes by reducing the expression of CSC markers [78]. Glasdegib (PF-04449913) was an FDA-approved SMO antagonist for the treatment of acute myeloid leukemia and launched in the USA in 2018 [79]. Glasdegib is also undergoing a Phase II clinical trial for the treatment of myelodysplastic syndrome and chronic myelomonocytic leukemia [80]. Although the development of SMO inhibitors is beneficial for cancer patients, monotherapy with each of the FDA-approved antagonists can cause SMO mutations in tumour tissues, leading to drug resistance [80]. Hence, novel therapeutic methods inhibiting SMO will need to be created in order to overcome this setback. The clinical development status of other SMO inhibitors is as follows [81]: (a) BMS-833923 is in a Phase I clinical trial for extensive-stage small cell lung cancer [82], (b) Itraconazole is in a Phase II clinical trial for prostate cancer [83], (c) Saridegib (IPI-926) is in a Phase I clinical trial for advanced and/or metastatic solid tumours [84], (d) LEQ-506 is in a Phase I clinical trial for advanced solid tumours [85], (e) Taladegib (LY2940680) is in a Phase I clinical trial for advanced solid tumours [86] and (f) TAK-441 is in a Phase I clinical trial for advanced nonhematologic malignancies [87]. In addition to the SMO inhibitors, arsenic trioxide is reported to inhibit the Hh signalling pathway and tumour growth by binding to GLI [81]. Arsenic trioxide is in a Phase II clinical trial for the treatment of advanced neuroblastoma or other childhood solid tumours [88]. Although there are no other GLI inhibitors undergoing clinical trials, in vitro and in vivo preclinical investigation into the GLI inhibitor GANT61 suggests its inhibitory activity against pancreatic cancer stem cells [89].

Figure 7.

Structures of Hedgehog signalling inhibitors in clinical development.

3.4 NF-κB signalling

The transcription factor nuclear factor kappa B (NF-κB) is a rapidly inducible transcription factor and a family of heterodimers or homodimers [90]. The heterodimers or homodimers are produced from different combinations of the five related proteins: p65/RelA, RelB, c-Rel, p105/p50 (NF-κB1) and p100/p52 (NF-κB2) [90]. The p50/p65 complex is thought to the most abundant form of NF-κB and serve main physiological functions [91]. The activation of NF-κB signalling induces the translocation of the transcription factor complexes from the cytoplasm to the nucleus [92]. The NF-κB signalling pathway diversifies into the canonical NF-κB signalling pathway and the noncanonical NF-κB signalling pathway [92]. The canonical NF-κB pathway is activated by the binding of ligands to their receptors such as the binding of (a) bacterial cell components to toll-like receptors (TLRs), (b) TNF-α to the TNF receptor (TNFR), (c) lipopolysaccharides to their respective receptors such as TLRs and (d) IL-1β to the IL-1 receptor (IL-1R) [93]. In response to the stimulation of these receptors, the kinase TGFβ-activated kinase 1 (TAK1) is activated, leading to the subsequent phosphorylation and activation of the IκB kinase (IKK) proteins [94]. The activated IKK proteins then phosphorylate IκB proteins. It induces the degradation of IκB, leading to the translocation of the activated p50/p60 complex into the nucleus [95]. The noncanonical NF-κB pathway is initiated by stimulation of receptors such as CD40, receptor activator for NF-κB (RANK), B cell activation factor (BAFF), TNFR2 and lymphotoxin β-receptor (LTBR) [94]. Subsequently, the kinase NF-κB-inducing kinase (NIK) is activated, resulting in the phosphorylation and activation of IKKα [94], which induces the phosphorylation of carboxy-terminal serine residues of p100 [94]. As a result, the C-terminal IκB-like structure of p100 is selectively degraded, which generates p52 and causes the p52-RelB complex to migrate to the nucleus [94].

A previous study shows that the production of cytokines, growth and angiogenic factors and proteases is promoted in the tumour development and progression, activating NF-κB signalling [96]. The NF-κB pathway is reported to contribute to self-renewal, maintenance and metastasis of CSCs [11]. Ovarian CSCs can display the enhanced capability of self-renewal, metastasis and maintenance due to the increased expression of RelA, RelB and IKKα [97]. In breast cancer, NIK expression is augmented, and the noncanonical NF-κB pathway is activated, leading to the self-renewal and metastasis of breast CSCs [98]. Regarding the development of therapeutic modalities to inhibit self-renewal, proliferation and metastasis of CSCs by targeting the NF-κB pathway (Figure 8A), Disulfiram is known to inhibit the activity of NF-κB in breast CSCs and is in a Phase II clinical trial for the treatment of metastatic breast cancer [99, 100]. Sulforaphane is suggested to prevent the translocation of p50/p65 and reduce the expression and transcriptional activity of p52/RelB, resulting in the inhibition of self-renewal of triple-negative breast CSCs [101]. Sulforaphane is in a Phase II clinical trial for the treatment of breast cancer [102]. The NF-κB pathway inhibitor curcumin is reported to impede self-renewal and metastasis of CSCs and is undergoing clinical trials for the treatment of breast cancer [103, 104, 105].

Figure 8.

Structures of CSC-related signalling inhibitors in clinical development. (A) NF-κB, (B) mTOR, (C) JAK/STAT, (D) ROCK.

3.5 mTOR signalling

The mammalian target of rapamycin (mTOR) signalling pathway plays a pivotal role in cellular growth and metabolism in mammalian cells within various environments [106]. The activation of mTOR signalling can promote cell survival by (a) increasing cellular metabolism and the synthesis of proteins and lipids and (b) blocking apoptotic pathways [106]. The mTOR pathway is initiated by the binding of growth factors to tyrosine kinase receptors in the cell membrane, leading to the phosphorylation and activation of phosphatidylinositol 3-kinase (PI3K) [107]. Subsequently, protein kinase B (AKT) is phosphorylated and activated, leading to the activation of mTOR [107]. It activates a variety of transcription factors and enables them to migrate to the nucleus in order to promote the transcription of the target genes [108]. mTOR is divided into two protein complexes. mTORC1 is a complex of RAPTOR, mLST8, PRAS40 and DEPTOR, and mTORC2 is a complex of RICTOR, mLST8, DEPTOR, mSin1 and PROCTOR [109, 110]. mTORC1 is reported to (a) activate the ribosomal protein S6K, which promotes protein synthesis, (b) facilitate lipid synthesis and mitochondrial biogenesis and (c) reduce autophagy [109, 110]. mTORC2 is known to promote actin cytoskeleton and cell migration by phosphorylating various proteins [109, 110].

Evidence suggests that the mTOR signalling pathway is correlated with the functions of CSCs. According to previous studies, activation of mTOR signalling contributes to (a) prostate cancer radioresistance due to the enhancement of CSC properties and (b) the tumourigenicity of breast CSCs [111, 112]. Dysregulation of the PI3K/Akt/mTOR signalling pathway is reported to enhance the expression of chemokine (C-X-C motif) receptor 4 (CXCR4), and CXCR4-mediated STAT3 signalling is then activated, promoting the self-renewal of CSCs in non-small-cell lung cancer [113]. Furthermore, the activity of mTOR through the PI3K feedback loop is associated with the survival of prostate CSCs [114].

The mTOR signalling pathway has been regarded as a promising candidate for therapeutic modalities targeting CSCs, and several mTOR inhibitors have been approved for the treatment of cancer or undergoing clinical development (Figure 8B) [115]. As demonstrated by a series of in vitro and in vivo experiments, the mTOR inhibitor Everolimus can suppress the expression and phosphorylation of AKT-1, and thereby inhibit the activity of HER2-overexpressing breast CSCs [116]. Everolimus has been approved by FDA for the treatment of advanced breast cancer [117]. The mTOR inhibitor Rapamycin is shown to suppress the properties of colon CSCs [118] and inhibit the stemness of haemangioma stem cells [119]. Rapamycin is undergoing clinical trials for advanced or metastatic colorectal cancer and infantile hepatic haemangioendothelioma [120, 121]. Previous reports show the inhibitory effects of the mTOR antagonist Metformin on breast and pancreatic CSCs [122, 123]. Clinical trials of Metformin are in progress or completed for the treatment of breast and pancreatic cancer [124, 125]. Considering previous studies show that mTOR signalling plays an important role in controlling the functions of CSCs, further investigation into a link between mTOR signalling and CSCs could lead to the development of better therapeutic strategies for cancer.

3.6 JAK/STAT signalling

The Janus kinase/signal transducer and activator of transcription (JAK/STAT) signalling pathway contributes to a variety of biological processes such as embryonic development, stem cell maintenance, haematopoiesis and inflammatory response [126]. It is known that the JAK family in mammals comprises four members (JAK1, JAK2, JAK3 and tyrosine kinase 2 (TYK2)) and that the STAT family in mammals consists of seven members (STAT1, STAT2, STAT3, STAT4, STAT5a, STAT5b and STAT6) [127]. The JAK/STAT signalling pathway is initiated by the binding of cytokines to their corresponding receptors [128]. The receptors then undergo dimerisation, and the JAKs bound to the receptors come close to each other, leading to the activation of the JAKs through the interaction of tyrosine phosphorylation [128]. The activated JAKs induce the phosphorylation of the tyrosine residues of the catalytic receptor, resulting in the formation of a docking site with the SH2 domain of the STAT protein [128]. Subsequently, the STAT protein bound to the receptor is phosphorylated and dimerised, culminating in the translocation of the activated STAT to the nucleus and the transcription of the target genes [128].

Literature precedent indicates that JAK/STAT signalling plays a role in controlling the properties of CSCs [129]. It has been suggested that (a) the self-renewal and survival of breast CSCs are facilitated by the persistent activation of JAK/STAT signalling [112], (b) IL-10-mediated JAK1/STAT1 signalling promotes the self-renewal and migration of non-small-cell lung cancer and colorectal CSCs [130, 131], (c) the OCT4-activated JAK1/STAT6 pathway is associated with the functions of ovarian CSCs [132], (d) OCT4, which is a gene downstream of IL-6-mediated JAK1/STAT6 signalling, is involved in the transformation of bulk cancer cells to CSCs in breast cancer [133] and (e) JAK2/STAT3 signalling plays a role in the regulation of the properties of breast and colorectal CSCs [134, 135, 136].

With respect to the development of therapeutic strategies to target JAK/STAT signalling for the treatment of cancer, the JAK1/2 inhibitor Ruxolitinib is shown to inhibit the functions of CSCs, leading to a decrease in the number of CSCs (Figure 8C) [137]. Ruxolitinib is undergoing clinical development for the treatment of solid tumours [138]. Many studies have shown that the JAK/STAT signalling pathway is correlated with the survival, self-renewal and metastasis of CSCs. Thus, more endeavours will be needed to develop novel therapeutic modalities targeting CSCs through the inhibition of JAK/STAT signalling.

3.7 ROCK signalling

The Rho-associated coiled-coil-containing protein kinase (ROCK) signalling pathway plays a pivotal role in various cellular activities such as cell survival and apoptosis [139]. It is known that there are two types of ROCK in mammals: ROCK1 and ROCK2 [139]. Diverse extracellular stimuli activate guanine nucleotide exchange factors (GEFs), leading to the conversion of Rho-GDP to Rho-GTP [140]. Rho-GTO subsequently activates ROCK1 and ROCK2, resulting in the phosphorylation of their substrates and the induction of a range of cellular responses [140].

Dysregulation of ROCK signalling is reported to be involved in the pathogenesis of a variety of diseases such as cancer [139]. According to an in vitro study, the properties of CSCs can be reduced by pharmacological inhibition of ROCK with the ROCK inhibitors ML7 or Y-27632, supporting the involvement of ROCK signalling in controlling the functions of CSCs [141]. With regard to the clinical development of ROCK inhibitors, the dual ROCK1/2 antagonist AT13148 is undergoing clinical trials for patients with advanced cancer [142, 143]. There is still sparce information on the roles of the ROCK signalling pathway in the regulation of the functions of CSCs. Thus, further investigation into a correlation between ROCK signalling and CSCs could be beneficial for the development of novel cancer treatments.

3.8 TGF-β signalling

The transforming growth factor β (TGF-β) signalling pathway contributes to diverse biological processes including cell proliferation and differentiation [11]. When dimeric TGF-β binds to the TGF-β receptor type-2 (TβRII), TβRII is dimerised with the TGF-β receptor type-1 (TβRI), leading to the phosphorylation and activation of the receptor-regulated SMADs (R-SMADs) SMAD2 and SMAD3 [11]. Subsequently, SMAD2 and SMAD3 undergo trimerisation with the common-partner SMAD, SMAD4 [11]. The trimer migrates to the nucleus and promotes the transcription of target genes [11].

It has been suggested that TGF-β may have contradictory functions in the properties of CSCs [144]. A study using a breast cancer xenograft model indicates that the activation of TGF-β signalling can reduce the number and the self-renewal potential of breast CSCs [145]. In addition, another in vivo study suggests that the number of CSCs in diffuse-type gastric carcinoma can be reduced by the activation of TGF-β signalling, leading to the suppression of tumour formation [146]. In contrast, it is reported that the activation of TGF-β signalling leads to an increase in CSC counts and the enhancement of CSC properties in various types of cancer such as breast cancer liver cancer, gastric cancer, skin cancer, glioblastoma and leukaemia [147, 148, 149, 150, 151, 152, 153, 154, 155, 156]. Considering these pieces of evidence, the therapeutic targeting of the TGF-β signalling pathway might be a promising strategy to eliminate CSCs. However, more work will be required to gain a better understanding of a link between TGF-β signalling and CSCs for the development of novel therapeutic modalities.

Advertisement

4. Conclusion

CSCs are a minor population in tumours, and their self-renewal capacity and differentiation potential contribute to tumour relapse, metastasis, chemoresistance and radioresistance. This chapter has provided a succinct summary of (a) major signalling pathways that are reported to be associated with the functions of CSCs and (b) clinical development of inhibitors targeting CSC-related signalling pathways for the purpose of encouraging research scientists (medicinal chemists, biologists, immunologists and others) to create new treatments (Table 1). Therapeutic targeting of CSCs via these signalling pathways has been considered to be a compelling strategy, and several small molecule inhibitors such as Vismodegib (GDC-0449), Sonidegib, Glasdegib (PF-04449913) and Everolimus have been approved by FDA for the treatment of cancer in the clinic. However, the development of such therapeutic interventions is challenging, and there is still vast scope for improvement. It is in part because signalling pathways interact with each other and because the CSC properties are thought to be controlled by the signalling network. AL/ML has been applied to the drug discovery, and it is reported that AL/ML is highly beneficial for target discovery, drug design and so on. These pieces of evidence can provide a scientific rational for applying AL/ML to the development of new therapeutic interventions targeting CSCs. The signal regulatory mechanisms of CSCs remain to be elucidated, and continuing studies of CSC-related pathways will lead to the creation of novel therapeutic modalities for various types of cancer.

Table 1.

Landscape of clinical development of inhibitors targeting CSC-related pathways.

References

  1. 1. Advancing cancer therapy. Nature Cancer. 2021;2:245-246. DOI: 10.1038/s43018-021-00192-x
  2. 2. Batlle E, Clevers H. Cancer stem cells revisited. Nature Medicine. 2017;23:1124-1134. DOI: 10.1038/nm.4409
  3. 3. Reya T, Morrison SJ, Clarke MF, Weissman IL. Stem cells, cancer, and cancer stem cells. Nature. 2001;414:105-111. DOI: 10.1038/35102167
  4. 4. Chen W, Dong J, Haiech J, Kilhoffer M-C, Zeniou M. Cancer stem cell quiescence and plasticity as major challenges in cancer therapy. Stem Cells International. 2016;2016:1740936. DOI: 10.1155/2016/1740936
  5. 5. Lapidot T, Sirard C, Vormoor J, Murdoch B, Hoang T, Caceres-Cortes J, et al. A cell initiating human acute myeloid leukaemia after transplantation into SCID mice. Nature. 1994;367:645-648. DOI: 10.1038/367645a0
  6. 6. Bonnet D, Dick JE. Human acute myeloid leukemia is organized as a hierarchy that originates from a primitive hematopoietic cell. Nature Medicine. 1997;3:730-737. DOI: 10.1038/nm0797-730
  7. 7. Zhao W, Li Y, Zhang X. Stemness-related markers in cancer. Cancer Translational Medicine. 2017;3:87-95. DOI: 10.4103/ctm.ctm_69_16
  8. 8. Matsui WH. Cancer stem cell signaling pathways. Medicine. 2016;95:S8-S19. DOI: 10.1097/MD.0000000000004765
  9. 9. Bjerkvig R, Tysnes BB, Aboody KS, Najbauer J, Terzis AJA. The origin of the cancer stem cell: Current controversies and new insights. Nature Reviews Cancer. 2005;5:899-904. DOI: 10.1038/nrc1740
  10. 10. Yu Z, Pestell TG, Lisanti MP, Pestellb RG. Cancer stem cells. The International Journal of Biochemistry & Cell Biology. 2012;44:2144-2151. DOI: 10.1016/j.biocel.2012.08.022
  11. 11. Yang L, Shi P, Zhao G, Xu J, Peng W, Zhang J, et al. Targeting cancer stem cell pathways for cancer therapy. Signal Transduction and Targeted Therapy. 2020;5:8. DOI: 10.1038/s41392-020-0110-5
  12. 12. Sun J-H, Luo Q , Liu L-L, Song G-B. Liver cancer stem cell markers: Progression and therapeutic implications. World Journal of Gastroenterology. 2016;22:3547-3557. DOI: 10.3748/wjg.v22.i13.3547
  13. 13. Zhou Y, Xia L, Wang H, Oyang L, Su M, Liu Q , et al. Cancer stem cells in progression of colorectal cancer. Oncotarget. 2018;9:33403-33415. DOI: 10.18632/oncotarget.23607
  14. 14. Ding Y, Gao H, Zhang Q. The biomarkers of leukemia stem cells in acute myeloid leukemia. Stem Cell Investigation. 2017;4:19. DOI: 10.21037/sci.2017.02.10
  15. 15. Abou-Antoun TJ, Dombrowski SM, Hale JS, Lathia JD. Brain cancer stem cells in adults and children: Cell biology and therapeutic implications. Neurotherapeutics. 2017;14:372-384. DOI: 10.1007/s13311-017-0524-0
  16. 16. Lathia J, Liu H, Matei D. The clinical impact of cancer stem cells. The Oncologist. 2020;25:123-131. DOI: 10.1634/theoncologist.2019-0517
  17. 17. Zhou B, Lin W, Long Y, Yang Y, Zhang H, Wu K, et al. Notch signaling pathway: Architecture, disease, and therapeutics. Signal Transduction and Targeted Therapy. 2022;24:95. DOI: 10.1038/s41392-022-00934-y
  18. 18. Hori K, Sen A, Artavanis-Tsakonas S. Notch signaling at a glance. Journal of Cell Science. 2013;126:2135-2140
  19. 19. Fiúza U-M, Arias AM. Cell and molecular biology of Notch. Journal of Endocrinology. 2007;194:459-474. DOI: 10.1677/JOE-07-0242
  20. 20. Nakata T, Shimizu H, Nagata S, Ito G, Fujii S, Suzuki K, et al. Indispensable role of Notch ligand-dependent signaling in the proliferation and stem cell niche maintenance of APC-deficient intestinal tumors. Biochemical and Biophysical Research Communications. 2017;482:1296-1303. DOI: 10.1016/j.bbrc.2016.12.031
  21. 21. Wang Z, Li Y, Banerjee S, Sarkar FH. Emerging role of Notch in stem cells and cancer. Cancer Letters. 2009;279:8-12. DOI: 10.1016/j.canlet.2008.09.030
  22. 22. Andersen P, Uosaki H, Shenje LT, Kwon C. Non-canonical Notch signaling: Emerging role and mechanism. Trends in Cell Biology. 2012;22:257-265. DOI: 10.1016/j.tcb.2012.02.003
  23. 23. Jeffries S, Capobianco AJ. Neoplastic transformation by Notch requires nuclear localization. Molecular and Cellular Biology. 2000;20:3928-3941. DOI: 10.1128/mcb.20.11.3928-3941.2000
  24. 24. Dumont E, Fuchs KP, Bommer G, Christoph B, Kremmer E, Kempkes B. Neoplastic transformation by Notch is independent of transcriptional activation by RBP-J signalling. Oncogene. 2000;19:556-561. DOI: 10.1038/sj.onc.1203352
  25. 25. Ayaz F, Osborne BA. Non-canonical notch signaling in cancer and immunity. Frontiers in Oncology. 2014;4:345. DOI: 10.3389/fonc.2014.00345
  26. 26. Purow B. Notch inhibition as a promising new approach to cancer therapy. Advances in Experimental Medicine and Biology. 2012;727:305-319. DOI: 10.1007/978-1-4614-0899-4_23
  27. 27. Cook N, Basu B, Smith D-M, Gopinathan A, Evans J, Steward WP, et al. A phase I trial of the γ-secretase inhibitor MK-0752 in combination with gemcitabine in patients with pancreatic ductal adenocarcinoma. British Journal of Cancer. 2018;118:793-801. DOI: 10.1038/bjc.2017.495
  28. 28. Takebe N, Nguyen D, Yang SX. Targeting notch signaling pathway in cancer: Clinical development advances and challenges. Pharmacology & Therapeutics. 2014;141:140-149. DOI: 10.1016/j.pharmthera.2013.09.005
  29. 29. Diaz-Padilla I, Wilson MK, Clarke BA, Hirte HW, Welch SA, Mackay HJ, et al. A phase II study of single-agent RO4929097, a gamma-secretase inhibitor of Notch signaling, in patients with recurrent platinum-resistant epithelial ovarian cancer: A study of the Princess Margaret, Chicago and California phase II consortia. Gynecologic Oncology. 2015;137:216-222. DOI: 10.1016/j.ygyno.2015.03.005
  30. 30. Lee SM, Moon J, Redman BG, Chidiac T, Flaherty LE, Zha Y, et al. Phase 2 study of RO4929097, a gamma-secretase inhibitor, in metastatic melanoma: SWOG 0933. Cancer. 2015;121:432-440. DOI: 10.1002/cncr.29055
  31. 31. Villalobos VM, Hall F, Jimeno A, Gore L, Kern K, Cesari R, et al. Long-term follow-up of desmoid fibromatosis treated with PF-03084014, an oral gamma secretase inhibitor. Annals of Surgical Oncology. 2018;25:768-775. DOI: 10.1245/s10434-017-6082-1
  32. 32. Kummar S, Coyne GOS, Do KT, Turkbey B, Meltzer PS, Polley E, et al. Clinical activity of the γ-secretase inhibitor PF-03084014 in adults with desmoid tumors (aggressive fibromatosis). Journal of Clinical Oncology. 2017;35:1561-1569. DOI: 10.1200/JCO.2016.71.199
  33. 33. Wu CX, Xu A, Zhang CC, Olson P, Chen L, Lee TK, et al. Notch inhibitor PF-03084014 inhibits hepatocellular carcinoma growth and metastasis via suppression of cancer stemness due to reduced activation of Notch1-Stat3. Molecular Cancer Therapeutics. 2017;6:1531-1543. DOI: 10.1158/1535-7163.MCT-17-0001
  34. 34. Morgan KM, Fischer BS, Lee FY, Shah JJ, Bertino JR, Rosenfeld J, et al. Gamma secretase inhibition by BMS-906024 enhances efficacy of paclitaxel in lung adenocarcinoma. Molecular Cancer Therapeutics. 2017;16:2759-2769. DOI: 10.1158/1535-7163.MCT-17-0439
  35. 35. Pant S, Jones SF, Kurkjian CD, Infante JR, Moore KN, Burris HA, et al. A first-in-human phase I study of the oral Notch inhibitor, LY900009, in patients with advanced cancer. European Journal of Cancer. 2016;56:1-9. DOI: 10.1016/j.ejca.2015.11.021
  36. 36. Miranda EL, Stathis A, Hess D, Racca F, Quon D, Rodon J, et al. Phase 1 study of CB-103, a novel first-in-class inhibitor of the CSL-NICD gene transcription factor complex in human cancers. Journal of Clinical Oncology. 2021;39:3020. DOI: 10.1200/JCO.2021.39.15_suppl.3020
  37. 37. Doi T, Tajimi M, Mori J, Asou H, Inoue K, Benhadji KA, et al. A phase 1 study of crenigacestat (LY3039478), the Notch inhibitor, in Japanese patients with advanced solid tumors. Investigational New Drugs. 2021;39:469-476. DOI: 10.1007/s10637-020-01001-5
  38. 38. Yang G, Zhou R, Guo X, Yan C, Lei J, Shi Y. Structural basis of gamma-secretase inhibition and modulation by small molecule drugs. Cell. 2021;184:521-533. DOI: 10.1016/j.cell.2020.11.049
  39. 39. Miao ZF, Xu H, Xu HM, Wang ZN, Zhao TT, Song YX, et al. DLL4 overexpression increases gastric cancer stem/progenitor cell self-renewal ability and correlates with poor clinical outcome via Notch-1 signaling pathway activation. Cancer Medicine. 2017;6:245-257. DOI: 10.1002/cam4.962
  40. 40. Chiorean EG, LoRusso P, Strother RM, Diamond JR, Younger A, Messersmith WA, et al. A phase I first-in-human study of Enoticumab (REGN421), a fully human delta-like ligand 4 (Dll4) monoclonal antibody in patients with advanced solid tumors. Clinical Cancer Research. 2015;21:2695-2703. DOI: 10.1158/1078-0432.CCR-14-2797
  41. 41. Murdoch B, Chadwick K, Martin M, Shojaei F, Shah KV, Gallacher L, et al. Wnt-5A augments repopulating capacity and primitive hematopoietic development of human blood stem cells in vivo. Proceedings of the National Academy of Sciences of the United States of America. 2003;100:3422-3427. DOI: 10.1073/pnas.0130233100
  42. 42. Ferrer-Mayorga G, Niell N, Cantero R, González-Sancho JM, Peso LD, Muñoz A, et al. Vitamin D and Wnt3A have additive and partially overlapping modulatory effects on gene expression and phenotype in human colon fibroblasts. Scientific Reports. 2019;9:8085. DOI: 10.1038/s41598-019-44574-9
  43. 43. Katoh M. Canonical and non-canonical WNT signaling in cancer stem cells and their niches: Cellular heterogeneity, omics reprogramming, targeted therapy and tumor plasticity (Review). International Journal of Oncology. 2017;51:1357-1369. DOI: 10.3892/ijo.2017.4129
  44. 44. Torres VI, Godoy JA, Inestrosa NC. Modulating Wnt signaling at the root: Porcupine and Wnt acylation. Pharmacology & Therapeutics. 2019;198:34-45. DOI: 10.1016/j.pharmthera.2019.02.009
  45. 45. Latres E, Chiaur DS, Pagano M. The human F box protein beta-Trcp associates with the Cul1/Skp1 complex and regulates the stability of beta-catenin. Oncogene. 1999;18:849-854. DOI: 10.1038/sj.onc.1202653
  46. 46. Pan T, Xu J, Zhu Y. Self-renewal molecular mechanisms of colorectal cancer stem cells. International Journal of Molecular Medicine. 2017;39:9-20. DOI: 10.3892/ijmm.2016.2815
  47. 47. Arnsdorf EJ, Tummala P, Jacobs CR. Non-canonical Wnt signaling and N-cadherin related bCatenin signaling play a role in mechanically induced osteogenic cell fate. PLoS One. 2009;4:e5388. DOI: 10.1371/journal.pone.0005388
  48. 48. MacDonald BT, Tamai K, He X. Wnt/β-catenin signaling: Components, mechanisms, and diseases. Developmental Cell. 2009;17:9-26. DOI: 10.1016/j.devcel.2009.06.016
  49. 49. Tree DRP, Shulman JM, Rousset R, Scott MP, Gubb D, Axelrod JD. Prickle mediates feedback amplification to generate asymmetric planar cell polarity signaling. Cell. 2002;109:371-381. DOI: 10.1016/s0092-8674(02)00715-8
  50. 50. Topol L, Jiang X, Choi H, Garrett-Beal L, Carolan PJ, Yang Y. Wnt-5a inhibits the canonical Wnt pathway by promoting GSK-3-independent beta-catenin degradation. Journal of Cell Biology. 2003;162:899-908. DOI: 10.1083/jcb.200303158
  51. 51. Kikuchi A, Yamamoto H, Sato A, Matsumoto S. New insights into the mechanism of Wnt signaling pathway activation. International Review of Cell and Molecular Biology. 2011;291:21-71. DOI: 10.1016/B978-0-12-386035-4.00002-1
  52. 52. Kim J, Kim DW, Chang W, Choe J, Kim J, Park C-S, et al. Wnt5a is secreted by follicular dendritic cells to protect germinal center B cells via Wnt/Ca2+/NFAT/NF-κB–B cell lymphoma 6 signaling. The Journal of Immunology. 2012;188:182-189. DOI: 10.4049/jimmunol.1102297
  53. 53. e Melo F d S, Vermeulen L. Wnt signaling in cancer stem cell biology. Cancers. 2016;8:60. DOI: 10.3390/cancers8070060
  54. 54. Liu Y, Qi X, Donnelly L, Elghobashi-Meinhardt N, Long T, Zhou RW, et al. Mechanisms and inhibition of Porcupine-mediated Wnt acylation. Nature. 2022;607:816-822. DOI: 10.1038/s41586-022-04952-2
  55. 55. Rodon J, Argilés G, Connolly RM, Vaishampayan U, Jonge M d, Garralda E, et al. Phase 1 study of single-agent WNT974, a first-in-class Porcupine inhibitor, in patients with advanced solid tumours. British Journal of Cancer. 2021;125:28-37. DOI: 10.1038/s41416-021-01389-8
  56. 56. Madan B, Ke Z, Harmston N, Ho SY, Frois AO, Alam J, et al. Wnt addiction of genetically defined cancers reversed by PORCN inhibition. Oncogene. 2016;35:2197-2207. DOI: 10.1038/onc.2015.280
  57. 57. Sasaki T, Kahn M. Inhibition of β-catenin/p300 interaction proximalizes mouse embryonic lung epithelium. Translational Respiratory Medicine. 2014;2:8. DOI: 10.1186/s40247-014-0008-1
  58. 58. Safety and Efficacy Study of PRI-724 in Subjects With Advanced Myeloid Malignancies. Available from: https://clinicaltrials.gov/ct2/show/NCT01606579
  59. 59. Pavlovic Z, Adams JJ, Blazer LL, Gakhal AK, Jarvik N, Steinhart Z, et al. A synthetic anti-frizzled antibody engineered for broadened specificity exhibits enhanced anti-tumor properties. MAbs. 2018;10:1157-1167. DOI: 10.1080/19420862.2018.1515565
  60. 60. A Study of Vantictumab (OMP-18R5) in Combination With Docetaxel in Patients With Previously Treated NSCLC. Available from: https://clinicaltrials.gov/ct2/show/NCT01957007
  61. 61. A Study of Vantictumab (OMP-18R5) in Combination With Nab-Paclitaxel and Gemcitabine in Previously Untreated Stage IV Pancreatic Cancer. Available from: https://clinicaltrials.gov/ct2/show/NCT02005315
  62. 62. A Study of Vantictumab (OMP-18R5) in Combination With Paclitaxel in Locally Recurrent or Metastatic Breast Cancer. Available from: https://clinicaltrials.gov/ct2/show/NCT01973309
  63. 63. Le PN, McDermott JD, Jimeno A. Targeting the Wnt pathway in human cancers: Therapeutic targeting with a focus on OMP-54F28. Pharmacology & Therapeutics. 2015;146:1-11. DOI: 10.1016/j.pharmthera.2014.08.005
  64. 64. A Dose Escalation Study of OMP-54F28 in Subjects With Solid Tumors. Available from: https://clinicaltrials.gov/ct2/show/NCT01608867
  65. 65. Dose Escalation Study of OMP-54F28 in Combination With Paclitaxel and Carboplatin in Patients With Recurrent Platinum-Sensitive Ovarian Cancer. Available from: https://clinicaltrials.gov/ct2/show/NCT02092363
  66. 66. Babu D, Fanelli A, Mellone S, Muniswamy R, Wasniewska M, Prodam F, et al. Novel GLI2 mutations identified in patients with Combined Pituitary Hormone Deficiency (CPHD) Evidence for a pathogenic effect by functional characterization. Clinical Endocrinology. 2019;90:449-456. DOI: 10.1111/cen.13914
  67. 67. Cochrane CR, Szczepny A, Watkins DN, Cain JE. Hedgehog signaling in the maintenance of cancer stem cells. Cancers. 2015;7:1554-1585. DOI: 10.3390/cancers7030851
  68. 68. Lee D-H, Lee S-Y, Oh SC. Hedgehog signaling pathway as a potential target in the treatment of advanced gastric cancer. Tumor Biology. 2017;39:1-10. DOI: 10.1177/1010428317692266
  69. 69. Han Y, Wang B, Cho YS, Zhu J, Wu J, Chen Y, et al. Phosphorylation of Ci/Gli by fused family kinases promotes Hedgehog signaling. Developmental Cell. 2019;50:610-626. DOI: 10.1016/j.devcel.2019.06.008
  70. 70. Yang Y, Li X, Wang T, Guo Q , Xi T, Zheng L. Emerging agents that target signaling pathways in cancer stem cells. Journal of Hematology and Oncology. 2020;13:60. DOI: 10.1186/s13045-020-00901-6
  71. 71. Ingram I: Vismodegib Granted FDA Approval for Treatment of Basal Cell Carcinoma. 2012. Available from: https://www.cancernetwork.com/view/vismodegib-granted-fda-approval-treatment-basal-cell-carcinoma
  72. 72. Li W, Yang H, Li X, Han L, Xu N, Shi A. Signaling pathway inhibitors target breast cancer stem cells in triple-negative breast cancer. Oncology Reports. 2019;41:437-446. DOI: 10.3892/or.2018.6805
  73. 73. Addition of Vismodegib to Neoadjuvant Chemotherapy in Triple Negative Breast Cancer Patients (SHH-CM). Available from: https://clinicaltrials.gov/ct2/show/NCT02694224
  74. 74. Singh BN, Fu J, Srivastava RK, Shankar S. Hedgehog signaling antagonist GDC-0449 (Vismodegib) inhibits pancreatic cancer stem cell characteristics: Molecular mechanisms. PLoS One. 2011;6:e27306. DOI: 10.1371/journal.pone.0027306
  75. 75. Wu C, Hu S, Cheng J, Wang G, Tao K. Smoothened antagonist GDC-0449 (Vismodegib) inhibits proliferation and triggers apoptosis in colon cancer cell lines. Experimental and Therapeutic Medicine. 2017;13:2529-2536. DOI: 10.3892/etm.2017.4282
  76. 76. Berlin J, Bendell JC, Hart LL, Firdaus I, Gore I, Hermann RC, et al. A randomized phase II trial of vismodegib versus placebo with FOLFOX or FOLFIRI and bevacizumab in patients with previously untreated metastatic colorectal cancer. Clinical Cancer Research. 2013;19:258-267. DOI: 10.1158/1078-0432.CCR-12-1800
  77. 77. Casey D, Demko S, Shord S, Zhao H, Chen H, He K, et al. FDA approval summary: sonidegib for locally advanced basal cell carcinoma. Clinical Cancer Research. 2017;23:2377-2381. DOI: 10.1158/1078-0432.CCR-16-2051
  78. 78. Cazet AS, Hui MN, Elsworth BL, Wu SZ, Roden D, Chan C-L, et al. Targeting stromal remodeling and cancer stem cell plasticity overcomes chemoresistance in triple negative breast cancer. Nature Communications. 2018;9:2897. DOI: 10.1038/s41467-018-05220-6
  79. 79. Norsworthy KJ, By K, Subramaniam S, Zhuang L, Valle PLD, Przepiorka D, et al. FDA approval summary: glasdegib for newly diagnosed acute myeloid leukemia. Clinical Cancer Research. 2019;25:6021-6025. DOI: 10.1158/1078-0432.CCR-19-0365
  80. 80. Du F-Y, Zhou Q-F, Sun W-J, Chen G-L. Targeting cancer stem cells in drug discovery: Current state and future perspectives. World Journal of Stem Cells. 2019;11:398-420. DOI: 10.4252/wjsc.v11.i7.398
  81. 81. Xie H, Paradise BD, Ma WW, Fernandez-Zapico ME. Recent advances in the clinical targeting of Hedgehog/GLI signaling in cancer. Cell. 2019;8:394. DOI: 10.3390/cells8050394
  82. 82. A Study of BMS-833923 With Carboplatin and Etoposide Followed by BMS-833923 Alone in Subjects With Extensive-Stage Small Cell Lung Cancer. Available from: https://clinicaltrials.gov/ct2/show/NCT00927875
  83. 83. Itraconazole in Treating Patients With Biochemically Relapsed Prostate Cancer. Available from: https://clinicaltrials.gov/ct2/show/NCT01787331
  84. 84. A Phase 1 Study of IPI-926 in Patients With Advanced and/or Metastatic Solid Tumor Malignancies. Available from: https://clinicaltrials.gov/ct2/show/NCT00761696
  85. 85. A Dose Finding and Safety Study of Oral LEQ506 in Patients With Advanced Solid Tumors. Available from: https://clinicaltrials.gov/ct2/show/NCT01106508
  86. 86. A Study of LY2940680 in Japanese Participants With Advanced Cancers. Available from: https://clinicaltrials.gov/ct2/show/NCT01919398
  87. 87. A Study of TAK-441 in Adult Patients With Advanced Nonhematologic Malignancies. Available from: https://clinicaltrials.gov/ct2/show/NCT01204073
  88. 88. Arsenic Trioxide in Treating Patients With Advanced Neuroblastoma or Other Childhood Solid Tumors. Available from: https://clinicaltrials.gov/ct2/show/NCT00024258
  89. 89. Fu J, Rodova M, Roy SK, Sharma J, Singh KP, Srivastava RK, et al. GANT-61 inhibits pancreatic cancer stem cell growth in vitro and in NOD/SCID/IL2R gamma null mice xenograft. Cancer Letters. 2013;330:22-32. DOI: 10.1016/j.canlet.2012.11.018
  90. 90. Oeckinghaus A, Ghosh S. The NF-κB family of transcription factors and its regulation. Cold Spring Harvor Perspectives in Biology. 2009;1:a000034. DOI: 10.1101/cshperspect.a00003410.1101/cshperspect.a000034
  91. 91. Siebenlist U, Franzoso G, Brown K. Structure, regulation and function of NF-kappa B. Annual Review of Cell Biology. 1994;10:405-455. DOI: 10.1146/annurev.cb.10.110194.002201
  92. 92. Novack DV. Role of NF-κB in the skeleton. Cell Research. 2011;21:169-182. DOI: 10.1038/cr.2010.159
  93. 93. Perkins ND, Gilmore TD. Good cop, bad cop: The different faces of NFkappaB. Cell Death & Differentiation. 2006;13:759-772. DOI: 10.1038/sj.cdd.4401838
  94. 94. Sun S-C. The non-canonical NF-κB pathway in immunity and inflammation. Nature Reviews Immunology. 2017;17:545-558. DOI: 10.1038/nri.2017.52
  95. 95. Hayden MS, Ghosh S. Shared principles in NF-kappaB signaling. Cell. 2008;132:344-362. DOI: 10.1016/j.cell.2008.01.020
  96. 96. Greten FR, Eckmann L, Greten TF, Park JM, Li Z-W, Egan LJ, et al. IKKbeta links inflammation and tumorigenesis in a mouse model of colitis-associated cancer. Cell. 2004;118:285-296. DOI: 10.1016/j.cell.2004.07.013
  97. 97. Gonzalez-Torres C, Gaytan-Cervantes J, Vazquez-Santillan K, Mandujano-Tinoco EA, Ceballos-Cancino G, Garcia-Venzor A, et al. NF-kappaB participates in the stem cell phenotype of ovarian cancer cells. Archives of Medical Research. 2017;48:343-351. DOI: 10.1016/j.arcmed.2017.08.001
  98. 98. Vazquez-Santillan K, Melendez-Zajgla J, Jimenez-Hernandez LE, Gaytan-Cervantes J, Muñoz-Galindo L, Piña-Sanchez P, et al. NF-kappaBeta-inducing kinase regulates stem cell phenotype in breast cancer. Scientific Reports. 2016;6:37340. DOI: 10.1038/srep37340
  99. 99. Han D, Wu G, Chang C, Zhu F, Xiao Y, Li Q , et al. Disulfiram inhibits TGF-β-induced epithelial–mesenchymal transition and stem-like features in breast cancer via ERK/NF-κB/Snail pathway. Oncotarget. 2015;6:40907-40919. DOI: 10.18632/oncotarget.5723
  100. 100. Phase II Trial of Disulfiram With Copper in Metastatic Breast Cancer (DISC). Available from: https://clinicaltrials.gov/ct2/show/NCT03323346
  101. 101. Burnett JP, Lim G, Li Y, Shah RB, Lim R, Paholak HJ, et al. Sulforaphane enhances the anticancer activity of taxanes against triple negative breast cancer by killing cancer stem cells. Cancer Letters. 2017;394:52-64. DOI: 10.1016/j.canlet.2017.02.023
  102. 102. Study to Evaluate the Effect of Sulforaphane in Broccoli Sprout Extract on Breast Tissue. Available from: https://clinicaltrials.gov/ct2/show/NCT00982319
  103. 103. Marquardt JU, Gomez-Quiroz L, Camacho LOA, Pinna F, Lee Y-H, Kitade M, et al. Curcumin effectively inhibits oncogenic NF-κB signaling and restrains stemness features in liver cancer. Journal of Hepatology. 2015;63:661-669. DOI: 10.1016/j.jhep.2015.04.018
  104. 104. Phase II Study of Curcumin vs Placebo for Chemotherapy-Treated Breast Cancer Patients Undergoing Radiotherapy. Available from: https://clinicaltrials.gov/ct2/show/NCT01740323
  105. 105. A "Window Trial" on Curcumin for Invasive Breast Cancer Primary Tumors. Available from: https://clinicaltrials.gov/ct2/show/NCT03980509
  106. 106. Laplante M, Sabatini DM. mTOR signaling at a glance. Journal of Cell Science. 2009;122:3589-3594. DOI: 10.1101/cshperspect.a011593
  107. 107. Li J, Kim SG, Blenis J. Rapamycin: One drug, many effects. Cell Metabolism. 2014;19:373-379. DOI: 10.1016/j.cmet.2014.01.001
  108. 108. Inoki K, Corradetti MN, Guan K-L. Dysregulation of the TSC-mTOR pathway in human disease. Nature Genetics. 2005;37:19-24. DOI: 10.1038/ng1494
  109. 109. Karagianni F, Pavlidis A, Malakou LS, Piperi C, Papadavid E. Predominant role of mTOR signaling in skin diseases with therapeutic potential. International Journal of Molecular Sciences. 2022;23:1693. DOI: 10.3390/ijms23031693
  110. 110. Papadopoli D, Boulay K, Kazak L, Pollak M, Mallette F, Topisirovic I, Hulea L: mTOR as a central regulator of lifespan and aging. F1000Research: 2019: 8: F1000 Faculty Rev-1998. doi: 10.12688/f1000research.17196.1
  111. 111. Chang L, Graham PH, Hao J, Ni J, Bucci J, Cozzi PJ, et al. Acquisition of epithelial-mesenchymal transition and cancer stem cell phenotypes is associated with activation of the PI3K/Akt/mTOR pathway in prostate cancer radioresistance. Cell Death & Disease. 2013;4:e875. DOI: 10.1038/cddis.2013.407
  112. 112. Zhou J, Wulfkuhle J, Zhang H, Gu P, Yang Y, Deng J, et al. Activation of the PTEN/mTOR/STAT3 pathway in breast cancer stem-like cells is required for viability and maintenance. Proceedings of the National Academy of Sciences of the United States of America. 2007;104:16158-16163. DOI: 10.1073/pnas.0702596104
  113. 113. Jung M-J, Rho J-K, Kim Y-M, Jung JE, Jin YB, Ko Y-G, et al. Upregulation of CXCR4 is functionally crucial for maintenance of stemness in drug-resistant non-small cell lung cancer cells. Oncogene. 2013;32:209-221. DOI: 10.1038/onc.2012.37
  114. 114. Marhold M, Tomasich E, El-Gazzar A, Heller G, Spittler A, Horvat R, et al. HIF-1alpha regulates mTOR signaling and viability of prostate cancer stem cells. Molecualr Cancer Research. 2015;13:556-564. DOI: 10.1158/1541-7786.MCR-14-0153-T
  115. 115. Zhu Y, Zhang X, Liu Y, Zhang S, Liu J, Ma Y, et al. Antitumor effect of the mTOR inhibitor everolimus in combination with trastuzumab on human breast cancer stem cells in vitro and in vivo. Tumor Biology. 2012;33:1349-1362. DOI: 10.1007/s13277-012-0383-6
  116. 116. Kim W-T, Ryu CJ. Cancer stem cell surface markers on normal stem cells. BMB Reports. 2017;50:285-298. DOI: 10.5483/bmbrep.2017.50.6.039
  117. 117. Leach B. FDA Approves Everolimus for Advanced Breast Cancer. 2012. Available from: https://www.onclive.com/view/fda-approves-everolimus-for-advanced-breast-cancer
  118. 118. Wang Y, Liu Y, Lu J, Zhang P, Wang Y, Xu Y, et al. Rapamycin inhibits FBXW7 loss-induced epithelial-mesenchymal transition and cancer stem cell-like characteristics in colorectal cancer cells. Biochemical and Biophysical Research Communications. 2013;434:352-356. DOI: 10.1016/j.bbrc.2013.03.077
  119. 119. Greenberger S, Yuan S, Walsh LA, Boscolo E, Kang K-T, Matthews B, et al. Rapamycin suppresses self-renewal and vasculogenic potential of stem cells isolated from infantile hemangioma. Journal of Investigative Dermatology. 2011;131:2467-2476. DOI: 10.1038/jid.2011.300
  120. 120. ABI-009 (Nab-rapamycin) in Combination With FOLFOX and Bevacizumab as First-line Therapy in Patients With Advanced or Metastatic Colorectal Cancer. Available from: https://clinicaltrials.gov/ct2/show/NCT03439462
  121. 121. Sirolimus in the Treatment for Infantile Hepatic Hemangioendothelioma(IEEH) (PRL-SRL-IHHE). Available from: https://clinicaltrials.gov/ct2/show/NCT04406870
  122. 122. Iliopoulos D, Hirsch HA, Struhl K. Metformin decreases the dose of chemotherapy for prolonging tumor remission in mouse xenografts involving multiple cancer cell types. Cancer Research. 2011;71:3196-3201. DOI: 10.1158/0008-5472.CAN-10-3471
  123. 123. Mohammed A, Janakiram NB, Brewer M, Ritchie RL, Marya A, Lightfoot S, et al. Antidiabetic drug metformin prevents progression of pancreatic cancer by targeting in part cancer stem cells and mTOR signaling. Translational Oncology. 2013;6:649-659. DOI: 10.1593/tlo.13556
  124. 124. A Phase III Randomized Trial of Metformin vs Placebo in Early Stage Breast Cancer. Available from: https://clinicaltrials.gov/ct2/show/NCT01101438
  125. 125. Metformin Combined With Chemotherapy for Pancreatic Cancer (GEM). Available from: https://clinicaltrials.gov/ct2/show/NCT01210911
  126. 126. Villarino AV, Kanno Y, O'Shea JJ. Mechanisms and consequences of Jak-STAT signaling in the immune system. Nature Immunology. 2017;18:374-384. DOI: 10.1038/ni.3691
  127. 127. Boehi F, Manetsch P, Hottiger MO. Interplay between ADP-ribosyltransferases and essential cell signaling pathways controls cellular responses. Cell Discovery. 2021;7:104. DOI: 10.1038/s41421-021-00323-9
  128. 128. Morris R, Kershaw NJ, Babon JJ. The molecular details of cytokine signaling via the JAK/STAT pathway. Protein Science. 2018;27:1984-2009. DOI: 10.1002/pro.3519
  129. 129. Wan S, Zhao E, Kryczek I, Vatan L, Sadovskaya A, Ludema G, et al. Tumor-associated macrophages produce interleukin 6 and signal via STAT3 to promote expansion of human hepatocellular carcinoma stem cells. Gastroenterology. 2014;147:1393-1404. DOI: 10.1053/j.gastro.2014.08.039
  130. 130. Yang L, Dong Y, Li Y, Wang D, Liu S, Wang D, et al. IL-10 derived from M2 macrophage promotes cancer stemness via JAK1/STAT1/NF-kappaB/Notch1 pathway in non-small cell lung cancer. International Journal of Cancer. 2019;145:1099-1110. DOI: 10.1002/ijc.32151
  131. 131. Jia H, Song L, Cong Q , Wang J, Xu H, Chu Y, et al. The LIM protein AJUBA promotes colorectal cancer cell survival through suppression of JAK1/STAT1/IFIT2 network. Oncogene. 2017;36:2655-2666. DOI: 10.1038/onc.2016.418
  132. 132. Ruan Z, Yang X, Cheng W. OCT4 accelerates tumorigenesis through activating JAK/STAT signaling in ovarian cancer side population cells. Cancer Management and Research. 2018;11:389-399. DOI: 10.2147/CMAR.S180418
  133. 133. Kim S-Y, Kang JW, Song X, Kim BK, Yoo YD, Kwon YT, et al. Role of the IL-6-JAK1-STAT3-Oct-4 pathway in the conversion of non-stem cancer cells into cancer stem-like cells. Cellular Signalling. 2013;25:961-969. DOI: 10.1016/j.cellsig.2013.01.007
  134. 134. Marotta LLC, Almendro V, Marusyk A, Shipitsin M, Schemme J, Walker SR, et al. The JAK2/STAT3 signaling pathway is required for growth of CD44(+)CD24(−) stem cell-like breast cancer cells in human tumors. Journal of Clinical Investigation. 2011;121:2723-2735. DOI: 10.1172/JCI44745
  135. 135. Zhang X, Hu F, Li G, Li G, Yang X, Liu L, et al. Human colorectal cancer-derived mesenchymal stem cells promote colorectal cancer progression through IL-6/JAK2/STAT3 signaling. Cell Death & Disease. 2018;9:25. DOI: 10.1038/s41419-017-0176-3
  136. 136. Zhou B, Damrauer JS, Bailey ST, Hadzic T, Jeong Y, Clark K, et al. Erythropoietin promotes breast tumorigenesis through tumor-initiating cell self-renewal. Journal of Clinical Investigation. 2014;124:553-563. DOI: 10.1172/JCI69804
  137. 137. Dolatabadi S, Jonasson E, Lindén M, Fereydouni B, Bäcksten K, Nilsson M, et al. JAK-STAT signalling controls cancer stem cell properties including chemotherapy resistance in myxoid liposarcoma. International Journal of Cancer. 2019;145:435-499. DOI: 10.1002/ijc.32123
  138. 138. An Open-Label Study to Enable Continued Treatment Access for Subjects Previously Enrolled in Studies of Ruxolitinib. Available from: https://clinicaltrials.gov/ct2/show/NCT02955940
  139. 139. Julian L, Olson MF. Rho-associated coiled-coil containing kinases (ROCK). Small GTPases. 2014;5:e29846. DOI: 10.4161/sgtp.29846
  140. 140. Mulherkar S, Tolias KF. RhoA-ROCK signaling as a therapeutic target in traumatic brain injury. Cell. 2020;9:245. DOI: 10.3390/cells9010245
  141. 141. Srinivasan S, Ashok V, Mohanty S, Das A, Das S, Kumar S, et al. Purwar R: Blockade of Rho-associated protein kinase (ROCK) inhibits the contractility and invasion potential of cancer stem like cells. Oncotarget. 2017;8:21418-21428. DOI: 10.18632/oncotarget.15248
  142. 142. Papadatos-Pastos D, Kumar R, Yap TA, Ruddle R, Decordova S, Halbert PJ, et al. A first-in-human study of the dual ROCK I/II inhibitor, AT13148, in patients with advanced cancers. Journal of Clinical Oncology. 2015;33:2566-2566. DOI: 10.1200/jco.2015.33.15_suppl.2566
  143. 143. Phase I Study of AT13148, a Novel AGC Kinase Inhibitor. Available from: https://clinicaltrials.gov/ct2/show/NCT01585701
  144. 144. Bellomo C, Caja L, Moustakas A. Transforming growth factor β as regulator of cancer stemness and metastasis. British Journal of Cancer. 2016;115:761-769. DOI: 10.1038/bjc.2016.255
  145. 145. Tang B, Yoo N, Vu M, Mamura M, Nam J-S, Ooshima A, et al. Transforming growth factor-beta can suppress tumorigenesis through effects on the putative cancer stem or early progenitor cell and committed progeny in a breast cancer xenograft model. Cancer Research. 2007;67:8643-8652. DOI: 10.1158/0008-5472.CAN-07-0982
  146. 146. Ehata S, Johansson E, Katayama R, Koike S, Watanabe A, Hoshino Y, et al. Transforming growth factor-β decreases the cancer-initiating cell population within diffuse-type gastric carcinoma cells. Oncogene. 2011;30:1693-1705. DOI: 10.1038/onc.2010.546
  147. 147. Bruna A, Greenwood W, Quesne JL, Teschendorff A, Miranda-Saavedra D, Rueda OM, et al. TGFβ induces the formation of tumour-initiating cells in claudinlow breast cancer. Nature Communications. 2012;3:1055. DOI: 10.1038/ncomms2039
  148. 148. Lo P-K, Kanojia D, Liu X, Singh UP, Berger FG, Wang Q , et al. CD49f and CD61 identify Her2/neu-induced mammary tumor-initiating cells that are potentially derived from luminal progenitors and maintained by the integrin-TGFβ signaling. Oncogene. 2012;31:2614-2626. DOI: 10.1038/onc.2011.439
  149. 149. Bhola NE, Balko JM, Dugger TC, Kuba MG, Sánchez V, Sanders M, et al. TGF-β inhibition enhances chemotherapy action against triple-negative breast cancer. Journal of Clinical Investigation. 2013;123:1348-1358. DOI: 10.1172/JCI65416
  150. 150. You H, Ding W, Rountree CB. Epigenetic regulation of cancer stem cell marker CD133 by transforming growth factor-beta. Hepatology. 2010;51:1635-1644. DOI: 10.1002/hep.23544
  151. 151. Mima K, Okabe H, Ishimoto T, Hayashi H, Nakagawa S, Kuroki H, et al. CD44s regulates the TGF-β-mediated mesenchymal phenotype and is associated with poor prognosis in patients with hepatocellular carcinoma. Cancer Research. 2012;72:3414-3423. DOI: 10.1158/0008-5472.CAN-12-0299
  152. 152. Hasegawa T, Yashiro M, Nishii T, Matsuoka J, Fuyuhiro Y, Morisaki T, et al. Cancer-associated fibroblasts might sustain the stemness of scirrhous gastric cancer cells via transforming growth factor-β signaling. International Journal of Cancer. 2014;134:1785-1795. DOI: 10.1002/ijc.28520
  153. 153. Oshimori N, Oristian D, Fuchs E. TGF-β promotes heterogeneity and drug resistance in squamous cell carcinoma. Cell. 2015;160:963-976. DOI: 10.1016/j.cell.2015.01.043
  154. 154. Ikushima H, Todo T, Ino Y, Takahashi M, Miyazawa K, Miyazono K. Autocrine TGF-beta signaling maintains tumorigenicity of glioma-initiating cells through Sry-related HMG-box factors. Cell Stem Cell. 2009;5:504-514. DOI: 10.1016/j.stem.2009.08.018
  155. 155. Peñuelas S, Anido J, Prieto-Sánchez RM, Folch G, Barba I, Cuartas I, et al. TGF-beta increases glioma-initiating cell self-renewal through the induction of LIF in human glioblastoma. Cancer Cell. 2009;15:315-327. DOI: 10.1016/j.ccr.2009.02.011
  156. 156. Naka K, Hoshii T, Muraguchi T, Tadokoro Y, Ooshio T, Kondo Y, et al. TGF-beta-FOXO signalling maintains leukaemia-initiating cells in chronic myeloid leukaemia. Nature. 2010;463:676-680. DOI: 10.1038/nature08734

Written By

Shin Mukai

Submitted: 06 October 2022 Reviewed: 09 November 2022 Published: 24 December 2022