Next Article in Journal
The Emerging Role of Glucagon-like Peptide-1 Receptor Agonists in the Management of Obesity-Related Heart Failure with Preserved Ejection Fraction: Benefits beyond What Scales Can Measure?
Previous Article in Journal
The Impact of Medication Regimen Adjustment Ratio on Adherence and Glycemic Control in Patients with Type 2 Diabetes and Mild Cognitive Impairment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cancer Stem Cells in Oral Squamous Cell Carcinoma: A Narrative Review on Experimental Characteristics and Methodological Challenges

by
Surendra Kumar Acharya
1,†,
Saptarsi Shai
2,†,
Yee Fan Choon
3,
Indrayadi Gunardi
4,
Firstine Kelsi Hartanto
4,
Kathreena Kadir
5,
Ajoy Roychoudhury
6,
Rahmi Amtha
4,* and
Vui King Vincent-Chong
7,*
1
Department of Oral Medicine, Radiology and Surgery, Faculty of Dentistry, Lincoln University College, Petaling Jaya 47301, Selangor, Malaysia
2
Center for Cell and Gene Therapy, Baylor College of Medicine, Texas Children’s Hospital, Houston, TX 77030, USA
3
Department of Oral and Maxillofacial Surgical Sciences, Faculty of Dentistry, MAHSA University, Jenjarom 42610, Selangor, Malaysia
4
Oral Medicine Department, Faculty of Dentistry, Universitas Trisakti, Jakarta 11440, Indonesia
5
Department of Oral and Maxillofacial Clinical Sciences, Faculty of Dentistry, University of Malaya, Kuala Lumpur 50603, Malaysia
6
Department of Oral and Maxillofacial Surgery, All India Institute of Medical Sciences, New Delhi 110029, India
7
Department of Oral Oncology, Roswell Park Comprehensive Cancer Center, Buffalo, NY 14263, USA
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Biomedicines 2024, 12(9), 2111; https://doi.org/10.3390/biomedicines12092111
Submission received: 14 July 2024 / Revised: 11 September 2024 / Accepted: 12 September 2024 / Published: 16 September 2024
(This article belongs to the Special Issue Novel Insights on Cancer Stem Cells)

Abstract

:
Cancer stem cells (CSCs) represent a subpopulation of cancer cells that are believed to initiate and drive cancer progression. In animal models, xenotransplanted CSCs have demonstrated the ability to produce tumors. Since their initial isolation in blood cancers, CSCs have been identified in various solid human cancers, including oral squamous cell carcinoma (OSCC). In addition to their tumorigenic properties, dysregulated stem-cell-related signaling pathways—Wnt family member (Wnt), neurogenic locus notch homolog protein (Notch), and hedgehog—have been shown to endow CSCs with characteristics like self-renewal, phenotypic plasticity, and chemoresistance, contributing to recurrence and treatment failure. Consequently, CSCs have become targets for new therapeutic agents, with some currently in different phases of clinical trials. Notably, small molecule inhibitors of the hedgehog signaling pathway, such as vismodegib and glasdegib, have been approved for the treatment of basal cell carcinoma and acute myeloid leukemia, respectively. Other strategies for eradicating CSCs include natural compounds, nano-drug delivery systems, targeting mitochondria and the CSC microenvironment, autophagy, hyperthermia, and immunotherapy. Despite the extensive documentation of CSCs in OSCC since its first demonstration in head and neck (HN) SCC in 2007, none of these novel pharmacological approaches have yet entered clinical trials for OSCC patients. This narrative review summarizes the in vivo and in vitro evidence of CSCs and CSC-related signaling pathways in OSCC, highlighting their role in promoting chemoresistance and immunotherapy resistance. Additionally, it addresses methodological challenges and discusses future research directions to improve experimental systems and advance CSC studies.

1. Introduction

In 2023, head and neck (HN) cancer, including the oral cavity and pharynx, remained the tenth most common malignancy worldwide among men, with more than 90% of these malignancies diagnosed as squamous cell carcinoma (SCC), known as HNSCC [1]. In South and Southeast Asia and the Western Pacific, cancer of the lip and oral cavity was a leading cause of cancer-related deaths among men. This type of cancer ranked 16th globally in both incidence and mortality, underscoring its significant impact on public health in these regions [2]. Standard conventional treatments for oral squamous cell carcinoma (OSCC) patients include surgery, radiotherapy, and chemotherapy [3,4]. Despite the use of adjuvant therapies, such as platinum-based chemotherapy/5-flurorouracil (5-FU) and cetuximab, locoregional recurrences and metastasis remain major factors for a dismal 5-year survival rate of 50% [5,6,7,8]. In addition, the economic and cost burden of treating head and neck cancer is significantly higher than other cancers [9]. The recent identification of a highly tumorigenic subgroup of cancer cells has received enormous attention [10,11,12,13,14,15]. These cells are known as cancer stem cells (CSCs), and they are postulated to be responsible for the recurrences, metastasis, and poor prognosis of cancers, including OSCC [16,17,18]. Elimination of CSCs theoretically improves the prognosis of the disease [19,20]. Indeed, many small molecules inhibitors targeting CSC-signaling pathways have been developed and are in different phases of clinical trials. Some of them, for instance, vismodegib and glasdegib, targeting smoothened (Smo) protein of the hedgehog signaling pathway, have been approved by the United States (U.S.) Food and Drug Administration (FDA) and launched for the treatment of basal cell carcinoma and acute myeloid leukemia, respectively [21,22,23]. Such advancement in treatment, however, is not observed in patients with OSCC, even though ample studies have demonstrated the existence of CSCs in OSCC. One of the proposed contributing factors for the slow progress in novel treatment in OSCC could be due to the lack of specificity in markers that have been utilized to isolate and characterize CSCs in OSCC [24]. The aim of this narrative review is, therefore, to present experimental evidence on the existence of CSCs in OSCC and discuss potential experimental systems to overcome methodological challenges to better understand CSCs in OSCC.

Search Strategy

Databases, including PubMed, Google Scholar, Ebsco, and Science Direct, were searched from 2007 (first report of CSCs in HNSCC) to 2024 using various combinations of the following keywords: “cancer stem cells in oral squamous cell carcinoma”, “cancer stem cells in head and neck squamous cell carcinoma”, “oral cancer stem cells”, “cancer stem-like cells”, “oral squamous cell carcinoma, OSCC”, “head and neck squamous cell carcinoma, HNSCC”, “cancer stem cells”, “cancer stem cells markers,”, “cancer stem cells pathways,” “chemoresistance and radioresistance in OSCC”, “chemoresistance and radioresistance in HNSCC”, “cancer stem cells targeted therapies”, and “cancer stem cells therapies”. Original experimental studies (both in vitro and in vivo), reviews, editorial letters, book chapters, opinions, and abstracts from the analyses published in English considering stem cell markers in HN-/OSCC were considered and included.

2. Cancer Stem Cells in Oral Squamous Cell Carcinoma

2.1. The CSCs Model

A couple of cancer evolution models have been put forth to elucidate heterogeneity in the cancer cell population: the CSCs model and Nowell’s clonal evolution model. As depicted in Figure 1, in the CSCs model, the cancer cell population is proposed to comprise (1) CSCs and (2) non-CSCs. The CSCs behave like normal stem cells in that they can either divide symmetrically through a self-renewal mechanism to propagate more CSCs or asymmetrically to give rise to more differentiated cells, i.e., non-CSCs. Hence, a malignant cellular hierarchy exists in the CSCs model in which the CSCs that occupy the apex are tumorigenic, as they continuously propagate tumors, in addition to giving rise to a differentiated, non-tumorigenic population, non-CSCs [16,25,26,27,28,29,30,31]. However, no such hierarchy exists in Nowell’s clonal evolution model. In this model, heterogeneity of the cancer cell population is attributed to the clonal expansion of cancer cells after acquiring genetic mutations, resulting in different phenotypic characteristics [32,33].

2.2. Definition of CSCs

As such, the CSCs model theoretically implies that, as long-lived resident cells in the normal tissues, stem cells are the targets for accumulation of genetic aberrations responsible for malignant transformation (Figure 1) [4,25,34]. Indeed, when induced with a carcinogen, 4-nitroquinoline 1-oxide (4-NQO), cell lineage tracing showed that normal stem cells in the basal layer of the tongue epithelium of transgenic mice were the cells of origin of carcinogenesis [35]. However, according to the American Association for Cancer Research (AACR) Workshop on CSCs, CSCs are not necessarily derived from normal stem cells [29], as tumors have been shown to arise from some “progenitor” or differentiated cells [36,37,38,39]. The AACR Workshop on CSCs defines CSCs as cancer cells that produce tumors; hence, xenotransplantation assay is the gold standard to demonstrate the tumorigenicity of CSCs. Cancer cells that are tumorigenic and form tumors when injected into mouse models are CSCs. Therefore, CSCs are defined only by xenotransplantation assay regardless of the cell of origin [29].

2.3. In Vivo Evidence for CSCs

Thus defined, CSCs were first demonstrated in human acute myeloid leukemia (AML). By using cell surface markers, CD34 and CD38, which identified progenitor and pluripotent stem cells in bone marrow, AML-initiating cells that produced leukemia when transplanted in immunodeficient mice were observed to be CD34+CD38 cells. Furthermore, CD34+CD38+ cells failed to produce leukemia in immunodeficient mice. Hence, CD34+CD38 cells fulfilled the definition of CSCs. Additionally, CD34+CD38 cells were able to differentiate into more differentiated leukemic blasts, indicating the retention of the cellular hierarchy in AML proposed by the CSCs theory [11,12]. After this successful demonstration of CSCs in AML, marker-based identification of CSCs is routinely used in CSCs study for solid tumors, as depicted in Figure 2. CSCs from solid tumors are identified by using fluorescence- or magnetic-beads-conjugated putative CSCs markers, such as CD44 or CD133. CSCs are then sorted and examined for in vitro characteristics, such as holoclone and sphere formations and, most importantly, tumor formation in xenotransplantation assay, as it is the gold standard to demonstrate CSCs [29]. With this approach, CSCs were subsequently demonstrated in breast cancer [13], glioblastoma [40], colorectal cancer [41], pancreatic cancer [42], and ovarian cancer [43]. The same methodological approach was used to demonstrate the existence of CSCs in HN- and OSCC. Previously reported putative CSCs markers for other solid tumors (CD44 in breast cancer and CD133 in brain tumor) were used to identify HN-/OSCC CSCs and then isolated and xenotransplanted in mice to observe for tumor formation. Table 1 chronologically summarizes the in vivo evidence of CSCs in HN- and OSCC from the first report in 2007 until 2024 [44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68].

2.4. In Vitro Characteristics of CSCs

2.4.1. Morphology

A monolayer culture of normal keratinocytes generated three different colony morphologies: (1) the holoclones, which are compact round colonies; (2) the paraclones, which are loose irregular colonies; and (3) the meroclones, which exhibit intermediate features; and these clones are derived from stem-, early-, and late-amplifying keratinocytes, respectively [69,70]. In the culture of human and mouse tongue epithelial cells, the results showed that holoclones were the stem cell compartment, as they readily generated holoclones at each passage, confirming a self-renewal property [71,72]. Such clonogenic morphologies were also reported in HN-/OSCC cell lines. Holoclones of HN-/OSCC cell lines also showed a stem cell property, as they were able to generate holoclones upon repassing. Furthermore, when examined for expression of putative HN-/OSCC stem cell markers like CD44 and CD133 (discussed in Section 2.4.3), the holoclones overexpressed these markers, but this expression was weak in the meroclones and paraclones. Using the marker-based sorting method as described in Figure 2, fresh HNSCC tumor samples derived from CD44+ cancer cells also formed holoclones, while CD44−/low cancer cells formed paraclones, indicating the retention of a malignant cellular hierarchy in HNSCC [50,73]. Holoclone formation is a consistent morphological feature of CSCs [74,75,76,77,78,79].

2.4.2. Sphere Formation

In a neural stem cells study, it has been shown that neural stem cells, such as ependymal cells, were able to proliferate as free-floating spheres, which differentiated into neurons, astrocytes, and oligodendrocytes when transferred to an adhesive substrate [80,81,82]. Hence, the ability to proliferate as spheres in non-adherent culture is used to assess stem cell activity [83,84]. Chen and colleagues showed that spheres generated from OSCC cell lines in a non-adhesive culture system were CSCs, as these spheres were shown to express stem cell markers, CD133 and aldehyde dehydrogenase (ALDH1) (discussed in Section 2.4.3), exhibit chemo- and radioresistance, and, more importantly, produce tumors in mice [85]. Moreover, marker-identified CSCs in HNSCC and OSCC have also been reported to consistently form spheres in suspension culture [86,87]. As such, sphere-forming assays are increasingly being used as an important tool to assess the potential of cells with stem cell traits, and, as will be discussed in details in Section 5, three-dimensional (3D) cell culture methods such as this may provide an alternative method for detecting CSCs than the conventional marker-based approach [88].

2.4.3. Markers

The use of markers has been the cornerstone in CSCs research. They are the means of identification, isolation, enrichment, and characterization of CSCs [89,90]. Evidently, a plethora of markers have been used to isolate CSCs from either primary OSCC tumors or cell lines (refer to Table 1). Here, we focus on and delineate the top three markers highlighted in a systemic review [91]: (a) CD44, (b) ALDH, and (c) CD133, which were routinely used to isolate and characterize CSCs in OSCC.
(a)
CD44
CD44 is a non-kinase cell surface glycoprotein that has been used as a marker for CSCs [92]. Based on CD44 expression, CSCs in breast cancer were demonstrated. As few as 100 CD44+CD24−/lowLineage cells formed tumors in mice as compared to the alternate phenotype [13]. Replicating this method of identification in breast cancer, CSCs in HNSCC were first reported by Prince and colleagues in 2007. The group showed that a distinct population of CD44+ and CD44 cancer cells was identifiable from resected primary HNSCC tumors. They demonstrated that CD44+ cells were CSCs, as they formed tumors in mice. In addition, CD44+ cancer cells also differentially expressed a stem-cell-related gene implicated in tumorigenesis, polycomb complex protein BMI-1 (BMI1) [44]. Since then, CD44 has become one of the most common putative markers used to identify CSCs in HN- and OSCC, either singly or in combination with other markers (discussed below) [93] (refer to Table 1). In a study to ascertain the role of CD44 expression in the survival of OSCC patients, Oliveira and colleagues reported that the absence of CD44 (and CD24) was associated with a better overall survival rate [94]. In contrast, expression of CD44 (and ALDH1, a phosphorylated signal transducer and activator of transcription 3 known as p-STAT3) was reported to be associated with a poor prognosis for HNSCC patients [48]. Alternative splicing of CD44 gene generates two families of CD44 isoforms: CD44v and CD44s [92]. To examine if CD44 merely serves as a marker of CSCs or plays a role in conferring CSCs traits to cancer cells, Zhang and colleagues showed that CD44s is strongly expressed in CSCs, and by knocking down CD44, the tumorigenicity of cancer cells was greatly reduced, with a concomitant downregulation of CD44 expression in these cells. Reintroduction of CD44s into cancer cells restored their tumor growth capacity, suggesting that CD44 is not merely a marker but also plays an essential role in promoting CSC traits [95]. Recently, Bai et al. [96] demonstrated that knocking down ephrin receptor A2 (EphA2) attenuated the CSC phenotype in OSCC by inhibiting the expression of CD44 and CD133 via the EphA2/Krüppel-like factor 4 (Klf4) axis in vitro. Additionally, one of the CD44 variants, CD44v3 (and CD24low) phenotype, was significantly correlated with post-operative lymph node metastasis and local recurrence in OSCC [97].
(b)
Aldehyde dehydrogenase (ALDH)
ALDH is an intracellular enzymatic family consisting of 19 isoforms that are localized in the cytoplasm, mitochondria, or nucleus and are responsible for oxidizing aldehydes to carboxylic acids [98]. Several of its isoforms were involved in retinoic acid (RA) cell signaling via RA production by oxidation of all-trans-retinal and 9-cis-retinal, which has been linked to the “stemness” of CSCs [99,100,101,102]. Of the 19 isoforms, expression of ALDH1 was reported to be a putative marker in HNSCC. ALDH1+-only cells possessed unequivocal CSCs traits, such as tumorigenicity, sphere formation, chemoresistance, and invasion, as CD44+CD24ALDH1+ cells. High ALDH1 expression was also associated with a poor prognosis [46]. Furthermore, Clay and colleagues demonstrated that ALDH+ cancer cells isolated from a primary OSCC tumor were ten times more tumorigenic than CD44+-only cells, strongly supporting the use of ALDH as a more discriminatory marker for CSCs in HN-/OSCC [47]. Additionally, knocking down of ALDH3A1 significantly reduced the cancer cells’ viability in cisplatin treatment, indicating ALDH conferred chemoresistance in OSCC [103]. The method of using ALDH to detect CSCs is different from the marker-based approach described in Figure 2. The ALDEFLUOR system is used to detect ALDH activity in CSCs, in which BODIPY-aminoacetaldehyde (BAAA) is used as the ALDH substrate and N,N-diethlyamionobenzalhyde (DEAB) as the negative to detect and isolate CSCs exhibiting high ALDH activity. The ALDEFLUOR system was reportedly able to detect the activities of isoforms ALDH1A1, ALDH1A2, ALDH1A3, ALDH1B1, ALDH2, ALDH3A1, ALDH3A2, ALDH3B1, and ALDH5A1. Whether or not ALDH plays an essential role in conferring CSC traits to cancer cells like CD44 is currently unreported, but ALDH as a therapeutic target in several cancers has entered different phases of clinical trials [104]. Additionally, overexpression of ALDH1 (and p75 neurotrophin receptor, p75NTR) at the tumor invasive front was reported to be an independent predictor of decreased survival and metastasis in OSCC [105].
(c)
CD133
Human CD133 (prominin-1) is a glycosylated protein with five transmembrane domains and two large extracellular loops [106]. Initially characterized as a marker for hematopoietic stem cells [107,108], CD133+ cancer cells isolated from human brain [14,40] and laryngeal [109] tumors were demonstrated to be CSCs. In OSCC, Chiou and colleagues demonstrated CD133 to be a CSC marker. Instead of employing CD133 in the conventional marker-based approach, the group used spheres formed (discussed in Section 2.4.2) in serum-free cultivation to assess the stem cell activity in OSCC cancer cells. Through this approach, the spheres (or oral cancer stem-like cells, OC-SLC, as they were termed) were demonstrated to be enriched with CSCs, as they more readily formed tumors in mice than the parental cells they were derived from. They also showed that these OC-SLC were stained positively for stem cell markers, such as octamer-binding transcription factor 4 (Oct4), nanog homeobox (Nanog), and CD133. Additionally, overexpression of these markers was reported to be associated with poor overall survival in oral cancer patients [45]. In another study, magnetically sorted CD133+ cells showed all CSC traits, such as tumor and sphere formation. These cells were also more resistant to paclitaxel, establishing CD133 as a CSC marker for OSCC [49]. Thereafter, more studies have used CD133 as a CSC marker in OSCC either alone or in combination with other markers (refer to Table 1). Moreover, CD133 may play a role in conferring CSC traits to cancer cells, as, in a study examining the function of CD133, silencing of CD133 attenuated a CSC population and tumorigenicity in OSCC. Treating CD133-knockdown CSC population with cisplatin reduced its proliferation and invasion, suggesting potential CD133-based therapies for OSCC [54].

3. Signaling Mechanisms of CSC

The CSC model views a tumor as an abnormal organ consisting of a hierarchy of cells, including self-renewing stem cells and highly proliferative progenitor cells, which differentiate to form the bulk of the tumor mass [16]. In HNSCC, the primary molecular characteristics of unchecked cell replication are the inactivation of the tumor suppressors p53 and retinoblastoma (RB). Additionally, frequent mutations in the epidermal growth factor receptor mitogen-activated protein kinase kinase 1 (EGFR-MEK), Notch, and phosphatidylinositol-4,5-bisphosphate 3-kinase catalytic subunit alpha/Akt serine/threonine kinase 2/phosphatase and tensin homolog (PI3K/Akt/PTEN) signaling pathways contribute to abnormal mitogenic signaling. Locoregional tumors are often amenable to surgical removal. However, many tumors are diagnosed at a locally advanced stage, which results in a poor prognosis despite multimodal treatments like surgery, radiotherapy, and chemotherapy, especially when combined with a human papilloma virus deoxyribonucleic acid (HPV-DNA)-negative status. Over 50% of patients with locally advanced HNSCC develop metastases or experience relapse, leading to survival rates of less than a year [110,111]. Preclinical and clinical findings indicate that CSCs are able to survive chemo- and radiotherapy and dynamically adapt to changing environmental conditions, e.g., hypoxia or a lack of nutrients [112,113,114]. They achieve this through various molecular pathways that support their survival and proliferation, contributing to therapy resistance. In the next section, we will delve deeper into these molecular pathways and mechanisms that enable CSCs to withstand conventional treatments and drive cancer progression. The precise regulation of stem cell functions is essential for normal biological activity. Several key developmental and signaling pathways are crucial in this regulatory process. These include the Janus-activated kinase/signal transducer and the activator of transcription (JAK/STAT), Wnt family member/beta-catenin (Wnt/β-catenin), PI3K/PTEN pathways. These pathways mediate various stem cell properties, such as self-renewal, cell fate decisions, survival, proliferation, and differentiation. It is not surprising that many of these critical pathways are dysregulated in cancer.

3.1. JAK/STAT Pathway

In extension to these phenomena, the JAK/STAT pathway has been shown to play a crucial role in regulating CSC. Inhibiting JAK/STAT signaling reduces the population of CSC-like cells resistant to the chemotherapy drug doxorubicin, thereby enhancing drug efficacy. However, some CSC-like cells remain resistant to doxorubicin despite JAK–STAT inhibition, suggesting the involvement of additional pathways. Microarray data highlight the importance of JAK/STAT signaling, alongside other pathways, like tumor necrosis factor alpha (TNFα) and KRAS proto-oncogene, GTPase (KRAS) signaling, in doxorubicin resistance [115]. Activation of JAK/STAT signaling in radioresistant colorectal cancer (CRC) tissues, particularly JAK2 overexpression in CRC stem cells, correlates with metastasis. JAK2/STAT3 signaling promotes tumor initiation and radioresistance by inhibiting apoptosis and enhancing clonogenic potential. Mechanistically, STAT3 binds to the cyclin D2 (CCND2) promoter, increasing its transcription and sustaining cancer stem cell growth post-radiotherapy by maintaining cell cycle integrity and reducing DNA damage accumulation. These findings suggest that JAK2/STAT3/CCND2 signaling tends to be a resistance mechanism in CRC, offering potential biomarkers and therapeutic avenues for improving outcomes in CRC patients post-radiotherapy [116].

3.2. Wnt/β-Catenin Pathway

The Wnt pathway plays a pivotal role in cancer stemness, with beta-catenin (β-catenin) activating telomerase reverse transcriptase (TERT) expression, promoting telomere maintenance. The leucine-rich repeat-containing G-protein coupled receptor 5 (Lrg5) marks intestinal stem cells, driving tumor growth when adenomatous polyposis coli (APC) is lost. The Rac family small GTPase 1 (RAC1) activation post-APC loss expands the Lgr5 population via reactive oxygen species/nuclear factor kappa-light-chain-enhancer of activated B cells/Wnt (ROS/NF-κB/Wnt) signaling. Tumor microenvironment (TME) factors, like hepatocyte growth factor (HGF) and periostin, enhance Wnt activity, sustaining tahe cancer stem cell phenotype. Non-coding RNAs like miR-146a stabilize β-catenin, while long non-coding transcription factor 7 (lncTCF7) activates TCF7, promoting self-renewal in liver CSCs. Intestinal stemness signatures correlate with a poor prognosis in colorectal cancer, reflecting a tumor differentiation status rather than Wnt-driven stem cell numbers [117]. Additionally, by inhibiting β-catenin signaling, XAV939 effectively curbed CSC progression in HNSCC and CRC cells, thereby mitigating CSC-induced chemical resistance [118].

3.3. PI3K/Akt/mTOR Pathway

The PI3K/Akt/mammalian target of rapamycin (mTOR) pathway plays a crucial role in maintaining stemness by inhibiting the MEK/extracellular signal-regulated kinases (ERK) signaling pathway. This pathway’s involvement extends beyond specific cancer types, also influencing the stemness of mesenchymal stem cells (MSCs). Stem cell markers like CD44, CD133, CD24, epithelial cell adhesion molecule (EpCAM), Lrg5, and ALDH1 are regulated by S6 kinase 1 (S6K1), a downstream component of PI3K/Akt/mTOR. Additionally, the inactivation of pathway inhibitors, such as PTEN, amplifies PI3K/Akt/mTOR signaling, leading to an expansion of the stem cell population, as observed in adult acute leukemia (ALL) [119]. In CSCs, ALDH activity, a primary property, is regulated by PI3K signaling and the sex-determining region Y-box 2 (Sox2). Sox2 expression, induced by PI3K signaling, significantly increases ALDH1A1 and the ALDH+ cell population by directly interacting with the ALDH1A1 promoter. Additionally, squamous tumor cells exhibit dynamic reactivation of PI3K/mTOR via EGFR/AXL receptor tyrosine kinase (AXL)/protein kinase C (PKC) following pharmacological inhibition of PI3Kα. This reactivation enhances Sox2 translation, subsequently upregulating genes associated with CSC stemness characteristics [120,121]. Moreover, Sox2 overexpression was found to restore transcriptional and translational levels of ALDH1A1, ALDH3A1, and CD44, which were suppressed after knocking down of ubiquitin-specific protease 14 (USP14), a well-established carcinogenic gene in HNSCC. The results suggested a master regulator role of Sox2 in the CSCs’ state [122]. Additionally, a study has shown PIK3CA overexpression enhances CSCs’ population in both murine and human HNSCC. However, CSC maintenance in PIK3CA-overexpressing HNSCC becomes independent of the PI3K pathway. Compensatory mechanisms, including ephrin receptors (Ephs), tropomyosin receptor kinases (TRKs), and proto-oncogene c-KIT (c-Kit) signaling activation, sustain CSCs’ population. Co-targeting these pathways may effectively eliminate PI3K-independent CSCs, offering potential for novel anti-CSC therapeutic strategies in HNSCC, particularly for patients with PIK3CA amplification [123]. Increasing evidence suggests that hyperactive or abnormal signaling within these pathways contributes to the survival of CSCs. CSCs are a relatively rare population of cancer cells with the ability to self-renew, differentiate, and generate serially transplantable heterogeneous tumors across various cancer types [124]. CSCs are located at the invasive fronts of HNSCC, particularly near blood vessels (perivascular niche). Signaling events initiated by endothelial cells are crucial for the survival and self-renewal of these stem cells.
Markers such as ALDH, CD133, and CD44 have been effectively used to identify highly tumorigenic CSCs in HNSCC. Recent evidence indicates that CSCs are highly resistant to conventional therapies and act as the primary drivers of local recurrence and metastatic spread [125]. The CSC model highlights a tumor’s hierarchical structure, with self-renewing stem cells driving tumor growth and therapy resistance. Dysregulated signaling pathways like PI3K/Akt/mTOR, JAK/STAT, and Wnt play pivotal roles in CSC maintenance and therapy response. Inhibition of these pathways, such as with XAV939-targeting β-catenin in HNSCC and colon cancer, presents promising anti-CSC therapeutic strategies. Additionally, compensatory mechanisms like Ephs, TRKs, and c-Kit signaling sustain the CSC population in PIK3CA-overexpressing HNSCC, suggesting potential co-targeting opportunities. CSCs, located at invasive fronts and perivascular niches, are highly resistant to therapy, driving recurrence and metastasis.

3.4. Epithelial Mesenchymal Transition (EMT)

EMT is a physiologic cellular program that plays an important role during embryogenesis. In EMT, epithelial cells exhibit (1) morphological changes, from a cobblestone-like appearance to spindle-shaped cells; (2) differentiation marker changes, from cytokeratin intermediate to vimentin filaments; and (3) functional changes, from stationary to motile cells that can invade through the extracellular matrix [126]. Distinct cell populations with marked CSCs and EMT characteristics had been reported in OSCC. In primary OSCC tumors and cell lines, Biddle and colleagues demonstrated two distinct CSC populations: (1) CD44highEpCAMhigh and (2) CD44highEpCAMlow. Even though both populations were CSCs, as they formed tumors in NOD scid gamma severe combined immunodeficiency (NOD-SCID) mice, the CD44highEpCAMhigh population was designated as non-EMT-CSCs, as this population exhibited more epithelial characteristics like cobblestone colony (holoclones) formation, more proliferative, and a lower rate of migration, with expression of epithelial-specific genes like E-cadherin and Keratin 15, while the CD44highEpCAMlow was designated as EMT-CSCs, as this population showed an elongated, fibroblast-like feature in a monolayer culture and a low proliferative rate but higher migratory ability, with overexpression of EMT markers like vimentin, Twist, Snail, and AXL. Furthermore, the study demonstrated that the cells of one population were able to generate cells of the other through EMT and mesenchymal epithelial transition (MET) processes and further identified a population with a MET ability restricted to another subpopulation, the CD44highEpCAMlow/−ALDH+ cells [50]. In a separate study, the group reported a new CSC marker, CD24, which identified a drug-resistant EMT CSC subpopulation, CD44highEpCAMlow/−CD24+ cells [56]. Other studies have also consistently documented such phenotypic plasticity in CSCs [59,65,127], raising questions on CSC heterogeneity and limitations in the methodologies for studying CSCs [24] (which will be discussed further in Section 4).

4. Cancer Stem Cells in Therapy

Resistance induced by the cancer cell remains a significant impediment to achieving effective therapeutic outcomes in HNSCC [128]. Despite the long-standing use of standard of care agents, such as cisplatin, their efficacy is limited, and they often lead to the development of chemoresistance [128]. Therefore, elucidating the molecular pathways that drive chemoresistance and targeting these pathways to re-sensitize tumors may be critical for advancing treatment strategies in HNSCC. While chemotherapeutic agents like cisplatin and 5-FU have been clinically approved for HNSCC, their therapeutic efficacy in metastatic and recurrent malignancy is markedly diminished due to acquired resistance [128]. Multiple factors contribute to this chemoresistance, with CSCs being a pivotal element. CSCs, characterized by their inherent self-renewal and plasticity, foster a more aggressive phenotype that is resistant to standard of care regimens. This CSCs subpopulation plays a critical role in immunosuppression, therapeutic resistance, metastasis, and invasion, ultimately leading to poor patient survival [34]. Addressing CSC-induced resistance through a deeper understanding of the underlying mechanisms and by targeting key pathways involved in chemoresistance holds promise for enhancing the efficacy of current therapeutic modalities and improving clinical outcomes in HNSCC [129]. Thus, Table 2 presents the current evidence on how CSCs and their downstream signaling are associated with chemoresistance, while also illustrating potential novel drug candidates to overcome CSC-induced chemoresistance in HNSCC preclinical models. Overall, approximately more than 80% of studies have exposed CSCs to a cisplatin regimen and characterized their resistance ability towards this chemotherapy due to the over-expression of CSC markers, such as β-catenin, polycomb complex protein BMI-1 (BMI1), ATP-binding cassette super-family G member 2 (ABCG2), CD44, ALDH1, Snail, CD10, Nanog, sex-determining region Y-box 2 (Sox2), octamer-binding transcription factor 4 (Oct4), CD133, neurogenic locus notch homolog protein 1 (Notch1), CD24, MYC oncogene (c-Myc), Krüppel-like factor 4 (Klf4), cyclin D1 (CCND1), forkhead box protein M1 (FOXM1), AlkB homolog 5, RNA demethylase (AKLBH5), CD271, CD326, and DEAD-box helicase family member (DDX3).
In 2013, Masui et al. [130] demonstrated that Snail-expressing HNSCC cell lines induced a stem-cell-like phenotype by increasing the expression of CD44 and ALDH1 and better resisted a cisplatin regimen as compared to non-Snail-expressing cell lines. Similarly, in 2015, Ota et al. [131] demonstrated a similar pattern observation that Snail expression enhances the chemoresistance to cisplatin in HNSCC cell lines. Khammanivong et al. [132] demonstrated that sphere cells with enrichment of CD44 and BMI1 CSC marker were highly resistant to cisplatin compared to monolayer cultures. Byun et al. [133] demonstrated that CD44+ HNSCC cells were relatively resistant to cisplatin and radiation, and knockdown of hypoxia-inducible factor (HIF-1) or Notch1 enhanced the chemosensitivity to cisplatin or radiation treatment as compared to the control. Lee et al. [134] demonstrated that Notch intracellular domain (NICD) activation enhanced sphere formation and also increased the expression of CD44, Oct4, and Sox2 markers. Knockdown of Notch1 enhanced the chemosensitivity to cisplatin treatment in HNSCC preclinical models. Xie et al. [135] reported that Sox8 enhances chemoresistance and upregulates CSC markers, while Sox8 knockdown increases chemosensitivity to cisplatin both in vitro and in vivo by reducing CD24+CD44+ CSC markers and tumor volume.
In Table 2, it can be seen that a few studies have utilized hyaluronic acid (HA, anti-CD44), inhibitors of apoptosis proteins (IAP), secreted frizzled-related protein 4 (sFRP4), curcumin, valproic acid (VPA), NCT-501, heat shock protein 90 (Hsp90) inhibitor, SB203580, XAV-939, SVC112, ketorolac, nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) inhibitor, and TVB-3166 to overcome CSCs chemoresistance. Bourguignon et al. [136] and Ref. [137] demonstrated both HA treatment of tumor cells or CD44v3highALDH1 high cells promote sphere formation and confer chemoresistance to cisplatin treatment. To overcome the cisplatin resistant, a combination of SM164 [136] or anti-CD44 with cisplatin enhanced the cell-killing activity by targeting CSCs in HNSCC. Similarly, another approach from Warrier et al.’s [138] study demonstrated that a combination of sFRP4 and cisplatin exhibited the lowest cell viability and highest apoptotic cell death as compared to monotherapy of sFRP4, cisplatin, and an untreated control in a cisplatin-resistance preclinical model. Apart from that, Basak et al. [139] demonstrated that CD44hi cells exhibit properties of CSC that confer cisplatin resistance. A combination of curcumin or curcumin-difluorinated (CDF), a synthetic analog of curcumin, with cisplatin increased the cell number reduction in CD44high CSCs’ cell population that confer cisplatin resistance. Using drug repurposing approaches, Lee et al. [140] demonstrated that a combination of VPA, a drug used for complex partial seizures, with cisplatin overcomes cisplatin resistance driven by CSCs in an HNSCC preclinical model. Kulsum et al. [141] focused on targeting ALDH1A1 in CSCs that promote chemoresistance to cisplatin treatment using NCT-501. Post-treatment with this inhibitor overcame chemoresistance by reducing the spheroid formation in vitro and reduced the tumor volume in vivo as compared to the untreated control. A similar trend was also observed in Subramanian et al.’s [142] study, which employed the Hsp90 inhibitor to reduce both CD44 and ALDH subpopulations in vitro and reduced tumor volume in vivo, which were resistant to cisplatin treatment. Few studies have been conducted by the Banerjee lab at Kalinga Institute of Industrial Technology using various inhibitors to target CSC markers in HNSCC to enhance chemosensitivity in preclinical models resistant to cisplatin [118,143,144,145]. For instance, Roy et al. [143,144] employed a mitogen-activated protein kinase (p38) inhibitor to target CSC markers in HNSCC, resulting in a decrease in the half maximal inhibitory concentration (IC50) value of cisplatin compared to untreated controls resistant to cisplatin treatment. They also demonstrated that the combination of XAV-939, a tankyrase inhibitor targeting β-catenin, with cisplatin synergistically attenuated chemoresistance by increasing DNA damage in vitro [118]. In 2020, Roy et al. [145] demonstrated that tetrahydroisoquinoline (THIQ)-mediated inhibition of CD44 sensitized HNSCC cells to cisplatin treatment. Similarly, Keysar et al. [146] demonstrated that SVC112, a translation elongation inhibitor, enhanced tumor reduction in vivo when combined with cisplatin and radiation regimens. Additionally, Shriwas et al. [147] showed that the combination of ketorolac, a bioactive inhibitor of DDX, with cisplatin enhanced the chemosensitivity of cisplatin-resistant cells by reducing the IC50 of cisplatin in vitro and inhibiting tumor growth in vivo. Recently, de Castro et al. [68] and Aquino et al. [66] demonstrated that the combination of an NF-κB inhibitor and a selective fatty acid synthase (FASN) inhibitor, respectively, enhanced the chemosensitivity of cisplatin in an in vitro model resistant to cisplatin treatment in HNSCC.
In Table 2, we also identify two novel drugs, namely, casein kinase 2 (CK2) inhibitor and MEDI0641, which were found to suppress CSC subpopulations in HNSCC. Lu et al. [148] demonstrated the CK2 inhibitor suppressed CSC markers expression, such as Nanog, Oct4, and Sox2 mRNA and protein expression, as well as the CSC side population phenotype. Similarly, Kerk et al. [149] demonstrated the ability of using an antibody-drug conjugate (ADC) to target the 5T4 oncofetal antigen known as MEDI0641 and reduce CD44 and ALDH cell populations, as well as reduce tumor growth and prevent tumor recurrence in an in vivo model of HNSCC.
5-FU is a commonly used chemotherapeutic agent for HNSCC patients, often employed alongside cisplatin or as a standalone therapy. Tabor et al. [150] demonstrated that the side population, a group of cells exhibiting stem-cell-like characteristics, showed resistance to 5-FU treatment compared to the non-side population cells. Other studies have shown that HNSCC cell lines exhibiting a CSC phenotype confer chemoresistance to monotherapy with cisplatin, 5-FU, or radiotherapy. For instance, Elkashty et al. [61] and Fukusumi et al. [151] demonstrated that the CD44+CD271+ and CD10 populations, respectively, exhibited a CSC phenotype and conferred resistance to monotherapy with cisplatin, 5-FU, or radiotherapy in vitro.
Table 2. Summary of findings on chemoresistance mechanisms and potential therapeutic strategies targeting CSCs in HNSCC.
Table 2. Summary of findings on chemoresistance mechanisms and potential therapeutic strategies targeting CSCs in HNSCC.
ReferencesMarker(s)Cell line(s)Drug(s)Findings
[150]β-catenin
BMI1
ABCG2
UM-SCC-10B5-Fluorouracil (5-FU)
(0–100 μM)
The side population cells were characterized with high expressions of BMI-1 and ABCG2 and a low expression of β-catenin and exhibited greater resistance to the chemotherapeutic agent 5-FU. BMI-1 suppresses p16Ink4a expression and enhances self-renewal abilities, whereas ABCG2 confers drug resistance through the efflux of chemotherapeutic drugs. Additionally, the decrease in β-catenin levels inhibits its translocation to the nucleus, thereby shielding cells from the cytotoxic impact of 5-FU.
[136]CD44v3
ALDH1
HSC-3HA (50 μg/mL) (or anti-CD44 antibody plus HA (50 μg/mL) or no HA)
Cisplatin (4 × 10−9 to 1.75 × 10−5 M)
IAP inhibitor (SM-164)
The interaction between HA and CD44v3 with Oct4-Sox2-Nanog signaling stimulates the production of miR-302, which subsequently results in the downregulation of AOF1/AOF2/DNMT1. This process enhances the survival and resistance to chemotherapy in tumor cell populations characterized by high levels of CD44v3 and ALDH1. Furthermore, the concurrent use of SM164 and cisplatin leads to increased efficacy in killing cancer cells by specifically targeting cancer stem cells in HNSCC.
[130]Snail
CD44
ALDH1
SAS
HSC-4
Cisplatin
(1–10 μM)
Induction of Snail increases expression of CD44 and ALDH1. The induction of EMT via Snail suppresses E-cadherin expression that leads HNSCC cells to adopt CSC-like phenotype and chemoresistance to cisplatin. Focusing on the strategic targeting of EMT-regulating Snail presents a potential benefit in cancer therapy, as suppressing EMT could effectively impede cancer progression and spread, while also preventing the development of cancer stem cells.
[151]CD10FaDu
Detroit562
Cisplatin (0.1–5  μM)
5-FU (0.5–50  μM)
A CD10+ positive subpopulation of HNSCC cells exhibited CSC-related properties, such as chemo- and radioresistance, a self-renewal capacity, and enhanced tumorigenicity. CD10 has been shown to be mechanistically associated with the increased expression of OCT3/4, a pivotal regulator of both stem cell self-renewal and oncogenic pathways. Silencing CD10 led to a reduction in OCT3/4 levels, implying that CD10 promotes cancer stem cell characteristics in HNSCC by enhancing OCT3/4 expression. These results indicate that targeting CD10 could potentially be a beneficial strategy for addressing the challenges of chemoresistance and radioresistance in treatment-resistant HNSCC.
[138]CD44+
ALDH+
Hep2
KB
sFRP4 (Wnt antagonist, 250 pg/mL)
Cisplatin (10 Mm)
A decrease in the expression of ABCG2 and ABCC4 was observed in sFRP4-drug-treated spheroids.
A combination of sFRP4 and cisplatin exhibits the lowest cell viability and highest apoptotic cell death as compared to monotherapy of sFRP4, cisplatin, and an untreated control.
[148]Nanog
Sox2
Oct4
UM-SCC-1 UM-SCC-46CX-4945 (CK2 inhibitor, 0.5, 1 and 5 μM)Inhibition of CK2 and activation of TAp73 repress the mRNA and protein expression of Nanog, Oct4, and Sox2, along with reducing the CSC side population phenotype. This process promotes tumor protein P73 (TAp73) expression, leading to growth arrest and the overexpression of pro-apoptotic genes like cyclin-dependent kinase inhibitor 1 (p21WAF1) and the p53 upregulated modulator of apoptosis (PUMA).
[132]CD44
BMI1
TR146
SCC-58
UMSCC-17B
Cisplatin (0 to 10 μM)Spheroids grown on Matrigel exhibited elevated expression of CD44 and BMI1, with a higher proportion of cells in the CD44high population, both of which are markers for HNSCC CSCs. These spheroids demonstrated significantly higher resistance to cisplatin compared to monolayer cultures, with the ALDH activity remaining either unchanged or slightly reduced in the spheroid cells. The increased expression of SMAD-specific E3 ubiquitin protein ligase 1 (SMURF1) in the spheroid cultures results in the suppression of traditional BMP signaling, specifically, the decreased activation of pSMAD1/5/8, which ultimately enhances resistance to chemotherapy. To combat this, addressing SMURF1 levels and reinstating BMP signaling could present a novel treatment strategy to stimulate differentiation and decrease the population of cancer stem cells, ultimately diminishing drug resistance and the likelihood of disease recurrence.
[139]CD44
CD133
ALDH
CCL-23
UM-SCC-1
Curcumin (25 μM)
Cisplatin (10–20 μM)
In vitro models resistant to cisplatin demonstrated heightened levels of IL-1β, IL-6, IL-8, IL-10, basic fibroblast growth factor (bFGF), vascular endothelial growth factor (VEGF), matrix metalloproteinase-1 (MMP-1), and matrix metalloproteinase-9 (MMP-9), which were associated with the activation of the AKT pathway (GSK-3β, p70S6K), thereby leading to the development of chemoresistance. The in vitro and in vivo inhibitory effect on CD44 clearly demonstrates that curcumin targets CSCs and could reduce CD44hi CSC accumulation with cisplatin treatment.
[140]Sox2
Oct4
K3
K5
Cisplatin (5 μM)
VPA (400 μM)
VPA disrupted the self-renewal of HNSCC CSCs by downregulating the expression of stem cell markers, such as Oct4, Sox2, and CD44. Combining VPA with cisplatin further diminished HNSCC CSC chemoresistance, likely by suppressing ABCC2 and ABCC6 expression and enhancing caspase-mediated apoptosis.
[131]SnailSAS
HSC-4
Cisplatin (1 μM)Snail overexpression induced CSC-like properties, and the expression of phospho-EGFR was enhanced in the Snail-transfected cells compared to the controls.
[134]Notch1SNU 1041Cisplatin (0–50 μM)Activation of Notch1 signaling is correlated with increased cell proliferation in HNSCC through its influence on cell cycle progression, while constitutive activation of NICD promotes CSC traits in differentiated HNSCC cells.
Reducing Notch1 signaling attenuates CSC traits in HNSCC-derived CSCs and enhances their chemosensitivity to cisplatin by suppressing the expression of ABCC2 and ABCG2 transporter genes.
[137]CD44v3high
ALDH1high
HSC-3Anti-CD44 antibody/HA (4 × 10−9 to 1.75 × 10−5 M)
Cisplatin (4 × 10−9 to 1.75 × 10−5 M)
HA treatment promotes sphere formation, self-renewal/growth, and differentiation, such as tumor cell invasion, in highly tumorigenic CD44v3highALDH1high tumor cells, indicating HA signaling’s role in regulating CSC properties.
[141]CD44
ALDH1A1
Cal27
Hep-2
NCT-501 (ALDH1A1 inhibitor, 0–80 nM)
Cisplatin (20 μM)
CSC-specific markers (CD44, ALDH1A1, CD133, Notch1), a drug efflux marker (ABCG2), and stem cell maintenance markers (Sox2, Nanog, Oct4) showed differential upregulation in resistant cell lines, correlating with increased spheroid formation and migration.
Silencing ALDH1A1 in cisplatin-resistant cell lines led to decreased multidrug resistance (MDR) gene expression, downregulation of CD44 and CD133, loss of self-renewal and migratory properties, and increased cisplatin sensitivity.
Post-treatment with NCT-501 reduced the tumor volume significantly as compared to the untreated control in an in vivo model.
[149]ALDHhigh
CD44high
UM-SCC-11B
UM-SCC-22B
(MEDI0641, Antibody-drug conjugate targets 5T4 oncofetal antigen, 0–1 μg/mL)A single dose of MEDI0641 led to long-term tumor regression in patient-derived xenograft (PDX) models of HNSCC, even after tumors were already established. MEDI0641 effectively targets and reduces the fraction of CSCs, which express high levels of 5T4, in both HNSCC cell lines in vitro and in vivo.
[142]CD44
ALDH
UMSCC- 22B
MDA1986
Hsp90 inhibitor
KU711 (20 to 40 μM) KU757 (1.0 to 2.5 μM)
Cisplatin (2μM)
Both Hsp90 inhibitors reduced sphere formation in vitro by decreasing the CD44 and ALDH populations, which were resistant to cisplatin treatment. Additionally, these inhibitors led to a reduction in tumor volume in vivo in cisplatin-resistant models.
[135]Sox2
Oct4
BMI1
ABCG2
CD44
CD24
Cal27
SCC9
Cisplatin (0−7 M to 10−5 M)Ectopic expression of Sox8 promotes chemoresistance, CSC properties, and EMT features in chemoresistant HNSCC cells. Knockdown of Sox8 inhibited chemoresistance, CSC properties, and EMT features. Post-treatment with cisplatin reduced the tumor volume in xenograft models with Sox8-knockdown cell lines compared to cisplatin alone or untreated controls.
[143]c-Myc
CD44
Oct4
Klf4
UPCI-SCC-131
Cal27
SB 203580 (p38 inhibitor, 10 μM)
Cisplatin (1–20 μM)
p38-inhibited cells exhibit a lower IC50 of cisplatin and reduced expression of CSC markers and increased apoptosis as compared to the untreated control.
[118]c-Myc
CD44
Oct4
Klf4
UPCI-SCC-131
Cal27
Cisplatin (1–20 μM) XAV-939 (Tankyrase inhibitor, 1–4 μM)The combination of cisplatin and XAV-939 significantly reduced the expression of the CSC marker Oct4, the nucleotide excision repair gene excision repair cross-complementing rodent repair deficiency, complementation group (ERCC1), and β-catenin. This combined treatment also heightened the genotoxic effects, as indicated by increased DNA damage and shortened telomere length. Together, cisplatin and XAV-939 contribute to genomic instability in HNSCC cells by amplifying DNA damage.
[146]c-Myc
CCND1
CD44
Sox2
036C
067C
013C
049C
Fadu
Detroit562
SVC112 (0–1000 nM)
Cisplatin (1 mg/kg once weekly)
SVC112, an elongation inhibitor, significantly suppressed the expression of c-Myc, CCND1, and Sox2 in HNSCC cells and CSCs across in vitro, ex vivo, and in vivo models. When combined with cisplatin and radiation therapy, SVC112 further reduced tumor growth in vivo.
[147]Sox2
Oct4
FOXM1
Nanog
ALKBH5
H357
SCC-9
SCC-4
Cisplatin (0–10 μM/2.5 mg/kg)
Ketorolac (0–2.5 μM/30 mg/kg)
A combination of cisplatin with ketorolac enhanced the chemosensitivity of cisplatin in cisplatin-resistant cell lines in in vitro and in vivo models by targeting the DDX3 and CSC subpopulations.
[145]CD44UPCI-SCC-131
Cal27
1,2,3,4 tetrahydroisoquinoline (THIQ) (1–5 mM)
Cisplatin (1–10 μM)
The impact of CD44 inhibition on the expression of proteins involved in the Wnt/β-catenin signaling pathway, as well as on other CSC markers like Oct4 and Klf4, was examined. Inhibiting CD44 led to reduced levels of β-catenin and phosphorylation of GSK3β (Ser 9), alongside decreased expression of other CSC markers. Treatment with 1 mM THIQ resulted in a significant reduction in CD44 expression in both cisplatin-resistant HNSCC cells and their parental counterparts. Additionally, a flow cytometric analysis revealed that 1 mM THIQ treatment decreased the percentage of CD44-positive cells.
[61]CD44+/CD271
CD44+/CD271+
SCC38 SCC12Cisplatin (2 µg/mL)
5-FU (32 µg/mL)
CD44+/CD271+ cells exhibited superior colony-forming and sphere-forming abilities compared to CD271 cells and their unsorted parental counterparts, and they also formed larger spheres. This subpopulation of CD44+CD271+ cells demonstrated the highest resistance to cisplatin, 5-FU, and radiation regimens when compared to CD44+CD271 cells and parental cell line models. The potential explanation for this phenomenon could stem from the ALDH1A1-mediated stimulation of the ATP-binding cassette subfamily B member 1 (ABCB1) drug-efflux pump, as well as survival proteins, such as AKT and BCL2. Focusing on the CD271 cell subset results in the ability to overcome chemotherapy resistance in HNSCC.
[144]c-Myc
CD44
Oct4
Klf4
SCC-131
Cal 27
SB 203580 (p38 inhibitor, 10 μM)
Cisplatin (1–10 μM)
Treatment with escalating doses of cisplatin resulted in an increased expression of CSC markers. In contrast, inhibition of p38 in cisplatin-resistant cells led to a reduced expression of CSC markers, as the cisplatin concentration increased. Additionally, p38 inhibition reduced both the size and sphere-forming ability of cisplatin-resistant cells compared to cells without p38 inhibition, suggesting that p38 plays a role in maintaining the CSC phenotype in HNSCC cells.
[133]CD44+SCC-15
SCC-25
Cisplatin (5 μM)CD44+ HNSCC cells exhibited relative resistance to cisplatin and radiation. However, knockdown of HIF-1α or Notch1 increased chemosensitivity to cisplatin and radiation compared to the control cells.
[68]ALDHhigh
CD44high
Cal 27
SCC9
Cisplatin (5 Μm)
NF-κB inhibitor
CBL0137 (0–10 μM)
Emetine (0–10 μM)
In CisR cell lines, overexpression and phosphorylation of NF-κB were observed. Treatment with NF-κB inhibitors reduced sphere formation and decreased the expression of NF-κB, chemokine (C-X-C motif) ligand 8 (CXCL8), and TNF-alpha mRNA. Combining cisplatin with NF-κB inhibitors further reduced the ALDHhigh/CD44high populations and the number of colonies formed compared to both the untreated control and cisplatin monotherapy.
[66]CD44
CD326
LN-1ATVB-3166 (0–150 μM)
Cisplatin (0–100 μM)
Three subpopulations known as CD44Low/CD326 (CSC-M1), CD44Low/CD326High (CSC-E), and CD44High/CD326 (CSC-M2) were isolated from LN-1A. Post-treatment with cisplatin showed that CSC-M1 has a lower IC50 compared to LN-1A, CSC-E, and CSC-M2, whereas both CSC-E and CSC-M2 have higher IC50 values compared to LN-1A, with CSC-M2 having the highest IC50. Exposure to TVB-3166 reduced the IC50 value for all three subpopulations compared to LN-1A.
The U.S. FDA has recently approved checkpoint inhibitors like pembrolizumab and nivolumab for recurrent metastatic HNSCC [152]. However, their response rate has been modest, at around 20% [153]. This indicates that intrinsic properties of the TME or specific cells within it contribute to resistance, driving recurrent and secondary tumor growth. A key factor in this resistance may be related to immune evasion mechanisms, where tumors downregulate peptide major histocompatibility complex molecules and employ other strategies to avoid immune detection [154]. CSCs are likely to play a significant role in this process, making it crucial to understand their resistance not only to standard therapies like chemotherapy and radiation but also to novel treatments, such as immune checkpoint inhibitors [155]. Gaining this knowledge can provide valuable insights into how we might target CSCs, re-sensitize immune cells, and ultimately enhance the therapeutic efficacy of checkpoint inhibitors, not only for HNSCC but also for other types of cancers.
The current therapeutic achievements of HNSCC have been very challenging due to the CSC hypothesis. A major reason behind this challenge is the role of CSCs to metastasize into distant organs and be resistant to therapies. One major component that drives a successful metastasis is the ability to invade the immune system [156]. The development and introduction of immunotherapy in HNSCC holds very promising as a supplement to the traditional standard of care, like surgery, chemotherapy, and radiation; however, the effect of this therapy has not reached its peak in the realm of HNSCC [157]. A major reason behind this setback is the ability of CSCs to recognize immune cells and mount an immunosuppressive environment within the TME and during its transit of metastasis [158]. Therefore, it is crucial to understand how CSCs drive the potency to suppress the immune system and how this interaction can be leveraged to gain further insights in developing novel therapies that can help in achieving better therapeutic success for patients diagnosed with HNSCC. It has been shown that the expression of CSC markers in patient tumor tissue is dependent on the number of tumor-infiltrating immune cells. The most important interaction between the T cell receptor on cytotoxic T cell lymphocytes (CTL) with the major histocompatibility complex (MHC) class I antigen is the one that drives the tumor cell recognition and cytotoxic phenomenon by T cells; however, it is suggested that the steps of this process of antigen processing and presenting are impaired in CSCs [159]. In accordance with the previous statement, recent studies have shown MHC downregulation in melanoma spheroid cells, leading to inhibition of the allogenic immune response of T cells. Similarly, defects in the MHC class I directed antigen presentation to CSC have no cytotoxic effect. Moreover, it has been studied that this phenomenon is not just restricted to CSCs but is also present in normal hematopoietic cells and epithelial cells that express Lgr5+, which is a stem cell marker in intestinal crypts and mammary glands and fails to induce antigen presentation [160,161,162]. It was also demonstrated that CSCs have been involved in the downregulation of the MHC molecule in HNSCC. HNSCC cells with CD44+ exhibit stem cell properties that have shown the downregulation of many common human leukocyte antigen (HLA) classes, like HLA-A2, HLA class II [163]. Furthermore, CD44+ HNSCC has been known to induce the immunosuppressive cells, like regulatory T cell (Treg) and myeloid-derived suppressor cells (MDSC), which have high immunosuppressive cytokines like interleukin-8 (IL-8) and transforming growth factor beta [158,163]. The expression of the immune checkpoint is well known by now to induce immunosuppression; however, its role in CSCs has not yet been conclusive enough. However, a recent study suggests that there can be other possible mechanisms that help CSCs to escape immune surveillance. CD276, which is a member of the B7 family of immune checkpoint proteins and is highly expressed in tumor cells, has also been seen to be expressed in CSCs in mouse HNSCC. When treating with anti-CD276 antibodies, it was observed that CSCs were eliminated by CD8+ T cells, which also inhibited tumor growth and nodal metastasis. Furthermore, RNA sequencing data showed the blockade of CD276 remodels SCC heterogeneity and reduced the EMT transition that is a crucial step in inducing metastasis [164,165].
HNSCC is characterized by its cold TME, which means the lack of efficient and sufficient tumor-infiltrating lymphocytes, which fosters a favorable microenvironment for CSCs to evade and metastasize. A recent study delved deeper into the aspect in the context of HNSCC, where they have characterized the presence of CD34+ cells producing a high level of IL-6, which participates in the development of TME and triggers its own malignant progression [166]. Other immune checkpoints that have played significant roles are CD276, leukocyte immunoglobulin-like receptor subfamily B member 2 (LILRB2), and CD47, which were upregulated in CSCs and led to a weakened or damaged host anti-tumor response. The presence of naïve CSCs appears to be more malignant and results in a worse prognosis [167]. The therapeutic landscape for HNSCC faces a significant challenge due to CSCs. CSCs can evade immune surveillance through different mechanisms, including downregulation of MHC and induction of immunosuppressive cells and cytokines. Recent studies suggest that targeting immune checkpoints like CD276 in CSCs may enhance anti-tumor responses and inhibit metastasis. Additionally, the cold TME in HNSCC further facilitates CSC evasion and metastasis, highlighting the importance of understanding and targeting these interactions for improved therapeutic outcomes.

5. Challenges and Future Directions

As evidenced in Table 1, the lack of specific CSCs markers is a challenge in the study of CSCs. The choice of markers is usually based on previous reports that have successfully demonstrated CSCs with a particular marker. For example, CD44 was first used as a marker to study CSCs in HNSCC because of its success in identifying breast CSCs [44]. Recently, by using previously reported putative CSCs markers, including CD44, CD24, and ALDH, Vipparthi and colleagues demonstrated the presence of four CSC subpopulations (CD44+CD24lowALDHhigh, CD44+CD24lowALDHlow, CD44+CD24highALDHhigh, CD44+CD24highALDHlow) in OSCC. All these subpopulations exhibited CSC traits by forming spheres in suspension culture. Moreover, self-renewal and differentiation were observed in these subpopulations; for instance, upon repassage, the CD44+CD24lowALDHhigh subpopulation gave rise to itself and the other three subpopulations, which overexpressed genes of differentiation. Additionally, the subpopulations also exhibited phenotypic plasticity upon exposure to cisplatin. At sub-lethal doses, the CD44+CD24highALDHlow cell phenotype was induced into the CD44+CD24highALDHhigh cell phenotype; the group further attributed the phenotypic change to overexpression of ABCG2 and Sox9 genes rather than apoptosis-mediated selection [168].
Hence, all together, these experimental characteristics of CSCs have indicated that (1) a CSC population is heterogeneous [167] and (2) any given CSC marker can only isolate a fraction of all CSCs [24]. Without specific markers, targeting CSCs for therapeutic elimination will be difficult [169]. As such, other experimental approaches should be looked into to better study CSCs. Here, we propose the use of 3D cell culture technologies to advance the study of CSCs in HN-/OSCC.

Three-Dimensional Cell Culture System in OSCC

Initially used in a neural stem cell study, a sphere formation assay is increasingly being used in CSC studies. In OSCC, CSCs identified and isolated by putative CSC markers consistently proliferated as free-floating spheres in suspension culture (refer to Table 1). Some authors have named these spheres orospheres [87]. As such, the development of non-adherent culture systems has provided an alternative method to study CSCs in addition to the aforementioned marker-based approach. Using OSCC cell lines, Chen and colleagues developed a non-adherent culture system to isolate OSCC-CSCs and reported that spheres generated with this method retained all the CSC traits, as they formed tumors more efficiently when injected into BALB/c nude mice, overexpressed CD133 and ALDH1, and were more chemoresistant than the parental cells [85]. Similarly, sorted CD44+ALDH+ HNSCC cancer cells cultured in ultralow attachment and soft-agar cultures grew as free-floating spheres, which can be propagated with little loss of CSC traits upon several passages [87]. To address the problem of marker specificity, Pozzi and colleagues demonstrated that a sphere formation assay could reliably identify and enrich CSCs from HNSCC for better characterization [55].
Spheres, orospheres, spheroids, multicellular tumor spheroids (MCTS), tumoroids, and organoids are terms used to name proliferating cancer cells grown as 3D cellular structures. Essentially, in these systems, cancer cells are first dissociated into a single-cell suspension and are then grown in scaffold-free or scaffold-based techniques. In a scaffold-free culture, cancer cells freely grow as 3D aggregates. In the scaffold-based technique, however, hydrogel-based support, which mimics the extracellular matrix, is used to support three-dimensional cell growth. Cancer cells grown as 3D structures better replicate the three-dimensional in situ complex characteristics of a tumor. The scaffold-based techniques also provide a platform for modification of the culture system by modifying the types of hydrogel-based supports to more closely replicate the in vivo TME [170,171].
As such, 3D cell culture methods are increasingly being used in HN-/OSCC. After successfully generating organoids derived from healthy oral mucosa, Driehuis and colleagues generated tumoroids from 31 fresh primary HNSCC samples with the use of basement membrane extract (BME) in the scaffold-based technique. The cancer cells grew as tumoroids in this culture system without contamination from immune, connective, or vascular tissues or normal epithelium cells. In addition, the tumoroids also recapitulated original genetic alterations in vitro and in vivo when xenografted. Furthermore, a biobank of tumor organoids was established and used as a tool for treatment outcome prediction and therapeutic exploration. The group reported that the organoids’ response to irradiation mimicked the patients’ response clinically when irradiated with a fraction of a dose corresponding to the treatment dose. The group also discovered that cyclin-dependent kinase inhibitor 2A (CDKN2A) null organoids demonstrated increased sensitivity to EZP01556, a protein arginine methyltransferase 5 (PRMT5) inhibitor, indicating a potential new treatment option for many HNSCC patients who harbored a loss of CDKN2A [172,173]. In another study using the scaffold-free technique, MCTSs derived from HN cancer in 384-well U-bottom, ultralow attachment microplates exhibited permeability barriers that are frequently seen in patient tumors. Fluorescent images of the MCTSs after exposure to chemotherapeutic agents demonstrated a similar pattern of drug penetration in patient tumors and mouse xenografts; the MCTSs exposed to chemoreagents showed penetration of the drugs in the cells of the outer layer of MCTSs, resulting in uneven drug distribution within the MCTSs. This ability to recapitulate drug penetration indicates MCTSs mimic physiological TME and can be used as a model for drug testing [174,175].
A similar approach has also been used in the study of CSCs. A stem-cell-enriched spheroid model (SCESM) was developed [176]. In comparison to adherent cells, the tumor spheroids thus generated showed an increased expression of stem markers, such as ALDH1A1, CD44, and CD133, and core transcription factors of the human stem cell pluripotency signaling pathway, such as Oct4, Sox2, and Nanog. It was also shown that the tumor spheroids were composed of fast proliferating cells on their outer rims and slow or non-proliferating cells at their cores. This is in corroboration with a metabolomic study on CSCs using OSCC-MCTSs, which reported that CSCs relied more on glycolytic than oxidative phosphorylation, with a reduction in fatty acid oxidation activity, suggesting that the CSCs were weakly proliferative [177]. Additionally, a histomorphological analysis of spheroids derived from OSCC cell lines demonstrated a consistent histomorphological pattern of OSCC with a central zone composed of highly pleomorphic and polyhedral cancer cells and more cohesive, flatter cancer cells on the outer rim [178].
One of the common uses for 3D cell culture models has been for drug screening and discovery. In a study using organoids generated from an HNSCC patient, Tanaka and colleagues demonstrated that the resistance displayed by these organoids to cisplatin and docetaxel coincided with recurrence in the patient receiving prior treatment with these chemo reagents [179]. In another study assessing the utility of spheroids from OSCC cell lines to cisplatin and cetuximab, Ono and colleagues reported that cancer cells grown as 3D structures better reflected drug responsiveness than a monolayer culture [114]. Using the 3D model, Zhao and colleagues further demonstrated the potential of targeting the lactate uptake pathway with monocarboxylate transporter 1 (MCT1) knockout in ablating OSCC-CSCs [180].
As cell culture technology continues to evolve, more advanced 3D cell culture models are developed. Ikeda-Motonakano and colleagues developed a high-throughput microfabricated microwell device, which accommodated the growth of 195 spheroids derived from OSCC cell lines, to study CSCs. The spheroids in this system displayed characteristics of CSCs with the formation of a tumor when xenografted, overexpressed CSCs markers such as CD44, and increased resistance to cisplatin, rendering this system useful for the screening of CSC-targeting drug candidates [181]. Recently, Chen and colleagues used a microfluidic chip to isolate a single cell from OSCC for CSC study. The group showed that the microfluidic-chip-assisted capture of OSCC cancer cells yielded significant sphere formation when cultured with the scaffold-free technique. These cells also exhibited CSC characteristics as previously discussed [182].
Taken together, as 3D cell culture technologies continue to evolve and incorporate other advancing technologies in tissue bioengineering, such as microfluidic organ-on-a-chip platform, these models will further replicate the physiological microenvironment more closely for the development of personalized medicine protocols for OSCC patients.

Author Contributions

R.A., S.K.A., Y.F.C., and V.K.V.-C. conceived the idea for this review. R.A., S.K.A., S.S., Y.F.C., I.G., F.K.H., K.K., A.R., and V.K.V.-C. performed the literature search for related articles, filtered through these articles, and found relevant ones that contributed to the targeting of cancer stem cells in OSCC. R.A., S.K.A., S.S., Y.F.C., I.G., F.K.H., and V.K.V.-C. wrote the manuscript. S.K.A., S.S., Y.F.C., I.G., and F.K.H. developed the summative table. S.K.A., S.S., Y.F.C., I.G., and F.K.H. created the figures. R.A., S.K.A., S.S., Y.F.C., I.G., F.K.H., and V.K.V.-C. worked on formatting the manuscript for publishing and made the final revisions. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Siegel, R.L.; Miller, K.D.; Wagle, N.S.; Jemal, A. Cancer statistics, 2023. CA Cancer J. Clin. 2023, 73, 17–48. [Google Scholar] [CrossRef] [PubMed]
  2. Bouvard, V.; Nethan, S.T.; Singh, D.; Warnakulasuriya, S.; Mehrotra, R.; Chaturvedi, A.K.; Chen, T.H.; Ayo-Yusuf, O.A.; Gupta, P.C.; Kerr, A.R.; et al. IARC Perspective on Oral Cancer Prevention. N. Engl. J. Med. 2022, 387, 1999–2005. [Google Scholar] [CrossRef] [PubMed]
  3. Forastiere, A.; Koch, W.; Trotti, A.; Sidransky, D. Head and neck cancer. N. Engl. J. Med. 2001, 345, 1890–1900. [Google Scholar] [CrossRef] [PubMed]
  4. Atashi, F.; Vahed, N.; Emamverdizadeh, P.; Fattahi, S.; Paya, L. Drug resistance against 5-fluorouracil and cisplatin in the treatment of head and neck squamous cell carcinoma: A systematic review. J. Dent. Res. Dent. Clin. Dent. Prospect. 2021, 15, 219–225. [Google Scholar] [CrossRef] [PubMed]
  5. Yao, M.; Chang, K.; Funk, G.F.; Lu, H.; Tan, H.; Wacha, J.; Dornfeld, K.J.; Buatti, J.M. The failure patterns of oral cavity squamous cell carcinoma after intensity-modulated radiotherapy-the university of iowa experience. Int. J. Radiat. Oncol. Biol. Phys. 2007, 67, 1332–1341. [Google Scholar] [CrossRef] [PubMed]
  6. Carvalho, A.L.; Nishimoto, I.N.; Califano, J.A.; Kowalski, L.P. Trends in incidence and prognosis for head and neck cancer in the United States: A site-specific analysis of the SEER database. Int. J. Cancer 2005, 114, 806–816. [Google Scholar] [CrossRef]
  7. Woolgar, J.A.; Rogers, S.; West, C.R.; Errington, R.D.; Brown, J.S.; Vaughan, E.D. Survival and patterns of recurrence in 200 oral cancer patients treated by radical surgery and neck dissection. Oral. Oncol. 1999, 35, 257–265. [Google Scholar] [CrossRef]
  8. Sola, A.M.; Johnson, D.E.; Grandis, J.R. Investigational multitargeted kinase inhibitors in development for head and neck neoplasms. Expert. Opin. Investig. Drugs 2019, 28, 351–363. [Google Scholar] [CrossRef]
  9. Jacobson, J.J.; Epstein, J.B.; Eichmiller, F.C.; Gibson, T.B.; Carls, G.S.; Vogtmann, E.; Wang, S.; Murphy, B. The cost burden of oral, oral pharyngeal, and salivary gland cancers in three groups: Commercial insurance, Medicare, and Medicaid. Head Neck Oncol. 2012, 4, 15. [Google Scholar] [CrossRef]
  10. Chen, J.; Li, Y.; Yu, T.S.; McKay, R.M.; Burns, D.K.; Kernie, S.G.; Parada, L.F. A restricted cell population propagates glioblastoma growth after chemotherapy. Nature 2012, 488, 522–526. [Google Scholar] [CrossRef]
  11. Lapidot, T.; Sirard, C.; Vormoor, J.; Murdoch, B.; Hoang, T.; Caceres-Cortes, J.; Minden, M.; Paterson, B.; Caligiuri, M.A.; Dick, J.E. A cell initiating human acute myeloid leukaemia after transplantation into SCID mice. Nature 1994, 367, 645–648. [Google Scholar] [CrossRef] [PubMed]
  12. Bonnet, D.; Dick, J.E. Human acute myeloid leukemia is organized as a hierarchy that originates from a primitive hematopoietic cell. Nat. Med. 1997, 3, 730–737. [Google Scholar] [CrossRef] [PubMed]
  13. Al-Hajj, M.; Wicha, M.S.; Benito-Hernandez, A.; Morrison, S.J.; Clarke, M.F. Prospective identification of tumorigenic breast cancer cells. Proc. Natl. Acad. Sci. USA 2003, 100, 3983–3988. [Google Scholar] [CrossRef]
  14. Singh, S.K.; Hawkins, C.; Clarke, I.D.; Squire, J.A.; Bayani, J.; Hide, T.; Henkelman, R.M.; Cusimano, M.D.; Dirks, P.B. Identification of human brain tumour initiating cells. Nature 2004, 432, 396–401. [Google Scholar] [CrossRef]
  15. Dalerba, P.; Dylla, S.J.; Park, I.K.; Liu, R.; Wang, X.; Cho, R.W.; Hoey, T.; Gurney, A.; Huang, E.H.; Simeone, D.M.; et al. Phenotypic characterization of human colorectal cancer stem cells. Proc. Natl. Acad. Sci. USA 2007, 104, 10158–10163. [Google Scholar] [CrossRef]
  16. Reya, T.; Morrison, S.J.; Clarke, M.F.; Weissman, I.L. Stem cells, cancer, and cancer stem cells. Nature 2001, 414, 105–111. [Google Scholar] [CrossRef]
  17. Zhang, Z.; Filho, M.S.; Nor, J.E. The biology of head and neck cancer stem cells. Oral. Oncol. 2012, 48, 1–9. [Google Scholar] [CrossRef]
  18. Dorna, D.; Paluszczak, J. Targeting cancer stem cells as a strategy for reducing chemotherapy resistance in head and neck cancers. J. Cancer Res. Clin. Oncol. 2023, 149, 13417–13435. [Google Scholar] [CrossRef] [PubMed]
  19. Wicha, M.S. Targeting self-renewal, an Achilles’ heel of cancer stem cells. Nat. Med. 2014, 20, 14–15. [Google Scholar] [CrossRef]
  20. Zhou, B.B.; Zhang, H.; Damelin, M.; Geles, K.G.; Grindley, J.C.; Dirks, P.B. Tumour-initiating cells: Challenges and opportunities for anticancer drug discovery. Nat. Rev. Drug Discov. 2009, 8, 806–823. [Google Scholar] [CrossRef]
  21. 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 J. Stem Cells 2019, 11, 398–420. [Google Scholar] [CrossRef] [PubMed]
  22. Nix, N.M.; Burdine, O.; Walker, M. Vismodegib: First-in-Class Hedgehog Pathway Inhibitor for Metastatic or Locally Advanced Basal Cell Carcinoma. J. Adv. Pract. Oncol. 2014, 5, 294–296. [Google Scholar] [CrossRef] [PubMed]
  23. Sarkaria, S.M.; Heaney, M.L. Glasdegib in newly diagnosed acute myeloid leukemia. Expert Rev. Anticancer Ther. 2021, 21, 573–581. [Google Scholar] [CrossRef] [PubMed]
  24. Lan, L.; Behrens, A. Are There Specific Cancer Stem Cell Markers? Cancer Res. 2023, 83, 170–172. [Google Scholar] [CrossRef]
  25. Valent, P.; Bonnet, D.; De Maria, R.; Lapidot, T.; Copland, M.; Melo, J.V.; Chomienne, C.; Ishikawa, F.; Schuringa, J.J.; Stassi, G.; et al. Cancer stem cell definitions and terminology: The devil is in the details. Nat. Rev. Cancer 2012, 12, 767–775. [Google Scholar] [CrossRef]
  26. Nguyen, L.V.; Vanner, R.; Dirks, P.; Eaves, C.J. Cancer stem cells: An evolving concept. Nat. Rev. Cancer 2012, 12, 133–143. [Google Scholar] [CrossRef]
  27. Blanpain, C. Tracing the cellular origin of cancer. Nat. Cell Biol. 2013, 15, 126–134. [Google Scholar] [CrossRef]
  28. Sell, S. Cellular origin of cancer: Dedifferentiation or stem cell maturation arrest? Environ. Health Perspect. 1993, 101 (Suppl. 5), 15–26. [Google Scholar] [CrossRef]
  29. Clarke, M.F.; Dick, J.E.; Dirks, P.B.; Eaves, C.J.; Jamieson, C.H.; Jones, D.L.; Visvader, J.; Weissman, I.L.; Wahl, G.M. Cancer stem cells--perspectives on current status and future directions: AACR Workshop on cancer stem cells. Cancer Res. 2006, 66, 9339–9344. [Google Scholar] [CrossRef]
  30. Magee, J.A.; Piskounova, E.; Morrison, S.J. Cancer stem cells: Impact, heterogeneity, and uncertainty. Cancer Cell 2012, 21, 283–296. [Google Scholar] [CrossRef]
  31. Meacham, C.E.; Morrison, S.J. Tumour heterogeneity and cancer cell plasticity. Nature 2013, 501, 328–337. [Google Scholar] [CrossRef] [PubMed]
  32. Nowell, P.C. The clonal evolution of tumor cell populations. Science 1976, 194, 23–28. [Google Scholar] [CrossRef] [PubMed]
  33. Nowell, P.C. Mechanisms of tumor progression. Cancer Res. 1986, 46, 2203–2207. [Google Scholar] [PubMed]
  34. Jordan, C.T.; Guzman, M.L.; Noble, M. Cancer stem cells. N. Engl. J. Med. 2006, 355, 1253–1261. [Google Scholar] [CrossRef]
  35. Tang, X.H.; Scognamiglio, T.; Gudas, L.J. Basal stem cells contribute to squamous cell carcinomas in the oral cavity. Carcinogenesis 2013, 34, 1158–1164. [Google Scholar] [CrossRef]
  36. Joseph, N.M.; Mosher, J.T.; Buchstaller, J.; Snider, P.; McKeever, P.E.; Lim, M.; Conway, S.J.; Parada, L.F.; Zhu, Y.; Morrison, S.J. The loss of Nf1 transiently promotes self-renewal but not tumorigenesis by neural crest stem cells. Cancer Cell 2008, 13, 129–140. [Google Scholar] [CrossRef] [PubMed]
  37. Liu, C.; Sage, J.C.; Miller, M.R.; Verhaak, R.G.; Hippenmeyer, S.; Vogel, H.; Foreman, O.; Bronson, R.T.; Nishiyama, A.; Luo, L.; et al. Mosaic analysis with double markers reveals tumor cell of origin in glioma. Cell 2011, 146, 209–221. [Google Scholar] [CrossRef]
  38. Wang, Y.; Yang, J.; Zheng, H.; Tomasek, G.J.; Zhang, P.; McKeever, P.E.; Lee, E.Y.; Zhu, Y. Expression of mutant p53 proteins implicates a lineage relationship between neural stem cells and malignant astrocytic glioma in a murine model. Cancer Cell 2009, 15, 514–526. [Google Scholar] [CrossRef]
  39. Alcantara Llaguno, S.; Chen, J.; Kwon, C.H.; Jackson, E.L.; Li, Y.; Burns, D.K.; Alvarez-Buylla, A.; Parada, L.F. Malignant astrocytomas originate from neural stem/progenitor cells in a somatic tumor suppressor mouse model. Cancer Cell 2009, 15, 45–56. [Google Scholar] [CrossRef]
  40. Singh, S.K.; Clarke, I.D.; Terasaki, M.; Bonn, V.E.; Hawkins, C.; Squire, J.; Dirks, P.B. Identification of a cancer stem cell in human brain tumors. Cancer Res. 2003, 63, 5821–5828. [Google Scholar]
  41. O’Brien, C.A.; Pollett, A.; Gallinger, S.; Dick, J.E. A human colon cancer cell capable of initiating tumour growth in immunodeficient mice. Nature 2007, 445, 106–110. [Google Scholar] [CrossRef]
  42. Li, C.; Heidt, D.G.; Dalerba, P.; Burant, C.F.; Zhang, L.; Adsay, V.; Wicha, M.; Clarke, M.F.; Simeone, D.M. Identification of pancreatic cancer stem cells. Cancer Res. 2007, 67, 1030–1037. [Google Scholar] [CrossRef] [PubMed]
  43. Curley, M.D.; Therrien, V.A.; Cummings, C.L.; Sergent, P.A.; Koulouris, C.R.; Friel, A.M.; Roberts, D.J.; Seiden, M.V.; Scadden, D.T.; Rueda, B.R.; et al. CD133 expression defines a tumor initiating cell population in primary human ovarian cancer. Stem Cells 2009, 27, 2875–2883. [Google Scholar] [CrossRef]
  44. Prince, M.E.; Sivanandan, R.; Kaczorowski, A.; Wolf, G.T.; Kaplan, M.J.; Dalerba, P.; Weissman, I.L.; Clarke, M.F.; Ailles, L.E. Identification of a subpopulation of cells with cancer stem cell properties in head and neck squamous cell carcinoma. Proc. Natl. Acad. Sci. USA 2007, 104, 973–978. [Google Scholar] [CrossRef]
  45. Chiou, S.H.; Yu, C.C.; Huang, C.Y.; Lin, S.C.; Liu, C.J.; Tsai, T.H.; Chou, S.H.; Chien, C.S.; Ku, H.H.; Lo, J.F. Positive correlations of Oct-4 and Nanog in oral cancer stem-like cells and high-grade oral squamous cell carcinoma. Clin. Cancer Res. 2008, 14, 4085–4095. [Google Scholar] [CrossRef] [PubMed]
  46. Chen, Y.C.; Chen, Y.W.; Hsu, H.S.; Tseng, L.M.; Huang, P.I.; Lu, K.H.; Chen, D.T.; Tai, L.K.; Yung, M.C.; Chang, S.C.; et al. Aldehyde dehydrogenase 1 is a putative marker for cancer stem cells in head and neck squamous cancer. Biochem. Biophys. Res. Commun. 2009, 385, 307–313. [Google Scholar] [CrossRef]
  47. Clay, M.R.; Tabor, M.; Owen, J.H.; Carey, T.E.; Bradford, C.R.; Wolf, G.T.; Wicha, M.S.; Prince, M.E. Single-marker identification of head and neck squamous cell carcinoma cancer stem cells with aldehyde dehydrogenase. Head Neck 2010, 32, 1195–1201. [Google Scholar] [CrossRef]
  48. Chen, Y.W.; Chen, K.H.; Huang, P.I.; Chen, Y.C.; Chiou, G.Y.; Lo, W.L.; Tseng, L.M.; Hsu, H.S.; Chang, K.W.; Chiou, S.H. Cucurbitacin I suppressed stem-like property and enhanced radiation-induced apoptosis in head and neck squamous carcinoma--derived CD44+ALDH1+ cells. Mol. Cancer Ther. 2010, 9, 2879–2892. [Google Scholar] [CrossRef] [PubMed]
  49. Zhang, Q.; Shi, S.; Yen, Y.; Brown, J.; Ta, J.Q.; Le, A.D. A subpopulation of CD133+ cancer stem-like cells characterized in human oral squamous cell carcinoma confer resistance to chemotherapy. Cancer Lett. 2010, 289, 151–160. [Google Scholar] [CrossRef]
  50. Biddle, A.; Liang, X.; Gammon, L.; Fazil, B.; Harper, L.J.; Emich, H.; Costea, D.E.; Mackenzie, I.C. Cancer stem cells in squamous cell carcinoma switch between two distinct phenotypes that are preferentially migratory or proliferative. Cancer Res. 2011, 71, 5317–5326. [Google Scholar] [CrossRef]
  51. Sun, Y.; Han, J.; Lu, Y.; Yang, X.; Fan, M. Biological characteristics of a cell subpopulation in tongue squamous cell carcinoma. Oral. Dis. 2012, 18, 169–177. [Google Scholar] [CrossRef] [PubMed]
  52. Han, J.; Fujisawa, T.; Husain, S.R.; Puri, R.K. Identification and characterization of cancer stem cells in human head and neck squamous cell carcinoma. BMC Cancer 2014, 14, 173. [Google Scholar] [CrossRef] [PubMed]
  53. Kumar, B.; Yadav, A.; Lang, J.C.; Teknos, T.N.; Kumar, P. Suberoylanilide hydroxamic acid (SAHA) reverses chemoresistance in head and neck cancer cells by targeting cancer stem cells via the downregulation of nanog. Genes Cancer 2015, 6, 169–181. [Google Scholar] [CrossRef]
  54. Yu, C.C.; Hu, F.W.; Yu, C.H.; Chou, M.Y. Targeting CD133 in the enhancement of chemosensitivity in oral squamous cell carcinoma-derived side population cancer stem cells. Head Neck 2016, 38 (Suppl. 1), E231–E238. [Google Scholar] [CrossRef]
  55. Pozzi, V.; Sartini, D.; Rocchetti, R.; Santarelli, A.; Rubini, C.; Morganti, S.; Giuliante, R.; Calabrese, S.; Di Ruscio, G.; Orlando, F.; et al. Identification and characterization of cancer stem cells from head and neck squamous cell carcinoma cell lines. Cell Physiol. Biochem. 2015, 36, 784–798. [Google Scholar] [CrossRef] [PubMed]
  56. Biddle, A.; Gammon, L.; Liang, X.; Costea, D.E.; Mackenzie, I.C. Phenotypic Plasticity Determines Cancer Stem Cell Therapeutic Resistance in Oral Squamous Cell Carcinoma. eBioMedicine 2016, 4, 138–145. [Google Scholar] [CrossRef]
  57. Kaseb, H.O.; Fohrer-Ting, H.; Lewis, D.W.; Lagasse, E.; Gollin, S.M. Identification, expansion and characterization of cancer cells with stem cell properties from head and neck squamous cell carcinomas. Exp. Cell Res. 2016, 348, 75–86. [Google Scholar] [CrossRef]
  58. Liu, C.M.; Peng, C.Y.; Liao, Y.W.; Lu, M.Y.; Tsai, M.L.; Yeh, J.C.; Yu, C.H.; Yu, C.C. Sulforaphane targets cancer stemness and tumor initiating properties in oral squamous cell carcinomas via miR-200c induction. J. Formos. Med. Assoc. 2017, 116, 41–48. [Google Scholar] [CrossRef] [PubMed]
  59. Xu, Q.; Zhang, Q.; Ishida, Y.; Hajjar, S.; Tang, X.; Shi, H.; Dang, C.V.; Le, A.D. EGF induces epithelial-mesenchymal transition and cancer stem-like cell properties in human oral cancer cells via promoting Warburg effect. Oncotarget 2017, 8, 9557–9571. [Google Scholar] [CrossRef]
  60. Chen, D.; Wu, M.; Li, Y.; Chang, I.; Yuan, Q.; Ekimyan-Salvo, M.; Deng, P.; Yu, B.; Yu, Y.; Dong, J.; et al. Targeting BMI1+ Cancer Stem Cells Overcomes Chemoresistance and Inhibits Metastases in Squamous Cell Carcinoma. Cell Stem Cell 2017, 20, 621–634.e6. [Google Scholar] [CrossRef]
  61. Elkashty, O.A.; Abu Elghanam, G.; Su, X.; Liu, Y.; Chauvin, P.J.; Tran, S.D. Cancer stem cells enrichment with surface markers CD271 and CD44 in human head and neck squamous cell carcinomas. Carcinogenesis 2020, 41, 458–466. [Google Scholar] [CrossRef] [PubMed]
  62. Ma, Z.; Zhang, C.; Liu, X.; Fang, F.; Liu, S.; Liao, X.; Tao, S.; Mai, H. Characterisation of a subpopulation of CD133+ cancer stem cells from Chinese patients with oral squamous cell carcinoma. Sci. Rep. 2020, 10, 8875. [Google Scholar] [CrossRef] [PubMed]
  63. You, X.; Zhou, Z.; Chen, W.; Wei, X.; Zhou, H.; Luo, W. MicroRNA-495 confers inhibitory effects on cancer stem cells in oral squamous cell carcinoma through the HOXC6-mediated TGF-beta signaling pathway. Stem Cell Res. Ther. 2020, 11, 117. [Google Scholar] [CrossRef]
  64. Amor, N.G.; Buzo, R.F.; Ortiz, R.C.; Lopes, N.M.; Saito, L.M.; Mackenzie, I.C.; Rodini, C.O. In vitro and in vivo characterization of cancer stem cell subpopulations in oral squamous cell carcinoma. J. Oral. Pathol. Med. 2021, 50, 52–59. [Google Scholar] [CrossRef]
  65. Ghuwalewala, S.; Ghatak, D.; Das, S.; Roy, S.; Das, P.; Butti, R.; Gorain, M.; Nath, S.; Kundu, G.C.; Roychoudhury, S. MiRNA-146a/AKT/beta-Catenin Activation Regulates Cancer Stem Cell Phenotype in Oral Squamous Cell Carcinoma by Targeting CD24. Front. Oncol. 2021, 11, 651692. [Google Scholar] [CrossRef]
  66. Aquino, I.G.; Cuadra-Zelaya, F.J.M.; Bizeli, A.L.V.; Palma, P.V.B.; Mariano, F.V.; Salo, T.; Coletta, R.D.; Bastos, D.C.; Graner, E. Isolation and phenotypic characterization of cancer stem cells from metastatic oral cancer cells. Oral. Dis. 2024; early view. [Google Scholar] [CrossRef]
  67. Huang, J.; Li, H.; Yang, Z.; Liu, R.; Li, Y.; Hu, Y.; Zhao, S.; Gao, X.; Yang, X.; Wei, J. SALL4 promotes cancer stem-like cell phenotype and radioresistance in oral squamous cell carcinomas via methyltransferase-like 3-mediated m6A modification. Cell Death Dis. 2024, 15, 139. [Google Scholar] [CrossRef]
  68. de Castro, L.R.; de Oliveira, L.D.; Milan, T.M.; Eskenazi, A.P.E.; Bighetti-Trevisan, R.L.; de Almeida, O.G.G.; Amorim, M.L.M.; Squarize, C.H.; Castilho, R.M.; de Almeida, L.O. Up-regulation of TNF-alpha/NFkB/SIRT1 axis drives aggressiveness and cancer stem cells accumulation in chemoresistant oral squamous cell carcinoma. J. Cell. Physiol. 2024, 239, e31164. [Google Scholar] [CrossRef] [PubMed]
  69. Barrandon, Y.; Green, H. Three clonal types of keratinocyte with different capacities for multiplication. Proc. Natl. Acad. Sci. USA 1987, 84, 2302–2306. [Google Scholar] [CrossRef] [PubMed]
  70. Rochat, A.; Barrandon, Y. Regeneration of epidermis from adult keratinocyte stem cells. In Essentials of Stem Cell Biology; Elsevier: Amsterdam, The Netherlands, 2009; pp. 551–560. [Google Scholar]
  71. Nakamura, T.; Endo, K.; Kinoshita, S. Identification of human oral keratinocyte stem/progenitor cells by neurotrophin receptor p75 and the role of neurotrophin/p75 signaling. Stem Cells 2007, 25, 628–638. [Google Scholar] [CrossRef]
  72. Luo, X.; Okubo, T.; Randell, S.; Hogan, B.L. Culture of endodermal stem/progenitor cells of the mouse tongue. In Vitro Cell. Dev. Biol. Anim. 2009, 45, 44–54. [Google Scholar] [CrossRef]
  73. Locke, M.; Heywood, M.; Fawell, S.; Mackenzie, I.C. Retention of intrinsic stem cell hierarchies in carcinoma-derived cell lines. Cancer Res. 2005, 65, 8944–8950. [Google Scholar] [CrossRef] [PubMed]
  74. Harper, L.J.; Piper, K.; Common, J.; Fortune, F.; Mackenzie, I.C. Stem cell patterns in cell lines derived from head and neck squamous cell carcinoma. J. Oral. Pathol. Med. 2007, 36, 594–603. [Google Scholar] [CrossRef] [PubMed]
  75. Zhang, K.; Waxman, D.J. PC3 prostate tumor-initiating cells with molecular profile FAM65Bhigh/MFI2low/LEF1low increase tumor angiogenesis. Mol. Cancer 2010, 9, 319. [Google Scholar] [CrossRef]
  76. Tan, L.; Sui, X.; Deng, H.; Ding, M. Holoclone forming cells from pancreatic cancer cells enrich tumor initiating cells and represent a novel model for study of cancer stem cells. PLoS ONE 2011, 6, e23383. [Google Scholar] [CrossRef] [PubMed]
  77. Beaver, C.M.; Ahmed, A.; Masters, J.R. Clonogenicity: Holoclones and meroclones contain stem cells. PLoS ONE 2014, 9, e89834. [Google Scholar] [CrossRef] [PubMed]
  78. Manley, E., Jr.; Waxman, D.J. H460 non-small cell lung cancer stem-like holoclones yield tumors with increased vascularity. Cancer Lett. 2014, 346, 63–73. [Google Scholar] [CrossRef]
  79. Liu, T.J.; Sun, B.C.; Zhao, X.L.; Zhao, X.M.; Sun, T.; Gu, Q.; Yao, Z.; Dong, X.Y.; Zhao, N.; Liu, N. CD133+ cells with cancer stem cell characteristics associates with vasculogenic mimicry in triple-negative breast cancer. Oncogene 2013, 32, 544–553. [Google Scholar] [CrossRef]
  80. Johansson, C.B.; Momma, S.; Clarke, D.L.; Risling, M.; Lendahl, U.; Frisen, J. Identification of a neural stem cell in the adult mammalian central nervous system. Cell 1999, 96, 25–34. [Google Scholar] [CrossRef]
  81. Bez, A.; Corsini, E.; Curti, D.; Biggiogera, M.; Colombo, A.; Nicosia, R.F.; Pagano, S.F.; Parati, E.A. Neurosphere and neurosphere-forming cells: Morphological and ultrastructural characterization. Brain Res. 2003, 993, 18–29. [Google Scholar] [CrossRef]
  82. Vescovi, A.L.; Reynolds, B.A.; Fraser, D.D.; Weiss, S. bFGF regulates the proliferative fate of unipotent (neuronal) and bipotent (neuronal/astroglial) EGF-generated CNS progenitor cells. Neuron 1993, 11, 951–966. [Google Scholar] [CrossRef]
  83. Pastrana, E.; Silva-Vargas, V.; Doetsch, F. Eyes wide open: A critical review of sphere-formation as an assay for stem cells. Cell Stem Cell 2011, 8, 486–498. [Google Scholar] [CrossRef] [PubMed]
  84. Bahmad, H.F.; Cheaito, K.; Chalhoub, R.M.; Hadadeh, O.; Monzer, A.; Ballout, F.; El-Hajj, A.; Mukherji, D.; Liu, Y.N.; Daoud, G.; et al. Sphere-Formation Assay: Three-Dimensional in vitro Culturing of Prostate Cancer Stem/Progenitor Sphere-Forming Cells. Front. Oncol. 2018, 8, 347. [Google Scholar] [CrossRef] [PubMed]
  85. Chen, S.F.; Chang, Y.C.; Nieh, S.; Liu, C.L.; Yang, C.Y.; Lin, Y.S. Nonadhesive culture system as a model of rapid sphere formation with cancer stem cell properties. PLoS ONE 2012, 7, e31864. [Google Scholar] [CrossRef] [PubMed]
  86. Okamoto, A.; Chikamatsu, K.; Sakakura, K.; Hatsushika, K.; Takahashi, G.; Masuyama, K. Expansion and characterization of cancer stem-like cells in squamous cell carcinoma of the head and neck. Oral. Oncol. 2009, 45, 633–639. [Google Scholar] [CrossRef]
  87. Krishnamurthy, S.; Nor, J.E. Orosphere assay: A method for propagation of head and neck cancer stem cells. Head Neck 2013, 35, 1015–1021. [Google Scholar] [CrossRef]
  88. Cianciosi, D.; Ansary, J.; Forbes-Hernandez, T.Y.; Regolo, L.; Quinzi, D.; Gracia Villar, S.; Garcia Villena, E.; Tutusaus Pifarre, K.; Alvarez-Suarez, J.M.; Battino, M.; et al. The Molecular Basis of Different Approaches for the Study of Cancer Stem Cells and the Advantages and Disadvantages of a Three-Dimensional Culture. Molecules 2021, 26, 2615. [Google Scholar] [CrossRef] [PubMed]
  89. Pronoy, T.U.H.; Islam, F.; Gopalan, V.; Lam, A.K. Surface Markers for the Identification of Cancer Stem Cells. Methods Mol. Biol. 2024, 2777, 51–69. [Google Scholar] [CrossRef]
  90. Gopalan, V.; Islam, F.; Lam, A.K. Surface Markers for the Identification of Cancer Stem Cells. Methods Mol. Biol. 2018, 1692, 17–29. [Google Scholar] [CrossRef]
  91. Singh, P.; Augustine, D.; Rao, R.S.; Patil, S.; Awan, K.H.; Sowmya, S.V.; Haragannavar, V.C.; Prasad, K. Role of cancer stem cells in head-and-neck squamous cell carcinoma—A systematic review. J. Carcinog. 2021, 20, 12. [Google Scholar] [CrossRef]
  92. Hassn Mesrati, M.; Syafruddin, S.E.; Mohtar, M.A.; Syahir, A. CD44: A Multifunctional Mediator of Cancer Progression. Biomolecules 2021, 11, 1850. [Google Scholar] [CrossRef]
  93. Ghuwalewala, S.; Ghatak, D.; Das, P.; Dey, S.; Sarkar, S.; Alam, N.; Panda, C.K.; Roychoudhury, S. CD44highCD24low molecular signature determines the Cancer Stem Cell and EMT phenotype in Oral Squamous Cell Carcinoma. Stem Cell Res. 2016, 16, 405–417. [Google Scholar] [CrossRef] [PubMed]
  94. Oliveira, L.R.; Oliveira-Costa, J.P.; Araujo, I.M.; Soave, D.F.; Zanetti, J.S.; Soares, F.A.; Zucoloto, S.; Ribeiro-Silva, A. Cancer stem cell immunophenotypes in oral squamous cell carcinoma. J. Oral. Pathol. Med. 2011, 40, 135–142. [Google Scholar] [CrossRef]
  95. Zhang, H.; Brown, R.L.; Wei, Y.; Zhao, P.; Liu, S.; Liu, X.; Deng, Y.; Hu, X.; Zhang, J.; Gao, X.D.; et al. CD44 splice isoform switching determines breast cancer stem cell state. Genes Dev. 2019, 33, 166–179. [Google Scholar] [CrossRef] [PubMed]
  96. Bai, J.; Chen, Y.; Sun, Y.; Wang, X.; Wang, Y.; Guo, S.; Shang, Z.; Shao, Z. EphA2 promotes the transcription of KLF4 to facilitate stemness in oral squamous cell carcinoma. Cell. Mol. Life Sci. 2024, 81, 278. [Google Scholar] [CrossRef]
  97. Todoroki, K.; Abe, Y.; Matsuo, K.; Nomura, H.; Kawahara, A.; Nakamura, Y.; Nakamura, M.; Seki, N.; Kusukawa, J. Prognostic effect of programmed cell death ligand 1/programmed cell death 1 expression in cancer stem cells of human oral squamous cell carcinoma. Oncol. Lett. 2024, 27, 79. [Google Scholar] [CrossRef]
  98. Marchitti, S.A.; Brocker, C.; Stagos, D.; Vasiliou, V. Non-P450 aldehyde oxidizing enzymes: The aldehyde dehydrogenase superfamily. Expert. Opin. Drug Metab. Toxicol. 2008, 4, 697–720. [Google Scholar] [CrossRef]
  99. Black, W.; Vasiliou, V. The aldehyde dehydrogenase gene superfamily resource center. Hum. Genom. 2009, 4, 136–142. [Google Scholar] [CrossRef] [PubMed]
  100. Ginestier, C.; Wicinski, J.; Cervera, N.; Monville, F.; Finetti, P.; Bertucci, F.; Wicha, M.S.; Birnbaum, D.; Charafe-Jauffret, E. Retinoid signaling regulates breast cancer stem cell differentiation. Cell Cycle 2009, 8, 3297–3302. [Google Scholar] [CrossRef]
  101. Tang, X.H.; Gudas, L.J. Retinoids, retinoic acid receptors, and cancer. Annu. Rev. Pathol. 2011, 6, 345–364. [Google Scholar] [CrossRef]
  102. Gudas, L.J.; Wagner, J.A. Retinoids regulate stem cell differentiation. J. Cell Physiol. 2011, 226, 322–330. [Google Scholar] [CrossRef]
  103. Kim, J.; Shin, J.H.; Chen, C.H.; Cruz, L.; Farnebo, L.; Yang, J.; Borges, P.; Kang, G.; Mochly-Rosen, D.; Sunwoo, J.B. Targeting aldehyde dehydrogenase activity in head and neck squamous cell carcinoma with a novel small molecule inhibitor. Oncotarget 2017, 8, 52345–52356. [Google Scholar] [CrossRef] [PubMed]
  104. Duan, J.J.; Cai, J.; Gao, L.; Yu, S.C. ALDEFLUOR activity, ALDH isoforms, and their clinical significance in cancers. J. Enzym. Inhib. Med. Chem. 2023, 38, 2166035. [Google Scholar] [CrossRef] [PubMed]
  105. Ortiz, R.C.; Amor, N.G.; Saito, L.M.; Santesso, M.R.; Lopes, N.M.; Buzo, R.F.; Fonseca, A.C.; Amaral-Silva, G.K.; Moyses, R.A.; Rodini, C.O. CSChighE-cadherinlow immunohistochemistry panel predicts poor prognosis in oral squamous cell carcinoma. Sci. Rep. 2024, 14, 10583. [Google Scholar] [CrossRef] [PubMed]
  106. Barzegar Behrooz, A.; Syahir, A.; Ahmad, S. CD133: Beyond a cancer stem cell biomarker. J. Drug Target. 2019, 27, 257–269. [Google Scholar] [CrossRef]
  107. Yin, A.H.; Miraglia, S.; Zanjani, E.D.; Almeida-Porada, G.; Ogawa, M.; Leary, A.G.; Olweus, J.; Kearney, J.; Buck, D.W. AC133, a novel marker for human hematopoietic stem and progenitor cells. Blood 1997, 90, 5002–5012. [Google Scholar] [CrossRef]
  108. Miraglia, S.; Godfrey, W.; Yin, A.H.; Atkins, K.; Warnke, R.; Holden, J.T.; Bray, R.A.; Waller, E.K.; Buck, D.W. A novel five-transmembrane hematopoietic stem cell antigen: Isolation, characterization, and molecular cloning. Blood 1997, 90, 5013–5021. [Google Scholar] [CrossRef]
  109. Zhou, L.; Wei, X.; Cheng, L.; Tian, J.; Jiang, J.J. CD133, one of the markers of cancer stem cells in Hep-2 cell line. Laryngoscope 2007, 117, 455–460. [Google Scholar] [CrossRef]
  110. Bhave, S.L.; Teknos, T.N.; Pan, Q. Molecular parameters of head and neck cancer metastasis. Crit. Rev. Eukaryot. Gene Expr. 2011, 21, 143–153. [Google Scholar] [CrossRef]
  111. Ionna, F.; Bossi, P.; Guida, A.; Alberti, A.; Muto, P.; Salzano, G.; Ottaiano, A.; Maglitto, F.; Leopardo, D.; De Felice, M.; et al. Recurrent/Metastatic Squamous Cell Carcinoma of the Head and Neck: A Big and Intriguing Challenge Which May Be Resolved by Integrated Treatments Combining Locoregional and Systemic Therapies. Cancers 2021, 13, 2371. [Google Scholar] [CrossRef]
  112. Peitzsch, C.; Nathansen, J.; Schniewind, S.I.; Schwarz, F.; Dubrovska, A. Cancer Stem Cells in Head and Neck Squamous Cell Carcinoma: Identification, Characterization and Clinical Implications. Cancers 2019, 11, 616. [Google Scholar] [CrossRef]
  113. Rossi, M.; Blasi, P. Multicellular Tumor Spheroids in Nanomedicine Research: A Perspective. Front. Med. Technol. 2022, 4, 909943. [Google Scholar] [CrossRef] [PubMed]
  114. Ono, K.; Sato, K.; Nakamura, T.; Yoshida, Y.; Murata, S.; Yoshida, K.; Kanemoto, H.; Umemori, K.; Kawai, H.; Obata, K.; et al. Reproduction of the Antitumor Effect of Cisplatin and Cetuximab Using a Three-dimensional Spheroid Model in Oral Cancer. Int. J. Med. Sci. 2022, 19, 1320–1333. [Google Scholar] [CrossRef] [PubMed]
  115. Dolatabadi, S.; Jonasson, E.; Linden, M.; Fereydouni, B.; Backsten, K.; Nilsson, M.; Martner, A.; Forootan, A.; Fagman, H.; Landberg, G.; et al. JAK-STAT signalling controls cancer stem cell properties including chemotherapy resistance in myxoid liposarcoma. Int. J. Cancer 2019, 145, 435–449. [Google Scholar] [CrossRef]
  116. Park, S.Y.; Lee, C.J.; Choi, J.H.; Kim, J.H.; Kim, J.W.; Kim, J.Y.; Nam, J.S. The JAK2/STAT3/CCND2 Axis promotes colorectal Cancer stem cell persistence and radioresistance. J. Exp. Clin. Cancer Res. 2019, 38, 399. [Google Scholar] [CrossRef]
  117. Zhan, T.; Rindtorff, N.; Boutros, M. Wnt signaling in cancer. Oncogene 2017, 36, 1461–1473. [Google Scholar] [CrossRef]
  118. Roy, S.; Roy, S.; Kar, M.; Chakraborty, A.; Kumar, A.; Delogu, F.; Asthana, S.; Hande, M.P.; Banerjee, B. Combined treatment with cisplatin and the tankyrase inhibitor XAV-939 increases cytotoxicity, abrogates cancer-stem-like cell phenotype and increases chemosensitivity of head-and-neck squamous-cell carcinoma cells. Mutat. Res. Genet. Toxicol. Environ. Mutagen. 2019, 846, 503084. [Google Scholar] [CrossRef] [PubMed]
  119. Karami Fath, M.; Ebrahimi, M.; Nourbakhsh, E.; Zia Hazara, A.; Mirzaei, A.; Shafieyari, S.; Salehi, A.; Hoseinzadeh, M.; Payandeh, Z.; Barati, G. PI3K/Akt/mTOR signaling pathway in cancer stem cells. Pathol. Res. Pract. 2022, 237, 154010. [Google Scholar] [CrossRef]
  120. Keysar, S.B.; Le, P.N.; Miller, B.; Jackson, B.C.; Eagles, J.R.; Nieto, C.; Kim, J.; Tang, B.; Glogowska, M.J.; Morton, J.J.; et al. Regulation of Head and Neck Squamous Cancer Stem Cells by PI3K and SOX2. J. Natl. Cancer Inst. 2017, 109, djw189. [Google Scholar] [CrossRef]
  121. Elkabets, M.; Pazarentzos, E.; Juric, D.; Sheng, Q.; Pelossof, R.A.; Brook, S.; Benzaken, A.O.; Rodon, J.; Morse, N.; Yan, J.J.; et al. AXL mediates resistance to PI3Kalpha inhibition by activating the EGFR/PKC/mTOR axis in head and neck and esophageal squamous cell carcinomas. Cancer Cell 2015, 27, 533–546. [Google Scholar] [CrossRef]
  122. Liu, C.; Zhou, S.; Tang, W. USP14 promotes the cancer stem-like cell properties of OSCC via promoting SOX2 deubiquitination. Oral. Dis. 2024; early view. [Google Scholar] [CrossRef]
  123. Chen, X.; Cao, Y.; Sedhom, W.; Lu, L.; Liu, Y.; Wang, H.; Oka, M.; Bornstein, S.; Said, S.; Song, J.; et al. Distinct roles of PIK3CA in the enrichment and maintenance of cancer stem cells in head and neck squamous cell carcinoma. Mol. Oncol. 2020, 14, 139–158. [Google Scholar] [CrossRef]
  124. Matsui, W.H. Cancer stem cell signaling pathways. Medicine 2016, 95, S8–S19. [Google Scholar] [CrossRef] [PubMed]
  125. Krishnamurthy, S.; Nor, J.E. Head and neck cancer stem cells. J. Dent. Res. 2012, 91, 334–340. [Google Scholar] [CrossRef] [PubMed]
  126. Yang, J.; Weinberg, R.A. Epithelial-mesenchymal transition: At the crossroads of development and tumor metastasis. Dev. Cell 2008, 14, 818–829. [Google Scholar] [CrossRef]
  127. Seino, S.; Shigeishi, H.; Hashikata, M.; Higashikawa, K.; Tobiume, K.; Uetsuki, R.; Ishida, Y.; Sasaki, K.; Naruse, T.; Rahman, M.Z.; et al. CD44high /ALDH1high head and neck squamous cell carcinoma cells exhibit mesenchymal characteristics and GSK3beta-dependent cancer stem cell properties. J. Oral. Pathol. Med. 2016, 45, 180–188. [Google Scholar] [CrossRef]
  128. Ortiz-Cuaran, S.; Bouaoud, J.; Karabajakian, A.; Fayette, J.; Saintigny, P. Precision Medicine Approaches to Overcome Resistance to Therapy in Head and Neck Cancers. Front. Oncol. 2021, 11, 614332. [Google Scholar] [CrossRef]
  129. Shahoumi, L.A. Oral Cancer Stem Cells: Therapeutic Implications and Challenges. Front. Oral. Health 2021, 2, 685236. [Google Scholar] [CrossRef]
  130. Masui, T.; Ota, I.; Yook, J.I.; Mikami, S.; Yane, K.; Yamanaka, T.; Hosoi, H. Snail-induced epithelial-mesenchymal transition promotes cancer stem cell-like phenotype in head and neck cancer cells. Int. J. Oncol. 2014, 44, 693–699. [Google Scholar] [CrossRef]
  131. Ota, I.; Masui, T.; Kurihara, M.; Yook, J.I.; Mikami, S.; Kimura, T.; Shimada, K.; Konishi, N.; Yane, K.; Yamanaka, T.; et al. Snail-induced EMT promotes cancer stem cell-like properties in head and neck cancer cells. Oncol. Rep. 2016, 35, 261–266. [Google Scholar] [CrossRef] [PubMed]
  132. Khammanivong, A.; Gopalakrishnan, R.; Dickerson, E.B. SMURF1 silencing diminishes a CD44-high cancer stem cell-like population in head and neck squamous cell carcinoma. Mol. Cancer 2014, 13, 260. [Google Scholar] [CrossRef]
  133. Byun, J.Y.; Huang, K.; Lee, J.S.; Huang, W.; Hu, L.; Zheng, X.; Tang, X.; Li, F.; Jo, D.G.; Song, X.; et al. Targeting HIF-1alpha/NOTCH1 pathway eliminates CD44+ cancer stem-like cell phenotypes, malignancy, and resistance to therapy in head and neck squamous cell carcinoma. Oncogene 2022, 41, 1352–1363. [Google Scholar] [CrossRef]
  134. Lee, S.H.; Do, S.I.; Lee, H.J.; Kang, H.J.; Koo, B.S.; Lim, Y.C. Notch1 signaling contributes to stemness in head and neck squamous cell carcinoma. Lab. Investig. 2016, 96, 508–516. [Google Scholar] [CrossRef] [PubMed]
  135. Xie, S.L.; Fan, S.; Zhang, S.Y.; Chen, W.X.; Li, Q.X.; Pan, G.K.; Zhang, H.Q.; Wang, W.W.; Weng, B.; Zhang, Z.; et al. SOX8 regulates cancer stem-like properties and cisplatin-induced EMT in tongue squamous cell carcinoma by acting on the Wnt/beta-catenin pathway. Int. J. Cancer 2018, 142, 1252–1265. [Google Scholar] [CrossRef] [PubMed]
  136. Bourguignon, L.Y.; Wong, G.; Earle, C.; Chen, L. Hyaluronan-CD44v3 interaction with Oct4-Sox2-Nanog promotes miR-302 expression leading to self-renewal, clonal formation, and cisplatin resistance in cancer stem cells from head and neck squamous cell carcinoma. J. Biol. Chem. 2012, 287, 32800–32824. [Google Scholar] [CrossRef] [PubMed]
  137. Bourguignon, L.Y.; Wong, G.; Shiina, M. Up-regulation of Histone Methyltransferase, DOT1L, by Matrix Hyaluronan Promotes MicroRNA-10 Expression Leading to Tumor Cell Invasion and Chemoresistance in Cancer Stem Cells from Head and Neck Squamous Cell Carcinoma. J. Biol. Chem. 2016, 291, 10571–10585. [Google Scholar] [CrossRef] [PubMed]
  138. Warrier, S.; Bhuvanalakshmi, G.; Arfuso, F.; Rajan, G.; Millward, M.; Dharmarajan, A. Cancer stem-like cells from head and neck cancers are chemosensitized by the Wnt antagonist, sFRP4, by inducing apoptosis, decreasing stemness, drug resistance and epithelial to mesenchymal transition. Cancer Gene Ther. 2014, 21, 381–388. [Google Scholar] [CrossRef]
  139. Basak, S.K.; Zinabadi, A.; Wu, A.W.; Venkatesan, N.; Duarte, V.M.; Kang, J.J.; Dalgard, C.L.; Srivastava, M.; Sarkar, F.H.; Wang, M.B.; et al. Liposome encapsulated curcumin-difluorinated (CDF) inhibits the growth of cisplatin resistant head and neck cancer stem cells. Oncotarget 2015, 6, 18504–18517. [Google Scholar] [CrossRef]
  140. Lee, S.H.; Nam, H.J.; Kang, H.J.; Samuels, T.L.; Johnston, N.; Lim, Y.C. Valproic acid suppresses the self-renewal and proliferation of head and neck cancer stem cells. Oncol. Rep. 2015, 34, 2065–2071. [Google Scholar] [CrossRef]
  141. Kulsum, S.; Sudheendra, H.V.; Pandian, R.; Ravindra, D.R.; Siddappa, G.; Chevour, P.; Ramachandran, B.; Sagar, M.; Jayaprakash, A.; Mehta, A. Cancer stem cell mediated acquired chemoresistance in head and neck cancer can be abrogated by aldehyde dehydrogenase 1 A1 inhibition. Mol. Carcinog. 2017, 56, 694–711. [Google Scholar] [CrossRef]
  142. Subramanian, C.; Kovatch, K.J.; Sim, M.W.; Wang, G.; Prince, M.E.; Carey, T.E.; Davis, R.; Blagg, B.S.J.; Cohen, M.S. Novel C-Terminal Heat Shock Protein 90 Inhibitors (KU711 and Ku757) Are Effective in Targeting Head and Neck Squamous Cell Carcinoma Cancer Stem cells. Neoplasia 2017, 19, 1003–1011. [Google Scholar] [CrossRef]
  143. Roy, S.; Roy, S.; Kar, M.; Padhi, S.; Saha, A.; Anuja, K.; Banerjee, B. Role of p38 MAPK in disease relapse and therapeutic resistance by maintenance of cancer stem cells in head and neck squamous cell carcinoma. J. Oral. Pathol. Med. 2018, 47, 492–501. [Google Scholar] [CrossRef]
  144. Roy, S.; Roy, S.; Anuja, K.; Thakur, S.; Akhter, Y.; Padhi, S.; Banerjee, B. p38 Mitogen-activated protein kinase modulates cisplatin resistance in Head and Neck Squamous Cell Carcinoma cells. Arch. Oral. Biol. 2021, 122, 104981. [Google Scholar] [CrossRef] [PubMed]
  145. Roy, S.; Kar, M.; Roy, S.; Padhi, S.; Kumar, A.; Thakur, S.; Akhter, Y.; Gatto, G.; Banerjee, B. Inhibition of CD44 sensitizes cisplatin-resistance and affects Wnt/beta-catenin signaling in HNSCC cells. Int. J. Biol. Macromol. 2020, 149, 501–512. [Google Scholar] [CrossRef] [PubMed]
  146. Keysar, S.B.; Gomes, N.; Miller, B.; Jackson, B.C.; Le, P.N.; Morton, J.J.; Reisinger, J.; Chimed, T.S.; Gomez, K.E.; Nieto, C.; et al. Inhibiting Translation Elongation with SVC112 Suppresses Cancer Stem Cells and Inhibits Growth in Head and Neck Squamous Carcinoma. Cancer Res. 2020, 80, 1183–1198. [Google Scholar] [CrossRef] [PubMed]
  147. Shriwas, O.; Priyadarshini, M.; Samal, S.K.; Rath, R.; Panda, S.; Das Majumdar, S.K.; Muduly, D.K.; Botlagunta, M.; Dash, R. DDX3 modulates cisplatin resistance in OSCC through ALKBH5-mediated m6A-demethylation of FOXM1 and NANOG. Apoptosis 2020, 25, 233–246. [Google Scholar] [CrossRef] [PubMed]
  148. Lu, H.; Yan, C.; Quan, X.X.; Yang, X.; Zhang, J.; Bian, Y.; Chen, Z.; Van Waes, C. CK2 phosphorylates and inhibits TAp73 tumor suppressor function to promote expression of cancer stem cell genes and phenotype in head and neck cancer. Neoplasia 2014, 16, 789–800. [Google Scholar] [CrossRef]
  149. Kerk, S.A.; Finkel, K.A.; Pearson, A.T.; Warner, K.A.; Zhang, Z.; Nor, F.; Wagner, V.P.; Vargas, P.A.; Wicha, M.S.; Hurt, E.M.; et al. 5T4-Targeted Therapy Ablates Cancer Stem Cells and Prevents Recurrence of Head and Neck Squamous Cell Carcinoma. Clin. Cancer Res. 2017, 23, 2516–2527. [Google Scholar] [CrossRef]
  150. Tabor, M.H.; Clay, M.R.; Owen, J.H.; Bradford, C.R.; Carey, T.E.; Wolf, G.T.; Prince, M.E. Head and neck cancer stem cells: The side population. Laryngoscope 2011, 121, 527–533. [Google Scholar] [CrossRef]
  151. Fukusumi, T.; Ishii, H.; Konno, M.; Yasui, T.; Nakahara, S.; Takenaka, Y.; Yamamoto, Y.; Nishikawa, S.; Kano, Y.; Ogawa, H.; et al. CD10 as a novel marker of therapeutic resistance and cancer stem cells in head and neck squamous cell carcinoma. Br. J. Cancer 2014, 111, 506–514. [Google Scholar] [CrossRef]
  152. Larkins, E.; Blumenthal, G.M.; Yuan, W.; He, K.; Sridhara, R.; Subramaniam, S.; Zhao, H.; Liu, C.; Yu, J.; Goldberg, K.B.; et al. FDA Approval Summary: Pembrolizumab for the Treatment of Recurrent or Metastatic Head and Neck Squamous Cell Carcinoma with Disease Progression on or After Platinum-Containing Chemotherapy. Oncologist 2017, 22, 873–878. [Google Scholar] [CrossRef]
  153. Ferris, R.L.; Blumenschein, G., Jr.; Fayette, J.; Guigay, J.; Colevas, A.D.; Licitra, L.; Harrington, K.; Kasper, S.; Vokes, E.E.; Even, C.; et al. Nivolumab for Recurrent Squamous-Cell Carcinoma of the Head and Neck. N. Engl. J. Med. 2016, 375, 1856–1867. [Google Scholar] [CrossRef]
  154. Taylor, B.C.; Balko, J.M. Mechanisms of MHC-I Downregulation and Role in Immunotherapy Response. Front. Immunol. 2022, 13, 844866. [Google Scholar] [CrossRef] [PubMed]
  155. Bayik, D.; Lathia, J.D. Cancer stem cell-immune cell crosstalk in tumour progression. Nat. Rev. Cancer 2021, 21, 526–536. [Google Scholar] [CrossRef] [PubMed]
  156. Qian, X.; Ma, C.; Nie, X.; Lu, J.; Lenarz, M.; Kaufmann, A.M.; Albers, A.E. Biology and immunology of cancer stem(-like) cells in head and neck cancer. Crit. Rev. Oncol. Hematol. 2015, 95, 337–345. [Google Scholar] [CrossRef] [PubMed]
  157. Park, J.C.; Krishnakumar, H.N.; Saladi, S.V. Current and Future Biomarkers for Immune Checkpoint Inhibitors in Head and Neck Squamous Cell Carcinoma. Curr. Oncol. 2022, 29, 4185–4198. [Google Scholar] [CrossRef]
  158. Liao, T.; Kaufmann, A.M.; Qian, X.; Sangvatanakul, V.; Chen, C.; Kube, T.; Zhang, G.; Albers, A.E. Susceptibility to cytotoxic T cell lysis of cancer stem cells derived from cervical and head and neck tumor cell lines. J. Cancer Res. Clin. Oncol. 2013, 139, 159–170. [Google Scholar] [CrossRef]
  159. Tsuchiya, H.; Shiota, G. Immune evasion by cancer stem cells. Regen. Ther. 2021, 17, 20–33. [Google Scholar] [CrossRef]
  160. Ramgolam, K.; Lauriol, J.; Lalou, C.; Lauden, L.; Michel, L.; de la Grange, P.; Khatib, A.M.; Aoudjit, F.; Charron, D.; Alcaide-Loridan, C.; et al. Melanoma spheroids grown under neural crest cell conditions are highly plastic migratory/invasive tumor cells endowed with immunomodulator function. PLoS ONE 2011, 6, e18784. [Google Scholar] [CrossRef]
  161. Iovino, F.; Meraviglia, S.; Spina, M.; Orlando, V.; Saladino, V.; Dieli, F.; Stassi, G.; Todaro, M. Immunotherapy targeting colon cancer stem cells. Immunotherapy 2011, 3, 97–106. [Google Scholar] [CrossRef]
  162. Agudo, J.; Park, E.S.; Rose, S.A.; Alibo, E.; Sweeney, R.; Dhainaut, M.; Kobayashi, K.S.; Sachidanandam, R.; Baccarini, A.; Merad, M.; et al. Quiescent Tissue Stem Cells Evade Immune Surveillance. Immunity 2018, 48, 271–285.e5. [Google Scholar] [CrossRef]
  163. Chikamatsu, K.; Takahashi, G.; Sakakura, K.; Ferrone, S.; Masuyama, K. Immunoregulatory properties of CD44+ cancer stem-like cells in squamous cell carcinoma of the head and neck. Head Neck 2011, 33, 208–215. [Google Scholar] [CrossRef]
  164. Wang, C.; Li, Y.; Jia, L.; Kim, J.K.; Li, J.; Deng, P.; Zhang, W.; Krebsbach, P.H.; Wang, C.Y. CD276 expression enables squamous cell carcinoma stem cells to evade immune surveillance. Cell Stem Cell 2021, 28, 1597–1613.e7. [Google Scholar] [CrossRef] [PubMed]
  165. Getu, A.A.; Tigabu, A.; Zhou, M.; Lu, J.; Fodstad, O.; Tan, M. New frontiers in immune checkpoint B7-H3 (CD276) research and drug development. Mol. Cancer 2023, 22, 43. [Google Scholar] [CrossRef] [PubMed]
  166. Nitsch, S.M.; Pries, R.; Wollenberg, B. Head and neck cancer triggers increased IL-6 production of CD34+ stem cells from human cord blood. In Vivo 2007, 21, 493–498. [Google Scholar] [PubMed]
  167. Xiao, M.; Zhang, X.; Zhang, D.; Deng, S.; Zheng, A.; Du, F.; Shen, J.; Yue, L.; Yi, T.; Xiao, Z.; et al. Complex interaction and heterogeneity among cancer stem cells in head and neck squamous cell carcinoma revealed by single-cell sequencing. Front. Immunol. 2022, 13, 1050951. [Google Scholar] [CrossRef] [PubMed]
  168. Vipparthi, K.; Hari, K.; Chakraborty, P.; Ghosh, S.; Patel, A.K.; Ghosh, A.; Biswas, N.K.; Sharan, R.; Arun, P.; Jolly, M.K.; et al. Emergence of hybrid states of stem-like cancer cells correlates with poor prognosis in oral cancer. iScience 2022, 25, 104317. [Google Scholar] [CrossRef]
  169. Rahman, M.; Deleyrolle, L.; Vedam-Mai, V.; Azari, H.; Abd-El-Barr, M.; Reynolds, B.A. The cancer stem cell hypothesis: Failures and pitfalls. Neurosurgery 2011, 68, 531–545, discussion 545. [Google Scholar] [CrossRef]
  170. Chitturi Suryaprakash, R.T.; Kujan, O.; Shearston, K.; Farah, C.S. Three-Dimensional Cell Culture Models to Investigate Oral Carcinogenesis: A Scoping Review. Int. J. Mol. Sci. 2020, 21, 9520. [Google Scholar] [CrossRef]
  171. Jensen, C.; Teng, Y. Is It Time to Start Transitioning From 2D to 3D Cell Culture? Front. Mol. Biosci. 2020, 7, 33. [Google Scholar] [CrossRef]
  172. Driehuis, E.; Kolders, S.; Spelier, S.; Lohmussaar, K.; Willems, S.M.; Devriese, L.A.; de Bree, R.; de Ruiter, E.J.; Korving, J.; Begthel, H.; et al. Oral Mucosal Organoids as a Potential Platform for Personalized Cancer Therapy. Cancer Discov. 2019, 9, 852–871. [Google Scholar] [CrossRef]
  173. Millen, R.; De Kort, W.W.B.; Koomen, M.; van Son, G.J.F.; Gobits, R.; Penning de Vries, B.; Begthel, H.; Zandvliet, M.; Doornaert, P.; Raaijmakers, C.P.J.; et al. Patient-derived head and neck cancer organoids allow treatment stratification and serve as a tool for biomarker validation and identification. Med 2023, 4, 290–310.e12. [Google Scholar] [CrossRef]
  174. Close, D.A.; Camarco, D.P.; Shan, F.; Kochanek, S.J.; Johnston, P.A. The Generation of Three-Dimensional Head and Neck Cancer Models for Drug Discovery in 384-Well Ultra-Low Attachment Microplates. Methods Mol. Biol. 2018, 1683, 355–369. [Google Scholar] [CrossRef] [PubMed]
  175. Shan, F.; Close, D.A.; Camarco, D.P.; Johnston, P.A. High-Content Screening Comparison of Cancer Drug Accumulation and Distribution in Two-Dimensional and Three-Dimensional Culture Models of Head and Neck Cancer. ASSAY. Drug Dev. Technol. 2018, 16, 27–50. [Google Scholar] [CrossRef] [PubMed]
  176. Gorican, L.; Gole, B.; Potocnik, U. Head and Neck Cancer Stem Cell-Enriched Spheroid Model for Anticancer Compound Screening. Cells 2020, 9, 1707. [Google Scholar] [CrossRef] [PubMed]
  177. Miao, Y.; Wang, P.; Huang, J.; Qi, X.; Liang, Y.; Zhao, W.; Wang, H.; Lyu, J.; Zhu, H. Metabolomics, Transcriptome and Single-Cell RNA Sequencing Analysis of the Metabolic Heterogeneity between Oral Cancer Stem Cells and Differentiated Cancer Cells. Cancers 2024, 16, 237. [Google Scholar] [CrossRef] [PubMed]
  178. Siquara da Rocha, L.O.; Souza, B.S.F.; Coletta, R.D.; Lambert, D.W.; Gurgel Rocha, C.A. Mapping Cell-in-Cell Structures in Oral Squamous Cell Carcinoma. Cells 2023, 12, 2418. [Google Scholar] [CrossRef]
  179. Tanaka, N.; Osman, A.A.; Takahashi, Y.; Lindemann, A.; Patel, A.A.; Zhao, M.; Takahashi, H.; Myers, J.N. Head and neck cancer organoids established by modification of the CTOS method can be used to predict in vivo drug sensitivity. Oral. Oncol. 2018, 87, 49–57. [Google Scholar] [CrossRef]
  180. Zhao, H.; Hu, C.Y.; Chen, W.M.; Huang, P. Lactate Promotes Cancer Stem-like Property of Oral Sequamous Cell Carcinoma. Curr. Med. Sci. 2019, 39, 403–409. [Google Scholar] [CrossRef]
  181. Ikeda-Motonakano, R.; Hirabayashi-Nishimuta, F.; Yada, N.; Yamasaki, R.; Nagai-Yoshioka, Y.; Usui, M.; Nakazawa, K.; Yoshiga, D.; Yoshioka, I.; Ariyoshi, W. Fabrication of a Three-Dimensional Spheroid Culture System for Oral Squamous Cell Carcinomas Using a Microfabricated Device. Cancers 2023, 15, 5162. [Google Scholar] [CrossRef]
  182. Chen, H.H.; Nguyen, T.V.; Shih, Y.H.; Chang, K.C.; Chiu, K.C.; Hsia, S.M.; Fuh, L.J.; Shieh, T.M. Combining microfluidic chip and low-attachment culture devices to isolate oral cancer stem cells. J. Dent. Sci. 2024, 19, 560–567. [Google Scholar] [CrossRef]
Figure 1. Evolution of cancer. (A) The cancer stem cell (CSC) theory of cancer evolution proposes the existence of a hierarchy in the heterogeneous cancer cell population in which cancer stem cells (CSCs) occupy the apex of the hierarchy. Through symmetrical and asymmetrical division in each cell cycle, CSCs give rise to other CSCs (through self-renewal) and non-CSCs, contributing to the heterogeneity in the cancer cell population. (B) In Nowell’s clonal theory of cancer evolution, all cancer cells are equally probable in acquiring more mutations, and each cell will expand clonally to form a heterogeneous cancer cell population.
Figure 1. Evolution of cancer. (A) The cancer stem cell (CSC) theory of cancer evolution proposes the existence of a hierarchy in the heterogeneous cancer cell population in which cancer stem cells (CSCs) occupy the apex of the hierarchy. Through symmetrical and asymmetrical division in each cell cycle, CSCs give rise to other CSCs (through self-renewal) and non-CSCs, contributing to the heterogeneity in the cancer cell population. (B) In Nowell’s clonal theory of cancer evolution, all cancer cells are equally probable in acquiring more mutations, and each cell will expand clonally to form a heterogeneous cancer cell population.
Biomedicines 12 02111 g001
Figure 2. CSCs in OSCC. Cancer cells harvested from primary OSCC or cancer cell lines consist of a heterogeneous cancer population of CSCs and non-CSCs. CSCs are identified and then sorted by using cell surface markers. CD133 and CD44 are commonly used cell surface markers to identify and sort OSCC CSCs using fluorescence-assisted cell sorting (FASC) or magnetic beads in magnetic sorting for further characterization of CSCs. When cultured as a monolayer, CSCs grow as tightly packed cell colonies (holoclones). In non-adherent culture, CSCs readily form tumor spheres. Most importantly, CSCs form tumors rapidly when xenotransplanted into an animal model and resist chemotherapy regimens.
Figure 2. CSCs in OSCC. Cancer cells harvested from primary OSCC or cancer cell lines consist of a heterogeneous cancer population of CSCs and non-CSCs. CSCs are identified and then sorted by using cell surface markers. CD133 and CD44 are commonly used cell surface markers to identify and sort OSCC CSCs using fluorescence-assisted cell sorting (FASC) or magnetic beads in magnetic sorting for further characterization of CSCs. When cultured as a monolayer, CSCs grow as tightly packed cell colonies (holoclones). In non-adherent culture, CSCs readily form tumor spheres. Most importantly, CSCs form tumors rapidly when xenotransplanted into an animal model and resist chemotherapy regimens.
Biomedicines 12 02111 g002
Table 1. Summary of in vivo evidence of CSCs in HNSCC.
Table 1. Summary of in vivo evidence of CSCs in HNSCC.
ReferencesMarkersCell LineMice StrainSite
[44]CD44+Primary HNSCC tumorsNOD/SCID or Rag2γDKOS.C. (Right and left flank)
[45]CD133+Primary OSCC tumors or cell lines BALB/c nude S.C. (Back region)
[46]CD44+CD24 ALDH+Primary HNSCC tumorsSCIDS.C. (Neck region)
[47]ALDHhighPrimary HNSCC tumorsNOD/SCIDS.C. (Site N/A)
[48]CD44+ ALDH+Primary HNSCC tumorsBALB/c nude S.C. (Neck region)
[49]CD133+OSCC cell linesAthymic NCr-nu/nuS.C. (Right and left midabdominal area)
[50]CD44highEpCAMhighOSCC cell linesNOD/SCIDOrthotopic (Tongue)
[51]CD44+CD133+OSCC cell linesBALB/C nudeS.C. (Roots of mouse limb)
[52]CD24+CD44+HNSCC cell linesAthymic nude immunodeficientS.C. (Dorsal flank)
[53]Cisplatin-resistant cellsHNSCC cell linesSCIDS.C. (Flank)
[54]Side populationOSCC cell linesBALB/cS.C. (Right and left midabdominal area)
[55]SpheresHNSCC and OSCC cell linesBALB/c nu/nu S.C. (Flank)
[56]CD44high/EpCAMlow/CD24+OSCC cell lineNOD/SCIDOrthotopic (Tongue)
[57]CD44+CD66Primary HNSCC tumorsRag-2/γc−/−S.C. (Neck region)
[58]CD44+ALDH1+OSCC cell linesBALB/c nu/nu S.C. (Right front axilla)
[59]CD44+CD24low/−OSCC cell linesAthymic NCr-nu/nuOrthotopic (Tongue)
[60]BMI1+Primary 4NQO-induced HNSCC mouse model C57BL/6, NOD/SCIDOrthotopic (Tongue)
[61]CD44+CD271HNSCC cell linesNU/NU nude (Crl:NU-Foxn1nu) miceOrthotopic (Tongue)
[62]CD133+Primary OSCC tumorsBALB/c miceS.C. (Exact site N/A)
[63]CD44+CD133+Primary OSCC tumorsBALB/c nudeS.C. (Exact site N/A)
[64]CD44highEpCAMhighOSCC cell linesNOD/SCID Orthotopic (Tongue)
[65]CD44highCD24lowOSCC cell linesNOD/SCID miceS.C. (Dorsal flank)
[66]CD44highCD326
CD44lowCD326high
CD44lowCD326
OSCC cell linesBALB/c nudeOrthotopic (Tongue)
[67]CD44+OSCC cell linesBALB/c nudeS.C. (Forelimb)
[68]Cisplatin-resistant cellsOSCC cell linesBALB/c nudeS.C. (Right and left back region)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Acharya, S.K.; Shai, S.; Choon, Y.F.; Gunardi, I.; Hartanto, F.K.; Kadir, K.; Roychoudhury, A.; Amtha, R.; Vincent-Chong, V.K. Cancer Stem Cells in Oral Squamous Cell Carcinoma: A Narrative Review on Experimental Characteristics and Methodological Challenges. Biomedicines 2024, 12, 2111. https://doi.org/10.3390/biomedicines12092111

AMA Style

Acharya SK, Shai S, Choon YF, Gunardi I, Hartanto FK, Kadir K, Roychoudhury A, Amtha R, Vincent-Chong VK. Cancer Stem Cells in Oral Squamous Cell Carcinoma: A Narrative Review on Experimental Characteristics and Methodological Challenges. Biomedicines. 2024; 12(9):2111. https://doi.org/10.3390/biomedicines12092111

Chicago/Turabian Style

Acharya, Surendra Kumar, Saptarsi Shai, Yee Fan Choon, Indrayadi Gunardi, Firstine Kelsi Hartanto, Kathreena Kadir, Ajoy Roychoudhury, Rahmi Amtha, and Vui King Vincent-Chong. 2024. "Cancer Stem Cells in Oral Squamous Cell Carcinoma: A Narrative Review on Experimental Characteristics and Methodological Challenges" Biomedicines 12, no. 9: 2111. https://doi.org/10.3390/biomedicines12092111

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

Article Metrics

Back to TopTop