Next Article in Journal
Migraine Pharmacological Treatment and Cognitive Impairment: Risks and Benefits
Next Article in Special Issue
Novel Thieno [2,3-b]pyridine Anticancer Compound Lowers Cancer Stem Cell Fraction Inducing Shift of Lipid to Glucose Metabolism
Previous Article in Journal
Cold-Induced Nuclear Import of CBF4 Regulates Freezing Tolerance
Previous Article in Special Issue
A Comprehensive Characterization of Stemness in Cell Lines and Primary Cells of Pancreatic Ductal Adenocarcinoma
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Chemoresistance-Related Stem Cell Signaling in Osteosarcoma and Its Plausible Contribution to Poor Therapeutic Response: A Discussion That Still Matters

by
Sara R. Martins-Neves
1,2,
Gabriela Sampaio-Ribeiro
1,2,3,4 and
Célia M. F. Gomes
1,2,3,4,*
1
iCBR-Coimbra Institute for Clinical and Biomedical Research, Faculty of Medicine, University of Coimbra, 3000-548 Coimbra, Portugal
2
Institute of Pharmacology and Experimental Therapeutics, Faculty of Medicine, University of Coimbra, 3000-548 Coimbra, Portugal
3
CIBB-Center for Innovative Biomedicine and Biotechnology, University of Coimbra, 3000-548 Coimbra, Portugal
4
CACC-Clinical Academic Center of Coimbra, 3000-075 Coimbra, Portugal
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(19), 11416; https://doi.org/10.3390/ijms231911416
Submission received: 20 August 2022 / Revised: 22 September 2022 / Accepted: 24 September 2022 / Published: 27 September 2022
(This article belongs to the Special Issue Cellular and Molecular Biology of Cancer Stem Cells)

Abstract

:
Osteosarcoma is amongst the most prevalent bone sarcomas and majorly afflicts children and adolescents. Therapeutic regimens based on the triad of doxorubicin, cisplatin and methotrexate have been used as the state-of-the-art approach to clinical treatment and management, with no significant improvements in the general outcomes since their inception in the early 1970s. This fact raises the following problematic questions: Why do some patients still relapse despite an initial good response to therapy? Why do nearly 30% of patients not respond to neoadjuvant therapies? Does residual persistent disease contribute to relapses and possible metastatic dissemination? Accumulating evidence suggests that chemoresistant cancer stem cells may be the major culprits contributing to those challenging clinical outcomes. Herein, we revisit the maneuvers that cancer stem cells devise for eluding cell killing by the classic cytotoxic therapies used in osteosarcoma, highlighting studies that demonstrate the complex crosstalk of signaling pathways that cancer stem cells can recruit to become chemoresistant.

1. Introduction

Osteosarcoma is very aggressive bone tumor with a dismal prognosis for poor responders to therapy and for metastasis-presenting patients [1]. Considering the histological observations of the existence of undifferentiated cellular subsets alongside with cells resembling more differentiated phenotypes, some studies propose a stem cell origin for osteosarcoma, including the possibility of tumor-promoting genetic hints that provide a hurdle to mesenchymal stem cells (MSCs) [2]. In fact, regardless of its origin, a plethora of evidence demonstrates that cancer stem cells (CSCs) exist within osteosarcomas and have a role as promising therapeutic targets [3].
Despite the difficulties in the identification of a consistent molecular phenotype for osteosarcoma CSCs, knowing the mechanisms by which stemness networks persist among tumor cells is particularly important for the recognition of new therapeutic targets [4]. Unpuzzling the signaling pathway players that determine chemoresistance and how they are molecularly intertwined with stemness signaling offers the possibility of effectively enhancing chemosensitivity for osteosarcoma. Therefore, in this review, we focus on the key mechanisms involved in resistance to chemotherapy in osteosarcoma CSCs, namely, detoxification systems (drug efflux transport and ALDH), survival-related pathways (ERK, AKT), adaptive metabolic routes, altered cell cycle and DNA repair, enhanced apoptosis and modulation of the tumor microenvironment (hypoxia, inflammation), by giving appropriate examples. We also discuss a few points regarding some of the immunotherapeutic options explored for osteosarcoma treatment, with a tentative focus on the few studies that specifically regard CSCs.

2. Epidemiology of Osteosarcoma and Rationale to Maintain the Discussion about Its Resistance to Chemotherapy

Osteosarcoma is the commonest type of bone cancer and a highly aggressive osteoid-depositing tumor affecting mainly children and adolescents [5]. Nearly 20% of patients have metastatic disease at presentation, mostly located in the lungs [6]. The treatment for osteosarcoma follows a multidisciplinary approach (Figure 1), with the essential standard therapy including (a) pre-operative or neoadjuvant chemotherapy, consisting of high-dose methotrexate (MTX), doxorubicin (DOX) and cisplatin (CIS); (b) local surgical resection; and (c) post-operative or adjuvant chemotherapy, administered when the degree of tumor necrosis after pre-operative chemotherapy is superior to 90% [7]. This chemotherapeutic regimen has improved the cure rate and long-term disease-free survival percentage for osteosarcoma patients, with localized lesions ranging from 5% to 20% in the pre-chemotherapy era to the 65% to 75% range observed nowadays [7,8,9]. Unfortunately, survival for patients with metastasis at initial diagnosis is still only 17–34% [10].
Pre-operative chemotherapy enables the early treatment of micrometastatic disease and facilitates the surgical resection by shrinking the tumor mass and decreasing vascularity. Limb-salvage surgery should occur after a defined time interval, with no advantage being observed with immediate surgery [13], aiming to preserve a functioning limb without increasing the risk of post-operative complications to the patient. Post-operative chemotherapy after surgical resection is normally performed in order to minimize the likelihood of local recurrence. It is clinically well-established that complete surgical resection of both the primary tumor and metastatic nodules is essential for survival [6,10].
Despite a good response to pre-operative therapy, local recurrence and metastatic disease occurs in 25–50% of surgically-treated patients without evidence of metastasis at diagnosis. This is most probably attributable to poor response to standard therapy and constitutes the major clinical problem: preventing the curing of high-grade osteosarcoma patients. Patients who develop metastatic disease after surgical resection have limited options; however, they frequently receive additional cytotoxic drugs to the standard drug regimen, such as zoledronate [14], gemcitabine and docetaxel [15], ifosfamide [16], regorafenib [17] or lenvatinib [18]. Nevertheless, in general, in these patients, neither the intensification of dose regimens [19,20,21,22] nor the addition of new drugs [23,24] has significantly modified their clinical outcomes.
Notwithstanding the exhaustive and incredible progress in the clinical research of new therapeutic targets and compounds in recent years, the attempts to increase the proportion of tumor necrosis by means of dosage intensification of pre-operative chemotherapy and the addition of new compounds to the classical triplet drug regimen (MTX, DOX and CIS), the cure rate and long-term disease-free survival percentage of osteosarcoma patients with localized disease has stagnated since the 1970s, in the 60% to 70% range, as previously indicated [25,26]. The cases of disease recurrence and the development of drug resistance in osteosarcoma provide a rationale for the exploration of both old and new insights that might contribute to a better identification of the precise mechanisms of resistance operating in this tumor and the pathways responsible for recurrence after a favorable response to therapy (Figure 2).

3. Cancer Stem Cells Contribute to Intra-Tumor Cellular Heterogeneity

Tumor cell heterogeneity is a well-recognized characteristic of osteosarcoma, constituting one of the main causes of treatment failure and may have at least two underlying causative tumor development models—the stochastic clonal evolution and the hierarchical cancer stem cell (CSC). The existence of different osteosarcoma clones is linked to dynamic genetic and epigenetic events, which subsequently determines differential patterns of cellular tumorigenicity [27,28]. However, analysis of histopathological specimens reveals that, in some cases, the tumors are organized in a hierarchical manner, with a leading CSC being the generator of phenotypic and functional heterogeneity. In fact, cells with stemness-related features have been found in several tumor types, including bone sarcomas [29,30], and are associated with treatment resistance, tumor relapses and recurrences, and disseminated metastatic disease [31]. In recent years, several studies suggest that the principles of both the clonal and the CSC models can be conciliated to better describe tumor heterogeneity, since virtually every cell within the tumor may convert from a non-tumorigenic to a tumorigenic state and is able to switch phenotypically into a stem-like cell. These transitions are sustained by the appropriate oncogenic insults or other microenvironmental stimuli, demonstrating the CSC’s intrinsic cell plasticity [32,33,34,35].

General Overview about the Identification of CSCs in Osteosarcoma

The identification of osteosarcoma CSCs has gained increasing attention over the last three decades. Owing to CSCs’ rarity and the absence of established specific markers for osteosarcoma, the characterization of CSCs has been done mostly based on functional criteria (Table 1), namely, with sphere-formation, Aldefluor™, side-population and surface marker flow cytometry-based assays. These methods do not mandatorily recognize a unique, exclusive set of CSCs, but do uncover the heterogeneous nature and phenotypic plasticity of osteosarcoma CSC sub-populations, with no methodology being better or more adequate than the other to identify those CSC subsets. In fact, based on our own previous experience with these techniques, when possible, combining, e.g., two of these functional assays may contribute to refine more bona-fide CSC subsets within the cell samples analyzed.
Cancer stem cell markers are used as permanent labels of stemness during self-renewal. Virtually every protein can serve as CSC biomarkers; however, not all are able to specifically identify a CSC population, as they might differ according to the source of CSCs and might change as the tumor evolves. Moreover, stem cells of different tissues are not all identical and the dissimilarities concerning, e.g., location, self-renewal and differentiation capabilities are often reflected by specific combinations of phenotypic markers.
Different combinations of markers constitute the basis for distinguishing a certain stem cell type from another one. Moreover, CSCs’ markers expression might be influenced by intrinsic tumor microenvironmental stimuli or modulated by therapeutics, as we previously discussed [74]. There is consensus that CSCs express many of the markers commonly used to identify normal stem cells (either embryonic or adult somatic stem cells) [75]. In general, these cell surface markers are very advantageous to identify and isolate CSC populations using the appropriate cell-sorting technologies and protocols.
Some markers are inclusive—that is, expressed by diverse types of CSCs—or exclusive, with the potential to be exploited for therapeutic targeting or as biomarkers. The first specific CSC markers were identified in hematological tumors, CD34+/CD38- in acute myeloid leukemia [76], while breast cancer was the first solid tumor for which a CD44+/CD133+/ALDH1+ phenotype was pointed as specific for breast CSCs [77,78].
Comprehensible descriptions of osteosarcoma CSC markers have been revised previously [79,80]. Notably, none of the herein- and therein-mentioned markers are unique to osteosarcoma CSCs. Indeed, CD surface markers presented in Table 1 either require a broader validation in osteosarcoma tissue cohorts or they have been detected also in other CSC types; for example, CD44/CD271 in melanoma [81] and CD117 in ovarian cancer [82]. Moreover, the identification of CSC surface markers has been more elusive in mesenchymal tumors than in tumors originating from other tissue types, in part due to the lack of agreement on the markers that unequivocally identify mesenchymal progenitor cells [83], the plausible osteosarcoma cells of origin [84]. Nevertheless, research towards specific molecular markers for osteosarcoma CSCs that may be therapeutically targeted is ongoing and justifies further investigation.

4. Mechanisms of CSC Resistance to Conventional Therapies

Several studies using cancer cell lines and pathological tissue specimens indicate that a complex network of mechanisms play a determinant role in cancer cell resistance to chemotherapeutics. The knowledge of the different chemoresistance-related pathways in osteosarcoma CSCs is then vital for the development of novel molecular targets that may enhance their chemosensitivity. Some of the mechanisms explored in this review seem exacerbated in osteosarcoma CSCs, which can undertake a quiescent state and activate signaling cascades, such as drug efflux transport or aldehyde dehydrogenase activity, and that may accompany apoptosis evasion, enhanced survival and DNA repair activation. Moreover, hypoxic and pro-inflammatory microenvironments surrounding CSCs also constitute facilitators of resistance to conventional therapies that target rapidly proliferating cells and induce DNA damage (Figure 3).

4.1. Chemoresistance Due to Detoxifying Mechanisms—Drug Efflux Transporters and Aldehyde Dehydrogenase Activity

4.1.1. Drug Efflux Transporters

Cancer cells can become resistant to multiple cytotoxic drugs through increased efflux of the drug from the cell, via the so-called ABC transporters. Overexpression of these molecular membrane pumps contributes to multidrug resistance, as they export a wide variety of drugs, including doxorubicin, cisplatin and methotrexate, which are clinically used in osteosarcoma. Multidrug resistance (MDR) may cause relapses to therapy, which together with metastatic dissemination is still a major contributor to death by cancer [85]. High expression of ABC transporters has been detected in several CSC types and correlated with resistance to diverse chemotherapeutics [86]. The study of the impact of ABC transporters in bone tumors has regained attention in recent years, since they are expressed by both normal tumor cells and CSCs, with ABCB1 and ABCG2 being of special interest. In fact, particularly ABCB1 expression has been positively correlated to the existence of metachronous lung metastases and a propensity to metastatic formation in non-responsive patients, as previously reviewed [87].
The expression and function of several ABC transporters seems to be exacerbated in osteosarcoma CSCs and linked to chemoresistance. ABGC2-high CSCs selected with 3-aminobenzamide had increased drug efflux ability [36]. We showed that the ABC inhibitor verapamil enhanced the cellular uptake of doxorubicin by P-glycoprotein- and BCRP-overexpressing CSCs [40]. This was encompassed by Bak upregulation and Bcl-2 suppression favoring CSC apoptosis [88]. We also found a positive correlation between ABCG2 expression and a side-population cell subset in nine cell lines [41]. Moreover, doxorubicin induced ABCG2 and ABCB1 expression paralleled by the activation of pluripotency markers and Wnt/β-catenin in bulk cells [50], while the inhibition of tankyrase with IWR-1 in CSC-enriched spheres led to downregulation of BCRP and P-glycoprotein, accompanied by a sensitization to doxorubicin-induced apoptosis [89]. Other studies showed that cisplatin-resistant CD133+ MG-63 cells express high levels of P-glycoprotein [90], and the same was observed in doxorubicin-resistant U2OS and MG-63 spheres [91]. Moreover, elevated ABCG1 and P-glycoprotein expression was found in doxorubicin-selected CSCs and in progeny derived from a single HOS cell, mediating resistance to doxorubicin and other drugs [92]. In 3D cultures of ABCB1-high/ABCA1-low chemo-immune-resistant cells, expression of the doxorubicin effluxer ABCB1 was upregulated by the Ras/ERK1/2/HIF-1α signaling axis, which suggests the existence of pathway crosstalk to reinforce an already chemoresistant phenotype in osteosarcoma cells [93]. ABCG2 transcriptional activity was also suppressed by the transcriptional regulator SMAR1, which increased the ABCG2 deacetylation level via HDAC2 and also decreased the ALDH activity [94]. miRNA-221 also seems to increase P-glycoprotein expression via the Stat3 pathway and promoting doxorubicin-resistance [95].
The modulation of ABC transporters activity to circumvent chemoresistance in osteosarcoma has been difficult to translate into successful clinical achievements, with controversial data being reported regarding ABCs as prognostic factors or significantly related to outcomes, as reviewed before [96]. However, considering that the first-line chemotherapeutics that are still indispensably used in osteosarcoma (doxorubicin and cisplatin) are substrates of several drug efflux pumps, revisiting their role as major contributors to MDR [97], exploring their pharmacogenetic and pharmacogenomic association with osteosarcoma survival and response to therapy [98] and admitting that ABCs may also be detrimental to osteosarcoma CSC survival, may altogether refine and renovate their clinical applications.

4.1.2. Aldehyde Dehydrogenase Activity

A plethora of studies have investigated the role of ALDH expression and activity in diverse tumor types and revealed that expression of ALDHs is involved in disease progression and metastatic dissemination [99], including osteosarcoma. ALDHs are actively involved in the chemoresistance of CSCs and their expression generally correlates with a poor prognosis [100]. ALDH enzymes mediate the conversion of xenobiotic and intracellular aldehydes, such as drugs, ethanol and vitamins, into their less harmful corresponding carboxylic acids, thereby acting as important mediators of defense against cytotoxic compounds that can induce DNA damage, inactivation of enzymes and even cell death. ALDHs are responsible for the metabolic regulation of retinoic acid and ROS, and this seems to underlie the functional roles of ALDHs in CSCs [101].
ALDH1 activity may be modulated by DKK-1, a Wnt antagonist, and ALDH1 can be involved in osteosarcoma cancer cell survival and resistance to chemotherapy. In this study, transcriptional activation of ALDH1 was dependent on the activation of the non-canonical Jun-mediated Wnt pathway, suggesting that signaling pathways other than those controlling self-renewal (e.g., Wnt/β-catenin signaling) can also participate in the modulation of ALDH activity [49]. Another study associated the resistance of Saos-2 and U2OS osteosarcoma cells to doxorubicin with activation of ALDH1/CD133-positive cells. This resistance phenotype was inhibited by forced expression of miR-143, which suggests that it may play a role in tumor suppression in osteosarcoma by counteracting stemness properties such as ALDH expression [61]. Retinal treatment preferentially affected ALDH-high cells by decreasing their proliferation, invasion capacity, and resistance to oxidative stress. These effects were more pronounced in highly metastatic osteosarcoma cells, accompanied by altered expression of metastasis-related genes and downregulation of Notch signaling markers [51].
Our previous experimental studies indicate that established osteosarcoma cell lines possess an Aldefluor-positive cell fraction that are SOX2-positive but KLF4-negative, and further enriched in CSCs isolated from spherical colonies [41]. Exposure of therapy-naïve cell lines to conventional drugs increased ALDH signaling as assessed by Aldefluor activity and ALDH1A1, ALDH2 and ALDH7A1 mRNA expression [50], while Wnt/β-catenin inhibition with the tankyrase inhibitor IWR-1 diminished the percentage of ALDH-positive cells and the previously mentioned ALDH isozymes [89]. Other authors have shown that miR-26a is downregulated in ALDH-positive ZOS and 143B cells, while its overexpression reduced ALDH activity via inhibition of another stem cell pathway, Jagged1/Notch [102].
Activity of ALDH can be inhibited with the FDA-approved drug disulfiram, which was used by the Weiss group to show its inhibitory effects in pulmonary metastasis formation in an immunocompetent Balb/c orthotopic mouse model, a result that was statistically equivalent to doxorubicin treatment [103]. This offers an alternative treatment route for osteosarcoma, in line with recent attempts of drug repurposing strategies based on gene expression signatures [104] and other in vitro modeling [105]. This was endeavored also for other drugs and tumor types [106], including the specific targeting of CSCs [107], and substantiates the possibility to refine the doses for the state-of-the-art approved drugs for osteosarcoma while decreasing their side-effects.
More recently, an elegant study using patient-derived xenograft models demonstrated that ALDH1-high xenografts with enhanced metastatic potential in vivo were sensitive to the β-catenin/transducin β-like protein 1 inhibitor tegavivint [52]. Moreover, it has been shown that ALDH activity was increased by the lncRNA THOR through enhanced SOX9 expression [108]. ALDH-positive cells were also sensitive to the apoptotic effects induced by the cell death ligand TRAIL and by leptomycin B [109]. Other studies also suggest that microRNAs may negatively regulate the expression of ALDH family members, such as miR-487b-3p and ALDH1A3, in vitro and in clinical samples [110], and miR-761 and ALDH1B1, in vitro and in vivo [111], which altogether substantiates the complex participation of ALDHs in the phenotypic behavior of osteosarcoma cells.

4.2. Chemoresistance Due to Enhanced Activity of ERK and AKT Survival-Related Signaling Pathways

The EGFR-Ras-Raf-MEK-ERK signaling network and the PI3K/PTEN/AKT signaling pathway controls cell growth and regulates cell survival, cell differentiation and apoptosis, thus being considered important targets for cancer therapy. Deregulation and oncogenic activation of these survival-related pathways is thus of utmost importance for CSCs and has been linked to stem cell features in other tumors [112]. Therefore, it is not surprising that the ERK and AKT pathways are upregulated in osteosarcoma and involved in tumorigenic features such as apoptosis resistance, in vivo tumorigenicity and EMT [113,114]. Some previous studies also pinpoint these two pathways as therapeutic targets [115,116]. Molecularly, the ERK pathway also seems to be involved in EMT and metastasis formation [117] and activated by the downregulation of p16 protein [118].
Several recent studies linked the activation of ERK and AKT pathways to stemness features in osteosarcoma. It is noteworthy that activation and modulation of these survival-related pathways strongly correlates with pluripotency-related transcriptional activity, depending also on the signaling cascades involved in cellular self-renewal, such as Wnt and Hedgehog, and correlates of EMT; they are also regulated by diverse types of microRNAs.
ERK is involved in the regulation of the Warburg effect via the physiological pathway regulator SLIT2/ROBO1 axis [119] and in cell proliferation and migration induced by the Wnt ligand WNT5A [120] and through Notch-induced ERK phosphorylation [121]. Other reports indicate that EMT in osteosarcoma is also regulated by ERK signaling [117,122]. ERK signaling also seems to participate in the acquisition of stemness features in osteosarcoma (expression of CD24, CD90, CD133, Nanog, SOX2, Oct4) mediated by miR-155; moreover, ERK inhibition with U0126 repressed expression of those markers [123]. Recently, Shimizu et al. reported that MEK inhibition with trametinib inhibited the cell cycle and induced apoptosis in non-adherent-growing U2OS cells. Moreover, trametinib decreased the size of primary tumors and circulating tumor cells in an in vivo mouse model [124].
The oncogenic long-non-coding MALAT1 was shown to activate the PI3K/Akt pathway via sponging miR-129-5p and promoted metastasis formation [125]. Activity of this pathway was also inhibited by the natural compound glaucocalyxin A, through the reduction in GLI1 activation and induction of apoptosis [126]. In the MNNG/HOS cell line, inhibition of the osteoblast regulator P2X7 hampered PI3K/Akt/GSK3β/β-catenin signaling and thereby inhibited stemness features and cell migration [127]. Moreover, PI3K/AKT also seems to be regulated by the lncRNA RNA FER1L4, via promoting apoptosis and suppressing EMT [128,129]. Recently, also TGF-β knockdown was associated to PI3K/Akt downregulation, suppressing viability and aggressiveness in osteosarcoma CSCs [130,131].

4.3. Chemoresistance Due to Altered Metabolism

Metabolic alterations occurring in cancer cells have been ascribed as an important hallmark of cancer [132]. In fact, despite that the specific characterization of metabolomics in CSCs has been scarce compared to regular cancer cells, diverse studies suggest that CSCs preferentially use glycolysis and have a declined oxidative phosphorylation activity. However, the location in which CSCs reside at the tumors (e.g., in normoxic vs. hypoxic regions) seems to be detrimental to the type of energy source that CSCs allocate to sustain their proliferative capacity and survival skills [133], which clearly suggests their intrinsic reprogramming capacity to adapt their bioenergetics response depending on the tumor microenvironment. Some studies specifically focused on the metabolic features of osteosarcoma CSCs are summarized below. For instance, Mizushima et al. suggest that indeed osteosarcoma CSCs follow the trend to be more aerobic glycolytic than to use oxidative phosphorylation, partly dependent on the antigen LIN28 [134].
Different types of mass spectrometry-based techniques have been used to explore the nature of osteosarcoma CSC metabolism. Zhong and colleagues used an ultra-high-performance liquid chromatography coupled with tandem Q-Exactive Orbitrap mass spectrometer to characterize the HOS-CSC isolated with the sphere-forming assay [135]. They found a significant upregulation of the metabolomics of several amino acids (alanine, aspartate, glutamate, arginine, proline, glutathione, cysteine and methionine) and a declined mitochondrial function and TCA cycle. Other authors used untargeted liquid chromatography with tandem mass spectrometry (LC–MS/MS) analysis to explore the metabolomics features of the osteosarcoma 143B and MG-63 spheres’ response to MTX [45]. They found that CSCs had alterations in the metabolomics of amino acid, fatty acid, energy and nucleic acid after treatment with MTX and emphasized the utility of mass spectrometry techniques to study the metabolomics features of CSCs. However, these authors did not specifically analyze whether the modulation of those metabolites upon MTX treatment had a pro- or anti-survival advantage to CSCs in that context.
3AB-OS-MG-63 has been used as an in vitro model to study osteosarcoma CSC metabolism. Palorini et al. analyzed the metabolic features of the 3AB-OS-selected CSCs and their parental MG-63 cells and extensively characterized the bioenergetics of CSCs. Compared to parental cells, it was demonstrated that 3AB-OS-CSC depend more on glycolytic metabolism, proliferate less when cultured in glucose-starvation conditions and have increased expression of lactate dehydrogenase A (LDHA). Moreover, their reduced mitochondrial respiration and fragmented mitochondria led the authors to suggest that CSCs possess metabolic similarities to normal stem cells [136]. Subsequent studies using this cell model substantiated the assumption that 3AB-OS-CSC rely more on glycolysis and MG-63 cells rely on glutamine oxidation. Moreover, cisplatin treatment in glutamine-depleted MG-63 resulted in additive inhibitory effects on cell survival, while promoting a pro-survival S-phase arrest in glucose-starved 3AB-OS-CSC. Of special note is the fact that when exposed to cisplatin in glucose-deprived conditions, CSCs switched their metabolic status, reprogramming to be more oxidative than glycolytic and to increase their mitochondrial functions [137]. These results demonstrate the plasticity of metabolic networks in CSCs and that contribute to circumvent cisplatin cytotoxicity.
Osteosarcoma, as a relatively undifferentiated bone sarcoma, retains some mesenchymal features, such as the capacity to respond to adipogenic stimuli. In this context, some authors explored the fatty acid metabolism of osteosarcoma CSCs using the anti-diabetic thiazolidinedione. This PPARγ agonist induced growth arrest and adipogenic differentiation in Sox2-high CSCs, via a mechanism involving cytoplasmic sequestration of the transcription factors SOX2 and YAP [138]. Other authors showed that the inhibition of TNIK (an essential factor for the transactivation of Wnt signal target genes) with NCB-0846 decreased the expression of stem cell genes (SOX2, NANOG, OCT4, MYC) and ALDH activity and also favored lipid biosynthesis, driving osteosarcoma cells’ differentiation into adipocyte-like cells, via induction of PPARγ [139]. These studies unveil the potential to modulate adipogenesis in osteosarcoma cells in order to affect their cell fate determination and promote their vulnerability to previously unexpected drugs that may be repurposed and used as new treatment strategies.
Accumulating evidence suggests that the anti-diabetic drug metformin may be specifically used to target osteosarcoma CSCs and to modulate their metabolic profile. Shang et al. found that metformin inhibited glucose uptake, lactate production and ATP production in HOS CSCs. Resistance of those cells to cisplatin was correlated to overexpression of the pyruvate kinase isoenzyme M2, with its downregulation reversing cisplatin resistance potentiated by metformin treatment [140]. Our own previous studies showed that metformin has deleterious effects in MNNG-HOS CSCs, potentiating low-dose DOX-induced cytotoxicity. Moreover, metformin induced mitochondrial stress by activating the energetic sensor AMPK and increasing [18F]-FDG uptake and lactate production in parental cells but not in quiescent CSCs [141]. Interestingly, Zhao et al. also observed activation of AMPK, which led to the reversal of the mTOR pathway and triggered autophagy. Metformin induced apoptosis in osteosarcoma CSCs through the collapse of the mitochondrial transmembrane potential, decreased ATP synthesis and increased ROS production [142]. Altogether, these studies suggest that osteosarcoma CSCs rely also on mitochondrial respiration for energy production and, when exposed to metformin, are captured in an energetic crisis. These metabolic alterations disturbed the homeostasis of stemness and pluripotency in the osteosarcoma CSCs both in vitro and in vivo, and corroborate an important role for metabolic modulators to chemosensitize this resistant cell subset within the tumors. Actually, more recently, a pre-clinical test using canine osteosarcoma CSCs also showed the capacity of metformin to inhibit mitochondrial function, by decreasing oxygen consumption and ATP production, while sensitizing canine CSCs to irradiation therapy [143]. Further studies are required for a better elucidation of the therapeutic potential of metabolic modulators and the mechanisms involved in the interference with stemness features leading to sensitivity to drugs in osteosarcoma CSCs.

4.4. Chemoresistance Due to Cell Cycle Arrest and Cellular Quiescence

Dormancy or quiescence in tumors may be derived from a defective angiogenic process, in which the expansion of tumor mass is limited due to the inability of recruitment of new and functional blood vessels, generating tumor masses poorly vascularized. This process associates with cell cycle arrest in which the blockade of tumor mass expansion results from the quiescent state of tumor cells. Microenvironmental stimuli or intracellular hits leading to increased cell proliferation may result in escape from dormancy and expansion of the tumor mass, leading to the emergence of clinically relevant disease [144]. Quiescence is molecularly regulated by cell cycle-related signaling [145], and by the tumor suppressors p53 and RB proteins, whose genetic alterations are detrimental in osteosarcomagenesis.
Quiescence is also a common characteristic of drug-resistant cells and has been linked with stem cell traits, since, like normal stem cells, CSCs are quiescent, slow-cycling cells and therefore circumvent the effects of high levels of intracellular reactive ROS, which accounts for their self-renewal capacity and drug resistance [146]. When entering a quiescent, cell cycle-arrested state, CSCs allocate themselves a period of time that may be used to activate and implement one or more of the survival signaling pathways mentioned in this review.
Quiescent cell populations found in osteosarcoma are dependent on angiogenic switches. Almog et al. identified a set of dormancy-associated microRNAs that regulated the phenotypic switch of dormant to fast-proliferating cells, specially miR-190 and miR-580. Moreover, a KHOS-24OS-based mouse model with angiogenic cells overexpressing miR190 remained quiescent during at least 120 days [147]. IGF2 or insulin exposure created an autophagic state of cellular dormancy in highly tumorigenic osteosarcoma cells cultured under serum-free conditions, which then conferred resistance to doxorubicin, cisplatin and methotrexate [148].
We observed that osteosarcoma CSC-enriched spheres were in a slow-proliferative state, being Ki-67-negative [41], along with a low uptake of the glucose analogue [18F]-FDG and accumulation of cells in the G0/G1 phase [40]. Avril et al. also identified OCT4/SOX2/NANOG-positive MNNG-HOS 3D spheres, which were arrested in the G0/G1 phase. Interestingly, the non-diving state of the spheres was not changed by MSC-conditioned media [149]. MiR-329 induced G0/G1 cell cycle arrest and inhibited cell proliferation and tumorigenicity in vivo, effects that seem to be mediated by the DNA repair protein Rad10 [150]. Other authors identified miR-34a, miR-93 and miR-200c as regulators of osteosarcoma dormancy, which could be delivered to fast proliferating Saos-2 and MG-63 cells with the nanocarrier dPG-NH2 to reduce their aggressiveness and migration abilities [151].
More recently, quantitative imaging techniques, such as holographic imaging cytometry, time-lapse microscopy and a contrast-based segmentation algorithm, were fine-tuned and used to monitor the transitions between angiogenic/non-angiogenic tumor states [152] and between diving/non-dividing cells, showing that non-dividing MG-63 cells did not constitute the CSC pool [153]. These techniques allow the observation of key cellular morphological behaviors that cancer cells reshuffle during their dormant and metastatic states, namely, altered cell motility speeds and cell cycle lengths, demonstrating that quiescent cells do not mandatorily represent stem cell subsets.

4.5. Chemoresistance Due to Enhanced DNA Repair

DNA-damaging agents, such as most conventional chemotherapeutics used in osteosarcoma treatment, elicit diverse types of lesions in the DNA molecules (e.g., single- and double-strand breaks). Cancer cells recognize those lesions and bypass the cytotoxic stress induced by anticancer agents, by activating various DNA repair pathways, such as nucleotide excision repair, base excision repair, homologous recombination repair and non-homologous end joining [154]. Several studies unraveled the molecular basis of these DNA repair pathways and provided a rationale for the development of DNA repair inhibitors, which have been demonstrating therapeutic benefits and synergizing with those DNA-damaging drugs [155,156]. The PARP inhibitor olaparib, for treating BRCA1 or BRCA2 mutated tumors, represents the first drug based on this principle [157]. Some studies suggest that a prompt activation of DNA damage response and enhanced DNA repair capacity are key contributors to CSCs’ increased resistance to therapy, due to their extraordinary ability to repair the genetic code compared with their offspring [158].
In osteosarcoma, suppression of Rad51, the key protein of DNA homologous recombination repair, correlated with cell cycle arrest and limited in vivo tumor growth, while also sensitizing osteosarcoma HOS cells to ionizing radiation and chemotherapy [159]. The nucleotide excision repair variants XPD rs13181 and rs1799793 are also related to higher event-free survival in osteosarcoma patients receiving neoadjuvant chemotherapy, and their expression provided a therapeutic advantage from cisplatin chemotherapy, probably by reducing DNA repair activity [160]. Other studies suggest that homologous recombination deficiency associates with MG-63 and ZK-58 cells sensitivity to the PARP inhibitor talazoparib alone or in combination with chemotherapeutics [161].
More recently, expression of the DNA damage repair inducer SIRT6 was associated with shorter overall survival in patients who received adjuvant chemotherapy. In vitro, SIRT6 overexpression contributed to doxorubicin resistance, but was blocked by the PARP inhibitor olaparib [162]. TH1579, an inhibitor of the MTH1 enzyme, which prevents the integration of oxidized nucleotides into DNA, decreased viability, cell cycle progression and induced apoptosis in osteosarcoma cells in vitro and in vivo [163]. Resistance to cisplatin was counteracted by downregulation of the DNA-dependent protein kinase catalytic subunit (DNA-PKcs) in PI3K/Akt/CD133+ cells, via a reduction in MARK2 [164], and also in combination with the checkpoint kinase 1 inhibitor AZD7762 [165] and the nucleotide excision repair pathway inhibitors NSC130813 and triptolide [166]. The role of small nucleolar RNAs has gained attention in recent years. Some of these RNAs (SNORD3A, SNORA13 and SNORA28) induce resistance to doxorubicin by modulating the expression of genes involved in DNA-damage sensing (GADD45A) and DNA repair (MYC) [167]. Other reports suggest that SNORA7A, modulated by the lncRNA H19, is oncogenic in osteosarcoma and correlates with poor patient survival [168].
Despite more studies being required to provide a more convincing understanding of the DNA repair pathways in osteosarcoma, promising pre-clinical results testing chemotherapy-potentiating DNA repair inhibitors in osteosarcoma and other tumors seems to hold potential for targeting both CSCs and their offspring. Moreover, considering that the DNA-damaging drugs doxorubicin and cisplatin are the most effective elected therapy for osteosarcoma [169], understanding their pharmacogenetic mechanisms may help to counteract their associated secondary side-effects and hold potential to development more tailored combinatorial therapies.

4.6. Enhanced Anti-Apoptotic Mechanisms

Programmed cell death or apoptosis occurs in diverse cellular phases, such as normal development, organogenesis, ageing or inflammatory response, and serves as a natural barrier to cancer. Apoptotic mechanisms involve a complex signaling cascade and are composed of both upstream regulators and downstream effector components [170]. Currently, two distinct, although tightly interconnected, signaling pathways control apoptotic cell death. In the intrinsic or mitochondrial pathway, the counterbalance between anti- (e.g., Bcl-2, Bcl-xL, Mcl-1) and pro-apoptotic (e.g., Bax, Bak) proteins of the Bcl-2 family determines the trigger to mitochondrial apoptosis [171,172]. Another distinct way of controlling apoptosis occurs due to the activation of cell-surface death receptors (e.g., DR3, DR5, TRAIL, TNF receptors), which are responsive to a diversity of death ligands (e.g., TRAIL, TNF, FAS) expressed by immunocompetent cells [172,173]. In case that the net chief signal is pro-apoptotic, the apoptotic program then culminates in the activation of normally latent proteases (effector caspases-3, -6 and -7), which will be responsible for the execution phase of apoptosis, in which the cell is progressively disassembled and its contents, so-called apoptotic bodies, engulfed by neighboring and phagocytic cells [174,175].
Osteosarcoma cells are no exception in what concerns evasion from apoptosis. Expression of key anti-apoptotic proteins, such as Bcl-2 [176], Bcl-xL [177] and survivin [178], was detected in cell lines and patient samples and correlated with enhanced cell survival and proliferation in vitro and also associated with poor prognosis and metastatic dissemination [179]. Conversely, activation of pro-apoptotic proteins, even if mediated by other survival-related molecules [180,181], promotes apoptosis in osteosarcoma cells. In fact, a plethora of studies reporting effects on proteins of the apoptotic signaling cascades is derived from the genetic or pharmacological modulation of several other signaling pathways, including self-renewal-related pathways. Nevertheless, experimental data suggest that direct inhibition of anti-apoptotic proteins may help to combat resistance to chemotherapy.
Despite that most chemotherapeutic drugs exert their biological effects through induction of apoptotic cell death, evasion from apoptosis seems exacerbated in CSCs [182,183]. In this regard, attempts have been made to inhibit the relevant anti-apoptotic proteins, namely, Bcl-2, Bcl-xL and Bcl-W, using the BH-3 mimetic ABT-737 [184,185,186]. These preliminary data provide a rationale for a wider exploratory use of inhibitors of anti-apoptotic proteins to target chemoresistance of tumor cells in osteosarcoma, as earlier suggested [187,188,189], as well as CSCs as previously endeavored in other tumor types [184,185,186,190].

4.7. Chemoresistance Conveyed by the Tumor Microenvironment—Hypoxia and Inflammation

The tumor microenvironment surrounding CSCs is detrimental for the preservation of their stemness state in the niche. For example, it has been shown that tumor-associated MSCs release extracellular vesicles containing pro-proliferative, anti-apoptotic and metastatic supportive molecules involving microRNAs and growth factors [191]. MSCs also revealed tropism towards osteosarcoma cells, favoring their aggressiveness, through the expression of chemotactic factors, such as MCP-1, GRO-α and TGF-β1, and transdifferentiation to cancer-associated fibroblasts (CAFs), which further increased the cytokine levels in the tumor microenvironment and promoted transendothelial cell migration [192]. Conversely, osteosarcoma cells themselves are also capable of inducing MSC differentiation into CAFs through Notch/Akt signaling [193]. Endothelial cells secreting exosomes are also promoters of osteosarcoma stemness through Notch signaling [194].

4.7.1. Hypoxia

Poorly organized networks of vascularization among solid tumors may determine the existence of hypoxic zones. The low oxygen tension present in these areas, despite generating toxic ROS, provides selective pressure for tumor growth and survival advantage for aggressive cells [195]. The central mediator of hypoxia, HIF-1, activates the transcription of genes involved in key aspects of cancer biology, representing an important therapeutic target. Moreover, hypoxia is considered an independent prognostic indicator of poor outcome and risk for metastasis development in osteosarcoma [196,197].
Expression of key hypoxia-related markers has been observed in osteosarcoma and was related to drug resistance [198], down-regulation of Wnt/β-catenin [199] and maintenance of the CSC phenotype in three-dimensional conditions [200]. Osteosarcoma as a solid tumor is highly susceptible to hypoxia activated pro-drugs such as TH-302, which has been shown to cooperate with doxorubicin against osteosarcoma-induced bone destruction and limiting the formation of pulmonary metastases [201]. The leading role of hypoxia in metastatic osteosarcoma dissemination seems to be mediated by the HIF-1α-CXCR4 pathway axis, which plays a crucial role during osteosarcoma cell migration [202,203]. Overexpression of the long non-coding RNA (lncRNA) HIF-2α-promoter upstream transcript (HIF2PUT) has gained attention as a regulatory marker of hypoxia in osteosarcoma. HIF2PUT inhibited in vitro osteosarcoma CSC proliferation and migration, the sphere-forming ability of MG-63 cells and decreased the CD133-positive cells [62], while others observed that HIF2PUT overexpression inhibited CSC migration and invasion through HIF2α downregulation [204]. Moreover, its expression levels were positively correlated with HIF-2α expression levels in primary tumor samples, predicting poor prognosis [205].
More recently, it was demonstrated that microvesicles derived from MSCs contributed to U2OS proliferation and migration under hypoxic conditions, which was associated with AKT and HIF-1α expression [206]. Hypoxia was suggested to induce dedifferentiation of MNNG/HOS cells, expression of Oct-4, Nanog and CD133, accelerated sphere formation and tumorigenesis in vivo—effects that were counterbalanced by the mTOR inhibitor rapamycin, which also suppressed HIF-1α expression [207]. Further studies are required concerning the role of hypoxia and oxidative stress in osteosarcoma CSCs’ biological features, in particular those studies involving adequate in vivo models that may uncover unanticipated markers of therapeutic resistance, such as HIF-2α previously reported in a humanized orthotopic model [208].

4.7.2. Inflammatory Networks

Multiple studies have indicated that MSCs produce soluble factors closely involved in cell proliferation. Consequently, these factors exert a proliferative and pro-inflammatory effect on cancer cells [209,210]. MSCs can also contribute to tumor progression by cooperating with the tumor cells to develop a suitable microenvironment [211].
Further, the MSCs can also transdifferentiate into CAFs that are considered essential stromal cells and crucial regulators for this favorable microenvironment creation and tumor progression [193,212]. By themselves, stromal cells are not malignant and maintain the tissue structure and function. However, when these cells acquire an active phenotype, they sustain cancer cell growth and tumor progression [213]. CAFs have a significant role in stimulating angiogenesis, cell proliferation, invasion and motility by releasing growth factors and cytokines, deregulating Notch and p53 signaling pathways, and producing matrix metalloproteinases [214]. Both MSCs and CAFs secrete several soluble factors, extracellular vesicles, chemokines and cytokines that stimulate tumor sustenance, growth and angiogenesis in osteosarcoma [215]. Within these factors, IL-6 plays a leading role in the inflammatory profile of this microenvironment.
Indeed, IL-6 is increasingly recognized as the soluble mediator linking chronic inflammation to cancer development, and its protein and mRNA are often overexpressed in serum and tumor samples from breast, bone, liver and colon cancers, both in humans and mice [216,217,218,219]. This cytokine secretion is also augmented by an increase in NF-kB expression and is described to have a paracrine effect on osteosarcoma cells via activation of the STAT3 pathway, which is a well-known activator of cells’ proliferation, enhancing tumor aggressiveness [215,220]. IL-6 is also described as responsible for the increased chemoresistance of human osteosarcoma cells [221] and exacerbates the invasive capacity of cells contributing to tumor development when secreted by CAFs. Furthermore, Cortini et al. found that IL-6 is responsible for CSC sphere growth, suggesting that it could be involved in the maintenance of the CSC population extremely involved in osteosarcoma relapse [215].
Other inflammatory mediators, such as TNF-α, IL-8, COX-2, are also encoded by NF-kB. In particular, when interacting with its receptors (TNFR1 and TNFR2), TNF-α induces cell survival and proliferation, an inflammatory response and an anti-apoptotic signal by NF-kB activation [214]. TNF-α is also an essential regulator of cancer-related inflammation by modulation of T cells, B cells and tumor-associated macrophages (TAMs) [222]. Specifically, several studies have reported TAMs as important immune cells implicated in various tumor-promoting tasks, including pro-inflammatory signaling, enhancement of angiogenesis, invasion, metastasis and therapy resistance [223,224]. Notably, the crosstalk between CAFs and T cells is reciprocal. Barnas et al. have described that activated T cells secrete factors that enhance the production of IL-6 by lung CAFs. Then, these activated fibroblasts induced the expression of IFN-γ and IL-17, both of which are known to impact the progression as well as the inhibition of tumor growth and metastasis [225].
TGF-β1 has also emerged as one of the most relevant cytokines secreted by osteosarcoma cells and a key player in self-renewal and maintenance of stemness features. It is described that the tumor maintains the stemness of MSC through the TGF/Smad3 pathway and that osteosarcoma cells, via TGF-β1 secretion, enhance the production of pro-tumorigenic cytokines, such as IL-6, in the nearby stroma [226]. Additionally, TGF-β was shown to increase ROS’ production in CAFs by the impairment of the respiratory transport chain, specifically acting on Complex IV via GS3K action. Furthermore, the Smad signaling contributes to maintaining ROS accumulation contributing to the inflammatory networks that sustain tumor progression [227].
Besides the normal functions of fibroblasts in matrix remodeling and secretion of ECM components, accumulating evidence identify CAFs as modulators of a tumor’s immune milieu. These activated fibroblasts operate towards a pro-inflammatory and immunosuppressive microenvironment because of reciprocal interactions between immune cells and CAFs mediated by the CAFs’ secretome. Despite many studies, the plasticity of CAFs and their contribution to chemoresistance and stemness features within tumors are still poorly understood. However, numerous lines of evidence imply CAFs as potential targets to new therapeutic approaches to inhibit tumor progression and therefore also warrant further investigation for osteosarcoma-directed targeting.

5. Immunotherapy in Osteosarcoma

Immunotherapy has been described as a promising strategy for advanced cancer treatment [228]. New therapies involving the application of tumor vaccines, immune modulators, genetically modified T cells, cytokines, immune checkpoint inhibitors or combination therapy can be used to decrease treatment side reactions, increase therapeutic effects and improve the quality of life of cancer patients [229,230].
Recently, multiple clinical trials have been conducted in various tumors to test the efficacy of immunomodulators, and several preclinical studies have also supported the application of immunotherapy in osteosarcoma, as previously reviewed [231]. Osteosarcoma is characterized by poor immune responses, which include low densities of tumor-infiltrating lymphocytes and increased anti-inflammatory responses within the tumor, leading to drug resistance and diminished overall patient survival [232]. Monoclonal antibodies, such as anti-cytotoxic T lymphocyte antigen-4 (CTLA-4), B7-H3, programmed cell receptor-1 (PD-1), and its ligand (PD-L1), have been developed to target immune checkpoints and have garnered widespread attention due to their excellent therapeutic effects on malignant tumors. However, most checkpoint inhibitors are less effective in treating solid tumors, including osteosarcoma [230,233]. PDL-1 and PD-1 expression have been considered potentially useful biomarkers for further studies [234]. However, anti-PD-1 antibodies, such as pembrolizumab, nivolumab and atezolizumab, showed limited therapeutic efficacy in osteosarcoma patients [234,235,236,237]. Lussier et al. revealed that T cells that infiltrated the osteosarcoma microenvironment could downregulate additional inhibitory receptors such as CTLA-4, which conspired to hinder tumor immunity [238]. They combined CTLA-4 with a PD-1 antibody in a K7M2 murine model of metastatic osteosarcoma, and the tumor progression was under control. Additionally, Helm et al. found that osteosarcoma mouse models treated with a combination of CTLA-4 and PD-1 immune checkpoint inhibitors showed a rise in CD8+ T cells, increasing the cytotoxicity in the tumor microenvironment and contributing to a better prognosis [239].
Recent research using chimeric antigen receptor (CAR) T cells has provided promising outcomes [240,241]. Their target is the immune checkpoint molecule B7-H3 (CD276), a tumor antigen that is significantly upregulated in osteosarcoma [242]. This molecule is described to play a crucial role in T cell function inhibition [243]. Despite the advantages of CAR T cell therapy and its increasing usage in clinical practice in lymphomas and acute lymphocytic leukemia [244], its applicability in solid tumors presents multiple challenges [245]. In terms of production, the viral transduction variation efficiency, as well as the high cost and time-consuming nature of the T cell-manufacturing procedure, is a problem. Moreover, another factor in the failure to combat solid tumors with CAR T cells is the restricted infiltration and poor persistence caused by a stiff osteoid bone tumor matrix and immunosuppressive components in the tumor microenvironment [231,246].
Increasing evidence suggests that CSCs play a critical role in innate and adaptive immunity, influencing the effectiveness of immunotherapy. The activation of stemness programs appears to limit antitumor immune responses by decreasing T cell infiltration and increasing the expression of immunosuppressive checkpoints, resulting in immunologically cold microenvironments [247,248]. Widespread negative associations between stemness and anticancer immunity have been described for most cancers, suggesting that targeting the stemness phenotype may render tumors more responsive to immune control [247,249]. Rainusso et al. identified a mechanism by which targeting HER2 with genetically modified T cells could eliminate osteosarcoma CSCs [250]. More recently, Shao et al. showed evidence that all-trans retinoic acid could indirectly limit CSC occurrence in osteosarcoma by the inhibition of the M2-like macrophages that activated the CSC phenotype [251]. Curcumol seems to inhibit the polarization of M2-like macrophages synergizing with cisplatin and reversing the drug-induced expression of ABCB1, ABCC1 and ABCG2 [252]. Guo et al. also demonstrated that a low CSCs score is correlated with increased immunocyte infiltration [249]. Overall, the tumor microenvironment is a problem in modern cancer research, which has a decisive function in the occurrence and progression of tumors. Its complete characterization, including extensive knowledge of the immune microenvironment, is crucial for immunotherapy implementation. which can contribute to patients who are refractory to therapy or who have relapsed, improving their prognosis.

6. Conclusions

The mechanisms of resistance explored in this revision confer resistance to doxorubicin, cisplatin and methotrexate, which are the main therapies used in osteosarcoma treatment. As mentioned before, this classic drug regimen still cannot be excluded from the main frontline treatments, since basically all the efforts made towards the discovery and development of alternative therapies have up to date failed in clinical trials. Moreover, as Harris and Hawkins have recently reviewed in IJMS, a number of drugs tested in diverse clinical trial phases have failed to significantly prolong the event-free survival or overall survival rates of poor responders to pre-operative chemotherapy or metastasis-presenting osteosarcoma patients (e.g., pirarubicin, pemetrexed, carboplatin, ifosfamide, etoposide, topotecan, thiotpeta, gemcitabine, L-alanosine, zoledronic acid, and several drugs targeting tyrosine kinases, among others) [253]. Nevertheless, a number of alternative drugs have been tested in osteosarcoma to target the resistance signaling pathways herein mentioned, as depicted in Figure 4, among many others. Fortunately, the rapid advances in the biological understanding of osteosarcoma, based on increasingly more robust pre-clinical models and access to rapid clinical testing, will certainly ameliorate the long-awaited outcomes of osteosarcoma patients.
The signaling pathways herein revisited clearly display inter-crosstalk and also modulate the activity and expression of other pathways involved in pluripotency and cell self-renewal. A precise detail of these latter pathways are a subject of discussion in another review in progress, but it is worth mentioning the relevance of Sox2, Nanog, Wnt, Notch and Stat3, along with diverse types of microRNAs. In general, the crosstalk with the pathways is harmonized to promote CSCs’ survival and aggressiveness, reinforcing their potential contribution to osteosarcoma resistance to chemotherapy and warranting further pre-clinical investigations focused on these special cell subsets.

Author Contributions

Conceptualization, S.R.M.-N. and C.M.F.G.; investigation, S.R.M.-N. and G.S.-R.; writing—original draft preparation, S.R.M.-N. and G.S.-R.; writing—review and editing, S.R.M.-N. and C.M.F.G.; funding acquisition, C.M.F.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Funds via FCT (Foundation for Science and Technology) through the Strategic Projects UIDB/04539/2020 and UIDP/04539/2020 (CIBB).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict 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. Odri, G.A.; Tchicaya-Bouanga, J.; Yoon, D.J.; Modrowski, D. Metastatic Progression of Osteosarcomas: A Review of Current Knowledge of Environmental versus Oncogenic Drivers. Cancers 2022, 14, 360. [Google Scholar] [CrossRef] [PubMed]
  2. Sarhadi, V.K.; Daddali, R.; Seppänen-Kaijansinkko, R. Mesenchymal Stem Cells and Extracellular Vesicles in Osteosarcoma Pathogenesis and Therapy. Int. J. Mol. Sci. 2021, 22, 11035. [Google Scholar] [CrossRef] [PubMed]
  3. Jubelin, C.; Muñoz-Garcia, J.; Cochonneau, D.; Moranton, E.; Heymann, M.F.; Heymann, D. Biological evidence of cancer stem-like cells and recurrent disease in osteosarcoma. Cancer Drug Resist. 2022, 5, 184–198. [Google Scholar] [CrossRef] [PubMed]
  4. Fernandes, I.; Melo-Alvim, C.; Lopes-Brás, R.; Esperança-Martins, M.; Costa, L. Osteosarcoma Pathogenesis Leads the Way to New Target Treatments. Int. J. Mol. Sci. 2021, 22, 813. [Google Scholar] [CrossRef]
  5. Siegel, R.L.; Miller, K.D.; Fuchs, H.E.; Jemal, A. Cancer Statistics, 2021. CA Cancer J. Clin. 2021, 71, 7–33. [Google Scholar] [CrossRef]
  6. de Campos Vieira Abib, S.; Chui, C.H.; Abdelhafeez, A.H.; Fernandez-Pineda, I.; Elgendy, A.; Karpelowsky, J.; Lobos, P.; Wijnen, M.; Fuchs, J.; Hayes, A.; et al. International Society of Paediatric Surgical Oncology (IPSO) Surgical Practice Guidelines. eCancer 2022, 16, 1356. [Google Scholar] [CrossRef]
  7. Ta, H.; Dass, C.; Choong, P.; Dunstan, D. Osteosarcoma treatment: State of the art. Cancer Metastasis Rev. 2009, 28, 247–263. [Google Scholar] [CrossRef]
  8. Ritter, J.; Bielack, S.S. Osteosarcoma. Ann. Oncol. 2010, 21, vii320–vii325. [Google Scholar] [CrossRef]
  9. Whelan, J.S.; Davis, L.E. Osteosarcoma, Chondrosarcoma, and Chordoma. J. Clin. Oncol. 2017, 36, 188–193. [Google Scholar] [CrossRef]
  10. Heaton, T.E.; Davidoff, A.M. Surgical treatment of pulmonary metastases in pediatric solid tumors. Semin. Pediatr. Surg. 2016, 25, 311–317. [Google Scholar] [CrossRef] [Green Version]
  11. Federman, N.; Bernthal, N.; Eilber, F.C.; Tap, W.D. The Multidisciplinary Management of Osteosarcoma. Curr. Treat. Options Oncol. 2009, 10, 82–93. [Google Scholar] [CrossRef] [PubMed]
  12. Jaffe, N.; Bruland, Ø.S.; Bielack, S. Pediatric and Adolescent Osteosarcoma, 1st ed.; Springer: New York, NY, USA, 2009. [Google Scholar]
  13. Goorin, A.M.; Schwartzentruber, D.J.; Devidas, M.; Gebhardt, M.C.; Ayala, A.G.; Harris, M.B.; Helman, L.J.; Grier, H.E.; Link, M.P. Presurgical Chemotherapy Compared With Immediate Surgery and Adjuvant Chemotherapy for Nonmetastatic Osteosarcoma: Pediatric Oncology Group Study POG-8651. J. Clin. Oncol. 2003, 21, 1574–1580. [Google Scholar] [CrossRef] [PubMed]
  14. Piperno-Neumann, S.; Le Deley, M.C.; Rédini, F.; Pacquement, H.; Marec-Bérard, P.; Petit, P.; Brisse, H.; Lervat, C.; Gentet, J.C.; Entz-Werlé, N.; et al. Zoledronate in combination with chemotherapy and surgery to treat osteosarcoma (OS2006): A randomised, multicentre, open-label, phase 3 trial. Lancet Oncol. 2016, 17, 1070–1080. [Google Scholar] [CrossRef]
  15. Lee, J.A.; Jeon, D.G.; Cho, W.H.; Song, W.S.; Yoon, H.S.; Park, H.J.; Park, B.K.; Choi, H.S.; Ahn, H.S.; Lee, J.W.; et al. Higher Gemcitabine Dose Was Associated With Better Outcome of Osteosarcoma Patients Receiving Gemcitabine-Docetaxel Chemotherapy. Pediatr. Blood Cancer 2016, 63, 1552–1556. [Google Scholar] [CrossRef]
  16. Palmerini, E.; Setola, E.; Grignani, G.; D’Ambrosio, L.; Comandone, A.; Righi, A.; Longhi, A.; Cesari, M.; Paioli, A.; Hakim, R.; et al. High Dose Ifosfamide in Relapsed and Unresectable High-Grade Osteosarcoma Patients: A Retrospective Series. Cells 2020, 9, 2389. [Google Scholar] [CrossRef]
  17. Davis, L.E.; Bolejack, V.; Ryan, C.W.; Ganjoo, K.N.; Loggers, E.T.; Chawla, S.; Agulnik, M.; Livingston, M.B.; Reed, D.; Keedy, V.; et al. Randomized Double-Blind Phase II Study of Regorafenib in Patients With Metastatic Osteosarcoma. J. Clin. Oncol. 2019, 37, 1424–1431. [Google Scholar] [CrossRef]
  18. Gaspar, N.; Venkatramani, R.; Hecker-Nolting, S.; Melcon, S.G.; Locatelli, F.; Bautista, F.; Longhi, A.; Lervat, C.; Entz-Werle, N.; Casanova, M.; et al. Lenvatinib with etoposide plus ifosfamide in patients with refractory or relapsed osteosarcoma (ITCC-050): A multicentre, open-label, multicohort, phase 1/2 study. Lancet Oncol. 2021, 22, 1312–1321. [Google Scholar] [CrossRef]
  19. Lewis, I.J.; Nooij, M.A.; Whelan, J.; Sydes, M.R.; Grimer, R.; Hogendoorn, P.C.W.; Memon, M.A.; Weeden, S.; Uscinska, B.M.; van Glabbeke, M.; et al. Improvement in Histologic Response But Not Survival in Osteosarcoma Patients Treated With Intensified Chemotherapy: A Randomized Phase III Trial of the European Osteosarcoma Intergroup. J. Natl. Cancer Inst. 2007, 99, 112–128. [Google Scholar] [CrossRef]
  20. Lewis, I.J.; Weeden, S.; Machin, D.; Stark, D.; Craft, A.W. Received Dose and Dose-Intensity of Chemotherapy and Outcome in Nonmetastatic Extremity Osteosarcoma. J. Clin. Oncol. 2000, 18, 4028–4037. [Google Scholar] [CrossRef]
  21. Marina, N.M.; Smeland, S.; Bielack, S.S.; Bernstein, M.; Jovic, G.; Krailo, M.D.; Hook, J.M.; Arndt, C.; van den Berg, H.; Brennan, B.; et al. Comparison of MAPIE versus MAP in patients with a poor response to preoperative chemotherapy for newly diagnosed high-grade osteosarcoma (EURAMOS-1): An open-label, international, randomised controlled trial. Lancet Oncol. 2016, 17, 1396–1408. [Google Scholar] [CrossRef] [Green Version]
  22. Zhang, Y.; He, Z.; Duan, Y.; Wang, C.; Kamar, S.; Shi, X.; Yang, J.; Yang, J.; Zhao, N.; Han, L.; et al. Does intensified chemotherapy increase survival outcomes of osteosarcoma patients? A meta-analysis. J. Bone Oncol. 2018, 12, 54–60. [Google Scholar] [CrossRef] [PubMed]
  23. Bielack, S.S.; Smeland, S.; Whelan, J.S.; Marina, N.; Jovic, G.; Hook, J.M.; Krailo, M.D.; Gebhardt, M.; Pápai, Z.; Meyer, J.; et al. Methotrexate, Doxorubicin, and Cisplatin (MAP) Plus Maintenance Pegylated Interferon Alfa-2b Versus MAP Alone in Patients With Resectable High-Grade Osteosarcoma and Good Histologic Response to Preoperative MAP: First Results of the EURAMOS-1 Good Response Randomized Controlled Trial. J. Clin. Oncol. 2015, 33, 2279–2287. [Google Scholar] [CrossRef] [PubMed]
  24. Meyers, P.A.; Schwartz, C.L.; Krailo, M.; Kleinerman, E.S.; Betcher, D.; Bernstein, M.L.; Conrad, E.; Ferguson, W.; Gebhardt, M.; Goorin, A.M.; et al. Osteosarcoma: A Randomized, Prospective Trial of the Addition of Ifosfamide and/or Muramyl Tripeptide to Cisplatin, Doxorubicin, and High-Dose Methotrexate. J. Clin. Oncol. 2005, 23, 2004–2011. [Google Scholar] [CrossRef] [PubMed]
  25. Ferrari, S.; Ruggieri, P.; Cefalo, G.; Tamburini, A.; Capanna, R.; Fagioli, F.; Comandone, A.; Bertulli, R.; Bisogno, G.; Palmerini, E.; et al. Neoadjuvant Chemotherapy With Methotrexate, Cisplatin, and Doxorubicin With or Without Ifosfamide in Nonmetastatic Osteosarcoma of the Extremity: An Italian Sarcoma Group Trial ISG/OS-1. J. Clin. Oncol. 2012, 30, 2112–2118. [Google Scholar] [CrossRef]
  26. Yang, Y.; Han, L.; He, Z.; Li, X.; Yang, S.; Yang, J.; Zhang, Y.; Li, D.; Yang, Y.; Yang, Z. Advances in limb salvage treatment of osteosarcoma. J. Bone Oncol. 2018, 10, 36–40. [Google Scholar] [CrossRef] [PubMed]
  27. Bousquet, M.; Noirot, C.; Accadbled, F.; Sales de Gauzy, J.; Castex, M.P.; Brousset, P.; Gomez-Brouchet, A. Whole-exome sequencing in osteosarcoma reveals important heterogeneity of genetic alterations. Ann. Oncol. 2016, 27, 738–744. [Google Scholar] [CrossRef]
  28. Gambera, S.; Abarrategi, A.; González-Camacho, F.; Morales-Molina, Á.; Roma, J.; Alfranca, A.; García-Castro, J. Clonal dynamics in osteosarcoma defined by RGB marking. Nat. Commun. 2018, 9, 3994. [Google Scholar] [CrossRef]
  29. Schiavone, K.; Garnier, D.; Heymann, M.F.; Heymann, D. The Heterogeneity of Osteosarcoma: The Role Played by Cancer Stem Cells. In Stem Cells Heterogeneity in Cancer; Birbrair, A., Ed.; Springer International Publishing: Cham, Switzerland, 2019; pp. 187–200. [Google Scholar]
  30. Izadpanah, S.; Shabani, P.; Aghebati-Maleki, A.; Baghbanzadeh, A.; Fotouhi, A.; Bisadi, A.; Aghebati-Maleki, L.; Baradaran, B. Prospects for the involvement of cancer stem cells in the pathogenesis of osteosarcoma. J. Cell Physiol. 2020, 235, 4167–4182. [Google Scholar] [CrossRef]
  31. Marquardt, S.; Solanki, M.; Spitschak, A.; Vera, J.; Pützer, B.M. Emerging functional markers for cancer stem cell-based therapies: Understanding signaling networks for targeting metastasis. Semin. Cancer Biol. 2018, 53, 90–109. [Google Scholar] [CrossRef]
  32. Kemper, K.; de Goeje, P.L.; Peeper, D.S.; van Amerongen, R. Phenotype Switching: Tumor Cell Plasticity as a Resistance Mechanism and Target for Therapy. Cancer Res. 2014, 74, 5937–5941. [Google Scholar] [CrossRef] [Green Version]
  33. Cabrera, M.C.; Hollingsworth, R.E.; Hurt, E.M. Cancer stem cell plasticity and tumor hierarchy. World J. Stem Cells 2015, 7, 27–36. [Google Scholar] [CrossRef] [PubMed]
  34. van Neerven, S.M.; Tieken, M.; Vermeulen, L.; Bijlsma, M.F. Bidirectional interconversion of stem and non-stem cancer cell populations: A reassessment of theoretical models for tumor heterogeneity. Mol. Cell Oncol. 2016, 3, e1098791. [Google Scholar] [CrossRef] [PubMed]
  35. Qin, S.; Jiang, J.; Lu, Y.; Nice, E.C.; Huang, C.; Zhang, J.; He, W. Emerging role of tumor cell plasticity in modifying therapeutic response. Signal Transduct. Target Ther. 2020, 5, 228. [Google Scholar] [CrossRef] [PubMed]
  36. Di Fiore, R.; Santulli, A.; Drago Ferrante, R.; Giuliano, M.; De Blasio, A.; Messina, C.; Pirozzi, G.; Tirino, V.; Tesoriere, G.; Vento, R. Identification and expansion of human osteosarcoma-cancer-stem cells by long-term 3-aminobenzamide treatment. J. Cell Physiol. 2009, 219, 301–313. [Google Scholar] [CrossRef]
  37. Fujji, H.; Honoki, K.; Tsujiuchi, T.; Kido, A.; Yoshitani, K.; Takakura, Y. Sphere-forming stem-like cell populations with drug resistance in human sarcoma cell lines. Int. J. Oncol. 2009, 34, 1381–1386. [Google Scholar] [CrossRef]
  38. Wang, L.; Park, P.; Lin, C.Y. Characterization of stem cell attributes in human osteosarcoma cell lines. Cancer Biol. Ther. 2009, 8, 543–552. [Google Scholar] [CrossRef]
  39. Saini, V.; Hose, C.D.; Monks, A.; Nagashima, K.; Han, B.; Newton, D.L.; Millione, A.; Shah, J.; Hollingshead, M.G.; Hite, K.M.; et al. Identification of CBX3 and ABCA5 as Putative Biomarkers for Tumor Stem Cells in Osteosarcoma. PLoS ONE 2012, 7, e41401. [Google Scholar] [CrossRef]
  40. Martins-Neves, S.R.; Lopes, A.; do Carmo, A.; Paiva, A.; Simoes, P.; Abrunhosa, A.; Gomes, C. Therapeutic implications of an enriched cancer stem-like cell population in a human osteosarcoma cell line. BMC Cancer 2012, 12, 139. [Google Scholar] [CrossRef]
  41. Martins-Neves, S.R.; Corver, W.E.; Paiva-Oliveira, D.I.; van den Akker, B.E.W.M.; Briaire-de-Bruijn, I.H.; Bovée, J.V.M.G.; Gomes, C.M.F.; Cleton-Jansen, A.-M. Osteosarcoma Stem Cells Have Active Wnt/β-Catenin and Overexpress SOX2 and KLF4. J. Cell Physiol. 2016, 231, 876–886. [Google Scholar] [CrossRef]
  42. Zhang, H.; Wu, H.; Zheng, J.; Yu, P.; Xu, L.; Jiang, P.; Gao, J.; Wang, H.; Zhang, Y. Transforming Growth Factor b1 Signal is Crucial for Dedifferentiation of Cancer Cells to Cancer Stem Cells in Osteosarcoma. Stem Cells 2013, 31, 433–446. [Google Scholar] [CrossRef]
  43. Salerno, M.; Avnet, S.; Bonuccelli, G.; Eramo, A.; De Maria, R.; Gambarotti, M.; Gamberi, G.; Baldini, N. Sphere-forming cell subsets with cancer stem cell properties in human musculoskeletal sarcomas. Int. J. Oncol. 2013, 43, 95–102. [Google Scholar] [CrossRef] [PubMed]
  44. Yu, L.; Fan, Z.; Fang, S.; Yang, J.; Gao, T.; Simões, B.M.; Eyre, R.; Guo, W.; Clarke, R.B. Cisplatin selects for stem-like cells in osteosarcoma by activating Notch signaling. Oncotarget 2016, 7, 33055–33068. [Google Scholar] [CrossRef] [PubMed]
  45. Wang, F.; Zhang, Z.; Li, Q.; Yu, T.; Ma, C. Untargeted LC-MS/MS analysis reveals metabolomics feature of osteosarcoma stem cell response to methotrexate. Cancer Cell Int. 2020, 20, 269. [Google Scholar] [CrossRef] [PubMed]
  46. Chen, X.; Zhang, Q.; Dang, X.; Song, T.; Wang, Y.; Yu, Z.; Zhang, S.; Fan, J.; Cong, F.; Zhang, W.; et al. Targeting the CtBP1-FOXM1 transcriptional complex with small molecules to overcome MDR1-mediated chemoresistance in osteosarcoma cancer stem cells. J. Cancer 2021, 12, 482–497. [Google Scholar] [CrossRef] [PubMed]
  47. Honoki, K.; Fujji, H.; Kubo, A.; Kido, A.; Mori, T.; Tanaka, Y.; Tsujiuchi, T. Possible involvement of stem-like populations with elevated ALDH1 in sarcomas for chemotherapeutic drug resistance. Oncol. Rep. 2010, 24, 501–505. [Google Scholar] [CrossRef]
  48. Wang, L.; Park, P.; Zhang, H.; La Marca, F.; Lin, C.-Y. Prospective identification of tumorigenic osteosarcoma cancer stem cells in OS99-1 cells based on high aldehyde dehydrogenase activity. Int. J. Cancer 2011, 128, 294–303. [Google Scholar] [CrossRef]
  49. Krause, U.; Ryan, D.M.; Clough, B.H.; Gregory, C.A. An unexpected role for a Wnt-inhibitor: Dickkopf-1 triggers a novel cancer survival mechanism through modulation of aldehyde-dehydrogenase-1 activity. Cell Death Dis. 2014, 5, e1093. [Google Scholar] [CrossRef]
  50. Martins-Neves, S.R.; Paiva-Oliveira, D.I.; Wijers-Koster, P.M.; Abrunhosa, A.J.; Fontes-Ribeiro, C.; Bovée, J.V.M.G.; Cleton-Jansen, A.M.; Gomes, C.M.F. Chemotherapy induces stemness in osteosarcoma cells through activation of Wnt/β-catenin signaling. Cancer Lett. 2016, 370, 286–295. [Google Scholar] [CrossRef]
  51. Mu, X.; Patel, S.; Mektepbayeva, D.; Mahjoub, A.; Huard, J.; Weiss, K. Retinal Targets ALDH Positive Cancer Stem Cell and Alters the Phenotype of Highly Metastatic Osteosarcoma Cells. Sarcoma 2015, 2015, 784954. [Google Scholar] [CrossRef]
  52. Nomura, M.; Rainusso, N.; Lee, Y.C.; Dawson, B.; Coarfa, C.; Han, R.; Larson, J.L.; Shuck, R.; Kurenbekova, L.; Yustein, J.T. Tegavivint and the β-Catenin/ALDH Axis in Chemotherapy-Resistant and Metastatic Osteosarcoma. J. Natl. Cancer Inst. 2019, 111, 1216–1227. [Google Scholar] [CrossRef]
  53. Yang, M.; Yan, M.; Zhang, R.; Li, J.; Luo, Z. Side population cells isolated from human osteosarcoma are enriched with tumor-initiating cells. Cancer Sci. 2011, 102, 1774–1781. [Google Scholar] [CrossRef] [PubMed]
  54. Yi, X.; Zhao, Y.; Qiao, L.; Jin, C.; Tian, J.; Li, Q. Aberrant Wnt/b-catenin signaling and elevated expression of stem cell proteins are associated with osteosarcoma side population cells of high tumorigenicity. Mol. Med. Rep. 2015, 12, 5042–5048. [Google Scholar] [CrossRef] [PubMed]
  55. Liu, L.-M.; Sun, H.-A.; Li, X.; Chen, Y.; Wei, B.-F.; Li, X.-J. Cluster of differentiation-44- and octamer-binding transcription factor-4-positive stem-like osteosarcoma cells involved in tumor development. Oncol. Lett. 2015, 10, 273–276. [Google Scholar] [CrossRef] [PubMed]
  56. Wang, Y.; Teng, J. Increased multi-drug resistance and reduced apoptosis in osteosarcoma side population cells are crucial factors for tumor recurrence. Exp. Ther. Med. 2016, 12, 81–86. [Google Scholar] [CrossRef] [PubMed]
  57. Ren, Y.M.; Duan, Y.H.; Sun, Y.B.; Yang, T.; Zhao, W.J.; Zhang, D.L.; Tian, Z.W.; Tian, M.Q. Exploring the key genes and pathways of side population cells in human osteosarcoma using gene expression array analysis. J. Orthop. Surg. Res. 2018, 13, 153. [Google Scholar] [CrossRef]
  58. Tirino, V.; Desiderio, V.; d’Aquino, R.; De Francesco, F.; Pirozzi, G.; Galderisi, U.; Cavaliere, C.; De Rosa, A.; Papaccio, G. Detection and Characterization of CD133+ Cancer Stem Cells in Human Solid Tumours. PLoS ONE 2008, 3, e3469. [Google Scholar] [CrossRef]
  59. He, A.; Qi, W.; Huang, Y.; Feng, T.; Chen, J.; Sun, Y.; Shen, Z.; Yao, Y. CD133 expression predicts lung metastasis and poor prognosis in osteosarcoma patients: A clinical and experimental study. Exp. Ther. Med. 2012, 4, 435–441. [Google Scholar] [CrossRef]
  60. Fujiwara, T.; Katsuda, T.; Hagiwara, K.; Kosaka, N.; Yoshioka, Y.; Takahashi, R.U.; Takeshita, F.; Kubota, D.; Kondo, T.; Ichikawa, H.; et al. Clinical Relevance and Therapeutic Significance of MicroRNA-133a Expression Profiles and Functions in Malignant Osteosarcoma-Initiating Cells. Stem Cells 2014, 32, 959–973. [Google Scholar] [CrossRef]
  61. Zhou, J.; Wu, S.; Chen, Y.; Zhao, J.; Zhang, K.; Wang, J.; Chen, S. microRNA-143 is associated with the survival of ALDH1+CD133+ osteosarcoma cells and the chemoresistance of osteosarcoma. Exp. Biol. Med. 2015, 240, 867–875. [Google Scholar] [CrossRef]
  62. Wang, Y.; Yao, J.; Meng, H.; Yu, Z.; Wang, Z.; Yuan, X.; Chen, H.; Wang, A. A novel long non-coding RNA, hypoxia-inducible factor-2a promoter upstream transcript, functions as an inhibitor of osteosarcoma stem cells in vitro. Mol. Med. Rep. 2015, 11, 2534–2540. [Google Scholar] [CrossRef] [Green Version]
  63. Chen, F.; Zeng, Y.; Qi, X.; Chen, Y.; Ge, Z.; Jiang, Z.; Zhang, X.; Dong, Y.; Chen, H.; Yu, Z. Targeted salinomycin delivery with EGFR and CD133 aptamers based dual-ligand lipid-polymer nanoparticles to both osteosarcoma cells and cancer stem cells. Nanomedicine 2018, 14, 2115–2127. [Google Scholar] [CrossRef]
  64. Levings, P.P.; McGarry, S.V.; Currie, T.P.; Nickerson, D.M.; McClellan, S.; Ghivizzani, S.C.; Steindler, D.A.; Gibbs, C.P. Expression of an Exogenous Human Oct-4 Promoter Identifies Tumor-Initiating Cells in Osteosarcoma. Cancer Res. 2009, 69, 5648–5655. [Google Scholar] [CrossRef] [PubMed]
  65. Adhikari, A.S.; Agarwal, N.; Wood, B.M.; Porretta, C.; Ruiz, B.; Pochampally, R.R.; Iwakuma, T. CD117 and Stro-1 Identify Osteosarcoma Tumor-Initiating Cells Associated with Metastasis and Drug Resistance. Cancer Res. 2010, 70, 4602–4612. [Google Scholar] [CrossRef] [PubMed]
  66. Pu, Y.; Zhao, F.; Wang, H.; Cai, W.; Gao, J.; Li, Y.; Cai, S. MiR-34a-5p promotes the multi-drug resistance of osteosarcoma by targeting the CD117 gene. Oncotarget 2016, 7, 28420–28434. [Google Scholar] [CrossRef] [PubMed]
  67. Rouleau, C.; Sancho, J.; Campos-Rivera, J.; Teicher, B.A. Endosialin expression in side populations in human sarcoma cell lines. Oncol. Lett. 2012, 3, 325–329. [Google Scholar] [CrossRef] [PubMed]
  68. Sun, D.; Liao, G.; Liu, K.; Jian, H. Endosialin-expressing bone sarcoma stem-like cells are highly tumor-initiating and invasive. Mol. Med. Rep. 2015, 12, 5665–5670. [Google Scholar] [CrossRef]
  69. Tian, J.; Li, X.; Si, M.; Liu, T.; Li, J. CD271+ Osteosarcoma Cells Display Stem-Like Properties. PLoS ONE 2014, 9, e98549. [Google Scholar] [CrossRef]
  70. Zhang, D.; Zhao, Q.; Sun, H.; Yin, L.; Wu, J.; Xu, J.; He, T.; Yang, C.; Liang, C. Defective autophagy leads to the suppression of stem-like features of CD271+ osteosarcoma cells. J Biomed. Sci. 2016, 23, 82. [Google Scholar] [CrossRef]
  71. Tian, J.; Gu, Y.; Li, Y.; Liu, T. CD271 antibody-functionalized HGNs for targeted photothermal therapy of osteosarcoma stem cells. Nanotechnology. 2020, 31, 305707. [Google Scholar] [CrossRef]
  72. Ren, T.; Piperdi, S.; Koirala, P.; Park, A.; Zhang, W.; Ivenitsky, D.; Zhang, Y.; Villanueva-Siles, E.; Hawkins, D.S.; Roth, M.; et al. CD49b inhibits osteogenic differentiation and plays an important role in osteosarcoma progression. Oncotarget 2017, 8, 87848–87859. [Google Scholar] [CrossRef] [Green Version]
  73. Zhou, Z.; Li, Y.; Kuang, M.; Wang, X.; Jia, Q.; Cao, J.; Hu, J.; Wu, S.; Wang, Z.; Xiao, J. The CD24+ cell subset promotes invasion and metastasis in human osteosarcoma. EBioMedicine 2020, 51, 102598. [Google Scholar] [CrossRef] [PubMed]
  74. Martins-Neves, S.R.; Cleton-Jansen, A.M.; Gomes, C.M.F. Therapy-induced enrichment of cancer stem-like cells in solid human tumors: Where do we stand? Pharmacol. Res. 2018, 137, 193–204. [Google Scholar] [CrossRef] [PubMed]
  75. Kim, W.T.; Ryu, C.J. Cancer stem cell surface markers on normal stem cells. BMB Rep. 2017, 50, 285–298. [Google Scholar] [CrossRef] [PubMed]
  76. 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]
  77. 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]
  78. Croker, A.K.; Goodale, D.; Chu, J.; Postenka, C.; Hedley, B.D.; Hess, D.A.; Allan, A.L. High aldehyde dehydrogenase and expression of cancer stem cell markers selects for breast cancer cells with enhanced malignant and metastatic ability. J. Cell Mol. Med. 2009, 13, 2236–2252. [Google Scholar] [CrossRef]
  79. Brown, H.K.; Tellez-Gabriel, M.; Heymann, D. Cancer stem cells in osteosarcoma. Cancer Lett. 2017, 386, 189–195. [Google Scholar] [CrossRef]
  80. Otoukesh, B.; Boddouhi, B.; Moghtadaei, M.; Kaghazian, P.; Kaghazian, M. Novel molecular insights and new therapeutic strategies in osteosarcoma. Cancer Cell Int. 2018, 18, 158. [Google Scholar] [CrossRef]
  81. 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]
  82. Fan, Y.; Cheng, H.; Liu, Y.; Liu, S.; Lowe, S.; Li, Y.; Bentley, R.; King, B.; Tuason, J.P.; Zhou, Q.; et al. Metformin anticancer: Reverses tumor hypoxia induced by bevacizumab and reduces the expression of cancer stem cell markers CD44/CD117 in human ovarian cancer SKOV3 cells. Front. Pharmacol. 2022, 13, 955984. [Google Scholar] [CrossRef]
  83. Samsonraj, R.M.; Raghunath, M.; Nurcombe, V.; Hui, J.H.; van Wijnen, A.J.; Cool, S.M. Concise Review: Multifaceted Characterization of Human Mesenchymal Stem Cells for Use in Regenerative Medicine. Stem Cells 2017, 6, 2173–2185. [Google Scholar] [CrossRef] [PubMed]
  84. Kannan, S.; Lock, I.; Ozenberger, B.B.; Jones, K.B. Genetic drivers and cells of origin in sarcomagenesis. J. Pathol. 2021, 254, 474–493. [Google Scholar] [CrossRef] [PubMed]
  85. Moitra, K.; Lou, H.; Dean, M. Multidrug Efflux Pumps and Cancer Stem Cells: Insights Into Multidrug Resistance and Therapeutic Development. Clin. Pharmacol. Ther. 2011, 89, 491–502. [Google Scholar] [CrossRef]
  86. Chen, Z.; Shi, T.; Zhang, L.; Zhu, P.; Deng, M.; Huang, C.; Hu, T.; Jiang, L.; Li, J. Mammalian drug efflux transporters of the ATP binding cassette (ABC) family in multidrug resistance: A review of the past decade. Cancer Lett. 2016, 370, 153–164. [Google Scholar] [CrossRef] [PubMed]
  87. Serra, M.; Hattinger, C.M.; Pasello, M.; Casotti, C.; Fantoni, L.; Riganti, C.; Manara, M.C. Impact of ABC Transporters in Osteosarcoma and Ewing’s Sarcoma: Which Are Involved in Chemoresistance and Which Are Not? Cells 2021, 10, 2461. [Google Scholar] [CrossRef]
  88. Gonçalves, C.; Martins-Neves, S.R.; Paiva-Oliveira, D.; Oliveira, V.E.B.; Fontes-Ribeiro, C.; Gomes, C.M.F. Sensitizing osteosarcoma stem cells to doxorubicin-induced apoptosis through retention of doxorubicin and modulation of apoptotic-related proteins. Life Sci. 2015, 130, 47–56. [Google Scholar] [CrossRef]
  89. Martins-Neves, S.R.; Paiva-Oliveira, D.I.; Fontes-Ribeiro, C.; Bovée, J.V.M.G.; Cleton-Jansen, A.M.; Gomes, C.M.F. IWR-1, a tankyrase inhibitor, attenuates Wnt/β-catenin signaling in cancer stem-like cells and inhibits in vivo the growth of a subcutaneous human osteosarcoma xenograft. Cancer Lett. 2018, 414, 1–15. [Google Scholar] [CrossRef]
  90. Li, K.; Li, X.; Tian, J.; Wang, H.; Pan, J.; Li, J. Downregulation of DNA-PKcs suppresses P-gp expression via inhibition of the Akt/NF-kB pathway in CD133-positive osteosarcoma MG-63 cells. Oncol. Rep. 2016, 36, 1973–1980. [Google Scholar] [CrossRef]
  91. Lee, Y.H.; Yang, H.W.; Yang, L.C.; Lu, M.Y.; Tsai, L.L.; Yang, S.F.; Huang, Y.F.; Chou, M.Y.; Yu, C.C.; Hu, F.W. DHFR and MDR1 upregulation is associated with chemoresistance in osteosarcoma stem-like cells. Oncol. Lett. 2017, 14, 171–179. [Google Scholar] [CrossRef] [PubMed]
  92. Roundhill, E.A.; Jabri, S.; Burchill, S.A. ABCG1 and Pgp identify drug resistant, self-renewing osteosarcoma cells. Cancer Lett. 2019, 453, 142–157. [Google Scholar] [CrossRef]
  93. Belisario, D.C.; Akman, M.; Godel, M.; Campani, V.; Patrizio, M.P.; Scotti, L.; Hattinger, C.M.; De Rosa, G.; Donadelli, M.; Serra, M.; et al. ABCA1/ABCB1 Ratio Determines Chemo- and Immune-Sensitivity in Human Osteosarcoma. Cells 2020, 9, 647. [Google Scholar] [CrossRef] [PubMed]
  94. Xu, H.; Liu, T.; Li, W.; Yao, Q. SMAR1 attenuates the stemness of osteosarcoma cells via through suppressing ABCG2 transcriptional activity. Environ. Toxicol. 2021. [Google Scholar] [CrossRef] [PubMed]
  95. Liu, Y.; Liu, X.; Yang, S. MicroRNA-221 Upregulates the Expression of P-gp and Bcl-2 by Activating the Stat3 Pathway to Promote Doxorubicin Resistance in Osteosarcoma Cells. Biol. Pharm. Bull. 2021, 44, 861–868. [Google Scholar] [CrossRef] [PubMed]
  96. Lilienthal, I.; Herold, N. Targeting Molecular Mechanisms Underlying Treatment Efficacy and Resistance in Osteosarcoma: A Review of Current and Future Strategies. Int. J. Mol. Sci. 2020, 21, 6885. [Google Scholar] [CrossRef]
  97. Robey, R.W.; Pluchino, K.M.; Hall, M.D.; Fojo, A.T.; Bates, S.E.; Gottesman, M.M. Revisiting the role of ABC transporters in multidrug-resistant cancer. Nat. Rev. Cancer 2018, 18, 452–464. [Google Scholar] [CrossRef]
  98. Hattinger, C.M.; Patrizio, M.P.; Luppi, S.; Serra, M. Pharmacogenomics and Pharmacogenetics in Osteosarcoma: Translational Studies and Clinical Impact. Int. J. Mol. Sci. 2020, 21, 4659. [Google Scholar] [CrossRef]
  99. Belayneh, B.; Weiss, K. The Role of ALDH in the Metastatic Potential of Osteosarcoma Cells and Potential ALDH Targets. Curr. Adv. Exp. Med. Biol. 2020, 1258, 157–166. [Google Scholar] [CrossRef]
  100. Douville, J.; Beaulieu, R.; Balicki, D. ALDH1 as a Functional Marker of Cancer Stem and Progenitor Cells. Stem Cells Dev. 2008, 18, 17–26. [Google Scholar] [CrossRef]
  101. Xu, X.; Chai, S.; Wang, P.; Zhang, C.; Yang, Y.; Yang, Y.; Wang, K. Aldehyde dehydrogenases and cancer stem cells. Cancer Lett. 2015, 369, 50–57. [Google Scholar] [CrossRef]
  102. Lu, J.; Song, G.; Tang, Q.; Yin, J.; Zou, C.; Zhao, Z.; Xie, X.; Xu, H.; Huang, G.; Wang, J.; et al. MiR-26a inhibits stem cell-like phenotype and tumor growth of osteosarcoma by targeting Jagged1. Oncogene 2017, 36, 231–241. [Google Scholar] [CrossRef]
  103. Crasto, J.A.; Fourman, M.S.; Morales-Restrepo, A.; Mahjoub, A.; Mandell, J.B.; Ramnath, K.; Tebbets, J.C.; Watters, R.J.; Weiss, K.R. Disulfiram reduces metastatic osteosarcoma tumor burden in an immunocompetent Balb/c or-thotopic mouse model. Oncotarget 2018, 9, 30163–30172. [Google Scholar] [CrossRef] [PubMed]
  104. Andrade, R.C.; Boroni, M.; Amazonas, M.K.; Vargas, F.R. New drug candidates for osteosarcoma: Drug repurposing based on gene expression signature. Comput. Biol. Med. 2021, 134, 104470. [Google Scholar] [CrossRef] [PubMed]
  105. Sheard, J.J.; Southam, A.D.; MacKay, H.L.; Ellington, M.A.; Snow, M.D.; Khanim, F.L.; Bunce, C.M.; Johnson, W.E. Combined bezafibrate, medroxyprogesterone acetate and valproic acid treatment inhibits osteosarcoma cell growth without adversely affecting normal mesenchymal stem cells. Biosci. Rep. 2021, 41, BSR20202505. [Google Scholar] [CrossRef]
  106. Bahmad, H.F.; Elajami, M.K.; El Zarif, T.; Bou-Gharios, J.; Abou-Antoun, T.; Abou-Kheir, W. Drug repurposing towards targeting cancer stem cells in pediatric brain tumors. Cancer Metastasis Rev. 2020, 39, 127–148. [Google Scholar] [CrossRef] [PubMed]
  107. Ralph, S.J.; Nozuhur, S.; ALHulais, R.A.; Rodríguez-Enríquez, S.; Moreno-Sánchez, R. Repurposing drugs as pro-oxidant redox modifiers to eliminate cancer stem cells and improve the treatment of advanced stage cancers. Med. Res. Rev. 2019, 39, 2397–2426. [Google Scholar] [CrossRef]
  108. Wu, H.; He, Y.; Chen, H.; Liu, Y.; Wei, B.; Chen, G.; Lin, H.; Lin, H. LncRNA THOR increases osteosarcoma cell stemness and migration by enhancing SOX9 mRNA stability. FEBS Open Bio 2019, 9, 781–790. [Google Scholar] [CrossRef]
  109. Phillips, K.L.; Wright, N.; McDermott, E.; Cross, N.A. TRAIL responses are enhanced by nuclear export inhibition in osteosarcoma. BioChem. Biophys. Res. Commun. 2019, 517, 383–389. [Google Scholar] [CrossRef]
  110. Cheng, M.; Duan, P.-G.; Gao, Z.-Z.; Dai, M. MicroRNA-487b-3p inhibits osteosarcoma chemoresistance and metastasis by targeting ALDH1A3. Oncol. Rep. 2020, 44, 2691–2700. [Google Scholar] [CrossRef]
  111. Wang, X.; Li, C.; Yao, W.; Tian, Z.; Liu, Z.; Ge, H. MicroRNA-761 suppresses tumor progression in osteosarcoma via negatively regulating ALDH1B1. Life Sci. 2020, 262, 118544. [Google Scholar] [CrossRef]
  112. Maehara, O.; Suda, G.; Natsuizaka, M.; Ohnishi, S.; Komatsu, Y.; Sato, F.; Nakai, M.; Sho, T.; Morikawa, K.; Ogawa, K.; et al. Fibroblast growth factor-2-mediated FGFR/Erk signaling supports maintenance of cancer stem-like cells in esophageal squamous cell carcinoma. Carcinogenesis 2017, 38, 1073–1083. [Google Scholar] [CrossRef] [Green Version]
  113. Pignochino, Y.; Grignani, G.; Cavalloni, G.; Motta, M.; Tapparo, M.; Bruno, S.; Bottos, A.; Gammaitoni, L.; Migliardi, G.; Camussi, G.; et al. Sorafenib blocks tumour growth, angiogenesis and metastatic potential in preclinical models of osteosarcoma through a mechanism potentially involving the inhibition of ERK1/2, MCL-1 and ezrin pathways. Mol. Cancer 2009, 8, 118. [Google Scholar] [CrossRef] [PubMed]
  114. Moriarity, B.S.; Otto, G.M.; Rahrmann, E.P.; Rathe, S.K.; Wolf, N.K.; Weg, M.T.; Manlove, L.A.; LaRue, R.S.; Temiz, N.A.; Molyneux, S.D.; et al. A Sleeping Beauty forward genetic screen identifies new genes and pathways driving osteosarcoma development and metastasis. Nat. Genet. 2015, 47, 615–624. [Google Scholar] [CrossRef] [PubMed]
  115. Kuijjer, M.L.; van den Akker, B.E.; Hilhorst, R.; Mommersteeg, M.; Buddingh, E.P.; Serra, M.; Bürger, H.; Hogendoorn, P.C.; Cleton-Jansen, A.M. Kinome and mRNA expression profiling of high-grade osteosarcoma cell lines implies Akt signaling as possible target for therapy. BMC Med. Genom. 2014, 7, 4. [Google Scholar] [CrossRef] [PubMed]
  116. Baranski, Z.; Booij, T.H.; Kuijjer, M.L.; de Jong, Y.; Cleton-Jansen, A.M.; Price, L.S.; van de Water, B.; Bovée, J.V.M.G.; Hogendoorn, P.C.; Danen, E.H. MEK inhibition induces apoptosis in osteosarcoma cells with constitutive ERK1/2 phosphorylation. Genes Cancer 2015, 6, 503–512. [Google Scholar] [CrossRef]
  117. Hou, C.H.; Lin, F.L.; Hou, S.M.; Liu, J.F. Cyr61 promotes epithelial-mesenchymal transition and tumor metastasis of osteosarcoma by Raf-1/MEK/ERK/Elk-1/TWIST-1 signaling pathway. Mol. Cancer 2014, 13, 236. [Google Scholar] [CrossRef]
  118. Silva, G.; Aboussekhra, A. p16INK4A inhibits the pro-metastatic potentials of osteosarcoma cells through targeting the ERK pathway and TGF-β1. Mol. Carcinog. 2016, 55, 525–536. [Google Scholar] [CrossRef]
  119. Zhao, S.J.; Shen, Y.F.; Li, Q.; He, Y.J.; Zhang, Y.K.; Hu, L.P.; Jiang, Y.Q.; Xu, N.W.; Wang, Y.J.; Li, J.; et al. SLIT2/ROBO1 axis contributes to the Warburg effect in osteosarcoma through activation of SRC/ERK/c-MYC/PFKFB2 pathway. Cell Death Dis. 2018, 9, 390. [Google Scholar] [CrossRef]
  120. Wang, X.; Zhao, X.; Yi, Z.; Ma, B.; Wang, H.; Pu, Y.; Wang, J.; Wang, S. WNT5A promotes migration and invasion of human osteosarcoma cells via SRC/ERK/MMP-14 pathway. Cell Biol. Int. 2018, 42, 598–607. [Google Scholar] [CrossRef]
  121. Qin, J.; Wang, R.; Zhao, C.; Wen, J.; Dong, H.; Wang, S.; Li, Y.; Zhao, Y.; Li, J.; Yang, Y.; et al. Notch signaling regulates osteosarcoma proliferation and migration through Erk phosphorylation. Tissue Cell 2019, 59, 51–61. [Google Scholar] [CrossRef]
  122. Lin, H.; Hao, Y.; Wan, X.; He, J.; Tong, Y. Baicalein inhibits cell development, metastasis and EMT and induces apoptosis by regulating ERK signaling pathway in osteosarcoma. J. Recept. Signal Transduct. Res. 2020, 40, 49–57. [Google Scholar] [CrossRef]
  123. Yao, J.; Lin, J.; He, L.; Huang, J.; Liu, Q. TNF-a/miR-155 axis induces the transformation of osteosarcoma cancer stem cells independent of TP53INP1. Gene 2020, 726, 144224. [Google Scholar] [CrossRef] [PubMed]
  124. Shimizu, T.; Kimura, K.; Sugihara, E.; Yamaguchi-Iwai, S.; Nobusue, H.; Sampetrean, O.; Otsuki, Y.; Fukuchi, Y.; Saitoh, K.; Kato, K.; et al. MEK inhibition preferentially suppresses anchorage-independent growth in osteosarcoma cells and decreases tumors in vivo. J. Orthop. Res. 2021, 39, 2732–2743. [Google Scholar] [CrossRef] [PubMed]
  125. Chen, Y.; Huang, W.; Sun, W.; Zheng, B.; Wang, C.; Luo, Z.; Wang, J.; Yan, W. LncRNA MALAT1 Promotes Cancer Metastasis in Osteosarcoma via Activation of the PI3K-Akt Signaling Pathway. Cell. Physiol. Biochem. 2018, 51, 1313–1326. [Google Scholar] [CrossRef] [PubMed]
  126. Zhu, J.; Sun, Y.; Lu, Y.; Jiang, X.; Ma, B.; Yu, L.; Zhang, J.; Dong, X.; Zhang, Q. Glaucocalyxin A exerts anticancer effect on osteosarcoma by inhibiting GLI1 nuclear translocation via regulating PI3K/Akt pathway. Cell Death Dis. 2018, 9, 708. [Google Scholar] [CrossRef]
  127. Zhang, Y.; Cheng, H.; Li, W.; Wu, H.; Yang, Y. Highly-expressed P2X7 receptor promotes growth and metastasis of human HOS/MNNG osteosarcoma cells via PI3K/Akt/GSK3b/b-catenin and mTOR/HIF1a/VEGF signaling. Int. J. Cancer 2019, 145, 1068–1082. [Google Scholar] [CrossRef]
  128. Ma, L.; Zhang, L.; Guo, A.; Liu, L.C.; Yu, F.; Diao, N.; Xu, C.; Wang, D. Overexpression of FER1L4 promotes the apoptosis and suppresses epithelial-mesenchymal transition and stemness markers via activating PI3K/AKT signaling pathway in osteosarcoma cells. Pathol. Res. Pract. 2019, 215, 152412. [Google Scholar] [CrossRef]
  129. Ye, F.; Tian, L.; Zhou, Q.; Feng, D. LncRNA FER1L4 induces apoptosis and suppresses EMT and the activation of PI3K/AKT pathway in osteosarcoma cells via inhibiting miR-18a-5p to promote SOCS5. Gene 2019, 721, 144093. [Google Scholar] [CrossRef]
  130. Ma, K.; Zhang, C.; Li, W. Gamabufotalin suppressed osteosarcoma stem cells through the TGF-β/periostin/PI3K/AKT pathway. Chem. Biol. Interact. 2020, 331, 109275. [Google Scholar] [CrossRef]
  131. Ma, K.; Zhang, C.; Li, W. TGF-β is associated with poor prognosis and promotes osteosarcoma progression via PI3K/Akt pathway activation. Cell Cycle 2020, 19, 2327–2339. [Google Scholar] [CrossRef]
  132. Ward, P.; Thompson, C. Metabolic Reprogramming: A Cancer Hallmark Even Warburg Did Not Anticipate. Cancer Cell 2012, 21, 297–308. [Google Scholar] [CrossRef] [Green Version]
  133. Snyder, V.; Reed-Newman, T.C.; Arnold, L.; Thomas, S.M.; Anant, S. Cancer Stem Cell Metabolism and Potential Therapeutic Targets. Front. Oncol. 2018, 8, 203. [Google Scholar] [CrossRef] [PubMed]
  134. Mizushima, E.; Tsukahara, T.; Emori, M.; Murata, K.; Akamatsu, A.; Shibayama, Y.; Hamada, S.; Watanabe, Y.; Kaya, M.; Hirohashi, Y.; et al. Osteosarcoma-initiating cells show high aerobic glycolysis and attenuation of oxidative phosphorylation mediated by LIN28B. Cancer Sci. 2020, 111, 36–46. [Google Scholar] [CrossRef] [PubMed]
  135. Zhong, Z.; Mao, S.; Lin, H.; Li, H.; Lin, J.; Lin, J.M. Alteration of intracellular metabolome in osteosarcoma stem cells revealed by liquid chromatography-tandem mass spectrometry. Talanta 2019, 204, 6–12. [Google Scholar] [CrossRef]
  136. Palorini, R.; Votta, G.; Balestrieri, C.; Monestiroli, A.; Olivieri, S.; Vento, R.; Chiaradonna, F. Energy Metabolism Characterization of a Novel Cancer Stem Cell-Like Line 3AB-OS. J. Cell Biochem. 2014, 115, 368–379. [Google Scholar] [CrossRef] [PubMed]
  137. Della Sala, G.; Pacelli, C.; Agriesti, F.; Laurenzana, I.; Tucci, F.; Tamma, M.; Capitanio, N.; Piccoli, C. Unveiling Metabolic Vulnerability and Plasticity of Human Osteosarcoma Stem and Differentiated Cells to Improve Cancer Therapy. Biomedicines 2021, 10, 28. [Google Scholar] [CrossRef]
  138. Basu-Roy, U.; Han, E.; Rattanakorn, K.; Gadi, A.; Verma, N.; Maurizi, G.; Gunaratne, P.H.; Coarfa, C.; Kennedy, O.D.; Garabedian, M.J.; et al. PPARg agonists promote differentiation of cancer stem cells by restraining YAP transcriptional activity. Oncotarget 2016, 7, 60954–60970. [Google Scholar] [CrossRef]
  139. Hirozane, T.; Masuda, M.; Sugano, T.; Sekita, T.; Goto, N.; Aoyama, T.; Sakagami, T.; Uno, Y.; Moriyama, H.; Sawa, M.; et al. Direct conversion of osteosarcoma to adipocytes by targeting TNIK. JCI Insight 2021, 6, e137245. [Google Scholar] [CrossRef]
  140. Shang, D.; Wu, J.; Guo, L.; Xu, Y.; Liu, L.; Lu, J. Metformin increases sensitivity of osteosarcoma stem cells to cisplatin by inhibiting expression of PKM2. Int. J. Oncol. 2017, 50, 1848–1856. [Google Scholar] [CrossRef]
  141. Paiva-Oliveira, D.I.; Martins-Neves, S.R.; Abrunhosa, A.J.; Fontes-Ribeiro, C.; Gomes, C.M.F. Therapeutic potential of the metabolic modulator Metformin on osteosarcoma cancer stem-like cells. Cancer Chemother. Pharm. 2018, 81, 49–63. [Google Scholar] [CrossRef]
  142. Zhao, B.; Luo, J.; Wang, Y.; Zhou, L.; Che, J.; Wang, F.; Peng, S.; Zhang, G.; Shang, P. Metformin Suppresses Self-Renewal Ability and Tumorigenicity of Osteosarcoma Stem Cells via Reactive Oxygen Species-Mediated Apoptosis and Autophagy. Oxid. Med. Cell. Longev. 2019, 2019, 9290728. [Google Scholar] [CrossRef]
  143. Deguchi, T.; Hosoya, K.; Kim, S.; Murase, Y.; Yamamoto, K.; Bo, T.; Yasui, H.; Inanami, O.; Okumura, M. Metformin preferentially enhances the radio-sensitivity of cancer stem-like cells with highly mitochondrial respiration ability in HMPOS. Mol. Ther. Oncolytics 2021, 22, 143–151. [Google Scholar] [CrossRef] [PubMed]
  144. Almog, N. Molecular mechanisms underlying tumor dormancy. Cancer Lett. 2010, 294, 139–146. [Google Scholar] [CrossRef] [PubMed]
  145. Cheung, T.H.; Rando, T.A. Molecular regulation of stem cell quiescence. Nat. Rev. Mol. Cell Biol. 2013, 14, 329–340. [Google Scholar] [CrossRef]
  146. Moore, N.; Lyle, S. Quiescent, Slow-Cycling Stem Cell Populations in Cancer: A Review of the Evidence and Discussion of Significance. J. Oncol. 2011, 2011, 396076. [Google Scholar] [CrossRef] [PubMed]
  147. Almog, N.; Ma, L.; Schwager, C.; Brinkmann, B.G.; Beheshti, A.; Vajkoczy, P.; Folkman, J.; Hlatky, L.; Abdollahi, A. Consensus micro RNAs governing the switch of dormant tumors to the fast-growing angiogenic phenotype. PLoS ONE 2012, 7, e44001. [Google Scholar] [CrossRef] [PubMed]
  148. Shimizu, T.; Sugihara, E.; Yamaguchi-Iwai, S.; Tamaki, S.; Koyama, Y.; Kamel, W.; Ueki, A.; Ishikawa, T.; Chiyoda, T.; Osuka, S.; et al. IGF2 Preserves Osteosarcoma Cell Survival by Creating an Autophagic State of Dormancy That Protects Cells against Chemotherapeutic Stress. Cancer Res. 2014, 74, 6531–6541. [Google Scholar] [CrossRef]
  149. Avril, P.; Le Nail, L.R.; Brennan, M.Á.; Rosset, P.; De Pinieux, G.; Layrolle, P.; Heymann, D.; Perrot, P.; Trichet, V. Mesenchymal stem cells increase proliferation but do not change quiescent state of osteosarcoma cells: Potential implications according to the tumor resection status. J. Bone Oncol. 2015, 5, 5–14. [Google Scholar] [CrossRef]
  150. Jiang, W.; Liu, J.; Xu, T.; Yu, X. MiR-329 suppresses osteosarcoma development by downregulating Rab10. FEBS Lett. 2016, 590, 2973–2981. [Google Scholar] [CrossRef]
  151. Tiram, G.; Segal, E.; Krivitsky, A.; Shreberk-Hassidim, R.; Ferber, S.; Ofek, P.; Udagawa, T.; Edry, L.; Shomron, N.; Roniger, M.; et al. Identification of Dormancy-Associated MicroRNAs for the Design of Osteosarcoma-Targeted Dendritic Polyglycerol Nanopolyplexes. ACS Nano 2016, 10, 2028–2045. [Google Scholar] [CrossRef]
  152. Guo, P.; Huang, J.; Moses, M.A. Characterization of dormant and active human cancer cells by quantitative phase imaging. Cytom. Part A 2017, 91, 424–432. [Google Scholar] [CrossRef] [Green Version]
  153. Dosch, J.; Hadley, E.; Wiese, C.; Soderberg, M.; Houwman, T.; Ding, K.; Kharazova, A.; Collins, J.L.; van Knippenberg, B.; Gregory, C.; et al. Time-lapse microscopic observation of non-dividing cells in cultured human osteosarcoma MG-63 cell line. Cell Cycle 2018, 17, 174–181. [Google Scholar] [CrossRef]
  154. Maugeri-Saccà, M.; Bartucci, M.; De Maria, R. DNA Damage Repair Pathways in Cancer Stem Cells. Mol. Cancer Ther. 2012, 11, 1627–1636. [Google Scholar] [CrossRef]
  155. Damia, G.; D’Incalci, M. Targeting DNA repair as a promising approach in cancer therapy. Eur. J. Cancer 2007, 43, 1791–1801. [Google Scholar] [CrossRef] [PubMed]
  156. Zhu, Y.; Hu, J.; Hu, Y.; Liu, W. Targeting DNA repair pathways: A novel approach to reduce cancer therapeutic resistance. Cancer Treat. Rev. 2009, 35, 590–596. [Google Scholar] [CrossRef] [PubMed]
  157. O’Connor, M. Targeting the DNA Damage Response in Cancer. Mol. Cell 2015, 60, 547–560. [Google Scholar] [CrossRef] [PubMed]
  158. Wang, Q.E. DNA damage responses in cancer stem cells: Implications for cancer therapeutic strategies. World J. Biol. Chem. 2015, 6, 57–64. [Google Scholar] [CrossRef]
  159. Du, L.Q.; Wang, Y.; Wang, H.; Cao, J.; Liu, Q.; Fan, F.Y. Knockdown of Rad51 expression induces radiation- and chemo-sensitivity in osteosarcoma cells. Med. Oncol. 2011, 28, 1481–1487. [Google Scholar] [CrossRef]
  160. Biason, P.; Hattinger, C.M.; Innocenti, F.; Talamini, R.; Alberghini, M.; Scotlandi, K.; Zanusso, C.; Serra, M.; Toffoli, G. Nucleotide excision repair gene variants and association with survival in osteosarcoma patients treated with neoadjuvant chemotherapy. Pharm. J. 2012, 12, 476–483. [Google Scholar] [CrossRef]
  161. Engert, F.; Kovac, M.; Baumhoer, D.; Nathrath, M.; Fulda, S. Osteosarcoma cells with genetic signatures of BRCAness are susceptible to the PARP inhibitor talazoparib alone or in combination with chemotherapeutics. Oncotarget 2016. [Google Scholar] [CrossRef]
  162. Zhang, Z.; Ha, S.H.; Moon, Y.J.; Hussein, U.K.; Song, Y.; Kim, K.M.; Park, S.H.; Park, H.S.; Park, B.H.; Ahn, A.R.; et al. Inhibition of SIRT6 potentiates the anti-tumor effect of doxorubicin through suppression of the DNA damage repair pathway in osteosarcoma. J Exp. Clin. Cancer Res. 2020, 39, 247. [Google Scholar] [CrossRef]
  163. Moukengue, B.; Brown, H.K.; Charrier, C.; Battaglia, S.; Baud’huin, M.; Quillard, T.; Pham, T.M.; Pateras, I.S.; Gorgoulis, V.G.; Helleday, T.; et al. TH1579, MTH1 inhibitor, delays tumour growth and inhibits metastases development in osteosarcoma model. EBioMedicine 2020, 53, 102704. [Google Scholar] [CrossRef]
  164. Xu, L.; Sun, Z.; Wei, X.; Tan, H.; Kong, P.; Li, Z.; Yang, Q.; Dai, E.; Li, J. The inhibition of MARK2 suppresses cisplatin resistance of osteosarcoma stem cells by regulating DNA damage and repair. J. Bone Oncol. 2020, 23, 100290. [Google Scholar] [CrossRef]
  165. Zhu, J.; Zou, H.; Yu, W.; Huang, Y.; Liu, B.; Li, T.; Liang, C.; Tao, H. Checkpoint kinase inhibitor AZD7762 enhance cisplatin-induced apoptosis in osteosarcoma cells. Cancer Cell Int. 2019, 19, 195. [Google Scholar] [CrossRef] [PubMed]
  166. Fanelli, M.; Tavanti, E.; Patrizio, M.P.; Vella, S.; Fernandez-Ramos, A.; Magagnoli, F.; Luppi, S.; Hattinger, C.M.; Serra, M. Cisplatin Resistance in Osteosarcoma: In vitro Validation of Candidate DNA Repair-Related Therapeutic Targets and Drugs for Tailored Treatments. Front. Oncol. 2020, 10, 331. [Google Scholar] [CrossRef] [PubMed]
  167. Godel, M.; Morena, D.; Ananthanarayanan, P.; Buondonno, I.; Ferrero, G.; Hattinger, C.M.; Di Nicolantonio, F.; Serra, M.; Taulli, R.; Cordero, F.; et al. Small Nucleolar RNAs Determine Resistance to Doxorubicin in Human Osteosarcoma. Int. J. Mol. Sci. 2020, 21, 4500. [Google Scholar] [CrossRef] [PubMed]
  168. Xu, A.; Huang, M.F.; Zhu, D.; Gingold, J.A.; Bazer, D.A.; Chang, B.; Wang, D.; Lai, C.C.; Lemischka, I.R.; Zhao, R.; et al. LncRNA H19 Suppresses Osteosarcomagenesis by Regulating snoRNAs and DNA Repair Protein Complexes. Front. Genet. 2021, 11, 1773. [Google Scholar] [CrossRef]
  169. Hattinger, C.M.; Patrizio, M.P.; Luppi, S.; Magagnoli, F.; Picci, P.; Serra, M. Current understanding of pharmacogenetic implications of DNA damaging drugs used in osteosarcoma treatment. Expert Opin. Drug Metab. Toxicol. 2019, 15, 299–311. [Google Scholar] [CrossRef]
  170. Plati, J.; Bucur, O.; Khosravi-Far, R. Apoptotic cell signaling in cancer progression and therapy. Integr. Biol. 2011, 3, 279–296. [Google Scholar] [CrossRef]
  171. Verena, J. The Intrinsic Apoptosis Pathways as a Target in Anticancer Therapy. Curr Pharm. Biotechnol. 2012, 13, 1426–1438. [Google Scholar]
  172. Baig, S.; Seevasant, I.; Mohamad, J.; Mukheem, A.; Huri, H.Z.; Kamarul, T. Potential of apoptotic pathway-targeted cancer therapeutic research: Where do we stand? Cell Death Dis. 2016, 7, e2058. [Google Scholar] [CrossRef]
  173. Ashkenazi, A. Targeting the extrinsic apoptotic pathway in cancer: Lessons learned and future directions. J. Clin. Investig. 2015, 125, 487–489. [Google Scholar] [CrossRef]
  174. Adams, J.M.; Cory, S. The Bcl-2 apoptotic switch in cancer development and therapy. Oncogene 2007, 26, 1324–1337. [Google Scholar] [CrossRef] [PubMed]
  175. Elmore, S. Apoptosis: A Review of Programmed Cell Death. Toxicol. Pathol. 2007, 35, 495–516. [Google Scholar] [CrossRef] [PubMed]
  176. Ferrari, S.; Bertoni, F.; Zanella, L.; Setola, E.; Bacchini, P.; Alberghini, M.; Versari, M.; Bacci, G. Evaluation of P-glycoprotein, HER-2/ErbB-2, p53, and Bcl-2 in primary tumor and metachronous lung metastases in patients with high-grade osteosarcoma. Cancer 2004, 100, 1936–1942. [Google Scholar] [CrossRef] [PubMed]
  177. Wang, Z.X.; Yang, J.S.; Pan, X.; Wang, J.R.; Li, J.; Yin, Y.M.; De, W. Functional and biological analysis of Bcl-xL expression in human osteosarcoma. Bone 2010, 47, 445–454. [Google Scholar] [CrossRef] [PubMed]
  178. Trieb, K.; Lehner, R.; Stulnig, T.; Sulzbacher, I.; Shroyer, K.R. Survivin expression in human osteosarcoma is a marker for survival. Eur. J. Surg. Oncol. 2003, 29, 379–382. [Google Scholar] [CrossRef] [PubMed]
  179. Wu, X.; Cai, Z.d.; Lou, L.m.; Zhu, Y.b. Expressions of p53, c-MYC, BCL-2 and apoptotic index in human osteosarcoma and their correlations with prognosis of patients. Cancer Epidemiol. 2012, 36, 212–216. [Google Scholar] [CrossRef]
  180. Eliseev, R.A.; Dong, Y.F.; Sampson, E.; Zuscik, M.J.; Schwarz, E.M.; O’Keefe, R.J.; Rosier, R.N.; Drissi, M.H. Runx2-mediated activation of the Bax gene increases osteosarcoma cell sensitivity to apoptosis. Oncogene 2008, 27, 3605–3614. [Google Scholar] [CrossRef]
  181. Wang, G.; Rong, J.; Zhou, Z.; Duo, J. A novel gene P28GANK confers multidrug resistance by modulating the expression of MDR-1, Bcl-2, and bax in osteosarcoma cells. Mol. Biol. 2010, 44, 898–906. [Google Scholar] [CrossRef]
  182. Fulda, S.; Pervaiz, S. Apoptosis signaling in cancer stem cells. Int. J. Biochem. Cell Biol. 2010, 42, 31–38. [Google Scholar] [CrossRef]
  183. Fulda, S. Regulation of apoptosis pathways in cancer stem cells. Cancer Lett. 2013, 338, 168–173. [Google Scholar] [CrossRef] [PubMed]
  184. Tagscherer, K.E.; Fassl, A.; Campos, B.; Farhadi, M.; Kraemer, A.; Bock, B.C.; Macher-Goeppinger, S.; Radlwimmer, B.; Wiestler, O.D.; Herold-Mende, C.; et al. Apoptosis-based treatment of glioblastomas with ABT-737, a novel small molecule inhibitor of Bcl-2 family proteins. Oncogene 2008, 27, 6646–6656. [Google Scholar] [CrossRef] [PubMed]
  185. Beurlet, S.; Omidvar, N.; Gorombei, P.; Krief, P.; Le Pogam, C.; Setterblad, N.; de la Grange, P.; Leboeuf, C.; Janin, A.; Noguera, M.E.; et al. BCL-2 inhibition with ABT-737 prolongs survival in an NRAS/BCL-2 mouse model of AML by targeting primitive LSK and progenitor cells. Blood 2013, 122, 2864–2876. [Google Scholar] [CrossRef] [PubMed]
  186. Baev, D.V.; Krawczyk, J.; O’Dwyer, M.; Szegezdi, E. The BH3-mimetic ABT-737 effectively kills acute myeloid leukemia initiating cells. Leuk. Res. Rep. 2014, 3, 79–82. [Google Scholar] [CrossRef] [PubMed]
  187. Zhang, F.; Yu, X.; Liu, X.; Zhou, T.; Nie, T.; Cheng, M.; Liu, H.; Dai, M.; Zhang, B. ABT-737 potentiates cisplatin-induced apoptosis in human osteosarcoma cells via the mitochondrial apoptotic pathway. Oncol. Rep. 2017, 38, 2301–2308. [Google Scholar] [CrossRef] [PubMed]
  188. Masuelli, L.; Benvenuto, M.; Izzi, V.; Zago, E.; Mattera, R.; Cerbelli, B.; Potenza, V.; Fazi, S.; Ciuffa, S.; Tresoldi, I.; et al. In vivo and in vitro inhibition of osteosarcoma growth by the pan Bcl-2 inhibitor AT-101. Investig. New Drugs 2020, 38, 675–689. [Google Scholar] [CrossRef]
  189. Kehr, S.; Haydn, T.; Bierbrauer, A.; Irmer, B.; Vogler, M.; Fulda, S. Targeting BCL-2 proteins in pediatric cancer: Dual inhibition of BCL-XL and MCL-1 leads to rapid induction of intrinsic apoptosis. Cancer Lett. 2020, 482, 19–32. [Google Scholar] [CrossRef]
  190. Zeuner, A.; Francescangeli, F.; Contavalli, P.; Zapparelli, G.; Apuzzo, T.; Eramo, A.; Baiocchi, M.; De Angelis, M.L.; Biffoni, M.; Sette, G.; et al. Elimination of quiescent/slow-proliferating cancer stem cells by Bcl-XL inhibition in non-small cell lung cancer. Cell Death Differ. 2014, 21, 1877–1888. [Google Scholar] [CrossRef]
  191. Vallabhaneni, K.C.; Hassler, M.Y.; Abraham, A.; Whitt, J.; Mo, Y.Y.; Atfi, A.; Pochampally, R. Mesenchymal Stem/Stromal Cells under Stress Increase Osteosarcoma Migration and Apoptosis Resistance via Extracellular Vesicle Mediated Communication. PLoS ONE 2016, 11, e0166027. [Google Scholar] [CrossRef]
  192. Pietrovito, L.; Leo, A.; Gori, V.; Lulli, M.; Parri, M.; Becherucci, V.; Piccini, L.; Bambi, F.; Taddei, M.L.; Chiarugi, P. Bone marrow-derived mesenchymal stem cells promote invasiveness and transendothelial migration of osteosarcoma cells via a mesenchymal to amoeboid transition. Mol. Oncol. 2018, 12, 659–676. [Google Scholar] [CrossRef]
  193. Wang, Y.M.; Wang, W.; Qiu, E.D. Osteosarcoma cells induce differentiation of mesenchymal stem cells into cancer associated fibroblasts through Notch and Akt signaling pathway. Int. J. Clin. Exp. Pathol. 2017, 10, 8479–8486. [Google Scholar] [PubMed]
  194. Yang, J.; Hu, Y.; Wang, L.; Sun, X.; Yu, L.; Guo, W. Human umbilical vein endothelial cells derived-exosomes promote osteosarcoma cell stemness by activating Notch signaling pathway. Bioengineered 2021, 12, 11007–11017. [Google Scholar] [CrossRef] [PubMed]
  195. Philip, B.; Ito, K.; Moreno-Sánchez, R.; Ralph, S.J. HIF expression and the role of hypoxic microenvironments within primary tumours as protective sites driving cancer stem cell renewal and metastatic progression. Carcinogenesis 2013, 34, 1699–1707. [Google Scholar] [CrossRef] [PubMed]
  196. Zhao, H.; Wu, Y.; Chen, Y.; Liu, H. Clinical significance of hypoxia-inducible factor 1 and VEGF-A in osteosarcoma. Int. J. Clin. Oncol. 2015, 20, 1233–1243. [Google Scholar] [CrossRef]
  197. Pierrevelcin, M.; Fuchs, Q.; Lhermitte, B.; Messé, M.; Guérin, E.; Weingertner, N.; Martin, S.; Lelong-Rebel, I.; Nazon, C.; Dontenwill, M.; et al. Focus on Hypoxia-Related Pathways in Pediatric Osteosarcomas and Their Druggability. Cells 2020, 9, 1998. [Google Scholar] [CrossRef]
  198. El Naggar, A.; Clarkson, P.; Zhang, F.; Mathers, J.; Tognon, C.; Sorensen, P.H. Expression and stability of hypoxia inducible factor 1α in osteosarcoma. Pediatr. Blood Cancer 2012, 59, 1215–1222. [Google Scholar] [CrossRef]
  199. Scholten, D.J.; Timmer, C.M.; Peacock, J.D.; Pelle, D.W.; Williams, B.O.; Steensma, M.R. Down Regulation of Wnt Signaling Mitigates Hypoxia-Induced Chemoresistance in Human Osteosarcoma Cells. PLoS ONE 2014, 9, e111431. [Google Scholar] [CrossRef]
  200. Gorgun, C.; Ozturk, S.; Gokalp, S.; Vatansever, S.; Gurhan, S.I.D.; Urkmez, A.S. Synergistic role of three dimensional niche and hypoxia on conservation of cancer stem cell phenotype. Int. J. Biol. Macromol. 2016, 90, 20–26. [Google Scholar] [CrossRef]
  201. Liapis, V.; Labrinidis, A.; Zinonos, I.; Hay, S.; Ponomarev, V.; Panagopoulos, V.; DeNichilo, M.; Ingman, W.; Atkins, G.J.; Findlay, D.M.; et al. Hypoxia-activated pro-drug TH-302 exhibits potent tumor suppressive activity and cooperates with chemotherapy against osteosarcoma. Cancer Lett. 2015, 357, 160–169. [Google Scholar] [CrossRef]
  202. Guo, M.; Cai, C.; Zhao, G.; Qiu, X.; Zhao, H.; Ma, Q.; Tian, L.; Li, X.; Hu, Y.; Liao, B.; et al. Hypoxia Promotes Migration and Induces CXCR4 Expression via HIF-1α Activation in Human Osteosarcoma. PLoS ONE 2014, 9, e90518. [Google Scholar] [CrossRef]
  203. Guan, G.; Zhang, Y.; Lu, Y.; Liu, L.; Shi, D.; Wen, Y.; Yang, L.; Ma, Q.; Liu, T.; Zhu, X.; et al. The HIF-1α/CXCR4 pathway supports hypoxia-induced metastasis of human osteosarcoma cells. Cancer Lett. 2015, 357, 254–264. [Google Scholar] [CrossRef] [PubMed]
  204. Zhao, D.; Wang, S.; Chu, X.; Han, D. LncRNA HIF2PUT inhibited osteosarcoma stem cells proliferation, migration and invasion by regulating HIF2 expression. Artif. Cells Nanomed. Biotechnol. 2019, 47, 1342–1348. [Google Scholar] [CrossRef] [PubMed]
  205. Li, W.; He, X.; Xue, R.; Zhang, Y.; Zhang, X.; Lu, J.; Zhang, Z.; Xue, L. Combined over-expression of the hypoxia-inducible factor 2a gene and its long non-coding RNA predicts unfavorable prognosis of patients with osteosarcoma. Pathol. Res. Pract. 2016, 212, 861–866. [Google Scholar] [CrossRef] [PubMed]
  206. Lin, S.; Zhu, B.; Huang, G.; Zeng, Q.; Wang, C. Microvesicles derived from human bone marrow mesenchymal stem cells promote U2OS cell growth under hypoxia: The role of PI3K/AKT and HIF-1α. Hum. Cell 2019, 32, 64–74. [Google Scholar] [CrossRef]
  207. Lin, J.; Wang, X.; Wang, X.; Wang, S.; Shen, R.; Yang, Y.; Xu, J.; Lin, J. Hypoxia increases the expression of stem cell markers in human osteosarcoma cells. Oncol. Lett. 2021, 21, 217. [Google Scholar] [CrossRef] [PubMed]
  208. Wagner, F.; Holzapfel, B.M.; Martine, L.C.; McGovern, J.; Lahr, C.A.; Boxberg, M.; Prodinger, P.M.; Gr+ñssel, S.; Loessner, D.; Hutmacher, D.W. A humanized bone microenvironment uncovers HIF2 alpha as a latent marker for osteosarcoma. Acta Biomater. 2019, 89, 372–381. [Google Scholar] [CrossRef]
  209. Zhao, W.; Qin, P.; Zhang, D.; Cui, X.; Gao, J.; Yu, Z.; Chai, Y.; Wang, J.; Li, J. Long non-coding RNA PVT1 encapsulated in bone marrow mesenchymal stem cell-derived exosomes promotes osteosarcoma growth and metastasis by stabilizing ERG and sponging miR-183-5p. Aging 2019, 11, 9581–9596. [Google Scholar] [CrossRef]
  210. Qin, F.; Tang, H.; Zhang, Y.; Zhang, Z.; Huang, P.; Zhu, J. Bone marrow-derived mesenchymal stem cell-derived exosomal microRNA-208a promotes osteosarcoma cell proliferation, migration, and invasion. J. Cell Physiol. 2020, 235, 4734–4745. [Google Scholar] [CrossRef]
  211. Zheng, Y.; Wang, G.; Chen, R.; Hua, Y.; Cai, Z. Mesenchymal stem cells in the osteosarcoma microenvironment: Their biological properties, influence on tumor growth, and therapeutic implications. Stem Cell Res. Ther. 2018, 9, 22. [Google Scholar] [CrossRef]
  212. Chang, X.; Ma, Z.; Zhu, G.; Lu, Y.; Yang, J. New perspective into mesenchymal stem cells: Molecular mechanisms regulating osteosarcoma. J. Bone Oncol. 2021, 29, 100372. [Google Scholar] [CrossRef]
  213. Liu, T.; Zhou, L.; Li, D.; Andl, T.; Zhang, Y. Cancer-Associated Fibroblasts Build and Secure the Tumor Microenvironment. Front. Cell Dev. Biol. 2019, 7, 60. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Belli, C.; Trapani, D.; Viale, G.; D’Amico, P.; Duso, B.A.; Della Vigna, P.; Orsi, F.; Curigliano, G. Targeting the microenvironment in solid tumors. Cancer Treat. Rev. 2018, 65, 22–32. [Google Scholar] [CrossRef] [PubMed]
  215. Cortini, M.; Massa, A.; Avnet, S.; Bonuccelli, G.; Baldini, N. Tumor-Activated Mesenchymal Stromal Cells Promote Osteosarcoma Stemness and Migratory Potential via IL-6 Secretion. PLoS ONE 2016, 11, e0166500. [Google Scholar] [CrossRef] [PubMed]
  216. Ma, Y.; Ren, Y.; Dai, Z.J.; Wu, C.J.; Ji, Y.H.; Xu, J. IL-6, IL-8 and TNF-a levels correlate with disease stage in breast cancer patients. Adv. Clin. Exp. Med. 2017, 26, 421–426. [Google Scholar] [CrossRef] [PubMed]
  217. Balamurugan, K.; Mendoza-Villanueva, D.; Sharan, S.; Summers, G.H.; Dobrolecki, L.E.; Lewis, M.T.; Sterneck, E. C/EBPd links IL-6 and HIF-1 signaling to promote breast cancer stem cell-associated phenotypes. Oncogene 2019, 38, 3765–3780. [Google Scholar] [CrossRef]
  218. Kern, L.; Mittenbühler, M.J.; Vesting, A.J.; Ostermann, A.L.; Wunderlich, C.M.; Wunderlich, F.T. Obesity-Induced TNFa and IL-6 Signaling: The Missing Link between Obesity and Inflammation-Driven Liver and Colorectal Cancers. Cancers 2019, 11, 24. [Google Scholar] [CrossRef]
  219. Rose-John, S. Local and systemic effects of interleukin-6 (IL-6) in inflammation and cancer. FEBS Lett. 2022, 596, 557–566. [Google Scholar] [CrossRef]
  220. Hoesel, B.; Schmid, J.A. The complexity of NF-kB signaling in inflammation and cancer. Mol. Cancer 2013, 12, 86. [Google Scholar] [CrossRef]
  221. Han, X.; Liu, F.; Zhang, C.; Ren, Z.; Li, L.; Wang, G. SIAH1/ZEB1/IL-6 axis is involved in doxorubicin (Dox) resistance of osteosarcoma cells. Biol. Chem. 2019, 400, 545–553. [Google Scholar] [CrossRef]
  222. Umemura, N.; Sugimoto, M.; Kitoh, Y.; Saio, M.; Sakagami, H. Metabolomic profiling of tumor-infiltrating macrophages during tumor growth. Cancer Immunol. Immunother. 2020, 69, 2357–2369. [Google Scholar] [CrossRef]
  223. Cassetta, L.; Fragkogianni, S.; Sims, A.H.; Swierczak, A.; Forrester, L.M.; Zhang, H.; Soong, D.Y.H.; Cotechini, T.; Anur, P.; Lin, E.Y.; et al. Human Tumor-Associated Macrophage and Monocyte Transcriptional Landscapes Reveal Cancer-Specific Reprogramming, Biomarkers, and Therapeutic Targets. Cancer Cell 2019, 35, 588–602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Monteran, L.; Erez, N. The Dark Side of Fibroblasts: Cancer-Associated Fibroblasts as Mediators of Immunosuppression in the Tumor Microenvironment. Front. Immunol. 2019, 10, 1835. [Google Scholar] [CrossRef] [PubMed]
  225. Barnas, J.L.; Simpson-Abelson, M.R.; Brooks, S.P.; Kelleher, R.J.; Bankert, R.B. Reciprocal Functional Modulation of the Activation of T Lymphocytes and Fibroblasts Derived from Human Solid Tumors. J. Immunol. 2010, 185, 2681–2692. [Google Scholar] [CrossRef] [PubMed]
  226. Tu, B.; Peng, Z.X.; Fan, Q.m.; Du, L.; Yan, W.; Tang, T.t. Osteosarcoma cells promote the production of pro-tumor cytokines in mesenchymal stem cells by inhibiting their osteogenic differentiation through the TGF-b/Smad2/3 pathway. Exp. Cell Res. 2014, 320, 164–173. [Google Scholar] [CrossRef]
  227. Angioni, R.; Sánchez-Rodríguez, R.; Viola, A.; Molon, B. TGF-b in Cancer: Metabolic Driver of the Tolerogenic Crosstalk in the Tumor Microenvironment. Cancers 2021, 13, 401. [Google Scholar] [CrossRef]
  228. Hamilton, P.T.; Anholt, B.R.; Nelson, B.H. Tumour immunotherapy: Lessons from predator-prey theory. Nat. Rev. Immunol. 2022, 1–11. [Google Scholar] [CrossRef]
  229. Gill, J.; Gorlick, R. Advancing therapy for osteosarcoma. Nat. Rev. Clin. Oncol. 2021, 18, 609–624. [Google Scholar] [CrossRef]
  230. Chen, C.; Xie, L.; Ren, T.; Huang, Y.; Xu, J.; Guo, W. Immunotherapy for osteosarcoma: Fundamental mechanism, rationale, and recent breakthroughs. Cancer Lett. 2021, 500, 1–10. [Google Scholar] [CrossRef]
  231. Lu, Y.; Zhang, J.; Chen, Y.; Kang, Y.; Liao, Z.; He, Y.; Zhang, C. Novel Immunotherapies for Osteosarcoma. Front. Oncol. 2022, 12, 546. [Google Scholar] [CrossRef]
  232. Wu, C.C.; Beird, H.C.; Andrew Livingston, J.; Advani, S.; Mitra, A.; Cao, S.; Reuben, A.; Ingram, D.; Wang, W.L.; Ju, Z.; et al. Immuno-genomic landscape of osteosarcoma. Nat. Commun. 2020, 11, 1008. [Google Scholar] [CrossRef]
  233. Ma, L.; Dichwalkar, T.; Chang, J.Y.H.; Cossette, B.; Garafola, D.; Zhang, A.Q.; Fichter, M.; Wang, C.; Liang, S.; Silva, M.; et al. Enhanced CAR-T cell activity against solid tumors by vaccine boosting through the chimeric receptor. Science 2019, 365, 162–168. [Google Scholar] [CrossRef] [PubMed]
  234. Alsaab, H.O.; Sau, S.; Alzhrani, R.; Tatiparti, K.; Bhise, K.; Kashaw, S.K.; Iyer, A.K. PD-1 and PD-L1 Checkpoint Signaling Inhibition for Cancer Immunotherapy: Mechanism, Combinations, and Clinical Outcome. Front. Pharmacol. 2017, 8, 561. [Google Scholar] [CrossRef] [PubMed]
  235. Tawbi, H.A.; Burgess, M.; Bolejack, V.; Van Tine, B.A.; Schuetze, S.M.; Hu, J.; D’Angelo, S.; Attia, S.; Riedel, R.F.; Priebat, D.A.; et al. Pembrolizumab in advanced soft-tissue sarcoma and bone sarcoma (SARC028): A multicentre, two-cohort, single-arm, open-label, phase 2 trial. Lancet Oncol. 2017, 18, 1493–1501. [Google Scholar] [CrossRef]
  236. Gao, W.; Zhou, J.; Ji, B. Evidence of Interleukin 21 Reduction in Osteosarcoma Patients Due to PD-1/PD-L1-Mediated Suppression of Follicular Helper T Cell Functionality. DNA Cell Biol. 2017, 36, 794–800. [Google Scholar] [CrossRef] [PubMed]
  237. Mochizuki, Y.; Tazawa, H.; Demiya, K.; Kure, M.; Kondo, H.; Komatsubara, T.; Sugiu, K.; Hasei, J.; Yoshida, A.; Kunisada, T.; et al. Telomerase-specific oncolytic immunotherapy for promoting efficacy of PD-1 blockade in osteosarcoma. Cancer Immunol. Immunother. 2021, 70, 1405–1417. [Google Scholar] [CrossRef] [PubMed]
  238. Lussier, D.M.; Johnson, J.L.; Hingorani, P.; Blattman, J.N. Combination immunotherapy with alpha-CTLA-4 and alpha-PD-L1 antibody blockade prevents immune escape and leads to complete control of metastatic osteosarcoma. J. Immunother. Cancer 2015, 3, 21. [Google Scholar] [CrossRef]
  239. Helm, A.; Tinganelli, W.; Simoniello, P.; Kurosawa, F.; Fournier, C.; Shimokawa, T.; Durante, M. Reduction of Lung Metastases in a Mouse Osteosarcoma Model Treated With Carbon Ions and Immune Checkpoint Inhibitors. Int. J. Radiat. Oncol. Biol. Phys. 2021, 109, 594–602. [Google Scholar] [CrossRef]
  240. Köksal, H.; Müller, E.; Inderberg, E.M.; Bruland, O.; Wälchli, S. Treating osteosarcoma with CAR T cells. Scand. J. Immunol. 2019, 89, e12741. [Google Scholar] [CrossRef]
  241. Lin, Z.; Wu, Z.; Luo, W. Chimeric Antigen Receptor T-Cell Therapy: The Light of Day for Osteosarcoma. Cancers 2021, 13, 4469. [Google Scholar] [CrossRef]
  242. Talbot, L.J.; Chabot, A.; Funk, A.; Nguyen, P.; Wagner, J.; Ross, A.; Tillman, H.; Davidoff, A.; Gottschalk, S.; DeRenzo, C. A Novel Orthotopic Implantation Technique for Osteosarcoma Produces Spontaneous Metastases and Illustrates Dose-Dependent Efficacy of B7-H3-CAR T Cells. Front. Immunol. 2021, 12, 691741. [Google Scholar] [CrossRef]
  243. Kiru, L.; Zlitni, A.; Tousley, A.M.; Dalton, G.N.; Wu, W.; Lafortune, F.; Liu, A.; Cunanan, K.M.; Nejadnik, H.; Sulchek, T.; et al. In vivo imaging of nanoparticle-labeled CAR T cells. Proc. Natl. Acad. Sci. USA 2022, 119, e2102363119. [Google Scholar] [CrossRef] [PubMed]
  244. Narkhede, M.; Mehta, A.; Ansell, S.M.; Goyal, G. CAR T-cell therapy in mature lymphoid malignancies: Clinical opportunities and challenges. Ann. Transl. Med. 2021, 9, 1036. [Google Scholar] [CrossRef] [PubMed]
  245. Patel, U.; Abernathy, J.; Savani, B.N.; Oluwole, O.; Sengsayadeth, S.; Dholaria, B. CAR T cell therapy in solid tumors: A review of current clinical trials. EJHaem 2022, 3, 24–31. [Google Scholar] [CrossRef]
  246. Balta, E.; Wabnitz, G.H.; Samstag, Y. Hijacked Immune Cells in the Tumor Microenvironment: Molecular Mechanisms of Immunosuppression and Cues to Improve T Cell-Based Immunotherapy of Solid Tumors. Int. J. Mol. Sci. 2021, 22, 5736. [Google Scholar] [CrossRef] [PubMed]
  247. Miranda, A.; Hamilton, P.T.; Zhang, A.W.; Pattnaik, S.; Becht, E.; Mezheyeuski, A.; Bruun, J.; Micke, P.; de Reynies, A.; Nelson, B.H. Cancer stemness, intratumoral heterogeneity, and immune response across cancers. Proc. Natl. Acad. Sci. USA 2019, 116, 9020–9029. [Google Scholar] [CrossRef] [PubMed]
  248. Dianat-Moghadam, H.; Mahari, A.; Salahlou, R.; Khalili, M.; Azizi, M.; Sadeghzadeh, H. Immune evader cancer stem cells direct the perspective approaches to cancer immunotherapy. Stem Cell Res. Ther. 2022, 13, 150. [Google Scholar] [CrossRef] [PubMed]
  249. Guo, L.; Yan, T.; Guo, W.; Niu, J.; Wang, W.; Ren, T.; Huang, Y.; Xu, J.; Wang, B. Molecular subtypes of osteosarcoma classified by cancer stem cell related genes define immunological cell infiltration and patient survival. Front. Immunol. 2022, 13, 986785. [Google Scholar] [CrossRef]
  250. Rainusso, N.; Brawley, V.S.; Ghazi, A.; Hicks, M.J.; Gottschalk, S.; Rosen, J.M.; Ahmed, N. Immunotherapy targeting HER2 with genetically modified T cells eliminates tumor-initiating cells in osteosarcoma. Cancer Gene Ther. 2012, 19, 212–217. [Google Scholar] [CrossRef]
  251. Shao, X.j.; Xiang, S.f.; Chen, Y.q.; Zhang, N.; Cao, J.; Zhu, H.; Yang, B.; Zhou, Q.; Ying, M.d.; He, Q.j. Inhibition of M2-like macrophages by all-trans retinoic acid prevents cancer initiation and stemness in osteosarcoma cells. Acta Pharmacol. Sin. 2019, 40, 1343–1350. [Google Scholar] [CrossRef]
  252. Wang, J.; Jin, J.; Chen, T.; Zhou, Q. Curcumol Synergizes with Cisplatin in Osteosarcoma by Inhibiting M2-like Polarization of Tumor-Associated Macrophages. Molecules 2022, 27, 4345. [Google Scholar] [CrossRef]
  253. Harris, M.A.; Hawkins, C.J. Recent and Ongoing Research into Metastatic Osteosarcoma Treatments. Int. J. Mol. Sci. 2022, 23, 3817. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Osteosarcoma therapeutic management. Doxorubicin (DOX), cisplatin (CIS) and methotrexate (MTX) are the first-line and the main chemotherapeutic drugs used in the treatment of osteosarcoma (diagram compiled from [6,11,12].
Figure 1. Osteosarcoma therapeutic management. Doxorubicin (DOX), cisplatin (CIS) and methotrexate (MTX) are the first-line and the main chemotherapeutic drugs used in the treatment of osteosarcoma (diagram compiled from [6,11,12].
Ijms 23 11416 g001
Figure 2. Rationale for the continuous discussion of the mechanisms of chemoresistance in osteosarcoma, derived from the plateaued outcomes and lack of benefits from the therapeutic interventions attempted in the modern clinical era (new drugs’ addition and dose intensification).
Figure 2. Rationale for the continuous discussion of the mechanisms of chemoresistance in osteosarcoma, derived from the plateaued outcomes and lack of benefits from the therapeutic interventions attempted in the modern clinical era (new drugs’ addition and dose intensification).
Ijms 23 11416 g002
Figure 3. Overview of the mechanisms of chemoresistance in osteosarcoma CSCs highlighted in this review. Supportive tumor microenvironmental conditions, modulated by hypoxia and inflammation, cooperate with detoxifying mechanisms, survival pathways activation and altered metabolism to induce a quiescent state, allowing DNA repair activation and culminating in evasion from apoptotic cell death.
Figure 3. Overview of the mechanisms of chemoresistance in osteosarcoma CSCs highlighted in this review. Supportive tumor microenvironmental conditions, modulated by hypoxia and inflammation, cooperate with detoxifying mechanisms, survival pathways activation and altered metabolism to induce a quiescent state, allowing DNA repair activation and culminating in evasion from apoptotic cell death.
Ijms 23 11416 g003
Figure 4. Graphical representation of the main pathways responsible for chemoresistance to the chemotherapies used in osteosarcoma (blue boxes). The conventional or classic drugs doxorubicin (DOX), cisplatin (CIS) and methotrexate (MTX) constitute the first-line treatment of osteosarcoma patients (red boxes) and often encounter tumor cell resistance mediated by those pathways. Alternative treatment strategies that can be applied in conjunction with classic drugs attempting to bypass drug resistance are represented in the green boxes.
Figure 4. Graphical representation of the main pathways responsible for chemoresistance to the chemotherapies used in osteosarcoma (blue boxes). The conventional or classic drugs doxorubicin (DOX), cisplatin (CIS) and methotrexate (MTX) constitute the first-line treatment of osteosarcoma patients (red boxes) and often encounter tumor cell resistance mediated by those pathways. Alternative treatment strategies that can be applied in conjunction with classic drugs attempting to bypass drug resistance are represented in the green boxes.
Ijms 23 11416 g004
Table 1. Overview of the functional experimental techniques used to isolate CSCs in osteosarcoma, their technical principles, possible drawbacks and stem cell-related characteristics they have identified in the mentioned studies.
Table 1. Overview of the functional experimental techniques used to isolate CSCs in osteosarcoma, their technical principles, possible drawbacks and stem cell-related characteristics they have identified in the mentioned studies.
MethodTechnical PrincipleExpert OpinionStem Cell Features Found [References]
Sphere-forming assays
-
most primitive and resilient cells survive the single-cell plating conditions in serum starvation culture systems;
-
suspended-growing conditions in non-adherent surfaces
-
experimental variability introduced over the years hampers data comparison, because of the diversity of cell density plating, use of mitogens and media supplements;
-
sphere-formation does not necessarily correlate to enhanced tumorigenic ability;
-
sphere assay mainly enriches for a population of stem and progenitor cells, together with more differentiated cells
-
expression of pluripotency-related markers [36,37,38,39,40];
-
Wnt/β-catenin activation [41];
-
resistance to chemotherapies [40,42,43,44,45,46]
Aldefluor™flow cytometry analysis of the intracellular enzymatic activity of aldehyde dehydrogenases
-
simple experimental kit, technically well-controlled;
-
needs specific flow cytometer filters which may not be available to all researchers;
-
some studies suggest that ALDH1A1 isoform is the major contributor of the positive phenotype, leaving the activity of other ALDH isoforms undetected;
-
can be difficult to separate a sufficient number of cells to conduct further biochemical experimental characterization
-
sphere-forming MG-63 cells resistant to doxorubicin, cisplatin [47];
-
stem cell marker expression and high tumorigenicity [48];
-
DKK-1 expression [49];
-
increased SOX2 expression [41];
-
expression of ALDH isozymes, such as ALDH1A1 [50];
-
metastatic dissemination [51,52]
Side-populationflow cytometry analysis of cellular extrusion of a vital dye (e.g., Hoechst-33342)Critical parameters:
-
preparation of a single-cell suspension;
-
concentration and possible toxicity of the vital dye used;
-
incubation method namely temperature and duration;
-
type and concentration of the ABC transporters’ inhibitor used to establish the negative controls;
-
accuracy of the discrimination of debris, dead and single cells;
-
quality of flow cytometry filters;
-
can be difficult to separate a sufficient number of cells to conduct further biochemical experimental characterization
-
tumorigenic capacity, expression of stemness-related markers [53];
-
Wnt activation [54];
-
CD44/Oct4 expression [55];
-
sphere-formation, drug resistance, clonogenicity [56], correlation with ABCG2 expression [41];
-
EIF4E/mTOR signaling and other genes involved in developmental processes [57]
Expression of specific surface markers (involved in e.g., cellular invasion, adhesion, and metastasis)sorting of phenotypically dissimilar cancer cell subsets based on the expression of a membrane protein using flow cytometry;
-
CSCs markers identified are based on those expressed by normal stem cells;
-
consider the possibility that sorted cancer cells reacquire their original markers;
-
inaccuracies in the sorting process itself;
-
CSC sorting in mesenchymal tumors based on surface marker expression has been more elusive than in other tumors and less consistent between research groups
-
CD133 associated with poor prognosis and chemoresistance [36,58,59,60,61,62,63];
-
CD29, CD90, CD105, CD44, ICAM-1, CD56 (mesenchymal signature) [64];
-
CD117 [44,65,66];
-
CBX3/ABCA5 [39];
-
CD248 [67,68];
-
CD271 [69,70,71];
-
osteoblastic differentiation markers CD49b [72], CD24 [73]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Martins-Neves, S.R.; Sampaio-Ribeiro, G.; Gomes, C.M.F. Chemoresistance-Related Stem Cell Signaling in Osteosarcoma and Its Plausible Contribution to Poor Therapeutic Response: A Discussion That Still Matters. Int. J. Mol. Sci. 2022, 23, 11416. https://doi.org/10.3390/ijms231911416

AMA Style

Martins-Neves SR, Sampaio-Ribeiro G, Gomes CMF. Chemoresistance-Related Stem Cell Signaling in Osteosarcoma and Its Plausible Contribution to Poor Therapeutic Response: A Discussion That Still Matters. International Journal of Molecular Sciences. 2022; 23(19):11416. https://doi.org/10.3390/ijms231911416

Chicago/Turabian Style

Martins-Neves, Sara R., Gabriela Sampaio-Ribeiro, and Célia M. F. Gomes. 2022. "Chemoresistance-Related Stem Cell Signaling in Osteosarcoma and Its Plausible Contribution to Poor Therapeutic Response: A Discussion That Still Matters" International Journal of Molecular Sciences 23, no. 19: 11416. https://doi.org/10.3390/ijms231911416

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