Skip to main content

Integrated regulation of chondrogenic differentiation in mesenchymal stem cells and differentiation of cancer cells

Abstract

Chondrogenesis is the formation of chondrocytes and cartilage tissues and starts with mesenchymal stem cell (MSC) recruitment and migration, condensation of progenitors, chondrocyte differentiation, and maturation. The chondrogenic differentiation of MSCs depends on co-regulation of many exogenous and endogenous factors including specific microenvironmental signals, non-coding RNAs, physical factors existed in culture condition, etc. Cancer stem cells (CSCs) exhibit self-renewal capacity, pluripotency and cellular plasticity, which have the potential to differentiate into post-mitotic and benign cells. Accumulating evidence has shown that CSCs can be induced to differentiate into various benign cells including adipocytes, fibrocytes, osteoblast, and so on. Retinoic acid has been widely used in the treatment of acute promyelocytic leukemia. Previous study confirmed that polyploid giant cancer cells, a type of cancer stem-like cells, could differentiate into adipocytes, osteocytes, and chondrocytes. In this review, we will summarize signaling pathways and cytokines in chondrogenic differentiation of MSCs. Understanding the molecular mechanism of chondrogenic differentiation of CSCs and cancer cells may provide new strategies for cancer treatment.

Introduction

The cartilage is a connective tissue composed of chondrocytes and their surrounding matrix, which mainly contains collagen type II and proteoglycans. Chondrogenic differentiation, the formation of chondrocytes and cartilage tissues, originates from the migration and condensation of mesenchymal stem cells (MSCs) [1]. Next, chondroprogenitor cells form, proliferate, and differentiate into chondrocytes [2]. Chondrocytes end up as resting cells to form the articular cartilage or undergo proliferation, hypertrophy, and apoptosis in a process termed endochondral ossification, thereby replacing the hypertrophic cartilage with bone [3]. Neural crest cells of the neural ectoderm, sclerotome of the paraxial mesoderm and somatopleure of the lateral plate mesoderm, which give rise to craniofacial bones, axial skeleton and skeleton of the limbs respectively, are the main source of mesenchymal stem cells [4]. Currently, MSCs are reported to be isolated and cultured from a wide range of tissues including adipose tissue [5], bone marrow [6], synovial membrane [7] and fetal appendages, such as the amniotic membrane, umbilical cord, and chorionic plate [8,9,10]. Because of their variable source and easy availability, MSCs have been widely used in cartilage tissue engineering [11]. Not only MSCs, but cancer stem cells (CSCs), a group of quiescent cell types that can drive cancer growth and reconstruct their heterogeneity, can differentiate into mesenchymal phenotypes, such as prostate cancer cell lines or melanoma cancer stem cells into osteocytes and adipocytes [12, 13]. In this review, we discuss the regulatory mechanism of MSCs differentiating into chondrogenic lineages, including signaling pathways, key proteins as well as other factors which have been shown to have an important role in chondrogenic differentiation.

Signaling pathways of chondrogenesis in mesenchymal stem cells

The proliferation and differentiation of mesenchymal cells to chondrocytes, or chondrogenic differentiation, are a complex process regulated by multiple elements, which contain intracellular proteins, receptor ligands, and transcription factors, and disruption in signaling can result in defective chondrocyte production. Signaling pathways involved in lineage determination of MSCs towards chondrogenesis is showed in Fig. 1.

Fig. 1
figure 1

Signaling pathways and proteins involved in chondrogenic differentiation of mesenchymal stem cells. R-SMAD-dependent TGF-β and Hedgehog pathways can promote the overexpression of Sox9, whereas Notch and Wnt pathways can inhibit chondrogenic transcription factor Sox9. FGF can act on R-SMAD-independent TGF-β pathway, thus playing a part in chondrogenic differentiation. There are other molecules that interact with Sox9, such as osteogenic transcription factor Runx2, peroxisome proliferator-activated receptor-gamma coactivator 1-alpha (PGC1-α) and the member of histone acetyltransferase family P300/CBP

TGF-β pathway exerts positive effects on chondrogenic differentiation

According to the interaction between distinct receptors and the ligands on the membrane, TGF-β family is divided into the TGF-β (transforming growth factor-β)/activin/nodal subfamily and BMP (bone morphogenetic protein)/growth differentiation factor (GDF)/anti-Mullerian hormone subfamily [14]. According to the participation of protein SMAD, intracellular pathways of TGF-β are separated into receptor SMAD (R-SMAD)-dependent and R-SMAD-independent [15]. The former is phosphorylated by type I kinase, thereby activating specific SMAD proteins complex that translocate into the nucleus to regulate the activation of target genes, such as chondrogenic transcription factor Sox9 [16, 17]. The latter includes mitogen-activated protein kinase (MAPK) pathways involving TGF-β-activated kinase 1 (TAK1) or extracellular signal-regulated kinase 1 and 2 [15]. The three ligands (TGF-β1, TGF-β2, and TGF-β3) have ability to induce chondrogenic differentiation, all of which bind to TGF-β type II receptor and TGF-β type I receptor and activate intercellular protein SMAD 2 and 3 [18]. TGF-βs stimulates the expression of cartilage-specific extracellular matrix proteins such as type II collagen and aggrecan [19, 20]. In a micro-pellet model, a single-day treatment of TGF-β1 was a sufficiency of stimulating bone marrow stromal cells (BMSCs) differentiation towards cartilage [21]. Amniotic MSCs overexpressing TGF-β1 expressed cartilage-specific genes and showed intense Safranin O and Alcian blue staining [22]. For cartilage tissue engineering, TGF-β1 is widely used to induce chondrogenic differentiation in MSCs [23, 24], but several studies have stated that TGF-β2 and TGF-β3 are more efficient for the chondrogenic induction with higher production of collagen II and aggrecan and glycosaminoglycan deposition [25, 26].

BMPs recruit MSCs, promote condensation or proliferation, and subsequently trigger their differentiation [27]. BMPs transduce signals through the formation of heteromeric complexes of BMP types II and I receptors and phosphorylation of intracellular protein SMAD 1, 5 and 8. Among the BMP isoforms, the most widely studied are BMP-2, BMP-3 (osteogenin), BMP-4, BMP-6, BMP-7 (also known as osteogenic protein-1), and BMP-9 [18, 28]. BMP-2/4 enhance the recruitment of mesenchymal precursors for cartilage condensations, and regulate the condensation size [29]. Human muscle-derived stem cells (hMDSCs) transfected with lenti-BMP2/GFP vector could enhance the capacity of hMDSCs to differentiate into cartilage, as confirmed by Alcian blue and Col2a1 positive staining [30]. BMP4, BMP11, BMP6, BMP7, BMP9, BMP13 and BMP14 have been reported to up-regulate chondrogenic markers Sox9, Sox5, and Sox6 [31].

In terms of R-SMAD-independent pathway, deletion of Tak1 in limb mesenchyme cells could result in the inactivation of the downstream MAPK target p38, as well as impaired the activation of the BMP/SMAD signaling pathway, which the differentiation of the chondrocyte lineage was interrupted [32]. Noggin subordinated with BMP signaling inhibitors and could be induced by TGF-β1 during the recruitment of progenitor cells into cartilage elements [33]. ERK1/2 is activated upon TGF-β1 stimulation or BMP2 administration and acts as the passive modulator of chondrogenic differentiation [34, 35], but this inhibition of chondrogenic differentiation was covered by the positive effect of R-SMAD-dependent pathway.

Hedgehog (Hh) pathway plays a relevant role in chondrogenic differentiation

Hh was first identified in Drosophila body plan [36]. Smoothened (Smo) is a receptor, which possesses an ability of activating intracellular signals repressed by Patched (Ptch) [37]. When secreted by the sending cell, Hh ligands bind to Ptch on the receiving cell, and this action can relieve the negative effect of Ptch on Smo, which initiates Hh signal transmission. In mammals, zinc finger proteins Gli1, Gli2, and Gli3 are involved in the transcriptional regulation [38], and sonic hedgehog (Shh), desert hedgehog (Dhh), and Indian hedgehog (Ihh) are ligands in Hh pathway [39]. Dhh is required for spermatogenesis and formation of neuronal sheaths [40], but rarely been reported in chondrogenic differentiation. There is the high-level expression of Ihh in proliferating limb bud mesodermal cells, which subsequently differentiate into osteo-chondroprogenitor after mesenchymal migration and condensation [41]. Compared to TGF-β1 and BMP-2, Ihh is also a potent inducer of chondrogenic differentiation of primary MSCs [42]. Shh takes part in morphogenesis of the muscle, hair, teeth, lung and gut, patterning of body limbs and cell fates of neural progenitors [43]. BMSCs transfected with Shh induced chondrogenic differentiation in the rotary cell culture system [44]. Exogenous Shh could induce the expression of the transcription factor Sox9 in the somatic tissue, and upregulated Sox9 expression level could induce robust chondrogenesis via BMP signals [45]. The Hh signaling antagonist HhAntag could modulate the BMP2-mediated the canonical SMAD1/5/8 and non-canonical p38/MAPK signaling [46] (Fig. 2).

Fig. 2
figure 2

Integrated regulation of chondrogenesis differentiation from mesenchymal stem cells. Red arrows indicate promoting effect and green arrows with lines indicate inhibiting effect

Notch pathway negatively regulates chondrogenic differentiation

Notch signaling works via cell–cell contact, where transmembrane ligands on one cell recognize transmembrane receptors on its adjacent cells. This interaction can activate intracellular proteases, such as ADAM metalloprotease and γ-secretase complex, thereby contribute to proteolytic cleavage of the receptor [47]. After cleavage, notch intracellular domain (NICD), the COOH-terminal portion of the receptor, is released and translocated to the nucleus to form a complex with the transcription factor CSL (also known as RBP-JK) and the transcriptional coactivator mastermind-like 1 [48]. The ultimate complex targets the hairy and enhancer of split (Hes) and Hes-related with YRPW motif (Hey) [49]. Unlike TGF-β and BMP pathways, Notch signaling inhibits chondrogenic differentiation via interactions between four receptors (Notch1–5) and at least five ligands including two members of the Jagged (JAG) family and three members of the delta-like family (Dll) [48, 50]. This suppression regulates the expression of the chondrogenic transcription factor Sox9, whereas it is irrelevant to the regulation of cartilage matrix catabolism [51, 52]. Cartilage regeneration induced by placenta-derived mesenchymal stromal cell (PMSC) could be promoted by inhibiting the Jagged1 (JAG1) peptides of Notch pathway [53, 54]. Notch inhibition of chondrogenic differentiation is regulated by the transcription factor Twist1, which cooperates with a putative NICD/RBPjK binding element in the promoter region [55]. Overexpression of ligand Dll2 could strongly promote the activation of p38/MAPK rather than extracellular signal-regulated kinase 1/2 and c-Jun N-terminal kinase, thereby inhibiting the chondrogenic differentiation of ATDC5 (a kind of cartilage cell line) [56].

Wnt pathway influences chondrogenic differentiation

Wnt pathway can be grouped into two categories based on the participation of β-catenin in signal transduction. One is canonical (β-catenin-dependent), and the other is non-canonical (β-catenin-independent) [57]. With respect to the β-catenin-dependent pathway, Wnt proteins bind a heterodimeric receptor complex, comprised of a Frizzled (Fzd) and an LRP5/6 protein [58]. This binding inhibits GSK-3β phosphorylation to β-catenin and leads to the stabilization of β-catenin. β-catenin in stable state accumulates in the nucleus and binds T cell-specific factor/lymphoid enhancer-binding factor (TCF/LEF), resulting in activation of target genes [59, 60]. In the deficiency of Wnt ligands, the cytoplasmic complex consisting of the scaffolding protein Axin, adenomatous polyposis coli, casein kinase 1, and GSK-3β can phosphorylate β-catenin, resulting in its ubiquitination and subsequent proteasomal degradation [61]. The non-canonical pathway is split into two branches, the Wnt/calcium pathway and the planar cell polarity [62]. β-catenin-mediated canonical Wnt signaling inhibits chondrogenic differentiation, while non-canonical pathways promote this differentiation [18, 63]. As the member of canonical Wnt pathway, Wnt3a plays a dual role on chondrogenic capacity of MSCs [64,65,66]. The non-canonical factor Wnt5a augments cartilage formation, collagen fiber rearrangement, and remarkably enhances glycosaminoglycan and collagen deposited in vivo [67]. The synergistic effect of Wnt5a and TGF-β3 stimulated the activation of p38/MAPK pathway, a positive regulator of chondrogenic differentiation [68]. TGF-β1-mediated MAP kinase activation could promote the accumulation of intracellular β-catenin, and increase the expression of Wnt7a and N-cadherin (mesenchymal condensation marker) [69].

Wnt inhibitors and activators also play a significant role in cartilage homeostasis and development. As a Wnt antagonist, FRZB blocked canonical Wnt signaling and increased the production and deposition of glycosaminoglycan with overexpression of chondrogenic markers (Sox9 and Col2a1) [70]. Lithium chloride and CHIR99021 (CHIR), commercially available Wnt agonists, not only stimulate MSC proliferation, but also enhance the chondrogenic capacity of MSCs [71].

Fibroblast growth factor (FGF) signaling pathway is associated with chondrogenic differentiation

FGF signaling pathway involves in many physiological processes comprising digit morphogenesis, limb organogenesis, cerebral development and metabolic homeostasis [72]. In terms of limb development, FGFs signaling is associated with mesenchymal condensation, chondrogenic differentiation and hypertrophy, mineral homeostasis and bone formation [73]. There are four members of FGF receptors (FGFR1–FGFR4) family with diverse FGF-binding capacities and at least 22 FGF ligands that can be subdivided into seven groups [74,75,76]. The FGF receptor (FGFR) commonly is composed of an intracellular receptor tyrosine kinase domain, a hydrophobic transmembrane region and extracellular ligand-binding domains [77]. In the presence of FGF or other ligands, FGFR kinases are released from autoinhibition and auto-phosphorylated, whose phosphorylated tyrosine residues serve as attachment sites for recruiting interacting proteins [78]. The downstream signaling pathways such as phosphoinositide 3-kinase/Akt (PI3K–AKT) and protein kinase C pathways can be activated and transduce information into nucleus [79]. In general, FGFR2, FGF2, FGF8, FGF9, and FGF18 have been reported to involve in chondrogenic differentiation [80, 81]. FGFR2 is an early marker of chondrogenesis, whose expression pattern is restrained in mesenchymal condensation region before occurrence of chondroprogenitor [81, 82]. A gain-of-function mutation (S252W) of FGFR2 in mice results in dwindling proliferation BMSCs, a decline in BMSC chondrogenic differentiation via inhibiting mineralization [83], and changes in both Wnt signals and MAPK expression during human synovium-derived stem cell chondrogenic differentiation [84]. Solchaga et al. illustrated that FGF-2 functioned in the regulation of chondrogenic differentiation through MAPK and Wnt signaling via DUSP 4/6 and Fzd7, respectively [85]. In addition to FGF-2, FGFR1 could cooperate with β-catenin and alter the lineage commitment of MSCs into chondrocytes [86]. In combination with transforming growth factor-beta (TGF-β), FGF9 and FGF18 stimulated early chondrogenic differentiation by shifting the chondrogenic program earlier [87].

Cytokines of chondrogenic differentiation in MSCs

Transcription factor—Sox9

Sox9, sex determining region Y-box 9 [88], pertains to the SRY-related high-mobility group (HMG) box (Sox) family, whose family proteins are a conserved group of transcriptional regulators [89]. The fate and terminal differentiation of chondrocyte are regulated by Sox9 via its accurate spatial and temporal expression pattern [89,90,91]. Sox9 can be detected in chondroprogenitors and mature chondrocytes, but not in hypertrophic chondrocytes, and Sox9 directly upregulates genes specifically expressed in precartilaginous condensation [92]. Heterozygous mutations of Sox9 in human were first ascertained as the causative factor of campomelic dysplasia, a skeletal deformity with disorder of sexual development and genital ambiguities [93]. Mice with heterozygous Sox9 mutant can be perinatal mortality and presented palatoschisis and crookedness of bony structures originated from cartilage precursors [94]. As a part of the SoxE family, Sox9 possesses a distinct dimerization domain located proximally to the HMG box, which can interact with the enhancers and promoters of the aggrecan gene (Acan) and type-II collagen gene (Col2a1) [95,96,97,98,99], and a unique transactivation domain [100]. Sox9 with downstream transcription factors of SoxD family—Sox5 and Sox6 can work as chondrogenic Sox Trio [91]. Unlike Sox9, Sox5 and Sox6 are similar to one another in structure and function [101]. Sox5-deleted or Sox6-deleted mice develop modest skeletal disorders, yet deletion of these two genes could suffer serious cartilage primordia [91]. Zhou et al. have reported that the strong transcriptional activity of the 48 bp minimal enhancer of Col2a1 gene appears just when Sox9, Sox5, and Sox6 (Sox5/6) bind together to the sites in the enhancer [102]. In addition to inducing overexpression of Col2a1 and Acan, Sox Trio inhibits hypertrophic gene expression and collagen type X deposition [103].

Cartilage matrix protein—Col2a1 and aggrecan

The collagen α1(II) gene, located at 12q13, is a key element for production of alpha-1 type II collagen (Col2a1) [104], which is a main extracellular matrix protein of cartilage. Mutations in Col2a1 can result in spondyloepiphyseal dysplasia and type II achondrogenesis. The former is an autosomal dominant disease characterized by limb and trunk shortness, pulmonary hypoplasia, abdominal enlargement with polyuria and edema [105, 106]. In chondrogenic differentiation, expression profiles of Col2a1 are similar to the expressions of Sox9 in all chondroprogenitor cells in mouse and chick embryonic development [107,108,109]. There is a 48-bp gens in intron 1 sequence of Col2a1 binding with Sox9, which is required for cartilage-specific expression [98, 102]. Besides SOX family, endogenous Hey-1 and Hes-1 can bind to N-box domains of the first intron of Col2a1, which is adjacent to the Sox9 enhancer binding site [110]. The response element of Runx1 has been identified in the 5-flanking regions of the Col2a1 promoter [111]. Transcription factors (Kruppel-like factor-4 and AT-rich interactive domain 5B) could combine with the E1 enhancer element in Col2a1 and regulate its expression [112]. Aggrecan (Acan) is a large chondroitin sulfate proteoglycan consisting of a polypeptide backbone covalently attached to one or more glycosaminoglycan chains in growth plate cartilage [113, 114]. Four enhancer elements were identified in the aggrecan gene, two of which (− 80 and − 62) showed individual chondrocyte developmental stage specificity. The other enhancers specific to chondrogenesis, + 28 and − 30, were not associated with chondrocyte type [95]. Additionally, transcription factor Sox9 binds to the first enhancer of the aggrecan gene with or without Sox5 and Sox6 [115, 116]. In addition to Sox9, PAX1/9-binding site partly overlaps with a Sox9-binding site and exhibits as a weak transactivator [117].

Non-coding RNA in chondrogenic differentiation

MiR-140 is involved in cartilage homeostasis and osteoarthritis development. miR-140-5p and miR-140-3p have been identified that are products of RNA Dicer excision at the 3′ and 5′ tail of the pre-miRNA, respectively [118]. This miRNA can bind to Sox9 directly or act on its upstream genes, such as RALA and SMAD3 [118,119,120]. Recent studies have confirmed the bioactivity of exosomes carrying miR-140 in chondrogenic differentiation of MSCs [121, 122]. MiR-23a and b are isoforms of miR-23 [123]. The former regulated by lncRNA SNHG5 decreases the expression of SOX6/SOX5 and inhibits chondrogenic differentiation, whereas miR-23a-3p silencing attenuates the differentiation effect of BMSCs [124, 125]. The latter induces differentiation into chondrocytes of hMSCs through the downregulation of protein kinase A signaling [126]. Recently, it was reported that FGF2 expression is negatively regulated by miR-23c, thereby affecting chondrogenic differentiation [127].

LncRNA DANCR, known as an anti-differentiation ncRNA, not only binds to protein-coding genes, but also regulates miRNAs, such as miR-1305 and miR-1275 in chondrogenesis [128,129,130,131,132]. Using CRISPR activation (CRISPRa) technology, DANCR improved adipose-derived stem cell chondrogenic differentiation by inhibiting miR-203a and miR-214 [133]. The lncRNA UCA1 has been proved to be relevant to several human cancers. The expression of UCA1 increased markedly along with chondrocyte differentiation [134]. UCA1 combined with miR-145-5p/SMAD5 and miR-124-3p/SMAD4 can regulate chondrogenic differentiation of MSCs [135].

According to the bioinformatics analysis of hybridization arrays, circRNAs associated with chondrogenic differentiation are derived from similar precursor genes, such as FKBP5, ZEB1, or SMYD3 [136]. However, no experimental studies have reported the specific role of circRNAs in the cartilage lineage differentiation of stem cells [137].

Biophysical factors in chondrogenic differentiation

Early stage of chondrogenic differentiation is enhanced and the expression of chondrogenic markers Col2a1, Sox9, and Acan are increased when MSCs are cultured under low oxygen tension [138, 139]. Besides the hypoxia factor, there are several mechanical cues in cartilage formation [140]. Fluid shear stress (ΔSS) is a potent regulator of chondrogenic differentiation, which is comparable to TGF-β1 induction [141, 142]. Hydrostatic pressure has an anabolic effect on MSCs chondrogenic differentiation, and the loading capacity and time, longitude of chondrogenic preprocessing prior to pressurizing, also have an effect on chondrogenic differentiation [143]. Lineage determination of MSCs is susceptible to material stiffness, which regulates this development through transforming growth factor beta (TGF-β) signaling pathway [144]. MSCs on softer substrates are prone to differentiation into chondrogenic lineage [145].

Chondrogenic differentiation of malignant tumors

Cancer stem cells (CSCs), also known as tumor-initiating cells, are a small fraction of cells inside tumor tissues. They can self-renew and differentiate and are responsible for relapses as well as resistance to chemoradiotherapy [146, 147]. The ability of cancer cells to switch from non-CSCs to CSCs and vice versa is called phenotypic plasticity [148]. CSC plasticity takes determination to malignancy population dynamic and promotes cancer cellular progression [149,150,151]. Therefore, many researchers have exploited the plasticity of CSCs as a promising therapeutic target. The differentiation therapy implies that CSCs could be induced into differentiate into matured cells, and also converted into non-stem cells that were sensitive to traditional anticancer reagents [152]. The mesenchymal differentiation capacities of CSCs are involved in nervous system neoplasms [153, 154], bone sarcoma [155, 156], and prostate and cervical cancers [12, 157]. Polyploid giant cancer cells (PGCCs) induced by cobalt chloride have the properties of CSC characteristics and a single PGCC can form tumor in nude mice. When cultured in adipogenic, osteogenic and chondrogenic induction medium, PGCCs could differentiate into adipocytes, osteocytes and chondrocytes, respectively [158].

Potential mechanism of CSCs inducible differentiation

Mesenchymal phenotype is a determinant for differentiation of CSC. Wilms tumors (WTs) are genetically heterogeneous kidney tumors. Five WT cell lines expressed MSC-specific surface proteins and had the ability to differentiation into adipogenic, chondrogenic, osteogenic and myogenic lineage [159]. ALDH + malignant phyllodes tumor cells could also be induced to differentiate into chondrocytes. CD133 + CSCs from osteosarcoma and Ewing’s sarcoma displayed the potency to differentiate into mesenchymal lineages, such as osteoblasts and chondrocytes [155, 156]. For epithelial-derived neoplasms, the transition between the epithelium and mesenchyme is a biological process called as epithelial–mesenchymal transition (EMT), in which epithelial cells obtain mesenchymal cell phenotypes and are related to mesenchymal differentiation [160]. Epithelial cells transduced with TGF-β1 retrovirus had the functional resemblance to MSCs, including a comparable antigenic phenotype (positive for CD44), the ability to lineage commitment (positive for Alcian blue of chondrocytes) [161]. It has been reported that the growth factor family (GF) regulated mesenchymal tri-lineage differentiation. Epidermal growth factor receptor (EGFR) inhibitors can induce the mesenchymal–epithelial transition and impact on EGF-induced EGFR signaling in breast CSC mesenchymal differentiation [162]. Transcriptional coactivator with PDZ-binding motif (TAZ) interacts with the transcription factor TEA domain family members (TEAD), which plays pivotal roles in EMT and cell growth. TAZ overexpression increases the expression of mesenchymal marker CD44 and the differentiation capability towards osteoblastic and chondrogenic lineage in glioma stem cells and murine neural stem cells [163].

Verhaak et al. described a robust gene expression-based molecular classification of GBM involved in proneural, neural, classical, and mesenchymal subtypes [164]. Analysis of The Cancer Genome Atlas (TCGA) demonstrated that RTVP-1 overexpressed in GBM which expressed mesenchymal phenotype, and silencing of RTVP-1 abrogated the chondrogenic differentiation of GBM cells in response to specific induction media. RTVP-1 promoter could bind C/EBP β, a master transcription factors that regulated mesenchymal transformation of GBM [165], and C/EBP β promoter activity could be suppressed by binding with Sox9 [166]. C/EBP β-expressing cells terminated in further increase of mesenchymal gene expression and obtained mesenchymal properties [167]. In gastric signet ring cell adenoma cancer, dysregulation of EMT-associated molecules and differentiation towards chondrocytes also occurred in KATO-III cell line (a cell line of gastric carcinoma) under the induction media [168].

The chondrogenic induction medium components in CSCs are same as those in MSCs, and include insulin-transferrin-selenium (ITS), ascorbate, TGF-β and dexamethasone. Exogenous TGF-β can stimulate TGF-β pathway, thereby accelerate the chondrogenic differentiation. Physiological levels of dexamethasone play a positive regulatory role in cartilage formation by directly interacting with the TGF-β signaling molecule, SMAD3 [169]. ITS is a nutritional supplement involved in glucose and proline metabolism during collagen synthesis. Ascorbate is a requisite cofactor in the production of collagen II and sulfated glycosaminoglycan [170]. The molecular mechanism of chondrogenic differentiation of CSCs may be similar to that of MSCs. Resveratrol inhibited MMP-induced chondrogenic differentiation via the p38 kinase and JNK pathways in chondrosarcoma cells [171]. The activation of p38/ERK/JNK pathways in the presence of TGF-β1 facilitated the expression of Sox9, collagen II and aggrecan [172]. MiR-200b-3p mediated chondrogenic differentiations of quercetin-induced pancreatic ductal adenocarcinoma CSCs by inhibiting Notch and activating Numb [173]. Sox9, Col2a1 and Acan was increased when transcription factor of notch pathway RBP-JK was restrained by DNA methyltransferase 3b [174].

Differentiation therapy of CSCs

Retinoic acid (RA) is a well-known modulator in skeleton development, of which functions through a class of nuclear hormone receptors, the retinoic acid receptors (RARs) and retinoid-X-receptors (RXRs), to regulate gene transcription [175]. RA has successfully used in the differentiation therapy of acute promyelocytic leukemia [176]. RA acts on the RARA moiety of PML-RARA fusion protein that will be degraded, and contributes leukemia cells to terminal differentiation and apoptosis [177]. RA inhibited the condensation and proliferation and dysregulated Sox9 and Col2a1 in a dose-dependent manner through inhibition of Shh/Gli3 pathway [178]. Subtype-specific RAR agonists as well as RA had a strong inhibitory effect on chondrogenic differentiation with decreased the expression levels of Col2a1 and Sox9 [179]. Palovarotene (PVO), a RAR selective agonist, attenuates overactivated BMP signaling and restored aberrant chondrogenic fate determination of osteochondroma cells [180].

Traditional Chinese medicine has been widely accepted as an alternative treatment for cancer [173, 181]. In glioblastoma stem cells, the expression levels of proteins associated with differentiation, such as glial fibrillary acidic protein (GFAP), Notch1 and Shh, were increased by β-elemene in vitro and in vivo [182, 183]. Antagonist of CXCR4 (PRX177561) could regulate the tumor microenvironment, decrease the expression of CSC markers Sox2, Twist, Nanog, and overexpress GFAP in glioma [184]. Treatment of CD133 + CSCs derived from human GBMs with BMPs decreased cell proliferation in vitro and differentiated into astrocytes [185]. As a typical potassium ionophore antibiotic, Salinomycin improves epithelial differentiation ability of breast CSCs by means of eliminating cancer cells from CSC features and blocking cell cycle [186]. The inhibitor of HDACs (Quisinostat) stimulates DOX-mediated cytotoxicity in breast CSCs as well as non-CSCs from basal-like, mesenchymal-like, and luminal-like breast cancer [187]. In hepatocellular carcinoma, DW14800, a novel inhibitor of protein arginine methyltransferase 5 (PRMT5), promotes the differentiation of CSCs by downregulating the expression of HNF4α and enhancing methylation of H4R3me2s [188]. Metformin, a well-known anti-diabetic drug, can decrease CSC marker expression (CD44 and Sox2) and increase the expression of differentiation markers (Kruppel-like factor 4 and MUC5AC) in gastric CSCs [189]. IDH1 and IDH2 mutations are frequent in many solid cancers, and the enzymes encoded by these mutation are endowed with new function that facilitate the accumulation of the oncometabolite D-2-hydroxyglutarate (D-2HG), which has significant impacts on epigenetic regulation, differentiation degree, and metabolic patterns [190]. The differentiation inducers targeted mutant isocitrate dehydrogenase IDH1 and IDH2 have been granted for clinic trials. Ivosidenib, an IDH1 inhibitor, showed improved progression-free survival versus placebo in the phase III clinical trial of patients with advanced cholangiocarcinoma [191]. In advanced glioma, patients treated with ivosidenib were associated with a favorable safety profile, prolonged progression-free survival [192].

Conclusion and perspective

Chondrogenic differentiation is a multifactorial and multistep process. Mesenchymal differentiation into chondrocytes is regulated by concurrent signaling pathways, among which the TGF-β pathway is the principal and the earliest signal in chondrogenic condensation. Engineered TGF-β superfamily ligands have been produced and tested for in vitro cartilage formation and used for commercial purposes. As transcriptomic and bioinformatics technologies have been developed, many of the identified ncRNAs have been validated for chondrogenic differentiation. LncRNA DANCR could improve mesenchymal differentiation to cartilage lineage by miRNAs (such as miR-203a and miR-214). CSCs, a small fraction of cancer cells with tumorigenesis, have ability to differentiate into non-CSCs because of phenotypic plasticity, which can be the basis for differentiation therapy. Retinoic acid (RA) has been accepted as the first-line treatment in acute promyelocytic leukemia. Retinoid signaling was also informed to play a negative role in chondrogenic differentiation. Therefore, it may be possible that retinoid antagonists directly transform malignant neoplasm into benign cartilage tissue. It has been reported that CSCs can differentiate along their lineage or undergo mesenchymal differentiation through EMT. CSCs could be induced to differentiate into adipogenic, osteogenic and chondrogenic cells in more than 10 kinds of malignant tumors. However, these studies just describe this differentiation phenomenon rather than clearly elucidate the mechanism underlying chondrogenic differentiation of CSCs. Understanding the molecular mechanism of induced differentiation of CSCs may be contribute to developing new drugs for cancer treatment.

Availability of data and materials

The authors declare that all data supporting the findings of this study are available within the article or contact the corresponding author upon reasonable request.

Abbreviations

MSC:

Mesenchymal stem cell

Sox:

Sex determining region Y-box

CSC:

Cancer stem cell

ECM:

Extracellular matrix

TGF-β:

Transforming growth factor-β

GDF:

Growth differentiation factor

MAPK:

Mitogen-activated protein kinase

TAK1:

TGF-β-activated kinase 1

BMSC:

Bone marrow stromal cell

BMP:

Bone morphogenetic protein

Col2a1:

Type-II collagen gene

Hh:

Hedgehog

Smo:

Smoothened

Ptch:

Patched

Shh:

Sonic hedgehog

Dhh:

Desert hedgehog

Ihh:

Indian hedgehog

NICD:

Notch intracellular domain

Hes:

Hairy and enhancer of split

Hey:

Hes-related with YRPW motif

Fzd:

Frizzled

GSK-3β:

Glycogen synthase kinase 3

FGF:

Fibroblast growth factor

FGFR:

FGF receptor

HMG:

High-mobility group

Acan:

Aggrecan

PGCCs:

Polyploid giant cancer cells

RA:

Retinoic acid

RAR:

Retinoic acid receptor

ITS:

Insulin-transferrin-selenium

GFAP:

Glial fibrillary acidic protein

GBM:

Glioblastoma

ATO:

Arsenic trioxide

IDH:

Isocitrate dehydrogenase

HDAC:

Histone deacetylase

EMT:

Epithelial–mesenchymal transition

References

  1. Hall BK, Miyake T. All for one and one for all: condensations and the initiation of skeletal development. BioEssays. 2000;22(2):138–47.

    Article  CAS  PubMed  Google Scholar 

  2. Goldring MB, Tsuchimochi K, Ijiri K. The control of chondrogenesis. J Cell Biochem. 2006;97(1):33–44.

    Article  CAS  PubMed  Google Scholar 

  3. Chijimatsu R, Saito T. Mechanisms of synovial joint and articular cartilage development. Cell Mol Life Sci. 2019;76(20):3939–52.

    Article  CAS  PubMed  Google Scholar 

  4. Tang X, Fan L, Pei M, Zeng L, Ge Z. Evolving concepts of chondrogenic differentiation: history, state-of-the-art and future perspectives. Eur Cell Mater. 2015;30:12–27.

    Article  CAS  PubMed  Google Scholar 

  5. Wang Y, Hu G, Hill RC, Dzieciatkowska M, Hansen KC, Zhang XB, Yan Z, Pei M. Matrix reverses immortalization-mediated stem cell fate determination. Biomaterials. 2021;265:120387.

    Article  CAS  PubMed  Google Scholar 

  6. Tian G, Jiang S, Li J, Wei F, Li X, Ding Y, Yang Z, Sun Z, Zha K, Wang F, et al. Cell-free decellularized cartilage extracellular matrix scaffolds combined with interleukin 4 promote osteochondral repair through immunomodulatory macrophages: in vitro and in vivo preclinical study. Acta Biomater. 2021;127:131–45.

    Article  CAS  PubMed  Google Scholar 

  7. Li N, Gao J, Mi L, Zhang G, Zhang L, Zhang N, Huo R, Hu J, Xu K. Synovial membrane mesenchymal stem cells: past life, current situation, and application in bone and joint diseases. Stem Cell Res Ther. 2020;11(1):381.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. You Q, Liu Z, Zhang J, Shen M, Li Y, Jin Y, Liu Y. Human amniotic mesenchymal stem cell sheets encapsulating cartilage particles facilitate repair of rabbit osteochondral defects. Am J Sports Med. 2020;48(3):599–611.

    Article  PubMed  Google Scholar 

  9. Liang H, Suo H, Wang Z, Feng W. Progress in the treatment of osteoarthritis with umbilical cord stem cells. Hum Cell. 2020;33(3):470–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Ma J, Wu J, Han L, Jiang X, Yan L, Hao J, Wang H. Comparative analysis of mesenchymal stem cells derived from amniotic membrane, umbilical cord, and chorionic plate under serum-free condition. Stem Cell Res Ther. 2019;10(1):19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Zhang R, Ma J, Han J, Zhang W, Ma J. Mesenchymal stem cell related therapies for cartilage lesions and osteoarthritis. Am J Transl Res. 2019;11(10):6275–89.

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Zhau HE, He H, Wang CY, Zayzafoon M, Morrissey C, Vessella RL, Marshall FF, Chung LWK, Wang R. Human prostate cancer harbors the stem cell properties of bone marrow mesenchymal stem cells. Clin Cancer Res. 2011;17(8):2159–69.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Fang D, Nguyen TK, Leishear K, Finko R, Kulp AN, Hotz S, Van Belle PA, Xu X, Elder DE, Herlyn M. A tumorigenic subpopulation with stem cell properties in melanomas. Cancer Res. 2005;65(20):9328–37.

    Article  CAS  PubMed  Google Scholar 

  14. Shi Y, Massague J. Mechanisms of TGF-beta signaling from cell membrane to the nucleus. Cell. 2003;113(6):685–700.

    Article  CAS  PubMed  Google Scholar 

  15. Thielen NGM, van der Kraan PM, van Caam APM. TGFbeta/BMP signaling pathway in cartilage homeostasis. Cells. 2019;8(9):969.

    Article  CAS  PubMed Central  Google Scholar 

  16. Lyons KM, Rosen V. BMPs, TGFβ, and border security at the interzone. Curr Top Dev Biol. 2019;133:153–70.

    Article  CAS  PubMed  Google Scholar 

  17. Grafe I, Alexander S, Peterson JR, Snider TN, Levi B, Lee B, Mishina Y. TGF-β family signaling in mesenchymal differentiation. Cold Spring Harb Perspect Biol. 2018;10(5):a022202.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  18. Zhou S, Chen S, Jiang Q, Pei M. Determinants of stem cell lineage differentiation toward chondrogenesis versus adipogenesis. Cell Mol Life Sci. 2019;76(9):1653–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Redini F, Galera P, Mauviel A, Loyau G, Pujol JP. Transforming growth factor beta stimulates collagen and glycosaminoglycan biosynthesis in cultured rabbit articular chondrocytes. FEBS Lett. 1988;234(1):172–6.

    Article  CAS  PubMed  Google Scholar 

  20. Wang G, Chen S, Xie Z, Shen S, Xu W, Chen W, Li X, Wu Y, Li L, Liu B, et al. TGFbeta attenuates cartilage extracellular matrix degradation via enhancing FBXO6-mediated MMP14 ubiquitination. Ann Rheum Dis. 2020;79(8):1111–20.

    Article  CAS  PubMed  Google Scholar 

  21. Futrega K, Robey PG, Klein TJ, Crawford RW, Doran MR. A single day of TGF-beta1 exposure activates chondrogenic and hypertrophic differentiation pathways in bone marrow-derived stromal cells. Commun Biol. 2021;4(1):29.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Chae D-S, Han JH, Park Y-J, Kim S-W. TGF-β1 overexpressing human MSCs generated using gene editing show robust therapeutic potential for treating collagen-induced arthritis. J Tissue Eng Regen Med. 2021;15(5):513–23.

    Article  CAS  PubMed  Google Scholar 

  23. Zhao L, Hantash BM. TGF-β1 regulates differentiation of bone marrow mesenchymal stem cells. Vitam Horm. 2011;87:127–41.

    Article  CAS  PubMed  Google Scholar 

  24. Kwon H, Brown WE, O’Leary SA, Hu JC, Athanasiou KA. Rejuvenation of extensively passaged human chondrocytes to engineer functional articular cartilage. Biofabrication. 2021;13:035002.

    Article  CAS  Google Scholar 

  25. de Araujo FV, Carrillo-Galvez AB, Martin F, Anderson P. TGF-beta and mesenchymal stromal cells in regenerative medicine, autoimmunity and cancer. Cytokine Growth Factor Rev. 2018;43:25–37.

    Article  CAS  Google Scholar 

  26. Wu M, Chen G, Li YP. TGF-beta and BMP signaling in osteoblast, skeletal development, and bone formation, homeostasis and disease. Bone Res. 2016;4:16009.

    Article  PubMed  PubMed Central  Google Scholar 

  27. Sampath TK, Vukicevic S. Biology of bone morphogenetic protein in bone repair and regeneration: a role for autologous blood coagulum as carrier. Bone. 2020;141:115602.

    Article  CAS  PubMed  Google Scholar 

  28. Deng ZH, Li YS, Gao X, Lei GH, Huard J. Bone morphogenetic proteins for articular cartilage regeneration. Osteoarthr Cartil. 2018;26(9):1153–61.

    Article  CAS  Google Scholar 

  29. Hall BK, Miyake T. Divide, accumulate, differentiate: cell condensation in skeletal development revisited. Int J Dev Biol. 1995;39(6):881–93.

    CAS  PubMed  Google Scholar 

  30. Gao X, Cheng H, Awada H, Tang Y, Amra S, Lu A, Sun X, Lv G, Huard C, Wang B, et al. A comparison of BMP2 delivery by coacervate and gene therapy for promoting human muscle-derived stem cell-mediated articular cartilage repair. Stem Cell Res Ther. 2019;10(1):346.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Zhang L, Luo Q, Shu Y, Zeng Z, Huang B, Feng Y, Zhang B, Wang X, Lei Y, Ye Z, et al. Transcriptomic landscape regulated by the 14 types of bone morphogenetic proteins (BMPs) in lineage commitment and differentiation of mesenchymal stem cells (MSCs). Genes Dis. 2019;6(3):258–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Gunnell LM, Jonason JH, Loiselle AE, Kohn A, Schwarz EM, Hilton MJ, O’Keefe RJ. TAK1 regulates cartilage and joint development via the MAPK and BMP signaling pathways. J Bone Miner Res. 2010;25(8):1784–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Ray A, Singh PNP, Sohaskey ML, Harland RM, Bandyopadhyay A. Precise spatial restriction of BMP signaling is essential for articular cartilage differentiation. Development. 2015;142(6):1169–79.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Seghatoleslami MR, Roman-Blas JA, Rainville AM, Modaressi R, Danielson KG, Tuan RS. Progression of chondrogenesis in C3H10T1/2 cells is associated with prolonged and tight regulation of ERK1/2. J Cell Biochem. 2003;88(6):1129–44.

    Article  CAS  PubMed  Google Scholar 

  35. Li J, Zhao Z, Liu J, Huang N, Long D, Wang J, Li X, Liu Y. MEK/ERK and p38 MAPK regulate chondrogenesis of rat bone marrow mesenchymal stem cells through delicate interaction with TGF-beta1/Smads pathway. Cell Prolif. 2010;43(4):333–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Ingham PW, McMahon AP. Hedgehog signaling in animal development: paradigms and principles. Genes Dev. 2001;15(23):3059–87.

    Article  CAS  PubMed  Google Scholar 

  37. Hu A, Song BL. The interplay of patched, smoothened and cholesterol in Hedgehog signaling. Curr Opin Cell Biol. 2019;61:31–8.

    Article  CAS  PubMed  Google Scholar 

  38. Petrov K, Wierbowski BM, Salic A. Sending and receiving Hedgehog Signals. Annu Rev Cell Dev Biol. 2017;33:145–68.

    Article  CAS  PubMed  Google Scholar 

  39. Qi X, Li X. Mechanistic Insights into the Generation and Transduction of Hedgehog Signaling. Trends Biochem Sci. 2020;45(5):397–410.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Ehlen HW, Buelens LA, Vortkamp A. Hedgehog signaling in skeletal development. Birth Defects Res C Embryo Today. 2006;78(3):267–79.

    Article  CAS  PubMed  Google Scholar 

  41. St-Jacques B, Hammerschmidt M, McMahon AP. Indian hedgehog signaling regulates proliferation and differentiation of chondrocytes and is essential for bone formation. Genes Dev. 1999;13(16):2072–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Steinert AF, Weissenberger M, Kunz M, Gilbert F, Ghivizzani SC, Göbel S, Jakob F, Nöth U, Rudert M. Indian hedgehog gene transfer is a chondrogenic inducer of human mesenchymal stem cells. Arthritis Res Ther. 2012;14(4):R168.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Ohba S. Hedgehog signaling in skeletal development: roles of Indian Hedgehog and the mode of its action. Int J Mol Sci. 2020;21(18):6665.

    Article  CAS  PubMed Central  Google Scholar 

  44. Chen L, Liu G, Li W, Wu X. Sonic hedgehog promotes chondrogenesis of rabbit bone marrow stem cells in a rotary cell culture system. BMC Dev Biol. 2019;19(1):18.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Zeng L, Kempf H, Murtaugh LC, Sato ME, Lassar AB. Shh establishes an Nkx3.2/Sox9 autoregulatory loop that is maintained by BMP signals to induce somitic chondrogenesis. Genes Dev. 2002;16(15):1990–2005.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Mundy C, Bello A, Sgariglia F, Koyama E, Pacifici M. HhAntag, a hedgehog signaling antagonist, suppresses chondrogenesis and modulates canonical and non-canonical BMP signaling. J Cell Physiol. 2016;231(5):1033–44.

    Article  CAS  PubMed  Google Scholar 

  47. Siebel C, Lendahl U. Notch signaling in development, tissue homeostasis, and disease. Physiol Rev. 2017;97(4):1235–94.

    Article  CAS  PubMed  Google Scholar 

  48. Guasto A, Cormier-Daire V. Signaling pathways in bone development and their related skeletal dysplasia. Int J Mol Sci. 2021;22(9):4321.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Zanotti S, Canalis E. Notch signaling and the skeleton. Endocr Rev. 2016;37(3):223–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Pakvasa M, Haravu P, Boachie-Mensah M, Jones A, Coalson E, Liao J, Zeng Z, Wu D, Qin K, Wu X, et al. Notch signaling: its essential roles in bone and craniofacial development. Genes Dis. 2021;8(1):8–24.

    Article  PubMed  Google Scholar 

  51. Rutkowski TP, Kohn A, Sharma D, Ren Y, Mirando AJ, Hilton MJ. HES factors regulate specific aspects of chondrogenesis and chondrocyte hypertrophy during cartilage development. J Cell Sci. 2016;129(11):2145–55.

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Chen S, Tao J, Bae Y, Jiang MM, Bertin T, Chen Y, Yang T, Lee B. Notch gain of function inhibits chondrocyte differentiation via Rbpj-dependent suppression of Sox9. J Bone Miner Res. 2013;28(3):649–59.

    Article  PubMed  CAS  Google Scholar 

  53. Sun J, Luo Z, Wang G, Wang Y, Wang Y, Olmedo M, Morandi MM, Barton S, Kevil CG, Shu B, et al. Notch ligand Jagged1 promotes mesenchymal stromal cell-based cartilage repair. Exp Mol Med. 2018;50(9):1–10.

    Article  PubMed  CAS  Google Scholar 

  54. Dishowitz MI, Terkhorn SP, Bostic SA, Hankenson KD. Notch signaling components are upregulated during both endochondral and intramembranous bone regeneration. J Orthop Res. 2012;30(2):296–303.

    Article  CAS  PubMed  Google Scholar 

  55. Tian Y, Xu Y, Fu Q, Chang M, Wang Y, Shang X, Wan C, Marymont JV, Dong Y. Notch inhibits chondrogenic differentiation of mesenchymal progenitor cells by targeting Twist1. Mol Cell Endocrinol. 2015;403:30–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Xu W, Wang Y, Zhao H, Fan B, Guo K, Cai M, Zhang S. Delta-like 2 negatively regulates chondrogenic differentiation. J Cell Physiol. 2018;233(9):6574–82.

    Article  CAS  PubMed  Google Scholar 

  57. Girardi F, Le Grand F. Wnt signaling in skeletal muscle development and regeneration. Prog Mol Biol Transl Sci. 2018;153:157–79.

    Article  CAS  PubMed  Google Scholar 

  58. Clevers H, Nusse R. Wnt/beta-catenin signaling and disease. Cell. 2012;149(6):1192–205.

    Article  CAS  PubMed  Google Scholar 

  59. Dash S, Trainor PA. The development, patterning and evolution of neural crest cell differentiation into cartilage and bone. Bone. 2020;137:115409.

    Article  CAS  PubMed  Google Scholar 

  60. Huang P, Yan R, Zhang X, Wang L, Ke X, Qu Y. Activating Wnt/beta-catenin signaling pathway for disease therapy: challenges and opportunities. Pharmacol Ther. 2019;196:79–90.

    Article  CAS  PubMed  Google Scholar 

  61. Lerner UH, Ohlsson C. The WNT system: background and its role in bone. J Intern Med. 2015;277(6):630–49.

    Article  CAS  PubMed  Google Scholar 

  62. Koni M, Pinnaro V, Brizzi MF. The Wnt signalling pathway: a tailored target in cancer. Int J Mol Sci. 2020;21(20):7697.

    Article  CAS  PubMed Central  Google Scholar 

  63. Usami Y, Gunawardena AT, Iwamoto M, Enomoto-Iwamoto M. Wnt signaling in cartilage development and diseases: lessons from animal studies. Lab Invest. 2016;96(2):186–96.

    Article  CAS  PubMed  Google Scholar 

  64. Qu F, Wang J, Xu N, Liu C, Li S, Wang N, Qi W, Li H, Li C, Geng Z, et al. WNT3A modulates chondrogenesis via canonical and non-canonical Wnt pathways in MSCs. Front Biosci (Landmark Ed). 2013;18:493–503.

    Article  CAS  Google Scholar 

  65. Narcisi R, Cleary MA, Brama PAJ, Hoogduijn MJ, Tüysüz N, ten Berge D, van Osch GJVM. Long-term expansion, enhanced chondrogenic potential, and suppression of endochondral ossification of adult human MSCs via WNT signaling modulation. Stem Cell Rep. 2015;4(3):459–72.

    Article  CAS  Google Scholar 

  66. Zhang Y, Huang X, Yuan Y. MicroRNA-410 promotes chondrogenic differentiation of human bone marrow mesenchymal stem cells through down-regulating Wnt3a. Am J Transl Res. 2017;9(1):136–45.

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Qi Y, Zhang W, Li G, Niu L, Zhang Y, Tang R, Feng G. An oriented-collagen scaffold including Wnt5a promotes osteochondral regeneration and cartilage interface integration in a rabbit model. FASEB J. 2020;34(8):11115–32.

    Article  CAS  PubMed  Google Scholar 

  68. Jin E-J, Park J-H, Lee S-Y, Chun J-S, Bang O-S, Kang S-S. Wnt-5a is involved in TGF-beta3-stimulated chondrogenic differentiation of chick wing bud mesenchymal cells. Int J Biochem Cell Biol. 2006;38(2):183–95.

    Article  CAS  PubMed  Google Scholar 

  69. Tuli R, Tuli S, Nandi S, Huang X, Manner PA, Hozack WJ, Danielson KG, Hall DJ, Tuan RS. Transforming growth factor-beta-mediated chondrogenesis of human mesenchymal progenitor cells involves N-cadherin and mitogen-activated protein kinase and Wnt signaling cross-talk. J Biol Chem. 2003;278(42):41227–36.

    Article  CAS  PubMed  Google Scholar 

  70. Xu R, Zhang F, Lu J, Wang K, Pan P, Sun Y, Zhang Y. Secreted frizzled-related protein 3 was genetically and functionally associated with developmental dysplasia of the hip. Aging (Albany NY). 2021;13(8):11281–95.

    Article  CAS  Google Scholar 

  71. Narcisi R, Arikan OH, Lehmann J, Ten Berge D, van Osch GJ. Differential effects of small molecule WNT agonists on the multilineage differentiation capacity of human mesenchymal stem cells. Tissue Eng Part A. 2016;22(21–22):1264–73.

    Article  CAS  PubMed  Google Scholar 

  72. Beenken A, Mohammadi M. The FGF family: biology, pathophysiology and therapy. Nat Rev Drug Discov. 2009;8(3):235–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Ornitz DM, Marie PJ. Fibroblast growth factor signaling in skeletal development and disease. Genes Dev. 2015;29(14):1463–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Johnson DE, Williams LT. Structural and functional diversity in the FGF receptor multigene family. Adv Cancer Res. 1993;60:1–41.

    CAS  PubMed  Google Scholar 

  75. Itoh N, Ornitz DM. Fibroblast growth factors: from molecular evolution to roles in development, metabolism and disease. J Biochem. 2011;149(2):121–30.

    Article  CAS  PubMed  Google Scholar 

  76. Ornitz DM, Itoh N. The Fibroblast Growth Factor signaling pathway. Wiley Interdiscip Rev Dev Biol. 2015;4(3):215–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Chen TM, Chen YH, Sun HS, Tsai SJ. Fibroblast growth factors: potential novel targets for regenerative therapy of osteoarthritis. Chin J Physiol. 2019;62(1):2–10.

    Article  CAS  PubMed  Google Scholar 

  78. Turner N, Grose R. Fibroblast growth factor signalling: from development to cancer. Nat Rev Cancer. 2010;10(2):116–29.

    Article  CAS  PubMed  Google Scholar 

  79. Katoh M. Fibroblast growth factor receptors as treatment targets in clinical oncology. Nat Rev Clin Oncol. 2019;16(2):105–22.

    Article  CAS  PubMed  Google Scholar 

  80. Su N, Jin M, Chen L. Role of FGF/FGFR signaling in skeletal development and homeostasis: learning from mouse models. Bone Res. 2014;2:14003.

    Article  PubMed  PubMed Central  Google Scholar 

  81. Xie Y, Zinkle A, Chen L, Mohammadi M. Fibroblast growth factor signalling in osteoarthritis and cartilage repair. Nat Rev Rheumatol. 2020;16(10):547–64.

    Article  CAS  PubMed  Google Scholar 

  82. Hellingman CA, Koevoet W, Kops N, Farrell E, Jahr H, Liu W, Baatenburg de Jong RJ, Frenz DA, van Osch GJVM. Fibroblast growth factor receptors in in vitro and in vivo chondrogenesis: relating tissue engineering using adult mesenchymal stem cells to embryonic development. Tissue Eng Part A. 2010;16(2):545–56.

    Article  CAS  PubMed  Google Scholar 

  83. Chen P, Zhang L, Weng T, Zhang S, Sun S, Chang M, Li Y, Zhang B, Zhang L. A Ser252Trp mutation in fibroblast growth factor receptor 2 (FGFR2) mimicking human Apert syndrome reveals an essential role for FGF signaling in the regulation of endochondral bone formation. PLoS ONE. 2014;9(1):e87311.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  84. Pizzute T, Li J, Zhang Y, Davis ME, Pei M. Fibroblast growth factor ligand dependent proliferation and chondrogenic differentiation of synovium-derived stem cells and concomitant adaptation of Wnt/mitogen-activated protein kinase signals. Tissue Eng Part A. 2016;22(15–16):1036–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Solchaga LA, Penick K, Porter JD, Goldberg VM, Caplan AI, Welter JF. FGF-2 enhances the mitotic and chondrogenic potentials of human adult bone marrow-derived mesenchymal stem cells. J Cell Physiol. 2005;203(2):398–409.

    Article  CAS  PubMed  Google Scholar 

  86. Maruyama T, Mirando AJ, Deng C-X, Hsu W. The balance of WNT and FGF signaling influences mesenchymal stem cell fate during skeletal development. Sci Signal. 2010;3(123):ra40.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  87. Correa D, Somoza RA, Lin P, Greenberg S, Rom E, Duesler L, Welter JF, Yayon A, Caplan AI. Sequential exposure to fibroblast growth factors (FGF) 2, 9 and 18 enhances hMSC chondrogenic differentiation. Osteoarthr Cartil. 2015;23(3):443–53.

    Article  CAS  Google Scholar 

  88. Nguyen JKB, Eames BF. Evolutionary repression of chondrogenic genes in the vertebrate osteoblast. FEBS J. 2020;287(20):4354–61.

    Article  CAS  PubMed  Google Scholar 

  89. Song H, Park KH. Regulation and function of SOX9 during cartilage development and regeneration. Semin Cancer Biol. 2020;67(Pt 1):12–23.

    Article  CAS  PubMed  Google Scholar 

  90. Panda M, Tripathi SK, Biswal BK. SOX9: an emerging driving factor from cancer progression to drug resistance. Biochim Biophys Acta Rev Cancer. 2021;1875(2): 188517.

    Article  CAS  PubMed  Google Scholar 

  91. Lefebvre V. Roles and regulation of SOX transcription factors in skeletogenesis. Curr Top Dev Biol. 2019;133:171–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Akiyama H, Chaboissier M-C, Martin JF, Schedl A, de Crombrugghe B. The transcription factor Sox9 has essential roles in successive steps of the chondrocyte differentiation pathway and is required for expression of Sox5 and Sox6. Genes Dev. 2002;16(21):2813–28.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Wagner T, Wirth J, Meyer J, Zabel B, Held M, Zimmer J, Pasantes J, Bricarelli FD, Keutel J, Hustert E, et al. Autosomal sex reversal and campomelic dysplasia are caused by mutations in and around the SRY-related gene SOX9. Cell. 1994;79(6):1111–20.

    Article  CAS  PubMed  Google Scholar 

  94. Bi W, Huang W, Whitworth DJ, Deng JM, Zhang Z, Behringer RR, de Crombrugghe B. Haploinsufficiency of Sox9 results in defective cartilage primordia and premature skeletal mineralization. Proc Natl Acad Sci USA. 2001;98(12):6698–703.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Li IMH, Liu K, Neal A, Clegg PD, De Val S, Bou-Gharios G. Differential tissue specific, temporal and spatial expression patterns of the Aggrecan gene is modulated by independent enhancer elements. Sci Rep. 2018;8(1):950.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  96. Bar Oz M, Kumar A, Elayyan J, Reich E, Binyamin M, Kandel L, Liebergall M, Steinmeyer J, Lefebvre V, Dvir-Ginzberg M. Acetylation reduces SOX9 nuclear entry and ACAN gene transactivation in human chondrocytes. Aging Cell. 2016;15(3):499–508.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Yasuda H, Oh CD, Chen D, de Crombrugghe B, Kim JH. A novel regulatory mechanism of type II collagen expression via a SOX9-dependent Enhancer in Intron 6. J Biol Chem. 2017;292(2):528–38.

    Article  CAS  PubMed  Google Scholar 

  98. Bell DM, Leung KK, Wheatley SC, Ng LJ, Zhou S, Ling KW, Sham MH, Koopman P, Tam PP, Cheah KS. SOX9 directly regulates the type-II collagen gene. Nat Genet. 1997;16(2):174–8.

    Article  CAS  PubMed  Google Scholar 

  99. Grimm D, Bauer J, Wise P, Kruger M, Simonsen U, Wehland M, Infanger M, Corydon TJ. The role of SOX family members in solid tumours and metastasis. Semin Cancer Biol. 2020;67(Pt 1):122–53.

    Article  CAS  PubMed  Google Scholar 

  100. Lefebvre V, Li P, de Crombrugghe B. A new long form of Sox5 (L-Sox5), Sox6 and Sox9 are coexpressed in chondrogenesis and cooperatively activate the type II collagen gene. EMBO J. 1998;17(19):5718–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Lefebvre V, Angelozzi M, Haseeb A. SOX9 in cartilage development and disease. Curr Opin Cell Biol. 2019;61:39–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Zhou G, Lefebvre V, Zhang Z, Eberspaecher H, de Crombrugghe B. Three high mobility group-like sequences within a 48-base pair enhancer of the Col2a1 gene are required for cartilage-specific expression in vivo. J Biol Chem. 1998;273(24):14989–97.

    Article  CAS  PubMed  Google Scholar 

  103. Raftery RM, Gonzalez Vazquez AG, Chen G, O’Brien FJ. Activation of the SOX-5, SOX-6, and SOX-9 trio of transcription factors using a gene-activated scaffold stimulates mesenchymal stromal cell chondrogenesis and inhibits endochondral ossification. Adv Healthc Mater. 2020;9(10):e1901827.

    Article  PubMed  CAS  Google Scholar 

  104. Liu YF, Chen WM, Lin YF, Yang RC, Lin MW, Li LH, Chang YH, Jou YS, Lin PY, Su JS, et al. Type II collagen gene variants and inherited osteonecrosis of the femoral head. N Engl J Med. 2005;352(22):2294–301.

    Article  CAS  PubMed  Google Scholar 

  105. Zhang B, Zhang Y, Wu N, Li J, Liu H, Wang J. Integrated analysis of COL2A1 variant data and classification of type II collagenopathies. Clin Genet. 2020;97(3):383–95.

    Article  CAS  PubMed  Google Scholar 

  106. Dogan P, Varal IG, Gorukmez O, Akkurt MO, Akdag A. Achondrogenesis type 2 in a newborn with a novel mutation on the COL2A1 gene. Balkan J Med Genet. 2019;22(1):89–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Ng LJ, Wheatley S, Muscat GE, Conway-Campbell J, Bowles J, Wright E, Bell DM, Tam PP, Cheah KS, Koopman P. SOX9 binds DNA, activates transcription, and coexpresses with type II collagen during chondrogenesis in the mouse. Dev Biol. 1997;183(1):108–21.

    Article  CAS  PubMed  Google Scholar 

  108. Zhao Q, Eberspaecher H, Lefebvre V, De Crombrugghe B. Parallel expression of Sox9 and Col2a1 in cells undergoing chondrogenesis. Dev Dyn. 1997;209(4):377–86.

    Article  CAS  PubMed  Google Scholar 

  109. Kulyk WM, Franklin JL, Hoffman LM. Sox9 expression during chondrogenesis in micromass cultures of embryonic limb mesenchyme. Exp Cell Res. 2000;255(2):327–32.

    Article  CAS  PubMed  Google Scholar 

  110. Grogan SP, Olee T, Hiraoka K, Lotz MK. Repression of chondrogenesis through binding of notch signaling proteins HES-1 and HEY-1 to N-box domains in the COL2A1 enhancer site. Arthritis Rheum. 2008;58(9):2754–63.

    Article  PubMed  PubMed Central  Google Scholar 

  111. Yano F, Hojo H, Ohba S, Fukai A, Hosaka Y, Ikeda T, Saito T, Hirata M, Chikuda H, Takato T, et al. A novel disease-modifying osteoarthritis drug candidate targeting Runx1. Ann Rheum Dis. 2013;72(5):748–53.

    Article  CAS  PubMed  Google Scholar 

  112. Zhang X, Nham GTH, Ito K, Shinomura T. Gene expression of type II collagen is regulated by direct interaction with Kruppel-like factor 4 and AT-rich interactive domain 5B. Gene. 2021;773:145381.

    Article  CAS  PubMed  Google Scholar 

  113. Knudson CB, Knudson W. Cartilage proteoglycans. Semin Cell Dev Biol. 2001;12(2):69–78.

    Article  CAS  PubMed  Google Scholar 

  114. Schwartz NB, Pirok EW 3rd, Mensch JR Jr, Domowicz MS. Domain organization, genomic structure, evolution, and regulation of expression of the aggrecan gene family. Prog Nucleic Acid Res Mol Biol. 1999;62:177–225.

    Article  CAS  PubMed  Google Scholar 

  115. Sekiya I, Tsuji K, Koopman P, Watanabe H, Yamada Y, Shinomiya K, Nifuji A, Noda M. SOX9 enhances aggrecan gene promoter/enhancer activity and is up-regulated by retinoic acid in a cartilage-derived cell line, TC6. J Biol Chem. 2000;275(15):10738–44.

    Article  CAS  PubMed  Google Scholar 

  116. Han Y, Lefebvre V. L-Sox5 and Sox6 drive expression of the aggrecan gene in cartilage by securing binding of Sox9 to a far-upstream enhancer. Mol Cell Biol. 2008;28(16):4999–5013.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Takimoto A, Kokubu C, Watanabe H, Sakuma T, Yamamoto T, Kondoh G, Hiraki Y, Shukunami C. Differential transactivation of the upstream aggrecan enhancer regulated by PAX1/9 depends on SOX9-driven transactivation. Sci Rep. 2019;9(1):4605.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  118. Duan L, Liang Y, Xu X, Xiao Y, Wang D. Recent progress on the role of miR-140 in cartilage matrix remodelling and its implications for osteoarthritis treatment. Arthritis Res Ther. 2020;22(1):194.

    Article  PubMed  PubMed Central  Google Scholar 

  119. Karlsen TA, Jakobsen RB, Mikkelsen TS, Brinchmann JE. microRNA-140 targets RALA and regulates chondrogenic differentiation of human mesenchymal stem cells by translational enhancement of SOX9 and ACAN. Stem Cells Dev. 2014;23(3):290–304.

    Article  CAS  PubMed  Google Scholar 

  120. Pais H, Nicolas FE, Soond SM, Swingler TE, Clark IM, Chantry A, Moulton V, Dalmay T. Analyzing mRNA expression identifies Smad3 as a microRNA-140 target regulated only at protein level. RNA. 2010;16(3):489–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Won Lee G, Thangavelu M, Joung Choi M, Yeong Shin E, Sol Kim H, Seon Baek J, Woon Jeong Y, Eun Song J, Carlomagno C, Miguel Oliveira J, et al. Exosome mediated transfer of miRNA-140 promotes enhanced chondrogenic differentiation of bone marrow stem cells for enhanced cartilage repair and regeneration. J Cell Biochem. 2020;121(7):3642–52.

    Article  PubMed  CAS  Google Scholar 

  122. Tao SC, Yuan T, Zhang YL, Yin WJ, Guo SC, Zhang CQ. Exosomes derived from miR-140-5p-overexpressing human synovial mesenchymal stem cells enhance cartilage tissue regeneration and prevent osteoarthritis of the knee in a rat model. Theranostics. 2017;7(1):180–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Lin ST, Huang Y, Zhang L, Heng MY, Ptacek LJ, Fu YH. MicroRNA-23a promotes myelination in the central nervous system. Proc Natl Acad Sci USA. 2013;110(43):17468–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Yang Z, Ren Z, She R, Ao J, Wa Q, Sun Z, Li B, Tian X. miR-23a-3p regulated by LncRNA SNHG5 suppresses the chondrogenic differentiation of human adipose-derived stem cells via targeting SOX6/SOX5. Cell Tissue Res. 2021;383(2):723–33.

    Article  CAS  PubMed  Google Scholar 

  125. Hu H, Dong L, Bu Z, Shen Y, Luo J, Zhang H, Zhao S, Lv F, Liu Z. miR-23a-3p-abundant small extracellular vesicles released from Gelma/nanoclay hydrogel for cartilage regeneration. J Extracell Vesicles. 2020;9(1):1778883.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Ham O, Song B-W, Lee S-Y, Choi E, Cha M-J, Lee CY, Park J-H, Kim I-K, Chang W, Lim S, et al. The role of microRNA-23b in the differentiation of MSC into chondrocyte by targeting protein kinase A signaling. Biomaterials. 2012;33(18):4500–7.

    Article  CAS  PubMed  Google Scholar 

  127. Shen PF, Wang B, Qu YX, Zheng C, Xu JD, Xie ZK, Ma Y. MicroRNA-23c inhibits articular cartilage damage recovery by regulating MSCs differentiation to chondrocytes via reducing FGF2. Eur Rev Med Pharmacol Sci. 2019;23(3):941–8.

    PubMed  Google Scholar 

  128. Kretz M, Webster DE, Flockhart RJ, Lee CS, Zehnder A, Lopez-Pajares V, Qu K, Zheng GX, Chow J, Kim GE, et al. Suppression of progenitor differentiation requires the long noncoding RNA ANCR. Genes Dev. 2012;26(4):338–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Zhang L, Chen S, Bao N, Yang C, Ti Y, Zhou L, Zhao J. Sox4 enhances chondrogenic differentiation and proliferation of human synovium-derived stem cell via activation of long noncoding RNA DANCR. J Mol Histol. 2015;46(6):467–73.

    Article  CAS  PubMed  Google Scholar 

  130. Zhang L, Yang C, Chen S, Wang G, Shi B, Tao X, Zhou L, Zhao J. Long noncoding RNA DANCR is a positive regulator of proliferation and chondrogenic differentiation in human synovium-derived stem cells. DNA Cell Biol. 2017;36(2):136–42.

    Article  CAS  PubMed  Google Scholar 

  131. Fang P, Zhang LX, Hu Y, Zhang L, Zhou LW. Long non-coding RNA DANCR induces chondrogenesis by regulating the miR-1275/MMP-13 axis in synovial fluid-derived mesenchymal stem cells. Eur Rev Med Pharmacol Sci. 2019;23(23):10459–69.

    CAS  PubMed  Google Scholar 

  132. Zhang L, Sun X, Chen S, Yang C, Shi B, Zhou L, Zhao J. Long noncoding RNA DANCR regulates miR-1305-Smad 4 axis to promote chondrogenic differentiation of human synovium-derived mesenchymal stem cells. Biosci Rep. 2017;37(4):BSR20170347.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Nguyen NTK, Chang YH, Truong VA, Hsu MN, Pham NN, Chang CW, Wu YH, Chang YH, Li H, Hu YC. CRISPR activation of long non-coding RNA DANCR promotes bone regeneration. Biomaterials. 2021;275:120965.

    Article  CAS  PubMed  Google Scholar 

  134. Ishikawa T, Nishida T, Ono M, Takarada T, Nguyen HT, Kurihara S, Furumatsu T, Murase Y, Takigawa M, Oohashi T, et al. Physiological role of urothelial cancer-associated one long noncoding RNA in human skeletogenic cell differentiation. J Cell Physiol. 2018;233(6):4825–40.

    Article  CAS  PubMed  Google Scholar 

  135. Shu T, He L, Wang X, Pang M, Yang B, Feng F, Wu Z, Liu C, Zhang S, Liu B, et al. Long noncoding RNA UCA1 promotes chondrogenic differentiation of human bone marrow mesenchymal stem cells via miRNA-145-5p/SMAD5 and miRNA-124-3p/SMAD4 axis. Biochem Biophys Res Commun. 2019;514(1):316–22.

    Article  CAS  PubMed  Google Scholar 

  136. Della Bella E, Menzel U, Basoli V, Tourbier C, Alini M, Stoddart MJ. Differential regulation of circRNA, miRNA, and piRNA during early osteogenic and chondrogenic differentiation of human mesenchymal stromal cells. Cells. 2020;9(2):398.

    Article  CAS  PubMed Central  Google Scholar 

  137. Wang Y, Jiang Z, Yu M, Yang G. Roles of circular RNAs in regulating the self-renewal and differentiation of adult stem cells. Differentiation. 2020;113:10–8.

    Article  CAS  PubMed  Google Scholar 

  138. Zha K, Sun Z, Yang Y, Chen M, Gao C, Fu L, Li H, Sui X, Guo Q, Liu S. Recent developed strategies for enhancing chondrogenic differentiation of MSC: impact on MSC-based therapy for cartilage regeneration. Stem Cells Int. 2021;2021:8830834.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  139. Pattappa G, Schewior R, Hofmeister I, Seja J, Zellner J, Johnstone B, Docheva D, Angele P. Physioxia has a beneficial effect on cartilage matrix production in interleukin-1 beta-inhibited mesenchymal stem cell chondrogenesis. Cells. 2019;8(8):936.

    Article  CAS  PubMed Central  Google Scholar 

  140. Fahy N, Alini M, Stoddart MJ. Mechanical stimulation of mesenchymal stem cells: implications for cartilage tissue engineering. J Orthop Res. 2018;36(1):52–63.

    PubMed  Google Scholar 

  141. Yue D, Zhang M, Lu J, Zhou J, Bai Y, Pan J. The rate of fluid shear stress is a potent regulator for the differentiation of mesenchymal stem cells. J Cell Physiol. 2019;234(9):16312–9.

    Article  CAS  Google Scholar 

  142. Arora S, Srinivasan A, Leung CM, Toh YC. Bio-mimicking shear stress environments for enhancing mesenchymal stem cell differentiation. Curr Stem Cell Res Ther. 2020;15(5):414–27.

    Article  CAS  PubMed  Google Scholar 

  143. Pattappa G, Zellner J, Johnstone B, Docheva D, Angele P. Cells under pressure—the relationship between hydrostatic pressure and mesenchymal stem cell chondrogenesis. Eur Cell Mater. 2019;37:360–81.

    Article  CAS  PubMed  Google Scholar 

  144. Selig M, Lauer JC, Hart ML, Rolauffs B. Mechanotransduction and stiffness-sensing: mechanisms and opportunities to control multiple molecular aspects of cell phenotype as a design cornerstone of cell-instructive biomaterials for articular cartilage repair. Int J Mol Sci. 2020;21(15):5399.

    Article  CAS  PubMed Central  Google Scholar 

  145. Lee J, Jeon O, Kong M, Abdeen AA, Shin JY, Lee HN, Lee YB, Sun W, Bandaru P, Alt DS, et al. Combinatorial screening of biochemical and physical signals for phenotypic regulation of stem cell-based cartilage tissue engineering. Sci Adv. 2020;6(21):eaaz5913.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Clarke MF. Clinical and therapeutic implications of cancer stem cells. N Engl J Med. 2019;380(23):2237–45.

    Article  CAS  PubMed  Google Scholar 

  147. Zhou HM, Zhang JG, Zhang X, Li Q. Targeting cancer stem cells for reversing therapy resistance: mechanism, signaling, and prospective agents. Signal Transduct Target Ther. 2021;6(1):62.

    Article  PubMed  PubMed Central  Google Scholar 

  148. Das PK, Pillai S, Rakib MA, Khanam JA, Gopalan V, Lam AKY, Islam F. Plasticity of cancer stem cell: origin and role in disease progression and therapy resistance. Stem Cell Rev Rep. 2020;16(2):397–412.

    Article  PubMed  Google Scholar 

  149. Cojoc M, Mabert K, Muders MH, Dubrovska A. A role for cancer stem cells in therapy resistance: cellular and molecular mechanisms. Semin Cancer Biol. 2015;31:16–27.

    Article  CAS  PubMed  Google Scholar 

  150. Sarmento-Ribeiro AB, Scorilas A, Goncalves AC, Efferth T, Trougakos IP. The emergence of drug resistance to targeted cancer therapies: clinical evidence. Drug Resist Updat. 2019;47:100646.

    Article  PubMed  Google Scholar 

  151. Garcia-Mayea Y, Mir C, Masson F, Paciucci R, Lleonart ME. Insights into new mechanisms and models of cancer stem cell multidrug resistance. Semin Cancer Biol. 2020;60:166–80.

    Article  CAS  PubMed  Google Scholar 

  152. Arima Y, Nobusue H, Saya H. Targeting of cancer stem cells by differentiation therapy. Cancer Sci. 2020;111(8):2689–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Liu Q, Nguyen DH, Dong Q, Shitaku P, Chung K, Liu OY, Tso JL, Liu JY, Konkankit V, Cloughesy TF, et al. Molecular properties of CD133+ glioblastoma stem cells derived from treatment-refractory recurrent brain tumors. J Neurooncol. 2009;94(1):1–19.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  154. Galli R, Binda E, Orfanelli U, Cipelletti B, Gritti A, De Vitis S, Fiocco R, Foroni C, Dimeco F, Vescovi A. Isolation and characterization of tumorigenic, stem-like neural precursors from human glioblastoma. Cancer Res. 2004;64(19):7011–21.

    Article  CAS  PubMed  Google Scholar 

  155. Suva ML, Riggi N, Stehle JC, Baumer K, Tercier S, Joseph JM, Suva D, Clement V, Provero P, Cironi L, et al. Identification of cancer stem cells in Ewing’s sarcoma. Cancer Res. 2009;69(5):1776–81.

    Article  CAS  PubMed  Google Scholar 

  156. Tirino V, Desiderio V, Paino F, De Rosa A, Papaccio F, Fazioli F, Pirozzi G, Papaccio G. Human primary bone sarcomas contain CD133+ cancer stem cells displaying high tumorigenicity in vivo. FASEB J. 2011;25(6):2022–30.

    Article  CAS  PubMed  Google Scholar 

  157. Wang L, Guo H, Lin C, Yang L, Wang X. Enrichment and characterization of cancer stem-like cells from a cervical cancer cell line. Mol Med Rep. 2014;9(6):2117–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Zhang S, Mercado-Uribe I, Xing Z, Sun B, Kuang J, Liu J. Generation of cancer stem-like cells through the formation of polyploid giant cancer cells. Oncogene. 2014;33(1):116–28.

    Article  CAS  PubMed  Google Scholar 

  159. Royer-Pokora B, Busch M, Beier M, Duhme C, de Torres C, Mora J, Brandt A, Royer H-D. Wilms tumor cells with WT1 mutations have characteristic features of mesenchymal stem cells and express molecular markers of paraxial mesoderm. Hum Mol Genet. 2010;19(9):1651–68.

    Article  CAS  PubMed  Google Scholar 

  160. Chen T, You Y, Jiang H, Wang ZZ. Epithelial-mesenchymal transition (EMT): a biological process in the development, stem cell differentiation, and tumorigenesis. J Cell Physiol. 2017;232(12):3261–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  161. Battula VL, Evans KW, Hollier BG, Shi Y, Marini FC, Ayyanan A, Wang RY, Brisken C, Guerra R, Andreeff M, et al. Epithelial-mesenchymal transition-derived cells exhibit multilineage differentiation potential similar to mesenchymal stem cells. Stem Cells. 2010;28(8):1435–45.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  162. Manupati K, Dhoke NR, Debnath T, Yeeravalli R, Guguloth K, Saeidpour S, De UC, Debnath S, Das A. Inhibiting epidermal growth factor receptor signalling potentiates mesenchymal–epithelial transition of breast cancer stem cells and their responsiveness to anticancer drugs. FEBS J. 2017;284(12):1830–54.

    Article  CAS  PubMed  Google Scholar 

  163. Bhat KP, Salazar KL, Balasubramaniyan V, Wani K, Heathcock L, Hollingsworth F, James JD, Gumin J, Diefes KL, Kim SH, et al. The transcriptional coactivator TAZ regulates mesenchymal differentiation in malignant glioma. Genes Dev. 2011;25(24):2594–609.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  164. Verhaak RGW, Hoadley KA, Purdom E, Wang V, Qi Y, Wilkerson MD, Miller CR, Ding L, Golub T, Mesirov JP, et al. Integrated genomic analysis identifies clinically relevant subtypes of glioblastoma characterized by abnormalities in PDGFRA, IDH1, EGFR, and NF1. Cancer Cell. 2010;17(1):98–110.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  165. Giladi ND, Ziv-Av A, Lee HK, Finniss S, Cazacu S, Xiang C, Waldman Ben-Asher H, deCarvalho A, Mikkelsen T, Poisson L, et al. RTVP-1 promotes mesenchymal transformation of glioma via a STAT-3/IL-6-dependent positive feedback loop. Oncotarget. 2015;6(26):22680–97.

    Article  PubMed  PubMed Central  Google Scholar 

  166. Wang Y, Sul HS. Pref-1 regulates mesenchymal cell commitment and differentiation through Sox9. Cell Metab. 2009;9(3):287–302.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  167. Carro MS, Lim WK, Alvarez MJ, Bollo RJ, Zhao X, Snyder EY, Sulman EP, Anne SL, Doetsch F, Colman H, et al. The transcriptional network for mesenchymal transformation of brain tumours. Nature. 2010;463(7279):318–25.

    Article  CAS  PubMed  Google Scholar 

  168. Shah S, Pocard M, Mirshahi M. Targeting the differentiation of gastric cancer cells (KATO-III) downregulates epithelial–mesenchymal and cancer stem cell markers. Oncol Rep. 2019;42(2):670–8.

    CAS  PubMed  PubMed Central  Google Scholar 

  169. Derfoul A, Perkins GL, Hall DJ, Tuan RS. Glucocorticoids promote chondrogenic differentiation of adult human mesenchymal stem cells by enhancing expression of cartilage extracellular matrix genes. Stem Cells. 2006;24(6):1487–95.

    Article  CAS  PubMed  Google Scholar 

  170. Cigan AD, Nims RJ, Albro MB, Esau JD, Dreyer MP, Vunjak-Novakovic G, Hung CT, Ateshian GA. Insulin, ascorbate, and glucose have a much greater influence than transferrin and selenous acid on the in vitro growth of engineered cartilage in chondrogenic media. Tissue Eng Part A. 2013;19(17–18):1941–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Gweon EJ, Kim S-J. Resveratrol attenuates matrix metalloproteinase-9 and -2-regulated differentiation of HTB94 chondrosarcoma cells through the p38 kinase and JNK pathways. Oncol Rep. 2014;32(1):71–8.

    Article  CAS  PubMed  Google Scholar 

  172. Ma N, Teng X, Zheng Q, Chen P. The regulatory mechanism of p38/MAPK in the chondrogenic differentiation from bone marrow mesenchymal stem cells. J Orthop Surg Res. 2019;14(1):434.

    Article  PubMed  PubMed Central  Google Scholar 

  173. Nwaeburu CC, Abukiwan A, Zhao Z, Herr I. Quercetin-induced miR-200b-3p regulates the mode of self-renewing divisions in pancreatic cancer. Mol Cancer. 2017;16(1):23.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  174. Ying J, Xu T, Wang C, Jin H, Tong P, Guan J, Abu-Amer Y, O’Keefe R, Shen J. Dnmt3b ablation impairs fracture repair through upregulation of Notch pathway. JCI Insight. 2020;5(3):e131816.

    Article  PubMed Central  Google Scholar 

  175. Underhill TM, Weston AD. Retinoids and their receptors in skeletal development. Microsc Res Tech. 1998;43(2):137–55.

    Article  CAS  PubMed  Google Scholar 

  176. de The H. Differentiation therapy revisited. Nat Rev Cancer. 2018;18(2):117–27.

    Article  PubMed  CAS  Google Scholar 

  177. Grignani F, Ferrucci PF, Testa U, Talamo G, Fagioli M, Alcalay M, Mencarelli A, Grignani F, Peschle C, Nicoletti I, et al. The acute promyelocytic leukemia-specific PML-RAR alpha fusion protein inhibits differentiation and promotes survival of myeloid precursor cells. Cell. 1993;74(3):423–31.

    Article  CAS  PubMed  Google Scholar 

  178. Hong Q, Li X-D, Xie P, Du S-X. All-trans-retinoic acid suppresses rat embryo hindlimb bud mesenchymal chondrogenesis by modulating HoxD9 expression. Bioengineered. 2021;12(1):3900–11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Sumitani Y, Uchibe K, Yoshida K, Weng Y, Guo J, Yuan H, Ikegame M, Kamioka H, Okamura H. Inhibitory effect of retinoic acid receptor agonists on in vitro chondrogenic differentiation. Anat Sci Int. 2020;95(2):202–8.

    Article  CAS  PubMed  Google Scholar 

  180. Inubushi T, Lemire I, Irie F, Yamaguchi Y. Palovarotene inhibits osteochondroma formation in a mouse model of multiple hereditary exostoses. J Bone Miner Res. 2018;33(4):658–66.

    Article  CAS  PubMed  Google Scholar 

  181. Naveen CR, Gaikwad S, Agrawal-Rajput R. Berberine induces neuronal differentiation through inhibition of cancer stemness and epithelial-mesenchymal transition in neuroblastoma cells. Phytomedicine. 2016;23(7):736–44.

    Article  CAS  PubMed  Google Scholar 

  182. Zhu TZ, Li XM, Luo LH, Song ZQ, Gao X, Li ZQ, Su JY, Liang GB. beta-elemene inhibits stemness, promotes differentiation and impairs chemoresistance to temozolomide in glioblastoma stem-like cells. Int J Oncol. 2014;45(2):699–709.

    Article  PubMed  CAS  Google Scholar 

  183. Zhu T, Li X, Luo L, Wang X, Li Z, Xie P, Gao X, Song Z, Su J, Liang G. Reversion of malignant phenotypes of human glioblastoma cells by β-elemene through β-catenin-mediated regulation of stemness-, differentiation- and epithelial-to-mesenchymal transition-related molecules. J Transl Med. 2015;13:356.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  184. Gravina GL, Mancini A, Colapietro A, Vitale F, Vetuschi A, Pompili S, Rossi G, Marampon F, Richardson PJ, Patient L, et al. The novel CXCR4 antagonist, PRX177561, reduces tumor cell proliferation and accelerates cancer stem cell differentiation in glioblastoma preclinical models. Tumour Biol. 2017;39(6):1010428317695528.

    Article  PubMed  CAS  Google Scholar 

  185. Sakariassen PØ, Immervoll H, Chekenya M. Cancer stem cells as mediators of treatment resistance in brain tumors: status and controversies. Neoplasia. 2007;9(11):882–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  186. Gupta PB, Onder TT, Jiang G, Tao K, Kuperwasser C, Weinberg RA, Lander ES. Identification of selective inhibitors of cancer stem cells by high-throughput screening. Cell. 2009;138(4):645–59.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  187. Hii L-W, Chung FF-L, Soo JS-S, Tan BS, Mai C-W, Leong C-O. Histone deacetylase (HDAC) inhibitors and doxorubicin combinations target both breast cancer stem cells and non-stem breast cancer cells simultaneously. Breast Cancer Res Treat. 2020;179(3):615–29.

    Article  CAS  PubMed  Google Scholar 

  188. Zheng BN, Ding CH, Chen SJ, Zhu K, Shao J, Feng J, Xu WP, Cai LY, Zhu CP, Duan W, et al. Targeting PRMT5 activity inhibits the malignancy of hepatocellular carcinoma by promoting the transcription of HNF4alpha. Theranostics. 2019;9(9):2606–17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. Courtois S, Durán RV, Giraud J, Sifré E, Izotte J, Mégraud F, Lehours P, Varon C, Bessède E. Metformin targets gastric cancer stem cells. Eur J Cancer. 2017;84:193–201.

    Article  CAS  PubMed  Google Scholar 

  190. Pirozzi CJ, Yan H. The implications of IDH mutations for cancer development and therapy. Nat Rev Clin Oncol. 2021;18(10):645–61.

    Article  CAS  PubMed  Google Scholar 

  191. Zhu AX, Macarulla T, Javle MM, Kelley RK, Lubner SJ, Adeva J, Cleary JM, Catenacci DVT, Borad MJ, Bridgewater JA, et al. Final overall survival efficacy results of ivosidenib for patients with advanced cholangiocarcinoma with IDH1 Mutation: the phase 3 randomized clinical ClarIDHy trial. JAMA Oncol. 2021;7(11):1669–77.

    Article  PubMed  Google Scholar 

  192. Mellinghoff IK, Ellingson BM, Touat M, Maher E, De La Fuente MI, Holdhoff M, Cote GM, Burris H, Janku F, Young RJ, et al. Ivosidenib in isocitrate dehydrogenase 1mutated advanced glioma. J Clin Oncol. 2020;38(29):3398–406.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We acknowledge Editage service for the manuscript language edit.

Funding

This work was supported in part by grants from the National Science Foundation of China (#82173283 and #82103088), and Foundation of committee on science and technology of Tianjin (#20JCYBJC01230). The funders had no roles in the design of the study, data collection, analysis and interpretation, or decision to write and publish the work.

Author information

Authors and Affiliations

Authors

Contributions

SZ designed the paper; contributed to manuscript writing; and approved the manuscript before submission. XY and FT contributed to manuscript writing; and approved the manuscript before submission. LF, RN and MY collected literatures and approved the manuscript before submission. SC and MY gave constructive comments on the manuscript, and approved the manuscript before submission. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Shiwu Zhang.

Ethics declarations

Ethics approval and consent to participate

This study was submitted to the ethics committee of Tianjin Union Medical Center for review and approval prior to the start of the clinical study. The use of human tissue samples was approved by the Hospital Review Board and the confidentiality of patient information was maintained.

Consent for publication

All the authors have approved the manuscript and agree with submission.

Competing interests

The authors declare that there is no conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yang, X., Tian, S., Fan, L. et al. Integrated regulation of chondrogenic differentiation in mesenchymal stem cells and differentiation of cancer cells. Cancer Cell Int 22, 169 (2022). https://doi.org/10.1186/s12935-022-02598-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12935-022-02598-8

Keywords