<<

RETRACTED www.nature.com/scientificreports

OPEN The Role of Scleraxis in Fate Determination of Mesenchymal Stem Cells for Tenocyte received: 19 March 2015 accepted: 02 June 2015 Differentiation Published: 20 August 2015 Yonghui Li1,*, Melissa Ramcharan1,2,*, Zuping Zhou1,*,†, Daniel J. Leong1,2,3,4, Takintope Akinbiyi1, Robert J. Majeska2 & Hui B. Sun1,3,4

Mesenchymal stem cells (MSCs) are pluripotent cells that primarily differentiate into osteocytes, chondrocytes, and adipocytes. Recent studies indicate that MSCs can also be induced to generate tenocyte-like cells; moreover, MSCs have been suggested to have great therapeutic potential for tendon pathologies. Yet the precise molecular cascades governing tenogenic differentiation of MSCs remain unclear. We demonstrate scleraxis, a critically involved in embryonic tendon development and formation, plays a pivotal role in the fate determination of MSC towards tenocyte differentiation. Using murine C3H10T1/2 pluripotent stem cells as a model system, we show scleraxis is extensively expressed in the early phase of bone morphogenetic (BMP)-12- triggered tenocytic differentiation. Once induced, scleraxis directly transactivates tendon lineage- related genes such as tenomodulin and suppresses osteogenic, chondrogenic, and adipogenic capabilities, thus committing C3H10T1/2 cells to differentiate into the specific tenocyte-like lineage, while eliminating plasticity for other lineages. We also reveal that mechanical loading-mediated tenocytic differentiation follows a similar pathway and that BMP-12 and cyclic uniaxial strain act in an additive fashion to augment the maximal response by activating signal transducer Smad8. These results provide critical insights into the determination of multipotent stem cells to the tenocyte lineage induced by both chemical and physical signals.

Mesenchymal stem cells (MSCs) can give rise to mesenchymal lineages such as osteocytes, chondrocytes, and adipocytes, and potentially transdifferentiate into non-mesenchymal cell types such as pancreatic cells and cardiomyocytes1–4. MSCs are considered the ideal source of cellular therapeutic agents for tendon repair5–8. Although increasing studies show that MSCs can be induced to differentiate into tenocytes5,9,10, the potential regulators that govern MSC fate for tenocyte differentiation in tendon neoformation and regeneration have not yet been defined. A number of growth factors, particularly members of the bone morphogenetic (BMP) family, have been shown to promote tenocytic differentiation from the multipotent MSCs9,11,12. Of the more than 20 BMP members13, BMP-12 appears to be the most potent tendon inducers identified thus far. BMP-12 gene transfer augments the repair of lacerated tendon9,14 and improves Achilles tendon healing15. In vitro, bone marrow MSCs transfected or stimulated with BMP-12 exhibit efficient generation of tenocyte-like cells6,16. In addition to chemical factors, stimulatory effects of

1Leni and Peter W. May Department of Orthopedics, Mount Sinai School of Medicine, New York, NY 10029. 2Department of Biomedical Engineering, City College of New York, New York, NY 10031. 3Department of Orthopedic Surgery, Albert Einstein College of Medicine, Bronx, NY 10461. 4Department of Radiation Oncology, Albert Einstein College of Medicine, Bronx, NY 10461. *These authors contributed equally to this work. †Current Address: Center for Stem cell and Regenerative Medicine Guangxi Normal University, Guilin 541004, China. Correspondence and requests for materials should be addressed to H.B.S. (email: [email protected])

SCiENTifiC RePorTS | 5:13149 | DOI: 10.1038/srep13149 1 www.nature.com/scientificreports/ RETRACTED

mechanical loading on the tenocytic differentiation have also been described17–19. Despite these efforts, however, the molecular mechanisms responsible for the directed differentiation of MSCs into tenocytes are not well understood. During embryogenesis, many transcriptional factors such as Six1 and Six2, EphA4, Eya1 and Eya2, follistatin, tenascin, mohawk, and scleraxis are recognized to regulate tendon development and for- mation20–22. Among them, scleraxis a class II basic helix-loop-helix (bHLH) transcription factor, has received particular attention23. Scleraxis (Scx) is expressed in both stem cells and differentiated cells of tendons. At E10.5 endogenous Scx can be readily detected in the developing limb buds and in a somatic compartment known as syndetome at the dorsolateral edge of the sclerotome24–26, and its expression is continued throughout later stages of tendon formation24. Conditional knockout of the Scx gene in mice causes a failure in the condensation of tendon stem cells and severe defects in tendon differentiation, and consequently a dramatic loss of tendons that are particularly responsible for transmitting muscu- loskeletal force in the limbs, tail and trunk27. These studies highlight an essential role of Scx in tendon development, and suggest that Scx may be critically involved in fate determination of adult MSCs to differentiate into tenocytes. Here we showed that Scx is a key regulator of MSC differentiation into the tenocyte lineage as evidenced by its critical involvement both in the activation of downstream genes and suppression of non-tenogenic pathways in C3H10T1/2 cells. Our study provides evidence suggesting that the tenocytic differentiation process of adult MSCs may recapture the molecular cascades operative in embryonic tendon formation. Results Scleraxis is induced by BMP-12 in C3H10T1/2 cells. To investigate the involvement of Scx in the tenogenic differentiation of MSCs, we first determined whether BMP-12 would induce Scx in C3H10T1/2 stem cells. As assessed by RT-qPCR, BMP-12 treatment led to strong expression of Scx and Tenomodulin (Tnmd), a marker related to later stage tendon formation28. The induction of both genes were dependent on the dose of BMP-12; increases in expression of Scx and Tnmd were detectable at 2.5 ng/ml and - imal (2.5 ~ 3-fold over untreated controls) at 10–25 ng/ml (Fig. 1A). Simultaneously, BMP-12 downregu- lated the expression of neucleosteimin (Nst), a pan-stem cell marker (Fig. 1B). Tnmd is subject to positive regulation by Scx in a tendon cell lineage-dependent manner28. Consistent with this behavior, induction of Tnmd lagged behind induction of Scx following BMP-12 treatment, as evidenced at the mRNA level by RT-qPCR (Fig. 1C) and at the protein level by immunostaining (Fig. 1D).

Scleraxis is required for tenomodulin expression in C3H10T1/2 cells. Based on the sequential expression of Scx and Tnmd pattern and previous findings that Scx stimulates Tnmd expression28, we next determined whether Tnmd expression in C3H10T1/2 cells could be directly regulated by Scx by examining Tnmd expression under conditions of Scx gain- and loss-of-function. In transient transfec- tion assays, C3H10T1/2 cells overexpressing Scx showed significantly upregulated Tnmd compared to untransfected and empty vector-transfected cells (Fig. 2A), providing evidence that simply increasing the level of Scx can mimic the inductive effect of BMP-12. On the other hand, when Scx expres- sion was suppressed by transfection with Scx siRNA, the BMP-12-dependent upregulation of both Scx and Tnmd seen in untransfected and scrambled siRNA-transfected cells was abolished (Fig. 2B). Furthermore, while transfection with Tnmd siRNA blocked BMP-12-induced Tnmd expression, it had no effect on expression of Scx (Fig. 2B). Similar results were also obtained at protein levels by Western blotting (Fig. 2C).

Scleraxis directly transactivates tenomodulin in C3H10T1/2 cells. Scx modulates the expression of target genes by binding to a consensus sequence (CANNTG) called E-box within promoters29. To test whether Scx might directly transactivate the transcription of Tnmd via this mechanism, we car- ried out electrophoretic mobility shift assays (EMSA) using olgionucleotide probes containing a putative E-box and sequences from the Tnmd promoter flanking both sides of the E-box. As shown in Fig. 3A, a protein-DNA complex was detected only when probes were incubated with nuclear extracts prepared from cells expressing high levels of Scx due to transfection (Fig. 3A, lane 3) or BMP-12-treatment (Fig. 3A, lane 4). The shifted bands became undetectable in the presence of excess unlabeled wild-type (Fig. 3A, lane 6, 7) or mutant probes (Fig. 3A, lane 9, 10). To identify specific E-box sites in the Tnmd promoter that interact with Scx, we further conducted chromatin immunoprecipitation (ChIP) assays. Of the five putative E-boxes found within the− 3000 bp region of the Tnmd promoter, E-boxes 1 to 4, but not E-box 5, demonstrated functional interactions with Scx (Fig. 3B).

Scleraxis promotes tenocytic differentiation by suppressing non-tenogenic potentials of C3H10T1/2 cells. Loss of Nst expression in C3H10T1/2 cells stimulated by BMP-12 suggests that tenocyte differentiation may be associated with loss of stem cell features, which are likely reflected by a decline in the capacity to differentiate along alternative pathways. Therefore, we next determined whether Scx was involved in the suppression of non-tenogenic potential. RT-qPCR analysis showed in C3H10T1/2 cells plated at colony-forming densities and treated with varying concentrations of

SCiENTifiC RePorTS | 5:13149 | DOI: 10.1038/srep13149 2 www.nature.com/scientificreports/ RETRACTED

Figure 1. Induction of scleraxis and tenomodulin by BMP-12 in C3H10T1/2 stem cells. (A) Upregulation of Scx and Tnmd. (B) Downregulation of Nst Expression. Cells were left untreated or treated with different concentrations of BMP-12 for 24 hours. mRNA levels were determined by RT-qPCR. Data are expressed as mean ± S.D. (n = 6). *p < 0.05. (C,D) Time course of Scx and Tnmd expression. Cells were treated with 10 ng/ml of BMP-12 for different lengths of time. mRNA (C) and protein (D) levels of gene expression were determined by RT-qPCR and immunostaining, respectively. Data in (C) are expressed as mean ± S.D. (n = 3).

BMP-12, Scx-expressing colonies increased in a BMP-12 dose-dependent manner (Supplementary Fig. S1), although not all of them expressed Scx. We further tested whether Scx-positive (Scx+) and Scx-negative (Scx−) colonies differ in their dif- ferentiation potentials using a single CFU-F re-plating assay (SCRA) (Fig. 4A). When cells from indi- vidual Scx-positive (Scx+) and Scx-negative (Scx−) colonies were re-plated in lineage-specific induction media, cytochemical staining for lineage specific markers revealed that the osteogenic, chondrogenic and adipogenic differentiation were largely suppressed in Scx+ colony-derived cells, but maintained in cells from Scx− colonies (Fig. 4B,C). These results were further confirmed by assessing mRNA levels of lineage-specific genes (Fig. 4D). To test a direct role for Scx and/or Tnmd in suppressing non-tenocytic activities, the loss of func- tion of both genes were examined. As shown in Fig. 5, marked increases in the expression of osteo-, chondro-, and adipose-specific genes were detected inScx siRNA transfected cells in the presence of induction media. The expression of these genes was increased in untransfected and scrambled siRNA transfected cells that received BMP-12 treatment to induce Scx prior to lineage induction, or in those without induction. These data show Scx is a negative regulator of the osteogenic, chondrogenic, and adi- pogenic pathways. Notably, silencing of Tnmd in the presence of high levels of Scx did not fully suppress non-tenogenic gene expression, especially chondrocyte- and adipocyte markers, showing that Tnmd may also participate in suppression of non-tenocytic potentials.

SCiENTifiC RePorTS | 5:13149 | DOI: 10.1038/srep13149 3 www.nature.com/scientificreports/ RETRACTED

Figure 2. Scleraxis-dependent expression of tenomodulin in C3H10T1/2 cells. (A) Overexpression of Scx upregulated Tnmd. Cells were left untransfected or transiently transfected with empty vectors or vectors containing Scx or Tnmd cDNA. mRNA levels were determined by RT-qPCR. Data are expressed as mean ± S.D. (n = 3). *p < 0.01. (B,C) Knockdown of Scx blocked BMP-12-induced Tnmd. C3H10T1/2 cells untransfected or transfected with scramble siRNA, or Scx siRNA, or Tnmd siRNA were left untreated or stimulated with BMP-12, and mRNA and protein levels of gene expression were determined by RT-qPCR and Western blot, respectively. Data in (B) are expressed as mean ± S.D. (n = 6). *p < 0.05.

Combining cyclic loading and BMP-12 enhances expression of Scx and formation of tenocyte-like cells. Mechanical stress has been shown to influence tenocytic differentiation from MSCs in vitro19,30 and from tendon stem and progenitor cells in vivo18,31–34. Therefore, we further tested whether mechanical loading also modulates tenocyte differentiation in C3H10T1/2, and whether such a physical stimulus can act synergistically or additively with BMP-12. We first assessed the effects of cyclic uniaxial strain loading on Scx expression. Compared to non-loaded controls, significant increases in Scx expression were observed in all loading-stimulated groups of cells with maximal expression in cells loaded at 5% strain (Supplementary Fig. S2A) at 0.5 Hz (Supplementary Fig. S2B). We next examined the combined effects of BMP-12 and stretch loading. Consistent with experiments described above, both mechanical loading and BMP-12 treatment resulted in significant increases in Scx mRNA expression though the latter provoked a more pronounced response. When the two stimuli were applied simultaneously, however, the increases in gene expression appeared to be additive (Fig. 6A). Toluidine blue staining revealed that cells treated with a combination of mechanical loading plus BMP- 12- appeared to resemble more closely mature tenocytes (Fig. 6B) as evidenced by a greater elongation (Fig. 6C) and cell alignment (Fig. 6D). The data show that physical and biochemical signals can interact cooperatively to facilitate differentiation into tenocytic cells.

Smad8 is an upstream signaling molecule of Scx required to mediate the teno-inductive response of cyclic loading and BMP-12. Members of the -regulated Smads (R-Smads) sub- class, Smad1/5/8, are the major signaling transducers for BMPs35. We monitored the phosphorylation of these proteins in cells under different treatment conditions. While BMP-12 and stretch loading each induced extensive phosphorylation of Smad 1/5/8, significantly increased R-Smad phosphorylation was observed over the two time-points in cells that had received both stimuli (Fig. 7A). The data demonstrate that BMP-12 and mechanical loading signaling may interact at the R-Smad site, thus leading to augmen- tation of tenocytic differentiation.

SCiENTifiC RePorTS | 5:13149 | DOI: 10.1038/srep13149 4 www.nature.com/scientificreports/ RETRACTED

Figure 3. Transactivation of tenomodulin by scleraxis in C3H10T1/2 cells. (A) Binding of Scx to Tnmd E-boxes assessed by electrophoretic mobility shift assays (EMSA) Olgionucleotide probes corresponding to a specific E-box consensus sequence were incubated with nuclear extracts from C3H10T1/2 cells with either Scx overexpression and/or BMP-12 treatment or left untreated (control). B( ) Binding of Scx to Tnmd E-boxes assessed by chromatin immunoprecipitation (ChIP) assays. Cross-linked chromatins were prepared by immuneprecipitated cellular lysates using anti-scleraxis antibodies and purified DNA was used to PCR amplify the biding regions of Tnmd promoter. The binding of Scx to all testedTnmd E-boxes except 5 was detected in BMP-12-treated cells. Upper panel, distribution and localization of E-boxes in Tnmd promoter.

To pinpoint the Smad molecule specifically involved in the mechanical loading and BMP-12 induc- tion of Scx, we assessed the effects of Smad1, Smad5 or Smad8 suppression. Phosphorylation of Smad 1/5/8 mediated by BMP-12, loading, or both, was almost entirely repressed upon Smad8 knockdown by siRNAs (Fig. 7B), concurrent with absence of Scx induction associated with the treatments (Fig. 7C). In contrast, Smad1 or Smad5 knockdown (Fig. 7D) had minimal impact on the overall phosphorylation of Smad 1/5/8 (Fig. 7E). Discussion Stem cell-based approaches are considered to have a great potential to improve healing of soft tissue injuries, which account for nearly half of the 33 million musculoskeletal injuries that occur in the United States each year36. The effective use of MSCs relies on a better understanding of the signaling pathways that control their differentiation into tenocytes. In the present study, we have demonstrated for the first time that scleraxis, a class B bHLH transcription factor, is critical in determining the fate of C3H10T1/2 stem cells to commit to a tenocyte lineage.

SCiENTifiC RePorTS | 5:13149 | DOI: 10.1038/srep13149 5 www.nature.com/scientificreports/ RETRACTED

Figure 4. Scleraxis induction is associated with loss of non-tenogenic potential of C3H10T1/2 cells. (A) Schematic procedures of single CFU-F replating assay (SCRA). (B) Cytochemical staining of osteo-(left panel), chondro-(middle panel), and adipo-(right panel) differentiation. C( ) Immunohistochemical staining of lineage-specific markers. Data shown in B and C are representatives of three experiments. D( ) mRNA expression of lineage-specific marker genes determined by RT-qPCR. Data are expressed as mean ± S.D. (n = 6). *p < 0.01.

We showed that the induction of Scx was a key early event occurring in the BMP-12-mediated ten- ocytic differentiation of C3H10T1/2 cells and required to regulate the expression of downstream gene Tnmd, a well characterized marker related to a late stage of tenocyte differentiation28. This conclusion was supported by several lines of evidence. First, treatment with recombinant BMP-12 proteins led to sequen- tial expression of Scx and Tnmd in C3H10T1/2 stem cells. Second, overexpression of Scx in C3H10T1/2 cells simulated the effect of BMP-12 on the upregulation ofTnmd whereas knockdown of Scx abolished BMP-12 induced Tnmd, indicating that the advent and presence of Scx is substantive to the ensuing expression of Tnmd. Third, altering the expression levels of Tnmd had minimal impact on Scx induction as demonstrated in comparable experiments on Tnmd gain- and loss-of-function. Finally, Scx directly transactivated Tnmd, a process entailing the recruitment and binding of Scx to the multiple selective E-boxes within the promoter region of Tnmd. Although we have not examined other later markers along the teno-differentiation such as type I collagen genes and aggrecan, Scx may also be poised to modulate the expression of these genes as well29,37,38. Our previous study demonstrated that BMP-12 treatment of rat adult MSCs derived from bone marrow (BM-MSCs) induces these adult stem cells undergoing tenocyte differentiation, as remarked by expression of not only Scx and Tnmd, but also other late teno-phenotypic markers collagen I and tenascin C9. The BMP-12 treated MSCs showed ability in augmenting tendon-like tissue formation and defect repair in vivo. Together, these studies suggest induction of Scx is a key early

SCiENTifiC RePorTS | 5:13149 | DOI: 10.1038/srep13149 6 www.nature.com/scientificreports/ RETRACTED

Figure 5. Scleraxis knockdown preserves multi-potent potentials in C3H10T1/2 cells. Untransfected C3H10T1/2 cells or cells transfected with scramble siRNA, or Scx siRNA, or Tnmd siRNA were left untreated or treated with BMP-12 for 24 hours and further incubated for designed length of time in the absence or presence of lineage induction media. The phenotypes Scx( and Tnmd expression) of respective groups of cells and mRNA expression of lineage-specific marker genes were assayed by RT-qPCR. Data are expressed as mean ± S.D. (n = 6). *p < 0.05.

event in BMP-12-mediated tenocyte differentiation in both embryonic-derived (i.e. C3H10T1/2) and adult (i.e. BM-MSCs) MSCs. In addition to its indispensable role in the control of downstream genes within the tenogenic path- way, Scx might also be directly responsible for the suppression of non-tenocytic differentiation poten- tials of C3H10T1/2 cells. Utilizing a sophisticated approach combining single colony-forming assay and RT-qPCR, we were first able to identify and successfully separate two types of colonies: Scx+ and Scx- colonies, and found that while the osteogenic, adipogenic, and chondrogenic capabilities were largely abrogated in Scx-positive cells, all of those distinct lineage potentials were, surprisingly, well maintained in Scx-negative cells. The inhibitory role of Scx was collaborated by the observation that upon repres- sion of Scx by siRNA, C3H10T1/2 cells were capable of retaining their multi-lineage potentials even in the presence of BMP-12, which was exactly the same as those observed with Scx- cells. In support of our data, reports documented that Scx and Sox9, a master transcription factor for chondrogenesis39,40, are co-expressed in sclerotome during the early stage of embryonic development, but Scx is markedly

SCiENTifiC RePorTS | 5:13149 | DOI: 10.1038/srep13149 7 www.nature.com/scientificreports/ RETRACTED

Figure 6. Compound effects of BMP-12 and cyclic loading on the tenocytic differentiation of C3H10T1/2 stem cells. (A) Induced expression of Scx by BMP-12 and cyclic stretch loading. Cells were left untreated or treated with BMP-12 alone, or cyclic loading alone, or BMP-12 plus cyclic loading. After incubation, cells were harvested and analyzed for Scx mRNA expression by RT-qPCR. Data are expressed as mean ± S.D. (n = 3). *p < 0.05. (B) BMP-12 and stretch loading induced cell morphological changes. Cells were treated as in (A) and stained with toluidine blue. (C) Nuclear aspect ratio: width vs length of nucleus; (D) Nuclear orientation angle: deviation of nuclear axis from average axis. *p < 0.05. C: Control, B: BMP-12 only, S: stretch loading only, B + S: BMP-12+stretch loading.

reduced in the center of Sox9-expresssed sclerotome following development41, underlining a potentially suppressive function for Scx in the chondrogenesis. It should be pointed out, however, that Scx alone appeared to be insufficient to drive MSC differentia- tion into tenocytes. Scx+ colony-derived cells still expressed osteo-, chondro-, and adipogenic potentials at low but detectably greater levels than un-induced controls (Fig. 4D). We observed a similar phenomenon in cells where only Tnmd, but not Scx, was silenced with siRNA (Scx+ Tnmd- cells) (Fig. 5). Most likely, the Scx+ colonies derived from single colony-forming assays may harbor Scx+ Tnmd− and Scx+ Tnmd+ sub-fractions, and the former are responsible for the low level of differentiation along non-tenogenic pathways. Given these considerations, we propose that the phenotypic plasticity of C3H10T1/2 cells becomes progressively limited during tenocytic differentiation, and that these changes occur in stages that are functionally determined by the sequential expression of Scx and Tnmd. In the earlier stage, Scx appears to play a central role both in triggering the expression of tenocyte-associated genes (Tnmd and possibly type I collagen, tenascin-C, type XIV collagen), as well as in suppressing non-tenogenic capacity. At the later (Scx+/Tnmd+) stage, cells become fully committed to the tenocyte phenotype and completely lose their capacity to differentiate into non-tenocyte lineages, at least under the conditions we examined. The critical role of Scx in tenocyte differentiation of C3H10T1/2 cells was not limited to induction by BMP-12, but was also required for mechanical load-triggered tenocytic differentiation. Moreover, combination of the biochemical and mechanical stimuli led to marked augmentation of the differentia- tion response incurred by either stimulus alone as evidenced by increased Scx expression and enhanced formation of tenocyte-like cells. Furthermore, we found that both BMP-12 and stretch loading induced extensive phosphorylation of Smad8 but differed in the time course of this biochemical process. The mediation of both BMP-12 and stretch loading responses by the Smad system provides a molecular basis to explain not only the crosstalk between these stimuli in controlling tenocyte differentiation, but also the downregulation of non-tenogenic potency by Scx since Smad8 has been shown to promote neotendon formation by inhibiting the osteogenic pathway induced by BMP-242. Together, this study suggests that Smad8 is a key signaling molecule whose activation is required to mediate Scx expression and ensuing tenocytic differentiation of C3H10T1/2 cells in response to BMP-12 and stretch loading. In summary, we have identified scleraxis as a central regulator responsible for the fate determination of C3H10T1/2 cells to differentiate into tenocytes and suppress alternate pathways for differentiating into other lineages. Both actions are induced by mechanical as well as chemical stimuli that are mediated intracellularly by phosphorylation of Smad8 and act in an additive fashion. These findings offer novel insights into the molecular cascades underlying tenocyte differentiation from pluripotent stem cells, and may contribute to the development of cell-based tissue engineering approaches useful for tendon repair.

SCiENTifiC RePorTS | 5:13149 | DOI: 10.1038/srep13149 8 www.nature.com/scientificreports/ RETRACTED

Figure 7. Dependency of Smad8 activation in BMP-12 and cyclic stretch loading signals. (A) Enhanced Smad 1/5/8 phosphorylation induced by combined treatment of BMP-12 and cyclic stretch loading. Cells under indicated treatment were harvested at 1 hour and 2 hours post-treatment to analyze the levels of Smad1/5/8 phosphorylation by Western blot. (B–E) Activation of Smad8 in C3H10T1/2 stem cells in response to BMP-12 and stretch loading. Cells were transfected with siRNA against Smad1, Smad5, Smad8, or control siRNA, (siScr) followed by treatment of BMP-12 or loading or both, and protein phosphorylation analyzed by Western blot. C: Control, B: BMP-12 only, S: stretch loading only, B + S: BMP-12+stretch loading, Scr: Scramble of siRNA, siS1: siRNA-Smad1, siS5: siRNA-Smad5, siS8: siRNA-Smad8.

Methods Cell culture and stimulation. Murine pluripotent mesenchymal stem cells C3H10T1/2 (ATCC CCL226)43 were grown as monolayers in DMEM with 10% fetal bovine serum. Twenty-four hours prior to BMP-12 treatment or mechanical loading, cells were starved for 12 hours in DMEM supplanted with 1% FBS. For BMP-12 treatment, cells were treated with indicated concentration of recombinant mouse BMP-12 (R&D Systems) or PBS (control) in starving media for an additional 12 hours. For mechanical loading, cells were subjected to uniaxial cyclic loading for 1 hour with indicated load levels using a cus- tom-designed cell-loading device connected to a computer controlled linear stepper motor (Anaheim Automation). For combined treatment experiments, cells were treated with BMP-12 for 11 hours and then further treated with BMP-12 for one more hour or subjected to uniaxial cyclic loading for 1 hour at indicated load levels. Following treatments, cells were maintained for an additional 3 to 7 days and collected for analysis of gene expression and morphological changes.

SCiENTifiC RePorTS | 5:13149 | DOI: 10.1038/srep13149 9 www.nature.com/scientificreports/ RETRACTED

Single CFU-F replating assay (SCRA). To determine the differentiation potential of Scx+ and Scx− cells derived BMP-12 treated stem cells, SCRA was performed as previously described (20). Briefly, cells were seeded at low density (1 × 103/100 mm culture dish) and cultured for two weeks. Individual colonies were collected with EDTA–trypsin. Cell suspensions from each colony were divided into four portions: one of them was used to assay Scx expression by RT-qPCR, and the remaining three portions of cells were replated in individual wells of a 48-well plate containing either osteogenic, chondrogenic, or adipogenic differentiation medium. The osteogenic induction cocktail consisted of growth media supplemented with 10 nM dexamethasone, 50 μ g/ml ascorbate-phosphate, and 10 mM β -glycerophosphate (Sigma). The adi- pogenic induction cocktail consisted of growth media supplemented with 0.5 uM hydrocortisone, 100 μ g/ ml isobytulmethylxanthine, and 60 μ M indomethacin (Sigma). For chondrogenic induction, cells were grown in a micromass culture. 1 × 106 cells were centrifuged in a polypropylene tube at 150 g for 5 min to form a cell pellet. The pellet was suspended in a chondrogenic induction cocktail consisting of growth media supplemented with 10 ng/ml TGF-β 3, 50 mg/ml insulin-transferrin-selenium (ITS) + Premix (Invitrogen), and 50 ug/mL ascorbate-phosphate. Cells were harvested after two weeks lineage induction and assayed for osteo-, chondro-, or adipo-phenotype markers gene expression using RT-qPCR or in some cases by immunocytochemical staining.

Immunocytochemical staining (ICC). Cells were fixed in 10% formaldehyde, treated with serum-free protein blocking solution (Dako), and then incubated overnight at 4 °C with antibodies against Scx (1:150, Abcam, Rabbit polyclonal), Tnmd (1:150, Santa Cruz Biotechnology, Goat Polyclonal), Runx2 (1:200, Abcam, Rabbit polyclonal), Sox9 (1:200 Santa Cruz Biotechnology, Rabbit polyclonal), PPARγ 2 (1:200, Abcam, Rabbit polyclonal), or aggrecan (1:200, Abcam, Rabbit polyclonal). Cells were then rinsed in PBS. For Scx and Tnmd staining, cells were incubated with anti-rabbit Alexafluor 488 (1:1000, Invitrogen) or anti-goat Alexafluor 568 (1:1000, Invitrogen). For the other antibodies, cells were incu- bated for 30 minutes with anti-rabbit secondary antibody conjugated with horseradish peroxidase (Dako) followed by DAB chromagen (Vector Laboratories) for 2 minutes. Negative control sections were pre- pared using irrelevant isotype matched primary antibodies (Dako).

Cytochemical Staining (CCS). Osteogenic, adipogenic and chondrogenic differentiation were assessed by staining with Alizarian Red S (Sigma) for calcium deposition, Oil Red O (Sigma) and Safranin-O, respectively44. Imaging was performed using a Zeiss inverted microscope, Axiocam camera and Axiovision 4.6 software.

Electrophoretic mobility shift assay (EMSA). Nuclear extracts of C3H10T1/2 cells—either trans- fected to overexpress Scx, treated with 10 ng/ml BMP-12 or left untreated—were incubated with olgio- nucleotide probes containing a specific E-box consensus sequence. Probe sequences were 18 bp long and contained the 6-mer E-box consensus sequence (CANNTG) flanked by 6-mer sequences on each side of the E-box to increase the probe’s specificity. Probes were labeled with [α -32P] dCTP (PerkinElmer Life Sciences) for detection. Specific binding was assessed by competition with an excess of the unlabeled, wild-type (CANNTG) probes and mutated (CANNTG→ AATTTA) probes.

Chromatin immunoprecipitation (ChIP) assay. C3H10T1/2 cells were treated with 10 ng/ml BMP- 12 to induce Scx expression or left untreated (control). The ChIP assay was performed using a com- mercial kit (USB Corporation) following the manufacturer’s instructions. Briefly, cells were treated with 1% formaldehyde to cross-link the scleraxis molecules bound to DNA, and cells were then lysed and sonicated. The samples were immuno-purified using Scx antibodies (Abcam). The subsequent samples were then heated to reverse DNA-protein cross-links and DNA was recovered by phenol/chloroform extraction and resuspended in 10 mM Tris-HCl buffer. Purified DNA was PCR amplified using primer sets designed to include individual E-box sequences with the following priemrs: E-box1 (from start codon): F, 5-CTGAGTTTGCGATTTCTGAGTT-3; R, 5-CTCGATTCTGTGCCTGTAACC-3. E-box2: F, 5-AAGAAAGCAAGCTGGAGGAA-3; R, 5-GAGCATCTGTACCCGGATGT-3. E-box3: F, 5-CGTTT CTTCTTCCTTCCTTCC-3; R, 5-ACCTCACCCTTCCCTGTCTT-3. E-box4: F 5-TGTATATTATGGT TTCCAATG-3; R, 5-GAAGGAAGGAAGAAGAAACG-3. E-box5: F, 5-TGTGTTTCTCCAGGGGAGTC- 3; R, 5-CACACACACACTGTGCCAAA-3.

Total RNA isolation, Reverse transcription and quantitative PCR (RT-qPCR). Total RNA was isolated using the RNeasy Mini Kit (Qiagen). Reverse transcription was carried out using oligodT and Superscript II (Invitrogen). RT-qPCR was performed with a SYBR Green PCR Master Mix (Applied Biosystems) using ABI Prism 7900HT sequence detection system (Applied Biosystems), and normal- ized with GAPDH and β -actin levels. The housekeeping genes have been confirmed to be not regulated by stimuli such as BMP-12. Data were analyzed using the threshold cycle (Ct)-relative quantification method. Primer pairs used in this study are listed in Supplementary Table S1.

Overexpression of Scx and Tnmd. Enforced expression of Scx and Tnmd were performed by trans- fecting C3H10T1/2 cells with vectors containing the coding region of murine Scx or Tnmd, or with no

SCiENTifiC RePorTS | 5:13149 | DOI: 10.1038/srep13149 10 www.nature.com/scientificreports/ RETRACTED

insert using Lipofectamine 2000 reagent (Invitrogen). Cells were then analyzed for Scx and Tnmd mRNA expression by RT-qPCR.

siRNA Transfections. siRNA transfections were carried out using Lipofectamine 2000 reagent (Invitrogen) according to the manufacturer’s instructions. C3H10T1/2 cells were transfected with either Scx-siRNA, Tnmd-siRNA, Smad1-siRNA, Smad5-siRNA, or Smad8-siRNA (Santa Cruz Biotechnologies, Inc., Santa Cruz, CA). Cells were then treated with 10 ng/ml BMP-12 for 12 hours prior to harvest for RT-qPCR or Western blot.

Western Blot. Cell lysates were prepared using RIPA lysis buffer (Pierce Technologies) and protease and phosphatase inhibitor (Thermo Scientific). Proteins were separated in 12% SDS-PAGE gel and trans- ferred onto a nitrocellulose membrane and blocked with milk protein. Membranes were incubated with anti-Scx (1:1000, Rabbit polyclonal, Santa Cruz Biotechnology), anti-Tnmd (1:1000, Rabbit polyclonal, Santa Cruz Biotechnology), anti-phospho-Smad 1/5/8 (1:1000, Rabbit polyclonal, Millipore), anti-Smad8 (1:1000, Rabbit polyclonal, Santa Cruz Biotechnology), or α -Tubulin (1:1000, Rabbit polyclonal, Santa Cruz Biotechnology) followed by incubation with anti-rabbit conjugated with horseradish peroxidase (Santa Cruz Technologies). Membranes were washed with TBST and immunoreactivity was normalized by chemiluminescence using ECL Plus Detection system (Amersham Biosciences).

Statistical analysis. The data were expressed as mean ± S.D. Statistical comparisons between two groups were performed using Student’s t-test or one-way ANOVA with Tukey post-hoc test (for RT-qPCR results). p < 0.05 was considered statistically significant.

References 1. Barzilay, R., Melamed, E. & Offen, D. Introducing transcription factors to multipotent mesenchymal stem cells: making transdifferentiation possible. Stem Cells 27, 2509–15 (2009). 2. Stappenbeck, T. S. & Miyoshi, H. The role of stromal stem cells in tissue regeneration and wound repair. Science 324, 1666–9 (2009). 3. Bianco, P. et al. The meaning, the sense and the significance: translating the science of mesenchymal stem cells into medicine. Nat Med 19, 35–42 (2013). 4. Pittenger, M. F. et al. Multilineage potential of adult human mesenchymal stem cells. Science 284, 143–7 (1999). 5. Chen, X. et al. Stepwise differentiation of human embryonic stem cells promotes tendon regeneration by secreting fetal tendon matrix and differentiation factors. Stem Cells 27, 1276–87 (2009). 6. Wang, Q. W., Chen, Z. L. & Piao, Y. J. Mesenchymal stem cells differentiate into tenocytes by bone morphogenetic protein (BMP) 12 gene transfer. J Biosci Bioeng 100, 418–22 (2005). 7. Awad, H. A. et al. Autologous mesenchymal stem cell-mediated repair of tendon. Tissue Eng 5, 267–77 (1999). 8. Ramdass, B. & Koka, P. S. Ligament and tendon repair through regeneration using mesenchymal stem cells. Curr Stem Cell Res Ther. 10, 84–8 (2014). 9. Lee, J. Y. et al. BMP-12 treatment of adult mesenchymal stem cells in vitro augments tendon-like tissue formation and defect repair in vivo. PLoS One 6, e17531 (2011). 10. Sassoon, A. A. et al. and bone marrow derived stromal cells: a comparison of tenocyte differentiation capabilities. J Orthop Res 30, 1710–8 (2012). 11. Wolfman, N. M. et al. Ectopic induction of tendon and ligament in rats by growth and differentiation factors 5, 6, and 7, members of the TGF-beta gene family. J Clin Invest 100, 321–30 (1997). 12. Yu, Y., Bliss, J. P., Bruce, W. J. & Walsh, W. R. Bone morphogenetic proteins and Smad expression in ovine tendon-bone healing. Arthroscopy 23, 205–10 (2007). 13. Wagner, D. O. et al. BMPs: from bone to body morphogenetic proteins. Sci Signal 3, mr1. 14. Lou, J., Tu, Y., Burns, M., Silva, M. J. & Manske, P. BMP-12 gene transfer augmentation of lacerated tendon repair. J Orthop Res 19, 1199–202 (2001). 15. Majewski, M. et al. Ex vivo adenoviral transfer of bone morphogenetic protein 12 (BMP-12) cDNA improves Achilles tendon healing in a rat model. Gene Ther 15, 1139–46 (2008). 16. Violini, S., Ramelli, P., Pisani, L. F., Gorni, C. & Mariani, P. Horse bone marrow mesenchymal stem cells express embryo stem cell markers and show the ability for tenogenic differentiation by in vitro exposure to BMP-12. BMC Cell Biol 10, 29 (2009). 17. Juncosa-Melvin, N. et al. Effects of mechanical stimulation on the biomechanics and histology of stem cell-collagen sponge constructs for rabbit patellar tendon repair. Tissue Eng 12, 2291–300 (2006). 18. Farng, E., Urdaneta, A. R., Barba, D., Esmende, S. & McAllister, D. R. The effects of GDF-5 and uniaxial strain on mesenchymal stem cells in 3-D culture. Clin Orthop Relat Res 466, 1930–7 (2008). 19. Chen, X. et al. Force and scleraxis synergistically promote the commitment of human ES cells derived MSCs to tenocytes. Sci Rep 2, 977 (2012). 20. Aslan, H., Kimelman-Bleich, N., Pelled, G. & Gazit, D. Molecular targets for tendon neoformation. J Clin Invest 118, 439–44 (2008). 21. Jelinsky, S. A., Archambault, J., Li, L. & Seeherman, H. Tendon-selective genes identified from rat and human musculoskeletal tissues. J Orthop Res 28, 289–97. 22. Onizuka, N. et al. The Mohawk transcription factor regulates the differentiation of tendons and volar plates. J Orthop Sci 19, 172–80 (2014). 23. Liu, H. et al. Crucial transcription factors in tendon development and differentiation: their potential for tendon regeneration. Cell Tissue Res 356, 287–98 (2014). 24. Schweitzer, R. et al. Analysis of the tendon cell fate using Scleraxis, a specific marker for tendons and ligaments. Development 128, 3855–66 (2001). 25. Brent, A. E., Schweitzer, R. & Tabin, C. J. A somitic compartment of tendon progenitors. Cell 113, 235–48 (2003). 26. Pryce, B. A., Brent, A. E., Murchison, N. D., Tabin, C. J. & Schweitzer, R. Generation of transgenic tendon reporters, ScxGFP and ScxAP, using regulatory elements of the scleraxis gene. Dev Dyn 236, 1677–82 (2007).

SCiENTifiC RePorTS | 5:13149 | DOI: 10.1038/srep13149 11 www.nature.com/scientificreports/ RETRACTED

27. Murchison, N. D. et al. Regulation of tendon differentiation by scleraxis distinguishes force-transmitting tendons from muscle- anchoring tendons. Development 134, 2697–708 (2007). 28. Shukunami, C., Takimoto, A., Oro, M. & Hiraki, Y. Scleraxis positively regulates the expression of tenomodulin, a differentiation marker of tenocytes. Dev Biol 298, 234–47 (2006). 29. Liu, Y. et al. Overexpression of a single helix-loop-helix-type transcription factor, scleraxis, enhances aggrecan gene expression in osteoblastic osteosarcoma ROS17/2.8 cells. J Biol Chem 272, 29880–5 (1997). 30. Chen, Y. J. et al. Effects of cyclic mechanical stretching on the mRNA expression of tendon/ligament-related and osteoblast- specific genes in human mesenchymal stem cells. Connect Tissue Res 49, 7–14 (2008). 31. Kuo, C. K. & Tuan, R. S. Mechanoactive tenogenic differentiation of human mesenchymal stem cells. Tissue Eng Part A 14, 1615–27 (2008). 32. Petrigliano, F. A. et al. The effects of local bFGF release and uniaxial strain on cellular adaptation and gene expression in a 3D environment: implications for ligament tissue engineering. Tissue Eng 13, 2721–31 (2007). 33. Virchenko, O. & Aspenberg, P. How can one platelet injection after tendon injury lead to a stronger tendon after 4 weeks? Interplay between early regeneration and mechanical stimulation. Acta Orthop 77, 806–12 (2006). 34. Zhang, J. & Wang, J. H. The effects of mechanical loading on tendons–an in vivo and in vitro model study. PLoS One 8, e71740 (2013). 35. Miyazawa, K., Shinozaki, M., Hara, T., Furuya, T. & Miyazono, K. Two major Smad pathways in TGF-beta superfamily signalling. Genes Cells 7, 1191–204 (2002). 36. Butler, D. L. & Awad, H. A. Perspectives on cell and collagen composites for tendon repair. Clin Orthop Relat Res Oct, S324–32 (1999). 37. Lejard, V. et al. Scleraxis and NFATc regulate the expression of the pro-alpha1(I) collagen gene in tendon fibroblasts. J Biol Chem 282, 17665–75 (2007). 38. Terraz, C., Brideau, G., Ronco, P. & Rossert, J. A combination of cis-acting elements is required to activate the pro-alpha 1(I) collagen promoter in tendon fibroblasts of transgenic mice. J Biol Chem 277, 19019–26 (2002). 39. Mori-Akiyama, Y., Akiyama, H., Rowitch, D. H. & de Crombrugghe, B. Sox9 is required for determination of the chondrogenic cell lineage in the cranial neural crest. Proc Natl Acad Sci USA 100, 9360–5 (2003). 40. Bernard, P. et al. Dimerization of SOX9 is required for chondrogenesis, but not for sex determination. Hum Mol Genet 12, 1755–65 (2003). 41. Asou, Y. et al. Coordinated expression of scleraxis and Sox9 genes during embryonic development of tendons and cartilage. J Orthop Res 20, 827–33 (2002). 42. Hoffmann, A. et al. Neotendon formation induced by manipulation of the Smad8 signalling pathway in mesenchymal stem cells. Journal of Clinical Investigation 116, 940–952 (2006). 43. Reznikoff, C. A., Brankow, D. W. & Heidelberger, C. Establishment and characterization of a cloned line of C3H mouse embryo cells sensitive to postconfluence inhibition of division. Cancer Res 33, 3231–8 (1973). 44. Bi, Y. et al. Identification of tendon stem/progenitor cells and the role of the extracellular matrix in their niche. Nat Med 13, 1219–27 (2007). Acknowledgments This study was supported by NIH grants AR050968, AR047628 (Sun H.B.). Author Contributions Y.L., M.R. and T.A. performed all the experiments. The study was designed by H.S. Z.Z., Y.L., M.R., D.L., T.A., R.M. and H.S. contributed to the scientific interpretation, reviewed, and made contributions to the final manuscript text and figures. Additional Information Supplementary information accompanies this paper at http://www.nature.com/srep Competing financial interests: The authors declare no competing financial interests. How to cite this article: Li, Y. et al. The Role of Scleraxis in Fate Determination of Mesenchymal Stem Cells for Tenocyte Differentiation. Sci. Rep. 5, 13149; doi: 10.1038/srep13149 (2015). This work is licensed under a Creative Commons Attribution 4.0 International License. The images or other third party material in this article are included in the article’s Creative Com- mons license, unless indicated otherwise in the credit line; if the material is not included under the Creative Commons license, users will need to obtain permission from the license holder to reproduce the material. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/

SCiENTifiC RePorTS | 5:13149 | DOI: 10.1038/srep13149 12 www.nature.com/scientificreports

OPEN Retraction Note: The Role of Scleraxis in Fate Determination of Mesenchymal Stem Cells for

Published: xx xx xxxx Tenocyte Differentiation Yonghui Li1, Melissa Ramcharan1,2, Zuping Zhou1,5, Daniel J. Leong1,2,3,4, Takintope Akinbiyi1, Robert J. Majeska2 & Hui B. Sun1,3,4

Retraction of: Scientific Reports https://doi.org/10.1038/srep13149, published online 20 August 2015

The Authors are retracting this Article.

During an Inquiry into an allegation of research misconduct at the Icahn School of Medicine at Mount Sinai, where the research was conducted, it was determined that the “BMP-12 + Stretch” panel depicted in Figure 6B had been falsified. As a result, the graphs based upon Fig. 6B were also rendered inaccurate. Since these data are critical to certain main conclusions of the study, and the authors feel they cannot guarantee the accuracy of those conclusions, a retraction of the paper is warranted.

Melissa Ramcharan, Daniel J. Leong, Takintope Akinbiyi, Robert J. Majeska and Hui B. Sun agree to the retraction. Yonghui Li and Zuping Zhou could not be reached.

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 Cre- ative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not per- mitted 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 license, visit http://creativecommons.org/licenses/by/4.0/.

© The Author(s) 2018

1Leni and Peter W. May Department of Orthopedics, Mount Sinai School of Medicine, New York, NY, 10029, USA. 2Department of Biomedical Engineering, City College of New York, New York, NY, 10031, USA. 3Department of Orthopedic Surgery, Albert Einstein College of Medicine, Bronx, NY, 10461, USA. 4Department of Radiation Oncology, Albert Einstein College of Medicine, Bronx, NY, 10461, USA. 5Present address: Center for Stem cell and Regenerative Medicine Guangxi Normal University, Guilin, 541004, China. Yonghui Li, Melissa Ramcharan and Zuping Zhou contributed equally. Correspondence and requests for materials should be addressed to H.B.S. (email: herb. [email protected])

SCientifiC REPOrTS | (2018) 8:16377 | DOI:10.1038/s41598-018-34658-3 1