Introduction

Glioblastoma (GBM) is the most common primary malignant brain tumor and is one of the most aggressive cancers [1]. Despite the current best practice of maximal safe surgery, chemotherapy, and radiotherapy, the outcome of GBM is still devastating, in part due to intratumoral heterogeneity. Intratumoral heterogeneity implies spatially dispersed genetically diverse populations of cancer cells, distinguishing themselves in a variety of biological functions, including cell proliferation, invasion, immunity, and metastasis. Phylogenic clonal evolution giving rise to differential cellular hierarchies during oncogenesis usually leads to intratumoral heterogeneity in human GBMs. Therefore, intratumoral heterogeneity can contribute to treatment resistance or immune evasion [2].

Epithelial to mesenchymal transition, an EMT process, is a highly controversial term in GBM tumorigenesis. In gliomagenesis and its progression, a non-classical EMT-like or glial to mesenchymal transition is proposed to be linked to the mesenchymal nature of the neural cells during the neurodevelopmental process [3]. Another study reported the onco-plasticity of EMT-like or mesenchymal to epithelial transition (MET)-like conversion mediated by the epigenetic changes in the tumor microenvironment around glioma stem cells [4]. The well-known classification of glioma subtypes by gene expression profiles also uses several EMT-related genes (CD44, MERTK, TGFB1, NOTCH2 etc.), which are enriched in the mesenchymal subtypes [5, 6]. A recent study attempted to classify the GBMs using a series of EMT-associated genes to predict the differences in prognosis [7]. In addition, EMT has been widely described as a key factor that drives cellular heterogeneity influenced by both genetic and non-genetic determinants and even a small number of cells undergoing this process can cause evident intratumoral heterogeneity in cancers [8, 9]. However, the role of EMT and its clinical significance in GBM is still not well established due to the lack of direct evidence of its evolution in clinical samples and the rarity of cell lines that are derived from the same tissue to perform the verifying experiments.

Here, we isolated and established two distinct cell lines derived from the same GBM tissue sample. These cell lines have genetically developed into different cell lines by subculture. We compared and analyzed the genomic and biological signatures of the two cell lines and found that one of the cell lines exhibits mesenchymal transition characteristics, unlike the other. Furthermore, we reanalyzed publicly available single-cell RNA-seq data to identify analogous clusters to our case and found four similar GBM samples with occult clones of mesenchymal transition. We also found recurrent GBM samples showing evolution of mesenchymal transition clones characterized by upregulation of the EMT genes identical to the cell line which was not identified in the initial sample. These findings provide direct evidence to support the subcellular evolution of mesenchymal transition in GBM, which can be utilized in future investigations into intratumoral heterogeneity and the mesenchymal transition process.

Materials and methods

Cell lines and culture condition

An IRB-approved written informed consent was obtained from the patient to use the samples and to establish cell lines for research purposes (Seoul National University Hospital IRB No. H-0507-509-153 and H-1102-098-357). Tissues were freshly frozen in liquid nitrogen immediately after resection, and white blood cells (WBCs) were extracted from whole blood. Both tissue and WBC samples were then stored at − 80 °C until later use. The cell lines were established using the core area of the fresh tumor tissue just after resection. Solid tumor tissue was carefully minced with scissors and isolated into small mixtures by pipetting. Appropriate amounts of delicate tumor tissue fragments were seeded into 25 cm2 cell culture flasks. The tumor cells were initially cultured in Opti-MEM medium supplemented with 5% heat-inactivated fetal bovine serum (FBS) (O5). Cultures were maintained in RPMI 1640 medium containing 10% heat-inactivated FBS (R10). Initial passages were completed when abundant tumor cell growth was observed, and consecutive passages were done every 1 or 2 weeks. During passaging, adherent cells were recovered by pipetting by treatment with trypsin while growth was subconfluent. Differential trypsinization was used to obtain a pure tumor cell population when stromal cell growth was noted in the initial cultures. Cultures were maintained in humidified incubators at 37 °C with 5% CO2 and 95% air. The established cell lines used in the present study are available at the Korea Cell Line Bank (https://cellbank.snu.ac.kr/). The cell lines were maintained in opti-DMEM medium supplemented with 20% fetal bovine serum (GIBCO).

Cell proliferation assay

Cell proliferation assays were performed with EZ-Cytox (Daeillab Service) on cells seeded at 1 × 103 cells/well in 96-well plates and cultured for the designated times. The absorbance of the plates was measured using a microplate reader (Molecular Devices) at a wavelength of 450 nm.

Colony-forming assay

To grow colonies, cells were seeded in 6-well plates at 1000 cells/well density and incubated at 37 °C in an atmosphere of 5% CO2 for 14 days. The cell colonies were fixed and stained with 0.05% crystal violet-methanol-acetic acid solution after 14 days. Stained colonies were scanned and score was calculated.

Western blot analysis

Western blot analysis was performed as previously described [10]. Antibodies against TGF-β (3711), Vimentin (5741), N-Cadherin (13,116), Claudin-1 (13,255), β-Catenin (8480), ZO-1 (8193), Snail (3879), Slug (9585), TCF/ZEB1 (3396), E-cadherin (3195), phosphorylated Smad2 at Ser465/467 (3108), Smad2 (5339), phosphorylated Smad3 at Ser423/425 (9520), Smad3 (9523), Smad4 (38,454) and ACTB (4967) were from Cell Signalling Technology; and HRP-conjugated IgGs (111-035-003 and 115-035-003) were from Jackson Immune Research. Immunoblots were visualized with a ChemiDoc XRS system (Bio-Rad). The density of the bands was measured using free image analyzer software (ImageJ V1.8x; National Institutes of Health, USA, http://rsb.info.nih.gov/ij/) [11].

Telomere repeat amplification protocol (TRAP) assay with ELISA

A TeloTAGGG Telomerase PCR ELISA PLUS kit (Roche) was used according to the manufacturer’s protocol to measure the telomerase activity. GBM tissues or cells were homogenized in ice-cold lysis buffer using an automill (Tokken). Briefly, after BCA protein quantification of the lysates, 10 µg proteins were incubated in a total volume of 50 µl reaction mixture at 25 °C for 30 min to allow the telomerase to add telomeric repeats to the end of the biotin-labeled primer. Then, PCR was conducted for 33 cycles of 94 °C for 30 s, 50 °C for 30 s, and 72 °C for 90 s, followed by an additional extension time of 10 min at 72 °C and then holding the PCR tubes at 4 °C if not used immediately. The TA was measured at a reference wavelength of 450 and 690 nm. The relative TA (RTA) of each sample was calculated according to the instructions of the TeloTAGGG Telomerase PCR ELISA PLUS Kit.

C-circle assay

C-circle detection was executed as previously described [12]. Briefly, 30 ng DNA was combined with 7.5 U Φ29 DNA polymerase (NEB), 0.2 mg/ml BSA, 1 mM each dATP, dGTP and dTTP, 10 μl 2X Φ29 buffer, 0.1% (v/v) Tween 20, and incubated at 30 °C for 4 h or 8 h followed by 20 min at 70 °C. Amplification products were transferred on a Hybond N + nylon membrane (Bio-Rad) and processed using the TeloTAGGG Telomere Length Assay Kit (Roche). Chemiluminescent signals were visualized with a ChemiDoc XRS system (Bio-Rad), and the intensity of the spots was quantified with ImageQuant TL software (Bio-Rad).

In vivo xenograft tumor growth assay

To observe the in vivo proliferation of patient-derived primary GBM cells, a subcutaneous xenograft mouse model was constructed using 5-week-old female Balb/c nude mice. Primary GBM cells were cultured in Opti-MEM (LS31985070, Gibco) supplemented with 5% FBS (S 001-01, Welgene) and 1 × antibiotic–antimycotic (15240-062, Thermo Fisher Scientific) at 37 °C in a humidified incubator in the presence of 5% CO2. In total, 3 × 106 cells were mixed with 100 µl of Opti-MEM and Matrigel mixture (Opti-MEM:Matrigel = 1:1) and subcutaneously injected into the mice. Tumor size was measured once a week using calipers, and tumor volume was calculated using the following formula: (length × width2)/2.

Telomere length fragmentation assay

Telomere lengths were determined by Southern blot using a TeloTAGGG Telomere Length Assay Kit (Roche) according to the manufacturer’s protocol. Briefly, 1 µg of each DNA sample was digested with RsaI and HinfI for overnight at 37 °C, electrophoresed on a 0.8% agarose gel at 50 V for 4 h, and transferred to a nylon membrane by Southern blotting. The membrane was blocked and hybridized overnight to a digoxigenin (DIG)-labeled probe specific for telomeric repeats. Then, it was incubated with anti-DIG-alkaline phosphatase (1:1000 dilution) for 30 min and processed using the substrate in the TeloTAGGG Telomere Length Assay kit (Roche). After chemiluminescence signals were visualized with a ChemiDoc XRS system (Bio-Rad), telomere length was calculated with Telo Tool version 1.3.

Whole exome sequencing and data analysis

Extracted DNAs from the samples were checked for quality control with Bioanalyzer. The sequencing library was prepared by random fragmentation of the DNA or cDNA followed by 5′ and 3′ adapter ligation. Library preparation was done using the SureSelectXT library prep kit. Whole exome sequencing (WES) was performed with the Illumina platform at Macrogen, Korea, and generated BCL binary was converted into raw FASTQ files utilizing the Illumina bcl2fastq package. Next FASTQ files were mapped to the reference genome (hg19) with BWA (0.7.12), and PCR duplicates were marked with Picard (1.130) [13]. Base recalibration and SNP and INDEL calling was done using GATK (v3.4.0) [14] We processed five different samples (blood, tumor, parent cell, subclone #5, and subclone #11) to obtain the WES data from the same patient. To track genomic alterations across the samples, the generated bam and vcf files were then analyzed with SuperFreq [15]. Typically, the SuperFreq program is designed for clonal analysis with samples related by timepoints. It tracks the SNP/INDEL and CNV from the samples and calls out clones or clusters according to the mutation changes. Since our samples do not contain any temporal information, we took advantage of the SuperFreq to track mutational differences and sample-specific mutations for both SNP/INDEL and CNV of our five samples. Mutation variations were clustered according to their variant allele frequency throughout the samples.

RNA-sequencing and data analysis

Extracted RNAs from the samples were checked for quality control with Bioanalyzer. Library preparation was conducted using a Lexogen Quantseq 3’ mRNAseq library prep kit. Single-end mRNA sequencing was performed with Illumina Nextseq500 at ebiogen, Korea. Generated FASTQ files were mapped to the reference genome (hg19) with Bowtie2. Normalization of the count was done with EdgeR and batch effects were removed with the help of the R package limma (doi: https://doi.org/10.1093/nar/gks042 and https://doi.org/10.1093/nar/gkv007) [16, 17]. We then compared the general RNA-seq expression profile of the genes across the samples with Pearson correlation analysis using R. Differential gene expression (DEG) analysis between subclone #11 and #5 were done using the ExDEGA software provided by ebiogen Korea. Significant upregulated genes in subclone #11 compared to subclone #5 was selected with the filtering criteria of P value < 0.05 and Log2 fold change 11/5 ≥ 2. Then functional clustering of the upregulated genes in subclone #11 was done using DAVID v6.8 [18]. For the RNA-seq of GBM longitudinal samples, a total of 55 pairs (initial and recurrent) of tumor samples RNA was extracted and the library was prepared with a SureSelectV6 mRNA seq library preparation kit. RNA-sequencing was performed with Illumina Hiseq2500 at Macrogen, Korea. FASTQ files were processed and normalization of the count was done as described above. Then the differences between gene expression were calculated by subtracting the normalized count of each initial sample from its paired recurrent sample.

Single cell RNA-seq analysis

We downloaded GBM single-cell RNA-seq data from four recently published studies (GSE117891, GSE131928, GSE125587, phs001287), which included a total of 71 samples from 65 patients [19,20,

Availability of data and materials

Data and materials for the study are included in the manuscript and supplementary information.

Abbreviations

GBM:

Glioblastomas

EMT:

Epithelial–mesenchymal transition

MET:

Mesenchymal to epithelial transition

EGFR:

Epidermal growth factor receptor

ALT:

Alternative lengthening of telomeres

TRF:

Telomere restriction fragment

CNV:

Copy number variation

WES:

Whole-exome sequencing

DEGs:

Differentially expressed genes

References

  1. Parker NR, Khong P, Parkinson JF, Howell VM, Wheeler HR (2015) Molecular heterogeneity in glioblastoma: potential clinical implications. Front Oncol 5:55. https://doi.org/10.3389/fonc.2015.00055

    Article  PubMed  PubMed Central  Google Scholar 

  2. Burrell RA, Swanton C (2014) Tumour heterogeneity and the evolution of polyclonal drug resistance. Mol Oncol 8:1095–1111. https://doi.org/10.1016/j.molonc.2014.06.005

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Mahabir R, Tanino M, Elmansuri A, Wang L, Kimura T, Itoh T et al (2014) Sustained elevation of Snail promotes glial-mesenchymal transition after irradiation in malignant glioma. Neurooncology 16:671–685. https://doi.org/10.1093/neuonc/not239

    Article  CAS  Google Scholar 

  4. Baysan M, Woolard K, Bozdag S, Riddick G, Kotliarova S, Cam MC et al (2014) Micro-environment causes reversible changes in DNA methylation and mRNA expression profiles in patient-derived glioma stem cells. PLoS One 9:e94045. https://doi.org/10.1371/journal.pone.0094045

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Verhaak RG, Hoadley KA, Purdom E, Wang V, Qi Y, Wilkerson MD et al (2010) Integrated genomic analysis identifies clinically relevant subtypes of glioblastoma characterized by abnormalities in PDGFRA, IDH1, EGFR, and NF1. Cancer Cell 17:98–110. https://doi.org/10.1016/j.ccr.2009.12.020

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Phillips HS, Kharbanda S, Chen R, Forrest WF, Soriano RH, Wu TD et al (2006) Molecular subclasses of high-grade glioma predict prognosis, delineate a pattern of disease progression, and resemble stages in neurogenesis. Cancer Cell 9:157–173. https://doi.org/10.1016/j.ccr.2006.02.019

    Article  CAS  PubMed  Google Scholar 

  7. Tao C, Huang K, Shi J, Hu Q, Li K, Zhu X (2020) Genomics and prognosis analysis of epithelial–mesenchymal transition in glioma. Front Oncol 10:183. https://doi.org/10.3389/fonc.2020.00183

    Article  PubMed  PubMed Central  Google Scholar 

  8. Neelakantan D, Drasin DJ, Ford HL (2015) Intratumoral heterogeneity: clonal cooperation in epithelial-to-mesenchymal transition and metastasis. Cell Adhes Migr 9:265–276. https://doi.org/10.4161/19336918.2014.972761

    Article  CAS  Google Scholar 

  9. Hinohara K, Polyak K (2019) Intratumoral heterogeneity: more than just mutations. Trends Cell Biol 29:569–579. https://doi.org/10.1016/j.tcb.2019.03.003

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Chung H, Jung YM, Shin DH, Lee JY, Oh MY, Kim HJ et al (2008) Anticancer effects of wogonin in both estrogen receptor-positive and -negative human breast cancer cell lines in vitro and in nude mice xenografts. Int J Cancer 122:816–822. https://doi.org/10.1002/ijc.23182

    Article  CAS  PubMed  Google Scholar 

  11. Schneider CA, Rasband WS, Eliceiri KW (2012) NIH Image to ImageJ: 25 years of image analysis. Nat Methods 9:671–675. https://doi.org/10.1038/nmeth.2089

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Henson JD, Lau LM, Koch S, Martin La Rotta N, Dagg RA, Reddel RR (2017) The C-circle assay for alternative-lengthening-of-telomeres activity. Methods 114:74–84. https://doi.org/10.1016/j.ymeth.2016.08.016

    Article  CAS  PubMed  Google Scholar 

  13. Li H, Durbin R (2010) Fast and accurate long-read alignment with Burrows–Wheeler transform. Bioinformatics 26:589–595. https://doi.org/10.1093/bioinformatics/btp698

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Van der Auwera GA, Carneiro MO, Hartl C, Poplin R, Del Angel G, Levy-Moonshine A et al (2013) From FastQ data to high confidence variant calls: the Genome Analysis Toolkit best practices pipeline. Curr Protoc Bioinform 43:11101–111033. https://doi.org/10.1002/0471250953.bi1110s43

    Article  Google Scholar 

  15. Flensburg C, Sargeant T, Oshlack A, Majewski IJ (2020) SuperFreq: integrated mutation detection and clonal tracking in cancer. PLoS Comput Biol 16:e1007603. https://doi.org/10.1371/journal.pcbi.1007603

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Robinson MD, McCarthy DJ, Smyth GK (2010) edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics 26:139–140. https://doi.org/10.1093/bioinformatics/btp616

    Article  CAS  PubMed  Google Scholar 

  17. Ritchie ME, Phipson B, Wu D, Hu Y, Law CW, Shi W et al (2015) limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res 43:e47. https://doi.org/10.1093/nar/gkv007

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Huang DW, Sherman BT, Lempicki RA (2009) Systematic and integrative analysis of large gene lists using DAVID bioinformatics resources. Nat Protoc 4:44–57

    Article  CAS  Google Scholar 

  19. Yu K, Hu Y, Wu F, Guo Q, Qian Z, Hu W et al (2020) Surveying brain tumor heterogeneity by single-cell RNA-sequencing of multi-sector biopsies. Natl Sci Rev 7:1306–1318. https://doi.org/10.1093/nsr/nwaa099

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Neftel C, Laffy J, Filbin MG, Hara T, Shore ME, Rahme GJ et al (2019) An integrative model of cellular states, plasticity, and genetics for glioblastoma. Cell 178(835–849):e21. https://doi.org/10.1016/j.cell.2019.06.024

    Article  CAS  Google Scholar 

  21. **ao Y, Kim D, Dura B, Zhang K, Yan R, Li H et al (2019) Ex vivo dynamics of human glioblastoma cells in a microvasculature-on-a-chip system correlates with tumor heterogeneity and subtypes. Adv Sci (Weinh) 6:1801531. https://doi.org/10.1002/advs.201801531

    Article  CAS  Google Scholar 

  22. Wang LB, Karpova A, Gritsenko MA, Kyle JE, Cao S, Li Y et al (2021) Proteogenomic and metabolomic characterization of human glioblastoma. Cancer Cell 39(509–528):e20. https://doi.org/10.1016/j.ccell.2021.01.006

    Article  CAS  Google Scholar 

  23. Stuart T, Butler A, Hoffman P, Hafemeister C, Papalexi E, Mauck WM 3rd et al (2019) Comprehensive integration of single-cell data. Cell 177(1888–1902):e21. https://doi.org/10.1016/j.cell.2019.05.031

    Article  CAS  Google Scholar 

  24. Tickle T, Tirosh I, Georgescu C, Brown M and Haas B. (2019). inferCNV of the Trinity CTAT Project

  25. Ku J-L, Park J-G (2005) Biology of SNU cell lines. Cancer Res Treat 37:1

    Article  PubMed  PubMed Central  Google Scholar 

  26. Behnan J, Finocchiaro G, Hanna G (2019) The landscape of the mesenchymal signature in brain tumours. Brain 142:847–866. https://doi.org/10.1093/brain/awz044

    Article  PubMed  PubMed Central  Google Scholar 

  27. Iwadate Y (2016) Epithelial–mesenchymal transition in glioblastoma progression. Oncol Lett 11:1615–1620. https://doi.org/10.3892/ol.2016.4113

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Polyak K, Weinberg RA (2009) Transitions between epithelial and mesenchymal states: acquisition of malignant and stem cell traits. Nat Rev Cancer 9:265–273. https://doi.org/10.1038/nrc2620

    Article  CAS  PubMed  Google Scholar 

  29. Mani SA, Guo W, Liao MJ, Eaton EN, Ayyanan A, Zhou AY et al (2008) The epithelial–mesenchymal transition generates cells with properties of stem cells. Cell 133:704–715. https://doi.org/10.1016/j.cell.2008.03.027

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Brennan CW, Verhaak RG, McKenna A, Campos B, Noushmehr H, Salama SR et al (2013) The somatic genomic landscape of glioblastoma. Cell 155:462–477. https://doi.org/10.1016/j.cell.2013.09.034

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Lafferty-Whyte K, Cairney CJ, Will MB, Serakinci N, Daidone MG, Zaffaroni N et al (2009) A gene expression signature classifying telomerase and ALT immortalization reveals an hTERT regulatory network and suggests a mesenchymal stem cell origin for ALT. Oncogene 28:3765–3774. https://doi.org/10.1038/onc.2009.238

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Henson JD, Reddel RR (2010) Assaying and investigating alternative lengthening of telomeres activity in human cells and cancers. FEBS Lett 584:3800–3811. https://doi.org/10.1016/j.febslet.2010.06.009

    Article  CAS  PubMed  Google Scholar 

  33. Mayba O, Gnad F, Peyton M, Zhang F, Walter K, Du P et al (2014) Integrative analysis of two cell lines derived from a non-small-lung cancer patient--a panomics approach. Pac Symp Biocomput. https://doi.org/10.1142/9789814583220_0008

    Article  PubMed  PubMed Central  Google Scholar 

  34. Band V, Zajchowski D, Swisshelm K, Trask D, Kulesa V, Cohen C et al (1990) Tumor progression in four mammary epithelial cell lines derived from the same patient. Cancer Res Treat 50:7351–7357

    CAS  Google Scholar 

  35. Navone NM, Olive M, Ozen M, Davis R, Troncoso P, Tu SM et al (1997) Establishment of two human prostate cancer cell lines derived from a single bone metastasis. Clin Cancer Res 3:2493–2500

    CAS  PubMed  Google Scholar 

  36. Colella B, Faienza F, Di Bartolomeo S (2019) EMT regulation by autophagy: a new perspective in glioblastoma biology. Cancers (Basel). https://doi.org/10.3390/cancers11030312

    Article  Google Scholar 

  37. Kubelt C, Hattermann K, Sebens S, Mehdorn HM, Held-Feindt J (2015) Epithelial-to-mesenchymal transition in paired human primary and recurrent glioblastomas. Int J Oncol 46:2515–2525. https://doi.org/10.3892/ijo.2015.2944

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

This study was supported by the Chung Yang, Cha Young Sun, & Jang Hi Joo Memorial Fund, the Bio & Medical Technology Development Program of the National Research Foundation (NRF) of Korea (NRF-2018M3A9H3021707, NRF-2020M3A9D5A01082439, and NRF-2019R1A2C2087198). And this research was partially supported by the Korea Health Technology R&D Project through the Korea Health Industry Development Institute (KHIDI), funded by the Ministry of Health & Welfare, Republic of Korea (HI21C0239).

Funding

This study was supported by the Chung Yang, Cha Young Sun, & Jang Hi Joo Memorial Fund, the Bio & Medical Technology Development Program of the National Research Foundation (NRF) of Korea (NRF-2018M3A9H3021707, NRF-2020M3A9D5A01082439, and NRF-2019R1A2C2087198). And this research was partially supported by the Korea Health Technology R&D Project through the Korea Health Industry Development Institute (KHIDI), funded by the Ministry of Health & Welfare, Republic of Korea (HI21C0239).

Author information

Authors and Affiliations

Authors

Contributions

SK, S-JP, TC, HL, KK, and CKP designed the study. SK, S-JP, TC, J-IH, JA, TYJ, HJY, Y-KS, J-LK, JBP, and JKH conducted the experiments. HL, KK, and CKP supervised the research. All authors discussed the results and the final manuscript.

Corresponding authors

Correspondence to Hwa** Lee, Kyoungmi Kim or Chul‑Kee Park.

Ethics declarations

Conflict of interest

The authors declare no competing interests.

Ethical approval

All experiments comply with the current South Korea laws.

Additional information

Publisher's Note

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

Supplementary Information

Below is the link to the electronic supplementary material.

Supplementary file1 (DOCX 5639 kb)

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/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kim, S., Park, SJ., Chowdhury, T. et al. Subcellular progression of mesenchymal transition identified by two discrete synchronous cell lines derived from the same glioblastoma. Cell. Mol. Life Sci. 79, 181 (2022). https://doi.org/10.1007/s00018-022-04188-3

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00018-022-04188-3

Keywords