Please wait a minute...
Frontiers of Medicine

ISSN 2095-0217

ISSN 2095-0225(Online)

CN 11-5983/R

Postal Subscription Code 80-967

2018 Impact Factor: 1.847

Front. Med.    2023, Vol. 17 Issue (6) : 1204-1218    https://doi.org/10.1007/s11684-023-1007-9
GID complex regulates the differentiation of neural stem cells by destabilizing TET2
Meiling Xia1,2, Rui Yan2, Wenjuan Wang2, Meng Zhang2, Zhigang Miao2, Bo Wan2(), Xingshun Xu1,2,3()
1. Department of Neurology, The First Affiliated Hospital of Soochow University, Suzhou 215006, China
2. Institute of Neuroscience, Soochow University, Suzhou 215006, China
3. Jiangsu Key Laboratory of Neuropsychiatric Diseases, Soochow University, Suzhou 215123, China
 Download: PDF(8440 KB)   HTML
 Export: BibTeX | EndNote | Reference Manager | ProCite | RefWorks
Abstract

Brain development requires a delicate balance between self-renewal and differentiation in neural stem cells (NSC), which rely on the precise regulation of gene expression. Ten-eleven translocation 2 (TET2) modulates gene expression by the hydroxymethylation of 5-methylcytosine in DNA as an important epigenetic factor and participates in the neuronal differentiation. Yet, the regulation of TET2 in the process of neuronal differentiation remains unknown. Here, the protein level of TET2 was reduced by the ubiquitin-proteasome pathway during NSC differentiation, in contrast to mRNA level. We identified that TET2 physically interacts with the core subunits of the glucose-induced degradation-deficient (GID) ubiquitin ligase complex, an evolutionarily conserved ubiquitin ligase complex and is ubiquitinated by itself. The protein levels of GID complex subunits increased reciprocally with TET2 level upon NSC differentiation. The silencing of the core subunits of the GID complex, including WDR26 and ARMC8, attenuated the ubiquitination and degradation of TET2, increased the global 5-hydroxymethylcytosine levels, and promoted the differentiation of the NSC. TET2 level increased in the brain of the Wdr26+/− mice. Our results illustrated that the GID complex negatively regulates TET2 protein stability, further modulates NSC differentiation, and represents a novel regulatory mechanism involved in brain development.

Keywords TET2      GID complex      neural stem cells      differentiation of neurons     
Corresponding Author(s): Bo Wan,Xingshun Xu   
Just Accepted Date: 28 July 2023   Online First Date: 13 September 2023    Issue Date: 06 February 2024
 Cite this article:   
Meiling Xia,Rui Yan,Wenjuan Wang, et al. GID complex regulates the differentiation of neural stem cells by destabilizing TET2[J]. Front. Med., 2023, 17(6): 1204-1218.
 URL:  
https://academic.hep.com.cn/fmd/EN/10.1007/s11684-023-1007-9
https://academic.hep.com.cn/fmd/EN/Y2023/V17/I6/1204
Fig.1  Ablation of Tet2 reduced the neural differentiation of NSC. (A,B) Representative immunofluorescence (IF) images (A) and relative quantification (B) of TUJ1+ immature neurons (green) in the DG of 2-month-old wild-type (WT) and Tet2 cKO mice. Scale bar, 50?μm; n = 4. (C,D) Representative IF images (C) and relative quantification (D) of SOX2+ NSC (green) in the DG of 2-month-old WT and Tet2 cKO mice. Scale bar, 50?μm; n = 4. (E,F) Representative IF images (E) and relative quantification (F) of PCNA+ proliferating NSC (green) in the DG of 2-month-old WT and Tet2 cKO mice. Scale bar, 50?μm; n = 4. (G,H) The neurospheres were extracted from E14.5-day-old WT and Tet2 cKO mice. After 5 days of incubation, neurosphere-derived single cells were cultured for 5 days to detect the number and diameters of neurospheres. After neurosphere-derived single cells were plated, the growth medium containing forskolin (1 μmol/L) and retina acid (1 μmol/L) was used to induce the differentiation of the NSC for 48 h for IF staining. Representative IF images (G) and relative quantification (H) of SOX2+KI67+ proliferating neurospheres (yellow). (I) Representative bright-field images of the second passage hippocampal neurospheres culture in vitro on day 5 and IF images of TUJ1+ immature neurons (green), and GFAP+ astrocytes (red) of WT and Tet2 cKO mice. Scale bar of bright-field images, 25?μm; scale bar of IF images, 50?μm; n = 3. (J,K) Quantification of the diameter (J) and relative number (K) of cultured neurospheres. (L) Quantification of relative TUJ1+ neurons after NSC differentiation was induced. Quantified data were normalized to the control group. The value was equal to 1. All data were presented as mean ± SEM. Unpaired two-tailed Student’s t-test was used. *P < 0.05; **P < 0.01.
Fig.2  Proteasome pathway is the main pathway that affects the stability of TET2. (A) The quantitative real-time PCR of Tet2 level before and after NSC differentiation; n = 5. (B,C) Representative Western blots (B) and quantitative analysis of TET2 (C) level before and after NSC differentiation; n = 4. (D,E) Representative Western blots (D) and quantitative analysis of FLAG-TET2 (E) in HEK293t cells. HEK293t cells were transfected with FLAG-TET2 for 24 h and then treated with cycloheximide (CHX; 50 µg/mL) alone or in combination with MG132 (10 µmol/L), Z-VAD-FMK (10 µmol/L), or calpeptin (20 µmol/L) for 12 h. Cell lysates were subjected to Western blotting with the FLAG antibody; n = 4. (F–I) Representative Western blots (F,H) and quantitative analysis of TET2 (G,I) in N2a (F,G) or HT22 cells (H,I). N2a or HT22 cells were treated with CHX (15 µg/mL for HT22 cells and 25 µg/mL for N2a cells) alone or in combination with MG132 (10 µmol/L), or Z-VAD-FMK (10 µmol/L), or calpeptin (20 µmol/L) for 12 h. Cell lysates were subjected to Western blotting with the TET2 antibody; n = 4. Quantified data were normalized to the control group. The value was equal to 1. All data were presented as mean ± SEM. Unpaired two-tailed Student’s t-test (A, C) and one-way ANOVA (E, G, I) were used. *P < 0.05; **P < 0.01; ***P < 0.001; ****P < 0.0001.
Fig.3  TET2 interacts with GID complex subunits WDR26 and ARMC8. (A,B) Using the Metascape website for functional analysis based on the mass spectra of TET2 interacting proteins. (C,D) The list of GID complex subunits that could potentially interact with TET2 were sieved out by cluster analysis. Mouse FLAG-TET2 and HA-WDR26/HA-ARMC8 expression plasmids were co-transfected into HEK293t cells, and using anti-FLAG magnetic beads for extraneous immunoprecipitation. Co-immunoprecipitation (Co-IP) assays was used to validate interaction of HA-WDR26 (E) or HA-ARMC8 (F) with FLAG-TET2 in HEK293t cells. (G) Endogenous Co-IP experiments of TET2 with WDR26 and ARMC8 were using N2a cells, and immunoblotting with TET2, WDR26 and ARMC8 antibody.
Fig.4  TET2 and GID complex was highly expressed in neurons. (A–C) Representative images of SOX2+ (red) and NESTIN+ (green) NSC (A), MAP2+ (green) neurons (B) and GFAP+ (red) astrocytes (C). NSC and neurons were extracted from E14.5-day-old wild-type mice and astrocytes from 2-day-old mice. (D) Relative mRNA levels of TET family members and GID complex members in astrocytes, NSC, and neurons were examined; n = 5–6. Quantified data were normalized to the control group. The value was equal to 1. All data were presented as mean ± SEM. One-way ANOVA was used. *P < 0.05; **P < 0.01; ***P < 0.001; ****P < 0.0001.
Fig.5  TET2-GID complex is involved in neuronal differentiation. (A) RT-PCR of TET family and GID complex members’ relative mRNA levels before and after NSC differentiation; n = 4. (B–E) Representative Western blots (B) and quantitative analysis of TET2 (C), WDR26 (D), and ARMC8 (E) in NSC before and after differentiation; n = 4. (F) RT-PCR of TET family and GID complex members’ relative mRNA levels before and after N2a cell differentiation; n = 4. (G–J) Representative Western blots (G) and quantitative analysis of TET2 (H), WDR26 (I), and ARMC8 (J) in N2a cells before and after differentiation; n = 4. (K) Representative Western blots of TET2, WDR26, and ARMC8 during N2a differentiation. NSC differentiation was induced by forskolin (1 μmol/L) and RA (1 μmol/L) for 48 h, and N2a cell differentiation was induced by RA (20 μmol/L) for 12 h. (L) Representative Western blots of TET2, WDR26, and ARMC8 in the hippocampi of mice from days 1, 7, 14, 21, and 30. Quantified data were normalized to the control group, and the value was equal to 1. All data were presented as mean ± SEM. Unpaired two-tailed Student’s t-test was used. *P < 0.05; **P < 0.01; ****P < 0.0001.
Fig.6  TET2 is the direct substrate of the GID complex. (A–C) Western blots confirmed the knockout of Armc8 (A), Wdr26 (B), or Tet2 (C) by sgRNAs in the N2a cells. (D,E) A rescue experiment was used to detect the effect of Armc8 KO on the degradation of TET2 with overexpression FLAG-ARMC8 plasmids in the Armc8 KO N2a cells. Wild-type (WT) N2a cells were transfected with vector plasmids (the first lane), and Armc8 KO N2a cells were transfected with vector or FLAG-ARMC8 plasmids (the second and third lanes). Representative Western blots (D) and quantitative analysis of TET2 and FLAG-ARMC8 (E); n = 3. (F,G) WT N2a cells were transfected with vector plasmids (the first lane), and the Wdr26 KO N2a cells were transfected with vector or HA-WDR26 plasmids (the second and third lanes). Representative Western blots (F) and quantitative analysis of TET2 and HA-WDR26 (G) with overexpression HA-WDR26 plasmids in the Wdr26 KO N2a cells; n = 3. (H) Representative Western blots of TET2 ubiquitination of control or ARMC8-overexpressed N2a cells. (I,J) Representative Western blots (I) and quantitative analysis (J) of control or Armc8 KO N2a cells, which were treated with 25 µg/mL cycloheximide (CHX) for different periods. (K) Representative Western blots of TET2 ubiquitination of control or WDR26-overexpressed N2a cells. (L,M) Representative Western blots (L) and quantitative analysis (M) of control or Wdr26-knockout N2a cells treated with 25 µg/mL CHX for different periods. (I, J, L and M) WT or Armc8 (I,J) or Wdr26 (L,M) knockout N2a cells were treated with 25 µg/mL CHX for different periods. The levels of the indicated proteins were examined by immunoblotting. Quantified data were normalized to the control group, and the value was equal to 1. All data were presented as mean ± SEM. One-way ANOVA was used. *P < 0.05; **P < 0.01.
Fig.7  TET2 protein increased in Wdr26 heterozygous knockout mice. (A–D) Representative Western blots (A) and quantitative analysis of TET1 (B), TET2 (C), and TET3 (D) in wild-type and Wdr26 heterozygous knockout mice; n = 3. Quantified data were normalized to the control group, and the value was equal to 1. All data were presented as mean ± SEM. Unpaired two-tailed Student’s t-test was used. *P < 0.05.
Fig.8  GID complex subunit deletion promotes the DNA 5-hydroxymethylation. (A,B and E) Representative 5hmC (A) and 5mC (B) dot blots and quantitative analysis (E) of wild-type (WT) and Armc8 KO N2a cells; n = 3. (C,D and F) Representative 5hmC (C) and 5mC (D) dot blots and quantitative analysis (F) of WT and Wdr26 KO N2a cells; n = 3. Quantified data were normalized to the control group, and the value was equal to 1. All data were presented as mean ± SEM. Two-way ANOVA was used. *P < 0.05.
Fig.9  GID complex subunit deletion promoted the process of neuronal differentiation. (A) Representative images of wild-type (WT), Armc8 KO, and Wdr26 KO N2a cells before and after differentiation. (B,C) Percentage of differentiated cells (B) and axon length (C) of N2a cells of WT and Wdr26 KO N2a cells; n = 3. (D, F) Western blots confirmed the knockdown of Armc8 (D) and Wdr26 (F) by sgRNAs in NSC. (E,G) RT-PCR of Tuj1 mRNA level of Armc8 (E) and Wdr26 (G) knockdown NSC after induced differentiation; n = 4. Quantified data were normalized to the control group, and the value was equal to 1. All data were presented as mean ± SEM. Unpaired two-tailed Student’s t-test (E,G) and two-way ANOVA (B,C) were used. *P < 0.05; **P < 0.01.
1 S Ito, L Shen, Q Dai, SC Wu, LB Collins, JA Swenberg, C He, Y Zhang. Tet proteins can convert 5-methylcytosine to 5-formylcytosine and 5-carboxylcytosine. Science 2011; 333(6047): 1300–1303
https://doi.org/10.1126/science.1210597
2 YF He, BZ Li, Z Li, P Liu, Y Wang, Q Tang, J Ding, Y Jia, Z Chen, L Li, Y Sun, X Li, Q Dai, CX Song, K Zhang, C He, GL Xu. Tet-mediated formation of 5-carboxylcytosine and its excision by TDG in mammalian DNA. Science 2011; 333(6047): 1303–1307
https://doi.org/10.1126/science.1210944
3 X Wu, Y Zhang. TET-mediated active DNA demethylation: mechanism, function and beyond. Nat Rev Genet 2017; 18(9): 517–534
https://doi.org/10.1038/nrg.2017.33
4 KP Koh, A Yabuuchi, S Rao, Y Huang, K Cunniff, J Nardone, A Laiho, M Tahiliani, CA Sommer, G Mostoslavsky, R Lahesmaa, SH Orkin, SJ Rodig, GQ Daley, A Rao. Tet1 and Tet2 regulate 5-hydroxymethylcytosine production and cell lineage specification in mouse embryonic stem cells. Cell Stem Cell 2011; 8(2): 200–213
https://doi.org/10.1016/j.stem.2011.01.008
5 MM Dawlaty, K Ganz, BE Powell, YC Hu, S Markoulaki, AW Cheng, Q Gao, J Kim, SW Choi, DC Page, R Jaenisch. Tet1 is dispensable for maintaining pluripotency and its loss is compatible with embryonic and postnatal development. Cell Stem Cell 2011; 9(2): 166–175
https://doi.org/10.1016/j.stem.2011.07.010
6 T Li, D Yang, J Li, Y Tang, J Yang, W Le. Critical role of Tet3 in neural progenitor cell maintenance and terminal differentiation. Mol Neurobiol 2015; 51(1): 142–154
https://doi.org/10.1007/s12035-014-8734-5
7 X Li, B Yao, L Chen, Y Kang, Y Li, Y Cheng, L Li, L Lin, Z Wang, M Wang, F Pan, Q Dai, W Zhang, H Wu, Q Shu, Z Qin, C He, M Xu, P Jin. Ten-eleven translocation 2 interacts with forkhead box O3 and regulates adult neurogenesis. Nat Commun 2017; 8: 15903
https://doi.org/10.1038/ncomms15903
8 Q Zhang, Q Hu, J Wang, Z Miao, Z Li, Y Zhao, B Wan, EG Allen, M Sun, P Jin, X Xu. Stress modulates Ahi1-dependent nuclear localization of ten-eleven translocation protein 2. Hum Mol Genet 2021; 30(22): 2149–2160
https://doi.org/10.1093/hmg/ddab179
9 L Li, M Miao, J Chen, Z Liu, W Li, Y Qiu, S Xu, Q Wang. Role of Ten eleven translocation-2 (Tet2) in modulating neuronal morphology and cognition in a mouse model of Alzheimer’s disease. J Neurochem 2021; 157(4): 993–1012
https://doi.org/10.1111/jnc.15234
10 Y Mi, X Gao, J Dai, Y Ma, L Xu, W Jin. A Novel function of TET2 in CNS: sustaining neuronal survival. Int J Mol Sci 2015; 16(9): 21846–21857
https://doi.org/10.3390/ijms160921846
11 Y Wang, Y Zhang. Regulation of TET protein stability by calpains. Cell Rep 2014; 6(2): 278–284
https://doi.org/10.1016/j.celrep.2013.12.031
12 J Cheng, S Guo, S Chen, SJ Mastriano, C Liu, AC D’Alessio, E Hysolli, Y Guo, H Yao, CM Megyola, D Li, J Liu, W Pan, CA Roden, XL Zhou, K Heydari, J Chen, IH Park, Y Ding, Y Zhang, J Lu. An extensive network of TET2-targeting microRNAs regulates malignant hematopoiesis. Cell Rep 2013; 5(2): 471–481
https://doi.org/10.1016/j.celrep.2013.08.050
13 D Wu, D Hu, H Chen, G Shi, IS Fetahu, F Wu, K Rabidou, R Fang, L Tan, S Xu, H Liu, C Argueta, L Zhang, F Mao, G Yan, J Chen, Z Dong, R Lv, Y Xu, M Wang, Y Ye, S Zhang, D Duquette, S Geng, C Yin, CG Lian, GF Murphy, GK Adler, R Garg, L Lynch, P Yang, Y Li, F Lan, J Fan, Y Shi, YG Shi. Glucose-regulated phosphorylation of TET2 by AMPK reveals a pathway linking diabetes to cancer. Nature 2018; 559(7715): 637–641
https://doi.org/10.1038/s41586-018-0350-5
14 M Ko, J An, HS Bandukwala, L Chavez, T Aijö, WA Pastor, MF Segal, H Li, KP Koh, H Lähdesmäki, PG Hogan, L Aravind, A Rao. Modulation of TET2 expression and 5-methylcytosine oxidation by the CXXC domain protein IDAX. Nature 2013; 497(7447): 122–126
https://doi.org/10.1038/nature12052
15 L Lv, Q Wang, Y Xu, LC Tsao, T Nakagawa, H Guo, L Su, Y Xiong. Vpr targets TET2 for degradation by CRL4VprBP E3 ligase to sustain IL-6 expression and enhance HIV-1 replication. Mol Cell 2018; 70(5): 961–970.e5
https://doi.org/10.1016/j.molcel.2018.05.007
16 O Santt, T Pfirrmann, B Braun, J Juretschke, P Kimmig, H Scheel, K Hofmann, M Thumm, DH Wolf. The yeast GID complex, a novel ubiquitin ligase (E3) involved in the regulation of carbohydrate metabolism. Mol Biol Cell 2008; 19(8): 3323–3333
https://doi.org/10.1091/mbc.e08-03-0328
17 F Lampert, D Stafa, A Goga, MV Soste, S Gilberto, N Olieric, P Picotti, M Stoffel, M Peter. The multi-subunit GID/CTLH E3 ubiquitin ligase promotes cell proliferation and targets the transcription factor Hbp1 for degradation. eLife 2018; 7: e35528
https://doi.org/10.7554/eLife.35528
18 CM Skraban, CF Wells, P Markose, MT Cho, AI Nesbitt, PYB Au, A Begtrup, JA Bernat, LM Bird, K Cao, Brouwer APM de, EH Denenberg, G Douglas, KM Gibson, K Grand, A Goldenberg, AM Innes, J Juusola, M Kempers, E Kinning, DM Markie, MM Owens, K Payne, R Person, R Pfundt, A Stocco, CLS Turner, NE Verbeek, LE Walsh, TC Warner, PG Wheeler, D Wieczorek, AB Wilkens, E; Deciphering Developmental Disorders Study; Kleefstra T Zonneveld-Huijssoon, SP Robertson, A Santani, Gassen KLI van, MA Deardorff. WDR26 haploinsufficiency causes a recognizable syndrome of intellectual disability, seizures, abnormal gait, and distinctive facial features. Am J Hum Genet 2017; 101(1): 139–148
https://doi.org/10.1016/j.ajhg.2017.06.002
19 A Córdova-Palomera, M Fatjó-Vilas, C Gastó, V Navarro, MO Krebs, L Fañanás. Genome-wide methylation study on depression: differential methylation and variable methylation in monozygotic twins. Transl Psychiatry 2015; 5(4): e557
https://doi.org/10.1038/tp.2015.49
20 C Tangsuwansri, T Saeliw, S Thongkorn, W Chonchaiya, K Suphapeetiporn, A Mutirangura, T Tencomnao, VW Hu, T Sarachana. Investigation of epigenetic regulatory networks associated with autism spectrum disorder (ASD) by integrated global LINE-1 methylation and gene expression profiling analyses. PLoS One 2018; 13(7): e0201071
https://doi.org/10.1371/journal.pone.0201071
21 C Dong, H Zhang, L Li, W Tempel, P Loppnau, J Min. Molecular basis of GID4-mediated recognition of degrons for the Pro/N-end rule pathway. Nat Chem Biol 2018; 14(5): 466–473
https://doi.org/10.1038/s41589-018-0036-1
22 YW Zhang, Z Wang, W Xie, Y Cai, L Xia, H Easwaran, J Luo, RC Yen, Y Li, SB Baylin. Acetylation enhances TET2 function in protecting against abnormal DNA methylation during oxidative stress. Mol Cell 2017; 65(2): 323–335
https://doi.org/10.1016/j.molcel.2016.12.013
23 D Guallar, X Bi, JA Pardavila, X Huang, C Saenz, X Shi, H Zhou, F Faiola, J Ding, P Haruehanroengra, F Yang, D Li, C Sanchez-Priego, A Saunders, F Pan, VJ Valdes, K Kelley, MG Blanco, L Chen, H Wang, J Sheng, M Xu, M Fidalgo, X Shen, J Wang. RNA-dependent chromatin targeting of TET2 for endogenous retrovirus control in pluripotent stem cells. Nat Genet 2018; 50(3): 443–451
https://doi.org/10.1038/s41588-018-0060-9
24 Z Sun, AV Smrcka, S Chen. WDR26 functions as a scaffolding protein to promote Gβγ-mediated phospholipase C β2 (PLCβ2) activation in leukocytes. J Biol Chem 2013; 288(23): 16715–16725
https://doi.org/10.1074/jbc.M113.462564
25 SJ Chen, X Wu, B Wadas, JH Oh, A Varshavsky. An N-end rule pathway that recognizes proline and destroys gluconeogenic enzymes. Science 2017; 355(6323): eaal3655
https://doi.org/10.1126/science.aal3655
26 H Hochgerner, A Zeisel, P Lönnerberg, S Linnarsson. Conserved properties of dentate gyrus neurogenesis across postnatal development revealed by single-cell RNA sequencing. Nat Neurosci 2018; 21(2): 290–299
https://doi.org/10.1038/s41593-017-0056-2
27 J Muhr, DW Hagey. The cell cycle and differentiation as integrated processes: cyclins and CDKs reciprocally regulate Sox and Notch to balance stem cell maintenance. BioEssays 2021; 43(7): e2000285
https://doi.org/10.1002/bies.202000285
28 CV Borlongan. Regenerative medicine during the pandemic period. Brain Circ 2021; 7(1): 1–2
https://doi.org/10.4103/bc.bc_22_21
29 LM Farkas, WB Huttner. The cell biology of neural stem and progenitor cells and its significance for their proliferation versus differentiation during mammalian brain development. Curr Opin Cell Biol 2008; 20(6): 707–715
https://doi.org/10.1016/j.ceb.2008.09.008
30 Z Guo, M Chen, Y Chao, C Cai, L Liu, L Zhao, L Li, QR Bai, Y Xu, W Niu, L Shi, Y Bi, D Ren, F Yuan, S Shi, Q Zeng, K Han, Y Shi, S Bian, G He. RGCC balances self-renewal and neuronal differentiation of neural stem cells in the developing mammalian neocortex. EMBO Rep 2021; 22(9): e51781
https://doi.org/10.15252/embr.202051781
31 EC Gilmore, CA Walsh. Genetic causes of microcephaly and lessons for neuronal development. Wiley Interdiscip Rev Dev Biol 2013; 2(4): 461–478
https://doi.org/10.1002/wdev.89
32 M Groszer, R Erickson, DD Scripture-Adams, JD Dougherty, J Le Belle, JA Zack, DH Geschwind, X Liu, HI Kornblum, H Wu. PTEN negatively regulates neural stem cell self-renewal by modulating G0-G1 cell cycle entry. Proc Natl Acad Sci USA 2006; 103(1): 111–116
https://doi.org/10.1073/pnas.0509939103
33 NM Wilpert, F Marguet, C Maillard, F Guimiot, J Martinovic, S Drunat, T Attié-Bitach, F Razavi, A Tessier, Y Capri, A Laquerrière, N Bahi-Buisson. Human neuropathology confirms projection neuron and interneuron defects and delayed oligodendrocyte production and maturation in FOXG1 syndrome. Eur J Med Genet 2021; 64(9): 104282
https://doi.org/10.1016/j.ejmg.2021.104282
34 R Gruber, Z Zhou, M Sukchev, T Joerss, PO Frappart, ZQ Wang. MCPH1 regulates the neuroprogenitor division mode by coupling the centrosomal cycle with mitotic entry through the Chk1-Cdc25 pathway. Nat Cell Biol 2011; 13(11): 1325–1334
https://doi.org/10.1038/ncb2342
35 T Maraldi, C Angeloni, C Prata, S Hrelia. NADPH oxidases: redox regulators of stem cell fate and function. antioxidants 2021; 10(6): 973
https://doi.org/10.3390/antiox10060973
36 RR Zhang, QY Cui, K Murai, YC Lim, ZD Smith, S Jin, P Ye, L Rosa, YK Lee, HP Wu, W Liu, ZM Xu, L Yang, YQ Ding, F Tang, A Meissner, C Ding, Y Shi, GL Xu. Tet1 regulates adult hippocampal neurogenesis and cognition. Cell Stem Cell 2013; 13(2): 237–245
https://doi.org/10.1016/j.stem.2013.05.006
37 K Izumi. Disorders of transcriptional regulation: an emerging category of multiple malformation syndromes. Mol Syndromol 2016; 7(5): 262–273
https://doi.org/10.1159/000448747
38 JE Bestman, LC Huang, J Lee-Osbourne, P Cheung, HT Cline. An in vivo screen to identify candidate neurogenic genes in the developing Xenopus visual system. Dev Biol 2015; 408(2): 269–291
https://doi.org/10.1016/j.ydbio.2015.03.010
39 M Nassan, Q Li, PE Croarkin, W Chen, CL Colby, M Veldic, SL McElroy, GD Jenkins, E Ryu, JM Cunningham, M Leboyer, MA Frye, JM Biernacka. A genome wide association study suggests the association of muskelin with early onset bipolar disorder: implications for a GABAergic epileptogenic neurogenesis model. J Affect Disord 2017; 208: 120–129
https://doi.org/10.1016/j.jad.2016.09.049
40 N Huffman, D Palmieri, V Coppola. The CTLH complex in cancer cell plasticity. J Oncol 2019; 2019: 4216750
https://doi.org/10.1155/2019/4216750
[1] FMD-23026-OF-XXS_suppl_1 Download
[1] Ming Hou, Suji Wang, Dandan Yu, Xinyi Lu, Xiansen Zhao, Zhangpeng Chen, Chao Yan. Cannabidiol prevents depressive-like behaviors through the modulation of neural stem cell differentiation[J]. Front. Med., 2022, 16(2): 227-239.
[2] Xiaolin Fan, Yanzhen Xiong, Yuan Wang. A reignited debate over the cell(s) of origin for glioblastoma and its clinical implications[J]. Front. Med., 2019, 13(5): 531-539.
Viewed
Full text


Abstract

Cited

  Shared   
  Discussed