リケラボ論文検索は、全国の大学リポジトリにある学位論文・教授論文を一括検索できる論文検索サービスです。

リケラボ 全国の大学リポジトリにある学位論文・教授論文を一括検索するならリケラボ論文検索大学・研究所にある論文を検索できる

リケラボ 全国の大学リポジトリにある学位論文・教授論文を一括検索するならリケラボ論文検索大学・研究所にある論文を検索できる

大学・研究所にある論文を検索できる 「MFF機能不全のヒト乳歯由来幹細胞から分化したドーパミン作動性ニューロンにおけるミトコンドリアのカルシウム誘発性酸化ストレスと発達障害」の論文概要。リケラボ論文検索は、全国の大学リポジトリにある学位論文・教授論文を一括検索できる論文検索サービスです。

コピーが完了しました

URLをコピーしました

論文の公開元へ論文の公開元へ
書き出し

MFF機能不全のヒト乳歯由来幹細胞から分化したドーパミン作動性ニューロンにおけるミトコンドリアのカルシウム誘発性酸化ストレスと発達障害

孫, 梟 SUN, XIAO ソン, キョウ 九州大学

2023.03.20

概要

九州大学学術情報リポジトリ
Kyushu University Institutional Repository

Mitochondrial Calcium-Triggered Oxidative
Stress and Developmental Defects in
Dopaminergic Neurons Differentiated from
Deciduous Teeth-Derived Dental Pulp Stem Cells
with MFF Insufficiency
孫, 梟

https://hdl.handle.net/2324/6787516
出版情報:Kyushu University, 2022, 博士(歯学), 課程博士
バージョン:
権利関係:© 2022bytheauthors. Licensee MDPI,Basel,Switzerland. This
articleisanopenaccessarticle distributed underthetermsand conditions oftheCreativeCommons
Attribution (CCBY)license.

(様式3)



名 : 孫 梟

論 文 名 :Mitochondrial Calcium-Triggered Oxidative Stress and Developmental Defects in Dopaminergic
Neurons Differentiated from Deciduous Teeth-Derived Dental Pulp Stem Cells with MFF Insufficiency
(MFF機能不全のヒト乳歯由来幹細胞から分化したドーパミン作動性ニューロンにおけるミトコンドリア
のカルシウム誘発性酸化ストレスと発達障害)


分 :甲

論 文 内 容 の 要 旨
Mitochondrial fission factor (MFF) is an adapter that targets dynamin-related protein 1 from the cytosol to the
mitochondria for fission. Loss-of-function MFF mutations cause encephalopathy due to defective mitochondrial and
peroxisomal fission 2 (EMPF2). To elucidate the molecular mechanisms that were involved, we analyzed the
functional effects of MFF depletion in deciduous teeth-derived dental pulp stem cells differentiating into
dopaminergic neurons (DNs). When treated with MFF-targeting small interfering RNA, DNs showed impaired
neurite outgrowth and reduced mitochondrial signals in neurites harboring elongated mitochondria. MFF silencing
also caused mitochondrial Ca2+ accumulation through accelerated Ca2+ influx from the endoplasmic reticulum (ER)
via the inositol 1,4,5-trisphosphate receptor. Mitochondrial Ca2+ overload led DNs to produce excessive reactive
oxygen species (ROS), and downregulated peroxisome proliferator-activated receptor-gamma co-activator-1 alpha
(PGC-1α). MFF was co-immunoprecipitated with voltage-dependent anion channel 1, an essential component of the
ER-mitochondrial Ca2+ transport system. Folic acid supplementation normalized ROS levels, PGC-1α mediated
mitochondrial biogenesis, and neurite outgrowth in MFF depleted DNs, without affecting their mitochondrial
morphology or Ca2+ levels. We propose that MFF negatively regulates the mitochondrial Ca2+ influx from the ER.
MFF-insufficiency recapitulated the EMPF2 neuropathology with increased oxidative stress and suppressed
mitochondrial biogenesis. ROS and mitochondrial biogenesis might be potential therapeutic targets for EMPF2.

この論文で使われている画像

参考文献

1.

2.

3.

4.

5.

6.

7.

8.

Westermann, B. Mitochondrial fusion and fission in cell life and death. Nat. Rev. Mol. Cell Biol. 2010, 12, 872884. [CrossRef]

Adebayo, M.; Singh, S.; Singh, A.P.; Dasgupta, S. Mitochondrial fusion and fission: The fine-tune balance for cellular homeostasis.

FASEB J. 2021, 6, e21620. [CrossRef]

Brookes, R.S.; Yoon, Y.; Robotham, J.L.; Anders, M.W.; Sheu, S.S. Calcium, ATP, and ROS: A mitochondrial love-hate triangle. Am.

J. Physiol. Cell Physiol. 2004, 287, C817–C833. [CrossRef]

Archer, S.L. Mitochondrial dynamics—Mitochondrial fission and fusion in human diseases. N. Engl. J. Med. 2013, 369, 2236–2251.

[CrossRef]

Yapa, N.M.B.; Lisnyak, V.; Reljic, B.; Ryan, M.T. Mitochondrial dynamics in health and disease. FEBS Lett. 2021, 595, 1184–1204.

[CrossRef]

Waterham, H.R.; Koster, J.; van Roermund, C.W.T.; Mooyer, P.A.W.; Wanders, R.J.A.; Leonard, J.V. A lethal defect of mitochondrial

and peroxisomal fission. N. Engl. J. Med. 2007, 356, 1736–1741. [CrossRef]

Sheffer, R.; Douiev, L.; Edvardson, S.; Shaag, A.; Tamimi, K.; Soiferman, D.; Meiner, V.; Saada, A. Postnatal microcephaly and pain

insensitivity due to a de novo heterozygous DNM1L mutation causing impaired mitochondrial fission and function. Am. J. Med.

Genet. 2016, 170, 1603–1607. [CrossRef]

Vanstone, J.R.; Smith, A.M.; McBride, S.; Naas, T.; Holcik, M.; Antoun, G.; Harper, M.-E.; Michaud, J.; Sell, E.; Chakraborty, P.; et al.

DNM1L-related mitochondrial fission defect presenting as refractory epilepsy. Eur. J. Hum. Genet. 2016, 24, 1084–1088. [CrossRef]

Antioxidants 2022, 11, 1361

9.

10.

11.

12.

13.

14.

15.

16.

17.

18.

19.

20.

21.

22.

23.

24.

25.

26.

27.

28.

29.

30.

31.

32.

33.

34.

16 of 18

Fahrner, J.A.; Liu, R.; Perry, M.S.; Klein, J.; Chan, D.C. A novel de novo dominant negative mutation in DNM1L impairs

mitochondrial fission and presents as childhood epileptic encephalopathy. Am. J. Med. Genet. A 2016, 170, 2002–2011. [CrossRef]

Chao, Y.-H.; Robak, L.A.; Xia, F.; Koenig, M.K.; Adesina, A.; Bacino, C.A.; Scaglia, F.; Bellen, H.J.; Wangler, M.F. Missense variants

in the middle domain of DNM1L in cases of infantile encephalopathy alter peroxisomes and mitochondria when assayed in

Drosophila. Hum. Mol. Genet. 2016, 25, 1846–1856. [CrossRef]

Yoon, G.; Malam, Z.; Paton, T.; Marshall, C.R.; Hyatt, E.; Ivakine, Z.; Scherer, S.W.; Lee, K.-S.; Hawkins, C.; Cohn, R.D. Lethal

disorder of mitochondrial fission caused by mutations in DNM1L. J. Pediat. 2016, 171, 313–316. [CrossRef]

Nasca, A.; Legati, A.; Baruffini, E.; Nolli, C.; Moroni, I.; Ardissone, A.; Goffrini, P.; Ghezzi, D. Biallelic mutations in DNM1L are

associated with a slowly progressive infantile encephalopathy. Hum. Mutat. 2016, 37, 898–903. [CrossRef]

Vandeleur, D.; Chen, C.V.; Huang, E.J.; Connolly, A.J.; Sanchez, H.; Moon-Grady, A.J. Novel and lethal case of cardiac involvement

in DNM1L mitochondrial encephalopathy. Am. J. Med. Genet. A 2019, 179, 2486–2489. [CrossRef]

Ishihara, N.; Nomura, M.; Jofuku, A.; Kato, H.; Suzuki, S.O.; Masuda, K.; Otera, H.; Nakanishi, Y.; Nonaka, I.; Goto, Y.; et al.

Mitochondrial fission factor Drp1 is essential for embryonic development and synapse formation in mice. Nat. Cell Biol. 2009,

11, 958–966. [CrossRef]

Gandre-Babbe, S.; van der Bliek, A.M. The novel tail-anchored membrane protein MFF controls mitochondrial and peroxisomal

fission in mammalian cells. Mol. Biol. Cell 2008, 19, 2402–2412. [CrossRef]

Wasiak, S.; Zunino, R.; McBride, H.M. Bax/Bak promote sumoylation of DRP1 and its stable association with mitochondria

during apoptotic cell death. J. Cell Biol. 2007, 177, 439–450. [CrossRef]

Karbowski, M.; Neutzner, A.; Youle, R.J. The mitochondrial E3 ubiquitin ligase MARCH5 is required for Drp1 dependent

mitochondrial division. J. Cell Biol. 2007, 178, 71–84. [CrossRef]

Otera, H.; Wang, C.; Cleland, M.M.; Setoguchi, K.; Yokota, S.; Youle, R.J.; Mihara, K. MFF is an essential factor for mitochondrial

recruitment of Drp1 during mitochondrial fission in mammalian cells. J. Cell Biol. 2010, 191, 1141–1158. [CrossRef]

Shamseldin, H.E.; Alshammari, M.; Al-Sheddi, T.; Salih, M.A.; Alkhalidi, H.; Kentab, A.; Repetto, G.M.; Hashem, M.; Alkuraya,

F.S. Genomic analysis of mitochondrial diseases in a consanguineous population reveals novel candidate disease genes. J. Med.

Genet. 2012, 49, 234–241. [CrossRef]

Koch, J.; Feichtinger, R.G.; Freisinger, P.; Pies, M.; Schrodl, F.; Iuso, A.; Sperl, W.; Mayr, J.A.; Prokisch, H.; Haack, T.B. Disturbed

mitochondrial and peroxisomal dynamics due to loss of MFF causes Leigh-like encephalopathy, optic atrophy and peripheral

neuropathy. J. Med. Genet. 2016, 53, 270–278. [CrossRef]

Nasca, A.; Nardecchia, F.; Commone, A.; Semeraro, M.; Legati, A.; Garavaglia, B.; Ghezzi, D.; Leuzzi, V. Clinical and biochemical

features in a patient with mitochondrial fission factor gene alteration. Front. Genet. 2018, 9, 625. [CrossRef]

Panda, I.; Ahmad, I.; Sagar, S.; Zahra, S.; Shamim, U.; Sharma, S.; Faruq, M. Encephalopathy due to defective mitochondrial and

peroxisomal fission 3 caused by a novel MFF gene mutation in a young child. Clin. Genet. 2020, 97, 933–937. [CrossRef]

Lewis, T.L., Jr.; Kwon, S.K.; Lee, A.; Shaw, R.; Polleux, F. MFF-dependent mitochondrial fission regulates presynaptic release and

axon branching by limiting axonal mitochondria size. Nat. Commun. 2018, 9, 5008. [CrossRef]

Feissner, R.F.; Skalska, J.; Gaum, W.E.; Sheu, S.S. Crosstalk signaling between mitochondrial Ca2+ and ROS. Front. Biosci. 2009,

14, 1197–1218. [CrossRef]

Baev, A.Y.; Vinokurov, A.Y.; Novikova, I.N.; Dremin, V.V.; Potapova, E.V.; Abramov, A.Y. Interaction of Mitochondrial Calcium

and ROS in Neurodegeneration. Cells 2022, 11, 706. [CrossRef]

Gronthos, S.; Mankani, M.; Brahim, J.; Robey, P.G.; Shi, S. Postnatal human dental pulp stem cells (DPSCs) in vitro and in vivo.

Proc. Natl. Acad. Sci. USA 2000, 97, 13625–13630. [CrossRef]

Miura, M.; Gronthos, S.; Zhao, M.; Lu, B.; Fisher, L.W.; Robey, P.G.; Shi, S. SHED: Stem cells from human exfoliated deciduous

teeth. Proc. Natl. Acad. Sci. USA 2003, 100, 5807–5812. [CrossRef]

Fujii, H.; Matsubara, K.; Sakai, K.; Ito, M.; Ohno, K.; Ueda, M.; Yamamoto, A. Dopaminergic differentiation of stem cells from

human deciduous teeth and their therapeutic benefits for Parkinsonian rats. Brain Res. 2015, 1613, 59–72. [CrossRef]

Kanafi, M.; Majumdar, D.; Bhonde, R.; Gupta, P.; Datta, I. Midbrain cues dictate differentiation of human dental pulp stem cells

towards functional dopaminergic neurons. J. Cell Physiol. 2014, 229, 1369–1377. [CrossRef]

Majumdar, D.; Kanafi, M.; Bhonde, R.; Gupta, P.; Datta, I. Differential neuronal plasticity of dental pulp stem cells from exfoliated

deciduous and permanent teeth towards dopaminergic neurons. J. Cell Physiol. 2016, 231, 2048–2063. [CrossRef]

Masuda, K.; Han, X.; Kato, H.; Sato, H.; Zhang, Y.; Sun, X.; Hirofuji, Y.; Yamaza, H.; Yamada, A.; Fukumoto, S. Dental pulp-derived

mesenchymal stem cells for modeling genetic disorders. Int. J. Mol. Sci. 2021, 22, 2269. [CrossRef]

Kato, H.; Pham, T.M.T.; Yamaza, H.; Masuda, K.; Hirofuji, Y.; Han, X.; Sato, H.; Taguchi, T.; Nonaka, K. Mitochondria regulate the

differentiation of stem cells from human exfoliated deciduous teeth. Cell Struct. Funct. 2017, 42, 105–116. [CrossRef]

Sun, X.; Kato, H.; Sato, H.; Han, X.; Hirofuji, Y.; Kato, A.; Sakai, Y.; Ohga, S.; Fukumoto, S.; Masuda, K. Dopamine-related

oxidative stress and mitochondrial dysfunction in dopaminergic neurons differentiated from deciduous teeth-derived stem cells

of children with Down syndrome. FASEB BioAdv. 2022, 4, 454–467. [CrossRef]

Schneider, C.A.; Rasband, W.S.; Eliceiri, K.W. NIH Image to ImageJ: 25 years of image analysis. Nat. Methods 2012, 9, 671–675.

[CrossRef]

Antioxidants 2022, 11, 1361

35.

36.

37.

38.

39.

40.

41.

42.

43.

44.

45.

46.

47.

48.

49.

50.

51.

52.

53.

54.

55.

56.

57.

58.

59.

17 of 18

Sun, X.; Kato, H.; Sato, H.; Torio, M.; Han, X.; Zhang, Y.; Hirofuji, Y.; Kato, A.; Sakai, Y.; Ohga, S.; et al. Impaired neurite

development and mitochondrial dysfunction associated with calcium accumulation in dopaminergic neurons differentiated from

the dental pulp stem cells of a patient with metatropic dysplasia. Biochem. Biophys. Rep. 2021, 26, 100968. [CrossRef]

Kageyama, Y.; Zhang, Z.; Roda, R.; Fukaya, M.; Wakabayashi, J.; Wakabayashi, N.; Kensler, T.W.; Reddy, P.H.; Iijima, M.; Sesaki,

H. Mitochondrial division ensures the survival of postmitotic neurons by suppressing oxidative damage. J. Cell Biol. 2012,

197, 535–551. [CrossRef]

Rossi, A.; Pizzo, P.; Filadi, R. Calcium, mitochondria and cell metabolism: A functional triangle in bioenergetics. Biochim. Biophys.

Acta Mol. Cell Res. 2019, 1866, 1068–1078. [CrossRef]

Garbincius, J.F.; Elrod, J.W. Mitochondrial calcium exchange in physiology and disease. Physiol. Rev. 2022, 102, 893–992.

[CrossRef]

Márta, K.; Hasan, P.; Rodríguez-Prados, M.; Paillard, M.; Hajnóczky, G. Pharmacological inhibition of the mitochondrial Ca2+

uniporter: Relevance for pathophysiology and human therapy. J. Mol. Cell Cardiol. 2021, 151, 135–144. [CrossRef]

Rizzuto, R.; Brini, M.; Murgia, M.; Pozzan, T. Microdomains with high Ca2+ close to IP3-sensitive channels that are sensed by

neighboring mitochondria. Science 1993, 262, 744–747. [CrossRef]

Gafni, J.; Munsch, J.A.; Lam, T.H.; Catlin, M.C.; Costa, L.G.; Molinski, T.F.; Pessah, I.N. Xestospongins: Potent membrane

permeable blockers of the inositol 1,4,5-trisphosphate receptor. Neuron 1997, 19, 723–733. [CrossRef]

Joshi, R.; Adhikari, S.; Patro, B.S.; Chattopadhyay, S.; Mukherjee, T. Free radical scavenging behavior of folic acid: Evidence for

possible antioxidant activity. Free Radic. Biol. Med. 2001, 30, 1390–1399. [CrossRef]

Mentch, S.J.; Locasale, J.W. One-carbon metabolism and epigenetics: Understanding the specificity. Ann. N. Y. Acad. Sci. 2016,

1363, 91–98. [CrossRef]

Zhang, Y.; Kato, H.; Sato, H.; Yamaza, H.; Hirofuji, Y.; Han, X.; Masuda, K.; Nonaka, K. Folic acid-mediated mitochondrial

activation for protection against oxidative stress in human dental pulp stem cells derived from deciduous teeth. Biochem. Biophys.

Res. Commun. 2019, 508, 850–856. [CrossRef]

Lee, S.; Min, K.T. The Interface between ER and Mitochondria: Molecular Compositions and Functions. Mol. Cells 2018,

41, 1000–1007.

Szabadkai, G.; Bianchi, K.; Varnai, P.; de Stefani, D.; Wieckowski, M.R.; Cavagna, D.; Nagy, A.I.; Balla, T.; Rizzuto, R. Chaperonemediated coupling of endoplasmic reticulum and mitochondrial Ca2+ channels. J. Cell Biol. 2006, 175, 901–911. [CrossRef]

Yamamoto, T.; Yamada, A.; Watanabe, M.; Yoshimura, Y.; Yamazaki, N.; Yoshimura, Y.; Yamauchi, T.; Kataoka, M.; Nagata, T.;

Terada, H.; et al. VDAC1, having a shorter N-terminus than VDAC2 but showing the same migration in an SDS-polyacrylamide

gel, is the predominant form expressed in mitochondria of various tissues. J. Proteome Res. 2006, 5, 3336–3344. [CrossRef]

Seo, J.H.; Chae, Y.C.; Kossenkov, A.V.; Lee, Y.G.; Tang, H.Y.; Agarwal, E.; Gabrilovich, D.I.; Languino, L.R.; Speicher, D.W.;

Shastrula, P.K.; et al. MFF Regulation of Mitochondrial Cell Death Is a Therapeutic Target in Cancer. Cancer Res. 2019,

79, 6215–6226. [CrossRef]

Hansford, R.G. Physiological role of mitochondrial Ca2+ transport. J. Bioenerg. Biomembr. 1994, 26, 495–508. [CrossRef]

Turrens, J.F. Mitochondrial formation of reactive oxygen species. J. Physiol. 2003, 552 Pt 2, 335–344. [CrossRef]

Palty, R.; Silverman, W.F.; Hershfinkel, M.; Caporale, T.; Sensi, S.L.; Parnis, J.; Nolte, C.; Fishman, D.; Shoshan-Barmatz, V.;

Herrmann, S.; et al. NCLX is an essential component of mitochondrial Na+ /Ca2+ ; exchange. Proc. Natl. Acad. Sci. USA 2010,

107, 436–441. [CrossRef] [PubMed]

Bernardi, P.; Azzone, G.F. Regulation of Ca2+ efflux in rat liver mitochondria. Role of membrane potential. Eur. J. Biochem. 1983,

134, 377–383. [CrossRef] [PubMed]

Thirupathi, A.; de Souza, C.T. Multi-regulatory network of ROS: The interconnection of ROS, PGC-1 alpha, and AMPK-SIRT1

during exercise. J. Physiol. Biochem. 2017, 73, 487–494. [CrossRef]

Bouchez, C.; Devin, A. Mitochondrial Biogenesis and Mitochondrial Reactive Oxygen Species (ROS): A Complex Relationship

Regulated by the cAMP/PKA Signaling Pathway. Cells 2019, 8, 287. [CrossRef]

St-Pierre, J.; Drori, S.; Uldry, M.; Silvaggi, J.M.; Rhee, J.; Jäger, S.; Handschin, C.; Zheng, K.; Lin, J.; Yang, W.; et al. Suppression of

reactive oxygen species and neurodegeneration by the PGC-1 transcriptional coactivators. Cell 2006, 127, 397–408. [CrossRef]

Garnier, A.; Fortin, D.; Deloménie, C.; Momken, I.; Veksler, V.; Ventura-Clapier, R. Depressed mitochondrial transcription factors

and oxidative capacity in rat failing cardiac and skeletal muscles. J. Physiol. 2003, 551, 491–501. [CrossRef]

Javadov, S.; Purdham, D.M.; Zeidan, A.; Karmazyn, M. NHE-1 inhibition improves cardiac mitochondrial function through

regulation of mitochondrial biogenesis during postinfarction remodeling. Am. J. Physiol. Heart Circ. Physiol. 2006,

291, H1722–H1730. [CrossRef]

Wang, B.; Sun, J.; Ma, Y.; Wu, G.; Tian, Y.; Shi, Y.; Le, G. Resveratrol preserves mitochondrial function, stimulates mitochondrial

biogenesis, and attenuates oxidative stress in regulatory T cells of mice fed a high-fat diet. J. Food Sci. 2014, 79, H1823–H1831.

[CrossRef]

Baldelli, S.; Aquilano, K.; Ciriolo, M.R. PGC-1↵ buffers ROS-mediated removal of mitochondria during myogenesis. Cell Death

Dis. 2014, 5, e1515. [CrossRef]

Antioxidants 2022, 11, 1361

60.

61.

18 of 18

Lee, W.C.; Li, L.C.; Chen, J.B.; Chang, H.W. Indoxyl sulfate-induced oxidative stress, mitochondrial dysfunction, and impaired

biogenesis are partly protected by vitamin C and N-acetylcysteine. Sci. World J. 2015, 2015, 620826. [CrossRef]

Takeichi, Y.; Miyazawa, T.; Sakamoto, S.; Hanada, Y.; Wang, L.; Gotoh, K.; Uchida, K.; Katsuhara, S.; Sakamoto, R.; Ishihara,

T.; et al. Non-alcoholic fatty liver disease in mice with hepatocyte-specific deletion of mitochondrial fission factor. Diabetologia

2021, 64, 2092–2107. [CrossRef] [PubMed]

Table S1. Alexa Fluor-conjugated secondary antibodies.

Figure S1. (a) DNs were stained with mouse monoclonal anti-Tom20 (#sc-17764; Santa Cruz Biotechnology, Paso Robles, CA, USA), rabbit polyclonal anti-MFF (#170909-1-AP; Proteintech, Rosemont, IL, USA), rabbit polyclonal anti-MID51 (#20164-1-AP; Proteintech), rabbit polyclonal antiMID49 (#16413-1-AP; Proteintech), and mouse monoclonal anti-DRP1 (#611113; BD Biosciences,

San Jose, CA, USA) antibodies. Scale bars = 10 μm. The boxed regions on the merged images are

shown at a greater magnification in the lower panels. Scale bars = 2 μm. (b) MFF protein levels

were measured by western blotting using anti-MFF and mouse monoclonal anti-α-tubulin (#sc32293; Santa Cruz Biotechnology) antibodies. The mean ± SEM was taken from three independent

experiments. ***p < 0.001. (c) DRP1, MID49, and MID51 protein levels were detected by western

blotting using the aforementioned antibodies.

2 of 8

Figure S2. NURR1 and TH mRNA expression in DNs were measured using RT-qPCR. (a) The primer set for NURR1 is forward 5'-GCACTTCGGCAGAGTTGAATGA-3' and reverse 5'GGTGGCTGTGTTGCTGGTAGTT-3'. (b) The primer set of TH was purchased from Bio-Rad

(#qHsaCED0001111; Bio-Rad, Hercules, CA, USA). Relative expression of the target gene was analyzed using the comparative threshold cycle method by normalizing to 18s expression. The mean ±

SEM was taken from three independent experiments. n.s., not significant.

3 of 8

Figure S3. (a) DNs were stained with Rhod-2 AM and MTG. Mitochondria in neurites were observed using confocal microscopy. Scale bar = 10 μm. To measure the mitochondrial Ca 2+ level per

mitochondrion in neurites, fluorescence intensity of 10 mitochondria in each case were measured

and Rhod-2 AM intensity was divided by that of MTG. The mean ± SEM was taken from three

independent experiments. **P < 0.01. (b) DNs were stained with MitoSOX Red and MTG. Mitochondria in neurites were observed using confocal microscopy. Scale bar = 5 μm. To measure the

ROS level per mitochondrion in neurites, the fluorescence intensities of 10 mitochondria in each

case were measured and MitoSOX Red intensity was divided by that of MTG. The mean ± SEM

was taken from three independent experiments. **P < 0.01.

4 of 8

Figure S4. (a) Cleaved caspase-3 expression in DNs was detected using western blotting. The protein sample of control cells treated with 5 μΜ actinomycin D (ActD) for 12 h was used as a positive control for apoptotic cells. (b) DNs were stained with mouse monoclonal anti-cleaved

caspase-3 antibody (#9664S; Cell Signaling Technology, Danvers, MA, USA) and DAPI. Scale bars

= 50 μm. The left graph shows the percentage of cleaved caspase-3-positive cells; 100 cells were

counted for each case. DNs were stained with trypan blue solution to distinguish between live and

dead cells. Graph on the right shows the number of living cells. The mean ± SEM was taken from

three independent experiments. n.s., not significant.

5 of 8

Figure S5. (a) SHEDs were transfected with negative control- (Ctrl-siR) and DRP1-siRNA (DRP1siR) and differentiated into DNs. The DRP1 siRNA sequences were as follows: sense 5'GUAAUACUGAGACUUUGUUdTdT-3' and antisense 5'-AACAAAGUCUCAGUAUUACdTdT3'. The control siRNA was purchased from Sigma-Aldrich, St. Louis, MO, USA (SIC001-10NMOL).

DRP1 and MFF levels were measured using western blotting with anti-MFF (#170909-1-AP; Proteintech), mouse monoclonal anti-DRP1 (#611113; BD Biosciences), and mouse monoclonal anti-αtubulin (#sc-32293; Santa Cruz Biotechnology) antibodies. The mean ± SEM was from three independent experiments. n.s., not significant, **p < 0.001. (b) DNs were stained with anti-Tom20 and

anti-TH antibodies and counterstained with DAPI. Mitochondrial proportion in neurites was observed using confocal microscopy. Scale bar = 5 μm. Mitochondrial length in neurites was measured for 10 mitochondria of each case (left graph), and the number of mitochondria per 50 μm

neurite was measured for 10 neurites of each case (right graph). The mean ± SEM was taken from

three independent experiments. ***p < 0.001. (c) DNs were stained with Rhod2-AM. Fluorescence

intensity was measured using a plate reader. The mean ± SEM was taken from three independent

experiments. n.s., not significant. (d) DNs were stained with MitoSOX Red, and the signal was

measured using flow cytometry. The mean ± SEM was taken from three independent experiments.

*p < 0.05.

6 of 8

Figure S6. DNs were treated with 1 μM Ru-R for the last 4 h of culture. (a, b) DNs were stained

with Rhod-2 AM or Fluo-4 AM. Fluorescence intensities of these fluorescent probes were measured using a plate reader. The mean ± SEM was taken from three independent experiments. n.s.,

not significant, *p < 0.05, **p < 0.01. (c) DNs were stained with MitoSOX Red, and the signal was

measured using flow cytometry. The mean ± SEM was taken from three independent experiments.

n.s., not significant, *p < 0.05.

Figure S7. Full scans of the western blots from Figure 4c. (a, b) IP was performed using an antiMFF and anti-VDAC1 antibodies. Immunoprecipitants were detected by western blotting (WB)

using the indicated antibodies.

7 of 8

Figure S8. Immunofluorescence analysis of mitochondria and ER contact. DNs were stained with

mouse monoclonal anti-Tom20 (#sc-17764; Santa Cruz Biotechnology) and rabbit polyclonal antiSec61B (#14648S-AP; Cell Signalling Technology) antibodies. Scale bar = 20 μm. The boxed regions

on the merged images are shown at a greater magnification in the lower panels. Scale bars = 2 μm.

ER and mitochondrial colocalization were analyzed through Mander's co-localization coefficients

using ImageJ software version 1.53 with JACop plugin. Fifteen DNs were analyzed in each case.

The mean ± SEM was taken from three independent experiments. ***p < 0.001.

8 of 8

Figure S9. (a) DNs were stained with Rhod-2 AM. The fluorescence intensity of the probe was

measured using a plate reader. The mean ± SEM was taken from three independent experiments.

n.s., not significant, *p < 0.05, **P < 0.01. (b) DNs were stained with anti-Tom20 and anti-TH antibodies and counterstained with DAPI. Scale bars = 10 μm. The boxed regions on the Tom20stained images are shown at a greater magnification in the lower panels. Scale bar = 5 μm. Average mitochondrial length in neurite was measured for 10 mitochondria of each case. The mean ±

SEM was taken from three independent experiments. n.s., not significant, ***p < 0.001.

...

参考文献をもっと見る

全国の大学の
卒論・修論・学位論文

一発検索!

この論文の関連論文を見る