Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Letter
  • Published:

Mutations in SLC25A46, encoding a UGO1-like protein, cause an optic atrophy spectrum disorder

Abstract

Dominant optic atrophy (DOA)1,2 and axonal peripheral neuropathy (Charcot-Marie-Tooth type 2, or CMT2)3 are hereditary neurodegenerative disorders most commonly caused by mutations in the canonical mitochondrial fusion genes OPA1 and MFN2, respectively4. In yeast, homologs of OPA1 (Mgm1) and MFN2 (Fzo1)5,6 work in concert with Ugo1, for which no human equivalent has been identified thus far7. By whole-exome sequencing of patients with optic atrophy and CMT2, we identified four families with recessive mutations in SLC25A46. We demonstrate that SLC25A46, like Ugo1, is a modified carrier protein that has been recruited to the outer mitochondrial membrane and interacts with the inner membrane remodeling protein mitofilin (Fcj1). Loss of function in cultured cells and in zebrafish unexpectedly leads to increased mitochondrial connectivity, while severely affecting the development and maintenance of neurons in the fish. The discovery of SLC25A46 strengthens the genetic overlap between optic atrophy and CMT2 while exemplifying a new class of modified solute transporters linked to mitochondrial dynamics.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Pedigrees and clinical features of optic atrophy plus syndromes with variants in SLC25A46.
Figure 2: SLC25A46 localizes to the outer mitochondrial membrane and interacts with mitofilin on the inner membrane.
Figure 3: SLC25A46 levels regulate mitochondrial morphology.
Figure 4: Patient-derived fibroblasts have hyperfilamentous mitochondria and are respiratory deficient.
Figure 5: slc25a46 knockdown in zebrafish affects the growth and maintenance of neurons.
Figure 6: Mitochondrial distribution and morphology in zebrafish motor neurons at 48 h.p.f.

Similar content being viewed by others

Accession codes

Accessions

NCBI Reference Sequence

References

  1. Delettre, C., Lenaers, G., Pelloquin, L., Belenguer, P. & Hamel, C.P. OPA1 (Kjer type) dominant optic atrophy: a novel mitochondrial disease. Mol. Genet. Metab. 75, 97–107 (2002).

    Article  CAS  Google Scholar 

  2. Alexander, C. et al. OPA1, encoding a dynamin-related GTPase, is mutated in autosomal dominant optic atrophy linked to chromosome 3q28. Nat. Genet. 26, 211–215 (2000).

    Article  CAS  Google Scholar 

  3. Züchner, S. et al. Mutations in the mitochondrial GTPase mitofusin 2 cause Charcot-Marie-Tooth neuropathy type 2A. Nat. Genet. 36, 449–451 (2004).

    Article  Google Scholar 

  4. Chan, D.C. Dissecting mitochondrial fusion. Dev. Cell 11, 592–594 (2006).

    Article  CAS  Google Scholar 

  5. Sesaki, H. & Jensen, R.E. UGO1 encodes an outer membrane protein required for mitochondrial fusion. J. Cell Biol. 152, 1123–1134 (2001).

    Article  CAS  Google Scholar 

  6. Sesaki, H. & Jensen, R.E. Ugo1p links the Fzo1p and Mgm1p GTPases for mitochondrial fusion. J. Biol. Chem. 279, 28298–28303 (2004).

    Article  CAS  Google Scholar 

  7. van der Bliek, A.M., Shen, Q. & Kawajiri, S. Mechanisms of mitochondrial fission and fusion. Cold Spring Harb. Perspect. Biol. 5, a011072 (2013).

    Article  Google Scholar 

  8. Braathen, G.J. Genetic epidemiology of Charcot-Marie-Tooth disease. Acta Neurol. Scand. Suppl. 126, iv–22 (2012).

    Article  Google Scholar 

  9. Züchner, S. et al. Axonal neuropathy with optic atrophy is caused by mutations in mitofusin 2. Ann. Neurol. 59, 276–281 (2006).

    Article  Google Scholar 

  10. Yu-Wai-Man, P., Griffiths, P.G., Hudson, G. & Chinnery, P.F. Inherited mitochondrial optic neuropathies. J. Med. Genet. 46, 145–158 (2009).

    Article  CAS  Google Scholar 

  11. Amati-Bonneau, P. et al. OPA1 mutations induce mitochondrial DNA instability and optic atrophy “plus” phenotypes. Brain 131, 338–351 (2008).

    Article  Google Scholar 

  12. Züchner, S. & Vance, J.M. Emerging pathways for hereditary axonopathies. J. Mol. Med. 83, 935–943 (2005).

    Article  Google Scholar 

  13. Gonzalez, M.A. et al. GEnomes Management Application (GEM.app): a new software tool for large-scale collaborative genome analysis. Hum. Mutat. 34, 842–846 (2013).

    Article  CAS  Google Scholar 

  14. Palmieri, F. The mitochondrial transporter family SLC25: identification, properties and physiopathology. Mol. Aspects Med. 34, 465–484 (2013).

    Article  CAS  Google Scholar 

  15. Haitina, T., Lindblom, J., Renström, T. & Fredriksson, R. Fourteen novel human members of mitochondrial solute carrier family 25 (SLC25) widely expressed in the central nervous system. Genomics 88, 779–790 (2006).

    Article  CAS  Google Scholar 

  16. Hoppins, S., Horner, J., Song, C., McCaffery, J.M. & Nunnari, J. Mitochondrial outer and inner membrane fusion requires a modified carrier protein. J. Cell Biol. 184, 569–581 (2009).

    Article  CAS  Google Scholar 

  17. Coonrod, E.M., Karren, M.A. & Shaw, J.M. Ugo1p is a multipass transmembrane protein with a single carrier domain required for mitochondrial fusion. Traffic 8, 500–511 (2007).

    Article  CAS  Google Scholar 

  18. Lamarca, V. et al. Two isoforms of PSAP/MTCH1 share two proapoptotic domains and multiple internal signals for import into the mitochondrial outer membrane. Am. J. Physiol. Cell Physiol. 293, C1347–C1361 (2007).

    Article  CAS  Google Scholar 

  19. Zaltsman, Y. et al. MTCH2/MIMP is a major facilitator of tBID recruitment to mitochondria. Nat. Cell Biol. 12, 553–562 (2010).

    Article  CAS  Google Scholar 

  20. Robinson, A.J., Kunji, E.R.S. & Gross, A. Mitochondrial carrier homolog 2 (MTCH2): the recruitment and evolution of a mitochondrial carrier protein to a critical player in apoptosis. Exp. Cell Res. 318, 1316–1323 (2012).

    Article  CAS  Google Scholar 

  21. Harner, M. et al. The mitochondrial contact site complex, a determinant of mitochondrial architecture. EMBO J. 30, 4356–4370 (2011).

    Article  CAS  Google Scholar 

  22. Karbowski, M. et al. Quantitation of mitochondrial dynamics by photolabeling of individual organelles shows that mitochondrial fusion is blocked during the Bax activation phase of apoptosis. J. Cell Biol. 164, 493–499 (2004).

    Article  CAS  Google Scholar 

  23. Polyakov, V.Y., Soukhomlinova, M.Y. & Fais, D. Fusion, fragmentation, and fission of mitochondria. Biochemistry (Mosc.) 68, 838–849 (2003).

    Article  CAS  Google Scholar 

  24. Zanna, C. et al. OPA1 mutations associated with dominant optic atrophy impair oxidative phosphorylation and mitochondrial fusion. Brain 131, 352–367 (2008).

    Article  Google Scholar 

  25. Song, Z., Ghochani, M., Mccaffery, J.M., Frey, T.G. & Chan, D.C. Mitofusins and OPA1 mediate sequential steps in mitochondrial membrane fusion. Mol. Biol. Cell 20, 3525–3532 (2009).

    Article  CAS  Google Scholar 

  26. Shamas-Din, A. et al. Multiple partners can kiss-and-run: Bax transfers between multiple membranes and permeabilizes those primed by tBid. Cell Death Dis. 5, e1277 (2014).

    Article  CAS  Google Scholar 

  27. Hirota, T. et al. Genome-wide association study identifies eight new susceptibility loci for atopic dermatitis in the Japanese population. Nat. Genet. 44, 1222–1226 (2012).

    Article  CAS  Google Scholar 

  28. Li, Y.-C. et al. Mitochondrial accumulation in the distal part of the initial segment of chicken spinal motoneurons. Brain Res. 1026, 235–243 (2004).

    Article  CAS  Google Scholar 

  29. Narendra, D.P. & Youle, R.J. Neurodegeneration: trouble in the cell's powerhouse. Nature 483, 418–419 (2012).

    Article  CAS  Google Scholar 

  30. Tu, Y.-T. & Barrientos, A. The human mitochondrial DEAD-box protein DDX28 resides in RNA granules and functions in mitoribosome assembly. Cell Rep. 10, 854–864 (2015).

    Article  CAS  Google Scholar 

  31. Fontanesi, F., Clemente, P. & Barrientos, A. Cox25 teams up with Mss51, Ssc1, and Cox14 to regulate mitochondrial cytochrome c oxidase subunit 1 expression and assembly in Saccharomyces cerevisiae. J. Biol. Chem. 286, 555–566 (2011).

    Article  CAS  Google Scholar 

  32. Divakaruni, A.S., Paradyse, A., Ferrick, D.A., Murphy, A.N. & Jastroch, M. Analysis and interpretation of microplate-based oxygen consumption and pH data. Methods Enzymol. 547, 309–354 (2014).

    Article  CAS  Google Scholar 

  33. Kucenas, S., Snell, H. & Appel, B. nkx2.2a promotes specification and differentiation of a myelinating subset of oligodendrocyte lineage cells in zebrafish. Neuron Glia Biol. 4, 71–81 (2008).

    Article  Google Scholar 

  34. Poulain, F.E. & Chien, C.-B. Proteoglycan-mediated axon degeneration corrects pre-target topographic sorting errors. Neuron 78, 49–56 (2013).

    Article  CAS  Google Scholar 

  35. Myers, P.Z., Eisen, J.S. & Westerfield, M. Development and axonal outgrowth of identified motoneurons in the zebrafish. J. Neurosci. 6, 2278–2289 (1986).

    Article  CAS  Google Scholar 

  36. Plucin´ska, G. et al. In vivo imaging of disease-related mitochondrial dynamics in a vertebrate model system. J. Neurosci. 32, 16203–16212 (2012).

    Article  Google Scholar 

  37. Prince, J., Nolen, T.G. & Coelho, L. Defensive ink pigment processing and secretion in Aplysia californica: concentration and storage of phycoerythrobilin in the ink gland. J. Exp. Biol. 201, 1595–1613 (1998).

    CAS  PubMed  Google Scholar 

  38. Rebelo, A.P., Williams, S.L. & Moraes, C.T. In vivo methylation of mtDNA reveals the dynamics of protein-mtDNA interactions. Nucleic Acids Res. 37, 6701–6715 (2009).

    Article  CAS  Google Scholar 

  39. Thisse, C. & Thisse, B. High-resolution in situ hybridization to whole-mount zebrafish embryos. Nat. Protoc. 3, 59–69 (2008).

    Article  CAS  Google Scholar 

  40. Burgess, H.A. & Granato, M. Modulation of locomotor activity in larval zebrafish during light adaptation. J. Exp. Biol. 210, 2526–2539 (2007).

    Article  Google Scholar 

Download references

Acknowledgements

We would like to thank all families for participating in this study. We thank A. Torroni and A. Olivieri for their kind help with the control population samples from Sardinia. From the University of Miami, we are thankful to R. Cepeda for his excellent care and maintenance of the zebrafish facility, the transmission electron microscopy class of J. Prince, and to R. Duncan, W. Browne and P. Tsouflas for advice concerning phylogenic analysis. We are thankful to F. Poulain from the University of Utah for providing the Islet2-b fish and B. Schmid from the German Center for Neurodegenerative Diseases (DZNE) for donating the mitofish construct. This study was supported by US National Institutes of Health (NIH) grants R01NS075764, 5R01NS072248 and U54NS065712 (to S.Z.) and by the CMT Association and the Muscular Dystrophy Association. A.J.A. received additional support from the Lois Pope Life Foundation development award. This study was partially supported by the Cincinnati Children's Hospital Medical Center (CCHMC) Research Foundation and the Division of Human Genetics (to T.H.), US NIH grant R01DC012564 (to Z.M.A.) and ERMION, the E-RARE European consortium on optic atrophies (to V.C.). A.H.N. and S.M.D. were funded by the UK Oxford Partnership Comprehensive Biomedical Research Centre with funding from the UK Department of Health's National Institute for Health Research (NIHR) Biomedical Research Centres funding scheme. R.S. was funded by the European Union (E-RARE JTC grants 'NEUROLIPID' 01GM1408B and 'HSP/CMT genetics' PIOF-GA-2012-326681) and the Interdisciplinary Center for Clinical Research (IZKF) Tübingen (grant 1970-0-0). A.A. is supported by the National Institute of Neurological Disorders and Stroke, and L.B.G. is supported by US NIH Cellular and Molecular Biology Training Grant T32GM007315, US NIH Medical Scientist Training Grant T32GM07863 and US NIH grant F30 NRSA NS092238. A.B. is supported by US NIH grants RO1GM071775, GM105781 and GM112179. F.F. is supported by a Development grant from the American Heart Association.

Author information

Authors and Affiliations

Authors

Contributions

A.H.N., S.Z. and J.E.D. initiated the project, which was subsequently developed and led jointly by J.E.D., S.Z., T.H. and V.C. Experiments were conceived and carried out by A.J.A., A.R., C.Z., N.P., A.V.S., F.T., F.F., A.B., R.B.H., L.B.G., A.M. and R.S., with financial support from S.Z., T.H., A.A. and J.P. Zebrafish experiments were carried out by A.J.A., I.J.C. and S.G., with financial support from S.Z. and J.E.D. Patient recruitment, collection and analysis of human DNA, and genetic data analysis were carried out by M.A.G., F.S., L.A., S.M.D., L.C., C.L.M., R. Liguori, Z.M.A., K.L.S., X.W., L.A.K., Y.P., C.E.P., C.A.P., E.K.S., H.H.Z., O.A.A.-R. and Y.Y. Functional MRI-MRS data were generated by R. Lodi and Z.M.A. A.J.A. wrote the manuscript with critical input from S.Z. and J.E.D.

Corresponding authors

Correspondence to Taosheng Huang or Stephan Zuchner.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Sanger sequence traces and Clustal alignments of identified mutations.

(a) Sequence traces of the probands and one unaffected family member of all four pedigrees. Heterozygous mutations are marked by standard FASTA ambiguity code (R = A or G; Y = C or T; W = A or T). For frameshifts, inserted sequences are shown in red. (b) Clustal alignments show that all affected residues are conserved in vertebrates, with the exception of the residue affect by frameshift p.N296fs*297.

Supplementary Figure 2 Magnetic resonance imaging and spectroscopy (MRI-MRS) data.

(a) Midsagittal MRIs showing significant cerebellar atrophy in three of the probands. (b) MRS data indicating decreased NAA and increased lactate peaks in affected patients, typical of metabolic disorders.

Supplementary Figure 3 Maximum-likelihood phylogenetic tree containing all predicted mitochondrial carriers of H. sapiens, C. elegans, S. cerevisiae and S. pombe.

(a) Full tree drawn with 1,000 bootstraps. The red arrow indicates the node at which the SLC25A46 and Ugo1 orthologs bifurcate. (b) Collapsed tree depicting the node with calculated branch support values out of 1,000 and respective branch lengths (blue). SLC25A46 is orthologous to Y40B1B.8 in C. elegans, while Ugo1 is orthologous to Ugo1p in S. pombe; however, the branch support value between Ugo1 and SLC25A46 is under 700 and is therefore insufficient evidence for orthology between these proteins.

Supplementary Figure 4 Unrooted neighbor-joining trees of the mitochondrial solute carrier families of S. cerevisiae and H. sapiens with branch lengths drawn to scale.

Typical transporters are 30–40% identical to the family consensus sequence, while SLC25A46 and Ugo1 are both less than 20% identical to their respective families, as represented by longer branch lengths. Mitochondrial carrier homologs (MTCH1 and MTCH2) are also highly divergent.

Supplementary Figure 5 Colocalization patterns of additional outer and inner membrane proteins and endogenous mitofilin immunoprecipitation.

(a) Immunostaining patterns of SLC25A46, ANT2, mitofilin and TOM20 in COS-7 cells (scale bars, 1 µm). Linear profiles represent fluorescence intensity along the 2-µm white line in the merged image. Outer membrane proteins (SLC25A46 and TOM20) colocalize more with each other than with the inner membrane proteins (ANT2 and mitofilin). (b) Coimmunoprecipitation assay in mitochondrial extracts from HEK293T cells cotransfected with SLC25A46-HA. Immunoprecipitations were performed either by antibody to HA conjugated to Sepharose beads or protein G–conjugated seahorse beads plus a monoclonal antibody against mitofilin. The bound and unbound material were collected and probed by immunostaining with antibodies against HA, mitofilin and MFN2. M, total mitochondria.

Supplementary Figure 6 Validation of gene knockdown and assorted negative results.

(a) RT-PCR from COS-7 cells transfected with siRNA against SLC25A46 48 h post-transfection (h.p.t.). (b) Real-time PCR of HeLa cells transfected with siRNA 48 h.p.t. (c) RT-PCR used to validate morpholino (MO)-mediated knockdown in zebrafish at 48 h.p.f. The 508-bp amplicon was confirmed to be the result of exon 3 skipping, leading to a frameshift and protein truncation. (d) Whole-cell lysates from COS-7 cells show no significant changes in ATP levels. (e) Total mitochondrial DNA amount was not significantly different in HeLa cells. (f) ATP assay in whole zebrafish embryos at 24 h.p.f. shows no significant differences.

Supplementary Figure 7 Mitochondrial fusion assay by the extinction of mitochondrial-targeted photoactivated GFP (Mito PA-GFP).

Mito PA-GFP was activated in an ROI (white circle in the preactivated image), and images were acquired up to 90 min after activation. There was no significant difference between the extinction rates of control and knockdown cells by two-tailed t test, suggesting that there is neither an increase nor decrease in the mitochondrial fusion rate. Scale bars, 10 µm.

Supplementary Figure 8 The expression of slc25a46 in zebrafish retina and effects of morpholino knockdown.

(ad) RNA In situ hybridization of zebrafish larvae at 72 h.p.f. (a) Control, sense probe shows background staining. (bd) The antisense probe shows that the slc25a46 transcript is highly enriched in the neuronal processes of retinal ganglion cells (RGCs), including those of the optic nerve (ON), optic tectum (OT) and inner plexiform layer (IPL). The slc25a46 transcript is also enriched in a portion of the inner nuclear layer (INL). (e) Retinal ganglion cell staining using antibody to ZN5 with DAB in embryos injected with 0.45 pmol of either scrambled control MO or slc25a46 MO at 96 h.p.f. (f) Quantification of the staining intensity of the inner plexiform layer (IPL) suggests that fewer dendrites are present in the knockdown. The 8-bit gray scale numbers 0 = white and 256 = black were converted to percentages whereby 0 = white and 100 = black. One-tailed Student’s t test was used to determine significance. Control MO, n = 6; 0.45 pmol slc25a46 MO, n = 6. Scale bars, 50 μm in ae. (g) Representative images of the gross morphology at 48 h.p.f. Scale bar, 1 mm. (h) Quantification of morphology: the curly phenotype was predominantly observed at the 0.45-pmol dosage. 0.45 pmol of control MO, n = 220; 0.45 pmol of slc25a46 MO, n = 223; 0.3 pmol of slc25a46 MO, n = 181. (i) Swim traces from three different larvae illustrating the angle of body curvature at 96 h.p.f.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–8, Supplementary Tables 1–5 and Supplementary Note. (PDF 2271 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Abrams, A., Hufnagel, R., Rebelo, A. et al. Mutations in SLC25A46, encoding a UGO1-like protein, cause an optic atrophy spectrum disorder. Nat Genet 47, 926–932 (2015). https://doi.org/10.1038/ng.3354

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ng.3354

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing