La maladie de Parkinson en France (serveur d'exploration)

Attention, ce site est en cours de développement !
Attention, site généré par des moyens informatiques à partir de corpus bruts.
Les informations ne sont donc pas validées.

Loss of functional OPA1 unbalances redox state: implications in dominant optic atrophy pathogenesis

Identifieur interne : 000687 ( Pmc/Corpus ); précédent : 000686; suivant : 000688

Loss of functional OPA1 unbalances redox state: implications in dominant optic atrophy pathogenesis

Auteurs : Aurélie M. C. Millet ; Ambre M. Bertholet ; Marlène Daloyau ; Pascal Reynier ; Anne Galinier ; Anne Devin ; Bernd Wissinguer ; Pascale Belenguer ; Noélie Davezac

Source :

RBID : PMC:4891995

Abstract

AbstractObjective

OPA1 mutations cause protein haploinsufficiency leading to dominant optic atrophy (DOA), an incurable retinopathy with variable severity. Up to 20% of patients also develop extraocular neurological complications. The mechanisms that cause this optic atrophy or its syndromic forms are still unknown. After identifying oxidative stress in a mouse model of the pathology, we sought to determine the consequences of OPA1 dysfunction on redox homeostasis.

Methods

Mitochondrial respiration, reactive oxygen species levels, antioxidant defenses, and cell death were characterized by biochemical and in situ approaches in both in vitro and in vivo models of OPA1 haploinsufficiency.

Results

A decrease in aconitase activity suggesting an increase in reactive oxygene species and an induction of antioxidant defenses was observed in cortices of a murine model as well as in OPA1 downregulated cortical neurons. This increase is associated with a decline in mitochondrial respiration in vitro. Upon exogenous oxidative stress, OPA1‐depleted neurons did not further exhibit upregulated antioxidant defenses but were more sensitive to cell death. Finally, low levels of antioxidant enzymes were found in fibroblasts from patients supporting their role as modifier factors.

Interpretation

Our study suggests that the pro‐oxidative state induced by OPA1 loss may contribute to DOA pathogenesis and that differences in antioxidant defenses can explain the variability in expressivity. Furthermore, antioxidants may be used as therapy as they could prevent or delay DOA symptoms in patients.


Url:
DOI: 10.1002/acn3.305
PubMed: 27547769
PubMed Central: 4891995

Links to Exploration step

PMC:4891995

Le document en format XML

<record>
<TEI>
<teiHeader>
<fileDesc>
<titleStmt>
<title xml:lang="en">Loss of functional
<styled-content style="fixed-case">OPA</styled-content>
1 unbalances redox state: implications in dominant optic atrophy pathogenesis</title>
<author>
<name sortKey="Millet, Aurelie M C" sort="Millet, Aurelie M C" uniqKey="Millet A" first="Aurélie M. C." last="Millet">Aurélie M. C. Millet</name>
<affiliation>
<nlm:aff id="acn3305-aff-0001"></nlm:aff>
</affiliation>
</author>
<author>
<name sortKey="Bertholet, Ambre M" sort="Bertholet, Ambre M" uniqKey="Bertholet A" first="Ambre M." last="Bertholet">Ambre M. Bertholet</name>
<affiliation>
<nlm:aff id="acn3305-aff-0001"></nlm:aff>
</affiliation>
</author>
<author>
<name sortKey="Daloyau, Marlene" sort="Daloyau, Marlene" uniqKey="Daloyau M" first="Marlène" last="Daloyau">Marlène Daloyau</name>
<affiliation>
<nlm:aff id="acn3305-aff-0001"></nlm:aff>
</affiliation>
</author>
<author>
<name sortKey="Reynier, Pascal" sort="Reynier, Pascal" uniqKey="Reynier P" first="Pascal" last="Reynier">Pascal Reynier</name>
<affiliation>
<nlm:aff id="acn3305-aff-0002"></nlm:aff>
</affiliation>
</author>
<author>
<name sortKey="Galinier, Anne" sort="Galinier, Anne" uniqKey="Galinier A" first="Anne" last="Galinier">Anne Galinier</name>
<affiliation>
<nlm:aff id="acn3305-aff-0003"></nlm:aff>
</affiliation>
</author>
<author>
<name sortKey="Devin, Anne" sort="Devin, Anne" uniqKey="Devin A" first="Anne" last="Devin">Anne Devin</name>
<affiliation>
<nlm:aff id="acn3305-aff-0004"></nlm:aff>
</affiliation>
</author>
<author>
<name sortKey="Wissinguer, Bernd" sort="Wissinguer, Bernd" uniqKey="Wissinguer B" first="Bernd" last="Wissinguer">Bernd Wissinguer</name>
<affiliation>
<nlm:aff id="acn3305-aff-0005"></nlm:aff>
</affiliation>
</author>
<author>
<name sortKey="Belenguer, Pascale" sort="Belenguer, Pascale" uniqKey="Belenguer P" first="Pascale" last="Belenguer">Pascale Belenguer</name>
<affiliation>
<nlm:aff id="acn3305-aff-0001"></nlm:aff>
</affiliation>
</author>
<author>
<name sortKey="Davezac, Noelie" sort="Davezac, Noelie" uniqKey="Davezac N" first="Noélie" last="Davezac">Noélie Davezac</name>
<affiliation>
<nlm:aff id="acn3305-aff-0001"></nlm:aff>
</affiliation>
</author>
</titleStmt>
<publicationStmt>
<idno type="wicri:source">PMC</idno>
<idno type="pmid">27547769</idno>
<idno type="pmc">4891995</idno>
<idno type="url">http://www.ncbi.nlm.nih.gov/pmc/articles/PMC4891995</idno>
<idno type="RBID">PMC:4891995</idno>
<idno type="doi">10.1002/acn3.305</idno>
<date when="2016">2016</date>
<idno type="wicri:Area/Pmc/Corpus">000687</idno>
<idno type="wicri:explorRef" wicri:stream="Pmc" wicri:step="Corpus" wicri:corpus="PMC">000687</idno>
</publicationStmt>
<sourceDesc>
<biblStruct>
<analytic>
<title xml:lang="en" level="a" type="main">Loss of functional
<styled-content style="fixed-case">OPA</styled-content>
1 unbalances redox state: implications in dominant optic atrophy pathogenesis</title>
<author>
<name sortKey="Millet, Aurelie M C" sort="Millet, Aurelie M C" uniqKey="Millet A" first="Aurélie M. C." last="Millet">Aurélie M. C. Millet</name>
<affiliation>
<nlm:aff id="acn3305-aff-0001"></nlm:aff>
</affiliation>
</author>
<author>
<name sortKey="Bertholet, Ambre M" sort="Bertholet, Ambre M" uniqKey="Bertholet A" first="Ambre M." last="Bertholet">Ambre M. Bertholet</name>
<affiliation>
<nlm:aff id="acn3305-aff-0001"></nlm:aff>
</affiliation>
</author>
<author>
<name sortKey="Daloyau, Marlene" sort="Daloyau, Marlene" uniqKey="Daloyau M" first="Marlène" last="Daloyau">Marlène Daloyau</name>
<affiliation>
<nlm:aff id="acn3305-aff-0001"></nlm:aff>
</affiliation>
</author>
<author>
<name sortKey="Reynier, Pascal" sort="Reynier, Pascal" uniqKey="Reynier P" first="Pascal" last="Reynier">Pascal Reynier</name>
<affiliation>
<nlm:aff id="acn3305-aff-0002"></nlm:aff>
</affiliation>
</author>
<author>
<name sortKey="Galinier, Anne" sort="Galinier, Anne" uniqKey="Galinier A" first="Anne" last="Galinier">Anne Galinier</name>
<affiliation>
<nlm:aff id="acn3305-aff-0003"></nlm:aff>
</affiliation>
</author>
<author>
<name sortKey="Devin, Anne" sort="Devin, Anne" uniqKey="Devin A" first="Anne" last="Devin">Anne Devin</name>
<affiliation>
<nlm:aff id="acn3305-aff-0004"></nlm:aff>
</affiliation>
</author>
<author>
<name sortKey="Wissinguer, Bernd" sort="Wissinguer, Bernd" uniqKey="Wissinguer B" first="Bernd" last="Wissinguer">Bernd Wissinguer</name>
<affiliation>
<nlm:aff id="acn3305-aff-0005"></nlm:aff>
</affiliation>
</author>
<author>
<name sortKey="Belenguer, Pascale" sort="Belenguer, Pascale" uniqKey="Belenguer P" first="Pascale" last="Belenguer">Pascale Belenguer</name>
<affiliation>
<nlm:aff id="acn3305-aff-0001"></nlm:aff>
</affiliation>
</author>
<author>
<name sortKey="Davezac, Noelie" sort="Davezac, Noelie" uniqKey="Davezac N" first="Noélie" last="Davezac">Noélie Davezac</name>
<affiliation>
<nlm:aff id="acn3305-aff-0001"></nlm:aff>
</affiliation>
</author>
</analytic>
<series>
<title level="j">Annals of Clinical and Translational Neurology</title>
<idno type="eISSN">2328-9503</idno>
<imprint>
<date when="2016">2016</date>
</imprint>
</series>
</biblStruct>
</sourceDesc>
</fileDesc>
<profileDesc>
<textClass></textClass>
</profileDesc>
</teiHeader>
<front>
<div type="abstract" xml:lang="en">
<title>Abstract</title>
<sec id="acn3305-sec-0001">
<title>Objective</title>
<p>
<italic>
<styled-content style="fixed-case">OPA</styled-content>
1</italic>
mutations cause protein haploinsufficiency leading to dominant optic atrophy (
<styled-content style="fixed-case">DOA</styled-content>
), an incurable retinopathy with variable severity. Up to 20% of patients also develop extraocular neurological complications. The mechanisms that cause this optic atrophy or its syndromic forms are still unknown. After identifying oxidative stress in a mouse model of the pathology, we sought to determine the consequences of
<styled-content style="fixed-case">OPA</styled-content>
1 dysfunction on redox homeostasis.</p>
</sec>
<sec id="acn3305-sec-0002">
<title>Methods</title>
<p>Mitochondrial respiration, reactive oxygen species levels, antioxidant defenses, and cell death were characterized by biochemical and in situ approaches in both in vitro and in vivo models of
<styled-content style="fixed-case">OPA</styled-content>
1 haploinsufficiency.</p>
</sec>
<sec id="acn3305-sec-0003">
<title>Results</title>
<p>A decrease in aconitase activity suggesting an increase in reactive oxygene species and an induction of antioxidant defenses was observed in cortices of a murine model as well as in
<styled-content style="fixed-case">OPA</styled-content>
1 downregulated cortical neurons. This increase is associated with a decline in mitochondrial respiration in vitro. Upon exogenous oxidative stress,
<styled-content style="fixed-case">OPA</styled-content>
1‐depleted neurons did not further exhibit upregulated antioxidant defenses but were more sensitive to cell death. Finally, low levels of antioxidant enzymes were found in fibroblasts from patients supporting their role as modifier factors.</p>
</sec>
<sec id="acn3305-sec-0004">
<title>Interpretation</title>
<p>Our study suggests that the pro‐oxidative state induced by
<styled-content style="fixed-case">OPA</styled-content>
1 loss may contribute to
<styled-content style="fixed-case">DOA</styled-content>
pathogenesis and that differences in antioxidant defenses can explain the variability in expressivity. Furthermore, antioxidants may be used as therapy as they could prevent or delay DOA symptoms in patients.</p>
</sec>
</div>
</front>
<back>
<div1 type="bibliography">
<listBibl>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
<biblStruct></biblStruct>
</listBibl>
</div1>
</back>
</TEI>
<pmc article-type="research-article">
<pmc-dir>properties open_access</pmc-dir>
<front>
<journal-meta>
<journal-id journal-id-type="nlm-ta">Ann Clin Transl Neurol</journal-id>
<journal-id journal-id-type="iso-abbrev">Ann Clin Transl Neurol</journal-id>
<journal-id journal-id-type="doi">10.1002/(ISSN)2328-9503</journal-id>
<journal-id journal-id-type="publisher-id">ACN3</journal-id>
<journal-title-group>
<journal-title>Annals of Clinical and Translational Neurology</journal-title>
</journal-title-group>
<issn pub-type="epub">2328-9503</issn>
<publisher>
<publisher-name>John Wiley and Sons Inc.</publisher-name>
<publisher-loc>Hoboken</publisher-loc>
</publisher>
</journal-meta>
<article-meta>
<article-id pub-id-type="pmid">27547769</article-id>
<article-id pub-id-type="pmc">4891995</article-id>
<article-id pub-id-type="doi">10.1002/acn3.305</article-id>
<article-id pub-id-type="publisher-id">ACN3305</article-id>
<article-categories>
<subj-group subj-group-type="overline">
<subject>Research Article</subject>
</subj-group>
<subj-group subj-group-type="heading">
<subject>Research Articles</subject>
</subj-group>
</article-categories>
<title-group>
<article-title>Loss of functional
<styled-content style="fixed-case">OPA</styled-content>
1 unbalances redox state: implications in dominant optic atrophy pathogenesis</article-title>
<alt-title alt-title-type="left-running-head">A. M. C. Millet
<italic>et al</italic>
.</alt-title>
</title-group>
<contrib-group>
<contrib id="acn3305-cr-0001" contrib-type="author">
<name>
<surname>Millet</surname>
<given-names>Aurélie M. C.</given-names>
</name>
<xref ref-type="aff" rid="acn3305-aff-0001">
<sup>1</sup>
</xref>
</contrib>
<contrib id="acn3305-cr-0002" contrib-type="author">
<name>
<surname>Bertholet</surname>
<given-names>Ambre M.</given-names>
</name>
<xref ref-type="aff" rid="acn3305-aff-0001">
<sup>1</sup>
</xref>
</contrib>
<contrib id="acn3305-cr-0003" contrib-type="author">
<name>
<surname>Daloyau</surname>
<given-names>Marlène</given-names>
</name>
<xref ref-type="aff" rid="acn3305-aff-0001">
<sup>1</sup>
</xref>
</contrib>
<contrib id="acn3305-cr-0004" contrib-type="author">
<name>
<surname>Reynier</surname>
<given-names>Pascal</given-names>
</name>
<xref ref-type="aff" rid="acn3305-aff-0002">
<sup>2</sup>
</xref>
</contrib>
<contrib id="acn3305-cr-0005" contrib-type="author">
<name>
<surname>Galinier</surname>
<given-names>Anne</given-names>
</name>
<xref ref-type="aff" rid="acn3305-aff-0003">
<sup>3</sup>
</xref>
</contrib>
<contrib id="acn3305-cr-0006" contrib-type="author">
<name>
<surname>Devin</surname>
<given-names>Anne</given-names>
</name>
<xref ref-type="aff" rid="acn3305-aff-0004">
<sup>4</sup>
</xref>
</contrib>
<contrib id="acn3305-cr-0007" contrib-type="author">
<name>
<surname>Wissinguer</surname>
<given-names>Bernd</given-names>
</name>
<xref ref-type="aff" rid="acn3305-aff-0005">
<sup>5</sup>
</xref>
</contrib>
<contrib id="acn3305-cr-0008" contrib-type="author">
<name>
<surname>Belenguer</surname>
<given-names>Pascale</given-names>
</name>
<xref ref-type="aff" rid="acn3305-aff-0001">
<sup>1</sup>
</xref>
<xref ref-type="author-notes" rid="acn3305-note-0001">
<sup></sup>
</xref>
</contrib>
<contrib id="acn3305-cr-0009" contrib-type="author" corresp="yes">
<name>
<surname>Davezac</surname>
<given-names>Noélie</given-names>
</name>
<xref ref-type="aff" rid="acn3305-aff-0001">
<sup>1</sup>
</xref>
<xref ref-type="author-notes" rid="acn3305-note-0001">
<sup></sup>
</xref>
</contrib>
</contrib-group>
<aff id="acn3305-aff-0001">
<label>
<sup>1</sup>
</label>
<named-content content-type="organisation-division">Center of Developmental Biology (CBD)/Research Center on Animal Cognition (CRCA)</named-content>
<named-content content-type="organisation-division">Center for Integrative Biology (CBI)</named-content>
<institution>Toulouse University, CNRS, UPS</institution>
<named-content content-type="city">Toulouse</named-content>
<country country="FR">France</country>
</aff>
<aff id="acn3305-aff-0002">
<label>
<sup>2</sup>
</label>
<institution>CNRS UMR 6214 Inserm UMR 1083 UFR Sciences médicales</institution>
<named-content content-type="street">Rue Haute de Reculee</named-content>
<named-content content-type="city">Angers Cedex 01</named-content>
<named-content content-type="post-code">49045</named-content>
<country country="FR">France</country>
</aff>
<aff id="acn3305-aff-0003">
<label>
<sup>3</sup>
</label>
<named-content content-type="organisation-division">Laboratoire de Biochimie Nutritionnelle “STROMALab”</named-content>
<institution>UMR UPS/CNRS/EFS 5273 Inserm U1031, CHU Rangueil</institution>
<named-content content-type="street">1 avenue Jean Poulhès</named-content>
<named-content content-type="city">Toulouse Cedex 9</named-content>
<named-content content-type="post-code">31059</named-content>
<country country="FR">France</country>
</aff>
<aff id="acn3305-aff-0004">
<label>
<sup>4</sup>
</label>
<named-content content-type="organisation-division">Laboratoire métabolisme énergétique cellulaire</named-content>
<institution>IBGC du CNRS</institution>
<named-content content-type="street">1 rue Camille Saint Saëns</named-content>
<named-content content-type="city">Bordeaux Cedex</named-content>
<named-content content-type="post-code">33077</named-content>
<country country="FR">France</country>
</aff>
<aff id="acn3305-aff-0005">
<label>
<sup>5</sup>
</label>
<named-content content-type="organisation-division">Centre for Ophthalmology</named-content>
<institution>University of Tübingen</institution>
<named-content content-type="street">Roentgenweg 11</named-content>
<named-content content-type="city">Tübingen</named-content>
<named-content content-type="post-code">D‐72076</named-content>
<country country="DE">Germany</country>
</aff>
<author-notes>
<corresp id="correspondenceTo">
<label>*</label>
<bold>Correspondence</bold>
<break></break>
Noélie Davezac, UMR 5547 Université Toulouse 3, 118 route de Narbonne, 31062 Toulouse Cedex 09, France. Tel: 0033(0)5 61 55 65 76; Fax: 0033(0)5 61 55 88 94; E‐mail:
<email>noelie.davezac@univ-tlse3.fr</email>
<break></break>
</corresp>
<fn id="acn3305-note-0001">
<label></label>
<p>Share senior authorship.</p>
</fn>
</author-notes>
<pub-date pub-type="epub">
<day>25</day>
<month>4</month>
<year>2016</year>
</pub-date>
<pub-date pub-type="collection">
<month>6</month>
<year>2016</year>
</pub-date>
<volume>3</volume>
<issue>6</issue>
<issue-id pub-id-type="doi">10.1111/acn3.2016.3.issue-6</issue-id>
<fpage>408</fpage>
<lpage>421</lpage>
<history>
<date date-type="received">
<day>06</day>
<month>10</month>
<year>2015</year>
</date>
<date date-type="rev-recd">
<day>04</day>
<month>3</month>
<year>2016</year>
</date>
<date date-type="accepted">
<day>05</day>
<month>3</month>
<year>2016</year>
</date>
</history>
<permissions>
<pmc-comment> © 2016 American Neurological Association </pmc-comment>
<copyright-statement content-type="article-copyright">© 2016 The Authors.
<italic>Annals of Clinical and Translational Neurology</italic>
published by Wiley Periodicals, Inc on behalf of American Neurological Association.</copyright-statement>
<license license-type="creativeCommonsBy-nc-nd">
<license-p>This is an open access article under the terms of the
<ext-link ext-link-type="uri" xlink:href="http://creativecommons.org/licenses/by-nc-nd/4.0/">Creative Commons Attribution‐NonCommercial‐NoDerivs</ext-link>
License, which permits use and distribution in any medium, provided the original work is properly cited, the use is non‐commercial and no modifications or adaptations are made.</license-p>
</license>
</permissions>
<self-uri content-type="pdf" xlink:type="simple" xlink:href="file:ACN3-3-408.pdf"></self-uri>
<abstract id="acn3305-abs-0001">
<title>Abstract</title>
<sec id="acn3305-sec-0001">
<title>Objective</title>
<p>
<italic>
<styled-content style="fixed-case">OPA</styled-content>
1</italic>
mutations cause protein haploinsufficiency leading to dominant optic atrophy (
<styled-content style="fixed-case">DOA</styled-content>
), an incurable retinopathy with variable severity. Up to 20% of patients also develop extraocular neurological complications. The mechanisms that cause this optic atrophy or its syndromic forms are still unknown. After identifying oxidative stress in a mouse model of the pathology, we sought to determine the consequences of
<styled-content style="fixed-case">OPA</styled-content>
1 dysfunction on redox homeostasis.</p>
</sec>
<sec id="acn3305-sec-0002">
<title>Methods</title>
<p>Mitochondrial respiration, reactive oxygen species levels, antioxidant defenses, and cell death were characterized by biochemical and in situ approaches in both in vitro and in vivo models of
<styled-content style="fixed-case">OPA</styled-content>
1 haploinsufficiency.</p>
</sec>
<sec id="acn3305-sec-0003">
<title>Results</title>
<p>A decrease in aconitase activity suggesting an increase in reactive oxygene species and an induction of antioxidant defenses was observed in cortices of a murine model as well as in
<styled-content style="fixed-case">OPA</styled-content>
1 downregulated cortical neurons. This increase is associated with a decline in mitochondrial respiration in vitro. Upon exogenous oxidative stress,
<styled-content style="fixed-case">OPA</styled-content>
1‐depleted neurons did not further exhibit upregulated antioxidant defenses but were more sensitive to cell death. Finally, low levels of antioxidant enzymes were found in fibroblasts from patients supporting their role as modifier factors.</p>
</sec>
<sec id="acn3305-sec-0004">
<title>Interpretation</title>
<p>Our study suggests that the pro‐oxidative state induced by
<styled-content style="fixed-case">OPA</styled-content>
1 loss may contribute to
<styled-content style="fixed-case">DOA</styled-content>
pathogenesis and that differences in antioxidant defenses can explain the variability in expressivity. Furthermore, antioxidants may be used as therapy as they could prevent or delay DOA symptoms in patients.</p>
</sec>
</abstract>
<funding-group>
<award-group>
<funding-source>Centre National de la Recherche Scientifique</funding-source>
</award-group>
<award-group>
<funding-source>Université Paul Sabatier</funding-source>
</award-group>
<award-group>
<funding-source>Rétina‐France</funding-source>
</award-group>
<award-group>
<funding-source>Union Nationale Des Aveugles et Déficients Visuels</funding-source>
</award-group>
<award-group>
<funding-source>Gueules Cassées Sourire quand même</funding-source>
</award-group>
<award-group>
<funding-source>Association contre les Maladies Mitochondriales</funding-source>
</award-group>
<award-group>
<funding-source>French Ministry for Research and Education</funding-source>
</award-group>
</funding-group>
<counts>
<page-count count="14"></page-count>
</counts>
<custom-meta-group>
<custom-meta>
<meta-name>source-schema-version-number</meta-name>
<meta-value>2.0</meta-value>
</custom-meta>
<custom-meta>
<meta-name>component-id</meta-name>
<meta-value>acn3305</meta-value>
</custom-meta>
<custom-meta>
<meta-name>cover-date</meta-name>
<meta-value>June 2016</meta-value>
</custom-meta>
<custom-meta>
<meta-name>details-of-publishers-convertor</meta-name>
<meta-value>Converter:WILEY_ML3GV2_TO_NLMPMC version:4.9.4 mode:remove_FC converted:11.08.2016</meta-value>
</custom-meta>
</custom-meta-group>
</article-meta>
</front>
<body>
<sec id="acn3305-sec-0005">
<title>Introduction</title>
<p>Dominant optic atrophy (DOA) is characterized by moderate to severe loss of visual acuity with insidious onset in early childhood. DOA penetrance may be as low as 40% and prevalence is 1:50,000 worldwide.
<xref rid="acn3305-bib-0001" ref-type="ref">1</xref>
Up to now, there is no effective treatment for this complex pathology. The majority of DOA patients (~75%) harbor mutation in the
<italic>OPA1</italic>
gene coding for a mitochondrial GTPase.
<xref rid="acn3305-bib-0002" ref-type="ref">2</xref>
There are 280 different
<italic>OPA1</italic>
mutations (http:/mitodyn.org), the majority of which result in premature termination ensuing OPA1 haploinsufficiency.
<xref rid="acn3305-bib-0001" ref-type="ref">1</xref>
There is marked inter‐ and intrafamilial variability in the rate of disease progression, and recent studies show a severe multisystemic disorder associated with some
<italic>OPA1</italic>
mutations (DOA
<sup>+</sup>
syndrome).
<xref rid="acn3305-bib-0003" ref-type="ref">3</xref>
,
<xref rid="acn3305-bib-0004" ref-type="ref">4</xref>
,
<xref rid="acn3305-bib-0005" ref-type="ref">5</xref>
,
<xref rid="acn3305-bib-0006" ref-type="ref">6</xref>
These patients present additional neurological complications, such as ataxia, sensorineural deafness, sensory motor neuropathy as well as chronic progressive external ophtalmoplegia and myopathy. Moreover, a recent case described two DOA families which also presented parkinsonism.
<xref rid="acn3305-bib-0007" ref-type="ref">7</xref>
Altogether these recent findings are of major pathophysiological importance, and highlight the widespread deleterious consequences of OPA1 mutations, not only for retinal ganglionic cells (RGCs) but also for other neuronal populations.
<xref rid="acn3305-bib-0006" ref-type="ref">6</xref>
,
<xref rid="acn3305-bib-0008" ref-type="ref">8</xref>
,
<xref rid="acn3305-bib-0009" ref-type="ref">9</xref>
</p>
<p>The
<italic>OPA1</italic>
gene encodes a mitochondrial inner membrane protein facing the intermembrane space
<xref rid="acn3305-bib-0010" ref-type="ref">10</xref>
with various functions including inner membrane fusion, cristae structuration, mitochondrial DNA maintenance, mitochondrial energetic modulation, and protection from apoptosis.
<xref rid="acn3305-bib-0010" ref-type="ref">10</xref>
Patients' skin fibroblasts, muscles, or lymphoblasts revealed impairments in mitochondrial morphology,
<xref rid="acn3305-bib-0011" ref-type="ref">11</xref>
,
<xref rid="acn3305-bib-0012" ref-type="ref">12</xref>
,
<xref rid="acn3305-bib-0013" ref-type="ref">13</xref>
respiration, and energetics, and also displayed a loss of mitochondrial DNA integrity and an increase in sensitivity to apoptosis.
<xref rid="acn3305-bib-0008" ref-type="ref">8</xref>
,
<xref rid="acn3305-bib-0014" ref-type="ref">14</xref>
,
<xref rid="acn3305-bib-0015" ref-type="ref">15</xref>
However, it is still not clear if DOA patients do or do not develop energetic defects (see for review refs.
<xref rid="acn3305-bib-0010" ref-type="ref">10</xref>
,
<xref rid="acn3305-bib-0016" ref-type="ref">16</xref>
). Mitochondria from DOA or DOA
<sup>+</sup>
mice models present an increase in mitochondrial fragmentation, cristae disorganization, mitophagy, and cytochrome c oxidase deficiency. In
<italic>Caenorhabditis elegans</italic>
, mutations of the
<italic>OPA1</italic>
gene ortholog,
<italic>eat‐3</italic>
, cause mitochondrial fragmentation and hypersensitivity to oxidative stress, but do not lead to cell death.
<xref rid="acn3305-bib-0017" ref-type="ref">17</xref>
Heterozygous mutations of OPA1 in
<italic>Drosophila melanogaster</italic>
result in a shortened life span, increased reactive oxygen species (ROS) production, sensitivity to oxidative stress, and defects in the activity of respiratory chain complexes.
<xref rid="acn3305-bib-0018" ref-type="ref">18</xref>
,
<xref rid="acn3305-bib-0019" ref-type="ref">19</xref>
These two DOA invertebrate models link the critical generation of ROS with OPA1 dysfunction. We sought to investigate if ROS production promotes DOA pathogenesis in mammals using various complementary models such as (1) a DOA mouse model,
<xref rid="acn3305-bib-0020" ref-type="ref">20</xref>
(2) a canonical
<italic>ex vivo</italic>
neuronal model of OPA1 haploinsufficiency,
<xref rid="acn3305-bib-0021" ref-type="ref">21</xref>
and (3) fibroblasts from DOA and DOA
<sup>+</sup>
patients.
<xref rid="acn3305-bib-0022" ref-type="ref">22</xref>
</p>
<p>Our results show that OPA1 haploinsufficiency lead to a decrease in mitochondrial respiration, accompanied by an increase in mitochondrial ROS production. This associated increase is linked to a decrease in aconitase activity and an increase in antioxidant defenses, in both primary cultured neurons and in mice. ROS production is countered by the activation of the nuclear factor (erythroid‐derived 2)‐like 2 (NRF2) transcription factor and its downstream targets. Loss of functional OPA1 imbalanced the redox state, because additional exogenous oxidative stress challenged both the antioxidant response and the viability of OPA1‐depleted neurons. In addition, we found altered expression of antioxidant genes at the protein level in fibroblasts from patients with DOA. On the basis of these new findings, we propose a model in which reduced functional OPA1 leads to an imbalance in cellular redox state, which sensitizes cells to exogenous pro‐oxidative stresses. We believe this phenomenon is a molecular mechanism underlying DOA pathogenesis.</p>
</sec>
<sec id="acn3305-sec-0006">
<title>Materials and Methods</title>
<sec id="acn3305-sec-0007">
<title>Cell culture</title>
<sec id="acn3305-sec-0008">
<title>Wistar rats</title>
<p>All animals (
<italic>n</italic>
= 45 pregnant rats, 350 embryos) in this study were ethically maintained and used as ethical laws of CNRS (Centre National de la Recherche Scientifique) and FRBT (Fédération de Recherche en Biologie de Toulouse). The animal facility called Genotoul Anexpo is maintained in accordance to F.E.L.A.S.A organism (Federation of European Laboratory Animal Science Associations). Animals were housed in groups of six and kept in a specific pathogen‐free and temperature‐controlled facility on a 12‐h light/dark cycle. Pregnant Wistar rats are delivered by Janvier company 48 h before being sacrificed and were handled in a calm room. Day 17 embryos were removed from pregnant Wistar rats (Janvier, France) under intraperitoneal pentobarbitol (Sigma‐Aldrich, UK) anesthesia. The rats were then immediately sacrificed. Cortices were dissected, enzymatically dissociated with papain (10 U/mL, Sigma‐Aldrich, UK), then exposed for 5 min in a solution that inactivated papain: DNAse I (Invitrogen, Thermo Fisher Scientific, Waltham, USA) and B27 (Gibco, Thermo Fisher Scientific, Waltham, USA), diluted in Phosphate Buffer Saline (PBS) 1X with
<sc>d</sc>
‐glucose (33 mmol/L, Sigma‐Aldrich, UK). Cells were dissociated by trituration and filtered through a membrane (70
<italic>μ</italic>
m, BD Falcon). Cells were then purified through a BSA solution (8%, Sigma‐Aldrich, UK) diluted in Neurobasal A‐25 (Invitrogen, Thermo Fisher Scientific, Waltham, USA). Dishes, with or without glass coverslips, were coated with poly‐
<sc>d</sc>
‐lysine (0.1 mg/mL, Sigma‐Aldrich, UK) 24 h prior to culturing. For each experiment, cortices from 8 to 12 embryos per rat are mixed. Experiments were reproduced three to eight times. Cultures were grown in Neurobasal
<sup>®</sup>
(Eurobio) supplemented with B27 (Invitrogen, Thermo Fisher Scientific, Waltham, USA), 2 mmol/L glutamine, 0.1% penicillin and streptomycin (Gibco, Thermo Fisher Scientific, Waltham, USA), 250 U/mL amphotericin (Invitrogen, Thermo Fisher Scientific, Waltham, USA), and 1 mmol/L lactic acid (Sigma‐Aldrich, UK) at a density of 6 × 10
<sup>5</sup>
cells per cm
<sup>2</sup>
.</p>
<p>Cortical neurons (1 × 10
<sup>6</sup>
) were electroporated using the Rat Neuron Nucleofector Kit (Amaxa, Lonza) according to the manufacturers' optimized protocol with 3
<italic>μ</italic>
g of control luciferase‐targeting (siCtrl, D‐001210‐02, Dharmacon GE, UK) or OPA1‐targeting (siOPA1, target sequence GAUUGUGCCUGACUUUAUA, Dharmacon GE, UK) small interfering RNA (Dharmacon GE, UK).</p>
</sec>
<sec id="acn3305-sec-0009">
<title>Patients' fibroblasts</title>
<p>Fibroblasts obtained by skin biopsy from consenting DOA patients and healthy volunteers were cultured in Dulbecco's modified Eagle's medium 4.5 g/L glucose (DMEM, Invitrogen, Thermo Fisher Scientific, Waltham, USA), supplemented with 10% Fetal Bovine Serum (FBS). penicillin (100 U/mL) and streptomycin (100 mg/mL), and maintained for up to 20 passages. The use of patient fibroblasts for research purpose in neurogenetics was approved by the “Comité de Protection des Personnes Ouest II – Angers” (No. CB 2014/02).</p>
</sec>
</sec>
<sec id="acn3305-sec-0010">
<title>
<italic>Opa1</italic>
<sup>
<italic>+/−</italic>
</sup>
mice</title>
<p>Mice were kept in a 12‐h light/12‐h dark cycle with food and water available ad libitum in full‐barrier facilities free of specific pathogens. Mouse breeding and all experimental procedures were performed according to ethical laws of CNRS (Centre National de la Recherche Scientifique) and Fédération de Recherche en Biologie de Toulouse (FRBT). The animal facility called Genotoul Anexpo is maintained in accordance to F.E.L.A.S.A organism (Federation of European Laboratory Animal Science Associations). Two sanitary controls are effectuated per year following the set point CO66. The B6;C3‐
<italic>Opa1</italic>
<sup>
<italic>329‐355del</italic>
</sup>
mouse strain (abbreviated
<italic>Opa1</italic>
<sup>
<italic>+/−</italic>
</sup>
mouse) has been previously described by Alavi and colleagues.
<xref rid="acn3305-bib-0013" ref-type="ref">13</xref>
,
<xref rid="acn3305-bib-0020" ref-type="ref">20</xref>
,
<xref rid="acn3305-bib-0023" ref-type="ref">23</xref>
Wild‐type (
<italic>Opa1</italic>
<sup>
<italic>+/+</italic>
</sup>
) littermates were systematically used as control. All examined mice were sacrificed either at 4 or 10 months old. After sacrifice, adult brains cortices were isolated and separated from hippocampus and cerebellum area. Each of those areas has been snap frozen in liquid nitrogen and stored at −80°C. Cortices were thawed on ice and 20 mg of tissue as been isolated and processed for immunoblot experiments.</p>
</sec>
<sec id="acn3305-sec-0011">
<title>Measurement of oxygen consumption and ATP/ADP levels</title>
<p>Oxygen consumption rates (OCR) were performed using the XF24 Extracellular Flux Analyser (Seahorse Bioscience, North Billerica, MA). Neurons (3 × 10
<sup>5</sup>
) transfected with siCtrl or siOPA1 were plated on XF24 microplates 6 days before OCR measurements. Dual‐analyte sensor cartridges were soaked in XF Calibrant Solution (Seahorse Biosciences) in 24‐wells cell culture microplates overnight at 37°C to hydrate the probes. One hour prior to experimentation, injection ports on the sensor cartridge were filled with oligomycin (0.6
<italic>μ</italic>
mol/L), carbonyl cyanide 4‐(trifluoromethoxy)phenylhydrazone (FCCP) (6
<italic>μ</italic>
mol/L), and rotenone (50 nmol/L) plus antimycin A (0.182
<italic>μ</italic>
mol/L). Plates were then loaded into the XF24 instrument for calibration. For oxygen consumption measurement, growth media of neurons were replaced with incubation media (DMEM supplemented with NaCl (143 mmol/L), phenol red (3 mg/mL), glucose (10 mmol/L), glutamine (2 mmol/L), and pyruvate (2 mmol/L) at pH 7.4, and kept at 37°C, 1 h prior experimentation. The XF24 microplate was then loaded into the Seahorse XF24 analyzer following the manufacturer's instructions.</p>
<p>ATP and ADP measurements in siOPA1‐ or siCtrl‐transfected neurons were determined using a bioluminescence technique using an ATP monitoring kit as described previously.
<xref rid="acn3305-bib-0024" ref-type="ref">24</xref>
</p>
</sec>
<sec id="acn3305-sec-0012">
<title>Immunoblot analysis</title>
<p>Neurons, human fibroblasts, and 20 mg of DOA mice models and control littermate mice cortices were lysed for 30 min in a buffer containing 50 mmol/L Tris‐HCL pH 7.5, 250 mmol/L NaCl, 5 mmol/L Ethylene Diamine Tetraacetic acid (EDTA), 5 mmol/L Ethylene Glycol Tetraacetic Acid (EGTA), 1 mmol/L dithiothreitol, 0.1% Triton X‐100, 0.1% SDS, 1% deoxycholate, 1% tergitol‐type NP‐40, and protease inhibitors (“Complete” protease inhibitor mixture, Roche Applied Science, UK). After sonication, cell lysates were centrifuged at 20,000 g at 4°C for 10 min. Lysate from frozen mice cortices were obtained using the same protocol, but adding a step of dounce homogenization.</p>
<p>Total protein concentration was determined in the supernatant using the Bradford protein assay (Bio‐Rad). Proteins (100–200
<italic>μ</italic>
g) were separated by SDS‐PAGE (8–15%) and transferred onto nitrocellulose membranes (Whatman, Protran, Sigma‐Aldrich, UK). Free binding sites were blocked with 5% nonfat dry milk, 0.2% Tween 20 in Tris buffer saline, pH 7.6 (blocking buffer). The membranes were probed with various primary antibodies: anti‐OPA1 (1/300, BD Bioscience, USA), anti‐actin (1/25000, Chemicon, Merck Millipore, UK), anti‐HSP60 (1/8000, Sigma‐Aldrich, UK), anti‐citrate synthase (1/3000, Abcam, Cambridge, USA), anti‐VDAC (1/1000, Abcam, Cambridge, USA), anti‐TOM20 (1/2000, Abcam, Cambridge, USA), anti‐aconitase (1/500, Abcam, Cambridge, USA), anti‐SOD1 and anti‐SOD2 (1/2000, Epitomics, Abcam, Cambridge, USA), anti‐catalase (1/3000, Abcam, Cambridge, USA), anti‐NQO1 (1/3000, Abcam, Cambridge, USA), anti‐mitofilin (1/1000 Abcam, Cambridge, USA), anti‐MFN1 (1/1000 , Abnova, Taipei City, Taiwan), and anti‐GSTP1 (1/8000, Oxford Biochemical Research, Euromedex, France), and incubated overnight at 4°C in blocking buffer. After chemiluminescent detection of horseradish peroxidase‐conjugated secondary antibody (1/10000, Abcam, Cambridge, USA), scanned photographic films were quantitatively analyzed using Image J software.</p>
</sec>
<sec id="acn3305-sec-0013">
<title>Immunocytochemistry</title>
<p>Neurons were fixed for 15 min at 37°C with PBS containing 3.7% formaldehyde and permeabilized for 15 min in PBS, 1% bovine serum albumin, 0.3% Triton™ X‐100, and for 10 min at −20°C in methanol. Nonspecific sites were blocked for 1 h in PBS containing 5% goat serum, 3% bovine serum albumin, and 0.5% Tween 20 (blocking solution). Polyclonal antibodies against NRF2 (1/50, Santa Cruz Biotechnology, Dallas, TX, USA) were incubated overnight at 4°C in blocking solution. Neurons were then incubated with Alexa fluor 488‐conjugated secondary antibodies (1/300, Molecular Probes), labeled with 0.25
<italic>μ</italic>
g/mL Hoechst in PBS over 5 min and mounted in Mowiol. Immunolabeling was visualized under a fluorescence microscope (Nikon Eclipse 80i or Zeiss 710 Big) and images were acquired using NIS‐Element (Nikon Digital Sight DUS2 camera) or ZEN 2011 software. Nucleus raw integrated densities (sum of pixel values) of NRF2 by
<italic>μ</italic>
m
<sup>2</sup>
in neurons were measured using Image J software and confocal images.</p>
</sec>
<sec id="acn3305-sec-0014">
<title>Aconitase and catalase activities and GSH/GSSG measurements</title>
<p>Aconitase activities measurements were performed using a protocol described previously.
<xref rid="acn3305-bib-0025" ref-type="ref">25</xref>
The photochrome was measured at 340 nm using the UVIKON Spectrophotometer 922. Catalase activity was determined by measuring decomposition of H
<sub>2</sub>
O
<sub>2</sub>
at 240 nm as described previously.
<xref rid="acn3305-bib-0026" ref-type="ref">26</xref>
Reduced (GSH) and oxidized (GSSG) measurements were performed by reverse‐phase high‐performance liquid chromatography as described previously.
<xref rid="acn3305-bib-0026" ref-type="ref">26</xref>
For both aconitase and catalase activities measurement, 20 mg of DOA mice and control mice cortices were used per replicate of experiment and 1 × 10
<sup>6</sup>
cortical neurons ex vivo transfected with either siCtrl or siOPA1 has been harvested for the assay.</p>
</sec>
<sec id="acn3305-sec-0015">
<title>Rotenone treatment and cellular viability</title>
<p>Acute oxidative stress was induced in DIV6 (6 days in vitro) cultured neurons by 1 h incubation in neurobasal medium containing 500 nmol/L rotenone (Sigma‐Aldrich, UK). Then, rotenone was removed and cells were incubated for 3 h in neurobasal medium.</p>
<p>Neuronal viability was estimated by trypan blue exclusion and counting picnotic nuclei after DAPI staining as described in Bertholet et al.
<xref rid="acn3305-bib-0021" ref-type="ref">21</xref>
</p>
</sec>
<sec id="acn3305-sec-0016">
<title>Statistical analysis</title>
<p>Most of the experiments were statistically treated with Student's paired
<italic>t</italic>
‐test because of the systematic comparison between siCtrl and siOPA1 neurons. OCRs between siCtrl and siOPA1 neurons were investigated using a two‐way analysis of variance (ANOVA) with a Sidak's multiple comparison test. Nucleus NRF2 raw integrated densities in siCtrl and siOPA1 neurons were carried out using a nonparametric test (Mann–Whitney test). The results presented in Figures
<xref rid="acn3305-fig-0001" ref-type="fig">1</xref>
and
<xref rid="acn3305-fig-0003" ref-type="fig">3</xref>
were assessed by a two‐way ANOVA with a Sidak's multiple comparison test.
<italic>P</italic>
values: *
<italic>P</italic>
< 0.05, **
<italic>P</italic>
< 0.01, ***
<italic>P</italic>
< 0.001, and ****
<italic>P</italic>
< 0.0001 were considered statistically significant.</p>
<fig fig-type="Figure" xml:lang="en" id="acn3305-fig-0001" orientation="portrait" position="float">
<label>Figure 1</label>
<caption>
<p>Redox state is imbalanced in a dominant optic atrophy (
<styled-content style="fixed-case">DOA</styled-content>
) mice model.
<bold>(A)</bold>
Histograms representative of
<styled-content style="fixed-case">OPA</styled-content>
1 protein quantities, assessed by immunoblot, in cortices of 4 and 10 months old control mice (gray bars) and
<italic>Opa1</italic>
<sup>+/−</sup>
mice (striped bars).
<styled-content style="fixed-case">OPA</styled-content>
1 quantities are significantly decreased in both
<italic>Opa1</italic>
<sup>
<italic>+/</italic>
</sup>
4 (0.70 ± 0.10 arbitrary units [
<styled-content style="fixed-case">AU</styled-content>
]) and 10 (0.50 ± 0.05 AU) months old mice when compare to 4 (1.09 ± 0.10 AU) and 10 (1.01 ± 0.09 AU) months old
<italic>Opa1</italic>
<sup>
<italic>+/+</italic>
</sup>
mice. Results are expressed as mean ± standard error of the mean (
<styled-content style="fixed-case">SEM</styled-content>
) for 4 and 10 months old mice (
<italic>n</italic>
= 6–9). Statistical significance was determined by a two‐way analysis of variance (
<styled-content style="fixed-case">ANOVA</styled-content>
), *
<italic>P</italic>
< 0.05, **
<italic>P</italic>
< 0.01.
<bold>(B)</bold>
Histogram representative of aconitase activities in cortices of 4 and 10 months old
<italic>Opa1</italic>
<sup>
<italic>+/</italic>
</sup>
(striped bars) and
<italic>Opa1</italic>
<sup>
<italic>+/+</italic>
</sup>
control mice (gray bars) cortices. Aconitase activity is significantly lower in 4 (1.56 ± 0.11 mU/mg) and 10 (0.97 ± 0.23 mU/mg) months old
<italic>Opa1</italic>
<sup>
<italic>+/</italic>
</sup>
mice when compared to 4 (2.95 ± 0.40 mU/mg) and 10 (2.24 ± 0.38 mU/mg) months old control
<italic>Opa1</italic>
<sup>
<italic>+/+</italic>
</sup>
mice, respectively. Aconitase proteins quantities are unchanged in
<styled-content style="fixed-case">DOA</styled-content>
mice model's cortices. Results are expressed as mean ± SEM (
<italic>n</italic>
= 9 for 4 months old mice and
<italic>n</italic>
= 6 for 10 months old mice). Statistical significance was determined by a two‐way
<styled-content style="fixed-case">ANOVA</styled-content>
, *
<italic>P</italic>
< 0.05.
<bold>(C)</bold>
Histogram representative of
<styled-content style="fixed-case">SOD</styled-content>
1,
<styled-content style="fixed-case">SOD</styled-content>
2, and catalase proteins quantities relative to actin, assessed by immunoblot, in 4 and 10 months old
<italic>Opa1</italic>
<sup>
<italic>+/</italic>
</sup>
(striped bars) and
<italic>Opa1</italic>
<sup>
<italic>+/+</italic>
</sup>
(gray bars) mice cortices.
<styled-content style="fixed-case">SOD</styled-content>
1 proteins quantities are significantly increased in 10 months old (1.48 ± 0.17 AU)
<italic>Opa1</italic>
<sup>
<italic>+/</italic>
</sup>
mice compared to 4 months old (0.76 ± 0.08 AU)
<italic>Opa1</italic>
<sup>
<italic>+/</italic>
</sup>
mice. Likewise,
<styled-content style="fixed-case">SOD</styled-content>
2 proteins quantities are significantly increased in 10 months old (1.55 ± 0.11 AU)
<italic>Opa1</italic>
<sup>
<italic>+/</italic>
</sup>
mice compared to 4 months old (0.98 ± 0.09 AU)
<italic>Opa1</italic>
<sup>
<italic>+/</italic>
</sup>
mice. Catalase quantity is unchanged in
<italic>Opa1</italic>
<sup>+/+</sup>
and
<italic>Opa1</italic>
<sup>
<italic>+/</italic>
</sup>
4 and 10 months old mice. Results are expressed as mean ± SEM (
<italic>n</italic>
= 9 for 4 months old mice and
<italic>n</italic>
= 6 for 10 months old mice). Statistical significance was determined using a two‐way
<styled-content style="fixed-case">ANOVA</styled-content>
, *
<italic>P</italic>
< 0.05, **
<italic>P</italic>
< 0.01.</p>
</caption>
<graphic id="nlm-graphic-1" xlink:href="ACN3-3-408-g001"></graphic>
</fig>
</sec>
</sec>
<sec id="acn3305-sec-0017">
<title>Results</title>
<sec id="acn3305-sec-0018">
<title>
<italic>Opa1</italic>
<sup>
<italic>+/</italic>
</sup>
transgenic mice cortices show imbalanced redox state</title>
<p>To address the question of the implication of ROS as a determinant of DOA pathogenesis in mammals, we investigated whether a redox imbalance could be evidenced in vivo in a mouse model of DOA.
<xref rid="acn3305-bib-0020" ref-type="ref">20</xref>
</p>
<p>As expected, a decrease in OPA1 levels was evidenced in
<italic>Opa1</italic>
<sup>
<italic>+/</italic>
</sup>
transgenic mice cortices compared to litter mate
<italic>Opa1</italic>
<sup>+/+</sup>
mice (Fig.
<xref rid="acn3305-fig-0001" ref-type="fig">1</xref>
A). Cortices from 4 and 10 months old DOA mice were analyzed for aconitase activity. Aconitase was shown to be highly sensitive to oxidation due to damaged FeS core and inhibition of its activity is routinely used as a signature of increased mitochondrial ROS production.
<xref rid="acn3305-bib-0027" ref-type="ref">27</xref>
,
<xref rid="acn3305-bib-0028" ref-type="ref">28</xref>
,
<xref rid="acn3305-bib-0029" ref-type="ref">29</xref>
Downregulation of OPA1 correlated with a 47% and 57% inhibition of aconitase activity in 4 and 10 months old
<italic>Opa1</italic>
<sup>
<italic>+/</italic>
</sup>
transgenic mice compared to litter mate
<italic>Opa1</italic>
<sup>+/+</sup>
mice, respectively (Fig.
<xref rid="acn3305-fig-0001" ref-type="fig">1</xref>
B). Moreover, aconitase activity was slightly decreased (38%) from 4 to 10 months in
<italic>Opa1</italic>
<sup>
<italic>+/</italic>
</sup>
transgenic mice (Fig.
<xref rid="acn3305-fig-0001" ref-type="fig">1</xref>
B). This drop could not be attributed to a change in protein quantity since aconitase protein levels were unchanged (Fig.
<xref rid="acn3305-fig-0001" ref-type="fig">1</xref>
B).</p>
<p>Decrease in aconitase activity strongly supports higher mitochondrial ROS production in
<italic>Opa1</italic>
<sup>+/−</sup>
mice. However, to reinforce our data we monitored antioxidant defenses induction as a consequence of increased ROS levels. The protein levels of superoxide dismutases 1 and 2 (SOD1 and SOD2), that catalyze the dismutation of superoxide anion into hydrogen peroxide, and the levels of catalase, detoxifying hydrogen peroxide into water and oxygen, were estimated by immunoblotting in 4 and 10 months old
<italic>Opa1</italic>
<sup>
<italic>+/</italic>
</sup>
mice cortices and control (Fig.
<xref rid="acn3305-fig-0001" ref-type="fig">1</xref>
C). The data reveal an increase in SOD1 and SOD2, while catalase is unchanged in 10 months
<italic>Opa1</italic>
<sup>
<italic>+/</italic>
</sup>
mice cortices when compared to 4 months
<italic>Opa1</italic>
<sup>
<italic>+/</italic>
</sup>
mice. No difference was observed in control mice (Fig.
<xref rid="acn3305-fig-0001" ref-type="fig">1</xref>
C).</p>
<p>Altogether these findings suggest that during the 6 months interval (between 4 and 10 months) there is an excess of mitochondrial ROS in
<italic>Opa1</italic>
<sup>
<italic>+/</italic>
</sup>
mice that may induce an antioxidant response.</p>
</sec>
<sec id="acn3305-sec-0019">
<title>OPA1 downregulation imbalances neuronal redox state in vitro and activates NRF2 pathway</title>
<p>To decipher the molecular mechanisms underlying the excess of ROS in DOA mice, we used a previously characterized in vitro neuronal model of OPA1 haploinsufficiency.
<xref rid="acn3305-bib-0021" ref-type="ref">21</xref>
</p>
<p>Primary cortical neurons transfected with siRNA directed against OPA1 mRNA (siOPA1) showed 70% decrease in OPA1 protein levels at day 6 post‐transfection when compared to neurons treated with control siRNA (siCtrl) (Fig.
<xref rid="acn3305-fig-0002" ref-type="fig">2</xref>
A). Interestingly, citrate synthase, HSP60, VDAC, and TOM20 levels were unchanged (Fig.
<xref rid="acn3305-fig-0002" ref-type="fig">2</xref>
A), showing that mitochondria quantity was not affected by OPA1 downregulation.</p>
<fig fig-type="Figure" xml:lang="en" id="acn3305-fig-0002" orientation="portrait" position="float">
<label>Figure 2</label>
<caption>
<p>
<styled-content style="fixed-case">OPA</styled-content>
1 downregulation decreases mitochondrial respiration, induces the nuclear translocation of
<styled-content style="fixed-case">NRF</styled-content>
2, and increases both catalase quantity and activity in cortical neurons ex vivo. (A) Representative immunoblots and histograms showing protein levels of
<styled-content style="fixed-case">OPA</styled-content>
1 (inner membrane), citrate synthase (matrix),
<styled-content style="fixed-case">HSP</styled-content>
60 (matrix),
<styled-content style="fixed-case">VDAC</styled-content>
(outer membrane), and
<styled-content style="fixed-case">TOM</styled-content>
20 (outer membrane) relative to actin in si
<styled-content style="fixed-case">OPA</styled-content>
1‐ (gray bars) and siCtrl‐transfected (white bars) neurons. Only
<styled-content style="fixed-case">OPA</styled-content>
1 protein quantity is drastically decreased (0.23 ± 0.06 AU) in si
<styled-content style="fixed-case">OPA</styled-content>
1 neurons when compared to controls (1.08 ± 0.16 AU). Results are expressed as mean ± SEM (
<italic>n</italic>
= 5–8). Statistical significance was determined by Student's paired
<italic>t</italic>
‐test, ***
<italic>P</italic>
< 0.001. (B) Oxygen consumption rates were measured at 6 days in vitro (
<styled-content style="fixed-case">DIV</styled-content>
6) in neurons transfected with control small interfering
<styled-content style="fixed-case">RNA</styled-content>
(siCtrl, black line) or small interfering
<styled-content style="fixed-case">RNA</styled-content>
against
<styled-content style="fixed-case">OPA</styled-content>
1 (si
<styled-content style="fixed-case">OPA</styled-content>
1, dotted line). Spontaneous mitochondrial respiration is significantly lower in si
<styled-content style="fixed-case">OPA</styled-content>
1‐transfected neurons (0.48 ± 0.07 pmol/min per
<italic>μ</italic>
g) than in controls (0.92 ± 0.11 pmol/min per
<italic>μ</italic>
g). After 0.6
<italic>μ</italic>
mol/L oligomycin injection, cell respiration is also significantly lower in si
<styled-content style="fixed-case">OPA</styled-content>
1 neurons (0.20 ± 0.03 pmol/min per
<italic>μ</italic>
g) than in controls (0.40 ± 0.06 pmol/min per
<italic>μ</italic>
g). After 6
<italic>μ</italic>
mol/L
<styled-content style="fixed-case">FCCP</styled-content>
injection, maximal respiration is significantly lower si
<styled-content style="fixed-case">OPA</styled-content>
1 neurons (0.44 ± 0.09 pmol/min per
<italic>μ</italic>
g) than in controls (0.59 ± 0.16 pmol/min per
<italic>μ</italic>
g). Finally, 50 nmol/L rotenone and 0.18
<italic>μ</italic>
mol/L antimycin A injections inhibit mitochondrial respiration. Results are expressed as mean ± SEM (
<italic>n</italic>
= 3). Statistical significance was determined by a two‐way analysis of variance (
<styled-content style="fixed-case">ANOVA</styled-content>
), **
<italic>P</italic>
< 0.01 and ***
<italic>P</italic>
< 0.001. (C)
<styled-content style="fixed-case">ATP</styled-content>
over
<styled-content style="fixed-case">ADP</styled-content>
ratio is unchanged in si
<styled-content style="fixed-case">OPA</styled-content>
1 neurons (gray bar) when compared to siCtrl‐treated (white bar) cells. Results are expressed as mean ± SEM (
<italic>n</italic>
= 3). (D) Aconitase activity is lower in si
<styled-content style="fixed-case">OPA</styled-content>
1‐treated neurons (2.2 ± 0.4 mU/mg) (gray bars) when compared to control cells (3.0 ± 0.5 mU/mg) (white bars). Results are expressed as mean ± SEM (
<italic>n</italic>
= 5). Statistical analysis was determined by Student's paired
<italic>t</italic>
‐test, *
<italic>P</italic>
< 0.05. Representative immunoblots and histogram showing that
<styled-content style="fixed-case">OPA</styled-content>
1 downregulation in neurons has no effect on the aconitase quantity relative to actin. Results are expressed as mean ± SEM (
<italic>n</italic>
= 6). Statistical significance was determined by Student's paired
<italic>t</italic>
‐test, *
<italic>P</italic>
< 0.05. (E) a: Representative micrographs of
<styled-content style="fixed-case">NRF</styled-content>
2 immunolabeling (green) and Hoechst
<styled-content style="fixed-case">DNA</styled-content>
staining (blue) in si
<styled-content style="fixed-case">OPA</styled-content>
1 and siCtrl neurons. b: Histogram represents fluorescence intensity (pixels sum/
<italic>μ</italic>
m
<sup>2</sup>
) of nuclear
<styled-content style="fixed-case">NRF</styled-content>
2 determined by Image J software, which is higher in si
<styled-content style="fixed-case">OPA</styled-content>
1 neurons (471.4 ± 8.5 pixels sum/
<italic>μ</italic>
m
<sup>2</sup>
) (gray bars) than in controls (412.8 ± 6.2 pixels sum/
<italic>μ</italic>
m
<sup>2</sup>
) (white bars). Results are expressed as mean ± SEM (≈800 cells per condition,
<italic>n</italic>
= 5). Statistical significance was determined by a two‐way
<styled-content style="fixed-case">ANOVA</styled-content>
, ***
<italic>P</italic>
< 0.001. Scale bar: 5
<italic>μ</italic>
m. (F) Catalase activity is increased in si
<styled-content style="fixed-case">OPA</styled-content>
1 neurons (5.56 ± 0.66
<italic>μ</italic>
mol/min per mg) when compared to controls (3.44 ± 0.69
<italic>μ</italic>
mol/min per mg). Results are expressed as mean ± SEM (
<italic>n</italic>
= 5). Representative immunoblots and protein quantities of catalase in si
<styled-content style="fixed-case">OPA</styled-content>
1 (gray bars) and siCtrl (white bars) neurons relative to actin. Catalase quantity is increased in si
<styled-content style="fixed-case">OPA</styled-content>
1 (1.6 ± 0.33 AU) when compared to controls (0.85 ± 0.14 AU). Results are expressed as mean ± SEM (
<italic>n</italic>
= 7). Statistical significance was determined by Student's paired
<italic>t</italic>
‐test, *
<italic>P</italic>
< 0.05, **
<italic>P</italic>
< 0.01.</p>
</caption>
<graphic id="nlm-graphic-3" xlink:href="ACN3-3-408-g002"></graphic>
</fig>
<p>While OPA1 downregulation is known to negatively impact respiration in various cell lines,
<xref rid="acn3305-bib-0030" ref-type="ref">30</xref>
,
<xref rid="acn3305-bib-0031" ref-type="ref">31</xref>
,
<xref rid="acn3305-bib-0032" ref-type="ref">32</xref>
we evaluated here for the first time the effect of OPA1 downregulation on respiration in neurons using the Seahorse XF24 analyzer (Seahorse Bioscience) (Fig.
<xref rid="acn3305-fig-0002" ref-type="fig">2</xref>
B). In siOPA1‐ and siCtrl‐transfected neurons, rotenone and antimycin treatment considerably inhibited the OCR, showing that more than 95% of respiration was due to mitochondria. In siCtrl‐transfected neurons, oligomycin inhibited ATP‐linked respiration, while addition of the protonophore FCCP, which uncouples oxidation and phosphorylation, resulted in maximal OCR. In siOPA1‐transfected neurons, spontaneous respiration was reduced by 32%, when compared to siCtrl‐treated neurons. Furthermore, the ATP‐linked respiration and the maximal OCR were reduced by 34% and 46%, respectively. Contrary to siCtrl‐transfected cells, the maximal OCR in siOPA1‐treated cells was not significantly different than spontaneous OCR.</p>
<p>Therefore, depletion of OPA1 in neurons induced a decrease in spontaneous, ATP‐linked respiration and maximal mitochondrial respiration without affecting the mitochondrial biomass. A decrease in mitochondrial respiration could lead to energetic failure. However, we showed that neither ATP nor ADP levels, and thus the ratio of ATP over ADP, changed upon OPA1 downregulation in cortical neurons (Fig.
<xref rid="acn3305-fig-0002" ref-type="fig">2</xref>
C). The ratio of ATP/ADP (and the amounts of these molecules) in the steady state is the result of synthesis and consumption. So if it remains stable while cellular respiration in phosphorylation conditions decreases, it could be the consequence of a consumption decrease to the same extent as synthesis. In the context of neurons it can be reflected by a decrease in the functionality of these cells. This was shown in Bertholet et al.
<xref rid="acn3305-bib-0021" ref-type="ref">21</xref>
by the reduction of synaptogenesis and dendritogenesis in OPA1 downregulated neurons.</p>
<p>Impaired MRC functioning could lead to an imbalance in the redox state because of increased electron leaks. In agreement, we showed that aconitase activity was reduced by 33% in siOPA1‐treated neurons (Fig.
<xref rid="acn3305-fig-0002" ref-type="fig">2</xref>
D), with no effect in aconitase protein quantity, suggesting that OPA1 downregulation would induce an increase in mitochondrial ROS levels in cultured neurons.
<xref rid="acn3305-bib-0021" ref-type="ref">21</xref>
</p>
<p>Again, to reinforce these data we analyzed the NRF2 transcription factor pathway which accounts in large parts for the oxidative stress responses, and monitored some of its antioxidant target genes by analyzing SOD1, SOD2, and catalase protein levels.
<xref rid="acn3305-bib-0033" ref-type="ref">33</xref>
Since a hallmark of NRF2 activation is its nuclear translocation, we studied its intracellular localization. In neurons ex vivo, the decreased quantity of OPA1 induced a 1.3‐fold increase in nuclear NRF2 fluorescence compared to siCtrl cells (Fig.
<xref rid="acn3305-fig-0002" ref-type="fig">2</xref>
E). In siOPA1‐transfected neurons, no statistical differences in SOD1 and 2 (data not shown) were detected, while we uncovered an increase in both catalase quantity (88%) and activity (61%) (Fig.
<xref rid="acn3305-fig-0002" ref-type="fig">2</xref>
F).</p>
<p>We thus concluded that downregulation of OPA1 induced NRF2 activation leading to upregulation of, at least, one of its target genes, catalase.</p>
</sec>
<sec id="acn3305-sec-0020">
<title>Supplemental oxidative stress challenges viability of OPA1‐downregulated neurons</title>
<p>Since OPA1 downregulation leads to increased mitochondrial ROS levels that induce antioxidant response, we addressed the impact of an additional acute oxidative stress in OPA1 downregulated cells. A supplemental stress would therefore be deleterious for cell viability. Neurons were thus incubated with rotenone, a potent inhibitor of respiratory complex I that induces a major oxidative stress.
<xref rid="acn3305-bib-0034" ref-type="ref">34</xref>
,
<xref rid="acn3305-bib-0035" ref-type="ref">35</xref>
We checked this hypothesis by estimating the neuronal viability by trypan blue exclusion assay (Fig.
<xref rid="acn3305-fig-0003" ref-type="fig">3</xref>
A) as well as the number of picnotic nuclei by DAPI staining (Fig.
<xref rid="acn3305-fig-0003" ref-type="fig">3</xref>
B) of siCtrl‐ and siOPA1‐transfected neurons upon acute treatment of rotenone (1h, and 500 nmol/L). siOPA1‐transfected cells showed 8% decreased viability when compared to siCtrl‐treated neurons (Fig.
<xref rid="acn3305-fig-0003" ref-type="fig">3</xref>
A). Accordingly, the number of picnotic nuclei is increased by 34% in siOPA1 versus siCtrl rotenone‐treated neurons (Fig.
<xref rid="acn3305-fig-0003" ref-type="fig">3</xref>
B). These results suggest that antioxidant defenses may be overwhelmed in OPA1 downregulated conditions. We checked this hypothesis by analyzing the NRF2 nuclear translocation (Fig.
<xref rid="acn3305-fig-0003" ref-type="fig">3</xref>
C), as well as SOD1, SOD2, and catalase protein quantities (Fig.
<xref rid="acn3305-fig-0003" ref-type="fig">3</xref>
D). Upon acute rotenone treatment, NRF2 translocation was increased by 1.2‐fold in siCtrl neurons, while it was unchanged in siOPA1 neurons (Fig.
<xref rid="acn3305-fig-0003" ref-type="fig">3</xref>
C). Accordingly, a 1.5‐fold increase in SOD1 protein quantity was observed in control cells, while no change occurred in siOPA1 neurons (Fig.
<xref rid="acn3305-fig-0003" ref-type="fig">3</xref>
D). No difference was observed for SOD2 protein quantities in siOPA1 and siCtrl neurons treated or not with rotenone (Fig.
<xref rid="acn3305-fig-0003" ref-type="fig">3</xref>
D). Moreover, catalase protein quantity, which is increased by 32% in basal condition in siOPA1 neurons when compared to control neurons, did not increase and even diminished by 32%, suggesting that downregulation of OPA1 results in maximal catalase induction that cannot be further increased upon rotenone‐induced oxidative stress.</p>
<fig fig-type="Figure" xml:lang="en" id="acn3305-fig-0003" orientation="portrait" position="float">
<label>Figure 3</label>
<caption>
<p>Oxidative stress challenges viability of
<styled-content style="fixed-case">OPA</styled-content>
1 downregulated neurons, which do not induce the
<styled-content style="fixed-case">NRF</styled-content>
2 pathway. (A) Trypan blue exclusion assay were performed after acute rotenone treatment (1 h, 500 nmol/L rotenone treatment) in neurons transfected with (gray) or without si
<styled-content style="fixed-case">OPA</styled-content>
1 (white). Viable cells, which exclude the coloring agent, are lower in si
<styled-content style="fixed-case">OPA</styled-content>
1 neurons treated with rotenone (75.37 ± 2.13%) than in control cells treated with rotenone (83.9 ± 1.82%). Results are expressed as mean ± SEM (
<italic>n</italic>
= 5). Statistical significance was determined by a two‐way analysis of variance (
<styled-content style="fixed-case">ANOVA</styled-content>
), *
<italic>P</italic>
< 0.05. (B) Percentage of picnotic nuclei determined after Hoechst staining is higher in si
<styled-content style="fixed-case">OPA</styled-content>
1 (gray bars, 20.78 ± 2.50%) than in siCtrl neurons (white bars, 13.71 ± 1.82%) upon rotenone treatment. Results are expressed as mean ± SEM (
<italic>n</italic>
= 4, 100 nuclei per conditions). Statistical significance was determined by a two‐way
<styled-content style="fixed-case">ANOVA</styled-content>
, *
<italic>P</italic>
< 0.05, ***
<italic>P</italic>
< 0.001, ****
<italic>P</italic>
< 0.0001. (C) Histogram representing
<styled-content style="fixed-case">NRF</styled-content>
2 nuclear fluorescence intensity (pixels sum/
<italic>μ</italic>
m
<sup>2</sup>
, Image J software), which is higher in si
<styled-content style="fixed-case">OPA</styled-content>
1 neurons (gray bars, 471.4 ± 8.5 pixels sum/
<italic>μ</italic>
m
<sup>2</sup>
) than in controls (white bars, 412.8 ± 6.1 pixels sum/
<italic>μ</italic>
m
<sup>2</sup>
) in basal condition. After rotenone treatment,
<styled-content style="fixed-case">NRF</styled-content>
2 nuclear fluorescence intensity is increased in controls (478.7 ± 10.4 pixels sum/
<italic>μ</italic>
m
<sup>2</sup>
), but not in si
<styled-content style="fixed-case">OPA</styled-content>
1 neurons (459.9 ± 8.8 pixels sum/
<italic>μ</italic>
m
<sup>2</sup>
). Results are expressed as mean ± SEM (≈800 cells per condition,
<italic>n</italic>
= 5). Statistical significance was determined by a two‐way
<styled-content style="fixed-case">ANOVA</styled-content>
, ****
<italic>P</italic>
< 0.0001. (D) Representative immunoblots and histograms of
<styled-content style="fixed-case">SOD</styled-content>
1,
<styled-content style="fixed-case">SOD</styled-content>
2, and catalase proteins quantities relative to actin in si
<styled-content style="fixed-case">OPA</styled-content>
1 (gray bars) and siCtrl neurons (white bars) with or without rotenone treatment. Upon stress,
<styled-content style="fixed-case">SOD</styled-content>
1 quantity is increased in controls (without rotenone: 1.10 ± 0.10 AU; with rotenone: 1.49 ± 0.11 AU), but not in si
<styled-content style="fixed-case">OPA</styled-content>
1‐transfected neurons (without rotenone: 1.15 ± 0.19 AU; with rotenone: 0.88 ± 0.15 AU).
<styled-content style="fixed-case">SOD</styled-content>
2 quantities are the same in siCtrl‐ and si
<styled-content style="fixed-case">OPA</styled-content>
1‐transfected neurons with or without rotenone. Upon rotenone treatment catalase protein level is unchanged in controls (without rotenone: 1.19 ± 0.17 AU; with rotenone: 1.22 ± 0.13 AU), while it is decreased in si
<styled-content style="fixed-case">OPA</styled-content>
1 neurons (without rotenone: 1.67 ± 0.25 AU; with rotenone: 1.05 ± 0.1 AU). Results are expressed as mean ± SEM (
<italic>n</italic>
= 7). Statistical significance was determined by a two‐way
<styled-content style="fixed-case">ANOVA</styled-content>
, *
<italic>P</italic>
< 0.05, ***
<italic>P</italic>
< 0.001.</p>
</caption>
<graphic id="nlm-graphic-5" xlink:href="ACN3-3-408-g003"></graphic>
</fig>
</sec>
<sec id="acn3305-sec-0021">
<title>Cellular antioxidant defenses are impaired in certain fibroblasts from DOA patients</title>
<p>We next addressed the question of antioxidant defenses in DOA patients' fibroblasts (Table S1). SOD1, SOD2, and catalase were studied by immunoblot analysis (Fig.
<xref rid="acn3305-fig-0004" ref-type="fig">4</xref>
A). Fibroblasts from five healthy volunteers (C1–C5) and eight DOA patients (P1–P8) were analyzed. Even though heterogeneity exists between healthy volunteers and DOA patients in antioxidant defenses, some patients showed altered expression of antioxidant genes at the protein level (Fig.
<xref rid="acn3305-fig-0004" ref-type="fig">4</xref>
A and Fig. S2). DOA patients P1, P3, P5, and P7 expressed low levels of SOD1 and SOD2 proteins, while catalase was not different (Fig.
<xref rid="acn3305-fig-0004" ref-type="fig">4</xref>
A). To note, the levels of these antioxidant proteins are not correlated with the age of patients (Table S1 and Fig. S3).</p>
<fig fig-type="Figure" xml:lang="en" id="acn3305-fig-0004" orientation="portrait" position="float">
<label>Figure 4</label>
<caption>
<p>Antioxidant defenses and
<styled-content style="fixed-case">OPA</styled-content>
1 proteins levels in fibroblasts from dominant optic atrophy (
<styled-content style="fixed-case">DOA</styled-content>
) patients. (A) Representative histograms of
<styled-content style="fixed-case">SOD</styled-content>
1,
<styled-content style="fixed-case">SOD</styled-content>
2, and catalase proteins levels relative to actin in five healthy volunteer's skin fibroblasts (
<styled-content style="fixed-case">CTRL</styled-content>
, C1–C5) and eight
<styled-content style="fixed-case">DOA</styled-content>
patient's skin fibroblasts (
<styled-content style="fixed-case">DOA</styled-content>
, P1–P8) ex vivo.
<styled-content style="fixed-case">SOD</styled-content>
1
<styled-content style="fixed-case">CTRL</styled-content>
: 1.15 ± 0.25 AU;
<styled-content style="fixed-case">SOD</styled-content>
1
<styled-content style="fixed-case">DOA</styled-content>
: 0.62 ± 0.16 AU;
<styled-content style="fixed-case">SOD</styled-content>
2
<styled-content style="fixed-case">CTRL</styled-content>
: 1.21 ± 0.19 AU;
<styled-content style="fixed-case">SOD</styled-content>
2
<styled-content style="fixed-case">DOA</styled-content>
: 0.59 ± 0.16 AU; catalase
<styled-content style="fixed-case">CTRL</styled-content>
: 0.82 ± 0.10 AU; catalase
<styled-content style="fixed-case">DOA</styled-content>
: 0.96 ± 0.10 AU. Results are expressed as mean ± SEM (
<italic>n</italic>
= 3–6 per
<styled-content style="fixed-case">DOA</styled-content>
and
<styled-content style="fixed-case">CTRL</styled-content>
fibroblasts). (B) Representative histograms of
<styled-content style="fixed-case">OPA</styled-content>
1 protein levels relative to actin in five healthy volunteer's skin fibroblasts (
<styled-content style="fixed-case">CTRL</styled-content>
, C1–C5) and eight
<styled-content style="fixed-case">DOA</styled-content>
patient's skin fibroblasts (
<styled-content style="fixed-case">DOA</styled-content>
, P1–P8) ex vivo. Histogram representative of the ratio of the long (l‐
<styled-content style="fixed-case">OPA</styled-content>
1) to the short (s‐
<styled-content style="fixed-case">OPA</styled-content>
1)
<styled-content style="fixed-case">OPA</styled-content>
1 isoforms. Results are expressed as mean ± SEM (
<italic>n</italic>
= 3–6 per
<styled-content style="fixed-case">DOA</styled-content>
and
<styled-content style="fixed-case">CTRL</styled-content>
fibroblasts) Statistical significance for the ratio of the long to the short
<styled-content style="fixed-case">OPA</styled-content>
1 isoforms was determined by a one‐way analysis of variance (ANOVA), **
<italic>P</italic>
< 0.01.</p>
</caption>
<graphic id="nlm-graphic-7" xlink:href="ACN3-3-408-g004"></graphic>
</fig>
<p>These results show actual heterogeneity of SODs content in patient fibroblasts. To determine if this heterogeneity could be linked to the status of OPA1 protein level in fibroblasts from DOA patients, we investigated the levels of both total OPA1 protein and of the ratio of the long (l‐OPA1) relative to the short (s‐OPA1) OPA1 isoforms (Fig.
<xref rid="acn3305-fig-0004" ref-type="fig">4</xref>
B). P1 and P3 presented a lower quantity of OPA1. P5 and P7 presented differences of the ratio l‐OPA1 over s‐OPA1.</p>
</sec>
</sec>
<sec id="acn3305-sec-0022">
<title>Discussion</title>
<p>In the present work, we clearly demonstrate that OPA1 dysfunction induces a decline in endogenous mitochondrial respiration in primary cultured neurons with a maximal mitochondrial respiration that does not exceed basal mitochondrial respiration. Defects in mitochondrial respiration have also been observed in several cell lines upon OPA1 knockdown. In agreement with our results, OPA1‐depleted MEFs also showed a severe reduction in endogenous respiration, no effect after uncoupling, and a decrease in oxygen consumption mediated by complexes I, II, and IV.
<xref rid="acn3305-bib-0032" ref-type="ref">32</xref>
Studies using fibroblasts from DOA patients support the observation of a coupling defect of oxidative phosphorylation due to a reduction in complex IV at the protein level and heightened complex V activity, with no change in ATP production in either case.
<xref rid="acn3305-bib-0012" ref-type="ref">12</xref>
Defective oxidative phosphorylation with lowered ATP production has been found in DOA patient tissue analyzed by MR spectroscopy.
<xref rid="acn3305-bib-0036" ref-type="ref">36</xref>
However, the existence and the nature of the mitochondrial oxidative phosphorylation dysfunction are still controversial since other studies did not find respiratory defects in lymphoblastoid lines
<xref rid="acn3305-bib-0031" ref-type="ref">31</xref>
or in muscle biopsies of patients with DOA.
<xref rid="acn3305-bib-0008" ref-type="ref">8</xref>
Along these lines, cardiomyocytes derived from
<italic>Opa1</italic>
<sup>+/−</sup>
mice also do not display any respiratory defects although mitochondrial morphology is altered.
<xref rid="acn3305-bib-0037" ref-type="ref">37</xref>
</p>
<p>Mitochondrial respiration impairment due to OPA1 loss could originate from either downregulation/deregulation of MRC, or disorganization of the mitochondrial inner membrane structure,
<xref rid="acn3305-bib-0038" ref-type="ref">38</xref>
or both. Indeed, we have previously shown that the protein levels of certain subunits of MRC were decreased in siOPA1‐transfected neurons.
<xref rid="acn3305-bib-0021" ref-type="ref">21</xref>
Besides, NDUFB8 and NDUFA9 from complex I, Core 1 and Core 2 from complex III, and Cox I and Cox VIc from complex IV decreased by 30–50%. Moreover, Zanna and colleagues have suggested a direct interaction between OPA1 and the respiratory chain complexes I, II, and III.
<xref rid="acn3305-bib-0014" ref-type="ref">14</xref>
On the other hand, a decrease in respiration in siOPA1‐treated neurons could also damage the spatial organization of the MRC.
<xref rid="acn3305-bib-0039" ref-type="ref">39</xref>
,
<xref rid="acn3305-bib-0040" ref-type="ref">40</xref>
The organization of respiratory chain complexes into supercomplexes (RCS) is essential for the assembly and stability of each complex
<xref rid="acn3305-bib-0041" ref-type="ref">41</xref>
,
<xref rid="acn3305-bib-0042" ref-type="ref">42</xref>
,
<xref rid="acn3305-bib-0043" ref-type="ref">43</xref>
and may influence the efficiency of electron transfer.
<xref rid="acn3305-bib-0044" ref-type="ref">44</xref>
,
<xref rid="acn3305-bib-0045" ref-type="ref">45</xref>
,
<xref rid="acn3305-bib-0046" ref-type="ref">46</xref>
Therefore, it is likely, as recently suggested, that the respiratory deficiency that we observed upon OPA1 downregulation in neurons results from disruption of RCS.
<xref rid="acn3305-bib-0047" ref-type="ref">47</xref>
</p>
<p>Our studies using primary cortical neurons did not reveal any significant changes in ATP or ADP levels. This may indicate that either OPA1 knockdown is not sufficient to induce an energetic failure or any decrease in mitochondrial ATP synthesis is offset by a decrease in global ATP consumption. In addition, we found an increase in mitochondrial ROS production revealed by a decrease in aconitase activity in OPA1‐depleted neurons, consistent with prior findings in OPA1‐mutant flies.
<xref rid="acn3305-bib-0048" ref-type="ref">48</xref>
We demonstrated that this increase in ROS production activates the NRF2 pathway, one of the primary mechanisms of cell detoxification.
<xref rid="acn3305-bib-0049" ref-type="ref">49</xref>
Indeed, downregulation of OPA1 in primary neurons induced relocalization of the NRF2 transcription factor into the nucleus, where it upregulated target genes bearing antioxidant responsive promoter elements, such as catalase. Moreover, upregulation of antioxidant defenses also confirms that OPA1 downregulation promotes higher ROS production. Additional exogenous oxidative stress caused by rotenone reduced neuronal viability. These results are consistent with previous experiments showing hypersensitivity to oxidative stress in
<italic>C. elegans</italic>
and in
<italic>D. melanogaster</italic>
.
<xref rid="acn3305-bib-0017" ref-type="ref">17</xref>
,
<xref rid="acn3305-bib-0018" ref-type="ref">18</xref>
Our data suggest that the viability impairment of siOPA1‐treated neurons upon acute oxidative stress is due to inefficient antioxidant defenses. SOD1, SOD2, and catalase levels did not change upon acute oxidative stress in siOPA1 rotenone‐treated cells, while SOD1 levels increased in rotenone‐treated control cells. Our experiments performed in
<italic>Opa1</italic>
<sup>
<italic>+/</italic>
</sup>
mutant mice also found decreased aconitase activity and increased antioxidant defenses in
<italic>Opa1</italic>
<sup>
<italic>+/</italic>
</sup>
mouse cortices compared to controls, again suggesting excess mitochondrial ROS production.</p>
<p>Our results thus underscore the critical role of NRF2 signaling in DOA pathogenesis. To note, even if we observed a difference in the NRF2‐induced target in our different DOA models, this is largely described in the literature (for review see ref.
<xref rid="acn3305-bib-0049" ref-type="ref">49</xref>
) as well as the different answers observed for the antioxidant defenses among the patient's fibroblasts for example for Friedreich Ataxia.
<xref rid="acn3305-bib-0050" ref-type="ref">50</xref>
Because many studies focused on fibroblasts to understand cellular attempts common to different cells and particularly neurodegenerative diseases,
<xref rid="acn3305-bib-0007" ref-type="ref">7</xref>
,
<xref rid="acn3305-bib-0014" ref-type="ref">14</xref>
,
<xref rid="acn3305-bib-0050" ref-type="ref">50</xref>
we decided in this study to compare the effect of OPA1 dysfunction in DOA patients fibroblasts and neuronal models. In all our models, the NRF2 pathway is associated with activation of SOD1 or SOD2 or catalase, but not GSTP1 and NQO1. Indeed, these last two enzymes remain constant for all the patients as well as in neurons treated by siOPA1 or siCtrl, which supports distinguishing SOD1 or SOD2 or catalase as reliable markers.</p>
<p>Our results also show that DOA pathogenesis may be related to both the quantity of OPA1 protein and the relative levels of the five isoforms of OPA1. P1 and P3 present a decrease in OPA1 protein levels, suggesting that, in these two cases, haploinsufficiency may contribute to DOA pathogenesis. P5 and P7 present lower ratios of l‐OPA1 to s‐OPA1, which might also contribute to the severity of this disease. Changes in the ratio of l‐OPA1 to s‐OPA1 were shown to impact OPA1 functions,
<xref rid="acn3305-bib-0010" ref-type="ref">10</xref>
and in a recent work, the level of l‐OPA1 was presented as a possible therapeutic target for ischemia–reperfusion injury‐related diseases at least for the retina because it could restore mitochondrial morphology.
<xref rid="acn3305-bib-0051" ref-type="ref">51</xref>
These results suggest that the protein profile of OPA1, beyond merely its genetic status, should be considered in future research.</p>
<p>As we showed here for DOA, dysfunctions in ROS metabolism may also underlie Leber's hereditary optic neuropathy,
<xref rid="acn3305-bib-0052" ref-type="ref">52</xref>
glaucoma,
<xref rid="acn3305-bib-0053" ref-type="ref">53</xref>
and other neurodegenerative pathologies such as Alzheimer's and Parkinson's diseases.
<xref rid="acn3305-bib-0054" ref-type="ref">54</xref>
In our model, the redox state imbalance induced by OPA1 haploinsufficiency could escalate into an oxidative stress state upon exposure to factors such as UV light, blue light, smoke, aging, or diet. Interestingly, some studies described an increase in sensitivity of RGCs to blue light stimulation and to hypoxia.
<xref rid="acn3305-bib-0055" ref-type="ref">55</xref>
,
<xref rid="acn3305-bib-0056" ref-type="ref">56</xref>
,
<xref rid="acn3305-bib-0057" ref-type="ref">57</xref>
Additionally, within the retina, mitochondrial DNA lesions reach remarkable levels after exposure to endogenous ROS generated by the mitochondria themselves or from photosensitizers
<xref rid="acn3305-bib-0058" ref-type="ref">58</xref>
(for review see ref.
<xref rid="acn3305-bib-0059" ref-type="ref">59</xref>
). In this framework, it is noteworthy that numerous studies suggest that increasing mitochondrial antioxidant defenses may rescue optic nerve defects.
<xref rid="acn3305-bib-0053" ref-type="ref">53</xref>
,
<xref rid="acn3305-bib-0060" ref-type="ref">60</xref>
,
<xref rid="acn3305-bib-0061" ref-type="ref">61</xref>
Cellular systems that protect against oxidants include antioxidant defenses enzymes and oxidant scavengers. In addition to direct antioxidants, a number of antioxidant compounds are weak pro‐oxidants with indirect antioxidant properties.
<xref rid="acn3305-bib-0062" ref-type="ref">62</xref>
Acting as weak pro‐oxidants, these molecules have limited ability to cause oxidative damage, yet they are potent enough to activate the NRF2‐mediated adaptive response, thereby protecting against existing or future high‐level stresses.
<xref rid="acn3305-bib-0063" ref-type="ref">63</xref>
Given that our findings that OPA1 downregulation activates NRF2 pathway, the pro‐oxidant compounds could be insufficient or dangerous therapeutic approaches for some DOA patients. Instead, direct testing of antioxidant compounds could offer a complementary pharmacological treatment for DOA. Accordingly, a recent study focused on idebenone, a benzoquinone that transfers electrons to complex III bypassing complex I, as a potential curative compound for DOA patients.
<xref rid="acn3305-bib-0064" ref-type="ref">64</xref>
Thus, it is of primary importance to identify patients able to induce an efficient NRF2 response. Our work could thus have a direct impact on both treatment and care of DOA patients. Indeed, the low expression of some antioxidant enzymes in several DOA patients could support their role as modifier factors for DOA.</p>
</sec>
<sec id="acn3305-sec-0024">
<title>Conflict of Interest</title>
<p>None declared.</p>
</sec>
<sec sec-type="supplementary-material">
<title>Supporting information</title>
<supplementary-material content-type="local-data">
<caption>
<p>
<bold>Figure S1.</bold>
Protein levels of MFN1 and mitofilin in OPA1 downregulated neurons. Representative immunoblots and histograms showing protein levels of mitofilin and mitofusin 1 (MFN1) relative to actin in siOPA1‐ (gray bars) and siCtrl‐transfected (white bars) neurons. Results are expressed as mean ± SEM (
<italic>n</italic>
= 5). Statistical significance was determined by Student's paired
<italic>t</italic>
‐test.</p>
</caption>
<media xlink:href="ACN3-3-408-s001.eps">
<caption>
<p>Click here for additional data file.</p>
</caption>
</media>
</supplementary-material>
<supplementary-material content-type="local-data">
<caption>
<p>
<bold>Figure S2.</bold>
DOA patient's fibroblasts and control's fibroblasts antioxidant proteins levels.
<bold>(A)</bold>
Representative immunoblots showing protein levels of SOD1, SOD2, catalase, GSTP1, NQO1, and actin, and histograms representing GSTP1 and NQO1 proteins quantities, in five healthy volunteer's (CTRL) fibroblasts (C1–C5) and in eight DOA patient's fibroblasts (P1–P8).
<bold>(B)</bold>
Representative immunoblots showing protein levels of OPA1 and actin.</p>
</caption>
<media xlink:href="ACN3-3-408-s002.eps">
<caption>
<p>Click here for additional data file.</p>
</caption>
</media>
</supplementary-material>
<supplementary-material content-type="local-data">
<caption>
<p>
<bold>Figure S3.</bold>
DOA patients' and controls' fibroblasts antioxidant proteins levels relative to their age at the moment of the biopsies. Antioxidant proteins levels (SOD1, SOD2, catalase, GSTP1, and NQO1) relative to the age of patients and controls' fibroblasts: the quantity of each protein is not correlated to the age of the persons.</p>
</caption>
<media xlink:href="ACN3-3-408-s003.eps">
<caption>
<p>Click here for additional data file.</p>
</caption>
</media>
</supplementary-material>
<supplementary-material content-type="local-data">
<caption>
<p>
<bold>Table S1.</bold>
Clinical symptoms and mutations of DOA patients.</p>
</caption>
<media xlink:href="ACN3-3-408-s004.docx">
<caption>
<p>Click here for additional data file.</p>
</caption>
</media>
</supplementary-material>
</sec>
</body>
<back>
<ack id="acn3305-sec-0023">
<title>Acknowledgments</title>
<p>We thank Brice Ronsin for technical assistance with Image J software and technical assistance in confocal analysis on the Toulouse RIO imaging platform. We thank the animal core facility at the Center of Developmental Biology. This project was supported by grants from the Centre National de la Recherche Scientifique, the Université Paul Sabatier, the “Rétina‐France,” the “Union Nationale Des Aveugles et Déficients Visuels,” the “Gueules Cassées Sourire quand même,” and the “Association contre les Maladies Mitochondriales.” A. Millet was funded by the French Ministry for Research and Education and the “Rétina‐France” for Ph.D. studies.</p>
</ack>
<ref-list content-type="cited-references" id="acn3305-bibl-0001">
<title>References</title>
<ref id="acn3305-bib-0001">
<label>1</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0001">
<string-name>
<surname>Amati‐Bonneau</surname>
<given-names>P</given-names>
</string-name>
,
<string-name>
<surname>Milea</surname>
<given-names>D</given-names>
</string-name>
,
<string-name>
<surname>Bonneau</surname>
<given-names>D</given-names>
</string-name>
, et al.
<article-title>OPA1‐associated disorders: phenotypes and pathophysiology</article-title>
.
<source>Int J Biochem Cell Biol</source>
<year>2009</year>
;
<volume>41</volume>
:
<fpage>1855</fpage>
<lpage>1865</lpage>
.
<pub-id pub-id-type="pmid">19389487</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0002">
<label>2</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0002">
<string-name>
<surname>Delettre</surname>
<given-names>C</given-names>
</string-name>
,
<string-name>
<surname>Lenaers</surname>
<given-names>G</given-names>
</string-name>
,
<string-name>
<surname>Griffoin</surname>
<given-names>JM</given-names>
</string-name>
, et al.
<article-title>Nuclear gene OPA1, encoding a mitochondrial dynamin‐related protein, is mutated in dominant optic atrophy</article-title>
[In Process Citation].
<source>Nat Genet</source>
<year>2000</year>
;
<volume>26</volume>
:
<fpage>207</fpage>
<lpage>210</lpage>
.
<pub-id pub-id-type="pmid">11017079</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0003">
<label>3</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0003">
<string-name>
<surname>Cohn</surname>
<given-names>AC</given-names>
</string-name>
,
<string-name>
<surname>Toomes</surname>
<given-names>C</given-names>
</string-name>
,
<string-name>
<surname>Potter</surname>
<given-names>C</given-names>
</string-name>
, et al.
<article-title>Autosomal dominant optic atrophy: penetrance and expressivity in patients with OPA1 mutations</article-title>
.
<source>Am J Ophthalmol</source>
<year>2007</year>
;
<volume>143</volume>
:
<fpage>656</fpage>
<lpage>662</lpage>
.
<pub-id pub-id-type="pmid">17306754</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0004">
<label>4</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0004">
<string-name>
<surname>Yu‐Wai‐Man</surname>
<given-names>P</given-names>
</string-name>
,
<string-name>
<surname>Griffiths</surname>
<given-names>PG</given-names>
</string-name>
,
<string-name>
<surname>Burke</surname>
<given-names>A</given-names>
</string-name>
, et al.
<article-title>The prevalence and natural history of dominant optic atrophy due to OPA1 mutations</article-title>
.
<source>Ophthalmology</source>
<year>2010</year>
;
<volume>117</volume>
:
<fpage>1538</fpage>
<lpage>1546</lpage>
, 46 e1.
<pub-id pub-id-type="pmid">20417570</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0005">
<label>5</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0005">
<string-name>
<surname>Amati‐Bonneau</surname>
<given-names>P</given-names>
</string-name>
,
<string-name>
<surname>Valentino</surname>
<given-names>ML</given-names>
</string-name>
,
<string-name>
<surname>Reynier</surname>
<given-names>P</given-names>
</string-name>
, et al.
<article-title>OPA1 mutations induce mitochondrial DNA instability and optic atrophy ‘plus' phenotypes</article-title>
.
<source>Brain</source>
<year>2008</year>
;
<volume>131</volume>
(
<issue>Pt 2</issue>
):
<fpage>338</fpage>
<lpage>351</lpage>
.
<pub-id pub-id-type="pmid">18158317</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0006">
<label>6</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0006">
<string-name>
<surname>Zeviani</surname>
<given-names>M.</given-names>
</string-name>
<article-title>OPA1 mutations and mitochondrial DNA damage: keeping the magic circle in shape</article-title>
.
<source>Brain</source>
<year>2008</year>
;
<volume>131</volume>
(
<issue>Pt 2</issue>
):
<fpage>314</fpage>
<lpage>317</lpage>
.
<pub-id pub-id-type="pmid">18222990</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0007">
<label>7</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0007">
<string-name>
<surname>Carelli</surname>
<given-names>V</given-names>
</string-name>
,
<string-name>
<surname>Musumeci</surname>
<given-names>O</given-names>
</string-name>
,
<string-name>
<surname>Caporali</surname>
<given-names>L</given-names>
</string-name>
, et al.
<article-title>Syndromic parkinsonism and dementia associated with OPA1 missense mutations</article-title>
.
<source>Ann Neurol</source>
<year>2015</year>
;
<volume>78</volume>
:
<fpage>21</fpage>
<lpage>38</lpage>
.
<pub-id pub-id-type="pmid">25820230</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0008">
<label>8</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0008">
<string-name>
<surname>Spinazzi</surname>
<given-names>M</given-names>
</string-name>
,
<string-name>
<surname>Cazzola</surname>
<given-names>S</given-names>
</string-name>
,
<string-name>
<surname>Bortolozzi</surname>
<given-names>M</given-names>
</string-name>
, et al.
<article-title>A novel deletion in the GTPase domain of OPA1 causes defects in mitochondrial morphology and distribution, but not in function</article-title>
.
<source>Hum Mol Genet</source>
<year>2008</year>
;
<volume>17</volume>
:
<fpage>3291</fpage>
<lpage>3302</lpage>
.
<pub-id pub-id-type="pmid">18678599</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0009">
<label>9</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0009">
<string-name>
<surname>Mackey</surname>
<given-names>DA</given-names>
</string-name>
,
<string-name>
<surname>Trounce</surname>
<given-names>I</given-names>
</string-name>
.
<article-title>Genetics: optic nerve genetics–more than meets the eye</article-title>
.
<source>Nat Rev Neurol</source>
<year>2010</year>
;
<volume>6</volume>
:
<fpage>357</fpage>
<lpage>358</lpage>
.
<pub-id pub-id-type="pmid">20639910</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0010">
<label>10</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0010">
<string-name>
<surname>Belenguer</surname>
<given-names>P</given-names>
</string-name>
,
<string-name>
<surname>Pellegrini</surname>
<given-names>L</given-names>
</string-name>
.
<article-title>The dynamin GTPase OPA1: more than mitochondria?</article-title>
<source>Biochim Biophys Acta</source>
<year>2013</year>
;
<volume>1833</volume>
:
<fpage>176</fpage>
<lpage>183</lpage>
.
<pub-id pub-id-type="pmid">22902477</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0011">
<label>11</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0011">
<string-name>
<surname>Olichon</surname>
<given-names>A</given-names>
</string-name>
,
<string-name>
<surname>Guillou</surname>
<given-names>E</given-names>
</string-name>
,
<string-name>
<surname>Delettre</surname>
<given-names>C</given-names>
</string-name>
, et al.
<article-title>Mitochondrial dynamics and disease, OPA1</article-title>
.
<source>Biochim Biophys Acta</source>
<year>2006</year>
;
<volume>1763</volume>
:
<fpage>500</fpage>
<lpage>509</lpage>
.
<pub-id pub-id-type="pmid">16737747</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0012">
<label>12</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0012">
<string-name>
<surname>Chevrollier</surname>
<given-names>A</given-names>
</string-name>
,
<string-name>
<surname>Guillet</surname>
<given-names>V</given-names>
</string-name>
,
<string-name>
<surname>Loiseau</surname>
<given-names>D</given-names>
</string-name>
, et al.
<article-title>Hereditary optic neuropathies share a common mitochondrial coupling defect</article-title>
.
<source>Ann Neurol</source>
<year>2008</year>
;
<volume>63</volume>
:
<fpage>794</fpage>
<lpage>798</lpage>
.
<pub-id pub-id-type="pmid">18496845</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0013">
<label>13</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0013">
<string-name>
<surname>Alavi</surname>
<given-names>MV</given-names>
</string-name>
,
<string-name>
<surname>Fuhrmann</surname>
<given-names>N</given-names>
</string-name>
,
<string-name>
<surname>Nguyen</surname>
<given-names>HP</given-names>
</string-name>
, et al.
<article-title>Subtle neurological and metabolic abnormalities in an Opa1 mouse model of autosomal dominant optic atrophy</article-title>
.
<source>Exp Neurol</source>
<year>2009</year>
;
<volume>220</volume>
:
<fpage>404</fpage>
<lpage>409</lpage>
.
<pub-id pub-id-type="pmid">19815013</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0014">
<label>14</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0014">
<string-name>
<surname>Zanna</surname>
<given-names>C</given-names>
</string-name>
,
<string-name>
<surname>Ghelli</surname>
<given-names>A</given-names>
</string-name>
,
<string-name>
<surname>Porcelli</surname>
<given-names>AM</given-names>
</string-name>
, et al.
<article-title>OPA1 mutations associated with dominant optic atrophy impair oxidative phosphorylation and mitochondrial fusion</article-title>
.
<source>Brain</source>
<year>2008</year>
;
<volume>131</volume>
(
<issue>Pt 2</issue>
):
<fpage>352</fpage>
<lpage>367</lpage>
.
<pub-id pub-id-type="pmid">18222991</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0015">
<label>15</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0015">
<string-name>
<surname>Yu‐Wai‐Man</surname>
<given-names>P</given-names>
</string-name>
,
<string-name>
<surname>Shankar</surname>
<given-names>SP</given-names>
</string-name>
,
<string-name>
<surname>Biousse</surname>
<given-names>V</given-names>
</string-name>
, et al.
<article-title>Genetic screening for OPA1 and OPA3 mutations in patients with suspected inherited optic neuropathies</article-title>
.
<source>Ophthalmology</source>
<year>2011</year>
;
<volume>118</volume>
:
<fpage>558</fpage>
<lpage>563</lpage>
.
<pub-id pub-id-type="pmid">21036400</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0016">
<label>16</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0016">
<string-name>
<surname>Landes</surname>
<given-names>T</given-names>
</string-name>
,
<string-name>
<surname>Leroy</surname>
<given-names>I</given-names>
</string-name>
,
<string-name>
<surname>Bertholet</surname>
<given-names>A</given-names>
</string-name>
, et al.
<article-title>OPA1 (dys)functions</article-title>
.
<source>Semin Cell Dev Biol</source>
<year>2010</year>
;
<volume>21</volume>
:
<fpage>593</fpage>
<lpage>598</lpage>
.
<pub-id pub-id-type="pmid">20045077</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0017">
<label>17</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0017">
<string-name>
<surname>Kanazawa</surname>
<given-names>T</given-names>
</string-name>
,
<string-name>
<surname>Zappaterra</surname>
<given-names>MD</given-names>
</string-name>
,
<string-name>
<surname>Hasegawa</surname>
<given-names>A</given-names>
</string-name>
, et al.
<article-title>The C. elegans Opa1 homologue EAT‐3 is essential for resistance to free radicals</article-title>
.
<source>PLoS Genet</source>
<year>2008</year>
;
<volume>4</volume>
:
<fpage>e1000022</fpage>
.
<pub-id pub-id-type="pmid">18454199</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0018">
<label>18</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0018">
<string-name>
<surname>Tang</surname>
<given-names>S</given-names>
</string-name>
,
<string-name>
<surname>Le</surname>
<given-names>PK</given-names>
</string-name>
,
<string-name>
<surname>Tse</surname>
<given-names>S</given-names>
</string-name>
, et al.
<article-title>Heterozygous mutation of Opa1 in Drosophila shortens lifespan mediated through increased reactive oxygen species production</article-title>
.
<source>PLoS One</source>
<year>2009</year>
;
<volume>4</volume>
:
<fpage>e4492</fpage>
.
<pub-id pub-id-type="pmid">19221591</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0019">
<label>19</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0019">
<string-name>
<surname>Shahrestani</surname>
<given-names>P</given-names>
</string-name>
,
<string-name>
<surname>Leung</surname>
<given-names>HT</given-names>
</string-name>
,
<string-name>
<surname>Le</surname>
<given-names>PK</given-names>
</string-name>
, et al.
<article-title>Heterozygous mutation of Drosophila Opa1 causes the development of multiple organ abnormalities in an age‐dependent and organ‐specific manner</article-title>
.
<source>PLoS One</source>
<year>2009</year>
;
<volume>4</volume>
:
<fpage>e6867</fpage>
.
<pub-id pub-id-type="pmid">19718456</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0020">
<label>20</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0020">
<string-name>
<surname>Alavi</surname>
<given-names>MV</given-names>
</string-name>
,
<string-name>
<surname>Bette</surname>
<given-names>S</given-names>
</string-name>
,
<string-name>
<surname>Schimpf</surname>
<given-names>S</given-names>
</string-name>
, et al.
<article-title>A splice site mutation in the murine Opa1 gene features pathology of autosomal dominant optic atrophy</article-title>
.
<source>Brain</source>
<year>2007</year>
;
<volume>130</volume>
(
<issue>Pt 4</issue>
):
<fpage>1029</fpage>
<lpage>1042</lpage>
.
<pub-id pub-id-type="pmid">17314202</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0021">
<label>21</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0021">
<string-name>
<surname>Bertholet</surname>
<given-names>AM</given-names>
</string-name>
,
<string-name>
<surname>Millet</surname>
<given-names>AM</given-names>
</string-name>
,
<string-name>
<surname>Guillermin</surname>
<given-names>O</given-names>
</string-name>
, et al.
<article-title>OPA1 loss of function affects in vitro neuronal maturation</article-title>
.
<source>Brain</source>
<year>2013</year>
;
<volume>136</volume>
(
<issue>Pt 5</issue>
):
<fpage>1518</fpage>
<lpage>1533</lpage>
.
<pub-id pub-id-type="pmid">23543485</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0022">
<label>22</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0022">
<string-name>
<surname>Chevrollier</surname>
<given-names>A</given-names>
</string-name>
,
<string-name>
<surname>Cassereau</surname>
<given-names>J</given-names>
</string-name>
,
<string-name>
<surname>Ferre</surname>
<given-names>M</given-names>
</string-name>
, et al.
<article-title>Standardized mitochondrial analysis gives new insights into mitochondrial dynamics and OPA1 function</article-title>
.
<source>Int J Biochem Cell Biol</source>
<year>2012</year>
;
<volume>44</volume>
:
<fpage>980</fpage>
<lpage>988</lpage>
.
<pub-id pub-id-type="pmid">22433900</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0023">
<label>23</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0023">
<string-name>
<surname>Heiduschka</surname>
<given-names>P</given-names>
</string-name>
,
<string-name>
<surname>Schnichels</surname>
<given-names>S</given-names>
</string-name>
,
<string-name>
<surname>Fuhrmann</surname>
<given-names>N</given-names>
</string-name>
, et al.
<article-title>Electrophysiological and histologic assessment of retinal ganglion cell fate in a mouse model for OPA1‐associated autosomal dominant optic atrophy</article-title>
.
<source>Invest Ophthalmol Vis Sci</source>
<year>2010</year>
;
<volume>51</volume>
:
<fpage>1424</fpage>
<lpage>1431</lpage>
.
<pub-id pub-id-type="pmid">19834041</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0024">
<label>24</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0024">
<string-name>
<surname>Devin</surname>
<given-names>A</given-names>
</string-name>
,
<string-name>
<surname>Nogueira</surname>
<given-names>V</given-names>
</string-name>
,
<string-name>
<surname>Averet</surname>
<given-names>N</given-names>
</string-name>
, et al.
<article-title>Profound effects of the general anesthetic etomidate on oxidative phosphorylation without effects on their yield</article-title>
.
<source>J Bioenerg Biomembr</source>
<year>2006</year>
;
<volume>38</volume>
:
<fpage>137</fpage>
<lpage>142</lpage>
.
<pub-id pub-id-type="pmid">17029016</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0025">
<label>25</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0025">
<string-name>
<surname>Colombani</surname>
<given-names>AL</given-names>
</string-name>
,
<string-name>
<surname>Carneiro</surname>
<given-names>L</given-names>
</string-name>
,
<string-name>
<surname>Benani</surname>
<given-names>A</given-names>
</string-name>
, et al.
<article-title>Enhanced hypothalamic glucose sensing in obesity: alteration of redox signaling</article-title>
.
<source>Diabetes</source>
<year>2009</year>
;
<volume>58</volume>
:
<fpage>2189</fpage>
<lpage>2197</lpage>
.
<pub-id pub-id-type="pmid">19581415</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0026">
<label>26</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0026">
<string-name>
<surname>Galinier</surname>
<given-names>A</given-names>
</string-name>
,
<string-name>
<surname>Carriere</surname>
<given-names>A</given-names>
</string-name>
,
<string-name>
<surname>Fernandez</surname>
<given-names>Y</given-names>
</string-name>
, et al.
<article-title>Adipose tissue proadipogenic redox changes in obesity</article-title>
.
<source>J Biol Chem</source>
<year>2006</year>
;
<volume>281</volume>
:
<fpage>12682</fpage>
<lpage>12687</lpage>
.
<pub-id pub-id-type="pmid">16377639</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0027">
<label>27</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0027">
<string-name>
<surname>Vincent</surname>
<given-names>AM</given-names>
</string-name>
,
<string-name>
<surname>McLean</surname>
<given-names>LL</given-names>
</string-name>
,
<string-name>
<surname>Backus</surname>
<given-names>C</given-names>
</string-name>
,
<string-name>
<surname>Feldman</surname>
<given-names>EL</given-names>
</string-name>
.
<article-title>Short‐term hyperglycemia produces oxidative damage and apoptosis in neurons</article-title>
.
<source>FASEB J</source>
<year>2005</year>
;
<volume>19</volume>
:
<fpage>638</fpage>
<lpage>640</lpage>
.
<pub-id pub-id-type="pmid">15677696</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0028">
<label>28</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0028">
<string-name>
<surname>Gardner</surname>
<given-names>PR</given-names>
</string-name>
,
<string-name>
<surname>Nguyen</surname>
<given-names>DD</given-names>
</string-name>
,
<string-name>
<surname>White</surname>
<given-names>CW</given-names>
</string-name>
.
<article-title>Aconitase is a sensitive and critical target of oxygen poisoning in cultured mammalian cells and in rat lungs</article-title>
.
<source>Proc Natl Acad Sci USA</source>
<year>1994</year>
;
<volume>91</volume>
:
<fpage>12248</fpage>
<lpage>12252</lpage>
.
<pub-id pub-id-type="pmid">7991614</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0029">
<label>29</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0029">
<string-name>
<surname>Kelly</surname>
<given-names>M</given-names>
</string-name>
,
<string-name>
<surname>Trudel</surname>
<given-names>S</given-names>
</string-name>
,
<string-name>
<surname>Brouillard</surname>
<given-names>F</given-names>
</string-name>
, et al.
<article-title>Cystic fibrosis transmembrane regulator inhibitors CFTR(inh)‐172 and GlyH‐101 target mitochondrial functions, independently of chloride channel inhibition</article-title>
.
<source>J Pharmacol Exp Ther</source>
<year>2010</year>
;
<volume>333</volume>
:
<fpage>60</fpage>
<lpage>69</lpage>
.
<pub-id pub-id-type="pmid">20051483</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0030">
<label>30</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0030">
<string-name>
<surname>Agier</surname>
<given-names>V</given-names>
</string-name>
,
<string-name>
<surname>Oliviero</surname>
<given-names>P</given-names>
</string-name>
,
<string-name>
<surname>Laine</surname>
<given-names>J</given-names>
</string-name>
, et al.
<article-title>Defective mitochondrial fusion, altered respiratory function, and distorted cristae structure in skin fibroblasts with heterozygous OPA1 mutations</article-title>
.
<source>Biochim Biophys Acta</source>
<year>2012</year>
;
<volume>1822</volume>
:
<fpage>1570</fpage>
<lpage>1580</lpage>
.
<pub-id pub-id-type="pmid">22800932</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0031">
<label>31</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0031">
<string-name>
<surname>Van Bergen</surname>
<given-names>NJ</given-names>
</string-name>
,
<string-name>
<surname>Crowston</surname>
<given-names>JG</given-names>
</string-name>
,
<string-name>
<surname>Kearns</surname>
<given-names>LS</given-names>
</string-name>
, et al.
<article-title>Mitochondrial oxidative phosphorylation compensation may preserve vision in patients with OPA1‐linked autosomal dominant optic atrophy</article-title>
.
<source>PLoS One</source>
<year>2011</year>
;
<volume>6</volume>
:
<fpage>e21347</fpage>
.
<pub-id pub-id-type="pmid">21731710</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0032">
<label>32</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0032">
<string-name>
<surname>Chen</surname>
<given-names>C</given-names>
</string-name>
,
<string-name>
<surname>Fossar</surname>
<given-names>N</given-names>
</string-name>
,
<string-name>
<surname>Weil</surname>
<given-names>D</given-names>
</string-name>
, et al.
<article-title>High frequency trans‐splicing in a cell line producing spliced and polyadenylated RNA polymerase I transcripts from an rDNA‐myc chimeric gene</article-title>
.
<source>Nucleic Acids Res</source>
<year>2005</year>
;
<volume>33</volume>
:
<fpage>2332</fpage>
<lpage>2342</lpage>
.
<pub-id pub-id-type="pmid">15849319</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0033">
<label>33</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0033">
<string-name>
<surname>Zhu</surname>
<given-names>H</given-names>
</string-name>
,
<string-name>
<surname>Itoh</surname>
<given-names>K</given-names>
</string-name>
,
<string-name>
<surname>Yamamoto</surname>
<given-names>M</given-names>
</string-name>
, et al.
<article-title>Role of Nrf2 signaling in regulation of antioxidants and phase 2 enzymes in cardiac fibroblasts: protection against reactive oxygen and nitrogen species‐induced cell injury</article-title>
.
<source>FEBS Lett</source>
<year>2005</year>
;
<volume>579</volume>
:
<fpage>3029</fpage>
<lpage>3036</lpage>
.
<pub-id pub-id-type="pmid">15896789</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0034">
<label>34</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0034">
<string-name>
<surname>Berndt</surname>
<given-names>N</given-names>
</string-name>
,
<string-name>
<surname>Holzhutter</surname>
<given-names>HG</given-names>
</string-name>
,
<string-name>
<surname>Bulik</surname>
<given-names>S</given-names>
</string-name>
.
<article-title>Implications of enzyme deficiencies on mitochondrial energy metabolism and reactive oxygen species formation of neurons involved in rotenone‐induced Parkinson's disease: a model‐based analysis</article-title>
.
<source>FEBS J</source>
<year>2013</year>
;
<volume>280</volume>
:
<fpage>5080</fpage>
<lpage>5093</lpage>
.
<pub-id pub-id-type="pmid">23937586</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0035">
<label>35</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0035">
<string-name>
<surname>Freestone</surname>
<given-names>PS</given-names>
</string-name>
,
<string-name>
<surname>Chung</surname>
<given-names>KK</given-names>
</string-name>
,
<string-name>
<surname>Guatteo</surname>
<given-names>E</given-names>
</string-name>
, et al.
<article-title>Acute action of rotenone on nigral dopaminergic neurons–involvement of reactive oxygen species and disruption of Ca2 + homeostasis</article-title>
.
<source>Eur J Neuorsci</source>
<year>2009</year>
;
<volume>30</volume>
:
<fpage>1849</fpage>
<lpage>1859</lpage>
.</mixed-citation>
</ref>
<ref id="acn3305-bib-0036">
<label>36</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0036">
<string-name>
<surname>Lodi</surname>
<given-names>R</given-names>
</string-name>
,
<string-name>
<surname>Tonon</surname>
<given-names>C</given-names>
</string-name>
,
<string-name>
<surname>Valentino</surname>
<given-names>ML</given-names>
</string-name>
, et al.
<article-title>Defective mitochondrial adenosine triphosphate production in skeletal muscle from patients with dominant optic atrophy due to OPA1 mutations</article-title>
.
<source>Arch Neurol</source>
<year>2011</year>
;
<volume>68</volume>
:
<fpage>67</fpage>
<lpage>73</lpage>
.
<pub-id pub-id-type="pmid">20837821</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0037">
<label>37</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0037">
<string-name>
<surname>Piquereau</surname>
<given-names>J</given-names>
</string-name>
,
<string-name>
<surname>Caffin</surname>
<given-names>F</given-names>
</string-name>
,
<string-name>
<surname>Novotova</surname>
<given-names>M</given-names>
</string-name>
, et al.
<article-title>Down‐regulation of OPA1 alters mouse mitochondrial morphology, PTP function, and cardiac adaptation to pressure overload</article-title>
.
<source>Cardiovasc Res</source>
<year>2012</year>
;
<volume>94</volume>
:
<fpage>408</fpage>
<lpage>417</lpage>
.
<pub-id pub-id-type="pmid">22406748</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0038">
<label>38</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0038">
<string-name>
<surname>Darshi</surname>
<given-names>M</given-names>
</string-name>
,
<string-name>
<surname>Mendiola</surname>
<given-names>VL</given-names>
</string-name>
,
<string-name>
<surname>Mackey</surname>
<given-names>MR</given-names>
</string-name>
, et al.
<article-title>ChChd3, an inner mitochondrial membrane protein, is essential for maintaining crista integrity and mitochondrial function</article-title>
.
<source>J Biol Chem</source>
<year>2011</year>
;
<volume>286</volume>
:
<fpage>2918</fpage>
<lpage>2932</lpage>
.
<pub-id pub-id-type="pmid">21081504</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0039">
<label>39</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0039">
<string-name>
<surname>Olichon</surname>
<given-names>A</given-names>
</string-name>
,
<string-name>
<surname>Baricault</surname>
<given-names>L</given-names>
</string-name>
,
<string-name>
<surname>Gas</surname>
<given-names>N</given-names>
</string-name>
, et al.
<article-title>Loss of OPA1 perturbates the mitochondrial inner membrane structure and integrity, leading to cytochrome c release and apoptosis</article-title>
.
<source>J Biol Chem</source>
<year>2003</year>
;
<volume>278</volume>
:
<fpage>7743</fpage>
<lpage>7746</lpage>
.
<pub-id pub-id-type="pmid">12509422</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0040">
<label>40</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0040">
<string-name>
<surname>Griparic</surname>
<given-names>L</given-names>
</string-name>
,
<string-name>
<surname>van der Wel</surname>
<given-names>NN</given-names>
</string-name>
,
<string-name>
<surname>Orozco</surname>
<given-names>IJ</given-names>
</string-name>
, et al.
<article-title>Loss of the intermembrane space protein Mgm1/OPA1 induces swelling and localized constrictions along the lengths of mitochondria</article-title>
.
<source>J Biol Chem</source>
<year>2004</year>
;
<volume>279</volume>
:
<fpage>18792</fpage>
<lpage>18798</lpage>
.
<pub-id pub-id-type="pmid">14970223</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0041">
<label>41</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0041">
<string-name>
<surname>Acin‐Perez</surname>
<given-names>R</given-names>
</string-name>
,
<string-name>
<surname>Bayona‐Bafaluy</surname>
<given-names>MP</given-names>
</string-name>
,
<string-name>
<surname>Fernandez‐Silva</surname>
<given-names>P</given-names>
</string-name>
, et al.
<article-title>Respiratory complex III is required to maintain complex I in mammalian mitochondria</article-title>
.
<source>Mol Cell</source>
<year>2004</year>
;
<volume>13</volume>
:
<fpage>805</fpage>
<lpage>815</lpage>
.
<pub-id pub-id-type="pmid">15053874</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0042">
<label>42</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0042">
<string-name>
<surname>Diaz</surname>
<given-names>F</given-names>
</string-name>
,
<string-name>
<surname>Fukui</surname>
<given-names>H</given-names>
</string-name>
,
<string-name>
<surname>Garcia</surname>
<given-names>S</given-names>
</string-name>
,
<string-name>
<surname>Moraes</surname>
<given-names>CT</given-names>
</string-name>
.
<article-title>Cytochrome c oxidase is required for the assembly/stability of respiratory complex I in mouse fibroblasts</article-title>
.
<source>Mol Cell Biol</source>
<year>2006</year>
;
<volume>26</volume>
:
<fpage>4872</fpage>
<lpage>4881</lpage>
.
<pub-id pub-id-type="pmid">16782876</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0043">
<label>43</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0043">
<string-name>
<surname>Li</surname>
<given-names>Y</given-names>
</string-name>
,
<string-name>
<surname>D'Aurelio</surname>
<given-names>M</given-names>
</string-name>
,
<string-name>
<surname>Deng</surname>
<given-names>JH</given-names>
</string-name>
, et al.
<article-title>An assembled complex IV maintains the stability and activity of complex I in mammalian mitochondria</article-title>
.
<source>J Biol Chem</source>
<year>2007</year>
;
<volume>282</volume>
:
<fpage>17557</fpage>
<lpage>17562</lpage>
.
<pub-id pub-id-type="pmid">17452320</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0044">
<label>44</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0044">
<string-name>
<surname>Acin‐Perez</surname>
<given-names>R</given-names>
</string-name>
,
<string-name>
<surname>Enriquez</surname>
<given-names>JA</given-names>
</string-name>
.
<article-title>The function of the respiratory supercomplexes: the plasticity model</article-title>
.
<source>Biochim Biophys Acta</source>
<year>2014</year>
;
<volume>1837</volume>
:
<fpage>444</fpage>
<lpage>450</lpage>
.
<pub-id pub-id-type="pmid">24368156</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0045">
<label>45</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0045">
<string-name>
<surname>Vartak</surname>
<given-names>R</given-names>
</string-name>
,
<string-name>
<surname>Porras</surname>
<given-names>CA</given-names>
</string-name>
,
<string-name>
<surname>Bai</surname>
<given-names>Y</given-names>
</string-name>
.
<article-title>Respiratory supercomplexes: structure, function and assembly</article-title>
.
<source>Protein Cell</source>
<year>2013</year>
;
<volume>4</volume>
:
<fpage>582</fpage>
<lpage>590</lpage>
.
<pub-id pub-id-type="pmid">23828195</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0046">
<label>46</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0046">
<string-name>
<surname>Lenaz</surname>
<given-names>G</given-names>
</string-name>
,
<string-name>
<surname>Genova</surname>
<given-names>ML</given-names>
</string-name>
.
<article-title>Supramolecular organisation of the mitochondrial respiratory chain: a new challenge for the mechanism and control of oxidative phosphorylation</article-title>
.
<source>Adv Exp Med Biol</source>
<year>2012</year>
;
<volume>748</volume>
:
<fpage>107</fpage>
<lpage>144</lpage>
.
<pub-id pub-id-type="pmid">22729856</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0047">
<label>47</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0047">
<string-name>
<surname>Cogliati</surname>
<given-names>S</given-names>
</string-name>
,
<string-name>
<surname>Frezza</surname>
<given-names>C</given-names>
</string-name>
,
<string-name>
<surname>Soriano</surname>
<given-names>ME</given-names>
</string-name>
, et al.
<article-title>Mitochondrial cristae shape determines respiratory chain supercomplexes assembly and respiratory efficiency</article-title>
.
<source>Cell</source>
<year>2013</year>
;
<volume>155</volume>
:
<fpage>160</fpage>
<lpage>171</lpage>
.
<pub-id pub-id-type="pmid">24055366</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0048">
<label>48</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0048">
<string-name>
<surname>Yarosh</surname>
<given-names>W</given-names>
</string-name>
,
<string-name>
<surname>Monserrate</surname>
<given-names>J</given-names>
</string-name>
,
<string-name>
<surname>Tong</surname>
<given-names>JJ</given-names>
</string-name>
, et al.
<article-title>The molecular mechanisms of OPA1‐mediated optic atrophy in Drosophila model and prospects for antioxidant treatment</article-title>
.
<source>PLoS Genet</source>
<year>2008</year>
;
<volume>4</volume>
:
<fpage>e6</fpage>
.
<pub-id pub-id-type="pmid">18193945</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0049">
<label>49</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0049">
<string-name>
<surname>Ma</surname>
<given-names>Q</given-names>
</string-name>
.
<article-title>Role of nrf2 in oxidative stress and toxicity</article-title>
.
<source>Annu Rev Pharmacol Toxicol</source>
<year>2013</year>
;
<volume>53</volume>
:
<fpage>401</fpage>
<lpage>426</lpage>
.
<pub-id pub-id-type="pmid">23294312</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0050">
<label>50</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0050">
<string-name>
<surname>Paupe</surname>
<given-names>V</given-names>
</string-name>
,
<string-name>
<surname>Dassa</surname>
<given-names>EP</given-names>
</string-name>
,
<string-name>
<surname>Goncalves</surname>
<given-names>S</given-names>
</string-name>
, et al.
<article-title>Impaired nuclear Nrf2 translocation undermines the oxidative stress response in Friedreich ataxia</article-title>
.
<source>PLoS One</source>
<year>2009</year>
;
<volume>4</volume>
:
<fpage>e4253</fpage>
.
<pub-id pub-id-type="pmid">19158945</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0051">
<label>51</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0051">
<string-name>
<surname>Sun</surname>
<given-names>Y</given-names>
</string-name>
,
<string-name>
<surname>Xue</surname>
<given-names>W</given-names>
</string-name>
,
<string-name>
<surname>Song</surname>
<given-names>Z</given-names>
</string-name>
, et al.
<article-title>Restoration of Opa1‐long isoform inhibits retinal injury‐induced neurodegeneration</article-title>
.
<source>J Mol Med (Berl)</source>
.
<year>2015</year>
;
<volume>94</volume>
:
<fpage>335</fpage>
<lpage>346</lpage>
.
<pub-id pub-id-type="pmid">26530815</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0052">
<label>52</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0052">
<string-name>
<surname>Yen</surname>
<given-names>MY</given-names>
</string-name>
,
<string-name>
<surname>Wang</surname>
<given-names>AG</given-names>
</string-name>
,
<string-name>
<surname>Wei</surname>
<given-names>YH</given-names>
</string-name>
.
<article-title>Leber's hereditary optic neuropathy: a multifactorial disease</article-title>
.
<source>Prog Retin Eye Res</source>
<year>2006</year>
;
<volume>25</volume>
:
<fpage>381</fpage>
<lpage>396</lpage>
.
<pub-id pub-id-type="pmid">16829155</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0053">
<label>53</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0053">
<string-name>
<surname>Lee</surname>
<given-names>D</given-names>
</string-name>
,
<string-name>
<surname>Shim</surname>
<given-names>MS</given-names>
</string-name>
,
<string-name>
<surname>Kim</surname>
<given-names>KY</given-names>
</string-name>
, et al.
<article-title>Coenzyme Q10 inhibits glutamate excitotoxicity and oxidative stress‐mediated mitochondrial alteration in a mouse model of glaucoma</article-title>
.
<source>Invest Ophthalmol Vis Sci</source>
<year>2014</year>
;
<volume>55</volume>
:
<fpage>993</fpage>
<lpage>1005</lpage>
.
<pub-id pub-id-type="pmid">24458150</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0054">
<label>54</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0054">
<string-name>
<surname>Yan</surname>
<given-names>MH</given-names>
</string-name>
,
<string-name>
<surname>Wang</surname>
<given-names>X</given-names>
</string-name>
,
<string-name>
<surname>Zhu</surname>
<given-names>X</given-names>
</string-name>
.
<article-title>Mitochondrial defects and oxidative stress in Alzheimer disease and Parkinson disease</article-title>
.
<source>Free Radic Biol Med</source>
<year>2013</year>
;
<volume>62</volume>
:
<fpage>90</fpage>
<lpage>101</lpage>
.
<pub-id pub-id-type="pmid">23200807</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0055">
<label>55</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0055">
<string-name>
<surname>Bennet</surname>
<given-names>D</given-names>
</string-name>
,
<string-name>
<surname>Kim</surname>
<given-names>MG</given-names>
</string-name>
,
<string-name>
<surname>Kim</surname>
<given-names>S</given-names>
</string-name>
.
<article-title>Light‐induced anatomical alterations in retinal cells</article-title>
.
<source>Anal Biochem</source>
<year>2013</year>
;
<volume>436</volume>
:
<fpage>84</fpage>
<lpage>92</lpage>
.
<pub-id pub-id-type="pmid">23376573</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0056">
<label>56</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0056">
<string-name>
<surname>Ji</surname>
<given-names>D</given-names>
</string-name>
,
<string-name>
<surname>Li</surname>
<given-names>GY</given-names>
</string-name>
,
<string-name>
<surname>Osborne</surname>
<given-names>NN</given-names>
</string-name>
.
<article-title>Nicotinamide attenuates retinal ischemia and light insults to neurones</article-title>
.
<source>Neurochem Int</source>
<year>2008</year>
;
<volume>52</volume>
:
<fpage>786</fpage>
<lpage>798</lpage>
.
<pub-id pub-id-type="pmid">17976861</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0057">
<label>57</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0057">
<string-name>
<surname>Kergoat</surname>
<given-names>H</given-names>
</string-name>
,
<string-name>
<surname>Herard</surname>
<given-names>ME</given-names>
</string-name>
,
<string-name>
<surname>Lemay</surname>
<given-names>M</given-names>
</string-name>
.
<article-title>RGC sensitivity to mild systemic hypoxia</article-title>
.
<source>Invest Ophthalmol Vis Sci</source>
<year>2006</year>
;
<volume>47</volume>
:
<fpage>5423</fpage>
<lpage>5427</lpage>
.
<pub-id pub-id-type="pmid">17122132</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0058">
<label>58</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0058">
<string-name>
<surname>Jarrett</surname>
<given-names>SG</given-names>
</string-name>
,
<string-name>
<surname>Lewin</surname>
<given-names>AS</given-names>
</string-name>
,
<string-name>
<surname>Boulton</surname>
<given-names>ME</given-names>
</string-name>
.
<article-title>The importance of mitochondria in age‐related and inherited eye disorders</article-title>
.
<source>Ophthalmic Res</source>
<year>2010</year>
;
<volume>44</volume>
:
<fpage>179</fpage>
<lpage>190</lpage>
.
<pub-id pub-id-type="pmid">20829642</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0059">
<label>59</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0059">
<string-name>
<surname>Sacca</surname>
<given-names>SC</given-names>
</string-name>
,
<string-name>
<surname>Roszkowska</surname>
<given-names>AM</given-names>
</string-name>
,
<string-name>
<surname>Izzotti</surname>
<given-names>A</given-names>
</string-name>
.
<article-title>Environmental light and endogenous antioxidants as the main determinants of non‐cancer ocular diseases</article-title>
.
<source>Mutat Res</source>
<year>2013</year>
;
<volume>752</volume>
:
<fpage>153</fpage>
<lpage>171</lpage>
.
<pub-id pub-id-type="pmid">23337404</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0060">
<label>60</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0060">
<string-name>
<surname>Qi</surname>
<given-names>X</given-names>
</string-name>
,
<string-name>
<surname>Sun</surname>
<given-names>L</given-names>
</string-name>
,
<string-name>
<surname>Hauswirth</surname>
<given-names>WW</given-names>
</string-name>
, et al.
<article-title>Use of mitochondrial antioxidant defenses for rescue of cells with a Leber hereditary optic neuropathy‐causing mutation</article-title>
.
<source>Arch Ophthalmol</source>
<year>2007</year>
;
<volume>125</volume>
:
<fpage>268</fpage>
<lpage>272</lpage>
.
<pub-id pub-id-type="pmid">17296905</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0061">
<label>61</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0061">
<string-name>
<surname>Bessero</surname>
<given-names>AC</given-names>
</string-name>
,
<string-name>
<surname>Clarke</surname>
<given-names>PG</given-names>
</string-name>
.
<article-title>Neuroprotection for optic nerve disorders</article-title>
.
<source>Curr Opin Neurol</source>
<year>2010</year>
;
<volume>23</volume>
:
<fpage>10</fpage>
<lpage>15</lpage>
.
<pub-id pub-id-type="pmid">19915465</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0062">
<label>62</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0062">
<string-name>
<surname>Son</surname>
<given-names>TG</given-names>
</string-name>
,
<string-name>
<surname>Camandola</surname>
<given-names>S</given-names>
</string-name>
,
<string-name>
<surname>Mattson</surname>
<given-names>MP</given-names>
</string-name>
.
<article-title>Hormetic dietary phytochemicals</article-title>
.
<source>NeuroMol Med</source>
<year>2008</year>
;
<volume>10</volume>
:
<fpage>236</fpage>
<lpage>246</lpage>
.</mixed-citation>
</ref>
<ref id="acn3305-bib-0063">
<label>63</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0063">
<string-name>
<surname>Zhang</surname>
<given-names>Q</given-names>
</string-name>
,
<string-name>
<surname>Pi</surname>
<given-names>J</given-names>
</string-name>
,
<string-name>
<surname>Woods</surname>
<given-names>CG</given-names>
</string-name>
,
<string-name>
<surname>Andersen</surname>
<given-names>ME</given-names>
</string-name>
.
<article-title>A systems biology perspective on Nrf2‐mediated antioxidant response</article-title>
.
<source>Toxicol Appl Pharmacol</source>
<year>2010</year>
;
<volume>244</volume>
:
<fpage>84</fpage>
<lpage>97</lpage>
.
<pub-id pub-id-type="pmid">19716833</pub-id>
</mixed-citation>
</ref>
<ref id="acn3305-bib-0064">
<label>64</label>
<mixed-citation publication-type="journal" id="acn3305-cit-0064">
<string-name>
<surname>Barboni</surname>
<given-names>P</given-names>
</string-name>
,
<string-name>
<surname>Valentino</surname>
<given-names>ML</given-names>
</string-name>
,
<string-name>
<surname>La Morgia</surname>
<given-names>C</given-names>
</string-name>
, et al.
<article-title>Idebenone treatment in patients with OPA1‐mutant dominant optic atrophy</article-title>
.
<source>Brain</source>
<year>2013</year>
;
<volume>136</volume>
(
<issue>Pt 2</issue>
):
<fpage>e231</fpage>
.
<pub-id pub-id-type="pmid">23388408</pub-id>
</mixed-citation>
</ref>
</ref-list>
</back>
</pmc>
</record>

Pour manipuler ce document sous Unix (Dilib)

EXPLOR_STEP=$WICRI_ROOT/Wicri/Sante/explor/ParkinsonFranceV1/Data/Pmc/Corpus
HfdSelect -h $EXPLOR_STEP/biblio.hfd -nk 000687 | SxmlIndent | more

Ou

HfdSelect -h $EXPLOR_AREA/Data/Pmc/Corpus/biblio.hfd -nk 000687 | SxmlIndent | more

Pour mettre un lien sur cette page dans le réseau Wicri

{{Explor lien
   |wiki=    Wicri/Sante
   |area=    ParkinsonFranceV1
   |flux=    Pmc
   |étape=   Corpus
   |type=    RBID
   |clé=     PMC:4891995
   |texte=   Loss of functional OPA1 unbalances redox state: implications in dominant optic atrophy pathogenesis
}}

Pour générer des pages wiki

HfdIndexSelect -h $EXPLOR_AREA/Data/Pmc/Corpus/RBID.i   -Sk "pubmed:27547769" \
       | HfdSelect -Kh $EXPLOR_AREA/Data/Pmc/Corpus/biblio.hfd   \
       | NlmPubMed2Wicri -a ParkinsonFranceV1 

Wicri

This area was generated with Dilib version V0.6.29.
Data generation: Wed May 17 19:46:39 2017. Site generation: Mon Mar 4 15:48:15 2024