You are on page 1of 30

Journal Pre-proof

Mitochondrial dysfunction in neurons in Friedreich's ataxia

Anna Stepanova, Jordi Magrané

PII: S1044-7431(19)30196-4
DOI: https://doi.org/10.1016/j.mcn.2019.103419
Reference: YMCNE 103419

To appear in: Molecular and Cellular Neuroscience

Received date: 22 June 2019


Revised date: 5 November 2019
Accepted date: 8 November 2019

Please cite this article as: A. Stepanova and J. Magrané, Mitochondrial dysfunction
in neurons in Friedreich's ataxia, Molecular and Cellular Neuroscience (2018),
https://doi.org/10.1016/j.mcn.2019.103419

This is a PDF file of an article that has undergone enhancements after acceptance, such
as the addition of a cover page and metadata, and formatting for readability, but it is
not yet the definitive version of record. This version will undergo additional copyediting,
typesetting and review before it is published in its final form, but we are providing this
version to give early visibility of the article. Please note that, during the production
process, errors may be discovered which could affect the content, and all legal disclaimers
that apply to the journal pertain.

© 2018 Published by Elsevier.


Journal Pre-proof

Mitochondrial dysfunction in neurons in Friedreich’s ataxia

Anna Stepanovaa and Jordi Magranéb,*

a
Department of Pediatrics, Columbia University Medical Center, New York, NY; Email:
aas2337@cumc.columbia.edu
b
Feil Family Brain and Mind Research Institute, Weill Cornell Medicine, New York, NY;
Email: jom2025@med.cornell.edu
* Corresponding author: 407 East 61st Street, Room RR-509; New York, NY 10065;

of
Phone: 646-962-8174.

ro
Abstract
-p
Friedreich’s ataxia is a multisystemic genetic disorder within the family of
re
mitochondrial diseases that is characterized by reduced levels of the essential
lP

mitochondrial protein frataxin. Based on clinical evidence, the peripheral nervous


system is affected early, neuronal dysfunction progresses towards the central nervous
system, and other organs (such as heart and pancreas) are affected later. However,
na

little attention has been given to the specific aspects of mitochondria function altered by
frataxin depletion in the nervous system. For years, commonly accepted views on
ur

mitochondria dysfunction in Friedreich’s ataxia stemmed from studies using non-


Jo

neuronal systems and may not apply to neurons, which have their own bioenergetic
needs and present a unique, extensive neurite network. Moreover, the basis of the
selective neuronal vulnerability, which primarily affects large sensory neurons in the
dorsal root ganglia, large principal neurons in the dentate nuclei of the cerebellum, and
pyramidal neurons in the cerebral cortex, remains elusive.
In order to identify potential misbeliefs in the field and highlight controversies, we
reviewed current knowledge on frataxin expression in different tissues, discussed the
molecular function of frataxin, and the consequences of its deficiency for mitochondria
structural and functional properties, with a focus on the nervous system.

1
Journal Pre-proof

Keywords: frataxin, sensory neurons, mitochondrial respiratory chain, calcium,


morphology, axonal transport, peripheral nervous system

Abbreviations: FRDA, Friedreich’s ataxia; FXN, human frataxin gene; Fxn, rodent
frataxin gene; DRG, dorsal root ganglia; ER, endoplasmic reticulum.

1. Introduction
Friedreich ataxia (FRDA) is an autosomal recessive disorder, affecting 1:40,000
individuals in the USA alone. It is typically diagnosed in childhood or early adulthood

of
and often progresses into a fatal outcome (Pandolfo, 2008). FRDA is caused by a

ro
trinucleotide guanine-adenine-adenine (GAA) repeat expansion in the first intron of the
frataxin (FXN) gene, which encodes an essential mitochondrial protein (Campuzano et

-p
al., 1997, 1996; Santoro et al., 1999). This repeat expansion impairs gene transcription
(Bidichandani et al., 1998) and results in reduced frataxin protein levels, which cause
re
mitochondrial dysfunction (González-Cabo and Palau, 2013) and is detrimental in the
lP

nervous system, heart, skeleton, and pancreas (Koeppen and Mazurkiewicz, 2013).
Some evidence suggests that the earliest pathology in FRDA may have a
neurodevelopmental component present in combination with atrophy of certain nervous
na

structures (Koeppen et al., 2017a, 2017b).


FRDA patients present a complex set of clinical features throughout the disease,
ur

including ataxia, cardiomyopathy, diabetes mellitus, dysarthria, hearing loss, scoliosis,


Jo

and visual loss (Abrahão et al., 2015; Pandolfo, 2009; Parkinson et al., 2013).
Distinctive characteristics of the FRDA disease phenotype are the progressive limb and
gait ataxia, sensory loss, and the absence of tendon reflexes in lower limbs (areflexia),
symptoms that are consistently observed at early stages of the disease (Harding, 1981;
Stephenson et al., 2015).
All FRDA patients produce less frataxin in their bodies; however, each cell type
presents different vulnerabilities to frataxin levels. Large-caliber myelinated sensory
neurons in the dorsal root ganglia (DRG), principal neurons of the dentate nucleus in
the cerebellum, and pyramidal neurons in the cerebral cortex constitute the primary
sites of nervous system degeneration (Caruso et al., 1987; Koeppen and Mazurkiewicz,

2
Journal Pre-proof

2013). If one argues that frataxin functions should be shared across organisms, tissues,
and even cell types, why else would different neurons show such divergence in their
frataxin-associated pathology, escalating in neuronal dysfunction and death in some
subsets, while others are mainly unaffected? Does this differential vulnerability reflect
differences in energy demands or other frataxin-mediated mitochondrial functions?
Alternatively, do neurons differ in cellular defense mechanisms that protect them from
mitochondrial dysfunction and associated toxicity? One of the challenges we face when
looking for answers to these questions is that many of the pathological changes
associated with frataxin deficits – such as reduced activity of aconitase and oxidative

of
phosphorylation complexes, mitochondrial iron accumulation, increased sensitivity to

ro
oxidative stress, increased mitochondrial DNA mutation rates, as well as the latest
evidence for ferroptotic pathway activation (Bhalla et al., 2016; Cotticelli et al., 2019;

-p
Martelli and Puccio, 2014; Rötig et al., 1997; Vaubel and Isaya, 2013) – have been
observed in yeast, in heart tissue, lymphocytes and fibroblasts from FRDA patients and
re
in mouse models, but not clearly demonstrated in the nervous system. Thus, in this
lP

review, we have critically reviewed the literature on Friedreich’s ataxia, mitochondria,


and the nervous system, and highlighted some contradictions and gaps in our current
knowledge.
na

2. Frataxin expression.
ur

Frataxin is a nuclear-encoded mitochondrial protein highly conserved among


Jo

eukaryotes (Campuzano et al., 1996) and present in certain bacteria (Gibson et al.,
1996). Maturation of frataxin precursor protein occurs within mitochondria and involves
two cleavages by a mitochondrial processing protease (Cavadini et al., 2000;
Schmucker et al., 2008). Once mature, frataxin remains within mitochondria (Babcock et
al., 1997; Campuzano et al., 1997; Koutnikova et al., 1997; Priller et al., 1997; Wilson
and Roof, 1997). A pool of extra-mitochondrial frataxin has been detected in human cell
lines (Acquaviva et al., 2005; Condò et al., 2006) and human erythrocytes (Guo et al.,
2018), but not in mice (Martelli et al., 2007).
The levels of frataxin mRNA differ among tissues in both humans and mice and may
even vary within cell types. Early studies demonstrated that adult expression of frataxin

3
Journal Pre-proof

in mice and humans is highest in the heart and spinal cord, detectable in liver, skeletal
muscle, pancreas, and cerebellum, while absent in the cerebral cortex (Campuzano et
al., 1996; Koutnikova et al., 1997), which mostly follows the sites of malfunction in
FRDA disease.
During mouse embryonic development, frataxin mRNA is first clearly detected at
embryonic day 10.5 (E10.5), and its expression peaks at E14.5 and continues into the
postnatal period. The highest levels are found in the thoracic and lumbar regions of the
spinal cord and DRG, and in non-nervous tissues with high metabolic rate, such as the
heart (Jiralerspong et al., 1997; Koutnikova et al., 1997). This embryonic pattern of

of
expression suggests high demands for frataxin during pre- and early postnatal

ro
development and points towards FRDA being a congenital disease. Indeed, complete
Fxn knockout in mice leads to embryonic lethality at E6.75-E7.5 (Cossée et al., 2000).

-p
Furthermore, evidence in humans suggests that the thinning of the spinal cord, smaller
DRGs, and incomplete entry of sensory axons into the spinal cord observed in FRDA
re
patients may result from an impaired neurodevelopmental process (Koeppen et al.,
lP

2017a, 2017b).
A few studies at a cellular level suggested that certain cell types are more
dependent on frataxin levels than others. For example, frataxin was required for
na

neuronal, but not for cardiomyocyte, differentiation (Santos, 2001). The silencing of Fxn
in the peripheral nervous system, but not in motor neurons, reduced the life span of
ur

adult Drosophila (Anderson et al., 2005). Finally, frataxin deficiency altered proliferation
Jo

and survival of a Schwann cell line, while other neuronal cell lines remained unaffected
(Lu et al., 2009). Despite this evidence, thorough studies comparing the relative levels
of frataxin expression among cellular types, especially within the nervous system, are
lacking.
Finally, the tissue- and cell type-specific pathology observed in FRDA may
additionally arise from differential expression of frataxin protein isoforms (Gakh et al.,
2010; Guo et al., 2018; Xia et al., 2012), microRNA regulation of frataxin expression
(Bandiera et al., 2013; Mahishi et al., 2012), or even be consequence of somatic
instability – that is, accumulation of longer trinucleotide GAA repeat expansions that

4
Journal Pre-proof

result in lower frataxin levels (Santoro et al., 1999) – in certain cell types, such as DRG
neurons (De Biase et al., 2007).

3. Functions of frataxin
In the first decade after the identification of the FXN gene in 1996 (Campuzano et
al., 1996), studies in yeast yielded valuable insights into the function of the protein and
the consequences of its deficiency, and it became evident that frataxin was involved in
the formation of iron-sulfur clusters (Foury and Cazzalini, 1997; Rötig et al., 1997). The
latter is a multi-step pathway requiring a protein complex to facilitate three sequential

of
stages: (1) the mobilization sulfur atoms from L-cysteine, (2) the incorporation of iron

ro
atoms, and (3) the transfer of an assembled iron-sulfur cluster to a carrier (a more
detailed overview can be found elsewhere (Das et al., 2019; Patra and Barondeau,

-p
2019)). While the major pathway of iron-sulfur cluster synthesis is localized in
mitochondria in yeast, an extra-mitochondrial pathway has been described in mammals
re
(Rouault, 2019). Over time the proposed role of frataxin has been changing. Initially,
lP

frataxin was seen as the donor of iron in stage (2) (Yoon and Cowan, 2003). Later
studies shifted attention towards the involvement of frataxin in stage (1) of iron-sulfur
cluster assembly pathway, unambiguously proving that frataxin activates a key enzyme
na

of this stage – cysteine desulfurase (Bridwell-Rabb et al., 2014; Maio and Rouault,
2015). Furthermore, a more recent study presented for the first time the structure of the
ur

human frataxin-bound iron-sulfur cluster assembly complex (Fox et al., 2019), which in
Jo

addition to a more functional study (Patra and Barondeau, 2019) uncovers the
molecular mechanism of the frataxin participation in the very initial steps of iron-sulfur
cluster formation, disregarding its role related to iron. Yet again the role of human
frataxin has been challenged by the same group (Das et al., 2019), which showed a
possibility for frataxin to play a part in the transfer of the iron-sulfur cluster to a carrier (in
this case, glutaredoxin) in stage (3). Thus, more than two decades after its discovery,
the exact role of frataxin remains not fully understood.
What could the functional meaning of the existence of different isoforms of frataxin
be? Extra-mitochondrial isoforms of frataxin, found only in humans so far, were
proposed to have various functions related to iron homeostasis in the cytosol, such as

5
Journal Pre-proof

formation (Guo et al., 2018; Kim et al., 2018) and regeneration (Xia et al., 2012) of iron-
sulfur clusters in a mitochondria-independent way. In contrast to bacterial and yeast
frataxin homologs, which can both store and deliver iron in a manner dependent on the
oligomerization of the protein, human frataxin exists in different isoforms, which need to
be orchestrated in order to cover different steps of iron homeostasis (Ahlgren et al.,
2017). One study (Xia et al., 2012) reported the differential expression of frataxin
isoforms in FRDA affected tissues and showed enrichment of frataxin isoforms with the
exon 1B in DRG, spinal cord, and cerebellum. Moreover, the expression level of this
frataxin isoform was affected to a greater extent in the cerebellum of FRDA patients. If

of
confirmed, the differential expression of frataxin isoforms could shed light on the

ro
susceptibility of the affected tissues to frataxin deficiency.

-p
4. Mitochondrial dysfunction in Friedreich’s ataxia
Based on its known molecular function and preferred mitochondria localization, we
re
would expect that frataxin deficiency results in impairment of proteins containing iron-
lP

sulfur clusters, such as respiratory chain complexes I-III and aconitase. But, does this
assumption hold true in the nervous system? In this section, we will critically review the
reported effects of frataxin insufficiency on mitochondria function with a focus on
na

neurons.
ur

4.1 Mitochondrial respiratory chain activities. There are conflicting reports on


Jo

which mitochondrial enzymatic activities are affected in FRDA. Our knowledge about
the consequences of frataxin reduction on the mitochondrial enzymatic activities stems
from the pioneering work using heart muscle tissues from FRDA patients, which
revealed decreased complexes I-III and aconitase activities (Bradley et al., 2000; Rötig
et al., 1997). While these studies led to the recognition of FRDA as a mitochondrial
disorder, the reported results have led to the impression that there is a universal
mechanism of mitochondria impairment following frataxin depletion. Even though severe
impairment of ATP metabolism was detected in the skeletal muscles of FRDA patients
(Lodi et al., 1999; Vorgerd et al., 2000), only mild, non-significant, differences were
identified in the respiratory chain and aconitase activities in mitochondria (Bradley et al.,

6
Journal Pre-proof

2000). Interestingly, normal activities were found in the cerebellum and DRG of the
same patients (Bradley et al., 2000).
In later studies, the mitochondrial activities were analyzed in other cell types not
directly related to the affected tissues: skin fibroblasts, lymphoblasts, and lymphocytes
(Heidari et al., 2009; Napoli et al., 2006). These and some other – animal – studies
faced the methodological challenges assaying complex I activity. Unless working with
the isolated enzyme, it is essential to demonstrate that the measured activity is specific
to complex I (e.g. by using its inhibitor such as rotenone), which is often omitted in the
studies relying only on NADH-autofluorescence as a measure of complex I activity

of
(Abeti et al., 2018a). Moreover, the activities of complex I with various artificial electron

ro
acceptors (i.e., different from quinone homologs), such as the NADH:ferricyanide
reductase, are not strictly specific to complex I and could not unequivocally reflect

-p
complex I deficiency as claimed (Heidari et al., 2009). Nonetheless, different groups
reported decreased activities of complex I and/or complex II in the cerebellum from the
re
KIKO (Miranda et al., 2002) and the YG8R (Al-Mahdawi et al., 2006)mouse models of
lP

FRDA(Abeti et al., 2016; Lin et al., 2017b), as well as in cerebellar granule neurons
from these mice and NSC34 cells differentiated into motor neurons after Fxn silencing
(Abeti et al., 2018a, 2016; Carletti et al., 2014). Yet, functional impairment of respiratory
na

complexes was not demonstrated in iPSC-derived neuronal progenitor cells (Bird et al.,
2014); however, the authors argue that such a model requires further tuning in order to
ur

challenge the cells to rely on mitochondrial metabolism.


Jo

The activity of aconitase is consistently reduced in various cellular and animal


models of FRDA (Anjomani Virmouni et al., 2015; Cherubini et al., 2015; Li et al., 2015;
Rötig et al., 1997; Sandi et al., 2014; Santoro et al., 2018), including the neuron/cardiac
muscle Fxn-knockout mouse (Puccio et al., 2001). This protein can be a marker for both
impaired iron-sulfur cluster formation (hence, the primary function of frataxin) and
increased oxidative stress due to its “moonlighting” ability to become iron regulatory
protein 1/2 (Cairo and Recalcati, 2007). Of note, there are two isoforms of aconitase –
cytosolic and mitochondrial – and the type of the measured activity (total, cytosolic or
mitochondrial) will depend on the sample preparation and assay conditions. Reported
values for the aconitase activity, unless measured in isolated mitochondria or as in-gel

7
Journal Pre-proof

activity after the electrophoresis (Zhao et al., 2017), should be interpreted as the
mitochondrial plus cytosolic aconitase activities.
A logical direct consequence of respiratory chain dysfunction is the increase
production of radical oxygen species. The significance of oxidative stress in the
development of FRDA, though, remains controversial and is extensively reviewed
elsewhere (Llorens et al., 2019; Lupoli et al., 2018). Oxidative stress was reported in
several non-neuronal laboratory models and patients’ fibroblasts and hearts; however it
was absent in a conditional neuronal Fxn knockout mouse model, which led to claim
that FRDA was a neurodegenerative disease not associated with oxidative damage

of
(Seznec et al., 2005). Moreover, for the past 15-20 years, both oxidative stress markers

ro
and reactive oxygen species probes have been revised because of the development of
new tools and approaches. Lupoli and colleagues (Lupoli et al., 2018) proposed some

-p
explanations for the discrepancies in the field of oxidative stress in FRDA: different time
points selected, different animal/cellular models studied, and different methodologies
re
used. The involvement of oxidative stress or alternative pathways in the
lP

neurodegeneration observed in FRDA in mice and flies has also been a matter of
debate (Chandran et al., 2017; Chen et al., 2016b, 2016a).
Another consequence of impaired mitochondrial respiratory chain is disruption of
na

ATP production, which is a crucial characteristic of mitochondrial disorders. First reports


that revealed impaired ATP production in FRDA heart and skeletal muscles derived
ur

from in vivo human studies using phosphorous nuclear magnetic resonance (Bunse et
Jo

al., 2003; Lodi et al., 1999; Vorgerd et al., 2000), which precisely measures the rate of
mitochondrial ATP production. Studies in cells, however, have yielded heterogenious
results because, depending on the experimental procedures, the reported values of
ATP content cannot reflect the dynamics of ATP synthesis, nor can the source of ATP –
glycolysis or oxidative phosphorylation – be distinguished. Thus, decreased ATP
content was found in fibroblasts and lymphoblasts derived from FRDA patients (Heidari
et al., 2009; Khdour et al., 2018; Li et al., 2015; Zhao et al., 2017). As for
cardiomyocytes, ATP content remained the same in primary rat ventricular myocytes
following Fxn silencing with short-hairpin RNA (Obis et al., 2014), and mitochondrial
ATP synthesis was not altered in iPSC-derived cardiomyocytes from FRDA patients

8
Journal Pre-proof

(Lee et al., 2014). When studying neuronal cell lines, the only reported depletion of ATP
content was shown in human neuroblastoma cells SH-SY5Y after FXN silencing
(Bolinches-Amorós et al., 2014). Therefore, with the currently available data, there is not
enough evidence that energy metabolism is disrupted in neuronal cells.
Keeping all the above in mind, a critical and in-depth exploration of the molecular
links between frataxin deficiency and mitochondrial and neuronal dysfunction is
necessary. Because its role in iron-sulfur cluster synthesis and its mitochondrial
localization, mitochondrial enzymatic activities have been assumed to be always altered
and followed by mitochondria dysfunction. However, as we have explained for the

of
nervous system, the consequences of frataxin deficiency cannot always be related to

ro
deficient mitochondrial enzymatic activities, and therefore we may need to start thinking
about the disease pathogenesis in a different way. For example, frataxin may still have

-p
unknown functions. Moreover, extra-mitochondrial isoforms of frataxin may be
responsible for the deficiency of a wide range of iron-sulfur cluster containing proteins
re
populating cytosolic pathways. And finally, the effects of frataxin deficiency may be
lP

originated at a gene level.

4.2. Calcium homeostasis in neurons. Calcium (Ca2+) not only has a role as an
na

intracellular messenger and regulates the activation of proteases and caspases, but it is
also critical for the generation of nerve action potentials, the regulation of axonal
ur

transport and distribution of mitochondria, and the synaptic activity, among other
Jo

neuronal functions. Thus, some of the neurological features observed in FRDA patients,
such as defects in nerve conduction velocity along sensory axons (Caruso et al., 1987;
Santoro et al., 1990), synaptic abnormalities in the cerebellum (Koeppen et al., 2015),
and neuronal degeneration (Koeppen and Mazurkiewicz, 2013) may be explained by
defects in Ca2+ homeostasis.
One of the earliest and most convincing associations between frataxin and
mitochondrial Ca2+ was obtained with mitochondria isolated from adipocyte-like cells
over-expressing frataxin (Ristow et al., 2000): in this study, frataxin expression activated
mitochondrial respiration, elevated mitochondrial membrane potential, and increased
ATP production; consequently, mitochondrial Ca2+ uptake increased and, potentially,

9
Journal Pre-proof

triggered tricarboxylic acid cycle activity (Ristow et al., 2000). Since this early
publication, though, few additional studies have attempted to address the
consequences of frataxin reduction on Ca2+ homeostasis in FRDA, and none of them
have been using in vivo animal models where alterations in Ca2+ transients reflect
neuronal dysfunction (Bai et al., 2017; Cichon et al., 2017).
In the past decade, three studies have placed Ca2+ homeostasis and signaling in the
center of the pathologic cascade of events observed in FRDA neurons (Mincheva-
Tasheva et al., 2014; Mollá et al., 2017; Purroy et al., 2018). Thus, in isolated
embryonic frataxin-deficient DRG sensory neurons, frataxin reduction led to a rise of

of
free intracellular Ca2+, and this cytosolic Ca2+ increase resulted in the activation of

ro
caspases, increased nitrosation, and activation of pro-apoptotic gene Bax, which
caused cytoskeleton protein alpha-fodrin fragmentation, axonal degeneration, and

-p
apoptosis, respectively (Mincheva-Tasheva et al., 2014) , all three events featured in
FRDA patients. Elevated Ca2+ levels also led to decreased levels of the mitochondrial
re
Na+/Ca2+ exchanger NCLX, which resulted in mitochondrial permeability transition pore
lP

activation and cell death (Purroy et al., 2018). In another study, adult DRG sensory
neurons isolated from the FRDA YG8R mouse model presented an increase in resting
cytosolic Ca2+ that was associated with a decrease in Ca2+ buffering capacity (Mollá et
na

al., 2017), possibly due to mitochondria dysfunction. Proteomic profiling of this cellular
model identified reduced expression of three Ca2+ binding proteins, calmodulin,
ur

calcineurin, and calpain (Mollá et al., 2019). Finally, in cerebellar granule neurons
Jo

isolated from YG8R mice, elevated cytosolic Ca2+ levels were also observed, together
with incomplete cellular Ca2+ buffering (Abeti et al., 2018b), which suggests a common
consequence of frataxin reduction in at least two different cell types. In this latter study,
the authors analyzed other critical Ca2+ stores, such as endoplasmic reticulum (ER),
and found they were also depleted of Ca2+ (Abeti et al., 2018b).
Despite all these research efforts, the regulation of Ca2+ fluxes from and to the
extracellular space and Ca2+ buffering through the ER and mitochondria have not been
studied in detail, and therefore, the relative contribution of mitochondria, ER, and
plasma membrane in Ca2+ homeostasis in FRDA neurons remains poorly understood.
Moreover, although mitochondrial bioenergetics defects seem a logical answer to Ca2+

10
Journal Pre-proof

abnormalities in FRDA, mitochondria Ca2+ transporters and ER-mitochondria


communication cannot be excluded. Finally, with the advance of genetically encoded
Ca2+ probes, some of the data generated using chemical probes may need to undergo a
critical review due to their side effects on cellular viability (Smith et al., 2018).

4.3. Mitochondria morphology and dynamics. It is accepted that disturbances in


mitochondrial homeostasis, such as the ones described in FRDA, compromise normal
cellular functions and often result in neurodegeneration (Area-Gomez et al., 2019).
However, for many years, an essential aspect of mitochondria biology in neurons – that

of
is, neuronal mitochondria form a highly dynamic and interconnected network that
undergoes axonal transport and remodeling by fusion and fission – was overlooked in

ro
many prominent diseases (Schon and Przedborski, 2011). In the case of FRDA, very

-p
few studies, none of which in mammalian systems, have systematically addressed the
involvement of mitochondria transport and dynamics in disease pathogenesis.
re
Appropriate distribution and supply of mitochondria along axons are necessary for
lP

the maintenance of neuronal architecture and activity, including synaptic plasticity and
function (Hollenbeck and Saxton, 2005). ATP/ADP ratio and Ca2+ concentration are
necessary signals that regulate mitochondrial transport (MacAskill et al., 2009; Mironov,
na

2007; Wang and Schwarz, 2009). Therefore, altered mitochondria bioenergetics and
Ca2+ homeostasis in FRDA neurons may impact axonal transport and distribution along
ur

axons and synapses, which could be particularly critical due to the extensive axonal
Jo

network of sensory neurons. The only published work focusing on mitochondrial axonal
transport in FRDA comes from Drosophila larvae studies (Shidara and Hollenbeck,
2010). Reduction of frataxin expression in their nervous system caused early decrease
of mitochondrial membrane potential, followed by defects in mitochondrial transport
(preferentially affecting the retrograde direction), abnormal accumulation of
mitochondria in the synapses, and dying-back neuropathy (Shidara and Hollenbeck,
2010). Unfortunately, no similar in vivo studies in other FRDA models have been
reported, and therefore the exact contribution of mitochondria transport and distribution
in the disease process remains an open question.

11
Journal Pre-proof

Regarding mitochondrial dynamics, a systematic study on genes involved in


mitochondrial fusion and fission (Mfn, Opa1 and Drp1), mitochondrial quality control
(Pink1 and Parkin) and mitochondrial biogenesis (PGC) identified silencing of Mfn
(mitofusin) as a suppressor of the FRDA phenotype in a glia frataxin-deficient
Drosophila model: in this case mitofusin participated in frataxin deficiency pathology by
promoting mitochondria-ER tethering and ER stress (Edenharter et al., 2018), but not
mitochondria fusion. The silencing of genes involved in anterograde mitochondrial
transport (Miro, Milton, and Khc) did not suppress the FRDA phenotype. However, the
genes involved in retrograde mitochondrial transport, which is preferentially affected in

of
neurons of a frataxin-deficient Drosophila model (Shidara and Hollenbeck, 2010) were

ro
not tested.
Mitochondria transport along axons and proper positioning of mitochondria at

-p
synaptic sites are dependent on cytoskeletal components such as tubulin and actin,
respectively. Despite its potential contribution to neuronal demise, studies on
re
cytoskeleton are also limited in FRDA. Thus, actin abnormalities were identified in
lP

FRDA fibroblasts (Pastore et al., 2003; Paupe et al., 2009), whereas depolymerized
tubulin levels increased in FRDA spinal cord autopsy specimens (Sparaco et al., 2009)
and Fxn-silenced NSC34 motor neurons (Piermarini et al., 2016). The only study using
na

cultured DRG sensory neurons revealed accumulations of neurofilaments, microtubules,


mitochondria, and vesicles in axon spheroids in old YG8R mice (Mollá et al., 2017).
ur

Because of its mitochondrial localization, a more direct consequence of frataxin


Jo

depletion may be the existence of abnormalities in mitochondria morphology,


ultrastructure, and number. Indeed, some of the events known to influence them (such
as fusion and fission, oxidative stress, mitochondrial biogenesis, and autophagic
degradation pathways) are abnormal in FRDA models. Thus, disorganized cristae were
observed in the sciatic nerves and DRG of a parvalbumin Fxn-knockout mouse (Piguet
et al., 2018) and in the inducible Fxn-knockdown mouse (Chandran et al., 2017), and
dendrites of Purkinje neurons in the KIKO mouse (Lin et al., 2017b). Moreover,
abnormal accumulation and ultrastructure of mitochondria was observed in FRDA iPSC-
derived cardiomyocytes (Hick et al., 2013) and in the hearts of three Fxn-knockout
mouse models (Chandran et al., 2017; Payne et al., 2011; Puccio et al., 2001; Seznec

12
Journal Pre-proof

et al., 2004), although no molecular mechanisms were proposed. On the contrary,


depletion of mitochondria was observed in the cerebellum of asymptomatic KIKO mice,
another FRDA mouse model, due to the downregulation of mitochondrial biogenesis
pathways (Lin et al., 2017a, 2017b). No abnormalities were identified in FRDA iPSC-
derived neurons (Hick et al., 2013), possibly because the differentiation approach used
did not produce the subset of neurons that are affected in the disease. Surprisingly, only
one report examined patient autopsy material and described an increased number of
mitochondria in the satellite cells of DRG in FRDA autopsies (Koeppen et al., 2014),
which was later confirmed in a larger cohort of patients (Koeppen et al., 2016). Frataxin

of
deficiency caused mitochondrial fragmentation in yeast (around 30% of cells presented

ro
fragmented or intermediate phenotypes), which was exacerbated by oxidative stress
(Lefevre et al., 2012); in the same study, however, no changes were observed in the

-p
mitochondrial network of fibroblasts from FRDA patients, unless challenged with
oxidative stress (Lefevre et al., 2012). Shorter mitochondria were also observed in
re
sensory neurons isolated from old YG8R mice (Mollá et al., 2017), in dendrites of
lP

Purkinje neurons from young KIKO mice (Lin et al., 2017b), but not in glial cells
(Edenharter et al., 2018) or motor axons (Shidara and Hollenbeck, 2010) of two frataxin
deficient Drosophila models. Only one report described longer mitochondria upon
na

frataxin reduction (Bolinches-Amorós et al., 2014), although the cells used were non-
differentiated SH-SY5Y cells (that is, no neurons were present). Therefore, the
ur

heterogeneity of all the results presented above clearly demonstrates that, not only it is
Jo

crucial to choose the relevant model for analysis (cells or structures known to be
affected in FRDA, such as cardiomyocytes, sensory neurons, cerebellum, versus non-
affected, such as motor neurons), but also that we need to be extremely cautious in
generalizing conclusions regarding frataxin deficiency.

5. Conclusions
Over the years, research in FRDA has used a wide variety of model organisms
(yeast, Drosophila, C.elegans, mice) and multiple cell models (primary cultured cells,
iPSC, stable cell lines). Moreover, different approaches to reduce frataxin levels (from
complete knockouts to either acute or chronic reduction of frataxin) have been

13
Journal Pre-proof

attempted to model the disease. While in many cases valuable information has been
provided, the heterogeneity of these models (regarding mitochondrial function, energy
sources, and metabolism) has also resulted in some contradictory observations and
unclear mechanistic explanations for the consequences of frataxin reduction in FRDA.
The evidence provided in this review points towards the need to place research
efforts on relevant neuronal types, when modeling neurodegeneration in FRDA, to
investigate mitochondria in their proper cellular context. Whenever possible,
experimental observations should be confirmed in animal models, which additionally
allow longitudinal studies and the study of neuron interactions with the environment

of
during disease pathogenesis.

ro
Competing interests’ statement: The authors have no competing interests to declare.

-p
Funding: This work was supported by the Friedreich’s Ataxia Research Alliance
re
(FARA).
lP

Acknowledgements: We would like to thank Ilana Seror and Juan Carlos Baiges
Salvadó for critically reading the manuscript.
na

References
ur

Abeti, R., Baccaro, A., Esteras, N., Giunti, P., 2018a. Novel Nrf2-Inducer Prevents
Jo

Mitochondrial Defects and Oxidative Stress in Friedreich’s Ataxia Models. Front.


Cell. Neurosci. 12, 188. https://doi.org/10.3389/fncel.2018.00188

Abeti, R., Brown, A.F., Maiolino, M., Patel, S., Giunti, P., 2018b. Calcium Deregulation:
Novel Insights to Understand Friedreich’s Ataxia Pathophysiology. Front. Cell.
Neurosci. 12, 1–13. https://doi.org/10.3389/fncel.2018.00264

Abeti, R., Parkinson, M.H., Hargreaves, I.P., Angelova, P.R., Sandi, C., Pook, M.A.,
Giunti, P., Abramov, A.Y., 2016. Mitochondrial energy imbalance and lipid
peroxidation cause cell death in friedreich’s ataxia. Cell Death Dis. 7, 1–11.
https://doi.org/10.1038/cddis.2016.111

14
Journal Pre-proof

Abrahão, A., Pedroso, J.L., Braga-Neto, P., Bor-Seng-Shu, E., de Carvalho Aguiar, P.,
Barsottini, O.G.P., 2015. Milestones in Friedreich ataxia: more than a century and
still learning. Neurogenetics. https://doi.org/10.1007/s10048-015-0439-z

Acquaviva, F., De Biase, I., Nezi, L., Ruggiero, G., Tatangelo, F., Pisano, C., Monticelli,
A., Garbi, C., Acquaviva, A.M., Cocozza, S., 2005. Extra-mitochondrial localisation
of frataxin and its association with IscU1 during enterocyte-like differentiation of the
human colon adenocarcinoma cell line Caco-2. J. Cell Sci. 118, 3917–3924.
https://doi.org/10.1242/jcs.02516

of
Ahlgren, E.C., Fekry, M., Wiemann, M., Söderberg, C.A., Bernfur, K., Gakh, O.,
Rasmussen, M., Højrup, P., Emanuelsson, C., Isaya, G., Al-Karadaghi, S., 2017.

ro
Iron-induced oligomerization of human FXN81-210 and bacterial CyaY frataxin and
the effect of iron chelators. PLoS One 12.
-p
https://doi.org/10.1371/journal.pone.0188937
re
Al-Mahdawi, S., Pinto, R.M., Varshney, D., Lawrence, L., Lowrie, M.B., Hughes, S.,
lP

Webster, Z., Blake, J., Cooper, J.M., King, R., Pook, M.A., 2006. GAA repeat
expansion mutation mouse models of Friedreich ataxia exhibit oxidative stress
leading to progressive neuronal and cardiac pathology. Genomics 88, 580–590.
na

https://doi.org/10.1016/j.ygeno.2006.06.015
ur

Anderson, P.R., Kirby, K., Hilliker, A.J., Phillips, J.P., 2005. RNAi-mediated suppression
of the mitochondrial iron chaperone, frataxin, in Drosophila. Hum. Mol. Genet. 14,
Jo

3397–3405. https://doi.org/10.1093/hmg/ddi367

Anjomani Virmouni, S., Al-Mahdawi, S., Sandi, C., Yasaei, H., Giunti, P., Slijepcevic, P.,
Pook, M.A., 2015. Identification of telomere dysfunction in Friedreich ataxia. Mol.
Neurodegener. 10. https://doi.org/10.1186/s13024-015-0019-6

Area-Gomez, E., Guardia-Laguarta, C., Schon, E.A., Przedborski, S., 2019.


Mitochondria, OxPhos, and neurodegeneration: cells are not just running out of
gas. J. Clin. Invest. 129, 34–45. https://doi.org/10.1172/JCI120848

Babcock, M., de Silva, D., Oaks, R., Davis-Kaplan, S., Jiralerspong, S., Montermini, L.,
Pandolfo, M., Kaplan, J., 1997. Regulation of mitochondrial iron accumulation by

15
Journal Pre-proof

Yfh1p, a putative homolog of frataxin. Science 276, 1709–1712.


https://doi.org/10.1126/science.276.5319.1709

Bai, Y., Li, M., Zhou, Y., Ma, L., Qiao, Q., Hu, W., Li, W., Wills, Z.P., Gan, W.B., 2017.
Abnormal dendritic calcium activity and synaptic depotentiation occur early in a
mouse model of Alzheimer’s disease. Mol. Neurodegener. 12.
https://doi.org/10.1186/s13024-017-0228-2

Bandiera, S., Cartault, F., Jannot, A.S., Hatem, E., Girard, M., Rifai, L., Loiseau, C.,
Munnich, A., Lyonnet, S., Henrion-Caude, A., 2013. Genetic Variations Creating

of
MicroRNA Target Sites in the FXN 3′-UTR Affect Frataxin Expression in Friedreich
Ataxia. PLoS One 8, e54791. https://doi.org/10.1371/journal.pone.0054791

ro
Bhalla, A.D., Khodadadi-Jamayran, A., Li, Y., Lynch, D.R., Napierala, M., 2016. Deep

Friedreich’s ataxia. Ann. Clin.


-p
sequencing of mitochondrial genomes reveals increased mutation load in
Transl. Neurol. 3, 523–536.
re
https://doi.org/10.1002/acn3.322
lP

Bidichandani, S.I., Ashizawa, T., Patel, P.I., 1998. The GAA triplet-repeat expansion in
friedreich ataxia interferes with transcription and may be associated with an
na

unusual DNA structure. Am. J. Hum. Genet. 62, 111–121.


https://doi.org/10.1086/301680
ur

Bird, M.J., Needham, K., Frazier, A.E., Rooijen, Jorienvan, Leung, J., Hough, S.,
Denham, M., Thornton, M.E., Parish, C.L., Nayagam, B.A., Pera, M., Thorburn,
Jo

D.R., Thompson, L.H., Dottori, M., van Rooijen, Jorien, Leung, J., Hough, S.,
Denham, M., Thornton, M.E., Parish, C.L., Nayagam, B.A., Pera, M., Thorburn,
D.R., Thompson, L.H., Dottori, M., Rooijen, Jorienvan, Leung, J., Hough, S.,
Denham, M., Thornton, M.E., Parish, C.L., Nayagam, B.A., Pera, M., Thorburn,
D.R., Thompson, L.H., Dottori, M., 2014. Functional characterization of Friedreich
ataxia iPS-derived neuronal progenitors and their integration in the adult brain.
PLoS One 9, e101718. https://doi.org/10.1371/journal.pone.0101718

Bolinches-Amorós, A., Mollá, B., Pla-Martín, D., Palau, F., González-Cabo, P., 2014.
Mitochondrial dysfunction induced by frataxin deficiency is associated with cellular

16
Journal Pre-proof

senescence and abnormal calcium metabolism. Front. Cell. Neurosci. 8, 124.


https://doi.org/10.3389/fncel.2014.00124

Bradley, J.L., Blake, J.C., Chamberlain, S., Thomas, P.K., Cooper, J.M., Schapira, A.H.,
2000. Clinical, biochemical and molecular genetic correlations in Friedreich’s
ataxia. Hum. Mol. Genet. 9, 275–282. https://doi.org/10.1093/hmg/9.2.275

Bridwell-Rabb, J., Fox, N.G., Tsai, C.-L., Winn, A.M., Barondeau, D.P., 2014. Human
Frataxin Activates Fe–S Cluster Biosynthesis by Facilitating Sulfur Transfer
Chemistry. Biochemistry 53, 4904–4913. https://doi.org/10.1021/bi500532e

of
Bunse, M., Bit-Avragim, N., Riefflin, A., Perrot, A., Schmidt, O., Kreuz, F.R., Dietz, R.,
Jung, W.I., Josef Osterziel, K., 2003. Cardiac energetics correlates to myocardial

ro
hypertrophy in Friedreich’s ataxia. Ann. Neurol. 53, 121–123.
https://doi.org/10.1002/ana.10419
-p
Cairo, G., Recalcati, S., 2007. Iron-regulatory proteins: Molecular biology and
re
pathophysiological implications. Expert Rev. Mol. Med. 9, 1–13.
lP

https://doi.org/10.1017/S1462399407000531

Campuzano, V., Montermini, L., Lutz, Y., Cova, L., Hindelang, C., Jiralerspong, S.,
na

Trottier, Y., Kish, S.J., Faucheux, B., Trouillas, P., Authier, F.J., Dürr, A., Mandel,
J.L., Vescovi, A., Pandolfo, M., Koenig, M., 1997. Frataxin is reduced in Friedreich
ur

ataxia patients and is associated with mitochondrial membranes. Hum. Mol. Genet.
6, 1771–1780. https://doi.org/10.1093/hmg/6.11.1771
Jo

Campuzano, V., Montermini, L., Molto, M.D., Pianese, L., Cossee, M., Cavalcanti, F.,
Monros, E., Rodius, F., Duclos, F., Monticelli, A., Zara, F., Canizares, J.,
Koutnikova, H., Bidichandani, S.I., Gellera, C., Brice, A., Trouillas, P., De Michele,
G., Filla, A., De Frutos, R., Palau, F., Patel, P.I., Di Donato, S., Mandel, J.-L.,
Cocozza, S., Koenig, M., Pandolfo, M., 1996. Friedreich’s Ataxia: Autosomal
Recessive Disease Caused by an Intronic GAA Triplet Repeat Expansion. Science
(80-. ). 271, 1423–1427. https://doi.org/10.1126/science.271.5254.1423

Carletti, B., Piermarini, E., Tozzi, G., Travaglini, L., Torraco, A., Pastore, A., Sparaco,
M., Petrillo, S., Carrozzo, R., Bertini, E., Piemonte, F., 2014. Frataxin Silencing

17
Journal Pre-proof

Inactivates Mitochondrial Complex I in NSC34 Motoneuronal Cells and Alters


Glutathione Homeostasis. Int. J. Mol. Sci. 15, 5789–5806.
https://doi.org/10.3390/ijms15045789

Caruso, G., Santoro, L., Perretti, A., Massini, R., Pelosi, L., Crisci, C., Ragno, M.,
Campanella, G., Filla, A., 1987. Friedreich’s ataxia: Electrophysiologic and
histologic findings in patients and relatives. Muscle Nerve 10, 503–515.
https://doi.org/10.1002/mus.880100604

Cavadini, P., Adamec, J., Taroni, F., Gakh, O., Isaya, G., 2000. Two-step processing of

of
human frataxin by mitochondrial processing peptidase: Precursor and intermediate
forms are cleaved at different rates. J. Biol. Chem. 275, 41469–41475.

ro
https://doi.org/10.1074/jbc.M006539200

-p
Chandran, V., Gao, K., Swarup, V., Versano, R., Dong, H., Jordan, M.C., Geschwind,
D.H., 2017. Inducible and reversible phenotypes in a novel mouse model of
re
Friedreich’s ataxia. Elife 6, 1–41. https://doi.org/10.7554/eLife.30054
lP

Chen, K., Ho, T.S.-Y., Lin, G., Tan, K.L., Rasband, M.N., Bellen, H.J., 2016a. Loss of
Frataxin activates the iron/sphingolipid/PDK1/Mef2 pathway in mammals. Elife 5.
na

https://doi.org/10.7554/eLife.20732

Chen, K., Lin, G., Haelterman, N.A., Ho, T.S.-Y., Li, T., Li, Z., Duraine, L., Graham,
ur

B.H., Jaiswal, M., Yamamoto, S., Rasband, M.N., Bellen, H.J., 2016b. Loss of
Frataxin induces iron toxicity, sphingolipid synthesis, and Pdk1/Mef2 activation,
Jo

leading to neurodegeneration. Elife 5. https://doi.org/10.7554/eLife.16043

Cherubini, F., Serio, D., Guccini, I., Fortuni, S., Arcuri, G., Condò, I., Rufini, A., Moiz, S.,
Camerini, S., Crescenzi, M., Testi, R., Malisan, F., 2015. Src inhibitors modulate
frataxin protein levels. Hum. Mol. Genet. 24, 4296–4305.
https://doi.org/10.1093/hmg/ddv162

Cichon, J., Blanck, T.J.J., Gan, W.B., Yang, G., 2017. Activation of cortical somatostatin
interneurons prevents the development of neuropathic pain. Nat. Neurosci. 20,
1122–1132. https://doi.org/10.1038/nn.4595

Condò, I., Ventura, N., Malisan, F., Tomassini, B., Testi, R., 2006. A Pool of

18
Journal Pre-proof

Extramitochondrial Frataxin That Promotes Cell Survival. J. Biol. Chem. 281,


16750–16756. https://doi.org/10.1074/jbc.M511960200

Cossée, M., Puccio, H., Gansmuller, A., Koutnikova, H., Dierich, A., LeMeur, M.,
Fischbeck, K., Dollé, P., Koenig, M., 2000. Inactivation of the Friedreich ataxia
mouse gene leads to early embryonic lethality without iron accumulation. Hum. Mol.
Genet. 9, 1219–1226. https://doi.org/10.1093/hmg/9.8.1219

Das, D., Patra, S., Bridwell-Rabb, J., Barondeau, D.P., 2019. Mechanism of frataxin
“bypass” in human iron–sulfur cluster biosynthesis with implications for Friedreich’s

of
ataxia. J. Biol. Chem. 294, 9276–9284. https://doi.org/10.1074/jbc.RA119.007716

De Biase, I., Rasmussen, A., Endres, D., Al-Mahdawi, S., Monticelli, A., Cocozza, S.,

ro
Pook, M., Bidichandani, S.I., 2007. Progressive GAA expansions in dorsal root
ganglia of Friedreich’s
https://doi.org/10.1002/ana.21052
ataxia
-p
patients. Ann. Neurol. 61, 55–60.
re
Edenharter, O., Schneuwly, S., Navarro, J.A., 2018. Mitofusin-dependent ER stress
lP

triggers glial dysfunction and nervous system degeneration in a drosophila model of


friedreich’s ataxia. Front. Mol. Neurosci. 11, 1–25.
na

https://doi.org/10.3389/fnmol.2018.00038

Foury, F., Cazzalini, O., 1997. Deletion of the yeast homologue of the human gene
ur

associated with Friedreich’s ataxia elicits iron accumulation in mitochondria. FEBS


Lett. 411, 373–377. https://doi.org/10.1016/S0014-5793(97)00734-5
Jo

Fox, N.G., Yu, X., Feng, X., Bailey, H.J., Martelli, A., Nabhan, J.F., Strain-Damerell, C.,
Bulawa, C., Yue, W.W., Han, S., 2019. Structure of the human frataxin-bound iron-
sulfur cluster assembly complex provides insight into its activation mechanism. Nat.
Commun. 10, 1–8. https://doi.org/10.1038/s41467-019-09989-y

Gakh, O., Bedekovics, T., Duncan, S.F., Smith IV, D.Y., Berkholz, D.S., Isaya, G.,
Smith, D.Y., Berkholz, D.S., Isaya, G., 2010. Normal and Friedreich Ataxia Cells
Express Different Isoforms of Frataxin with Complementary Roles in Iron-Sulfur
Cluster Assembly. J. Biol. Chem. 285, 38486–38501.
https://doi.org/10.1074/jbc.M110.145144

19
Journal Pre-proof

Gibson, T.J., Koonin, E. V, Musco, G., Pastore, A., Bork, P., 1996. Friedreich’s ataxia
protein: Phylogenetic evidence for mitochondrial dysfunction. Trends Neurosci. 19,
465–468. https://doi.org/10.1016/S0166-2236(96)20054-2

González-Cabo, P., Palau, F., 2013. Mitochondrial pathophysiology in Friedreich’s


ataxia. J. Neurochem. https://doi.org/10.1111/jnc.12303

Cotticelli, M.G., Xia, S., Lin, D., Lee, T., Terrab, L., Wipf, P., Huryn, D.M., Wilson, R.B.,
2019. Ferroptosis as a novel therapeutic target for Friedreich’s ataxia. J.
Pharmacol. Exp. Ther. https://doi.org/10.1124/jpet.118.252759

of
Guo, L., Wang, Q., Weng, L., Hauser, L.A., Strawser, C.J., Mesaros, C., Lynch, D.R.,
Blair, I.A., 2018. Characterization of a new N-terminally acetylated extra-

ro
mitochondrial isoform of frataxin in human erythrocytes. Sci. Rep. 8.
https://doi.org/10.1038/s41598-018-35346-y
-p
Harding, A.E., 1981. Friedreich’s ataxia: A clinical and genetic study of 90 families with
re
an analysis of early diagnostic criteria and intrafamilial clustering of clinical
lP

features. Brain 104, 589–620. https://doi.org/10.1093/brain/104.3.589

Heidari, M.M., Houshmand, M., Hosseinkhani, S., Nafissi, S., Khatami, M., 2009.
na

Complex I and ATP Content Deficiency in Lymphocytes from Friedreich’s Ataxia.


Can. J. Neurol. Sci. / J. Can. des Sci. Neurol. 36, 26–31.
ur

https://doi.org/10.1017/S0317167100006260
Jo

Hick, A., Wattenhofer-Donze, M., Chintawar, S., Tropel, P., Simard, J.P., Vaucamps, N.,
Gall, D., Lambot, L., Andre, C., Reutenauer, L., Rai, M., Teletin, M., Messaddeq,
N., Schiffmann, S.N., Viville, S., Pearson, C.E., Pandolfo, M., Puccio, H., 2013.
Neurons and cardiomyocytes derived from induced pluripotent stem cells as a
model for mitochondrial defects in Friedreich’s ataxia. Dis. Model. Mech. 6, 608–
621. https://doi.org/10.1242/dmm.010900

Hollenbeck, P.J., Saxton, W.M., 2005. The axonal transport of mitochondria. J. Cell Sci.
118, 5411–5419. https://doi.org/10.1242/jcs.02745

Jiralerspong, S., Liu, Y., Montermini, L., Stifani, S., Pandolfo, M., 1997. Frataxin shows
developmentally regulated tissue-specific expression in the mouse embryo.

20
Journal Pre-proof

Neurobiol. Dis. 4, 103–113. https://doi.org/10.1006/nbdi.1997.0139

Khdour, O.M., Bandyopadhyay, I., Visavadiya, N.P., Roy Chowdhury, S., Hecht, S.M.,
2018. Phenothiazine antioxidants increase mitochondrial biogenesis and frataxin
levels in Friedreich’s ataxia cells. Medchemcomm 9, 1491–1501.
https://doi.org/10.1039/c8md00274f

Kim, K.S., Maio, N., Singh, A., Rouault, T.A., 2018. Cytosolic HSC20 integrates de novo
iron–sulfur cluster biogenesis with the CIAO1-mediated transfer to recipients. Hum.
Mol. Genet. 27, 837–852. https://doi.org/10.1093/hmg/ddy004

of
Koeppen, A.H., Becker, A.B., Qian, J., Feustel, P.J., 2017a. Friedreich Ataxia:
Hypoplasia of Spinal Cord and Dorsal Root Ganglia. J. Neuropathol. Exp. Neurol.

ro
76, 101–108. https://doi.org/10.1093/jnen/nlw111

-p
Koeppen, A.H., Becker, A.B., Qian, J., Gelman, B.B., Mazurkiewicz, J.E., 2017b.
Friedreich Ataxia: Developmental Failure of the Dorsal Root Entry Zone. J.
re
Neuropathol. Exp. Neurol. 76, 969–977. https://doi.org/10.1093/jnen/nlx087
lP

Koeppen, A.H., Kuntzsch, E.C., Bjork, S.T., Ramirez, R.L., Mazurkiewicz, J.E., Feustel,
P.J., 2014. Friedreich ataxia: Metal dysmetabolism in dorsal root ganglia. Acta
na

Neuropathol. Commun. 2. https://doi.org/10.1186/2051-5960-1-26

Koeppen, A.H., Mazurkiewicz, J.E., 2013. Friedreich ataxia: Neuropathology revised. J.


ur

Neuropathol. Exp. Neurol. https://doi.org/10.1097/NEN.0b013e31827e5762


Jo

Koeppen, A.H., Ramirez, R.L., Becker, A.B., Feustel, P.J., Mazurkiewicz, J.E., 2015.
Friedreich ataxia: Failure of GABA-ergic and glycinergic synaptic transmission in
the dentate nucleus. J. Neuropathol. Exp. Neurol. 74, 166–176.
https://doi.org/10.1097/NEN.0000000000000160

Koeppen, A.H., Ramirez, R.L., Becker, A.B., Mazurkiewicz, J.E., 2016. Dorsal root
ganglia in Friedreich ataxia: satellite cell proliferation and inflammation. Acta
Neuropathol. Commun. 4, 46. https://doi.org/10.1186/s40478-016-0288-5

Koutnikova, H., Campuzano, V., Foury, F., Dollé, P., Cazzalini, O., Koenig, M., 1997.
Studies of human, mouse and yeast homologues indicate a mitochondrial function
for frataxin. Nat. Genet. 16, 345–351. https://doi.org/10.1038/ng0897-345
21
Journal Pre-proof

Lee, Y.-K., Ho, P.W.-L., Schick, R., Lau, Y.-M., Lai, W.-H., Zhou, T., Li, Y., Ng, K.-M.,
Ho, S.-L., Esteban, M.A., Binah, O., Tse, H.-F., Siu, C.-W., 2014. Modeling of
Friedreich ataxia-related iron overloading cardiomyopathy using patient-specific-
induced pluripotent stem cells. Pflugers Arch. 466, 1831–1844.
https://doi.org/10.1007/s00424-013-1414-x

Lefevre, S., Sliwa, D., Rustin, P., Camadro, J.M., Santos, R., 2012. Oxidative stress
induces mitochondrial fragmentation in frataxin-deficient cells. Biochem. Biophys.
Res. Commun. 418, 336–341. https://doi.org/10.1016/j.bbrc.2012.01.022

of
Li, Y., Polak, U., Bhalla, A.D., Rozwadowska, N., Butler, J.S., Lynch, D.R., Dent, S.Y.R.,
Napierala, M., 2015. Excision of expanded GAA repeats alleviates the molecular

ro
phenotype of friedreich’s ataxia. Mol. Ther. 23, 1055–1065.
https://doi.org/10.1038/mt.2015.41
-p
Lin, H., Magrane, J., Clark, E.M., Halawani, S.M., Warren, N., Rattelle, A., Lynch, D.R.,
re
2017a. Early VGLUT1-specific parallel fiber synaptic deficits and dysregulated
lP

cerebellar circuit in the KIKO mouse model of Friedreich ataxia. DMM Dis. Model.
Mech. 10, 1529–1538. https://doi.org/10.1242/dmm.030049
na

Lin, H., Magrane, J., Rattelle, A., Stepanova, A., Galkin, A., Clark, E.M., Dong, Y.N.,
Halawani, S.M., Lynch, D.R., 2017b. Early cerebellar deficits in mitochondrial
ur

biogenesis and respiratory chain complexes in the KIKO mouse model of Friedreich
ataxia. Dis. Model. Mech. 10, 1343–1352. https://doi.org/10.1242/dmm.030502
Jo

Llorens, J.V., Soriano, S., Calap-Quintana, P., Gonzalez-Cabo, P., Moltó, M.D., 2019.
The Role of Iron in Friedreich’s Ataxia: Insights From Studies in Human Tissues
and Cellular and Animal Models. Front. Neurosci. 13, 75.
https://doi.org/10.3389/fnins.2019.00075

Lodi, R, Cooper, J.M., Bradley, J.L., Manners, D., Styles, P., Taylor, D.J., Schapira,
A.H., 1999. Deficit of in vivo mitochondrial ATP production in patients with
Friedreich ataxia. Proc. Natl. Acad. Sci. U. S. A. 96, 11492–11495.
https://doi.org/10.1073/pnas.96.20.11492

Lu, C., Schoenfeld, R., Shan, Y., Tsai, H.J., Hammock, B., Cortopassi, G., 2009.

22
Journal Pre-proof

Frataxin deficiency induces Schwann cell inflammation and death. Biochim.


Biophys. Acta - Mol. Basis Dis. 1792, 1052–1061.
https://doi.org/10.1016/j.bbadis.2009.07.011

Lupoli, F., Vannocci, T., Longo, G., Niccolai, N., Pastore, A., 2018. The role of oxidative
stress in Friedreich’s ataxia. FEBS Lett. 592, 718–727.
https://doi.org/10.1002/1873-3468.12928

MacAskill, A.F., Rinholm, J.E., Twelvetrees, A.E., Arancibia-Carcamo, I.L., Muir, J.,
Fransson, A., Aspenstrom, P., Attwell, D., Kittler, J.T., 2009. Miro1 Is a Calcium

of
Sensor for Glutamate Receptor-Dependent Localization of Mitochondria at
Synapses. Neuron 61, 541–555. https://doi.org/10.1016/j.neuron.2009.01.030

ro
Mahishi, L.H., Hart, R.P., Lynch, D.R., Ratan, R.R., 2012. miR-886-3p levels are
elevated in Friedreich ataxia.
https://doi.org/10.1523/JNEUROSCI.0059-12.2012
-p J. Neurosci. 32, 9369–9373.
re
Maio, N., Rouault, T.A., 2015. Iron –sulfur cluster biogenesis in mammalian cells: New
lP

insights into the molecular mechanisms of cluster delivery. Biochim. Biophys. Acta -
Mol. Cell Res. 1853, 1493–1512. https://doi.org/10.1016/j.bbamcr.2014.09.009
na

Martelli, A., Puccio, H., 2014. Dysregulation of cellular iron metabolism in Friedreich
ataxia: from primary iron-sulfur cluster deficit to mitochondrial iron accumulation.
ur

Front. Pharmacol. 5. https://doi.org/10.3389/fphar.2014.00130


Jo

Martelli, A., Wattenhofer-Donzé, M., Schmucker, S., Bouvet, S., Reutenauer, L., Puccio,
H., 2007. Frataxin is essential for extramitochondrial Fe–S cluster proteins in
mammalian tissues. Hum. Mol. Genet. 16, 2651–2658.
https://doi.org/10.1093/hmg/ddm163

Mincheva-Tasheva, S., Obis, E., Tamarit, J., Ros, J., 2014. Apoptotic cell death and
altered calcium homeostasis caused by frataxin depletion in dorsal root ganglia
neurons can be prevented by BH4 domain of Bcl-xL protein. Hum. Mol. Genet. 23,
1829–1841. https://doi.org/10.1093/hmg/ddt576

Miranda, C.J., Santos, M.M., Ohshima, K., Smith, J., Li, L., Bunting, M., Cossée, M.,
Koenig, M., Sequeiros, J., Kaplan, J., Pandolfo, M., 2002. Frataxin knockin mouse.

23
Journal Pre-proof

FEBS Lett. 512, 291–297. https://doi.org/10.1016/S0014-5793(02)02251-2

Mironov, S.L., 2007. ADP regulates movements of mitochondria in neurons. Biophys. J.


92, 2944–2952. https://doi.org/10.1529/biophysj.106.092981

Mollá, B., Muñoz-Lasso, D.C., Calap, P., Fernandez-Vilata, A., de la Iglesia-Vaya, M.,
Pallardó, F. V., Moltó, M.D., Palau, F., Gonzalez-Cabo, P., 2019.
Phosphodiesterase Inhibitors Revert Axonal Dystrophy in Friedreich’s Ataxia
Mouse Model. Neurotherapeutics 16, 432–449. https://doi.org/10.1007/s13311-018-
00706-z

of
Mollá, B., Muñoz-Lasso, D.C., Riveiro, F., Bolinches-Amorós, A., Pallardó, F. V.,
Fernandez-Vilata, A., de la Iglesia-Vaya, M., Palau, F., Gonzalez-Cabo, P., 2017.

ro
Reversible axonal dystrophy by calcium modulation in frataxin-deficient sensory
neurons of YG8R mice.
https://doi.org/10.3389/fnmol.2017.00264
-p
Front. Mol. Neurosci. 10, 1–15.
re
Napoli, E., Taroni, F., Cortopassi, G.A., 2006. Frataxin, Iron–Sulfur Clusters, Heme,
lP

ROS, and Aging. Antioxid. Redox Signal. 8, 506–516.


https://doi.org/10.1089/ars.2006.8.506
na

Obis, È., Irazusta, V., Sanchís, D., Ros, J., Tamarit, J., 2014. Frataxin deficiency in
neonatal rat ventricular myocytes targets mitochondria and lipid metabolism. Free
ur

Radic. Biol. Med. https://doi.org/10.1016/j.freeradbiomed.2014.04.016


Jo

Pandolfo, M., 2009. Friedreich ataxia: The clinical picture. J. Neurol.


https://doi.org/10.1007/s00415-009-1002-3

Pandolfo, M., 2008. Friedreich ataxia. Arch. Neurol. 65, 1296–1303.


https://doi.org/10.1001/archneur.65.10.1296

Parkinson, M.H., Boesch, S., Nachbauer, W., Mariotti, C., Giunti, P., 2013. Clinical
features of Friedreich’s ataxia: classical and atypical phenotypes. J. Neurochem.
126 Suppl, 103–117. https://doi.org/10.1111/jnc.12317

Pastore, A., Tozzi, G., Gaeta, L.M., Bertini, E., Serafini, V., Di Cesare, S., Bonetto, V.,
Casoni, F., Carrozzo, R., Federici, G., Piemonte, F., 2003. Actin Glutathionylation
Increases in Fibroblasts of Patients with Friedreich’s Ataxia: A potential role in the
24
Journal Pre-proof

pathogenesis of the disease. J. Biol. Chem. 278, 42588–42595.


https://doi.org/10.1074/jbc.M301872200

Patra, S., Barondeau, D.P., 2019. Mechanism of activation of the human cysteine
desulfurase complex by frataxin. Proc. Natl. Acad. Sci. 116, 19421–19430.
https://doi.org/10.1073/pnas.1909535116

Paupe, V., Dassa, E.P., Goncalves, S., Auchère, F., Lönn, M., Holmgren, A., Rustin, P.,
2009. Impaired nuclear Nrf2 translocation undermines the oxidative stress
response in Friedreich ataxia. PLoS One 4.

of
https://doi.org/10.1371/journal.pone.0004253

Payne, R.M., Pride, P.M., Babbey, C.M., 2011. Cardiomyopathy of Friedreich’s ataxia:

ro
Use of mouse models to understand human disease and guide therapeutic

-p
development. Pediatr. Cardiol. 32, 366–378. https://doi.org/10.1007/s00246-011-
9943-6
re
Piermarini, E., Cartelli, D., Pastore, A., Tozzi, G., Compagnucci, C., Giorda, E.,
lP

D’Amico, J., Petrini, S., Bertini, E., Cappelletti, G., Piemonte, F., 2016. Frataxin
silencing alters microtubule stability in motor neurons: implications for Friedreich’s
na

ataxia. Hum. Mol. Genet. 25, 4288–4301. https://doi.org/10.1093/hmg/ddw260

Piguet, F., de Montigny, C., Vaucamps, N., Reutenauer, L., Eisenmann, A., Puccio, H.,
ur

2018. Rapid and Complete Reversal of Sensory Ataxia by Gene Therapy in a Novel
Model of Friedreich Ataxia. Mol. Ther. 26, 1940–1952.
Jo

https://doi.org/10.1016/j.ymthe.2018.05.006

Priller, J., Scherzer, C.R., Faber, P.W., MacDonald, M.E., Young, A.B., 1997. Frataxin
gene of Friedreich’s ataxia is targeted to mitochondria. Ann. Neurol. 42, 265–269.
https://doi.org/10.1002/ana.410420222

Puccio, H., Simon, D., Cossée, M., Criqui-Filipe, P., Tiziano, F., Melki, J., Hindelang, C.,
Matyas, R., Rustin, P., Koenig, M., 2001. Mouse models for Friedreich ataxia
exhibit cardiomyopathy, sensory nerve defect and Fe-S enzyme deficiency followed
by intramitochondrial iron deposits. Nat. Genet. 27, 181–186.
https://doi.org/10.1038/84818

25
Journal Pre-proof

Purroy, R., Britti, E., Delaspre, F., Tamarit, J., Ros, J., 2018. Mitochondrial pore opening
and loss of Ca2+ exchanger NCLX levels occur after frataxin depletion. Biochim.
Biophys. Acta - Mol. Basis Dis. 1864, 618–631.
https://doi.org/10.1016/j.bbadis.2017.12.005

Ristow, M., Pfister, M.F., Yee, A.J., Schubert, M., Michael, L., Zhang, C.-Y.C.Y., Ueki,
K., Michael, M.D., Lowell, B.B., Kahn, C.R., 2000. Frataxin activates mitochondrial
energy conversion and oxidative phosphorylation. Proc. Natl. Acad. Sci. 97, 12239–
12243. https://doi.org/10.1073/pnas.220403797

of
Rötig, A., de Lonlay, P., Chretien, D., Foury, F., Koenig, M., Sidi, D., Munnich, A.,
Rustin, P., 1997. Aconitase and mitochondrial iron-sulphur protein deficiency in

ro
Friedreich ataxia. Nat. Genet. 17, 215–217. https://doi.org/10.1038/ng1097-215

-p
Rouault, T.A., 2019. The indispensable role of mammalian iron sulfur proteins in
function and regulation of multiple diverse metabolic pathways. BioMetals 32, 343–
re
353. https://doi.org/10.1007/s10534-019-00191-7
lP

Sandi, C., Sandi, M., Jassal, H., Ezzatizadeh, V., Anjomani-Virmouni, S., Al-Mahdawi,
S., Pook, M.A., 2014. Generation and characterisation of Friedreich ataxia YG8R
na

mouse fibroblast and neural stem cell models. PLoS One 9, e89488.
https://doi.org/10.1371/journal.pone.0089488
ur

Santoro, A., Anjomani Virmouni, S., Paradies, E., Villalobos Coa, V.L., Al-Mahdawi, S.,
Khoo, M., Porcelli, V., Vozza, A., Perrone, M., Denora, N., Taroni, F., Merla, G.,
Jo

Palmieri, L., Pook, M.A., Marobbio, C.M.T.T., Virmouni, S.A., Paradies, E., Coa,
V.L.V., Al-Mahdawi, S., Khoo, M., Porcelli, V., Vozza, A., Perrone, M., Denora, N.,
Taroni, F., Merla, G., Palmieri, L., Pook, M.A., Marobbio, C.M.T.T., 2018. Effect of
diazoxide on friedreich ataxia models. Hum. Mol. Genet. 27, 992–1001.
https://doi.org/10.1093/hmg/ddy016

Santoro, L., De Michele, G., Perretti, A., Crisci, C., Cocozza, S., Cavalcanti, F., Ragno,
M., Monticelli, A., Filla, A., Caruso, G., 1999. Relation between trinucleotide GAA
repeat length and sensory neuropathy in Friedreich’s ataxia. J. Neurol. Neurosurg.
Psychiatry 66, 93–96. https://doi.org/10.1136/jnnp.66.1.93

26
Journal Pre-proof

Santoro, L., Perretti, A., Crisci, C., Ragno, M., Massini, R., Filla, A., De Michele, G.,
Caruso, G., 1990. Electrophysiological and histological follow‐up study in 15
Friedreich’s ataxia patients. Muscle Nerve 13, 536–540.
https://doi.org/10.1002/mus.880130610

Santos, M.M., 2001. Frataxin deficiency enhances apoptosis in cells differentiating into
neuroectoderm. Hum. Mol. Genet. 10, 1935–1944.
https://doi.org/10.1093/hmg/10.18.1935

Schmucker, S., Argentini, M., Carelle-Calmels, N., Martelli, A., Puccio, H., 2008. The in

of
vivo mitochondrial two-step maturation of human frataxin. Hum. Mol. Genet. 17,
3521–3531. https://doi.org/10.1093/hmg/ddn244

ro
Schon, E.A., Przedborski, S., 2011. Mitochondria: The Next (Neurode)Generation.

-p
Neuron. https://doi.org/10.1016/j.neuron.2011.06.003

Seznec, H., Simon, D., Bouton, C., Reutenauer, L., Hertzog, A., Golik, P., Procaccio, V.,
re
Patel, M., Drapier, J.-C.C., Koenig, M., Puccio, H., 2005. Friedreich ataxia: The
lP

oxidative stress paradox. Hum. Mol. Genet. 14, 463–474.


https://doi.org/10.1093/hmg/ddi042
na

Seznec, H., Simon, D., Monassier, L., Criqui-Filipe, P., Gansmuller, A., Rustin, P.,
Koening, M., Puccio, H., 2004. Idebenone delays the onset of cardiac functional
ur

alteration without correction of Fe-S enzymes deficit in a mouse model for


Friedreich ataxia. Hum. Mol. Genet. 13, 1017–1024.
Jo

https://doi.org/10.1093/hmg/ddh114

Shidara, Y., Hollenbeck, P.J., 2010. Defects in mitochondrial axonal transport and
membrane potential without increased reactive oxygen species production in a
Drosophila mdel of Friedreich aaxia. J. Neurosci.
https://doi.org/10.1523/JNEUROSCI.0529-10.2010

Smith, N.A., Kress, B.T., Lu, Y., Chandler-Militello, D., Benraiss, A., Nedergaard, M.,
2018. Fluorescent Ca2+ indicators directly inhibit the Na,K-ATPase and disrupt
cellular functions. Sci. Signal. 11. https://doi.org/10.1126/scisignal.aal2039

Sparaco, M., Gaeta, L.M., Santorelli, F.M., Passarelli, C., Tozzi, G., Bertini, E.,

27
Journal Pre-proof

Simonati, A., Scaravilli, F., Taroni, F., Duyckaerts, C., Feleppa, M., Piemonte, F.,
2009. Friedreich’s ataxia: oxidative stress and cytoskeletal abnormalities. J. Neurol.
Sci. 287, 111–118. https://doi.org/10.1016/j.jns.2009.08.052

Stephenson, J., Zesiewicz, T., Gooch, C., Wecker, L., Sullivan, K., Jahan, I., Kim, S.H.,
2015. Gait and balance in adults with Friedreich’s ataxia. Gait Posture 41, 603–
607. https://doi.org/10.1016/j.gaitpost.2015.01.002

Vaubel, R.A., Isaya, G., 2013. Iron-sulfur cluster synthesis, iron homeostasis and
oxidative stress in Friedreich ataxia. Mol. Cell. Neurosci. 55, 50–61.

of
https://doi.org/10.1016/j.mcn.2012.08.003

Vorgerd, M., Schöls, L., Hardt, C., Ristow, M., Epplen, J.T.T., Zange, J., 2000.

ro
Mitochondrial impairment of human muscle in Friedreich ataxia in vivo.

-p
Neuromuscul. Disord. 10, 430–435. https://doi.org/10.1016/s0960-8966(00)00108-5

Wang, X., Schwarz, T.L., 2009. The mechanism of Ca2+ -dependent regulation of
re
kinesin-mediated mitochondrial motility. Cell 136, 163–174.
lP

https://doi.org/10.1016/j.cell.2008.11.046

Wilson, R.B., Roof, D.M., 1997. Respiratory deficiency due to loss of mitochondrial dna
na

in yeast lacking the frataxin homologue. Nat. Genet. 16, 352–357.


https://doi.org/10.1038/ng0897-352
ur

Xia, H., Cao, Y., Dai, X., Marelja, Z., Zhou, D., Mo, R., Al-Mahdawi, S., Pook, M.A.,
Jo

Leimkühler, S., Rouault, T.A., Li, K., 2012. Novel Frataxin Isoforms May Contribute
to the Pathological Mechanism of Friedreich Ataxia. PLoS One 7, e47847.
https://doi.org/10.1371/journal.pone.0047847

Yoon, T., Cowan, J.A., 2003. Iron-sulfur cluster biosynthesis. Characterization of


frataxin as an iron donor for assembly of [2Fe-2S] clusters in ISU-type proteins. J.
Am. Chem. Soc. 125, 6078–6084. https://doi.org/10.1021/ja027967i

Zhao, H., Li, H., Hao, S., Chen, J., Wu, J., Song, C., Zhang, M., Qiao, T., Li, K., 2017.
Peptide SS-31 upregulates frataxin expression and improves the quality of
mitochondria: implications in the treatment of Friedreich ataxia. Sci. Rep. 7, 9840.
https://doi.org/10.1038/s41598-017-10320-2

28
Journal Pre-proof

Highlights
- Basis of selective neuronal vulnerability in Friedreich’s ataxia remains elusive.
- Certain aspects of mitochondria function altered by frataxin depletion in the nervous
system have been overlooked.
- Heterogenicity of models of Friedreich’s ataxia has resulted in contradictory
observations and unclear mechanistic explanations.

of
ro
-p
re
lP
na
ur
Jo

29

You might also like