ArticlePDF AvailableLiterature Review

Mitochondrial CHCHD2 and CHCHD10: Roles in Neurological Diseases and Therapeutic Implications

Authors:

Abstract and Figures

CHCHD2 mutations have been identified in various neurological diseases such as Parkinson’s disease (PD), frontotemporal dementia (FTD), and Alzheimer’s disease (AD). It is also the first mitochondrial gene whose mutations lead to PD. CHCHD10 is a homolog of CHCHD2; similar to CHCHD2, various mutations of CHCHD10 have been identified in a broad spectrum of neurological disorders, including FTD and AD, with a high frequency of CHCHD10 mutations found in motor neuron diseases. Functionally, CHCHD2 and CHCHD10 have been demonstrated to interact with each other in mitochondria. Recent studies link the biological functions of CHCHD2 to the MICOS complex (mitochondrial inner membrane organizing system). Multiple experimental models suggest that CHCHD2 maintains mitochondrial cristae and disease-associated CHCHD2 mutations function in a loss-of-function manner. However, both CHCHD2 and CHCHD10 knockout mouse models appear phenotypically normal, with no obvious mitochondrial defects. Strategies to maintain or enhance mitochondria cristae could provide opportunities to correct the associated cellular defects in disease state and unravel potential novel targets for CHCHD2-linked neurological conditions.
The pathophysiological roles of CHCHD2/CHCHD10 and potential intervening strategies for CHCHD2-related diseases. Various in vivo, such as flies and mice, and in vitro, such as cell lines and hiPSC models, have been used to investigate the pathophysiological role of CHCHD2 and its homologue CHCHD10. Both proteins are located in mitochondria. CHCHD2 and CHCHD10 form homodimer along with CHCHD2-CHCHD10 heterodimer to maintain MICOS. MICOS is a protein complex located in the mitochondrial inner membrane (IM) that determines the number and shape of crista junctions with IM-specific lipids such as cardiolipin. The mammalian MICOS are composed of key components Mitofilin/Mic60 and MINOS1/ Mic10, with other components including CHCHD3 and CHCHD6. CHCHD10 was previously believed to associate with MICOS. CHCHD2 mutation causes the impairment of CHCHD2-CHCHD10 complex, leading to destabilized MICOS and eventually mitochondrial dysfunction including reduced mitochondrial oxygen consumption rate and OXPHOS complexes. Strategies to enhance or stabilize MICOS complex or CHCHD2 could ameliorate or repair the mitochondrial defects caused by CHCHD2 mutations. In line with that, we found the FDA-approved peptidyl drug Elamipretide (MTP-131), which specifically targets the IM lipid cardiolipin, can ameliorate the cristae defects and mitochondrial dysfunction caused by CHCHD2 mutation. Several questions remain on the pathophysiological roles of CHCHD2/CHCHD10, for example: (1) Is there any mouse models on CHCHD2? (2) How do CHCHD2/CHCHD10 work with MICOS to maintain mitochondrial cristae? See the text for detailed description. OM = mitochondrial outer membrane; IMS = mitochondrial internal membrane space; IM = mitochondrial inner membrane; mtDNA = mitochondrial DNA; TFAM = transcriptional factor A, mitochondrial, which is a key activator of mitochondrial transcription as well as a participant in mtDNA replication.
… 
Content may be subject to copyright.
https://doi.org/10.1177/1073858419871214
The Neuroscientist
1 –15
© The Author(s) 2019
Article reuse guidelines:
sagepub.com/journals-permissions
DOI: 10.1177/1073858419871214
journals.sagepub.com/home/nro
Review
Neurodegenerative disorders, comprising of a broad spec-
trum of diseases including Alzheimer’s disease (AD),
Parkinson’s disease (PD), amyotrophic lateral sclerosis
(ALS), and frontotemporal dementia (FTD), present a
major challenge in the aging population. Diverse neurode-
generative disorders display selective dysfunction and
neuron death in different brain regions. One common
pathological hallmark shared by all those disorders is
mitochondrial dysfunction. Whether mitochondrial dys-
function is the cause or the consequence of pathogenic
process has been often debated. Current evidence suggests
that mitochondrial dysfunction contributes to multiple
pathological process of human diseases in that mitochon-
dria participate in almost all aspects of cellular function,
whereas mitochondrial dysfunction also directly causes
multiple inherited diseases and is heavily implicated in
age-related pathology including neurodegeneration.
Mitochondria and Mitochondrial
Genome
Mitochondria are distinct and powerful intracellular
organelles, playing a vital role in generating cellular
energy via oxidative phosphorylation. Mitochondria are
particularly important for energy-intensive, postmitotic
cells such as neurons, cardiac cells, and muscle cells. A
continual supply of energy from mitochondria is critical
to neuron function in that human brain consumes 20% of
the body’s total energy with only ~2% of total body mass.
Apart from energy generation, mitochondria perform
critical metabolic functions, harbor proapoptotic factors,
and are the major sources of damaging reactive oxygen
species.
Mitochondrion is unique in that it is the only cell organ-
elle under the control of both nuclear DNA (nDNA) and
its own mitochondrial DNA (mtDNA). mtDNA replicates
independent of cell cycle, and even replicates in postmi-
totic cells such as neurons. Such feature makes mtDNA
existing in multiple numbers of copies, varying from hun-
dreds to thousands, in different populations of cells.
871214NROXXX10.1177/1073858419871214The NeuroscientistZhou et al.
review-article2019
1Neuroscience Research laboratory, National Neuroscience Institute,
Duke NUS Medical School, Singapore
2Department of Neurology, Singapore General Hospital, Singapore
Corresponding Author:
Eng-King Tan, National Neuroscience Institute, Duke-NUS Medical
School, Singapore.
Email: gnrtek@sgh.com.sg
Mitochondrial CHCHD2 and CHCHD10:
Roles in Neurological Diseases and
Therapeutic Implications
Wei Zhou1, Dongrui Ma2, and Eng-King Tan1,2
Abstract
CHCHD2 mutations have been identified in various neurological diseases such as Parkinson’s disease (PD),
frontotemporal dementia (FTD), and Alzheimer’s disease (AD). It is also the first mitochondrial gene whose mutations
lead to PD. CHCHD10 is a homolog of CHCHD2; similar to CHCHD2, various mutations of CHCHD10 have been
identified in a broad spectrum of neurological disorders, including FTD and AD, with a high frequency of CHCHD10
mutations found in motor neuron diseases. Functionally, CHCHD2 and CHCHD10 have been demonstrated to
interact with each other in mitochondria. Recent studies link the biological functions of CHCHD2 to the MICOS
complex (mitochondrial inner membrane organizing system). Multiple experimental models suggest that CHCHD2
maintains mitochondrial cristae and disease-associated CHCHD2 mutations function in a loss-of-function manner.
However, both CHCHD2 and CHCHD10 knockout mouse models appear phenotypically normal, with no obvious
mitochondrial defects. Strategies to maintain or enhance mitochondria cristae could provide opportunities to correct
the associated cellular defects in disease state and unravel potential novel targets for CHCHD2-linked neurological
conditions.
Keywords
CHCHD2, CHCHD10, neurological disorders, iPSC, animal models
2 The Neuroscientist 00(0)
Human nDNA is 3.3 billion bp, while mtDNA is small and
circular DNA including 16,569 nucleotides. mtDNA con-
tains 37 genes, including 22 tRNAs, 2 rRNAs, and 13
mRNAs encoding 13 electron transport chain proteins.
Due to its small size, mtDNA is unable to produce all
mitochondrial proteins, which is over 1000, needed for its
functionality; thus, mitochondria rely heavily on imported
nuclear gene products. Mitochondrial genome compasses
additional over 1000 nDNA mitochondrial genes. Since
mitochondria involve both nDNA and mtDNA-encoded
proteins, mutations on either nDNA mitochondrial genes
or mtDNA genes that disrupt mitochondrial function lead
to a broad spectrum of clinically and genetically diverse
and chronic human diseases, which can affect any parts of
the body including neurons, muscles, eyes, and livers
(Fig. 1).
Next-generation sequencing enables the discovery of
novel mutations of mitochondrial genes. However, many
pathogenic candidate genes remain uncharacterized,
leaving unmet diagnostic and treatment need for patients
whose defective gene has not yet been established. In
2014, mutations of mitochondrial gene CHCHD10,
encoded by nDNA, were first reported in human patients
with FTD and ALS. In 2015, mutations of mitochondrial
Figure 1. Genetic mutations in mitochondrial genome cause mitochondrial dysfunction, leading to cellular dysfunction
manifested in different organs of patients with diverse pathology. Mitochondria are unique cellular organelles that are responsible
for cellular respiration, apoptosis, ROS generation, and so on. Mitochondria play a vital role in generating cellular energy,
and therefore are particularly important for cells that are highly energy demanding. Mitochondrial function is maintained by
mitochondrial genes that are encoded by either nuclear DNA or mitochondrial DNA. Genetic mutations in mitochondrial
genome lead to mitochondrial dysfunction in different types of cells, tissues, and organs, especially in neurons and neurological
system. In human patients, such manifested defects could be highly variable and presented in various organs or multiple organs
that may not be apparently linked.
Zhou et al. 3
gene CHCHD2, a homologue of CHCHD10 and encoded
by nDNA, were identified in PD patients. Subsequently,
CHCHD2 or CHCHD10 mutations have been reported in
different population of patients with PD, ALS, FTD, and
AD. In particular, CHCHD2 is the first mitochondrial
gene whose mutations lead to PD and this could provide
an opportunity to unravel novel therapeutic targets. Here
we provide a concise review of the clinical associations
between mutations of mitochondrial genes CHCHD2 and
CHCHD10 and a broad spectrum of neurological disor-
ders, discuss the pathophysiological roles of CHCHD2/
CHCHD10 based on various cellular and animal models,
and highlight the potential therapeutic targets for these
disorders.
CHCHD2 and Human Neurological
Disorders
CHCHD2 is a member of a family of proteins containing
CHCH (coiled-coil-helix-coiled-coil-helix) domain (Fig.
2A). CHCH domain is characterized by a pair of cyste-
ines separated by nine amino acids, named CX9C motifs
(Fig. 2B). CHCH domain adopts a CHCH fold stabilized
by two disulfide bonds that are formed by the four cyste-
ines in the twin CX9C motifs (i.e., two pairs of cysteines
each separated by nine residues) (Fig. 2C). The majority
of CHCH domain-containing proteins, such as CHCHD2,
have a single CHCH domain, while the rest, such as
CHCHD5, have two. The family of CHCH domain-con-
taining proteins is evolutionarily conserved with diverse
physiological functions. Many CHCH domain-contain-
ing proteins are localized in the mitochondria (Modjtahedi
and others 2016).
In addition to a C-terminal CHCH domain, CHCHD2
has an N-terminal mitochondrial targeting sequence.
Thirteen disease-related amino acid residues span the
whole CHCHD2 protein, 151 amino acids in length
(illustrated in Fig. 2A; details listed in Table 1). Most of
those residues are conserved in evolution, while some are
changed in invertebrates such as Drosophila (Fig. 2B).
The missense mutations Thr61Ile and Arg145Gln of
CHCHD2 were reported in inherited late-onset autoso-
mal dominant PD cases in a large Japanese family
(Funayama and others 2015). Thr61Ile was later reported
in a Chinese autosomal dominant PD family (Shi and oth-
ers 2016). Ala32Thr, Pro34Leu, and Ile80Val were identi-
fied in PD patients with European ancestry (Jansen and
others 2015). A nonsense heterozygous variant of
CHCHD2 (Gln126X), leading to a truncated protein, was
identified in a German PD patient (age at onset >40
years) (Koschmidder and others 2016). CHCHD2
Arg8His (Ikeda and others 2017) and Ala79Ser (Yang
and others 2019) were identified in sporadic PD cases.
CHCHD2 Val66Met was reported in an Italian patient
with multiple system atrophy (Nicoletti and others 2017).
The variant Pro2Leu was identified in PD, essential
tremor, AD, and FTD (Foo and others 2015; Funayama
and others 2015; Shi and others 2016; Wu and others
2016). The rare missense variants Ser5Arg, Ala32Thr,
and Ser85Arg of CHCHD2 were identified in AD and
FTD (Che and others 2018). Most disease-related
CHCHD2 mutations are heterozygous, with the excep-
tion of Ala71Pro. Homozygous CHCHD2 Ala71Pro,
along with a rare homozygous Asp211Asn mutation of
TOP1MT (mitochondrial topoisomerase I), was reported
in a 30-year-old Caucasian patient with early-onset PD
(Lee and others 2018b).
Pathophysiological Roles of
CHCHD2/CHCHD10 in Cellular
Models
CHCHD2 is located in both mitochondria and nuclei
(Aras and others 2015; Liu and others 2015; Zhou and
others 2018a). Both the N-terminal mitochondria target-
ing sequence and the C-terminal CHCH domain contrib-
ute to CHCHD2’s mitochondrial location. A truncation
mutation, CHCHD2 Q126X, which loses its CHCH
domain, is located in the cytosol (Zhou and others 2018a).
Endogenous CHCHD2 is primarily located in mitochon-
dria (Liu and others 2015; Zhou and others 2018a). Upon
apoptotic stimuli, CHCHD2 loses its mitochondrial local-
ization and seems to be translocated to the nucleus (Liu
and others 2015). In the nucleus, CHCHD2 functions as a
transcription factor. CHCHD2 transcriptionally regulates
the expression of COX (cytochrome c oxidase) subunit 4
isoform 2 (COX4I2) and CHCHD2 itself upon oxidative
or hypoxic stress (Aras and others 2013; Aras and others
2015). CHCHD2 responds to different stressors.
CHCHD2 was upregulated at both the transcript and pro-
tein levels by mitochondrial unfolded protein stress
(mtUPR) caused by the overexpression of truncated mito-
chondrial matrix protein ornithine transcarbamylase
(ΔOTC) in Drosophila (Meng and others 2017).
Since mitochondria play a significant role in neuron
system, there is much interest in understanding
CHCHD2’s mitochondrial role(s). CHCHD2 promotes
mitochondrial function. Knockdown of CHCHD2
reduced the activity of mitochondrial respiratory com-
plex IV and, to some extent, the activity of complex I
(Baughman and others 2009). Knockdown of CHCHD2
reduced mitochondrial oxygen consumption in various
cellular models (Baughman and others 2009; Zhou and
others 2018a). We and others also found that overexpres-
sion of CHCHD2 WT in cell lines did not enhance or
reduce the mitochondrial oxygen consumption rate. We
generated isogenic human embryonic stem cells (hESCs)
harboring heterozygous PD-associated CHCHD2 mutants
4 The Neuroscientist 00(0)
Figure 2. The schematic view of disease-related CHCHD2 mutations in human CHCHD2 protein. (A) Disease-associated
CHCHD2 mutations were labeled on the functional domains of human CHCHD2. There is an N-terminal mitochondrial targeting
sequence and C-terminal CHCH (coiled-coil-helix-coiled-coil-helix) domain in the human CHCHD2 protein. T61I (Thr61Ile),
Q126X (Gln126X), and R145Q (Arg145Gln), whose functions have been experimentally proven, are labeled red. Common risk
variants such as P2L (Pro2Leu), S5R (Ser5Arg), A32T (Ala32Thr), P34L (Pro34Leu), V66M (Val66Met), I80L (Ile80Val), and S85R
(Ser85Arg) are labeled yellow. A71P (Ala71Pro), labeled with *, was reported as a homozygous mutation in an early-onset PD
(continued)
Zhou et al. 5
Q126X or R145Q and highlighted that PD-associated
CHCHD2 Q126X and R145Q led to mitochondrial dys-
function with a reduction in both mitochondrial oxygen
consumption and the abundancy of components of the
mitochondrial respiratory complex (Zhou and others
2018a). A human induced pluripotent stem cell (hiPSC)
line derived from fibroblasts of a patient carrying hetero-
zygous CHCHD2 T61I was reported (Wang and others
2018), although mitochondrial oxygen consumption rates
from either patient-derived fibroblasts or hiPSCs are yet
to be shown. The generation of patient-derived fibro-
blasts or hiPSCs carrying CHCHD2 mutations with
proper control and careful characterization of those
patient-derived lines will provide more evidence.
CHCHD2 inhibits mitochondrial apoptosis.
Downregulation of CHCHD2 significantly increased cel-
lular ROS (Aras and others 2015). Transient knockdown
of CHCHD2 sensitized cells to multiple apoptotic stim-
uli. CHCHD2 interacts with Bcl-xL to regulate Bax local-
ization, activation, and oligomerization. When cells were
challenged by apoptotic stressors, CHCHD2 decreased in
mitochondria, thus losing its binding to Bcl-xL and allow-
ing apoptosis to proceed (Liu and others 2015). CHCHD2
was also shown to bind to cytochrome c along with
MICS1, a member of the Bax inhibitor-1 superfamily
(Meng and others 2017).
CHCHD2 and Mitophagy
Mitochondrial integrity and homeostasis are vital to main-
taining proper mitochondrial function. When terminally
damaged, mitochondria are degraded through the process
of targeted mitochondrial autophagy (mitophagy) to pre-
vent potentially catastrophic consequences. PINK1/Parkin
signaling is responsible for mitochondrial quality control
via mitophagy in the pathogenesis of PD. It has been spec-
ulated that mitochondrial CHCHD2 may be involved in
the process of PINK/Parkin-mediated mitophagy; never-
theless, evidence supporting a direct role of CHCHD2 in
PINK1/Parkin-mediated mitophagy is lacking. In
Drosophila model, loss of CHCHD2 did not further
enhance the mitochondrial defects caused by loss of
PINK1 or Parkin, suggesting that CHCHD2 may play a
role independent of Parkin/PINK1 (Meng and others
2017). On the other hand, PINK1 or Parkin
overexpression decreased the mitochondrial integrity of
dCHCHD2-deficient flies (Meng and others 2017). As
PINK1 or Parkin overexpression could activate mitoph-
agy, these results suggest that PINK1/Parkin recognized
and cleared the dysfunctional mitochondria caused by loss
of CHCHD2 in flies. Furthermore, in vivo evidence of
mitophagy mediated by PINK1/Parkin seems inconsistent
even though PINK1 and Parkin’s role in mitophagy is well
established. A recent study reported no substantial impact
on basal mitophagy in PINK1 or Parkin null flies, even
though these flies exhibit locomotor defects and dopami-
nergic neuron loss (Lee and others 2018a). Studies using
mice deficient in PINK1 also showed that basal mitoph-
agy is unaffected by loss of PINK1 (McWilliams and oth-
ers 2018). However, other investigators reported that
physiological mitophagy increased with age in Drosophila
(Cornelissen and others 2018), and mitophagy was abro-
gated by PINK1 or Parkin deficiency in vivo. Further
studies in mammalian systems may provide further in-
depth insights into the in vivo or in vitro role of mitophagy
as well as CHCHD2’s relationship with mitophagy.
CHCHD2, CHCHD10, and MICOS
Structurally, mitochondria are composed of a matrix sur-
rounded by two layers of lipid-based membranes, the
mitochondrial outer membrane (OM) and the mitochon-
drial inner membrane (IM) (Fig. 3). The matrix contains
mtDNA and enzymes for various metabolic pathways,
including the TCA cycle and fatty acid oxidation. OM,
which separates the mitochondria from the cytosol, is
responsible for communication between mitochondria
and other membranous organelles. IM shows a complex
ultrastructure with multiple cristae protruding into the
matrix (Fig. 3). The cristae provide an extended mem-
brane surface for multiple enzymes and cofactors com-
posing the mitochondrial respiratory chain, named
complexes I to V (complex I, NADH ubiquinone reduc-
tase; complex II, succinate ubiquinone reductase; com-
plex III, ubiquinol cytochrome c reductase; complex IV,
cytochrome c oxidase; complex V, F1FO-ATP synthase).
Therefore, the cristae are essential for OXPHOS enzymes
to produce ATP.
Such delicate mitochondrial cristae are maintained by
protein modulators and specific phospholipids in IM,
case, along with a homozygous TOP1MT Asp211Asn mutation. The remaining mutations were heterozygous in various human
neurodegenerative disorders. (B) Multiple sequence alignment of the CHCHD2 protein from H. sapiens (human), M. musculus
(mouse), R. norvegicus (rat), C. griseus (hamster), D. rerio (zebrafish), and D. melanogaster (fly, CG5010). Cysteines in the CHCH
domain are highlighted in blue; T61, Q126, and R145 residues are highlighted in red; the remaining disease-linked residues are
highlighted in yellow. Two CX9C motifs are labeled. (C) A typical coiled coil-helix-coiled coil helix (CHCH) fold formed by twin
CX9C motifs, in which two disulfide bonds (labeled blue) are formed by the four cysteines. The representative illustration was
adapted from the solution NMR (nuclear magnetic resonance) structure of human CHCHD4/Mia40, a membrane of the CHCH
domain-containing protein family. Protein Data Bank ID: 2L0Y.
Figure 2. (continued)
6 The Neuroscientist 00(0)
Table 1. Summary of Amino Acid Mutations of CHCHD2/CHCHD10 with the Associated Human Neurological Disorders.
Protein Mutation Genotype Associated Diseases Ethnicity Population Reference
CHCHD2 Pro2Leu +/− Sporadic PD Japanese (Funayama and others 2015)
Early-onset and sporadic PD Chinese (Foo and others 2015)
Sporadic PD Chinese (Shi and others 2016)
ET Chinese (Wu and others 2016)
AD, FTD Chinese (Che and others 2018)
Ser5Arg +/− AD, FTD Chinese (Che and others 2018)
Arg8His +/− Sporadic PD Japanese (Ikeda and others 2017)
Ala32Thr +/− PD Western European
ancestry
(Jansen and others 2015)
AD, FTD Chinese (Che and others 2018)
Pro34Leu +/− PD Western European
ancestry
(Jansen and others 2015)
Thr61Ile +/− Late-onset autosomal dominant PD Japanese (Funayama and others 2015)
Autosomal dominant PD Chinese (Shi and others 2016)
Val66Met +/− MSA Italian (Nicoletti and others 2017)
Ala71Pro −/− Early-onset PD, along with a homozygous TOP1MT
Asp211Asn mutation
Caucasian (Lee and others 2018b)
Ala79Ser +/− Sporadic PD Chinese (Yang and others 2019)
Ile80Val +/− PD Western European
ancestry
(Jansen and others 2015)
Ser85Arg +/− AD, FTD Chinese (Che and others 2018)
Gln126X +/− Early-onset PD German (Koschmidder and others 2016)
Arg145Gln +/− Late-onset autosomal dominant PD Japanese (Funayama and others 2015)
CHCHD10 Pro12Ser +/− ALS Spanish (Dols-Icardo and others 2015)
Arg15Leu/Ser +/− R15L-familial ALS German (Müller and others 2014)
R15S with G58R in cis-autosomal dominant
mitochondrial myopathy
Hispanic (Ajroud-Driss and others 2015)
R15L-familial ALS USA (Johnson and others 2014)
R15L-family with a history suggestive of autosomal-
dominant motor neuron disorder
German (Kurzwelly and others 2015)
R15L-sporadic ALS Caucasian (Zhang and others 2015)
His22Tyr +/− Sporadic FTD Chinese (Jiao and others 2016)
Pro23Thr/Ser/
Leu
+/− P23T-FTLD Italian (Zhang and others 2015)
−/− P23S-Sporadic FTD Chinese (Jiao and others 2016)
+/− P23L-Familial ALS Chinese (Shen and others 2017)
Pro24Leu +/− FTD Chinese (Che and others 2017)
Ser30Leu +/− Sporadic PD Chinese (Zhou and others 2018b)
Ala32Asp +/− Sporadic FTD Chinese (Jiao and others 2016)
Pro34Ser +/− FTD-ALS French (Chaussenot and others 2014)
Sporadic ALS Italian (Chio and others 2015)
PD Caucasian (Zhang and others 2015)
AD
Ala35Asp +/− FTLD Italian (Zhang and others 2015)
Sporadic late-onset AD Chinese (Xiao and others 2016)
Val57Glu +/− Sporadic FTD Chinese (Jiao and others 2016)
Gly58Arg +/− G58R with R15S in cis-autosomal dominant
mitochondrial myopathy
Hispanic (Ajroud-Driss and others 2015)
Ser59Leu +/− Late-onset FTD-ALS with cerebellar ataxia and
mitochondrial myopathy
French (Bannwarth and others 2014)
FTD and FTD-ALS Spanish (Chaussenot and others 2014)
Gly66Val +/− Familial ALS Finnish (Müller and others 2014)
Late-onset SMAJ Caucasian (Penttila and others 2015)
Axonal CMT2 European (Auranen and others 2015)
Pro80Leu +/− Sporadic early-onset ALS and muscle mitochondrial
pathology
Italian (Ronchi and others 2015)
ALS Caucasian (Zhang and others 2015)
Gln82X +/− FTD Spanish (Dols-Icardo and others 2015)
Tyr92Cys +/− Sporadic ALS Chinese (Zhou and others 2017)
Gln102His +/− Sporadic ALS Chinese (Zhou and others 2017)
Q108P/X +/− FTD Belgium (Perrone and others 2017)
PD = Parkinson’s disease; ET = essential tremor; AD = Alzheimer’s disease; FTD = frontotemporal dementia; MSA = multiple system atrophy; ALS = amyotrophic
lateral sclerosis; SMAJ = late-onset spinal motor neuropathy.
Zhou et al. 7
Figure 3. The pathophysiological roles of CHCHD2/CHCHD10 and potential intervening strategies for CHCHD2-related
diseases. Various in vivo, such as flies and mice, and in vitro, such as cell lines and hiPSC models, have been used to investigate
the pathophysiological role of CHCHD2 and its homologue CHCHD10. Both proteins are located in mitochondria. CHCHD2
and CHCHD10 form homodimer along with CHCHD2-CHCHD10 heterodimer to maintain MICOS. MICOS is a protein
complex located in the mitochondrial inner membrane (IM) that determines the number and shape of crista junctions with
IM-specific lipids such as cardiolipin. The mammalian MICOS are composed of key components Mitofilin/Mic60 and MINOS1/
Mic10, with other components including CHCHD3 and CHCHD6. CHCHD10 was previously believed to associate with MICOS.
CHCHD2 mutation causes the impairment of CHCHD2-CHCHD10 complex, leading to destabilized MICOS and eventually
mitochondrial dysfunction including reduced mitochondrial oxygen consumption rate and OXPHOS complexes. Strategies to
enhance or stabilize MICOS complex or CHCHD2 could ameliorate or repair the mitochondrial defects caused by CHCHD2
mutations. In line with that, we found the FDA-approved peptidyl drug Elamipretide (MTP-131), which specifically targets the
IM lipid cardiolipin, can ameliorate the cristae defects and mitochondrial dysfunction caused by CHCHD2 mutation. Several
questions remain on the pathophysiological roles of CHCHD2/CHCHD10, for example: (1) Is there any mouse models on
CHCHD2? (2) How do CHCHD2/CHCHD10 work with MICOS to maintain mitochondrial cristae? See the text for detailed
description. OM = mitochondrial outer membrane; IMS = mitochondrial internal membrane space; IM = mitochondrial
inner membrane; mtDNA = mitochondrial DNA; TFAM = transcriptional factor A, mitochondrial, which is a key activator of
mitochondrial transcription as well as a participant in mtDNA replication.
8 The Neuroscientist 00(0)
such as cardiolipin (Ikon and Ryan 2017). A large protein
complex named MICOS (the mitochondrial contact site
and cristae organizing system) is essential for mitochon-
drial membrane architecture and the formation of cristae
junctions (Fig. 3). The MICOS complex is highly con-
served from yeast to human and comprises at least 7 com-
ponents in mammals: Mitofilin/Mic60, CHCHD3/Mic19,
CHCHD6/Mic25, APOO/Mic26, APOOL/Mic27, QIL1/
Mic13, and MINOS1/Mic10. The list of the components
of MICOS complex may grow, as multiple pieces of evi-
dence suggest the presence of unidentified components of
MICOS. These components form two distinct MICOS
subcomplexes marked by the core components Mitofilin/
Mic60 and MINOS1/Mic10. The Mitofilin/Mic60 sub-
complex and MINOS1/Mic10 subcomplex play different
roles in maintaining contact sites between IM and OM
and in the formation of stable cristae junctions, respec-
tively (Wollweber and others 2017). Previous studies
revealed that complete absence or impairment of any
MICOS components caused drastic alterations of mito-
chondrial cristae and ultimately mitochondrial dysfunc-
tion, and the loss of one subunit of MICOS also prevented
the stable accumulation of other MICOS components in
mitochondria (van der Laan and others 2016; Wollweber
and others 2017). Disruption of the components of
MICOS has been associated with numbers of neurologi-
cal disorders (van der Laan and others 2016).
Recent studies link the biological functions of
CHCHD2 to MICOS complex. Such linkage starts from
CHCHD10, a homolog of CHCHD2. CHCHD10, with a
C-terminal CHCH domain, shares 53% protein sequence
identity with CHCHD2 (Fig. 4). It is speculated that both
proteins play similar and/or redundant biological roles;
furthermore, fast-moving research reveals rather contro-
versial conclusions on their roles based on different cel-
lular or animal models.
CHCHD10 mutations on 16 amino acid residues,
across its whole protein sequence, 142 amino acid in
length (illustrated in Fig. 4; details listed in Table 1), have
been identified in a broad spectrum of complicated neu-
rological disorders, including the motor neuron disease
ALS (Bannwarth and others 2014; Chaussenot and others
2014; Chio and others 2015; Dols-Icardo and others
2015; Johnson and others 2014; Kurzwelly and others
2015; Marroquin and others 2016; Müller and others
2014; Perrone and others 2017; Ronchi and others 2015;
Shen and others 2017; Teyssou and others 2016; Zhang
and others 2015; Zhou and others 2017), mitochondrial
myopathy (Ajroud-Driss and others 2015), late-onset spi-
nal motor neuronopathy (SMAJ) (Penttila and others
2015), Charcot-Marie-Tooth disease type 2 (CMT2)
(Auranen and others 2015), cognitive decline similar to
FTD (Bannwarth and others 2014; Chaussenot and others
2014; Che and others 2017; Dols-Icardo and others 2015;
Jiao and others 2016; Perrone and others 2017), PD
(Zhang and others 2015; Zhou and others 2018b), AD
(Xiao and others 2016; Zhang and others 2015), and cer-
ebellar ataxia (Bannwarth and others 2014). With a 53%
sequence identity and 64% sequence similarity (by Blast)
between CHCHD2 and CHCHD10, we found that 11 out
of 13 disease-associated CHCHD2 residues are con-
served in CHCHD10, and 12 out of 16 disease-associated
CHCHD10 residues are conserved in CHCHD2 (Fig.
4C), hinting a comparable pathological role of CHCHD2
and CHCHD10.
Similar to other CHCH domain-containing proteins,
CHCHD10 is a mitochondrial protein. Efforts were made
to investigate the mechanisms of CHCHD10’s patho-
physiological role. In 2016, Genin and colleagues, using
patient-derived fibroblasts and cell lines, reported that
disease-associated CHCHD10 S59L mutations caused
the loss of mitochondrial cristae and disassembly of
MICOS complex. They found that CHCHD10 bound to
Mitofilin/Mic60 in MICOS in HeLa cells and that
CHCHD10 comigrated with Mitofilin/Mic60, CHCHD3/
Mic19, and CHCHD6/Mic25 in human fibroblast lysates
(Genin and others 2016). The authors did not address
whether CHCHD10 S59L has impaired binding to
Mitofilin/Mic60 compared with CHCHD10 wildtype and
how CHCHD10 S59L causes the loss of mitochondrial
cristae.
In 2017, Meng and colleagues reported that CHCHD2
loss in flies caused mildly distorted mitochondrial cristae
in indirect flight muscles (Meng and others 2017). They
found that CHCHD2 bound to itself to form a homodi-
mer. In early 2018, two back-to-back papers reported a
CHCHD2-CHCHD10 interaction underlying CHCHD10-
associated ALS pathogenesis (Burstein and others 2018;
Straub and others 2018). However, authors of both papers
reported the failure to identify a CHCHD10-Mitofilin/
Mic60 interaction either in human fibroblasts (Straub and
others 2018) or in mouse heart mitochondria (Burstein
and others 2018). Similarly, we reported an endogenous
CHCHD2-CHCHD10 interaction but no endogenous
CHCHD10-Mitofilin/Mic60 bindings in dopaminergic
SK-N-SH cells and human brain lysates (Zhou and others
2018a). While the CHCHD2-CHCHD10 interaction has
been corroborated by many studies (Burstein and others
2018; Huang and others 2018; Purandare and others
2018; Straub and others 2018; Zhou and others 2018a),
many investigators were unable to validate the
CHCHD10-Mitofilin/Mic60 interaction. Moving ahead,
several interesting questions remain to be addressed: (1)
Is MICOS the primary target through which CHCHD10
fulfills its biological function? (2) Is MICOS the primary
target through which CHCHD2 fulfills its biological
function? (3) Does the CHCHD2-CHCHD10 complex
play any role in MICOS complex? (4) If the answer to
Zhou et al. 9
Figure 4. The schematic view of disease-related CHCHD10 mutations in human CHCHD10 protein. (A) Disease-associated
CHCHD10 mutants were labeled on the functional domains of the human CHCHD10 protein. There is an N-terminal mitochondrial
targeting sequence and a C-terminal CHCH (coiled-coil-helix-coiled-coil-helix) domain in the human CHCHD10 protein. Mutations/
variants identified in various human neurological disorders are labeled orange. Y92C and Q102C, labeled with *, were found in a
(continued)
10 The Neuroscientist 00(0)
Question 3 is yes, how does the CHCHD2-CHCHD10
complex work with MICOS components? If the answer is
no, how is the CHCHD2-CHCHD10 complex implicated
in the pathogenesis of various neurological disorders?
Regarding CHCHD10 and CHCHD2’s role in MICOS,
Straub and colleagues found that neither CHCHD10 nor
CHCHD2 comigrated with key components of the
MICOS complex, Mitofilin/Mic60 or CHCHD3/Mic19,
by using cell lysates from human fibroblasts (Burstein
and others 2018), similar to Genin and colleagues (Genin
and others 2016); however, these observations were
sharply opposite to those from Genin and colleagues in
2016. Straub and colleagues, therefore, concluded that
CHCHD10 did not associate with MICOS. Separately,
Huang and colleagues generated stable HeLa and
HEK293 cell lines deficient in CHCHD2, CHCHD10, or
both and observed very subtle abnormal mitochondrial
ultrastructure in CHCHD2/CHCHD10 double knockout
cell lines and no impairment of MICOS components in
either CHCHD2 or CHCHD10 knockout cell lines
(Huang and others 2018). Therefore, it was concluded
that both CHCHD2 and CHCHD10 were dispensable for
the MICOS complex. Genin and colleagues further
reported that disease-associated CHCHD10 G66V did
not impair the MICOS complex by using patient-derived
fibroblasts (Genin and others 2018) after they reported
MICOS defects in patient cells carrying CHCHD10 S59L
(Genin and others 2016). This observation complicated
CHCHD10’s role in MICOS as G66 is in proximity with
S59 in CHCHD10 (Fig. 4A). Taken together, multiple
lines of evidence seems to point to a non-MICOS-related
role of CHCHD10, and possibly CHCHD2.
We investigated the mechanism(s) of CHCHD2 muta-
tions associated with PD (Zhou and others 2018a). We
chose dopaminergic SK-N-SH cell line, which resembles
some physiological features of dopaminergic neurons.
We engaged super-resolution microscopy (STED), which
provides a resolution smaller than the wavelength of
light, and found that CHCHD2 showed a unique distribu-
tion pattern along mitochondria that was very similar to
the distribution pattern of Mitofilin/Mic60 revealed pre-
viously by STED. High-resolution imaging showed some
overlapping signals for CHCHD2 and Mitofilin/Mic60,
suggesting that they might not bind to each other. We
generated isogenic lines harboring PD-related CHCHD2
mutants R145Q or Q126X on pluripotent hESCs line H9.
hESCs can be differentiated into neural progenitor cells
(NPCs) and dopaminergic neurons under proper culture
conditions.
By using SK-N-SH, hESCs and NPCs, we found that
CHCHD2 knockdown caused distorted mitochondrial
cristae, destabilized MICOS and resulted in mitochon-
drial dysfunction. The same observations, as well as
reduced CHCHD2 itself, were made in isogenic hESCs
and derived NPCs carrying CHCHD2 R145Q or Q126X.
With these results, we found a CHCHD2-CHCHD10
interaction but not a CHCHD10-Mitofilin/Mic60 interac-
tion. PD-associated CHCHD2 mutants impaired their
binding to CHCHD10. To investigate whether the defects
caused by CHCHD2 mutants were indeed mediated by
CHCHD10, we examined SK-N-SH cells with CHCHD10
transient knockdown. We found that transient knockdown
of CHCHD10 in SK-N-SH caused a destabilized MICOS
complex and distorted mitochondrial ultrastructure.
CHCHD10 knockdown caused the reduction of CHCHD2
and vice versa. We showed colocalization of CHCHD2
and CHCHD10 in mouse brain sections, and both
CHCHD2 and CHCHD10 appeared as “rail-like” clusters
at the rim of mitochondria by super resolution micros-
copy, suggesting that both proteins were in proximity to
MICOS. To identify how CHCHD2 was involved in
MICOS, we tried to determine the endogenous protein-
protein interaction without any crosslinker, but we failed
to find any interaction between CHCHD2 and Mitofilin/
Mic60, MINOS1/Mic10, CHCHD3/Mic19, or CHCHD6/
Mic25. As the current understanding of MICOS compo-
nents is still limited and additional unknown subunits of
the mammalian MICOS have been suggested (van der
Laan and others 2016), we speculate that unidentified
MICOS components may bridge the CHCHD2-
CHCHD10 complex to known MICOS components. For
example, CHCHD2 binds to MICS1, a protein maintain-
ing cristae structure (Meng and others 2017). MICS1 is a
protein yet to be understood in terms of whether MICS1
maintains cristae structure through MICOS and whether
CHCHD10 associates with the MICS1-CHCHD2 com-
plex. Further study is needed to extend the understanding
of MICOS.
Previous studies on CHCHD10 knockdown or knock-
out showed little or mild defects (Burstein and others
longer transcript of human CHCHD10. (B) Multiple sequence alignment of the CHCHD10 protein from H. sapiens (human), M.
musculus (mouse), R. norvegicus (rat), and D. rerio (zebrafish). A potential fly ortholog of human CHCHD10, CG31007, was included
in the alignment and labeled D. melanogaster (fly). Cysteines within the CHCH domain are highlighted in blue; disease-linked residues
are highlighted in red. Two CX9C motifs are labeled. (C) Sequence alignment of the CHCHD2 protein and the CHCHD10
protein from H. sapiens (human). Cysteines within the CHCH domain are highlighted in blue; disease-linked CHCHD2 residues are
highlighted in green; disease-linked CHCHD10 residues are highlighted in red. Conserved residues in CHCHD2 or CHCHD10,
which were reported in neurological disorders, are highlighted in yellow with rectangles.
Figure 4. (continued)
Zhou et al. 11
2018; Huang and others 2018) in MICOS, while ours
showed significant defects in MICOS. One reason could
be cell type-specific effects. For example, dopaminergic
cells could be a more relevant model than HEK293/HeLa
cells. Another reason could be the different gene manipu-
lation strategy. During gene silencing, especially when
the target gene is indispensable, loss of one gene could be
compensated by another gene with overlapping function
to maintain the fitness of living organisms (El-Brolosy
and Stainier 2017). When generating stable lines or
knockout animals, chronic cultured cells could gradually
activate cellular mechanisms to compensate for the lack
of the target gene, while transient knockdown of the tar-
geted gene shows instant cellular responses upon losing
the target gene; therefore, the results of stable gene silenc-
ing may be compromised compared with those of tran-
sient genetic manipulation. With such a scenario, chronic
cellular or animal models harboring knock-in mutants
may be a better choice for phenocopy genetic mutation-
related human diseases.
hiPSC technology allows the isolation of patient-
derived cells that carry all genetic alterations that cause
disease. Those patient-origin cells have a unique advan-
tage to recapitulate disease-relevant phenotypes, provid-
ing an experimental system to study the pathogenesis of
the disease in vitro. However, one major challenging
issue of patient-derived hiPSCs is the variability between
hiPSC lines of a given lineage (Khurana and others 2015).
Such variation between lines is unpredictable and is
believed to be mostly caused by the differences in genetic
backgrounds as well as the reprogramming strategy. In
addition, the process of deriving fibroblast subclones
from parental fibroblasts may also introduce variations,
as both fibroblast subclones and derived clonal hiPSCs
were reported to contain similar numbers of de novo vari-
ants compared with their parental fibroblasts (Kwon and
others 2017). Therefore, to eliminate variations between
lines, it would be better to generate isogenic pairs of dis-
ease-specific and control hiPSCs that differ exclusively at
the disease-causing mutation to define subtle disease-
relevant differences. CRISPR/Cas9 allows relatively easy
genome editing, whereas the specificity of Cas9 and its
off-target effects need to be carefully monitored and ade-
quately addressed in isogenic pairs of hiPSCs.
Animal Models of CHCHD2/
CHCHD10
The discovery of inherited forms of PD led to the genera-
tion of novel genetic animal models of PD. Although flies
are distantly related to humans in evolutionary terms,
many neuronal circuits in the vertebrates and flies appear
to be analogous, with both vertebrates and flies having
dopaminergic neurons and circuits. It is therefore
possible that PD-associated pathogenic mutations may
have similar effects in both flies and humans. Fly models
with manipulated PD genes, such as PINK1, Parkin,
Fbxo7, TRAP1, LRRK2, and α-synuclein, have success-
fully phenocopied many characteristics of PD, including
reduced locomotion, the reduction of dopaminergic neu-
rons in the brain, and mitochondrial and synaptic pheno-
types, suggesting that fly nervous system could provide
unique opportunities to elucidate neurophysiological per-
turbations caused by genetic mutations (West and others
2015).
Tio and colleagues reported transgenic fly models with
overexpressed human CHCHD2 WT, PD-related CHCHD2
mutants T61I, R145Q or risk-variant P2L (Tio and others
2017). They reported that all transgenic fly lines exhibited
PD-related phenotypes such as locomotor dysfunction,
dopaminergic neuron degeneration, lifespan reduction,
mitochondrial dysfunction, and impairment in synaptic
transmission. Among all transgenic lines, CHCHD2 T61I
and R145Q overexpression showed more severe phenotypes
than others. As CHCHD2 generally promotes mitochondrial
function, the toxicity observed in flies with CHCHD2 WT
overexpression may not be caused by CHCHD2 per se.
Meng and colleagues carried out a series of experi-
ments on flies by manipulating Drosophila CHCHD2,
CG5010 (Meng and others 2017). They reported that flies
without dCHCHD2 were largely normal at a younger
stage but had a shorter life span compared to the control
group. The loss of dCHCHD2 led to oxidative stress,
dopaminergic neuron loss, and motor dysfunction with
age. CHCHD2 loss in flies caused distorted mitochon-
drial cristae in the indirect flight muscles and impaired
oxygen respiration in the mitochondria. The authors
found that these PD-associated phenotypes can be res-
cued by the overexpression of the translation inhibitor
4E-BP and by the introduction of human CHCHD2 but
not its PD-associated T61I or R145Q mutants. Their
study on flies provides evidence on CHCHD2’s role in
PD pathogenesis in vivo, and rescue experiments showed
that CHCHD2 T61I and R145Q acted as loss-of-function
mutants. There is currently no fly study on CHCHD10.
An uncharacterized protein CG31007 could be a potential
fly ortholog of CHCHD10. CG31007 shares 43% amino
acid sequence identity and 64% sequence similarity (by
Blast) with human CHCHD10 (Fig. 4B). On the other
hand, Woo and others, showed CHCHD10 exerts a pro-
tective role in mitochondria and synaptic integrity by
using C. elegans model, and they reported FTD/ALS-
related CHCHD10 R15L and S59L exhibit loss-of-func-
tion phenotypes (Woo and others 2017).
Similar to many other mouse models of genetic PD,
which have had limited success in showing disease-
related phenotypes, CHCHD2 knockout mouse models
appear normal. CHCHD2-null mice developed normally
12 The Neuroscientist 00(0)
with no distorted mitochondrial ultrastructure observed
in CHCHD2 null MEFs (mouse embryonic fibroblasts)
(Meng and others 2017). CHCHD10 knockout mice were
also phenotypically normal, with no mitochondrial
defects in the brain, heart, or skeletal muscles (Burstein
and others 2018). Such normal phenotypes could be due
to, at least partially, genetic compensation as discussed
above. In line with that, knock-in mouse models with
disease-associated mutations may recapitulate some
pathological features of human diseases. Recent studies
on transgenic mice harboring CHCHD10 S59L demon-
strated features of motor neuron disease with mitochon-
drial defects and motor-neuron death (Anderson and
others 2019; Genin and others 2019). The in vivo studies
on CHCHD2/CHCHD10 are summarized in Table 2.
Drugs Targeting CHCHD2
Multiple experimental models suggest that CHCHD2
maintains mitochondrial cristae and disease-associated
CHCHD2 mutations function in a loss-of-function man-
ner. Therefore, strategies to maintain or enhance mito-
chondrial cristae as well as CHCHD2 could provide
opportunities to correct CHCHD2 mutations-caused cel-
lular defects (Fig. 3). We sought a potential strategy to
correct defects caused by CHCHD2 mutations using iso-
genic NPCs carrying heterozygous PD-related CHCHD2
R145Q as an in vitro model for PD. By testing various
candidate compounds, we reported that mitochondrial
defects caused by CHCHD2 R145Q can be ameliorated
by Elamipretide/MTP-131 (Zhou and others 2018a).
Elamipretide/MTP-131 (Bendavia) is a cell-permeable
and mitochondria-targeting peptide that protects mito-
chondrial cristae by crossing the mitochondrial outer
membrane and interacting with phospholipid cardiolipin
in the IM. Cardiolipin plays a crucial role in cristae for-
mation, mitochondrial fusion, mtDNA integrity, and oxi-
dative phosphorylation (Ikon and Ryan 2017).
Elamipretide/MTP-131 is currently being tested in clini-
cal trials for ischemia reperfusion injury, mitochondrial
myopathies, and heart failure. The US Food and Drug
Administration has granted fast track designation for
Elamipretide/MTP-131 as a treatment for people with
rare mitochondrial diseases (2017–2018). Considering
the good safety profile and known pharmacokinetics and
pharmacodynamics of Elamipretide/MTP-131, it could
be used as a potential treatment for CHCHD2 mutation-
related neuronal disorders. Further vigorous testing of
Elamipretide/MTP-131 in a proper in vivo model of
CHCHD2 will provide stronger evidence to use this drug
for patients with neurological disorders associated with
CHCHD2 mutations. Moreover, Elamipretide/MTP-131
directly targets cardiolipin, but not CHCHD2. Further
investigation on alternative strategies to enhance/stabi-
lize CHCHD2 directly, thereby to stabilize MICOS, could
be more efficient to correct the defects and benefit rele-
vant patients.
Current Limitations and Challenges
CHCHD2/CHCHD10 are important mitochondrial pro-
teins whose mutations have been associated with various
neurological diseases among different populations.
Several important questions regarding their mechanisms
and relevant models are yet to be answered: (1) Does
CHCHD2/CHCHD10 play any role in mitophagy? (2)
How do CHCHD2/CHCHD10 work with MICOS to
maintain mitochondrial cristae? (3) Are there any dopa-
minergic neurons carrying CHCHD2/CHCHD10 muta-
tions? As mitochondrial function is vital for hiPSCs to
Table 2. Summary of Animal Models on CHCHD2/CHCHD10 Studies.
Animal Models CHCHD2 CHCHD10
Drosophila dCHCHD2 null flies show mitochondrial
defects in muscle, with neuron defects
(Meng and others 2017)
Not availanle
Overexpression of human WT, T61I, R145Q,
P2L in flies all show neuron toxicity (Tio
and others 2017)
C. elegans Not available Loss of C. elegans har-1, the closest orthologue to
mammalian CHCHD10, impairs mitochondria
and movement (Woo and others 2017)
Mouse Knockout Normal development; no mitochondrial
defects in MEFs (Meng and others 2017)
Normal without disease pathology (Burstein and
others 2018)
Knockin Not available CHCHD10S59L/+ mimic the mitochondrial
myopathy with mtDNA instability displayed by
the patients (Anderson and others 2019; Genin
and others 2019)
Zhou et al. 13
differentiate into functional neurons, it could be challeng-
ing to derive dopaminergic neurons from hiPSC or hESC
carrying CHCHD2/CHCHD10 mutations. (4) Are there
any hiPSC lines derived from human patients carrying
CHCHD2/CHCHD10 mutations with corrected
CHCHD2/CHCHD10 mutations as controls? (5) Are
there any CHCHD2 mouse models phenocopying PD or
related neurological disorders? As discussed before, a
knockin mouse model harboring diseases-related
CHCHD2 mutations may work similarly as knockin
mouse model harboring diseases-related CHCHD10
mutations. (6) Even with the similarity between CHCHD2
and CHCHD10, they obviously perform differential cel-
lular functions and are under possible varied regulatory
modes. Further characterization of functions and regula-
tion of CHCHD2 and CHCHD10 will shed new light on
the pathology associated with both proteins. We antici-
pate the answers to the above-mentioned questions will
facilitate the discovery of novel targeted therapy for neu-
rological patients who carry mutations of CHCHD2, and
possibly of CHCHD10.
Declaration of Conflicting Interests
The author(s) declared no potential conflicts of interest with respect
to the research, authorship, and/or publication of this article.
Funding
The author(s) disclosed receipt of the following financial sup-
port for the research, authorship, and/or publication of this
article: The authors’ work is supported by NMRC (National
Medical Research Council) Parkinson’s Disease Trans-
lational Clinical Program Grant, Parkinson’s Disease Large
Collaborative Grant (SPARK II) and STaR Award (EK-T).
ORCID iD
Wei Zhou https://orcid.org/0000-0002-0872-6223
References
Ajroud-Driss S, Fecto F, Ajroud K, Lalani I, Calvo SE, Mootha
VK, and others. 2015. Mutation in the novel nuclear-
encoded mitochondrial protein CHCHD10 in a fam-
ily with autosomal dominant mitochondrial myopathy.
Neurogenetics 16(1):1–9.
Anderson CJ, Bredvik K, Burstein SR, Davis C, Meadows SM,
Dash J, and others. 2019. ALS/FTD mutant CHCHD10 mice
reveal a tissue-specific toxic gain-of-function and mito-
chondrial stress response. Acta Neuropathol 138(1):103–21.
Aras S, Bai M, Lee I, Springett R, Huttemann M, Grossman
LI. 2015. MNRR1 (formerly CHCHD2) is a bi-organellar
regulator of mitochondrial metabolism. Mitochondrion
20:43–51.
Aras S, Pak O, Sommer N, Finley R Jr, Huttemann M,
Weissmann N, and others. 2013. Oxygen-dependent expres-
sion of cytochrome c oxidase subunit 4-2 gene expression
is mediated by transcription factors RBPJ, CXXC5 and
CHCHD2. Nucleic Acids Res 41(4):2255–66.
Auranen M, Ylikallio E, Shcherbii M, Paetau A, Kiuru-Enari
S, Toppila JP, and others. 2015. CHCHD10 variant
p(Gly66Val) causes axonal Charcot-Marie-Tooth disease.
Neurol Genet 1(1):e1.
Bannwarth S, Ait-El-Mkadem S, Chaussenot A, Genin EC,
Lacas-Gervais S, Fragaki K, and others. 2014. A mitochon-
drial origin for frontotemporal dementia and amyotrophic
lateral sclerosis through CHCHD10 involvement. Brain
137(pt 8):2329–45.
Baughman JM, Nilsson R, Gohil VM, Arlow DH, Gauhar Z,
Mootha VK. 2009. A computational screen for regulators
of oxidative phosphorylation implicates SLIRP in mito-
chondrial RNA homeostasis. PLoS Genet 5(8):e1000590.
Burstein SR, Valsecchi F, Kawamata H, Bourens M, Zeng R,
Zuberi A, and others. 2018. In vitro and in vivo studies of
the ALS-FTLD protein CHCHD10 reveal novel mitochon-
drial topology and protein interactions. Hum Mol Genet
27(1):160–77.
Chaussenot A, Le Ber I, Ait-El-Mkadem S, Camuzat A, de
Septenville A, Bannwarth S, and others. 2014. Screening of
CHCHD10 in a French cohort confirms the involvement of
this gene in frontotemporal dementia with amyotrophic lat-
eral sclerosis patients. Neurobiol Aging 35(12):2884. e1–4.
Che XQ, Zhao QH, Huang Y, Li X, Ren RJ, Chen SD, and
others. 2017. Genetic features of MAPT, GRN, C9orf72
and CHCHD10 gene mutations in Chinese patients
with frontotemporal dementia. Curr Alzheimer Res
14(10):1102–8.
Che XQ, Zhao QH, Huang Y, Li X, Ren RJ, Chen SD, and
others. 2018. Mutation screening of the CHCHD2 gene
for Alzheimer’s disease and frontotemporal demen-
tia in Chinese mainland population. J Alzheimers Dis
61(4):1283–8.
Chio A, Mora G, Sabatelli M, Caponnetto C, Traynor BJ,
Johnson JO, and others. 2015. CHCH10 mutations in an
Italian cohort of familial and sporadic amyotrophic lateral
sclerosis patients. Neurobiol Aging 36(4):1767.e3–6.
Cornelissen T, Vilain S, Vints K, Gounko N, Verstreken P,
Vandenberghe W. 2018. Deficiency of parkin and PINK1
impairs age-dependent mitophagy in Drosophila. Elife 7.
doi:10.7554/eLife.35878
Dols-Icardo O, Nebot I, Gorostidi A, Ortega-Cubero S,
Hernandez I, Rojas-Garcia R, and others. 2015. Analysis
of the CHCHD10 gene in patients with frontotemporal
dementia and amyotrophic lateral sclerosis from Spain.
Brain 138(pt 12):e400.
El-Brolosy MA, Stainier DYR. 2017. Genetic compensation:
a phenomenon in search of mechanisms. PLoS Genet
13(7):e1006780.
Foo JN, Liu J, Tan EK. 2015. CHCHD2 and Parkinson’s dis-
ease. Lancet Neurol 14(7):681–2.
Funayama M, Ohe K, Amo T, Furuya N, Yamaguchi J, Saiki S,
and others. 2015. CHCHD2 mutations in autosomal domi-
nant late-onset Parkinson’s disease: a genome-wide linkage
and sequencing study. Lancet Neurol 14(3):274–82.
Genin EC, Bannwarth S, Lespinasse F, Ortega-Vila B, Fragaki
K, Itoh K, and others. 2018. Loss of MICOS complex
integrity and mitochondrial damage, but not TDP-43 mito-
chondrial localisation, are likely associated with severity
of CHCHD10-related diseases. Neurobiol Dis 119:159–71.
14 The Neuroscientist 00(0)
Genin EC, Madji Hounoum B, Bannwarth S, Fragaki K,
Lacas-Gervais S, Mauri-Crouzet A, and others. 2019.
Mitochondrial defect in muscle precedes neuromuscu-
lar junction degeneration and motor neuron death in
CHCHD10(S59L/+) mouse. Acta Neuropathol 138(1):
123–45.
Genin EC, Plutino M, Bannwarth S, Villa E, Cisneros-Barroso
E, Roy M, and others. 2016. CHCHD10 mutations promote
loss of mitochondrial cristae junctions with impaired mito-
chondrial genome maintenance and inhibition of apoptosis.
EMBO Mol Med 8(1):58–72.
Huang X, Wu BP, Nguyen D, Liu YT, Marani M, Hench J, and
others. 2018. CHCHD2 accumulates in distressed mito-
chondria and facilitates oligomerization of CHCHD10.
Hum Mol Genet 28(2):349.
Ikeda A, Matsushima T, Daida K, Nakajima S, Conedera S, Li
Y, and others. 2017. A novel mutation of CHCHD2 p.R8H
in a sporadic case of Parkinson’s disease. Parkinsonism
Relat Disord 34:66–8.
Ikon N, Ryan RO. 2017. Cardiolipin and mitochondrial cristae
organization. Biochim Biophys Acta 1859(6):1156–63.
Jansen IE, Bras JM, Lesage S, Schulte C, Gibbs JR, Nalls
MA, and others. 2015. CHCHD2 and Parkinson’s disease.
Lancet Neurol 14(7):678–9.
Jiao B, Xiao T, Hou L, Gu X, Zhou Y, Zhou L, and others.
2016. High prevalence of CHCHD10 mutation in patients
with frontotemporal dementia from China. Brain 139(pt
4):e21.
Johnson JO, Glynn SM, Gibbs JR, Nalls MA, Sabatelli M,
Restagno G, and others. 2014. Mutations in the CHCHD10
gene are a common cause of familial amyotrophic lateral
sclerosis. Brain 137(pt 12):e311.
Khurana V, Tardiff DF, Chung CY, Lindquist S. 2015. Toward
stem cell-based phenotypic screens for neurodegenerative
diseases. Nat Rev Neurol 11(6):339–50.
Koschmidder E, Weissbach A, Bruggemann N, Kasten M, Klein
C, Lohmann K. 2016. A nonsense mutation in CHCHD2 in
a patient with Parkinson disease. Neurology 86(6):577–9.
Kurzwelly D, Kruger S, Biskup S, Heneka MT. 2015. A distinct
clinical phenotype in a German kindred with motor neu-
ron disease carrying a CHCHD10 mutation. Brain 138(pt
9):e376.
Kwon EM, Connelly JP, Hansen NF, Donovan FX, Winkler
T, Davis BW, and others. 2017. iPSCs and fibroblast sub-
clones from the same fibroblast population contain compa-
rable levels of sequence variations. Proc Natl Acad Sci U S
A 114(8):1964–9.
Lee JJ, Sanchez-Martinez A, Zarate AM, Beninca C, Mayor U,
Clague MJ, and others. 2018a. Basal mitophagy is wide-
spread in Drosophila but minimally affected by loss of
Pink1 or parkin. J Cell Biol 217(5):1613–22.
Lee RG, Sedghi M, Salari M, Shearwood AJ, Stentenbach M,
Kariminejad A, and others. 2018b. Early-onset Parkinson
disease caused by a mutation in CHCHD2 and mitochon-
drial dysfunction. Neurol Genet 4(5):e276.
Liu Y, Clegg HV, Leslie PL, Di J, Tollini LA, He Y, and oth-
ers. 2015. CHCHD2 inhibits apoptosis by interacting
with Bcl-x L to regulate Bax activation. Cell Death Differ
22(6):1035–46.
Marroquin N, Stranz S, Muller K, Wieland T, Ruf WP,
Brockmann SJ, and others. 2016. Screening for CHCHD10
mutations in a large cohort of sporadic ALS patients: no
evidence for pathogenicity of the p.P34S variant. Brain
139(pt 2):e8.
McWilliams TG, Prescott AR, Montava-Garriga L, Ball G,
Singh F, Barini E, and others. 2018. Basal mitophagy
occurs independently of PINK1 in mouse tissues of high
metabolic demand. Cell Metab 27(2):439–49. e5.
Meng H, Yamashita C, Shiba-Fukushima K, Inoshita T,
Funayama M, Sato S, and others. 2017. Loss of Parkinson’s
disease-associated protein CHCHD2 affects mitochon-
drial crista structure and destabilizes cytochrome c. Nat
Commun 8:15500.
Modjtahedi N, Tokatlidis K, Dessen P, Kroemer G. 2016.
Mitochondrial proteins containing coiled-coil-helix-coiled-
coil-helix (CHCH) domains in health and disease. Trends
Biochem Sci 41(3):245–60.
Müller K, Andersen PM, Hübers A, Marroquin N, Volk AE,
Danzer KM, and others. 2014. Two novel mutations in con-
served codons indicate that CHCHD10 is a gene associated
with motor neuron disease. Brain 137(pt 12):e309.
Nicoletti G, Gagliardi M, Procopio R, Iannello G, Morelli M,
Annesi G, and others. 2017. A new CHCHD2 mutation
identified in a southern Italy patient with multiple system
atrophy. Parkinsonism Relat Disord 47:91–3.
Penttila S, Jokela M, Bouquin H, Saukkonen AM, Toivanen
J, Udd B. 2015. Late onset spinal motor neuronopathy is
caused by mutation in CHCHD10. Ann Neurol 77(1):163–
72.
Perrone F, Nguyen HP, Van Mossevelde S, Moisse M, Sieben
A, Santens P, and others. 2017. Investigating the role of
ALS genes CHCHD10 and TUBA4A in Belgian FTD-ALS
spectrum patients. Neurobiol Aging 51:177.e9–16.
Purandare N, Somayajulu M, Huttemann M, Grossman LI,
Aras S. 2018. The cellular stress proteins CHCHD10 and
MNRR1 (CHCHD2): partners in mitochondrial and nuclear
function and dysfunction. J Biol Chem 293(17):6517–29.
Ronchi D, Riboldi G, Del Bo R, Ticozzi N, Scarlato M,
Galimberti D, and others. 2015. CHCHD10 mutations in
Italian patients with sporadic amyotrophic lateral sclerosis.
Brain 138(pt 8):e372.
Shen S, He J, Tang L, Zhang N, Fan D. 2017. CHCHD10
mutations in patients with amyotrophic lateral sclerosis in
Mainland China. Neurobiol Aging 54:214.e7–10.
Shi CH, Mao CY, Zhang SY, Yang J, Song B, Wu P, and others.
2016. CHCHD2 gene mutations in familial and sporadic
Parkinson’s disease. Neurobiol Aging 38:217.e9–13.
Straub IR, Janer A, Weraarpachai W, Zinman L, Robertson
J, Rogaeva E, and others. 2018. Loss of CHCHD10-
CHCHD2 complexes required for respiration underlies the
pathogenicity of a CHCHD10 mutation in ALS. Hum Mol
Genet 27(1):178–89.
Teyssou E, Chartier L, Albert M, Bouscary A, Antoine JC,
Camdessanche JP, and others. 2016. Genetic analysis of
Zhou et al. 15
CHCHD10 in French familial amyotrophic lateral sclerosis
patients. Neurobiol Aging 42:218.e1–3.
Tio M, Wen R, Lim YL, Zukifli ZHB, Xie S, Ho P, and others.
2017. Varied pathological and therapeutic response effects
associated with CHCHD2 mutant and risk variants. Hum
Mutat 38(8):978–87.
van der Laan M, Horvath SE, Pfanner N. 2016. Mitochondrial
contact site and cristae organizing system. Curr Opin Cell
Biol 41:33–42.
Wang Y, Wang Z, Sun H, Mao C, Yang J, Liu Y, and others.
2018. Generation of induced pluripotent stem cell line
(ZZUi007-A) from a 52-year-old patient with a novel
CHCHD2 gene mutation in Parkinson’s disease. Stem Cell
Res 32:87–90.
West RJ, Furmston R, Williams CA, Elliott CJ. 2015.
Neurophysiology of Drosophila models of Parkinson’s dis-
ease. Parkinsons Dis 2015:381281.
Wollweber F, von der Malsburg K, van der Laan M. 2017.
Mitochondrial contact site and cristae organizing system: a
central player in membrane shaping and crosstalk. Biochim
Biophys Acta 1864(9):1481–9.
Woo JA, Liu T, Trotter C, Fang CC, De Narvaez E, LePochat P,
and others. 2017. Loss of function CHCHD10 mutations in
cytoplasmic TDP-43 accumulation and synaptic integrity.
Nat Commun 8:15558.
Wu H, Lu X, Cen Z, Xie F, Zheng X, Chen Y, and others. 2016.
Genetic analysis of the CHCHD2 gene in Chinese patients
with familial essential tremor. Neurosci Lett 634:104–6.
Xiao T, Jiao B, Zhang W, Pan C, Wei J, Liu X, and others. 2016.
Identification of CHCHD10 mutation in Chinese patients
with Alzheimer disease. Mol Neurobiol 54(7):5243–7.
Yang N, Zhao Y, Liu Z, Zhang R, He Y, Zhou Y, and others.
2019. Systematically analyzing rare variants of autoso-
mal-dominant genes for sporadic Parkinson’s disease in a
Chinese cohort. Neurobiol Aging 76:215.e1–7.
Zhang M, Xi Z, Zinman L, Bruni AC, Maletta RG, Curcio SA,
and others. 2015. Mutation analysis of CHCHD10 in differ-
ent neurodegenerative diseases. Brain 138(pt 9):e380.
Zhou Q, Chen Y, Wei Q, Cao B, Wu Y, Zhao B, and oth-
ers. 2017. Mutation screening of the CHCHD10 gene in
Chinese patients with amyotrophic lateral sclerosis. Mol
Neurobiol 54(5):3189–94.
Zhou W, Ma D, Sun AX, Tran HD, Ma DL, Singh BK, and
others. 2018a. PD-linked CHCHD2 mutations impair
CHCHD10 and MICOS complex leading to mitochondria
dysfunction. Hum Mol Genet 28(7):1100–16.
Zhou X, Liu Z, Guo J, Sun Q, Xu Q, Yan X, and others. 2018b.
Identification of CHCHD10 variants in Chinese patients
with Parkinson’s disease. Parkinsonism Relat Disord
47:96–7.
... The CHCH domain is characterized by a pair of cysteines separated by nine amino acids, knowns as the CX9C motifs (Liu et al., 2020a). The CHCHD-containing protein family is highly conserved in various physiological functions (Zhou et al., 2020). CHCH mutation or deletion on cysteine residues will lead to a loss of mitochondrial functions associated with targeting and position. ...
... Interestingly, a single knockout of CHCHD10 has no effect on the number of OPA-1, although a double knockout of CHCHD2/10 worsens the clearance of L-OPA1 (Liu et al., 2020b). DRP1, which promotes mitochondrial fission, is significantly reduced in CHCHD2 single knockout and CHCHD2/10 double knockout mice (Zhou et al., 2020). Thus, these findings suggest that the loss of CHCHD2/CHCHD10 may reduce mitochondrial dynamics. ...
... Thus, these findings suggest that the loss of CHCHD2/CHCHD10 may reduce mitochondrial dynamics. Other mitochondrial dynamics-related proteins, including MFN1, MFN2, and MFF, exhibit relatively similar levels of expression in control and CHCHD2/CHCHD10 knockdown cells (Zhou et al., 2020). ...
Article
Full-text available
CHCHD2 and CHCHD10 are homolog mitochondrial proteins that play key roles in the neurological, cardiovascular, and reproductive systems. They are also involved in the mitochondrial metabolic process. Although previous research has concentrated on their functions within mitochondria, their functions within apoptosis, synaptic plasticity, cell migration as well as lipid metabolism remain to be concluded. The review highlights the different roles played by CHCHD2 and/or CHCHD10 binding to various target proteins (such as OPA-1, OMA-1, PINK, and TDP43) and reveals their non-negligible effects in cognitive impairments and motor neuron diseases. This review focuses on the functions of CHCHD2 and/or CHCHD10. This review reveals protective effects and mechanisms of CHCHD2 and CHCHD10 in neurodegenerative diseases characterized by cognitive and motor deficits, such as frontotemporal dementia (FTD), Lewy body dementia (LBD), Parkinson’s disease (PD) and amyotrophic lateral sclerosis (ALS). However, there are numerous specific mechanisms that have yet to be elucidated, and additional research into these mechanisms is required.
... The coiled-coil-helix-coiled-coil-helix domain (CHCHD)-containing proteins are identified as a kind of conserved nucleus-encoded small mitochondrial proteins [9], among which have various important pathophysiological roles. Recently, CHCHD2 has been suggested to be associated with motor dysfunction and pathogenesis of Parkinson's disease (PD) [10,11], particularly in mitochondrial homeostasis [12,13]. Specifically, hepatocyte CHCHD2 contributed to liver fibrosis in nonalcoholic fatty liver disease (NAFLD) [14], whereas CHCHD3 was found to improve mitochondrial function via increasing mitochondrial ROS and promoting ferroptosis, by which modulated tumorigenesis [15]. ...
... Specifically, hepatocyte CHCHD2 contributed to liver fibrosis in nonalcoholic fatty liver disease (NAFLD) [14], whereas CHCHD3 was found to improve mitochondrial function via increasing mitochondrial ROS and promoting ferroptosis, by which modulated tumorigenesis [15]. In line with CHCHD2, CHCHD10 has been verified to be involved in multiple neurological alterations including Parkinson's disease (PD), frontotemporal dementia (FTD), as well as Alzheimer's disease (AD) [13,16]. CHCHD10 also serves as a mitochondrial modulator in adipose browning [17] and thermogenesis of adipocytes [18]. ...
Article
Full-text available
Background Pulmonary arterial hypertension (PAH) is a highly prevalent cardiopulmonary disorder characterized by vascular remodeling and increased resistance in pulmonary artery. Mitochondrial coiled–coil–helix–coiled–coil–helix domain (CHCHD)-containing proteins have various important pathophysiological roles. However, the functional roles of CHCHD proteins in hypoxic PAH is still ambiguous. Here, we aimed to investigate the role of CHCHD4 in hypoxic PAH and provide new insight into the mechanism driving the development of PAH. Methods Serotype 1 adeno‐associated viral vector (AAV) carrying Chchd4 was intratracheally injected to overexpress CHCHD4 in Sprague Dawley (SD) rats. The Normoxia groups of animals were housed at 21% O2. Hypoxia groups were housed at 10% O2, for 8 h/day for 4 consecutive weeks. Hemodynamic and histological characteristics are investigated in PAH. Primary pulmonary artery smooth muscle cells of rats (PASMCs) are used to assess how CHCHD4 affects proliferation and migration. Results We found CHCHD4 was significantly downregulated among CHCHD proteins in hypoxic PASMCs and lung tissues from hypoxic PAH rats. AAV1-induced CHCHD4 elevation conspicuously alleviates vascular remodeling and pulmonary artery resistance, and orchestrates mitochondrial oxidative phosphorylation in PASMCs. Moreover, we found overexpression of CHCHD4 impeded proliferation and migration of PASMCs. Mechanistically, through lung tissues bulk RNA-sequencing (RNA-seq), we further identified CHCHD4 modulated mitochondrial dynamics by directly interacting with SAM50, a barrel protein on mitochondrial outer membrane surface. Furthermore, knockdown of SAM50 reversed the biological effects of CHCHD4 overexpression in isolated PASMCs. Conclusions Collectively, our data demonstrated that CHCHD4 elevation orchestrates mitochondrial oxidative phosphorylation and antagonizes aberrant PASMC cell growth and migration, thereby disturbing hypoxic PAH, which could serve as a promising therapeutic target for PAH treatment. Graphical Abstract
... Although the physiological and pathological functions of the CHCHD family of proteins remain poorly understood, recent research has highlighted their close association with the development of neurodegenerative diseases [13]. Genetic analysis and experimental evidence indicate a strong correlation between CHCHD2 gene mutations and Parkinson's disease (PD) [14] and Lewy body disease (LBD) [15]. ...
Article
Full-text available
Huntington disease (HD) is a neurodegenerative disease caused by the abnormal expansion of a polyglutamine tract resulting from a mutation in the HTT gene. Oxidative stress has been identified as a significant contributing factor to the development of HD and other neurodegenerative diseases, and targeting anti-oxidative stress has emerged as a potential therapeutic approach. CHCHD2 is a mitochondria-related protein involved in regulating cell migration, anti-oxidative stress, and anti-apoptosis. Although CHCHD2 is highly expressed in HD cells, its specific role in the pathogenesis of HD remains uncertain. We postulate that the up-regulation of CHCHD2 in HD models represents a compensatory protective response against mitochondrial dysfunction and oxidative stress associated with HD. To investigate this hypothesis, we employed HD mouse striatal cells and human induced pluripotent stem cells (hiPSCs) as models to examine the effects of CHCHD2 overexpression (CHCHD2-OE) or knockdown (CHCHD2-KD) on the HD phenotype. Our findings demonstrate that CHCHD2 is crucial for maintaining cell survival in both HD mouse striatal cells and hiPSCs-derived neurons. Our study demonstrates that CHCHD2 up-regulation in HD serves as a compensatory protective response against oxidative stress, suggesting a potential anti-oxidative strategy for the treatment of HD.
... Recently, two missense mutations (T61I and R145Q) and one splice site mutation (c.300 + 5G > A) in CHCHD2 have been identified in autosomal dominant familial PD [78] . The mechanism of function in MICOS complex includes the regulation of mitochondrial apoptosis and mitophagy [79] . ...
Article
Full-text available
Mitochondrial dysfunction can lead to degeneration in the central nervous system. F1Fo-ATPase catalyzes most of the intracellular ATP synthesis which plays an essential role in cellular energy supply. The dimerized assembly of F1Fo-ATPase underlies the rotational catalytic function and regulates the mechanisms of oxidative phosphorylation. F1Fo-ATPase dysfunction is involved in a variety of neurological diseases, including epilepsy, Alzheimer's disease, and Parkinson’s disease. Dysregulated expression, activity, and localization of F1Fo-ATPase subunits and the interactions with pathogenic proteins result in decreased F1Fo-ATPase activity and ATP production, and aggravated oxidative stress.
... Mitochondrial cristae are propagations of the IMM that increase the available surface for the respiratory chain, hosting enzymatic complexes and making the OXPHOS more efficient. They are thus required to be functional for the correct localization of OXPHOS enzymes and for ATP production [50]. ...
Article
Full-text available
Amyotrophic lateral sclerosis (ALS) is a fatal neurodegenerative disease characterized by progressive loss of the upper and lower motor neurons. Despite the increasing effort in understanding the etiopathology of ALS, it still remains an obscure disease, and no therapies are currently available to halt its progression. Following the discovery of the first gene associated with familial forms of ALS, Cu–Zn superoxide dismutase, it appeared evident that mitochondria were key elements in the onset of the pathology. However, as more and more ALS-related genes were discovered, the attention shifted from mitochondria impairment to other biological functions such as protein aggregation and RNA metabolism. In recent years, mitochondria have again earned central, mechanistic roles in the pathology, due to accumulating evidence of their derangement in ALS animal models and patients, often resulting in the dysregulation of the energetic metabolism. In this review, we first provide an update of the last lustrum on the molecular mechanisms by which the most well-known ALS-related proteins affect mitochondrial functions and cellular bioenergetics. Next, we focus on evidence gathered from human specimens and advance the concept of a cellular-specific mitochondrial “metabolic threshold”, which may appear pivotal in ALS pathogenesis.
Article
Mitochondria are dynamic cellular organelles with complex roles in metabolism and signalling. Primary mitochondrial disorders are a group of approximately 400 monogenic disorders arising from pathogenic genetic variants impacting mitochondrial structure, ultrastructure and/or function. Amongst these disorders, defects of complex lipid biosynthesis, especially of the unique mitochondrial membrane lipid cardiolipin, and membrane biology are an emerging group characterised by clinical heterogeneity, but with recurrent features including cardiomyopathy, encephalopathy, neurodegeneration, neuropathy and 3‐methylglutaconic aciduria. This review discusses lipid synthesis in the mitochondrial membrane, the mitochondrial contact site and cristae organising system (MICOS), mitochondrial dynamics and trafficking, and the disorders associated with defects of each of these processes. We highlight overlapping functions of proteins involved in lipid biosynthesis and protein import into the mitochondria, pointing to an overarching coordination and synchronisation of mitochondrial functions. This review also focuses on membrane interactions between mitochondria and other organelles, namely the endoplasmic reticulum, peroxisomes, lysosomes and lipid droplets. We signpost disorders of these membrane interactions that may explain the observation of secondary mitochondrial dysfunction in heterogeneous pathological processes. Disruption of these organellar interactions ultimately impairs cellular homeostasis and organismal health, highlighting the central role of mitochondria in human health and disease.
Article
An extensive review is presented on mitochondrial structure and function, mitochondrial proteins, the outer and inner membranes, cristae, the role of F 1 F O -ATP synthase, the mitochondrial contact site and cristae organizing system (MICOS), the sorting and assembly machinery morphology and function, and phospholipids, in particular cardiolipin. Aspects of mitochondrial regulation under physiological and pathological conditions are outlined, in particular the role of dysregulated MICOS protein subunit Mic60 in Parkinson’s disease, the relations between mitochondrial quality control and proteins, and mitochondria as signaling organelles. A mathematical modeling approach of cristae and MICOS using mechanical beam theory is introduced and outlined. The proposed modeling is based on the premise that an optimization framework can be used for a better understanding of critical mitochondrial function and also to better map certain experiments and clinical interventions.
Article
Many different intrinsically disordered proteins and proteins with intrinsically disordered regions are associated with neurodegenerative diseases. These types of proteins including amyloid-β, tau, α-synuclein, CHCHD2, CHCHD10, and G-protein coupled receptors are increasingly becoming evaluated as potential drug targets in the pharmaceutical-based treatment approaches. Here, we focus on the neurobiology of this class of proteins, which lie at the center of numerous neurodegenerative diseases, such as Alzheimer’s and Parkinson’s diseases, Huntington’s disease, amyotrophic lateral sclerosis, frontotemporal dementia, Charcot–Marie–Tooth diseases, spinal muscular atrophy, and mitochondrial myopathy. Furthermore, we discuss the current treatment design strategies involving intrinsically disordered proteins and proteins with intrinsically disordered regions in neurodegenerative diseases. In addition, we emphasize that although the G-protein coupled receptors are traditionally investigated using structural biology-based models and approaches, current studies show that these receptors are proteins with intrinsically disordered regions and therefore they require new ways for their analysis.
Article
Background Parkinson’s disease (PD) is associated with coiled-coil-helix-coiled-coil-helix domain containing 2 (CHCHD2) downregulation, which has been linked to reduced cyclocytase activity and increased levels of oxygen free radicals, leading to mitochondrial fragmentation and apoptosis. Little is known about how CHCHD2 normally functions in the cell and, therefore, how its downregulation may contribute to PD. Objective This study aimed to identify such target genes using chromatin immunoprecipitation sequencing from SH-SY5Y human neuroblastoma cells treated with neurotoxin 1-methyl-4-phenylpyridinium (MPP+) as a PD model. Methods In this study, we established a MPP+ -reated SH-SY5Y cell model and evaluated the effects of CHCHD2 overexpression on cell proliferation and apoptosis. At the same time, we used high-throughput chromatin immunoprecipitation sequencing to identify its downstream target gene in SH-SY5Y cells. In addition, we verified the possible downstream target genes and discussed their mechanisms. Results The expression level of α-synuclein increased in SH-SY5Y cells treated with MPP+, while the protein expression level of CHCHD2 decreased significantly, especially after 24 h of treatment. Chip-IP results showed that CHCHD2 may regulate potential target genes such as HDX, ACP1, RAVER2, C1orf229, RN7SL130, GNPTG, erythroid 2 Like 2 (NFE2L2), required for cell differentiation 1 homologue (RQCD1), solute carrier family 5 member 7 (SLA5A7), and N-Acetyltransferase 8 Like (NAT8L). NFE2L2 and RQCD1 were validated as targets using PCR and western blotting of immunoprecipitates, and these two genes together with SLA5A7 and NAT8L were upregulated in SH-SY5Y cells overexpressing CHCHD2. Downregulation of CHCHD2 may contribute to PD by leading to inadequate expression of NFE2L2 and RQCD1 as well as, potentially, SLA5A7 and NAT8L. Conclusion Our results suggest that CHCHD2 plays a protective role by maintaining mitochondrial homeostasis and promoting proliferation in neurons. In this study, the changes of CHCHD2 and downstream target genes such as NFE2L2/RQCD1 may have potential application prospects in the future. These findings provide leads to explore PD pathogenesis and potential treatments.
Article
Full-text available
Mutations in coiled-coil-helix–coiled-coil-helix domain containing 10 (CHCHD10), a mitochondrial protein of unknown function, cause a disease spectrum with clinical features of motor neuron disease, dementia, myopathy and cardiomyopathy. To investigate the pathogenic mechanisms of CHCHD10, we generated mutant knock-in mice harboring the mouse-equivalent of a disease-associated human S59L mutation, S55L in the endogenous mouse gene. CHCHD10S55L mice develop progressive motor deficits, myopathy, cardiomyopathy and accelerated mortality. Critically, CHCHD10 accumulates in aggregates with its paralog CHCHD2 specifically in affected tissues of CHCHD10S55L mice, leading to aberrant organelle morphology and function. Aggregates induce a potent mitochondrial integrated stress response (mtISR) through mTORC1 activation, with elevation of stress-induced transcription factors, secretion of myokines, upregulated serine and one-carbon metabolism, and downregulation of respiratory chain enzymes. Conversely, CHCHD10 ablation does not induce disease pathology or activate the mtISR, indicating that CHCHD10S55L-dependent disease pathology is not caused by loss-of-function. Overall, CHCHD10S55L mice recapitulate crucial aspects of human disease and reveal a novel toxic gain-of-function mechanism through maladaptive mtISR and metabolic dysregulation.
Article
Full-text available
Recently, we provided genetic basis showing that mitochondrial dysfunction can trigger motor neuron degeneration, through identification of CHCHD10 encoding a mitochondrial protein. We reported patients, carrying the p.Ser59Leu heterozygous mutation in CHCHD10, from a large family with a mitochondrial myopathy associated with motor neuron disease (MND). Rapidly, our group and others reported CHCHD10 mutations in amyotrophic lateral sclerosis (ALS), frontotemporal dementia-ALS and other neurodegenerative diseases. Here, we generated knock-in (KI) mice, carrying the p.Ser59Leu mutation, that mimic the mitochondrial myopathy with mtDNA instability displayed by the patients from our original family. Before 14 months of age, all KI mice developed a fatal mitochondrial cardiomyopathy associated with enhanced mitophagy. CHCHD10S59L/+ mice also displayed neuromuscular junction (NMJ) and motor neuron degeneration with hyper-fragmentation of the motor end plate and moderate but significant motor neuron loss in lumbar spinal cord at the end stage of the disease. At this stage, we observed TDP-43 cytoplasmic aggregates in spinal neurons. We also showed that motor neurons differentiated from human iPSC carrying the p.Ser59Leu mutation were much more sensitive to Staurosporine or glutamate-induced caspase activation than control cells. These data confirm that mitochondrial deficiency associated with CHCHD10 mutations can be at the origin of MND. CHCHD10 is highly expressed in the NMJ post-synaptic part. Importantly, the fragmentation of the motor end plate was associated with abnormal CHCHD10 expression that was also observed closed to NMJs which were morphologically normal. Furthermore, we found OXPHOS deficiency in muscle of CHCHD10S59L/+ mice at 3 months of age in the absence of neuron loss in spinal cord. Our data show that the pathological effects of the p.Ser59Leu mutation target muscle prior to NMJ and motor neurons. They likely lead to OXPHOS deficiency, loss of cristae junctions and destabilization of internal membrane structure within mitochondria at motor end plate of NMJ, impairing neurotransmission. These data are in favor with a key role for muscle in MND associated with CHCHD10 mutations.
Article
Full-text available
CHCHD2 mutations were linked with autosomal dominant Parkinson’s disease (PD) and recently, Alzheimer’s disease/ Frontotemporal dementia. In current study, we generated isogenic human stem cell (hESC) lines harboring PD-associated CHCHD2 mutation R145Q or Q126X via CRISPR-Cas9 method, aiming to unravel pathophysiologic mechanism and seek potential intervention strategy against CHCHD2 mutant-caused defects. By engaging super resolution microscopy, we identified a physical proximity and similar distribution pattern of CHCHD2 along mitochondria with MICOS (mitochondrial inner membrane organizing system), a large protein complex maintaining mitochondria cristae. Isogenic hESCs and differentiated neural progenitor cells (NPCs) harboring CHCHD2 R145Q or Q126X mutation, showed impaired mitochondria function, reduced CHCHD2 and MICOS components, and exhibited nearly hollow mitochondria with reduced cristae. Furthermore, PD-linked CHCHD2 mutations lost their interaction with CHCHD10, while transient knockdown of either CHCHD2 or CHCHD10 reduced MICOS and mitochondria cristae. Importantly, a specific mitochondria-targeted peptide Elamipretide (MTP-131), now tested in phase 3 clinical trials for mitochondrial diseases was found to enhance CHCHD2 with MICOS and mitochondria oxidative phosphorylation enzymes in isogenic NPCs harboring heterozygous R145Q, suggesting that Elamipretide is able to attenuate CHCHD2 R145Q-induced mitochondria dysfunction. Taken together, our results suggested CHCHD2-CHCHD10 complex may be a novel therapeutic target for PD and related neurodegenerative disorders and Elamipretide may benefit CHCHD2 mutation-linked PD.
Article
Full-text available
Objective: Our goal was to identify the gene(s) associated with an early-onset form of Parkinson disease (PD) and the molecular defects associated with this mutation. Methods: We combined whole-exome sequencing and functional genomics to identify the genes associated with early-onset PD. We used fluorescence microscopy, cell, and mitochondrial biology measurements to identify the molecular defects resulting from the identified mutation. Results: Here, we report an association of a homozygous variant in CHCHD2, encoding coiled-coil-helix-coiled-coil-helix domain containing protein 2, a mitochondrial protein of unknown function, with an early-onset form of PD in a 26-year-old Caucasian woman. The CHCHD2 mutation in PD patient fibroblasts causes fragmentation of the mitochondrial reticular morphology and results in reduced oxidative phosphorylation at complex I and complex IV. Although patient cells could maintain a proton motive force, reactive oxygen species production was increased, which correlated with an increased metabolic rate. Conclusions: Our findings implicate CHCHD2 in the pathogenesis of recessive early-onset PD, expanding the repertoire of mitochondrial proteins that play a direct role in this disease.
Article
Full-text available
CHCHD2 mutation has been reported as a potential cause of a rare form of familial Parkinson's disease. Recently, a novel CHCHD2 mutation was identified in a family with Parkinson's disease. The dermal fibroblasts of the patient were obtained and successfully transformed into induced pluripotent stem cells(iPSCs), employing episomal plasmids expressing OCT3/4, SOX2, KLF4, LIN28, and L-MYC. Our model may offer a good platform for further research on the pathomechanism, drug testing, and gene therapy of this disease. Resource table: RESOURCE UTILITY: CHCHD2 mutation has been shown to be associated with Parkinson's disease (PD) (Shi et al., 2016). Induced pluripotent stem cells (iPSCs), generated from a patient harboring a CHCHD2 mutation, may provide an ideal cell model for exploring the pathogenesis of this disease and aid in drug screening. Resource details: Parkinson's disease (PD) is one of the most common neurodegenerative disorders, characterized by resting tremors, muscular rigidity, bradykinesia, and postural instability. Previous studies have revealed that parkinsonism can be caused by mutations in several genes including parkin, PTEN-induced putative kinase protein 1 (PINK1), parkinsonism-associated deglycase (DJ1), and ATPase 13A2 (ATP13A2) (Bonifati, 2014). In this study, a novel CHCHD2 mutation was identified in a family with Parkinson's disease (Shi et al., 2016), and the fibroblasts of the patient were successfully transformed into iPSCs. Episomal plasmids were used to generate the ZZUi007-A iPSC line (Fig. 1A). Pluripotency markers were examined via immunocytochemical staining using antibodies against human OCT-4, TRA-1-60 and Nanog (Fig. 1B). Flow cytometric analysis showed that more than 99% of the cells expressed OCT-4 and TRA-1-60 (Fig. 1C). The karyotype of CHCHD2-01 iPSCs was numerically and structurally normal (Fig. 1D). The mutation (c.182C > T; p.Thr61Ile) in CHCHD2 was confirmed by Sanger sequencing in the newly established iPSC line (Fig. 1E). Episomal plasmids were detected by polymerase chain reaction (PCR) using episomal plasmid-specific primers and disappeared from passage 15 (Fig. 1F). Furthermore, the iPSC line had the potential to differentiate into cells of all three germ layers in vivo (Fig. 1G).
Article
Full-text available
Mutations in paralogous mitochondrial proteins CHCHD2 and CHCHD10 cause autosomal dominant Parkinson Disease (PD) and Amyotrophic Lateral Sclerosis/Frontotemporal Dementia (ALS/FTD), respectively. Using newly generated CHCHD2, CHCHD10, and CHCHD2/10 double knockout cell lines, we find that the proteins are partially functionally redundant and similarly distributed throughout mitochondrial cristae. Contrary to these parallels, both proteins form heterodimers through a bioenergetically regulated mechanism that relies on key differences in the proteins' stability as well as mutual affinity: CHCHD2 is stabilized by loss of mitochondrial membrane potential; CHCHD10 oligomerization requires CHCHD2. Exploiting the dependence of CHCHD10 oligomerization on CHCHD2, we developed a heterodimer incorporation assay and demonstrate that CHCHD2 and CHCHD10 with disease-causing mutations readily form heterodimers. As we also find that both proteins are highly expressed in human Substantia nigra and cortical pyramidal neurons, mutant CHCHD2 and CHCHD10 may directly interact with their wild-type paralogs in the context of PD and ALS/FTD pathogenesis. Together, these findings demonstrate that differences in the stability and mutual affinity of CHCHD2 and CHCHD10 regulate their heterodimerization in response to mitochondrial distress, revealing an unanticipated link between PD and ALS/FTD pathogenesis.
Article
Full-text available
ELife digest Parkinson’s disease is a brain disorder where certain nerve cells slowly die, and the symptoms gradually worsen over time. While the risk of developing the condition increases with age, in certain patients the illness is caused by defects in two proteins, PINK1 and parkin. PINK1 and parkin help to manage mitochondria, the compartments in our cells that create molecules that serve as the energy currency for nearly all biological processes. When mitochondria get damaged, they release harmful substances that can kill their host cell. To prevent this, PINK1 and parkin can start a process known as mitophagy, which allows the cell to safely dispose of these dangerous mitochondria. Yet, mitophagy that is triggered by PINK1 and parkin has only been observed in cells grown in the laboratory; there is very little direct evidence that it also takes place in living organisms. If this mechanism does not happen in animals, then it is probably not relevant to Parkinson’s disease. Here, Cornelissen et al. genetically engineered fruit flies that carry a fluorescent marker which helps to track when and where damaged mitochondria are destroyed by a cell. The experiments revealed that mitophagy took place in muscles and in brain tissues. As the animals grew older, mitophagy became more frequent. However, this increase in mitophagy was not seen in insects that did not have PINK1 and parkin. These results showed that the role of PINK1 and parkin in mitophagy is not restricted to cells grown artificially. The fruit flies designed by Cornelissen et al. will be useful to investigate how PINK1 and parkin keep cells healthy by disposing of harmful mitochondria in living organisms. Ultimately, this may help to develop treatments that slow down the development of Parkinson’s disease.
Article
Studies have shown that rare variants of Mendelian genes for Parkinson's disease (PD) contribute to sporadic PD in the Caucasian population, which lacked confirmation in the Chinese population. Because the autosomal-dominant PD (AD-PD) had a phenotype closely resembling sporadic PD, we performed a systematic analysis of 7 AD-PD genes (SNCA, LRRK2, GIGYF2, VPS35, EIF4G1, DNAJC13, and CHCHD2) in 1456 Chinese sporadic PD patients and 1568 controls. Overall, 72 rare variants were identified, 7 of which were classified as likely pathogenic, 63 of which were categorized as of uncertain significance, and 2 of them were predicted to be likely benign. These AD-PD genes represented a clear enrichment of rare variants in PD patients from a burden analysis (p = 0.003), and significant differences could still be observed when likely pathogenic variants were removed (p = 0.027). The gene-based association testing also reached significance for LRRK2 (p = 0.004) and remained statistically significant after the Bonferroni correction. This report suggested that rare variants of AD-PD genes had a role in the Chinese sporadic PD cohort, especially for those rare variants of LRRK2.
Article
Following the involvement of CHCHD10 in FrontoTemporal-Dementia-Amyotrophic Lateral Sclerosis (FTD-ALS) clinical spectrum, a founder mutation (p.Gly66Val) in the same gene was identified in Finnish families with late-onset spinal motor neuronopathy (SMAJ). SMAJ is a slowly progressive form of spinal muscular atrophy with a life expectancy within normal range. In order to understand why the p.Ser59Leu mutation, responsible for severe FTD-ALS, and the p.Gly66Val mutation could lead to different levels of severity, we compared their effects in patient cells. Unlike affected individuals bearing the p.Ser59Leu mutation, patients presenting with SMAJ phenotype have neither mitochondrial myopathy nor mtDNA instability. The expression of CHCHD10S59L mutant allele leads to disassembly of mitochondrial contact site and cristae organizing system (MICOS) with mitochondrial dysfunction and loss of cristae in patient fibroblasts. We also show that G66V fibroblasts do not display the loss of MICOS complex integrity and mitochondrial damage found in S59L cells. However, S59L and G66V fibroblasts show comparable accumulation of phosphorylated mitochondrial TDP-43 suggesting that the severity of phenotype and mitochondrial damage do not depend on mitochondrial TDP-43 localization. The expression of the CHCHD10G66V allele is responsible for mitochondrial network fragmentation and decreased sensitivity towards apoptotic stimuli, but with a less severe effect than that found in cells expressing the CHCHD10S59L allele. Taken together, our data show that cellular phenotypes associated with p.Ser59Leu and p.Gly66Val mutations in CHCHD10 are different; loss of MICOS complex integrity and mitochondrial dysfunction, but not TDP-43 mitochondrial localization, being likely essential to develop a severe motor neuron disease.