ArticlePDF AvailableLiterature Review

Abstract and Figures

Alternative splicing is a crucial mechanism of gene expression regulation that enormously increases the coding potential of our genome and represents an intermediate step between messenger RNA (mRNA) transcription and protein posttranslational modifications. Alternative splicing occupies a central position in the development and functions of the nervous system. Therefore, its deregulation frequently leads to several neurological human disorders. In the present review, we provide an updated overview on the impact of alternative splicing in Parkinson's disease (PD), the second most common neurodegenerative disorder worldwide. We will describe the alternative splicing of major PD-linked genes by collecting the current evidences about this intricate and not carefully explored aspect. Assessing the role of this mechanism on PD pathobiology may represent a central step toward an improved understanding of this complex disease.
Content may be subject to copyright.
REVIEW ARTICLE
Splicing: is there an alternative contribution
to Parkinsons disease?
Vale n t i n a L a C o gnata
1,3
&Velia DAgata
3
&Francesca Cavalcanti
2
&
Sebastiano Cavallaro
1,2
Received: 8 January 2015 /Accepted: 4 May 2015
#The Author(s) 2015. This article is published with open access at Springerlink.com
Abstract Alternative splicing is a crucial mechanism of
gene expression regulation that enormously increases the
coding potential of our genome and represents an inter-
mediate step between messenger RNA (mRNA) transcrip-
tion and protein posttranslational modifications. Alterna-
tive splicing occupies a central position in the develop-
ment and functions of the nervous system. Therefore, its
deregulation frequently leads to several neurological hu-
man disorders. In the present review, we provide an up-
dated overview on the impact of alternative splicing in
Parkinsons disease (PD), the second most common neu-
rodegenerative disorder worldwide. We will describe the
alternative splicing of major PD-linked genes by
collecting the current evidences about this intricate and
not carefully explored aspect. Assessing the role of this
mechanism on PD pathobiology may represent a central
step toward an improved understanding of this complex
disease.
Keywords Parkinsons disease .Alternative splicing .
PD genes .mRNA splice transcripts .Protein isoforms
Introduction
The flow of genetic information from DNA to RNA to protein
has traditionally been considered the central dogma of molec-
ular biology. Additional steps of regulation are currently well
known, greatly expanding this simplistic framework and re-
vealing the complex network that controls gene expression
[1]. One of these steps is represented by alternative splicing
(AS), whereby a single gene gives rise to multiple messenger
RNA (mRNA) transcripts and protein isoforms with different
functional properties [1]. It is estimated that 94 % of human
protein-coding genes are alternatively spliced [2,3], and the
main site of alternative splicing events is the central nervous
system [4,5].
The alternative splicing process consists in the removal of
the intronic regions from the RNA primary transcript and si-
multaneous assembly of the exonic regions in different com-
binations to form a mature mRNA, which is then
polyadenilated, exported to the cytoplasm, and translated into
protein. The accuracy and efficiency of pre-mRNA splicing
process depend on a range of constitutive DNA sequence
motifs: the donor and the acceptor splice sites, the lariat
branch point, the polypyrimidine tract, and splicing enhancers
and silencers (Fig. 1, panel a). These motifs are recognized by
a large macromolecular splicing machinery (called the
spliceosome), which models the pre-mRNAwhile RNA poly-
merase II synthesizes it in the nucleus. The splicing machinery
includes five spliceosomal uridine-rich small nuclear ribonu-
cleoproteins (snRNPs) (U1, U2, U4, U5, and U6) and several
non-snRNP protein splicing factors such as the serine/arginine
(SR)-rich protein family and hnRNP proteins [6,7]. The splic-
ing reaction relies on two transesterification steps that occur
within the highly dynamic splicing machine. The stepwise
molecular mechanisms of the splicing reaction are detailed
in Fig. 1(panel a).
*Sebastiano Cavallaro
sebastiano.cavallaro@cnr.it
1
Institute of Neurological Sciences, Italian National Research
Council, Via Paolo Gaifami 18, 95125 Catania, Sicily, Italy
2
Institute of Neurological Sciences, Italian National Research
Council, 87050 Piano Lago di Mangone, Cosenza, Calabria, Italy
3
Department of Biomedical and Biotechnological Sciences, Section of
Human Anatomy and Histology, University of Catania, Catania, Italy
Neurogenetics
DOI 10.1007/s10048-015-0449-x
Alternative splicing works as an onoff switch in gene
expression. It affects the expression levels, stability, half-life
[via the nonsense-mediated mRNA decay (NMD)], and local-
ization of the RNA messengers. It has also the potential to
generate several protein isoforms with different biological
properties, proteinprotein interactions, subcellular localiza-
tion, signaling pathway, or catalytic ability. During the last
years, great efforts have been made to decipher the intricate
alternative splicing code. Five major alternative splicing
events (i.e., cassette exons, use of alternative acceptor and/or
donor sites, intron retention, and mutually exclusive exons)
have been described up to now and are detailed in Fig. 1
(panel b) [2,8]. However, how the spliceosome recognizes
alternative exons and decides which exons to include remains
not fully understood. Undoubtedly, there is more diversity in
splice transcript variants than in protein isoforms. Although
this is still not clear, different variants encode the same pro-
tein, but probably translate it with different efficiencies [9].
The finely tuned splicing regulatory network can easily
undergo alterations. An aberrant alternative splicing may arise
from changes in regulatory sequences required for correct pre-
mRNA processing (the so-called cis-acting mutations), as well
as from mutations that affect components necessary for splic-
ing regulation (trans-acting mutations). Cis-andtrans-
splicing aberrations represent direct causative agents of dis-
ease or more subtle contributions to the determinants of dis-
ease susceptibility or modulators of disease severity. An ex-
tensive range of neurological diseases has been already asso-
ciated to both splicing defects, including Alzheimersdisease,
retinitis pigmentosa, spinal muscular atrophy, muscular dys-
trophy, neurofibromatosis, and fragile X-associated tremor/
ataxia syndrome [1012,1,13]. In this broad neurological
disorder scenario, the relevance of alternative splicing in
Parkinsons disease (PD) is not still clear, and the splicing
mechanisms that regulate PD-related genes remain mostly
unknown.
Here, we provide an updated overview of the current
knowledge about the impact of alternative splicing on
Parkinsons disease. Firstly, we will take into account the most
common PD-related genes Bone by one^by analyzing their
alternative transcripts currently known and their involvement
in this disease. Then, we will describe the few studies that
have globally analyzed the changes of splice variant expres-
sion in PD patients through genome-wide RNA expression
approaches. Finally, we will briefly describe the current evi-
dences about the alternative splicing modulation in PD
through noncoding RNAs [microRNA (miRNA) and long
noncoding RNA (lcnRNA)].
Fig. 1 The alternative splicing mechanism. aFour main conserved DNA
sequence motifs allow the splicing mechanism: the donor splice site GU
(5SS), the acceptor splice site AG (3SS), the lariat branch point (A)
located upstream of the acceptor site and the polypyrimidine tract (PPT)
placed between the acceptor site and the branch point. The splicing
machinery includes mainly five spliceosomal uridine-rich small nuclear
ribonucleoproteins (snRNPs) (U1,U2,U4,U5,andU6) and further aux-
iliary RNA binding proteins. During the first step of spliceosome assem-
bly, U1 snRNP base pairs with the 5splice site of the pre-mRNA (E
complex), whereas U2 base pairs with the branch point (A complex). Then
the tri-snRNP complex U4, U5, and U6 associates with the forming
spliceosome (B complex), and both U1 and U4 are ejected. This allows
U6 to replace U1 at the 5splice site (C complex) and leads to a U6U2
interaction that gets close together the 5splice site and the branch point,
allowing for a transesterification step. At the end, U5 brings near the two
exons, joining them through a second transesterification reaction. bFive
major alternative splicing events are currently known: exon skipping/
inclusion, use of alternative 3splice site, use of alternative 5splice site,
mutually exclusive exons, and intron retention. In blue, are represented
the constitutive exons. Yellow and red represent the alternatively spliced
exons. The splicing events rely on the interplay between the constitutive
splicing motifs, the splicing regulatory sequences, the RNA secondary
structures, the components of the spliceosome, and further auxiliary
RNA-binding proteins. However, how the spliceosome decides which
exons to include remains currently not clear
Neurogenetics
Genetics of Parkinsonsdisease
PD is the second most common neurodegenerative disorder
worldwide, characterized by resting tremor, bradykinesia,
stiffness of movement, and postural instability. These symp-
toms are derived from the progressive loss of neurons from the
substantia nigra pars compacta, coupled with an accumula-
tion of intraneuronal aggregates called Lewy bodies.
Despite significant progresses in the understanding of PD
pathogenesis, the exact etiology of PD remains unknown.
Over the past 15 years, an even more detailed knowledge of
the genetic factors that contribute to PD has emerged through
different research strategies [14,15]. Linkage mapping analy-
sis, genome-wide association studies (GWAS), and next-
generation sequencing technologies are revealing an increas-
ing number of locus and genes strongly linked to either auto-
somaldominant(SNCA-PARK1,LRRK2-PARK8,VPS35-
PARK17, and GBA), or typical recessive (PARKIN-PARK2,
PINK1-PARK6,andDJ1-PARK7) and atypical recessive
(ATP13A2-PA R K 9 ,PLA2G6-PARK14,andFBXO7-PARK15)
or X-linked (ATP6A2 and TAF1) forms of disease. For the
sake of completeness, we mention here further monogenic
loci, not confirmed genes, or risk factor genes (i.e., PARK3,
UCHL1,PA R K 1 0 ,GIGYF2,PA R K 1 2 ,HTRA2,PA R K 1 6 ,
EIF4G1,DNAJ,HLA-DR,GAK-DGKQ,SYNJ1,andGBAP1)
[1517]. Furthermore, a large-scale meta-analysis of genome-
wide association data is revealing a wide range of additional
loci having genome-wide significant association [18]. How-
ever, we will overlook their discussion because of the few data
in the literature regarding their splicing regulation in patho-
logical conditions.
In the next paragraphs, we will describe the alternative
spliced mRNAvariants of PD genes and the current scientific
data demonstrating their involvement in PD pathogenesis. For
a more complete picture, we have also added some further
implicated genes (SRRM2,MAO-B,SNCAIP,MAPT,and
GBA), indicated as other PD-related genes, which are not di-
rectly causative genes, but whose splicing regulation seems to
be altered in PD states.
Autosomal dominant PD genes
SNCA
Alpha-synuclein, encoded by SNCA gene, is a small, natively
unfolded presynaptic protein linked to PD [19]. Aggregates of
alpha-synuclein protein represent the neuropathological hall-
mark lesions of PD and constitute the major components of
Lewy bodies. Genetically, mutations in SNCA gene were the
first to be associated with PD family inheritance. Missense
mutations in coding regions (Ala53Thr, Ala30Pro, and
Glu46Lys), single nucleotide substitution in 3untranslated
region (3UTR), and dose-dependent genomic multiplications
(duplications or triplications) of the gene cause both mono-
genic and sporadic forms of PD [20,19,21]. Some point
mutations in splice donor sites have also been reported
(IVS2+ 9A>C) [22].
SNCA gene maps to chromosome 4q22.1 and contains six
exons spanning about 114 kb [21]. The set of mRNAs pro-
duced by SNCA gene includes the full-length transcript, com-
monly known as SNCA-140 from the amino acidic length of
the encoded protein, and corresponds to SNCA-001, SNCA-
002, SNCA-003, SNCA-006, and SNCA-008 mRNAs from
Ensembl library (Table 1and Fig. 2). Further additional splic-
ing variants, known as SNCA-126, SNCA-112, and SNCA-
98 and corresponding to (i) SNCA-004, SNCA-203, SNCA-
201, (ii) SNCA-005, SNCA-202, and (iii) SNCA-010, respec-
tively, are generated by in-frame excision of exons 3, 5, or
both (Table 1and Fig. 2). Two additional splice variants
(SNCA-009 and SNCA-007) are generated from an inner tran-
scription start and encode proteins of 115 and 97 amino acids,
respectively (Table 1and Fig. 2). SNCA-140, SNCA-126, and
SNCA-112 are expressed in a broad spectrum of human tis-
sues, while SNCA-98 seems to be a brain-specific splice var-
iant with varying expression levels in different areas of fetal
and adult brain [23].
The expression profile of SNCA-140, SNCA-126, SNCA-
112, and SNCA-98 splice variants is different in the various
brain areas under normal and pathological states. Compared to
healthy controls, in PD frontal cortex, all these four transcripts
are overexpressed, with significant upregulation of SNCA-
126 [24]. In PD substantia nigra, only the three shorter tran-
scripts have been observed significantly overexpressed [25,
26], while higher SNCA-112 and SNCA-98 levels are also
present in the cerebellum [25]. Different expression profiles
of SNCA variants also occur in other forms of neurodegener-
ative disorders. Both SNCA-140 and SNCA-126 downregu-
lation and SNCA-98 overexpression have been reported in
dementia with Lewy bodies and Alzheimersdisease,while
SNCA-112 is upregulated in dementia with Lewy bodies and
downregulated in Alzheimersdisease[27,28,24].
Some interesting data emerge on SNCA-112 variant. An
association between PD risk-associated single nucleotide
polymorphisms (SNPs) within the 3region of SNCA gene
and higher SNCA-112 ratio level has been observed in about
100 of frontal cortex samples. These data reveal the cis-regu-
latory effect of these mutations on splicing mechanism [29].
The expression of SNCA-112 is also abundantly induced by
some parkinsonism mimetics (MPP+, rotenone) and related
oxidants [30]. However, the reason for these effects remains
unclear.
In addition to splice variants, specific RNA transcript iso-
forms of SNCA with an extended 3untranslated region have
been described and appear selectively linked to pathological
processes [31]. However, this review is focusing only on the
mRNA splice variants; thus, their discussion will be omitted.
Neurogenetics
The 140 amino acid isoform is a small protein with a mo-
lecular weight of 14.5 kDa. It is composed of three distinct
regions: (1) an amino terminus containing amphipathic helices
conferring the propensity to bind membranes; (2) a central
hydrophobic region, the so-called non-Ab component
(NAC), which confers the b-sheet potential; and (3) an acid
glutamatergic carboxyl terminus that is highly negatively
charged and prone to be unstructured [19]. Structural changes
in the shorter splicing isoforms can be predicted as a result of
exon skipping events. SNCA-126-predicted isoform shows
interruption of the N-terminal proteinmembrane interaction
domain [32]; SNCA-112 is significantly shorter in the un-
structured C-terminal [32], while SNCA-98 isoform results
in a truncated protein consisting almost only of the central
region containing NAC [23]. Recently, a lower aggregation
propensity of the shorter isoforms has been demonstrated
in vitro [33]. In addition, morphology studies by using elec-
tron microscopy have shown straight fibrils for SNCA-140,
shorter fibrils mostly arranged in parallel arrays for SNCA-
126, and circular structures for SNCA-98 [33]. These data
open new insights regarding the formation of Lewy bodies
induced by alpha-synuclein.
Numerous functions of alpha-synuclein have been pro-
posed, counting molecular chaperone, regulator of dopamine
uptake and homeostasis, inhibitor of phospholipase D2,
downregulator of p53 pathway [32],andpromoterofthe
SNARE-complex assembling [34]. Unfortunately, nothing is
known about the specific pathophysiological roles of each
alpha-synuclein isoform and their relative posttranslational
modifications (i.e., phosphorylations, nitration, sumoylation,
Tabl e 1 Alternative splice
variants of human autosomal
dominant PD genes
Gene name Transcript number Ensembl name Genbank accession number Protein length
SNCA 1. SNCA-003 NM_001146055 140 aa
2. SNCA-202 NM_007308 112 aa
3. SNCA-203 126 aa
4. SNCA-201 126 aa
5. SNCA-005 112 aa
6. SNCA-001 NM_001146054 140 aa
7. SNCA-002 NM_000345 140 aa
8. SNCA-008 140 aa
9. SNCA-006 140 aa
10. SNCA-004 126 aa
11. SNCA-010 98 aa
12. SNCA-009 115 aa
13. SNCA-007 67 aa
LRRK2 1. LRRK2-002 1271 aa
2. LRRK2-004 NM_198578 2527 aa
3. LRRK2-005 207 aa
4. LRRK2-001 521 aa
5. LRRK2-003 No protein
6. LRRK2-006 No protein
7. LRRK2-007 No protein
VPS35 1. VPS35-001 NM_018206 796 aa
2. VPS35-002 48 aa
3. VPS35-011 No protein
4. VPS35-012 No protein
5. VPS35-006 No protein
6. VPS35-003 No protein
7. VPS35-010 No protein
8. VPS35-005 47 aa
9. VPS35-008 41 aa
10. VPS35-007 No protein
11. VPS35-004 No protein
Gene name, Ensembl transcript names, GenBank accession numbers, and relative encoded amino acidic protein
length of splice variants are reported in the table. Number in the column BTranscript number^identifies the
transcript in Fig. 2
Neurogenetics
oxidation, glycosylation, cleavage, and ubiquitination), which
are known to play a key role in SNCA functions and regula-
tion [32].
LRRK2
LRRK2 encodes for leucine-rich repeat kinase 2 (or dardarin),
which is a large 2527 amino acid multidomain protein. The
protein consists of multiple conserved well-defined domains
including a small GTPase-like domain (Ras of complex pro-
teins or ROC), a domain of unknown function termed the C-
terminal of ROC (COR), a kinase domain, as well as several
protein interaction domains [e.g., the leucine-rich repeat
(LRR), the WD40 domain, the ankyrin repeat domain, and
the armadillo repeat region]. The precise physiological func-
tion of LRRK2 is unknown. However, LRRK2 seems impli-
cated in different cellular functions as neurite outgrowth, cy-
toskeletal maintenance, vesicle trafficking, and autophagic
protein degradation [35].
The LRRK2 gene spans a genomic region of 144 kb, with
51 exons, and harbors the most common mutations linked to
both autosomal dominant inherited late-onset and sporadic
PD. The missense mutations known so far are spread over
the whole LRRK2 gene and affect all functional domains.
Some mutations have much higher frequencies than others,
such as Gly2019Ser and mutations altering codon Arg1441,
respectively, in the kinase and ROC domains. In addition,
several unclear pathogenic mutations affecting splice sites
have been observed (IVS19+5_8delGTAA, IVS25-8delT,
IVS27-9C>T, IVS30-6C> T, IVS31+3A> G, IVS32+ 14G>
A, IVS33+6 T>A, IVS37-9A>G, IVS38+7C>T, IVS46-
14 T> A, and IVS46-8delT) [36,22,3743].
In addition to the full-length transcript (LRRK2-004), fur-
ther LRRK2 shorter transcripts are deposited in Ensembl li-
brary (Table 1and Fig. 2). Despite the existence of these
transcripts, there are currently no data analyzing the splicing
profile of this gene in PD states. Recently, a gene expression
and splicing analysis of the LRRK2 locus have been carried on
[44]. Both exon array and RT-PCR methods confirm the exis-
tence of an isoform with spliced out exons 3233 in the
substantia nigra and an isoform with exon 32 alone spliced
out in the occipital cortex,medulla,andcerebellum of healthy
humans [44].
Further evidences on LRRK2 splicing have been observed
by Giesert and collaborators [45], who have conducted a study
in various brain regions and organs from adult mice. In this
regard, it should be considered that LRRK2 is highly con-
served in human and mouse and that several transgenic animal
models have been created. Giesert et al. [45] have identified
two LRRK2 splice variants: one with skipped exon 5, primar-
ily expressed in astrocytes, and another truncated variant ter-
minating with an alternative exon 42a barely detectable in the
microglia but highly expressed in neurons and astrocytes.
Protein-structure predictions reveal that the loss of exon 5
may generate a smaller protein with changed affinity of bind-
ing partners, while the alternative exon 42a may leadto chang-
es of its enzymatic activity. In addition, the protein-interaction
domain WD40 would also be absent in such truncation.
Fig. 2 Structures of the
alternative splicing variants of
human dominant PD genes.
Structures of the described
mRNA splicing variants are
represented in the figure as
reported in Ensembl library
(http://www.ensembl.org/index.
html). On the left, each variant is
indicated with a number
corresponding to that indicated in
Tab le 1.LRRK2 gene is illustrated
in 5-3sense, while SNCA and
VPS35 genes are illustrated in
antisense corresponding to their
3-5sense transcription
Neurogenetics
Interestingly, the deletion of this domain in the Zebrafish
LRRK2 ortholog (zLRRK2) causes parkinsonism-like pheno-
type including loss of dopaminergic neurons in diencephalon
and locomotion defects [46]. Further studies will need to as-
sess the involvement of LRRK2 alternative splice variants in
PD.
VPS35
In 2011, two groups reported the identification of the same
missense mutation (p.Asp620Asn) in the vacuolar protein
sorting 35 (VPS35) gene as a novel cause of autosomal dom-
inant PD [47,48]. VPS35 was the first PD gene found by a
direct whole exome sequencing in large families of Austrian
and Swiss origins. An in-depth sequence analyses of all cod-
ing, noncoding, and exonintron boundaries VPS35 genetic
regions have been performed in a large well-characterized
cohort of Lewy body disorders, including PD patients, PD
with dementia, and dementia with Lewy bodies [49]. In addi-
tion to three novel missense mutations, silent and intronic
variations, predicted to activate cryptic splice sites, have been
observed in the patients group but not in controls. However,
the pathogenicity of these mutations was not completely con-
clusive since these mutations were not supported by segrega-
tion analysis in family relatives [49].
Various spliced transcript variants of this gene are reported
in Ensembl library (Table 1and Fig. 2), but the majority of
them are processed for degradation and do not encode
proteins.
Autosomal recessive PD genes
Early-onset typical PD genes
PARK2 Mutations in PA R K 2 gene (also known as PA R K 2
parkin RBR E3 ubiquitin-protein ligase) are the most common
cause (50 % of cases) of autosomal recessive juvenile parkin-
sonism (AR-JP), a form of early-onset parkinsonism charac-
terized by good and prolonged response to levodopa and a
benign, slow course. PA R K 2 mutations also explain ~15 %
of the sporadic cases with onset before 45 [50,51] and act
as susceptibility alleles for late-onset forms of Parkinsons
disease (2 % of cases) [52]. Along with about 200 mutations
currently identified in PARK2 coding region, several point
mutations in splice acceptor or donor sites (introns 1, 6, 7,
10, 12, 13, and 16) have been identified in PD patients
[5357,22,58,59].
PA R K 2 gene spans more than 1.38 Mb of genomic DNA in
the long arm of chromosome 6 (6q25.2q27) and contains 12
exons, which are alternatively spliced to produce at least 11
different splicing variants (Table 2and Fig. 3)[59]. The full-
length PA R K 2 transcript (PARK2-004) encodes a protein of
465 amino acids (parkin) [60,61,59] acting in numerous
molecular pathways (protein turnover, stress response, mito-
chondrial homeostasis, mitophagy, mitochondrial DNA stabil-
ity, metabolism, cell growth, and survival) [62]. Multiple
parkin isoforms likely arising from PAR K 2 splicing variants
have been observed in different brain areas through Western
blot studies [9].
The extensive alternative splicing of PARK2 is differently
regulated both at transcript and protein level in tissues and cells
[63,64,24,6568]. Distinct expression patterns of PA RK 2
splice variants emerge in human brain regions [69] and leuko-
cytes [68], in rat brain, in neuronal and glial cells [67],andina
wide variety of mouse tissues (brain, heart, lung, liver, skeletal
muscle, kidney, and testis) [64]. At the protein level, PA R K 2
protein isoforms show a differential distribution in human leu-
kocytes [70] and aged brain [71], as well as in different rat and
mouse nervous system areas (cerebral cortex/diencephalons,
hippocampus, cerebellum, brainstem, striatum, spinal cord,
and substantia nigra), peripheral tissues (heart, liver, spleen,
pancreas, and kidney), and developmental stages [7276].
Emerging evidences support the importance of PARK2
splice variant expression changes in disease development.
Differential expression of PARK2 transcripts have been iden-
tified in the frontal cortex of Parkinsonsdisease,puredemen-
tia with Lewy bodies, common Lewy body disease, and
Alzheimers disease patients, compared to controls [65,24].
Particularly, two PARK2 splicing variants are significantly
overexpressed in PD [65]. Another study reports both an in-
crease in the expression level of a parkin splice variant and a
decrease of the wild type between PD patients and healthy
controls [66]. The differential and disease-specific expression
profiles of PA R K 2 alternative splice variants suggest a role for
splicing deregulation in the development of neurodegenera-
tive disorders.
PINK1 Homozygous or compound heterozygous loss-of-
function mutations in PTEN-induced putative kinase 1
(PINK1) are the second most frequent cause of autosomal
recessive early-onset parkinsonism. Mutation frequency
varies geographically from 1 to 9 % depending on ethnic
background [77]. The PINK1 mutation spectrum involves
nonsense and missense mutations, insertions, or deletions,
and whole gene or single/multiple exon copy number variants
located across the entire gene [78].
PINK1 gene maps in the short arm of chromosome 1
(1p36.12), encompassing ~18 kb of genomic DNA. Its coding
sequence is spread over eight exons. In addition to the full
length (PINK1-001), two shorter variants exist but do not
produce proteins (Table 2and Fig. 3).
Some interesting findings emerge regarding the splicing
regulation of exon 7. A 23-bp deletion disrupting the splice
acceptor site of exon 7 has been detected in a sporadic parkin-
sonian patient, producing several aberrant mRNAs [79].
Moreover, whole exon 7 deletion and a novel U1-dependent
Neurogenetics
5splice-site mutation in exon 7 have been found in a large
Spanish family with PD members [80].
The PINK1 protein is a putative serine/threonine kinase of
581 amino acids involved in mitochondrial response to cellu-
lar and oxidative stress [81]. It has been demonstrated that at
least two isoforms are expressed in the human brain: a full-
length protein of ~63 kDa and an N-terminally truncated iso-
form of 52 kDa [77,8284]. An additional isoform of approx-
imately 45 kDa has been suggested, although it has not been
extensively studied [85]. The 52-kDa isoform seems to orig-
inate by enzymatic cleavage of PARL [86]; however, the exact
nature of the isoforms, the precise reason for the cleavage, and
the functional roles of these three different isoforms require
further studies.
Tabl e 2 Alternative splice variants of human autosomal recessive PD
genes
Gene
name
Transcript
number
Ensembl name Genbank accession
number
Protein
length
PA R K 2 1. PARK2-004 NM_004562 465 aa
2. PARK2-005 NM_013987 437 aa
3. PARK2-006 NM_013988 316 aa
4. PARK2-001 274 aa
5. PARK2-003 274 aa
6. PARK2-007 218 aa
7. PARK2-201 176 aa
8. PARK2-204 87 aa
9. PARK2-002 368 aa
10. PARK2-202 74 aa
11. PARK2-203 201 aa
PINK1 1. PINK1-001 NM_032409 581 aa
2. PINK1-002 No protein
3. PINK1-003 No protein
DJ1 1. PARK7-004 189 aa
2. PARK7-002 NM_001123377;
NM_007262
189 aa
3. PARK7-007 No protein
4. PARK7-001 189 aa
5. PARK7-003 169 aa
6. PARK7-008 No protein
7. PARK7-005 189 aa
8. PARK7-006 189 aa
9. PARK7-009 No protein
10. PARK7-010 160 aa
ATP13A2 1. ATP13A2-001 NM_022089 1180 aa
2. ATP13A2-002 NM_001141974 1158 aa
3. ATP13A2-005 NM_001141973 1175 aa
4. ATP13A2-004 No protein
5. ATP13A2-003 No protein
6. ATP13A2-010 191 aa
7. ATP13A2-007 398 aa
8. ATP13A2-014 258 aa
9. ATP13A2-009 321 aa
10. ATP13A2-006 No protein
11. ATP13A2-201 228 aa
12. ATP13A2-013 No protein
13. ATP13A2-011 190 aa
14. ATP13A2-012 191 aa
15. ATP13A2-008 188 aa
PLA2G6 1. PLA2G6-001 NM_003560 806 aa
2. PLA2G6-201 NM_001004426 752 aa
3. PLA2G6-002 NM_001199562 752 aa
4. PLA2G6-025 No protein
5. PLA2G6-021 No protein
6. PLA2G6-014 166 aa
7. PLA2G6-024 No protein
8. PLA2G6-013 No protein
Tabl e 2 (continued)
Gene
name
Transcript
number
Ensembl name Genbank accession
number
Protein
length
9. PLA2G6-026 168 aa
10. PLA2G6-015 120 aa
11. PLA2G6-010 99 aa
12. PLA2G6-023 226 aa
13. PLA2G6-009 No protein
14. PLA2G6-027 51 aa
15. PLA2G6-022 No protein
16. PLA2G6-012 151 aa
17. PLA2G6-019 124 aa
18. PLA2G6-011 No protein
19. PLA2G6-020 No protein
20. PLA2G6-008 No protein
21. PLA2G6-016 229 aa
22. PLA2G6-005 99 aa
23. PLA2G6-018 157 aa
24. PLA2G6-003 99 aa
25. PLA2G6-017 197 aa
26. PLA2G6-007 80 aa
27. PLA2G6-004 No protein
28. PLA2G6-006 No protein
FBXO7 1. FBXO7-003 41 aa
2. FBXO7-001 NM_012179 522 aa
3. FBXO7-004 49 aa
4. FBXO7-005 No protein
5. FBXO7-006 129 aa
6. FBXO7-002 NM_001033024;
NM_001257990
408 aa
7. FBXO7-007 54 aa
8. FBXO7-008 No protein
9. FBXO7-010 No protein
Gene name, Ensembl transcript names, GenBank accession numbers, and
relative encoded amino acidic protein length of splice variants are report-
ed in the table. Number in the column BTranscript number^identifies the
transcript in Fig. 3
Neurogenetics
DJ1 Mutations in the DJ1 (also known as PA R K 7 )geneare
the less common cause of autosomal recessive parkinsonism
(~1 % of early-onset PD) [87,88]. A large homozygous dele-
tion and a missense mutation (L166P) in DJ-1gene were first
identified in both Italian and Dutch consanguineous families
[89,90]. Additional mutations have been collected in other
PD families and include missense mutations in coding and
promoter regions, frame shifts, copy number variations [91,
88], and splice site alterations [92,93].
DJ-1gene maps to chromosome 1 (1p36.23) and includes
seven exons. Several spliced transcript variants have been
identified encoding the same protein (Table 2and Fig. 3).
Two shorter transcripts (the first lacking exon 4 and the second
starting in an inner transcription point) encode for smaller
proteins (Table 2and Fig. 3).
The product of DJ-1gene is a highly conserved protein of
189 amino acids belonging to the peptidase C56 family [94]. It
is a multifunctional protein, acting as a positive regulator of
transcription, redox-sensitive chaperone, sensor for oxidative
stress, and apparently protects neurons from ROS-induced
apoptosis [95,96]. In the human brain and peripheral blood,
several DJ-1isoforms exist and differ on their isoelectric point
(pI)[97100]. The relative abundance of these different DJ-1
isoforms appears to be altered in PD, and therefore, blood DJ-
1isoforms have been proposed as potential biomarkers for
Parkinsons disease [101]. The different pI of each variant
are believed to result from posttranslational modifications that
alter the intrinsic charge of the protein [101]. Interestingly, it
has been demonstrated that one of the major binding partners
of DJ-1 in dopaminergic neuronal cells is the splicing factor
proline/glutamine-rich (SFPQ protein) [96,102]. SFPQ, orig-
inally identified as a polypyrimidine tract-binding protein, is
part of the spliceosome C complex and is required for in vitro
splicing of pre-mRNA [96,102]. DJ-1 binding to SFPQ mod-
ulates its transcriptional activity and, therefore, tunes its effect
on splicing regulation. DJ-1 mutations could reverberate on its
downstream targets, including the splicing factor SFPQ and
altering the splicing control.
Juvenile atypical PD genes
ATP13A2 ATP13A2 mutations are associated with Kufor
Rakeb syndrome, a form of recessively levodopa-responsive
inherited atypical parkinsonism [103]. It encodes a large pro-
tein belonging to the ATPase transmembrane transporters, and
recently, it has been identified as a potent modifier of the
toxicity induced by alpha-synuclein [104].
ATP13A2 is composed of 29 exons and lies on chromo-
some 1 covering about 26 kb of genomic DNA. One of the
first identified disease-causing mutations was a guanine-to-
adenine transition in the donor splice site of exon 13, leading
to the skipping of exon 13 and resulting in a deletion of part of
the third transmembrane domain [105].
According to data repositories, at least 15 alternatively
spliced transcripts are expressed in humans (Table 2and
Fig. 3). The longest transcripts are ATP13A2-001,
ATP13A2-002, and ATP13A2-005. Transcript variants
ATP13A2-001 and ATP13A2-005 differ only in a nucleotide
segment on exon 5, while transcript variant ATP13A2-002
lacks exons 22 and 28. The ATP13A2 mRNA is highly
Fig. 3 Structures of the alternative splicing variants of human recessive
PD genes. Structures of the described mRNA splicing variants are
represented in the figure as reported in Ensembl library (http://www.
ensembl.org/index.html). On the left, each variant is indicated with a
number corresponding to that indicated in Table 2. All transcripts are
illustrated in 5-3sense, except PA R K 2 ,ATP13A2,andPLA2G6 genes,
which are illustrated in antisense corresponding to their 3′–5sense
transcription
Neurogenetics
expressed in the brain, particularly in the substantia nigra of
patients with classical late-onset PD [91]. However, nothing is
known about the splicing expression profiles of this gene in
PD and healthy subjects.
The products of these transcripts have been studied at the
protein level. The isoform 1 encoded by ATP13A2-001 is a
protein of 1180 amino acids with ten transmembrane domains.
Isoform 2 encoded by ATP13A2-005 contains a five amino
acid deletion near the N-terminus, while isoform 3, encoded
by ATP13A2-002, contains two deletions, generating a highly
diverged C-terminus [105]. Functional studies have shown
that the isoform 1 is located in the lysosome membrane,
whereas the isoform 3 protein is retained in the endoplasmic
reticulum and rapidly degraded by the proteasome. In addi-
tion, both isoforms 1 and 3 are eliminated via the endoplasmic
reticulum-associated degradation pathway [105].
PLA2G6 Recessive mutations in the phospholipase A2
group VI (PLA2G6) gene have been initially described
as the cause of infantile neuroaxonal dystrophy and neu-
rodegeneration associated with brain iron accumulation.
Recently, this gene has also been associated with a partic-
ular parkinsonian phenotype, consisting of levodopa-
responsive dystonia, pyramidal signs, and cognitive/
psychiatric features, with onset in early adulthood [106].
Among PLA2G6 identified mutations, the c.1077G>A
mutation at the last nucleotide of exon 7 (apparently a
synonymous mutation) stands out as a cause of abnormal
mRNA splicing. This single nucleotide substitution causes
the activation of a cryptic splice site producing a 4-bp
deleted transcript with altered frame shift in leukocytes
[106].
PLA2G6 gene maps on chromosome 22 (q13.1), covering
70 kb of genomic DNA. Several transcript variants encoding
multiple isoforms have been described up to now (Table 2and
Fig. 3). The longest PLA2G6 mRNA PLA2G6-001 includes
17 exonic regions and encodes the 85/88 kDa calcium-
independent phospholipase known as A2 isoform a. The other
Tabl e 3 Alternative splice
variants of human X-linked PD
genes
Gene name Transcript number Ensembl name Genbank accession number Protein length
TAF 1 1. TAF1-201 NM_001286074 1895 aa
2. TAF1-009 NM_138923 1872 aa
3. TAF1-008 NM_004606 1893 aa
4. TAF1-014 458 aa
5. TAF1-010 490 aa
6. TAF1-021 No protein
7. TAF1-011 No protein
8. TAF1-013 No protein
9. TAF1-012 No protein
10. TAF1-015 No protein
11. TAF1-018 No protein
12. TAF1-016 No protein
13. TAF1-022 No protein
14. TAF1-023 No protein
15. TAF1-005 No protein
16. TAF1-006 No protein
17. TAF1-020 No protein
18. TAF1-019 No protein
19. TAF1-017 279 aa
20. TAF1-007 150 aa
ATP6AP2 1. ATP6AP2-004 NM_005765 350 aa
2. ATP6AP2-007 203 aa
3. ATP6AP2-005 No protein
4. ATP6AP2-006 243 aa
5. ATP6AP2-001 259 aa
6. ATP6AP2-003 No protein
7. ATP6AP2-002 No protein
Gene name, Ensembl transcript names, GenBank accession numbers, and relative encoded amino acidic protein
length of splice variants are reported in the table. The number in the column BTranscript number^identifies the
transcript in Fig. 4
Neurogenetics
two long transcripts (PLA2G6-002 and PLA2G6-201) differ
in the start point, both lack of exon 9 and encode the same
protein, called isoform b. The expression profile of this gene
in healthy and disease states remains unknown.
FBXO7 Mutations in the F-box only protein 7 (FBXO7)
gene cause parkinsonian pyramidal disease (PPD- or
PARK15-associated parkinsonism), an autosomal reces-
sive neurodegenerative disease with juvenile onset, se-
vere levodopa-response, and additional pyramidal signs.
Some pathogenic mutations have been identified
(R378G, R498X, and T22M) including a compound het-
erozygous mutation (IVS7+1G/T) that removes the in-
variable splice donor of intron 7 and may disrupt FBXO7
messenger RNA splicing [107109].
The FBXO7 gene, mapped on chromosome 22q12.3, con-
tains nine exons spanning about 24.1 kb. It encodes a 522
amino acid protein consisting ofseveral domains [108], which
target proteins for ubiquitination [108]. Alternatively spliced
transcript variants of this gene have been identified (Table 2
and Fig. 3)[107]. FBXO7-001 is the longest and more abun-
dant transcript ubiquitously expressed [110], particularly in
skin fibroblasts [111]. FBXO7-002 arises from an inner alter-
native exon 1,differs in the start codon, and produces a shorter
isoform. Both these encoded protein isoforms have been de-
tected in cells [111].
X-linked parkinsonism
X-linked dystonia parkinsonism (XDP) is an X-linked reces-
sive adult-onset movement disorder characterized by both
dystonia and parkinsonism. TATA-box binding protein-associ-
ated factor 1 (TAF1) gene, located in the disease locus
Xq13.1, has been reported as the first related XPD gene, har-
boring disease-specific single-nucleotide changes and a small
deletion within the multiple transcript panel [112]. This gene
is part of a complex region of DNA (the TAF1/DYT3 multiple
transcript systems), which encompasses the exonic regions of
TAF1 gene and further additional downstream exons [112,
113]. This system includes multiple different transcription
start sites and encodes multiple spliced transcripts and iso-
forms (Table 3and Fig. 4)[112].
Recently, the ATP6AP2 gene (Table 3and Fig. 4) has been
proposed as a novel gene for X-linked parkinsonism with
spasticity (XPDS) by exome sequencing analysis [114]. A
silent mutation (p.S115S) in the ATP6AP2 gene has been
identified in one affected individual, resulting in the aberrant
splicing of ATP6AP2 mRNA and the overexpression of a
minor splice isoform [114]. Noteworthy, the ATP6AP2 is an
essential accessory component of the vacuolar ATPase re-
quired for lysosomal degradative functions and autophagy, a
pathwayfrequentlyaffectedinPD.
Other PD-related genes
SNCAIP
Synphilin-1, encoded by SNCAIP gene, is a presynaptic pro-
tein containing several proteinprotein interaction motifs, in-
cluding ankyrin-like repeats, a coiled-coil domain, and an
ATP/GTP-binding domain [115]. It interacts strongly with
alpha-synuclein in neuronal tissue and may play a role in the
formation of Lewy bodies during neurodegeneration. It is also
implicated in parkinsonism as one of the parkin substrates. In
addition, some studies have identified SNCAIP sequence var-
iants in PD patients and have suggested it as a candidate PD
gene [116,117].
SNCAIP gene maps on chromosome 5 (5q23.2) and
spans about 152 kb of genomic DNA. Although the data-
base Gene reports SNCAIP composed of 11 exons, addi-
tional exonic regions emerge by aligning the sequence of
the gene with each transcript. To date, at least 22 alterna-
tive spliced transcript variants have been identified
(Table 4and Fig. 5), but the most studied are synphilin-1
and 1A. Synphilin-1 (SNCAIP-001) is the full-length tran-
script, while synphilin-1A variant is a shorter form
(SNCAIP-201). The latter lacks exons 4 and 5 and contains
an extra exon located between exons 10 and 11. Synphilin-
1A isoform is thought to be involved in the pathogenesis of
PD and may play an important role in the formation of
Lewy bodies [118120]. Interestingly, synphilin-1A pro-
tein shows enhanced aggregation properties, which cause
neuronal toxicity [118120].
Fig. 4 Structures of the alternative splicing variants of human X-linked
PD genes. Structures of the described mRNA splicing variants are repre-
sented in the figure as reported in Ensembl library (http://www.ensembl.
org/index.html). On the left, each variant is indicated with a number
corresponding to that indicated in Table 3. All transcripts are illustrated
in 5-3sense
Neurogenetics
The mRNA expression levels of synphilin 1, 1A, and other
two additional synphilin variants have been simultaneously
Tabl e 4 Alternative splice variants of other human PD-related genes
Gene
name
Transcript
number
Ensembl name Genbank accession
number
Protein
length
SNCAIP 1. SNCAIP-019 135 aa
2. SNCAIP-003 66 aa
3. SNCAIP-016 161 aa
4. SNCAIP-017 98 aa
5. SNCAIP-010 858 aa
6. SNCAIP-001 NM_005460 919 aa
7. SNCAIP-204 113 a a
8. SNCAIP-201 NM_001242935 603 aa
9. SNCAIP-004 66 aa
10. SNCAIP-018 68 aa
11. SNCAIP-002 1016 aa
12. SNCAIP-006 62 aa
13. SNCAIP-005 No protein
14. SNCAIP-007 66 aa
15. SNCAIP-203 88 aa
16. SNCAIP-202 62 aa
17. SNCAIP-012 66 aa
18. SNCAIP-011 88 aa
19. SNCAIP-009 588 aa
20. SNCAIP-008 113 aa
21. SNCAIP-015 14 aa
22. SNCAIP-013 No protein
MAO-B1. MAOB-001 NM_000898 520 aa
2. MAOB-002 No protein
3. MAOB-004 No protein
GBA 1. GBA-011 No protein
2. GBA-001 NM_000157 536 aa
3. GBA-002 NM_001005741;
NM_001005742
536 aa
4. GBA-003 No protein
5. GBA-009 No protein
6. GBA-015 NM_001171812 487 aa
7. GBA-016 NM_001171811 449 aa
8. GBA-005 No protein
9. GBA-012 No protein
10. GBA-006 No protein
11. GBA-010 No protein
12. GBA-014 No protein
13. GBA-007 No protein
14. GBA-013 No protein
MAPT 1. MAPT-204 NM_005910 441 aa
2. MAPT-202 NM_001123067 412 aa
3. MAPT-201 NM_016835 758 aa
4. MAPT-205 NM_001203251;
NM_001203252
410 aa
5. MAPT-203 NM_001123066 776 aa
6. MAPT-013 No protein
7. MAPT-001 NM_016841 352 aa
8. MAPT-002 NM_016834 383 aa
Tabl e 4 (continued)
Gene
name
Transcript
number
Ensembl name Genbank accession
number
Protein
length
9. MAPT-006 410 aa
10. MAPT-007 441 aa
11. MAPT-008 758 aa
12. MAPT-004 776 aa
13. MAPT-003 412 aa
14. MAPT-009 341 aa
15. MAPT-014 No protein
16. MAPT-011 59 aa
17. MAPT-010 No protein
18. MAPT-012 No protein
SRRM2 1. SRRM2-001 NM_016333 2752 aa
2. SRRM2-201 311aa
3. SRRM2-003 1018 aa
4. SRRM2-006 297 aa
5. SRRM2-004 No protein
6. SRRM2-007 895 aa
7. SRRM2-011 No protein
8. SRRM2-028 94 aa
9. SRRM2-012 115 aa
10. SRRM2-013 251 aa
11. SRRM2-029 No protein
12. SRRM2-014 No protein
13. SRRM2-030 No protein
14. SRRM2-015 No protein
15. SRRM2-016 No protein
16. SRRM2-017 No protein
17. SRRM2-009 No protein
18. SRRM2-018 No protein
19. SRRM2-019 No protein
20. SRRM2-022 No protein
21. SRRM2-020 184 aa
22. SRRM2-021 No protein
23. SRRM2-023 No protein
24. SRRM2-024 No protein
25. SRRM2-025 No protein
26. SRRM2-026 No protein
27. SRRM2-010 No protein
28. SRRM2-027 No protein
29. SRRM2-031 78 aa
30 SRRM2-032 41 aa
31. SRRM2-033 No protein
Gene name, Ensembl transcript names, GenBank accession numbers and
relative encoded amino acidic protein length of splice variants are report-
ed in the table. Number in the column BTranscript number^identifies the
transcript in Fig. 5
Neurogenetics
investigated in the frontal cortex of PD patients. Their overall
overexpression has been demonstrated when compared to
healthy controls [24,65].
MAPT
MAPT gene encodes the microtubule-associated protein tau, a
protein involved in microtubule assembly and stability [121].
It is located on chromosome 17q21 and contains 15 exons. It
gives rise to multiple splice transcripts (Table 4and Fig. 5)
which are differentially expressed in human tissues [11]. In the
adult human central nervous system, MAPT splicing generates
six tau isoforms composed of either three or four microtubule-
binding repeat motifs in the C-terminal (3R- and 4R-tau).
A number of mutations within and around MAPT exon 10
disrupt exonic and intronic splicing elements as well as the
formation of an RNA stem-loop structure at the 5splice site
(which normally functions to restrict spliceosome assembly).
This event results in an altered ratio of 3R/4R isoforms [10,
13]. The disruption of the balance between them results in
hyperphosphorylation and aggregation of tau proteins into
neurofibrillary tangles, causing the frontotemporal dementia
with parkinsonism linked to chromosome 17 (FTDP-17) [10,
13]. These data support a direct relationship between aberrant
alternative splicing of tau and neuropathology.
GBA
Mutations in β-glucocerebrosidase (GBA) gene cause Gauch-
er disease, a lysosomal storage disease characterized by an
accumulation of glucocerebrosides. Some studies have iden-
tified GBA genetic variants as significant risk factors for the
development of PD [122,15].
GBA gene is located in a gene-rich region on chromosome
1q21. It spans 10.4 kb and contains 12 exons. Currently, there
are four annotated alternative transcripts encoding proteins
(GBA-001, GBA-002, GBA-015, and GBA-016; Table 4
and Fig. 5). Two of them originate from an alternative
Fig. 5 Structures of the alternative splicing variants of the other human
PD-related genes. Structures of the described mRNA splicing variants are
represented in the figure as reported in Ensembl library (http://www.
ensembl.org/index.html). On the left, each variant is indicated with a
number corresponding to that indicated in Table 4. All transcripts are
illustrated in 5-3sense, except MAO-Band GBA genes, which are
illustrated in antisense corresponding to their 3-5sense transcription
Neurogenetics
promoter located 2.6 kb upstream of the first ATG [123]. All
transcripts share the same start codon, with the exception of
GBA-012, whose open-reading frame starts upon exon 4 and
produces a shorter protein isoform. Further transcripts are pro-
duced, but they do not encode proteins. The GBA splicing
profile has not been studied still, and it is unknown if its
alternative splicing is involved in PD.
MAO-B
MAO-Bgene is located on chromosome X and includes 15
exons (Table 4and Fig. 5). Although it is not a confirmed
susceptibility gene [18], increased levels of monoamine oxi-
dase B (MAO) mRNA and enzymatic activity have been re-
ported in platelets from patients with both Parkinsons and
Alzheimers diseases [124]. Furthermore, it is well established
that MAO-B inhibitors delay progression of both pathologies
[125,126].
Several DNA polymorphisms in the MAO-Bgene have
been described in populations with distinct ethnic back-
grounds [124]. A SNP common in all ethnic groups and asso-
ciated with two-fold risk of PD is the G/A dimorphism in
intron 13 sequence [127129]. This SNP does not change
the coding sequence and does not affect the consensus accep-
tor and donor sites. However, it has been demonstrated the G/
A dimorphism in intron 13 sequence creates a splicing en-
hancer that stimulates intron 13 removal and a spliceosomal
complex assembly and alters splicing factorsbinding site
efficiency [124].
SRRM2
Along with cis-acting elements, alternative splicing regulation
relies on trans-splicing factors including the serine/arginine
(SR) proteins. One of these proteins, the RNA splicing factor
SRRM2 (or serine/arginine repetitive matrix 2), has been iden-
tified as the only gene that stood out as differentially expressed
in multiple gene expression PD datasets [130].
SRRM2 gene generates two main alternative splicing tran-
scripts different at their 3end (Table 4and Fig. 5). The full-
length isoform SRRM2-001 contains 15 exons, while the
shorter isoform SRRM2-003 contains 11 exons and lacks
exons 1215. These two isoforms are differentially expressed
in postmortem PD brain regions [130]. The shorter transcript
was upregulated in the substantia nigra but unchanged in the
amygdala of PD patients versus healthy controls. On the con-
trary, the longer transcript was downregulated in both
substantia nigra and amygdala of PDs as compared to con-
trols [130]. Furthermore, in the peripheral blood of patients
with PD, SRRM2 short isoform is overexpressed, while the
expression of longest isoform is reduced [130].
Genome-wide RNA expression analysis reveals
global alternative splicing changes in PD
Although a Bgene-by-gene^approach may simplify splicing
analysis, global alternative splicing changes in PD have to be
considered. The majority of the whole gene expression array
studies in PD brain regions have unfortunately looked at a
single transcript per gene, ignoring the multiple transcripts
generated by alternative splicing [14]. Nonetheless, mRNA
splicing has been identified as a mechanism significantly al-
tered in cortical neurons of PD patients [131].
In order to investigate the splicing expression changes,
some studies have used exon arrays. This kind of approach,
enabling better monitoring and detection of the alternative
splicing events, has allowed to observe significant changes
in overall gene splicing in PD blood cells compared to healthy
controls [130,132]. Another exon array study has been con-
ducted in blood of advanced PD patients prior to and follow-
ing deep brain stimulation neurosurgery, a technique that effi-
ciently improves the motor symptoms of PD [133]. This anal-
ysis has showed preliminary results suggesting brain electrical
stimulation may correlate with significant profile changes in
nonsense-mediated mRNA decay (an mRNA surveillance
process that detects and selectively degrades splice transcripts
harboring premature termination codons) in blood cell tran-
scripts [133]. Potashkin et al. [134]havealsousedspecific
splice variant microarrays in PD patients in order to identify
mRNAs splice transcripts as molecular biomarkers for an ear-
ly PD diagnosis. Through this approach, they identified 13
splice variants with an altered expression in early-stage PD
patients versus healthy controls [134,135].
A recent technology to better study splicing defects is deep
sequencing of RNA (RNAseq) [14]. The advantage of
RNAseq is that it is theoretically feasible to measure both
RNA expression levels and modifications such as splicing.
In addition, RNAseq gives the possibility of discovering novel
transcripts. Whole transcriptome RNAseq data have been ob-
tained from blood leukocytes of PD patientspredeep and
postdeep brain stimulation treatment [136]. This approach
has enabled to discover novel human exons and junctions in
protein-coding RNA molecules, as well as a large range of
differential splicing events pretreatment and posttreatment
compared to healthy controls [136]. Although this is the first
study using in-depth PD transcriptome sequencing, RNAseq
represents a promising technique to better study PD alterna-
tive splicing.
The role of miRNA and lncRNA in PD alternative
splicing modulation
A large number of alternative exon regions have been predict-
ed as binding sites of microRNAs (miRNAs). The latter is a
Neurogenetics
class of small noncoding RNA molecules, which mainly act as
posttranscriptional modulators of multiple target genes by par-
tial sequence complementarity. Through this mechanism, they
may also influence splicing process.
The interplay between miRNA differential expression
and alternative splicing modification in PD has been re-
cently investigated [137]. Parallel changes in miRNA pro-
files and their spliced targets have been observed in PD
leukocytes and PD-relevant brain regions (including the
substantia nigra as well as the frontal lobe). This study
was conducted through coupled analysis of small RNA
sequencing data, splice junction arrays, and exon arrays
[137].
Another novel fascinating class of RNAs with unknown
functions is long noncoding RNAs (lncRNAs), defined as
transcripts of over 200 nucleotides. The GENCODE non-
coding RNA set collects all lncRNAs known so far, in-
cluding several spliced transcript shorter than 200 bp.
LncRNA profiling has been recently assessed in PD leu-
kocytes predeep and postdeep brain stimulation via
RNAseq [136]. This survey allowed to identify some
lncRNAs overexpressed in PD and inversely decreased
following deep brain stimulation [136]. Differentially
expressed lncRNA includes the spliceosome component
U1, supporting the hypothesis of disease-involved splicing
modulations [136].
The identification of existing networks between noncoding
mRNAs and alternative splicing modifications represents an
important step forward the road to understanding the molecu-
lar basis of PD.
Conclusions
Alternative splicing is a highly harmonized process, based
on a combination of DNA sequence motifs, intronic and
exonic elements, regulatory factors, and temporal and spa-
tial signaling pathways. Mutations that disrupt any of these
critical features may alter the finely tuned splicing process-
es, upsetting the production or functions of the encoded
proteins, and finally causing human diseases. Assessing
the alternative splicing modulation of PD-related genes
represents an important point to understand PD molecular
etiology. Future studies, both with the standard or the new
currently available large-scale techniques, will offer a com-
plete data pool of the alternative splicing events in PD and
will provide new possible insights in order to develop strat-
egies for PD therapy and diagnosis.
Acknowledgments This study was supported by the international
Ph.D. program in Neuroscience of the University of Catania. We also
gratefully acknowledge Cristina Calì, Alfia Corsino, Maria Patrizia
DAngelo, and Francesco Marino for their administrative and technical
support.
Open Access This article is distributed under the terms of the Creative
Commons Attribution 4.0 International License (http://
creativecommons.org/licenses/by/4.0/), which permits unrestricted use,
distribution, and reproduction in any medium, provided you give appro-
priate credit to the original author(s) and the source, provide a link to the
Creative Commons license, and indicate if changes were made.
References
1. Ward AJ, Cooper TA (2010) The pathobiology of splicing. J
Pathol 220(2):152163. doi:10.1002/path.2649
2. Yap K, Makeyev EV (2013) Regulation of gene expression in
mammalian nervous system through alternative pre-mRNA splic-
ing coupled with RNA quality control mechanisms. Mol Cell
Neurosci 56:420428. doi:10.1016/j.mcn.2013.01.003
3. Calarco JA, Zhen M, Blencowe BJ (2011) Networking in a global
world: establishing functional connections between neural splic-
ing regulators and their target transcripts. RNA 17(5):775791.
doi:10.1261/rna.2603911
4. Li Q, Lee JA, Black DL (2007) Neuronal regulation of alternative
pre-mRNA splicing. Nat Rev Neurosci 8(11):819831. doi:10.
1038/nrn2237
5. Pan Q, Shai O, Lee LJ, Frey BJ, Blencowe BJ (2008) Deep sur-
veying of alternative splicing complexity in the human tran-
scriptome by high-throughput sequencing. Nat Genet 40(12):
14131415. doi:10.1038/ng.259
6. Behzadnia N, Golas MM, Hartmuth K, Sander B, Kastner B,
Deckert J, Dube P, Will CL, Urlaub H, Stark H, Luhrmann R
(2007) Composition and three-dimensional EM structure of dou-
ble affinity-purified, human prespliceosomal A complexes.
EMBO J 26(6):17371748. doi:10.1038/sj.emboj.7601631
7. Shefer K, Sperling J, Sperling R (2014) The supraspliceosomea
multi-task machine for regulated pre-mRNA processing in the cell
nucleus. Comput Struct Biotechnol J 11(19):113122. doi:10.
1016/j.csbj.2014.09.008
8. Gamazon ER, Stranger BE (2014) Genomics of alternative splic-
ing: evolution, development and pathophysiology. Hum Genet
133(6):679687. doi:10.1007/s00439-013-1411-3
9. Scuderi S, La Cognata V, Drago F, Cavallaro S, DAgata V (2014)
Alternative splicing generates different parkin protein isoforms:
evidences in human, rat, and mouse brain. Biomed Res Int 2014:
690796. doi:10.1155/2014/690796
10. Licatalosi DD, Darnell RB (2006) Splicing regulation in neuro-
logic disease. Neuron 52(1):93101. doi:10.1016/j.neuron.2006.
09.017
11. Tazi J, Bakkour N, Stamm S (2009) Alternative splicing and dis-
ease. Biochim Biophys Acta 1792(1):1426. doi:10.1016/j.
bbadis.2008.09.017
12. Singh RK, Cooper TA (2012) Pre-mRNA splicing in disease and
therapeutics. Trends Mol Med 18(8):472482. doi:10.1016/j.
molmed.2012.06.006
13. Cooper TA, Wan L, Dreyfuss G (2009) RNA and disease. Cell
136(4):777793. doi:10.1016/j.cell.2009.02.011
14. Lewis PA, Cookson MR (2012) Gene expression in the
Parkinsons disease brain. Brain Res Bull 88(4):302312. doi:
10.1016/j.brainresbull.2011.11.016
15. Bonifati V (2014) Genetics of Parkinsonsdiseasestate of the
art, 2013. Parkinsonism Relat Disord 20(1):S23S28. doi:10.
1016/S1353-8020(13)70009-9
Neurogenetics
16. Lubbe S, Morris HR (2014) Recent advances in Parkinsonsdis-
ease genetics. J Neurol 261(2):259266. doi:10.1007/s00415-
013-7003-2
17. Trinh J, Farrer M (2013) Advances in the genetics of Parkinson
disease. Nat Rev Neurol 9(8):445454. doi:10.1038/nrneurol.
2013.132
18. Nalls MA, Pankratz N, Lill CM, Do CB, Hernandez DG, Saad M,
DeStefano AL, Kara E, Bras J, Sharma M, Schulte C, Keller MF,
Arepalli S, Letson C, Edsall C, Stefansson H, Liu X, Pliner H, Lee
JH, Cheng R, International Parkinsons Disease Genomics C,
Parkinsons Study Group Parkinsons Research: The Organized
GI, andMe, GenePd, NeuroGenetics Research C, Hussman
Institute of Human G, Ashkenazi Jewish Dataset I, Cohorts for
H, Aging Research in Genetic E, North American Brain
Expression C, United Kingdom Brain Expression C, Greek
Parkinsons Disease C, Alzheimer Genetic Analysis G, Ikram
MA, Ioannidis JP, Hadjigeorgiou GM, Bis JC, Martinez M,
Perlmutter JS, Goate A, Marder K, Fiske B, Sutherland M,
Xiromerisiou G, Myers RH, Clark LN, Stefansson K, Hardy JA,
Heutink P, Chen H, Wood NW, Houlden H, Payami H, Brice A,
Scott WK, Gasser T, Bertram L, Eriksson N, Foroud T, Singleton
AB (2014) Large-scale meta-analysis of genome-wide association
data identifies six new risk loci for Parkinsons disease. Nat Genet
46(9):989993. doi:10.1038/ng.3043
19. Stefanis L (2012) Alpha-synuclein in Parkinsonsdisease.Cold
Spring Harb Perspect Med 2(2):a009399. doi:10.1101/
cshperspect.a009399
20. Pihlstrom L, Toft M (2011) Genetic variability in SNCA and
Parkinsons disease. Neurogenetics 12(4):283293. doi:10.1007/
s10048-011-0292-7
21. Deng H, Yuan L (2014) Genetic variants and animal models in
SNCA and Parkinson disease.Ageing Res Rev 15C:161176. doi:
10.1016/j.arr.2014.04.002
22. Nuytemans K, Meeus B, Crosiers D, Brouwers N, Goossens D,
Engelborghs S, Pals P, Pickut B, Van den Broeck M, Corsmit E,
Cras P, De Deyn PP, Del-Favero J, Van Broeckhoven C, Theuns J
(2009) Relative contribution of simple mutations vs. copy number
variations in five Parkinson disease genes in the Belgian popula-
tion. Hum Mutat 30(7):10541061. doi:10.1002/humu.21007
23. Beyer K, Domingo-Sabat M, Lao JI, Carrato C, Ferrer I, Ariza A
(2008) Identification and characterization of a new alpha-
synuclein isoform and its role in Lewy body diseases.
Neurogenetics 9(1):1523. doi:10.1007/s10048-007-0106-0
24. Beyer K, Domingo-Sabat M, Humbert J, Carrato C, Ferrer I, Ariza
A (2008) Differential expression of alpha-synuclein, parkin, and
synphilin-1 isoforms in Lewy body disease. Neurogenetics 9(3):
163172. doi:10.1007/s10048-008-0124-6
25. McLean JR, Hallett PJ, Cooper O, Stanley M, Isacson O (2012)
Transcript expression levels of full-length alpha-synuclein and its
three alternatively spliced variants in Parkinsons disease brain
regions and in a transgenic mouse model of alpha-synuclein over-
expression. Mol Cell Neurosci 49(2):230239. doi:10.1016/j.mcn.
2011.11.006
26. Cardo LF, Coto E, de Mena L, Ribacoba R, Mata IF,
Menendez M, Moris G, Alvarez V (2014) Alpha-synuclein
transcript isoforms in three different brain regions from
Parkinsons disease and healthy subjects in relation to the
SNCA rs356165/rs11931074 polymorphisms. Neurosci Lett
562:4549. doi:10.1016/j.neulet.2014.01.009
27. Beyer K, Lao JI, Carrato C, Mate JL, Lopez D, Ferrer I, Ariza A
(2004) Differential expression of alpha-synuclein isoforms in de-
mentia with Lewy bodies. Neuropathol Appl Neurobiol 30(6):
601607. doi:10.1111/j.1365-2990.2004.00572.x
28. Beyer K, Humbert J, Ferrer A, Lao JI, Carrato C, Lopez D, Ferrer
I, Ariza A (2006) Low alpha-synuclein 126 mRNA levels in
dementia with Lewy bodies and Alzheimer disease. Neuroreport
17(12):13271330. doi:10.1097/01.wnr.0000224773.66904.e7
29. McCarthy JJ, Linnertz C, Saucier L, Burke JR, Hulette CM,
Welsh-Bohmer KA, Chiba-Falek O (2011) The effect of SNCA
3region on the levels of SNCA-112 splicing variant.
Neurogenetics 12(1):5964. doi:10.1007/s10048-010-0263-4
30. Kalivendi SV, Yedlapudi D, Hillard CJ, Kalyanaraman B (2010)
Oxidants induce alternative splicing of alpha-synuclein: implica-
tions for Parkinsons disease. Free Radic Biol Med 48(3):377
383. doi:10.1016/j.freeradbiomed.2009.10.045
31. Rhinn H, Qiang L, Yamashita T, Rhee D, Zolin A, Vanti W,
Abeliovich A (2012) Alternative alpha-synuclein transcript usage
as a convergent mechanism in Parkinsons disease pathology. Nat
Commun 3:1084. doi:10.1038/ncomms2032
32. Beyer K (2006) Alpha-synuclein structure, posttranslational mod-
ification and alternative splicing as aggregation enhancers. Acta
Neuropathol 112(3):237251. doi:10.1007/s00401-006-0104-6
33. Bungeroth M, Appenzeller S, Regulin A, Volker W, Lorenzen I,
Grotzinger J, Pendziwiat M, Kuhlenbaumer G (2014) Differential
aggregation properties of alpha-synuclein isoforms. Neurobiol
Aging 35(8):19131919. doi:10.1016/j.neurobiolaging.2014.02.
009
34. Burre J, Sharma M, Tsetsenis T, Buchman V, Etherton MR,
Sudhof TC (2010) Alpha-synuclein promotes SNARE-complex
assembly in vivo and in vitro. Science 329(5999):16631667. doi:
10.1126/science.1195227
35. Rideout HJ, Stefanis L (2014) The neurobiology of LRRK2 and
its role in the pathogenesis of Parkinsons disease. Neurochem Res
39(3):576592. doi:10.1007/s11064-013-1073-5
36. Johnson J, Paisan-Ruiz C, Lopez G, Crews C, Britton A, Malkani
R, Evans EW, McInerney-Leo A, Jain S, Nussbaum RL, Foote
KD, Mandel RJ, Crawley A, Reimsnider S, Fernandez HH, Okun
MS, Gwinn-Hardy K, Singleton AB (2007) Comprehensive
screening of a North American Parkinsons disease cohort for
LRRK2 mutation. Neurodegener Dis 4(5):386391. doi:10.1159/
000105160
37. Skipper L, Shen H, Chua E, Bonnard C, Kolatkar P, Tan LC,
Jamora RD, Puvan K, Puong KY, Zhao Y, Pavanni R, Wong
MC, Yuen Y, Farrer M, Liu JJ, Tan EK (2005) Analysis of
LRRK2 functional domains in nondominant Parkinson disease.
Neurology 65(8):13191321. doi:10.1212/01.wnl.0000180517.
70572.37
38. Zabetian CP, Samii A, Mosley AD, Roberts JW, Leis BC, Yearout
D, Raskind WH, Griffith A (2005) A clinic-based study of the
LRRK2 gene in Parkinson disease yields new mutations.
Neurology 65(5):741744. doi:10.1212/01.wnl.0000172630.
22804.73
39. Shojaee S, Sina F, Farboodi N, Fazlali Z, Ghazavi F, Ghorashi SA,
Parsa K, Sadeghi H, Shahidi GA, Ronaghi M, Elahi E (2009) A
clinic-based screening of mutations in exons 31, 34, 35, 41, and 48
of LRRK2 in Iranian Parkinsons disease patients. Mov Disord
24(7):10231027. doi:10.1002/mds.22503
40. Di Fonzo A, Tassorelli C, De Mari M, Chien HF, Ferreira J, Rohe
CF, Riboldazzi G, Antonini A, Albani G, Mauro A, Marconi R,
Abbruzzese G, Lopiano L, Fincati E, Guidi M, Marini P, Stocchi F,
Onofrj M, Toni V, Tinazzi M, Fabbrini G, Lamberti P, Vanacore N,
Meco G, Leitner P, Uitti RJ, Wszolek ZK, Gasser T, Simons EJ,
Breedveld GJ, Goldwurm S, Pezzoli G, Sampaio C, Barbosa E,
Martignoni E, Oostra BA, Bonifati V, Italian Parkinson
s Genetics
N (2006) Comprehensive analysis of the LRRK2 gene in sixty
families with Parkinsons disease. Eur J Hum Genet 14(3):322
331. doi:10.1038/sj.ejhg.5201539
41. Grimes DA, Racacho L, Han F, Panisset M, Bulman DE (2007)
LRRK2 screening in a Canadian Parkinsons disease cohort. Can J
Neurol Sci 34(3):336338
Neurogenetics
42. Paisan-Ruiz C, Nath P, Washecka N, Gibbs JR, Singleton AB
(2008) Comprehensive analysis of LRRK2 in publicly available
Parkinsons disease cases and neurologically normal controls.
Hum Mutat 29(4):485490. doi:10.1002/humu.20668
43. Lesage S, Condroyer C, Lannuzel A, Lohmann E, Troiano A,
Tison F, Damier P, Thobois S, Ouvrard-Hernandez AM, Rivaud-
Pechoux S, Brefel-Courbon C, Destee A, Tranchant C, Romana
M, Leclere L, Durr A, Brice A, French Parkinsons Disease
Genetics Study G (2009) Molecular analyses of the LRRK2 gene
in European and North African autosomal dominant Parkinsons
disease. J Med Genet 46(7):458464. doi:10.1136/jmg.2008.
062612
44. Trabzuni D, Ryten M, Emmett W, Ramasamy A, Lackner KJ,
Zeller T, Walker R, Smith C, Lewis PA, Mamais A, de Silva R,
Vandrovcova J, International Parkinson Disease Genomics C,
Hernandez D, Nalls MA, Sharma M, Garnier S, Lesage S,
Simon-Sanchez J, Gasser T, Heutink P, Brice A, Singleton A,
Cai H, Schadt E, Wood NW, Bandopadhyay R, Weale ME,
Hardy J, Plagnol V (2013) Fine-mapping, gene expression and
splicing analysis of the disease associated LRRK2 locus. PLoS
ONE 8(8):e70724. doi:10.1371/journal.pone.0070724
45. Giesert F, Hofmann A, Burger A, Zerle J, Kloos K, Hafen U, Ernst
L, Zhang J, Vogt-Weisenhorn DM, Wurst W (2013) Expression
analysis of Lrrk1, Lrrk2 and Lrrk2 splice variants in mice. PLoS
ONE 8(5):e63778. doi:10.1371/journal.pone.0063778
46. Sheng D, Qu D, Kwok KH, Ng SS, Lim AY, Aw SS, Lee CW,
Sung WK, Tan EK, Lufkin T, Jesuthasan S, Sinnakaruppan M, Liu
J (2010) Deletion of the WD40 domain of LRRK2 in zebrafish
causes parkinsonism-like loss of neurons and locomotive defect.
PLoS Genet 6(4):e1000914. doi:10.1371/journal.pgen.1000914
47. Vilarino-Guell C, Wider C, Ross OA, Dachsel JC, Kachergus JM,
Lincoln SJ, Soto-Ortolaza AI, Cobb SA, Wilhoite GJ, Bacon JA,
Behrouz B, Melrose HL, Hentati E, Puschmann A, Evans DM,
Conibear E, Wasserman WW, Aasly JO, Burkhard PR, Djaldetti
R, Ghika J, Hentati F, Krygowska-Wajs A, Lynch T, Melamed E,
Rajput A, Rajput AH, Solida A, Wu RM, Uitti RJ, Wszolek ZK,
Vingerhoets F, Farrer MJ (2011) VPS35 mutations in Parkinson
disease. Am J Hum Genet 89(1):162167. doi:10.1016/j.ajhg.
2011.06.001
48. Zimprich A, Benet-Pages A, Struhal W, Graf E, Eck SH, Offman
MN, Haubenberger D, Spielberger S, Schulte EC, Lichtner P,
Rossle SC, Klopp N, Wolf E, Seppi K, Pirker W, Presslauer S,
Mollenhauer B, Katzenschlager R, Foki T, Hotzy C, Reinthaler E,
Harutyunyan A, Kralovics R, Peters A, Zimprich F, Brucke T,
Poewe W, Auff E, Trenkwalder C, Rost B, Ransmayr G,
Winkelmann J, Meitinger T, Strom TM (2011) A mutation in
VPS35, encoding a subunit of the retromer complex, causes late-
onset Parkinson disease. Am J Hum Genet 89(1):168175. doi:10.
1016/j.ajhg.2011.06.008
49. Verstraeten A, Wauters E, Crosiers D, Meeus B, Corsmit E, Elinck
E, Mattheijssens M, Peeters K, Cras P, Pickut B, Vandenberghe R,
Engelborghs S, De Deyn PP, Van Broeckhoven C (1844) Theuns J
(2012) Contribution of VPS35 genetic variability to LBD in the
FlandersBelgian population. Neurobiol Aging 33(8):e1811
e1843. doi:10.1016/j.neurobiolaging.2012.01.006
50. Bonifati V (2012) Autosomal recessive parkinsonism.
Parkinsonism Relat Disord 18(1):S4S6. doi:10.1016/S1353-
8020(11)70004-9
51. Lucking CB, Durr A, Bonifati V, Vaughan J, De Michele G,
Gasser T, Harhangi BS, Meco G, Denefle P, Wood NW, Agid Y,
Brice A, French Parkinsons Disease Genetics Study G, European
Consortium on Genetic Susceptibility in Parkinsons D (2000)
Association between early-onset Parkinsons disease and muta-
tions in the parkin gene. N Engl J Med 342(21):15601567. doi:
10.1056/NEJM200005253422103
52. Oliveira SA, Scott WK, Martin ER, Nance MA, Watts RL, Hubble
JP, Koller WC, Pahwa R, Stern MB, Hiner BC, Ondo WG, Allen
FH Jr, Scott BL, Goetz CG, Small GW, Mastaglia F, Stajich JM,
ZhangF,BoozeMW,WinnMP,MiddletonLT,HainesJL,
Pericak-Vance MA, Vance JM (2003) Parkin mutations and sus-
ceptibility alleles in late-onset Parkinsons disease. Ann Neurol
53(5):624629. doi:10.1002/ana.10524
53. Illarioshkin SN, Periquet M, Rawal N, Lucking CB,
Zagorovskaya TB, Slominsky PA, Miloserdova OV, Markova
ED, Limborska SA, Ivanova-Smolenskaya IA, Brice A (2003)
Mutation analysis of the parkin gene in Russian families with
autosomal recessive juvenile parkinsonism. Mov Disord 18(8):
914919. doi:10.1002/mds.10467
54. Pigullo S, De Luca A, Barone P, Marchese R, Bellone E,
Colosimo A, Scaglione C, Martinelli P, Di Maria E, Pizzuti A,
Abbruzzese G, Dallapiccola B, Ajmar F, Mandich P (2004)
Mutational analysis of parkin gene by denaturing high-
performance liquid chromatography (DHPLC) in essential tremor.
Parkinsonism Relat Disord 10(6):357362. doi:10.1016/j.
parkreldis.2004.04.012
55. Scherfler C, Khan NL, Pavese N, Eunson L, Graham E, Lees AJ,
Quinn NP, Wood NW, Brooks DJ, Piccini PP (2004) Striatal and
cortical pre- and postsynaptic dopaminergic dysfunction in spo-
radic parkin-linked parkinsonism. Brain 127(Pt 6):13321342.
doi:10.1093/brain/awh150
56. Bertoli-Avella AM, Giroud-Benitez JL, Akyol A, Barbosa E,
Schaap O, van der Linde HC, Martignoni E, Lopiano L,
Lamberti P, Fincati E, Antonini A, Stocchi F, Montagna P,
Squitieri F, Marini P, Abbruzzese G, Fabbrini G, Marconi R,
Dalla Libera A, Trianni G, Guidi M, De Gaetano A, Boff
Maegawa G, De Leo A, Gallai V, de Rosa G, Vanacore N, Meco
G, van Duijn CM, Oostra BA, Heutink P, Bonifati V, Italian
Parkinson Genetics N (2005) Novel parkin mutations detected in
patients with early-onset Parkinsons disease. Mov Disord 20(4):
424431. doi:10.1002/mds.20343
57. Bardien S, Keyser R, Yako Y, Lombard D, Carr J (2009)
Molecular analysis of the parkin gene in South African patients
diagnosed with Parkinsons disease. Parkinsonism Relat Disord
15(2):116121. doi:10.1016/j.parkreldis.2008.04.005
58. Pankratz N, Kissell DK, Pauciulo MW, Halter CA, Rudolph A,
Pfeiffer RF, Marder KS, Foroud T, Nichols WC, Parkinson Study
Group PI (2009) Parkin dosage mutations have greater pathoge-
nicity in familial PD than simple sequence mutations. Neurology
73(4):279286. doi:10.1212/WNL.0b013e3181af7a33
59. La Cognata V, Iemmolo R, DAgata V, Scuderi S, Drago F, Zappia
M, Cavallaro S (2014) Increasing the coding potential of genomes
through alternative splicing: the case of PARK2 gene. Curr Genomics
15(3):203216. doi:10.2174/1389202915666140426003342
60. Kitada T, Asakawa S, Hattori N, Matsumine H, Yamamura Y,
Minoshima S, Yokochi M, Mizuno Y, Shimizu N (1998)
Mutations in the parkin gene cause autosomal recessive juvenile
parkinsonism. Nature 392(6676):605608
61. Matsumine H, Saito M, Shimoda-Matsubayashi S, Tanaka H,
Ishikawa A, Nakagawa-Hattori Y, Yokochi M, Kobayashi T,
Igarashi S, Takano H, Sanpei K, Koike R, Mori H, Kondo T,
Mizutani Y, Schaffer AA, Yamamura Y, Nakamura S, Kuzuhara
S, Tsuji S, Mizuno Y (1997) Localization of a gene for an auto-
somal recessive form of juvenile parkinsonism to chromosome
6q25.227. Am J Hum Genet 60(3):588596
62. Xu L, Lin DC, Yin D, Koeffler HP (2014) An emerging role of
PARK2 in cancer. J Mol Med (Berl) 92(1):3142. doi:10.1007/
s00109-013-1107-0
63. Ikeuchi K, Marusawa H, Fujiwara M, Matsumoto Y, Endo Y,
Watanabe T, Iwai A, Sakai Y, Takahashi R, Chiba T (2009)
Attenuation of proteolysis-mediated cyclin E regulation by
Neurogenetics
alternatively spliced parkin in human colorectal cancers. Int J
Cancer 125(9):20292035. doi:10.1002/ijc.24565
64. Kitada T, Asakawa S, Minoshima S, Mizuno Y, Shimizu N (2000)
Molecular cloning, gene expression, and identification of a splic-
ing variant of the mouse parkin gene. Mamm Genome 11(6):417
421
65. Humbert J, Beyer K, Carrato C, Mate JL, Ferrer I, Ariza A (2007)
Parkin and synphilin-1 isoform expression changes in Lewy body
diseases. Neurobiol Dis 26(3):681687. doi:10.1016/j.nbd.2007.
03.007
66. Tan EK, Shen H, Tan JM, Lim KL, Fook-Chong S, Hu WP,
Paterson MC, Chandran VR, Yew K, Tan C, Yuen Y, Pavanni R,
Wong MC, Puvan K, Zhao Y (2005) Differential expression of
splice variant and wild-type parkin in sporadic Parkinsonsdis-
ease. Neurogenetics 6(4):179184. doi:10.1007/s10048-005-
0001-5
67. Dagata V, Cavallaro S (2004) Parkin transcript variants in rat and
human brain. Neurochem Res 29(9):17151724
68. Sunada Y, Saito F, Matsumura K, Shimizu T (1998) Differential
expression of the parkin gene in the human brain and peripheral
leukocytes. Neurosci Lett 254(3):180182
69. Solano SM, Miller DW, Augood SJ, Young AB, Penney JB Jr
(2000) Expression of alpha-synuclein, parkin, and ubiquitin
carboxy-terminal hydrolase L1 mRNA in human brain: genes as-
sociated with familial Parkinsons disease. Ann Neurol 47(2):201
210
70. Kasap M, Akpinar G, Sazci A, Idrisoglu HA, Vahaboglu H (2009)
Evidence for the presence of full-length PARK2 mRNA and
parkin protein in human blood. Neurosci Lett 460(3):196200.
doi:10.1016/j.neulet.2009.05.079
71. Pawlyk AC, Giasson BI, Sampathu DM, Perez FA, Lim KL,
Dawson VL, Dawson TM, Palmiter RD, Trojanowski JQ, Lee
VM (2003) Novel monoclonal antibodies demonstrate biochemi-
cal variation of brain parkin with age. J Biol Chem 278(48):
4812048128. doi:10.1074/jbc.M306889200
72. Horowitz JM, Myers J, Stachowiak MK, Torres G (1999)
Identification and distribution of Parkin in rat brain. Neuroreport
10(16):33933397
73. Stichel CC, Augustin M, Kuhn K, Zhu XR, Engels P, Ullmer C,
Lubbert H (2000) Parkin expression in the adult mouse brain. Eur
J Neurosci 12(12):41814194
74. Gu WJ, Abbas N, Lagunes MZ, Parent A, Pradier L, Bohme GA,
Agid Y, Hirsch EC, Raisman-Vozari R, Brice A (2000) Cloning of
rat parkin cDNA and distributionofparkininratbrain.J
Neurochem 74(4):17731776
75. Huynh DP, Dy M, Nguyen D, Kiehl TR, Pulst SM (2001)
Differential expression and tissue distribution of parkin isoforms
during mouse development. Brain Res Dev Brain Res 130(2):
173181
76. DAgata V, Grimaldi M, Pascale A, Cavallaro S (2000) Regional
and cellular expression of the parkin gene in the rat cerebral cor-
tex. Eur J Neurosci 12(10):35833588
77. Pogson JH, Ivatt RM, Whitworth AJ (2011) Molecular mecha-
nisms of PINK1-related neurodegeneration. Curr Neurol
Neurosci Rep 11(3):283290. doi:10.1007/s11910-011-0187-x
78. Nuytemans K, Theuns J, Cruts M, Van Broeckhoven C (2010)
Genetic etiology of Parkinson disease associated with mutations
in the SNCA, PARK2, PINK1, PARK7, and LRRK2 genes: a
mutation update. Hum Mutat 31(7):763780. doi:10.1002/humu.
21277
79. Marongiu R, Brancati F, Antonini A, Ialongo T, Ceccarini C,
Scarciolla O, Capalbo A, Benti R, Pezzoli G, Dallapiccola B,
Goldwurm S, Valente EM (2007) Whole gene deletion and splic-
ing mutations expand the PINK1 genotypic spectrum. Hum Mutat
28(1):98. doi:10.1002/humu.9472
80. Samaranch L, Lorenzo-Betancor O, Arbelo JM, Ferrer I, Lorenzo
E, Irigoyen J, Pastor MA, Marrero C, Isla C, Herrera-Henriquez J,
Pastor P (2010) PINK1-linked parkinsonism is associated with
Lewy body pathology. Brain 133(Pt 4):11281142. doi:10.1093/
brain/awq051
81. Valente EM, Abou-Sleiman PM, Caputo V, Muqit MM, Harvey
K, Gispert S, Ali Z, Del Turco D, Bentivoglio AR, Healy DG,
Albanese A, Nussbaum R, Gonzalez-Maldonado R, Deller T,
Salvi S, Cortelli P, Gilks WP, Latchman DS, Harvey RJ,
Dallapiccola B, Auburger G, Wood NW (2004) Hereditary
early-onset Parkinsons disease caused by mutations in PINK1.
Science 304(5674):11581160. doi:10.1126/science.1096284
82. Silvestri L, Caputo V, Bellacchio E, Atorino L, Dallapiccola B,
Valente EM, Casari G (2005) Mitochondrial import and enzymatic
activity of PINK1 mutants associated to recessive parkinsonism.
Hum Mol Genet 14(22):34773492. doi:10.1093/hmg/ddi377
83. Beilina A, Van Der Brug M, Ahmad R, Kesavapany S, Miller DW,
Petsko GA, Cookson MR (2005) Mutations in PTEN-induced
putative kinase 1 associated with recessive parkinsonism have
differential effects on protein stability. Proc Natl Acad Sci U S A
102(16):57035708. doi:10.1073/pnas.0500617102
84. dAmora M, Angelini C, MarcoliM,CervettoC,KitadaT,
Vallarino M (2011) Expression of PINK1 in the brain, eye and
ear of mouse during embryonic development. J Chem Neuroanat
41(2):7385. doi:10.1016/j.jchemneu.2010.11.004
85. Lin W, Kang UJ (2008) Characterization of PINK1 processing,
stability, and subcellular localization. J Neurochem 106(1):464
474. doi:10.1111/j.1471-4159.2008.05398.x
86. Deas E, Plun-Favreau H, Gandhi S, Desmond H,Kjaer S, Loh SH,
Renton AE, Harvey RJ, Whitworth AJ, Martins LM, Abramov
AY, Wood NW (2011) PINK1 cleavage at position A103 by the
mitochondrial protease PARL. Hum Mol Genet 20(5):867879.
doi:10.1093/hmg/ddq526
87. Lockhart PJ, Lincoln S, Hulihan M, Kachergus J, Wilkes K,
Bisceglio G, Mash DC, Farrer MJ (2004) DJ-1 mutations are a
rare cause of recessively inherited early onset parkinsonism medi-
ated by loss of protein function. J Med Genet 41(3):e22
88. Moore DJ, West AB, Dawson VL, Dawson TM (2005) Molecular
pathophysiology of Parkinsons disease. Annu Rev Neurosci 28:
5787. doi:10.1146/annurev.neuro.28.061604.135718
89. Bonifati V, Rizzu P, Squitieri F, Krieger E, Vanacore N, van
Swieten JC, Brice A, van Duijn CM, Oostra B, Meco G,
Heutink P (2003) DJ-1( PARK7), a novel gene for autosomal
recessive, early onset parkinsonism. Neurol Sci 24(3):159160.
doi:10.1007/s10072-003-0108-0
90. van Duijn CM, Dekker MC, Bonifati V, Galjaard RJ, Houwing-
Duistermaat JJ, Snijders PJ, Testers L, Breedveld GJ, Horstink M,
Sandkuijl LA, van Swieten JC, Oostra BA, Heutink P (2001)
Park7, a novel locus for autosomal recessive early-onset parkin-
sonism, on chromosome 1p36. Am J Hum Genet 69(3):629634.
doi:10.1086/322996
91. Corti O, Lesage S, Brice A (2011) What genetics tells us about the
causes and mechanisms of Parkinsons disease. Physiol Rev 91(4):
11611218. doi:10.1152/physrev.00022.2010
92. Tarantino P, Civitelli D, Annesi F, De Marco EV, Rocca FE,
Pugliese P, Nicoletti G, Carrideo S, Provenzano G, Annesi G,
Quattrone A (2009) Compound heterozygosity in DJ-1 gene
non-coding portion related to parkinsonism. Parkinsonism Relat
Disord 15(4):324326. doi:10.1016/j.parkreldis.2008.07.001
93. Hedrich K, Djarmati A, Schafer N, Hering R, Wellenbrock C,
Weiss PH, Hilker R, Vieregge P, Ozelius LJ, Heutink P, Bonifati
V, Schwinger E, Lang AE, Noth J, Bressman SB, Pramstaller PP,
Riess O, Klein C (2004) DJ-1 (PARK7) mutations are less fre-
quent than parkin (PARK2) mutations in early-onset Parkinson
disease. Neurology 62(3):389394
Neurogenetics
94. Lev N, Roncevic D, Ickowicz D, Melamed E, Offen D (2006)
Role of DJ-1 in Parkinsons disease. J Mol Neurosci 29(3):215
225
95. Ariga H, Takahashi-Niki K, Kato I, Maita H, Niki T, Iguchi-Ariga
SM (2013) Neuroprotective function of DJ-1 in Parkinsonsdis-
ease. Oxid Med Cell Longev 2013:683920. doi:10.1155/2013/
683920
96. Xu J, Zhong N, Wang H, Elias JE, Kim CY, Woldman I, Pifl C,
Gygi SP, Geula C, Yankner BA (2005) The Parkinsons disease-
associated DJ-1 protein is a transcriptional co-activator that pro-
tects against neuronal apoptosis. Hum Mol Genet 14(9):1231
1241. doi:10.1093/hmg/ddi134
97. Bandopadhyay R, Kingsbury AE, Cookson MR, Reid AR, Evans
IM, Hope AD, Pittman AM, Lashley T, Canet-Aviles R, Miller
DW, McLendon C, Strand C, Leonard AJ, Abou-Sleiman PM,
Healy DG, Ariga H, Wood NW, de Silva R, Revesz T, Hardy
JA, Lees AJ (2004) The expression of DJ-1 (PARK7) in normal
human CNS and idiopathic Parkinsons disease. Brain 127(Pt 2):
420430. doi:10.1093/brain/awh054
98. Besong Agbo D, Klafki H, Poschmann G, Seyfarth K, Genius J,
Janssen C, Stuhler K, Wurst W, Meyer HE, Klingenspor M,
Wiltfang J (2013) Development of a capillary isoelectric focusing
immunoassay to measure DJ-1 isoforms in biological samples.
Anal Biochem 443(2):197204. doi:10.1016/j.ab.2013.09.013
99. Kumaran R, Kingsbury A, Coulter I, Lashley T, Williams D, de
Silva R, Mann D, Revesz T, Lees A, Bandopadhyay R (2007) DJ-
1 (PARK7) is associated with 3R and 4R tau neuronal and glial
inclusions in neurodegenerative disorders. Neurobiol Dis 28(1):
122132. doi:10.1016/j.nbd.2007.07.012
100. Kumaran R, Vandrovcova J, Luk C, Sharma S, Renton A, Wood
NW, Hardy JA, Lees AJ, Bandopadhyay R (2009) Differential DJ-
1 gene expression in Parkinsons disease. Neurobiol Dis 36(2):
393400. doi:10.1016/j.nbd.2009.08.011
101. Lin X, Cook TJ, Zabetian CP, Leverenz JB, Peskind ER, Hu SC,
Cain KC, Pan C,Edgar JS, Goodlett DR, Racette BA, Checkoway
H, Montine TJ, Shi M, Zhang J (2012) DJ-1 isoforms in whole
blood as potential biomarkers of Parkinson disease. Sci Rep 2:954.
doi:10.1038/srep00954
102. Zhong N, Kim CY, Rizzu P, Geula C, Porter DR, Pothos EN,
Squitieri F, Heutink P, Xu J (2006) DJ-1 transcriptionally up-
regulates the human tyrosine hydroxylase by inhibiting the
sumoylation of pyrimidine tract-binding protein-associated splic-
ing factor. J Biol Chem 281(30):2094020948. doi:10.1074/jbc.
M601935200
103. Vilarino-Guell C, Soto AI, Lincoln SJ, Ben Yahmed S, Kefi M,
Heckman MG, Hulihan MM, Chai H, Diehl NN, Amouri R,
Rajput A, Mash DC, Dickson DW, Middleton LT, Gibson RA,
Hentati F, Farrer MJ (2009) ATP13A2 variability in Parkinson
disease. Hum Mutat 30(3):406410. doi:10.1002/humu.20877
104. Murphy KE, Cottle L, Gysbers AM, Cooper AA, Halliday GM
(2013) ATP13A2 (PARK9) protein levels are reduced in brain
tissue of cases with Lewy bodies. Acta Neuropathol Commun
1(1):11. doi:10.1186/2051-5960-1-11
105. Ugolino J, Fang S, Kubisch C, Monteiro MJ (2011) Mutant
Atp13a2 proteins involved in parkinsonism are degraded by ER-
associated degradation and sensitize cells to ER-stress induced
cell death. Hum Mol Genet 20(18):35653577. doi:10.1093/
hmg/ddr274
106. Lu CS, Lai SC, Wu RM, Weng YH, Huang CL, Chen RS, Chang
HC, Wu-Chou YH, Yeh TH (2012) PLA2G6 mutations in
PARK14-linked young-onset parkinsonism and sporadic
Parkinsons disease. Am J Med Genet B Neuropsychiatr Genet
159B(2):183191. doi:10.1002/ajmg.b.32012
107. Di Fonzo A, Dekker MC, Montagna P, Baruzzi A, Yonova EH,
Correia Guedes L, Szczerbinska A,Zhao T, Dubbel-Hulsman LO,
Wouters CH, de Graaff E, Oyen WJ, Simons EJ, Breedveld GJ,
Oostra BA, Horstink MW, Bonifati V (2009) FBXO7 mutations
cause autosomal recessive, early-onset parkinsonian-pyramidal
syndrome. Neurology 72(3):240245. doi:10.1212/01.wnl.
0000338144.10967.2b
108. Deng H, Liang H, Jankovic J (2013) F-box only protein 7 gene in
parkinsonian-pyramidal disease. JAMA Neurol 70(1):2024. doi:
10.1001/jamaneurol.2013.572
109. Gomez-Garre P, Jesus S, Carrillo F, Caceres-Redondo MT,
Huertas-Fernandez I, Bernal-Bernal I, Bonilla-Toribio M,
Vargas-Gonzalez L, Carballo M, Mir P (2014) Systematic muta-
tional analysis of FBXO7 in a Parkinsons disease population
from southern Spain. Neurobiol Aging 35(3):727. e57. doi:10.
1016/j.neurobiolaging.2013.09.011
110. Ilyin GP, Rialland M, Pigeon C, Guguen-Guillouzo C (2000)
cDNA cloning and expression analysis of new members of the
mammalian F-box protein family. Genomics 67(1):4047. doi:
10.1006/geno.2000.6211
111. Zhao T, De Graaff E, Breedveld GJ, Loda A, Severijnen LA,
Wouters CH, Verheijen FW, Dekker MC, Montagna P,
Willemsen R, Oostra BA, Bonifati V (2011) Loss of nuclear ac-
tivity of the FBXO7 protein in patients with parkinsonian-
pyramidal syndrome (PARK15). PLoS ONE 6(2):e16983. doi:
10.1371/journal.pone.0016983
112. Nolte D, Niemann S, Muller U (2003) Specific sequence changes
in multiple transcript system DYT3 are associated with X-linked
dystonia parkinsonism. Proc Natl Acad Sci U S A 100(18):10347
10352. doi:10.1073/pnas.1831949100
113. Herzfeld T, Nolte D, Grznarova M, Hofmann A, Schultze JL,
Muller U (2013) X-linked dystonia parkinsonism syndrome
(XDP, lubag): disease-specific sequence change DSC3 in TAF1/
DYT3 affects genes in vesicular transport and dopamine metabo-
lism. Hum Mol Genet 22(5):941951. doi:10.1093/hmg/dds499
114. Korvatska O, Strand NS, Berndt JD, Strovas T, Chen DH,
Leverenz JB, Kiianitsa K, Mata IF, Karakoc E, Greenup JL,
Bonkowski E, Chuang J, Moon RT, Eichler EE, Nickerson DA,
Zabetian CP, Kraemer BC, Bird TD, Raskind WH (2013) Altered
splicing of ATP6AP2 causes X-linked parkinsonism with spastic-
ity (XPDS). Hum Mol Genet 22(16):32593268. doi:10.1093/
hmg/ddt180
115. Myhre R, Klungland H, Farrer MJ, Aasly JO (2008) Genetic as-
sociation study of synphilin-1 in idiopathic Parkinsons disease.
BMC Med Genet 9:19. doi:10.1186/1471-2350-9-19
116. Keyser RJ, Oppon E, Carr JA, Bardien S (2011) Identification of
Parkinsons disease candidate genes using CAESAR and screen-
ing of MAPT and SNCAIP in South African Parkinsonsdisease
patients. J Neural Transm 118(6):889897. doi:10.1007/s00702-
011-0591-z
117. Marx FP, Holzmann C, Strauss KM, Li L, Eberhardt O, Gerhardt
E, Cookson MR, Hernandez D, Farrer MJ, Kachergus J,
Engelender S, Ross CA, Berger K, Schols L, Schulz JB, Riess
O, Kruger R (2003) Identification and functional characterization
of a novel R621C mutation in the synphilin-1 gene in Parkinsons
disease. Hum Mol Genet 12(11):12231231
118. Eyal A, Szargel R, Avraham E, Liani E, Haskin J, Rott R,
Engelender S (2006) Synphilin-1A: an aggregation-prone isoform
of synphilin-1 that causes neuronal death and is present in aggre-
gates from alpha-synucleinopathy patients. Proc Natl Acad Sci U
S A 103(15):59175922. doi:10.1073/pnas.0509707103
119. Eyal A, Engelender S (2006)Synphilin isoforms and the search for
a cellular model of Lewy body formation in Parkinsons disease.
Cell Cycle 5(18):20822086
120. Szargel R, Rott R, Engelender S (2008) Synphilin-1 isoforms in
Parkinsons disease: regulation by phosphorylation and
ubiquitylation. Cell Mol Life Sci 65(1):8088. doi:10.1007/
s00018-007-7343-0
Neurogenetics
121. Lei P, Ayton S, Finkelstein DI, Adlard PA, Masters CL, Bush AI
(2010) Tau protein: relevance to Parkinsonsdisease.IntJ
Biochem Cell Biol 42(11):17751778. doi:10.1016/j.biocel.
2010.07.016
122. Spatola M, Wider C (2014) Genetics of Parkinsons disease: the
yield. Parkinsonism Relat Disord 20(1):S35S38. doi:10.1016/
S1353-8020(13)70011-7
123. Svobodova E, Mrazova L, Luksan O, Elstein D, Zimran A,
Stolnaya L, Minks J, Eberova J, Dvorakova L, Jirsa M,
Hrebicek M (2011) Glucocerebrosidase gene has an alternative
upstream promoter, which has features and expression character-
istic of housekeeping genes. Blood Cells Mol Dis 46(3):239245.
doi:10.1016/j.bcmd.2010.12.011
124. Jakubauskiene E, Janaviciute V, Peciuliene I, Soderkvist P,
Kanopka A (2012) G/A polymorphism in intronic sequence af-
fects the processing of MAO-B gene in patients with Parkinson
disease. FEBS Lett 586(20):36983704. doi:10.1016/j.febslet.
2012.08.028
125. Thomas T (2000) Monoamine oxidase-B inhibitors in the treat-
ment of Alzheimers disease. Neurobiol Aging 21(2):343348
126. Stern G (1998) Neuroprotection by selegiline and other MAO
inhibitors. J Neural Transm Suppl 52:99107
127. Ho SL, Kapadi AL, Ramsden DB, Williams AC (1995) An
allelic association study of monoamine oxidase B in
Parkinsons disease. Ann Neurol 37(3):403405. doi:10.
1002/ana.410370318
128. Kurth JH, Kurth MC, Poduslo SE, Schwankhaus JD (1993)
Association of a monoamine oxidase B allele with Parkinsons
disease. Ann Neurol 33(4):368372. doi:10.1002/ana.410330406
129. Sobell JL, Lind TJ, Hebrink DD, Heston LL, Sommer SS (1997)
Screening the monoamine oxidase B gene in 100 male patients
with schizophrenia: a cluster of polymorphisms in African-
Americans but lack of functionally significant sequence changes.
Am J Med Genet 74(1):4449
130. Shehadeh LA, Yu K, Wang L, Guevara A, Singer C, Vance J,
Papapetropoulos S (2010) SRRM2, a potential blood biomarker
revealing high alternative splicing in Parkinsonsdisease.PLoS
ONE 5(2):e9104. doi:10.1371/journal.pone.0009104
131. Stamper C, Siegel A, Liang WS, Pearson JV, Stephan DA, Shill H,
Connor D, Caviness JN, Sabbagh M, Beach TG, Adler CH,
Dunckley T (2008) Neuronal gene expression correlates of
Parkinsons disease with dementia. Mov Disord 23(11):1588
1595. doi:10.1002/mds.22184
132. Soreq L, Bergman H, Israel Z, Soreq H (2012) Exon arrays reveal
alternative splicing aberrations in Parkinsons disease leukocytes.
Neurodegener Dis 10(14):203206. doi:10.1159/000332598
133. Soreq L, Bergman H, Israel Z, Soreq H (2013) Deep brain stimu-
lation modulates nonsense-mediated RNA decay in Parkinsons
patients leukocytes. BMC Genomics 14:478. doi:10.1186/1471-
2164-14-478
134. Potashkin JA, Santiago JA, Ravina BM, Watts A, Leontovich AA
(2012) Biosignatures for Parkinsons disease and atypical parkin-
sonian disorders patients. PLoS ONE 7(8):e43595. doi:10.1371/
journal.pone.0043595
135. Santiago JA, Scherzer CR, Harvard Biomarker S, Potashkin
JA (2013) Specific splice variants are associated with
Parkinsons disease. Mov Disord 28(12):17241727. doi:10.
1002/mds.25635
136. Soreq L, Guffanti A, Salomonis N, Simchovitz A, Israel Z,
Bergman H, Soreq H (2014) Long non-coding RNA and alterna-
tive splicing modulations in Parkinsons leukocytes identified by
RNA sequencing. PLoS Comput Biol 10(3):e1003517. doi:10.
1371/journal.pcbi.1003517
137. Soreq L, Salomonis N, Bronstein M, Greenberg DS, Israel Z,
Bergman H, Soreq H (2013) Small RNA sequencing-
microarray analyses in Parkinson leukocytes reveal deep
brain stimulation-induced splicing changes that classify brain
region transcriptomes. Front Mol Neurosci 6:10. doi:10.3389/
fnmol.2013.00010
Neurogenetics
... In this way, such a structural modification results in impairment of interaction between MALAT1 and SRSF2, leading to the altered splicing and the expression of target genes under its control [49]. Pre-mRNA splicing events play a key role in neuronal differentiation as well as synaptic maturation for normal brain development, and also their dysregulations have been proven in the pathogenesis of Parkinson's [54][55][56]. Therefore, it seems that the rs3200401C>T is an important functional SNP with significant biological effects on lncRNA MALAT1, especially in neurons. ...
Article
Full-text available
Research findings show that genetic susceptibility to sporadic Parkinson’s disease (PD), a common neurodegenerative disorder, is determined through gene variation of loci involved in its development and pathogenesis. A growing body of strong evidence has revealed that dysfunction of long non-coding RNAs (lncRNAs) plays key roles in the pathogenesis and progression of PD through impairing neuronal signaling pathways, but little is known about the relationship between their variants and PD susceptibility. In this research, we intended to study the relationship between functional SNPs rs12826786C>T, rs3200401C>T, and rs6931097G>A in the key lncRNAs stimulating neuroapoptosis and neuroinflammation in PD, including HOTAIR, MALAT1, and lincRNA-P21, respectively, with susceptibility to PD as well as its clinical symptoms. The population of this study consisted of 240 individuals, including 120 controls and 120 cases, and the sample taken from them was peripheral blood. Genotyping of the target SNPs was done using PCR-RFLP. We found that the healthy individuals carry more T allele of MALAT1-rs3200401C>T compared to the patients (P= 0.019). Furthermore, it was observed that in the dominant genetic model, subjects with genotypes carrying the T allele have a lower risk of PD (OR= 0.530; CI= 0.296–0.950; P= 0.033). Regarding the lincRNA-P21-rs6931097G>A, we observed a significant protective relationship between its GA (OR= 0.144; CI= 0.030–0.680; P= 0.014) and AA (OR= 0.195; CI= 00.047–0.799; P= 0.023) genotypes with the manifestation of tremor and bradykinesia symptoms, respectively. Furthermore, the findings indicated that the minor TT genotype of HOTAIR-rs12826786C>T was significantly associated with a reduced risk of bradykinesia symptoms (OR= 0.147; CI= 0.039–0.555; P= 0.005). Collectively, these findings suggest that MALAT1-rs3200401C>T may be an important lncRNA SNP against the development of PD, while the other two SNPs show protective effects on the clinical manifestations of PD in a way that lincRNA-P21-rs6931097G>A has a protective effect against the occurrence of tremor and bradykinesia symptoms in PD patients, and HOTAIR -rs12826786C>T indicates a protective effect against the display of bradykinesia feature. Therefore, they can have valuable potential as biomarkers for clinical evaluations of this disease.
... Regarding PD, this protein was identified as a splicing regulator engaged in the prevention of the SNCA-112 (112-syn; SNCA is the gene coding for α-syn) generation. Of note, 112-syn is an α-syn spliced isoform without exon 5, which is overexpressed in the brain of PD patients and has been linked with PD progression (La Cognata et al., 2015;Thonda et al., 2022). A downregulation of RBMX was shown in HEK-293 cells treated with the neurotoxin 1-methyl-4-phenyl pyridinium ion (MPP + ), a well-known inducer of parkinsonism, suggesting the pivotal role of this protein in PD pathophysiology. ...
Article
Full-text available
Aging represents the major risk factor for the onset and/or progression of various disorders including neuro-degenerative diseases, metabolic disorders, and bone-related defects. As the average age of the population is predicted to exponentially increase in the coming years, understanding the molecular mechanisms underlying the development of aging-related diseases and the discovery of new therapeutic approaches remain pivotal. Well-reported hallmarks of aging are cellular senescence, genome instability, autophagy impairment, mitochondria dysfunction, dysbiosis, telomere attrition, metabolic dysregulation, epigenetic alterations, low-grade chronic inflammation, stem cell exhaustion, altered cell-to-cell communication and impaired proteostasis. With few exceptions, however, many of the molecular players implicated within these processes as well as their role in disease development remain largely unknown. RNA binding proteins (RBPs) are known to regulate gene expression by dictating at post-transcriptional level the fate of nascent transcripts. Their activity ranges from directing primary mRNA maturation and trafficking to modulation of transcript stability and/or translation. Accumulating evidence has shown that RBPs are emerging as key regulators of aging and aging-related diseases, with the potential to become new diagnostic and therapeutic tools to prevent or delay aging processes. In this review, we summarize the role of RBPs in promoting cellular senescence and we highlight their dysregulation in the pathogenesis and progression of the main aging-related diseases, with the aim of encouraging further investigations that will help to better disclose this novel and captivating molecular scenario.
... Among the most signi cant DSGs identi ed were known TDP-43 splicing targets (e.g., Kcnip2, Pdp1, Poldipp3, Pp bp1, Dnajc5. Tmem2, Sort1), 24,94,97 transcripts associated with particular neurodegenerative diseases (e.g., Mapt, Atxn1, Lrrk2), [98][99][100][101][102][103][104] and also many robustly altered transcripts that are poorly characterized but linked to neurodegeneration (e.g., Nrxn3, Nos1, Arfgef2, Arhgap10, Lrp8, Smarca4, Rims2). Of all identi ed DSGs, the most substantially altered transcript was Sort1, encoding the Sortilin-1 (Sort1) ...
Preprint
Full-text available
TDP-43 proteinopathies including frontotemporal lobar degeneration (FTLD) and amyotrophic lateral sclerosis (ALS) are neurodegenerative disorders characterized by aggregation and mislocalization of the nucleic-acid binding protein TDP-43 and subsequent neuronal dysfunction. Here, we developed endogenous models of sporadic TDP-43 proteinopathy based on the principle that disease-associated TDP-43 acetylation at lysine 145 (K145) alters TDP-43 conformation, impairs RNA-binding capacity, and induces downstream mis-regulation of target genes. Expression of acetylation-mimic TDP-43 K145Q resulted in stress-induced nuclear TDP-43 foci and loss of TDP-43 function in primary mouse and human induced pluripotent stem cell (hiPSC)-derived cortical neurons. Mice harboring the TDP-43 K145Q mutation recapitulated key hallmarks of FTLD, including progressive TDP-43 phosphorylation and insolubility, TDP-43 mis-localization, transcriptomic and splicing alterations, and cognitive dysfunction. Our study supports a model in which TDP-43 acetylation drives neuronal dysfunction and cognitive decline through aberrant splicing and transcription of critical genes that regulate synaptic plasticity and stress response signaling. The neurodegenerative cascade initiated by TDP-43 acetylation recapitulates many aspects of human FTLD and provides a new paradigm to further interrogate TDP-43 proteinopathies.
Article
Eukaryotic elongation factor 1A1 (eEF1A1) is an RNA-binding protein that is associated with PARK2 activity in cells, suggesting a possible role in Parkinson’s disease (PD). To clear whether eEF1A1 plays a role in PD through transcriptional or posttranscriptional regulation. The GSE68719 dataset was downloaded from the GEO database, and the RNA-seq data of all brain tissue autopsies were obtained from 29 PD patients and 44 neurologically normal control subjects. To inhibit eEF1A1 from being expressed in U251 cells, siRNA was transfected into those cells, and RNA-seq high-throughput sequencing was used to determine the differentially expressed genes (DEGs) and differentially alternative splicing events (ASEs) resulting from eEF1A1 knockdown. eEF1A1 was significantly overexpressed in PD brain tissue in the BA9 area. GO and KEGG enrichment analyses revealed that eEF1A1 knockdown significantly upregulated the expression of the genes CXCL10, NGF, PTX3, IL6, ST6GALNAC3, NUPR1, TNFRSF21, and CXCL2 and upregulated the alternative splicing of the genes ACOT7, DDX10, SHMT2, MYEF2, and NDUFAF5. These genes were enriched in pathways related to PD pathogenesis, such as apoptosis, inflammatory response, and mitochondrial dysfunction. The results suggesting that eEF1A1 involved in the development of PD by regulating the differential expression and alternative splicing of genes, providing a theoretical basis for subsequent research.
Article
The role of alternative splicing in Chronic Obstructive Pulmonary Disease (COPD) is still largely unknown. We aimed to investigate the differences in alternatively splicing events between patients with mild-to-moderate and severe COPD compared to non-COPD controls and to identify splicing factors associated with aberrant alternative splicing in COPD. For this purpose, we performed genome-wide RNA-seq analysis of bronchial brushings from 23 mild-to-moderate, 121 severe COPD patients, and 23 non-COPD controls. We found a significant difference in the frequency of alternative splicing events in mild-to-moderate and severe COPD compared to non-COPD controls. There were from 2x to 8x (depending on event type) more differential alternative splicing events in the severe than in the mild-to-moderate stage. The samples from severe COPD patients showed less intron retention and more exon skipping. Interestingly, the transcript levels of the top 10 differentially expressed splicing factors were significantly correlated with the percentage of many alternatively spliced transcripts in severe COPD. The aberrant alternative splicing in severe COPD was predicted to increase the overall protein-coding capacity of gene products. In conclusion, we observed large and significant differences in alternative splicing between bronchial samples of COPD and control individuals, with more events observed in severe than in mild-to-moderate COPD. The changes in the expression of several splicing factors correlated with prevalence of alternative splicing in severe COPD. Alternative splicing can indirectly impact gene expression by changing the relative abundance of protein-coding isoforms potentially influencing pathophysiological changes. The presented results provide a better understanding of COPD-related alternative splicing changes.
Article
Full-text available
TDP-43 proteinopathies including frontotemporal lobar degeneration (FTLD) and amyotrophic lateral sclerosis (ALS) are neurodegenerative disorders characterized by aggregation and mislocalization of the nucleic acid-binding protein TDP-43 and subsequent neuronal dysfunction. Here, we developed endogenous models of sporadic TDP-43 proteinopathy based on the principle that disease-associated TDP-43 acetylation at lysine 145 (K145) alters TDP-43 conformation, impairs RNA-binding capacity, and induces downstream mis-regulation of target genes. Expression of acetylation-mimic TDP-43 K145Q resulted in stress-induced nuclear TDP-43 foci and loss of TDP-43 function in primary mouse and human-induced pluripotent stem cell (hiPSC)-derived cortical neurons. Mice harboring the TDP-43 K145Q mutation recapitulated key hallmarks of FTLD, including progressive TDP-43 phosphorylation and insolubility, TDP-43 mis-localization, transcriptomic and splicing alterations, and cognitive dysfunction. Our study supports a model in which TDP-43 acetylation drives neuronal dysfunction and cognitive decline through aberrant splicing and transcription of critical genes that regulate synaptic plasticity and stress response signaling. The neurodegenerative cascade initiated by TDP-43 acetylation recapitulates many aspects of human FTLD and provides a new paradigm to further interrogate TDP-43 proteinopathies.
Preprint
Full-text available
TDP-43 proteinopathies including frontotemporal lobar degeneration (FTLD) and amyotrophic lateral sclerosis (ALS) are neurodegenerative disorders characterized by aggregation and mislocalization of the nucleic-acid binding protein TDP-43 and subsequent neuronal dysfunction. Here, we developed endogenous models of sporadic TDP-43 proteinopathy based on the principle that disease-associated TDP-43 acetylation at lysine 145 (K145) alters TDP-43 conformation, impairs RNA-binding capacity, and induces downstream mis-regulation of target genes. Expression of acetylation-mimic TDP-43K145Q resulted in stress-induced nuclear TDP-43 foci and loss of TDP-43 function in primary mouse and human induced pluripotent stem cell (hiPSC)-derived cortical neurons. Mice harboring the TDP-43K145Q mutation recapitulated key hallmarks of FTLD, including progressive TDP-43 phosphorylation and insolubility, TDP-43 mis-localization, transcriptomic and splicing alterations, and cognitive dysfunction. Our study supports a model in which TDP-43 acetylation drives neuronal dysfunction and cognitive decline through aberrant splicing and transcription of critical genes that regulate synaptic plasticity and stress response signaling. The neurodegenerative cascade initiated by TDP-43 acetylation recapitulates many aspects of human FTLD and provides a new paradigm to further interrogate TDP-43 proteinopathies.
Article
Receptor-interacting protein 2 (RIP2) plays a critical role in the transduction of many signaling pathways and is associated with many diseases. Alternative splicing (AS) is an essential and ubiquitous regulatory mechanism of gene expression that contributes to distinct transcript variants and many different kinds of proteins. In this present study, we characterized genome-wide AS events in wild-type chicken macrophages (WT) and RIP2 overexpression/knockdown chicken macrophages (oeRIP2/shRIP2) by high-throughput RNA sequencing technology. A total of 1901, 2061, and 817 differentially expressed (DE) AS genes were identified in the comparison of oeRIP2 vs. WT, oeRIP2 vs. shRIP2, and shRIP2 vs. WT, respectively. These DE AS genes participated in many important KEGG pathways, including regulation of autophagy, Wnt signaling pathway, Ubiquitin mediated proteolysis, MAPK signaling pathway, and Focal adhesion, etc. In conclusion, this research provided a broad atlas of the genome-wide scale of the AS event landscape in RIP2 overexpression/knockdown and wild-type chicken macrophages. This research also provides the theoretical basis of the gene network related to RIP2.
Preprint
Full-text available
Eukaryotic Elongation Factor 1A1 (EEF1A1) is an RNA-binding protein that is associated with PARK2 activity in cells, predicting a possible important role in Parkinson`s disease (PD). However, it is not clear whether EEF1A1 will play a role in PD through transcriptional or post-transcriptional regulation. In this study, the GSE68719 data set was downloaded from the GEO database, and the RNA-seq data of BA9 brain tissue autopsy were obtained from 29 PD patients and 44 neurologically normal control subjects. The analysis found that EEF1A1 was significantly over-expressed in PD brain tissue. Then, to inhibit EEF1A1 from being expressed in U251 cells, siRNA was transfected into those cells, and the RNA-seq high-throughput sequencing was used to determine the Differentially Expressed Genes (DEGs) and differentially alternative splicing events (ASEs) resulting from the EEF1A1 knockdown. Further, GO and KEGG enrichment analysis revealed that EEF1A1 knockdown significantly up-regulated the expression of genes CXCL10, NGF, PTX3, IL6, ST6GALNAC3, NUPR1, TNFRSF21, CXCL2, and up-regulated alternative splicing of genes ACOT7, DDX10, SHMT2, MYEF2, NDUFAF5. These genes were enriched to pathways related to PD pathogenesis, such as apoptosis, inflammatory response, and mitochondrial dysfunction, suggesting that EEF1A1 may be involved in the development of PD by regulating differential expression and alternative splicing of genes. This study preliminarily investigated the potential role and mechanism of EEF1A1 in the pathogenesis of PD, providing a theoretical basis for subsequent research.
Article
Full-text available
Pre-mRNA splicing of Pol II transcripts is executed in the mammalian cell nucleus within a huge (21 MDa) and highly dynamic RNP machine – the supraspliceosome. It is composed of four splicing active native spliceosomes, each resembling an in vitro assembled spliceosome, which are connected by the pre-mRNA. Supraspliceosomes harbor protein splicing factors and all the five-spliceosomal U snRNPs. Recent analysis of specific supraspliceosomes at defined splicing stages revealed that they harbor all five spliceosomal U snRNAs at all splicing stages. Supraspliceosomes harbor additional pre-mRNA processing components, such as the 5’-end and 3'-end processing components, and the RNA editing enzymes ADAR1 and ADAR2. The structure of the native spliceosome, at a resolution of 20 Å, was determined by cryo-EM. A unique spatial arrangement of the spliceosomal U snRNPs within the native spliceosome emerged from in-silico studies, localizing the five U snRNPs mostly within its large subunit, and sheltering the active core components deep within the spliceosomal cavity. The supraspliceosome provides a platform for coordinating the numerous processing steps that the pre-mRNA undergoes: 5’ and 3’-end processing activities, RNA editing, constitutive and alternative splicing, and processing of intronic microRNAs. It also harbor a quality control mechanism termed Suppression of Splicing (SOS) that, under normal growth conditions, suppresses splicing at abundant intronic latent 5’ splice sites in a reading frame-dependent fashion. Notably, changes in these regulatory processing activities are associated with human disease and cancer. These findings emphasize the supraspliceosome as a multi-task master regulator of pre-mRNA processing in the cell nucleus.
Article
Full-text available
Parkinson protein 2, E3 ubiquitin protein ligase ( PARK2 ) gene mutations are the most frequent causes of autosomal recessive early onset Parkinson’s disease and juvenile Parkinson disease. Parkin deficiency has also been linked to other human pathologies, for example, sporadic Parkinson disease, Alzheimer disease, autism, and cancer. PARK2 primary transcript undergoes an extensive alternative splicing, which enhances transcriptomic diversification. To date several PARK2 splice variants have been identified; however, the expression and distribution of parkin isoforms have not been deeply investigated yet. Here, the currently known PARK2 gene transcripts and relative predicted encoded proteins in human, rat, and mouse are reviewed. By analyzing the literature, we highlight the existing data showing the presence of multiple parkin isoforms in the brain. Their expression emerges from conflicting results regarding the electrophoretic mobility of the protein, but it is also assumed from discrepant observations on the cellular and tissue distribution of parkin. Although the characterization of each predicted isoforms is complex, since they often diverge only for few amino acids, analysis of their expression patterns in the brain might account for the different pathogenetic effects linked to PARK2 gene mutations.
Article
Full-text available
We conducted a meta-analysis of Parkinson's disease genome-wide association studies using a common set of 7,893,274 variants across 13,708 cases and 95,282 controls. Twenty-six loci were identified as having genome-wide significant association; these and 6 additional previously reported loci were then tested in an independent set of 5,353 cases and 5,551 controls. Of the 32 tested SNPs, 24 replicated, including 6 newly identified loci. Conditional analyses within loci showed that four loci, including GBA, GAK-DGKQ, SNCA and the HLA region, contain a secondary independent risk variant. In total, we identified and replicated 28 independent risk variants for Parkinson's disease across 24 loci. Although the effect of each individual locus was small, risk profile analysis showed substantial cumulative risk in a comparison of the highest and lowest quintiles of genetic risk (odds ratio (OR) = 3.31, 95% confidence interval (CI) = 2.55-4.30; P = 2 × 10(-16)). We also show six risk loci associated with proximal gene expression or DNA methylation.
Article
Full-text available
The completion of the Human Genome Project aroused renewed interest in alternative splicing, an efficient and widespread mechanism that generates multiple protein isoforms from individual genes. Although our knowledge about alternative splicing is growing exponentially, its real impact on cellular life is still to be clarified. Connecting all splicing features (genes, splice transcripts, isoforms, and relative functions) may be useful to resolve this tangle. Herein, we will start from the case of a single gene, Parkinson protein 2, E3 ubiquitin protein ligase (PARK2), one of the largest in our genome. This gene is implicated in the pathogenesis of autosomal recessive juvenile Parkinsonism and it has been recently linked to cancer, leprosy, autism, type 2 diabetes mellitus and Alzheimer's disease. PARK2 primary transcript undergoes an extensive alternative splicing, which enhances transcriptomic diversification and protein diversity in tissues and cells. This review will provide an update of all human PARK2 alternative splice transcripts and isoforms presently known, and correlate them to those in rat and mouse, two common animal models for studying human disease genes. Alternative splicing relies upon a complex process that could be easily altered by both cis and trans-acting mutations. Although the contribution of PARK2 splicing in human disease remains to be fully explored, some evidences show disruption of this versatile form of genetic regulation may have pathological consequences.
Article
Full-text available
The continuously prolonged human lifespan is accompanied by increase in neurodegenerative diseases incidence, calling for the development of inexpensive blood-based diagnostics. Analyzing blood cell transcripts by RNA-Seq is a robust means to identify novel biomarkers that rapidly becomes a commonplace. However, there is lack of tools to discover novel exons, junctions and splicing events and to precisely and sensitively assess differential splicing through RNA-Seq data analysis and across RNA-Seq platforms. Here, we present a new and comprehensive computational workflow for whole-transcriptome RNA-Seq analysis, using an updated version of the software AltAnalyze, to identify both known and novel high-confidence alternative splicing events, and to integrate them with both protein-domains and microRNA binding annotations. We applied the novel workflow on RNA-Seq data from Parkinson's disease (PD) patients' leukocytes pre- and post- Deep Brain Stimulation (DBS) treatment and compared to healthy controls. Disease-mediated changes included decreased usage of alternative promoters and N-termini, 5'-end variations and mutually-exclusive exons. The PD regulated FUS and HNRNP A/B included prion-like domains regulated regions. We also present here a workflow to identify and analyze long non-coding RNAs (lncRNAs) via RNA-Seq data. We identified reduced lncRNA expression and selective PD-induced changes in 13 of over 6,000 detected leukocyte lncRNAs, four of which were inversely altered post-DBS. These included the U1 spliceosomal lncRNA and RP11-462G22.1, each entailing sequence complementarity to numerous microRNAs. Analysis of RNA-Seq from PD and unaffected controls brains revealed over 7,000 brain-expressed lncRNAs, of which 3,495 were co-expressed in the leukocytes including U1, which showed both leukocyte and brain increases. Furthermore, qRT-PCR validations confirmed these co-increases in PD leukocytes and two brain regions, the amygdala and substantia-nigra, compared to controls. This novel workflow allows deep multi-level inspection of RNA-Seq datasets and provides a comprehensive new resource for understanding disease transcriptome modifications in PD and other neurodegenerative diseases.
Article
The discovery of genes implicated in familial forms of Parkinson’s disease (PD) has provided new insights into the molecular events leading to neurodegeneration. Clinically, patients with genetically determined PD can be difficult to distinguish from those with sporadic PD. Monogenic causes include autosomal dominantly (SNCA, LRRK2, VPS35, EIF4G1) as well as recessively (PARK2, PINK1, DJ-1) inherited mutations. Additional recessive forms of parkinsonism present with atypical signs, including very early disease onset, dystonia, dementia and pyramidal signs. New techniques in the search for phenotype-associated genes (next-generation sequencing, genome-wide association studies) have expanded the spectrum of both monogenic PD and variants that alter risk to develop PD. Examples of risk genes include the two lysosomal enzyme coding genes GBA and SMPD1, which are associated with a 5-fold and 9-fold increased risk of PD, respectively. It is hoped that further knowledge of the genetic makeup of PD will allow designing treatments that alter the course of the disease.
Article
Pathologic aggregation of α-synuclein is a central process in the pathogenesis of Parkinson's disease. The α-synuclein gene (SNCA) encodes at least 4 different α-synuclein isoforms through alternative splicing (SNCA140, SNCA126, SNCA112, SNCA98). Differential expression of α-synuclein isoforms has been shown in Lewy body diseases. In contrast to the canonical α-synuclein isoform of 140 amino acid residues (SNCA140), which has been investigated in detail, little is known about the properties of the 3 alternative isoforms. We have investigated the aggregation properties of all 4 isoforms in cultured cells and analyzed fibril-formation of 3 isoforms (SNCA140, SNCA126, and SNCA98) in vitro by electron microscopy. Each of the 3 alternative isoforms aggregates significantly less than the canonical isoform SNCA140. Electron microscopy showed that SNCA140 formed the well-known relatively straight fibrils while SNCA126 formed shorter fibrils, which were arranged in parallel fibril bundles and SNCA98 formed annular structures. Expression analysis of α-synuclein isoforms in different human brain regions demonstrated low expression levels of the alternative isoforms in comparison to the canonical SNCA140 isoform. These findings demonstrate that α-synuclein isoforms differ qualitatively and quantitatively in their aggregation properties. The biological consequences of these findings remain to be explored in vitro and in vivo.