You are on page 1of 13

Biomedicine & Pharmacotherapy 111 (2019) 765–777

Contents lists available at ScienceDirect

Biomedicine & Pharmacotherapy


journal homepage: www.elsevier.com/locate/biopha

Review

Emerging pathways to neurodegeneration: Dissecting the critical molecular T


mechanisms in Alzheimer’s disease, Parkinson’s disease
Sean Hong Tana, Venkatanaidu Karrib, Nicole Wuen Rong Taya, Kuan Hui Changa, Hui Yen Aha,

Phui Qi Nga, Hui San Hoa, Hsiao Wai Keha, Mayuren Candasamyc,
a
School of Pharmacy, International Medical University, No 126, Jalan Jalil Perkasa 19, Bukit Jalil 57000, Kuala Lumpur, Malaysia
b
Department of Toxicogenomics, Faculty of Health, Medicines, Life Sciences, Maastricht University, Maastricht, Netherlands
c
Department of Life Sciences, School of Pharmacy, International Medical University, No 126, Jalan Jalil Perkasa 19, Bukit Jalil 57000, Kuala Lumpur, Malaysia

A R T I C LE I N FO A B S T R A C T

Keywords: Neurodegenerative diseases are usually sporadic in nature and commonly influenced by a wide range of genetic,
Alzheimer’s disease life style and environmental factors. A unifying feature of Alzheimer’s disease (AD) and Parkinson’s disease (PD)
Parkinson’s disease is the abnormal accumulation and processing of mutant or damaged intra and extracellular proteins; this leads to
Ubiquitin proteasome neuronal vulnerability and dysfunction in the brain. Through a detailed review of ubiquitin proteasome, mRNA
mRNA splicing
splicing, mitochondrial dysfunction, and oxidative stress pathway interrelation on neurodegeneration can im-
Mitochondrial dysfunction
Oxidative stress
prove the understanding of the disease mechanism. The identified pathways common to AD and PD nominate
Biomarker promising new targets for further studies, and as well as biomarkers. These insights suggested would likely
provide major stimuli for developing unified treatment approaches to combat neurodegeneration. More broadly,
pathways can serve as vehicles for integrating findings from diverse studies of neurodegeneration. The evidence
examined in this review provides a brief overview of the current literature on significant pathways in promoting
in AD, PD. Additionally, these insights suggest that biomarkers and treatment strategies may require simulta-
neous targeting of multiple components.

1. Introduction disease (PD), Progressive supranuclear palsy (PSP) and dementia with
Lewy bodies (DLB) [3]. Among these diseases, AD and PD are the most
Neurodegeneration is described as the loss of specific type of neu- prevalent neurodegenerative diseases where AD is known as an esca-
rons, pattern and distribution of the neurons, leading to central nervous lating dementia, while PD is primarily characterized as a movement
system (CNS) dysfunction progressively and often untreatable [1,2]. disorder.
Neurodegenerative diseases include a variety of pathological patterns The prevalence of AD has elevated to an estimated 40 million pa-
and clinical presentations, such as Alzheimer’s disease (AD), Parkinson tients globally [4]. AD running in the family, known as familial AD,

Abbreviations: %OH, hydroxyl radicals hydroxyl radicals; 4-HNE, 4-hydroxynonenal; AD, Alzheimer’s disease; APOE- ε4, εε4 allele of apolipoprotein; APP, amyloid
precursor protein; ARJP, autosomal recessive juvenile parkinsonism; AS, alternative splicing; ATP, adenosine triphosphate; ATP13A2, adenosine triphosphatase
13A2; Aβ, beta-amyloid; BACE, beta-secretase; Ca2+, calcium; CNS, central nervous system; DJ-1, protein deglycase 1; Drp1, dynamin-related protein 1; E1, ubiquitin
activating; E2, ubiquitin conjugating; E3, ubiquitin ligating; ER-α, estrogen receptor-α; ETC, electron transport chain; FBP1, far upstream sequence element-binding
protein 1; FBXO7, F-box protein 7; Fis1, mitochondrial fission protein 1; GABA B, γ-aminobutyric acid B; GBA1, glucocerebrosidase 1; GFAP, glial fibrillary acidic
protein; GPCR51, G-protein-coupled receptor 51; GSH, glutathione; H2O2, hydrogen peroxide; HT, Huntington’s disease; KGDHC, alpha-ketoglutarate dehydrogenase
complex; LB, Lewy body; LRRK2, leucine-rich repeat kinase 2; MAO-A, monoamine oxidase-A; MAO-B, monoamine oxidase-B; MAPT, microtubule associated protein
tau; MDA, malondialdehyde; Mfn1, Mitofusins 1; Mfn2, Mitofusins 2; mRNA, messenger ribonucleic acid; O2−, superoxide; OMM, outer membrane of mitochondria;
Opa1, optic atrophy protein 1; Pael-R, Parkin-associated endothelin-like receptor; PARK2, Parkinson disease protein 2; PARK6, Parkinson disease protein 6; PARK7,
Parkinson disease protein 7; PD, Parkinson’s disease; PINK1, PTEN-induced putative kinase 1; PKC, protein kinase C; PLA2G6, phospholipase A2G6; PRKN, Parkin;
PS1, presenilin 1; PS2, presenilin 2; PSP, progressive supranuclear palsy; PTP, permeability transition pore; RAGE, receptor for advanced glycation products; ROS,
reactive oxygen species; SNCA, alpha-synuclein; SNCAIP, synuclein alpha interacting protein; SNPC, substantia nigra pars compacta; SOD, superoxide dismutase;
SRRM2, serine/arginine repetitive matrix 2; TFAM, transcription factor A, mitochondrial; TRX, thioredoxin; TV12, transcript variant 12; TV3, transcript variant 3;
UCH-L1, ubiquitin carboxy-terminal hydrolase L1; UPS, ubiquitin proteasome system; VDACs, voltage-dependent anions channels

Corresponding author.
E-mail address: mayurencandasamy@imu.edu.my (M. Candasamy).

https://doi.org/10.1016/j.biopha.2018.12.101
Received 13 October 2018; Received in revised form 18 December 2018; Accepted 23 December 2018
0753-3322/ © 2018 The Authors. Published by Elsevier Masson SAS. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/BY-NC-ND/4.0/).
S.H. Tan et al. Biomedicine & Pharmacotherapy 111 (2019) 765–777

represents less than 5% of AD cases; whereas the mostly seen AD, first cognitive complaint by DLB, PDD and AD patients [24]. On the
known as sporadic AD occupies for more than 90% of AD cases [5]. other hand, dementia, cognitive fluctuations and visual hallucinations
Although genetic mutation can be one of the causative factors of AD, are common clinical features found in both DLB and PD. Others im-
most studies have proposed that the interaction between lifestyle and portant clinical presentation for both DLB and PD include executive
environmental factors can also lead to AD [1]. The irreversible brain dysfunctions and visual-spatial abnormalities [25]. DLB and PD dif-
damage in AD progresses slowly and eventually impairs the memory, ferent as bradykinesia and rigidity are more common in DLB [26]. DLB
perception and behaviour [6]. For instance, patients with advanced AD presented with greater deficiencies of attention while in PD it is less
shows verbal episodic memory and behavioural troubles [7]. Also, the common [27]. However, executive functions and visual symptoms are
presence of intracellular neurofibrillary tangles and extracellular claimed to be more impaired in PD compared to DLB probably due to
neuritic plaques are the key pathological features of AD [8]. Since reduce metabolism in visual pathways in PD [28,29]. In DLB, halluci-
molecular studies of AD began in the early 1980s, many scientists and nations occur spontaneously, which is reported to be related to Lewy
healthcare professionals have delved into all aspects of this complex, Bodies in temporal lobe of brains [30], while hallucinations in PD often
multifactorial syndrome [4]. The most prevalent hypothesis is the beta- occur after dopaminergic therapy [25,31]. There is report claimed that
amyloid (Aβ) cascade theory. This theory hypothesised that increased rate of cognitive decline is faster in DLB compared to PD and AD
production of insoluble extracellular Aβ leads to Aβ aggregation and [32,33]. Clinical data reported psychiatric symptoms especially delu-
deposition manifesting as plaques in the brain. The plaques are believed sions are more common in DLB compared to PD and AD [24].
leads to dementia and eventually death [9]. The theory holds that an However, this review mainly focuses on two of the most common
amyloid-related mechanism that prunes neuronal connections in the neurodegenerative diseases, AD and PD as they share some common
brain in the fast-growth phase of early life may be triggered by ageing- clinical pathological and epidemiological features. Thus, the objective
related processes in later life to cause the neuronal withering of AD. The is to improve the understanding of the role of critical biological path-
ongoing aggregation of Aβ which occurred due to alternations of sig- ways such as UPS, aberrant alternative splicing, mitochondrial dys-
nalling-protein such as glycogen synthase kinase-3β, fyn kinase and function and oxidative stress impact in AD, PD.
cyclin-dependent kinase (CDK5) could lead to Aβ oligomers formation,
Aβ1-42 particularly neurotoxic when aggregated in the brain [10]. The
2. Factors that contribute to the development of the
second one is tau hypothesis, in this theory, hyperphosphorylated tau
neurodegenerative diseases
begins to pair with other threads of tau. Eventually, they form neuro-
fibrillary tangles inside nerve cell bodies destroys the neurons [11]. On
The development of neurodegenerative diseases may result from a
the other hand, the mutation in APP, tau, presenillin-1, presenillin-2
number of factors: aging, genetics and environmental factors (Fig. 1).
and ApoE4 are involved in aggregation of Aβ as well [12]. The for-
mation of Aβ oligomers alter the signalling pathways and glutamate
receptors, as a result, neurogenesis defect and synaptic loss happened in 2.1. Aging
neurons which produce glutamate or acetylcholine as neurotransmitters
[13]. Ageing is a major risk factor of neurodegenerative diseases. Ageing
PD is the second most common neurodegenerative disorder after not only makes patients more prone to neurodegenerative diseases, but
Alzheimer’s disease. Parkinson disease typically develops between the also impairs their abilities of self-repair. The rising incidence of age-
ages of 55 and 65 years and occurs in 2% of people over the age of 60 related neurodegenerative diseases has become an important health
years, rising to 3.5% at age 85–89 years [14]. It is a complex disorder, concern in the world. Aging is associated with functional impairments
described by the loss of dopamine producing neurons in substantia such as dementia and motor neuron disability in neurodegenerative
nigra pars compacta (SNPC) as well as the presence of Lewy bodies in diseases such as AD and PD [34]. In the cellular level, ageing is asso-
the SNPC and locus coeruleus leading to poor motor control [1,15]. The ciated with accumulating oxidative stress, declining mitochondrial
evidences on dementia with Lewy bodies (DLB) and PD as two syn- function, impaired DNA repair and decreased tissue regeneration take
dromes within the same disease spectrum. Both are pathologically part in the development of AD and PD [35]. However, the increase of
characterized by neuronal loss and Lewy bodies, consisting of inclu- free radicals in the brain is not caused by elevation of oxidative stress.
sions of α-synuclein [16]. Studies shown that the activities of superoxide dismutase (SOD), cata-
PD is caused by deterioration of the dopaminergic neurons in the lase, glutathione peroxidase and glutathione reductase were reduced in
extrapyramidal tract of the midbrain. There is also accumulation of α- affected brain regions in AD, PD [36]. As mitochondrial efficiency de-
synuclein proteins, known as Lewy bodies, in the central, autonomic, clines with age, a number of observations have led researchers to
and peripheral nervous system as the hallmarks of the PD [17]. The speculate that mitochondrial dysfunction may contribute to the for-
intracellular aggregated α-synuclein makes a hole in the neuronal mation of misfolded protein aggregates and subsequently lead to AD
membrane killing neuronal through oxidative stress, excitotoxicity, and PD [37,38].
energy failure and neuroinflammation [18]. Thus, PD patients are
presented with clinical manifestations such as resting tremor, muscle
rigidity, postural instability, bradykinesia and a mask-like facial ex-
pression [19]. They cannot perform well in sequence learning, motor
sequence and switching between different tasks as well as verbal flu-
ency [20]. Alike with AD, 90 to 95% of the PD cases are sporadic PD,
whereas 5 to 10% of cases are familial PD [21]. The neuropathology of
PD is poorly understood until it has been considered as the prototypical
nongenetic disorder [13]. Gene mutations in alpha-synuclein (SNCA),
parkin (PRKN), ubiquitin carboxy-terminal hydrolase L1 (UCH-L1),
PTEN-induced kinase 1 (PINK1) and protein deglycase 1 (DJ-1) genes
are associated with PD as they result from the impairment of the bio-
logical pathways such as the ubiquitin proteasome system (UPS) and
mitochondrial dysfunction [22,23].
Several studies suggest that DLB, PD and AD overlap in term of
memory impairment. One clinical data reported memory changes as Fig. 1. Factors contributing to the development of AD and PD.

766
S.H. Tan et al. Biomedicine & Pharmacotherapy 111 (2019) 765–777

Table 1
Proteins that have a function in major neurodegenerative diseases.
Disease Genetic Causes Inheritance Pathogenesis References

AD APP Dominant Give rise to Aβ formation, the primary senile plaque. [49]
PS1 and PS2 Dominant Increased Aβ42/Aβ40 ratio. [50]
PD α-synuclein Dominant Genetic polymorphisms in the gene increases tendency to form Lewy bodies. [51]
LRRK2 Dominant A kinase. Function unknown. [52,53]
PARKIN Recessive Mutation in the gene causes impaired protein degradation. [45]
PINK1 Recessive A kinase localised to mitochondria. Function unknown. Seems to protect against cell death. [47]
DJ-1 Recessive Protects the cell against oxidant- induced cell death. [46]

2.2. Genetics [8,56]. Exposure to toxic heavy metals such as Pb, MeHg, Cd, and As
has been documented to cause neurotoxicity which eventually leads to
Family studies have established that genetic factors play a crucial neurodegenerative diseases [57].
role in AD, especially in younger onset cases (< 65 years) [39]. Familial
forms of AD follow an autosomal dominant inheritance pattern which 3. The possible underlying molecular mechanisms in AD, PD
result from mutations in three genes (APP, presenilin-1, and presenilin- neurodegeneration
2) that are responsible for Aβ plaque formation (Table 1). Non-familial
AD/ sporadic AD typically occurs at the age of 65 years or older, ac- Relatively little is known regarding the mechanisms of AD and PD
counting for most of the cases of AD [40]. It has been associated most pathogenesis. It has been suggested that impairments of UPS, dysre-
consistently with ε4 allele of apolipoprotein (APOE- ε4) gene which is a gulation in alternative splicing, mitochondrial dysfunction [69], and
very low-density lipoprotein carrier required for APOE- ε4 deposition decreased protection against oxidative stress and apoptosis have fun-
[41]. Carriers of the APOE- ε4 allele have reduced AD ages at onset, damental roles in neurodegenerative diseases such as AD and PD. These
with 3-fold and 15-fold risk excesses observed in heterozygotes and defects are normally resulting from genetic, environmental risk factors
homozygotes respectively [42]. The vast majority of AD is sporadic, [70].
with no obvious genetic component, suggesting that other mechanisms
are responsible. 4. The role of the ubiquitin-proteasome system (UPS) in
Specific causal mutations appear to be more prominent in early- neurodegenerative diseases
onset PD which affect at least five genetic loci (Table 1). The initial
discoveries were mutations of the gene encoding SNCA protein which UPS is a mechanism for protein degradation. It plays an important
have been related to early-onset of PD [43]. Mutations in leucine-rich role in regulating wide variety of cellular function [71]. The UPS is
repeat kinase 2 (LRRK2) gene have been associated with late-onset PD essential for the non-lysosomal degradation and clearance of short-
which may also contribute to non-familial PD [44]. Identified muta- lived, misfolded, mutant, and damaged proteins in eukaryotic cells.
tions in other genes include Parkinson disease protein 2 (PARKIN or Structural and functional deficits in the 26S and 20S proteasomes may
PARK2), PTEN-induced putative kinase 1 (PINK1 or PARK6) and DJ-1 lead to AD, and PD [72].
(PARK7), all of which follow a recessive inheritance mode [45–47].
Since 90% of PD cases are sporadic and cannot be attributed to only 4.1. Alzheimer’s disease
genetic factors, this suggests that PD has a multifactorial aetiology [48].
Many evidences proposed that AD might due to defect in UPS
2.3. Environmental factors function (Fig. 2). Evidence suggested that there are increased amount of
misfolded and aggregated protein accumulates in specific structure in
It is known that the aetiology of PD, AD is multifactorial (Table 2), AD, such as plaques and tangles [73]. Also, certain amount of ubiquitin
and there is evidence that potential external factors including lifestyle mutant proteins are shown in plaques and tangles which could be the
and chemical exposures are linked with the risk of the onset of these mediator of Aβ-induced neurotoxicity [74]. Aggregated oxidized pro-
diseases. Environmental factors such heavy metals, pesticides have teins in brain with AD could further reduce proteasome activity, which
been involved in PD, AD development due to their ability to disrupt the is responsible to destroy damaged proteins [73,75].
neuronal functions [54,55]. Much ongoing research is focusing on Reported studies showed that the relationship between Aβ and
metals exposure as potential risk factors for neurodegenerative diseases. proteasome pathway, where the accumulation of Aβ could diminish
Although a direct casual role for some metals such as lead (Pb) alu- proteasome function [75]. The balance between Aβ production and
minium (Al), iron (Fe), copper (Cu), zinc (Zn), mercury (Mg), arsenic degradation play a crucial role in the accumulation of Aβ in neurons, it
(As), and cadmium (Cd) has not yet been definitively demonstrated, indicates UPS is needed for Aβ metabolite clearance [76–78]. For in-
epidemiological evidence suggests that increased levels of these metals stance, both Aβ40 and Aβ42 oligomers affect proteasome activity by
in the brain may link to the development and progression of AD and PD notably reducing the chymotrypsin-like activity, trypsin-like activity

Table 2
Some of the proposed environmental factors for the major neurodegenerative diseases.
Environmental Factor(s) Associated With References

AD
• Metals (Pb, Hg, As, Cd, Al, Fe, Cu, Zn) Increased risk, inconclusive results [54,55,58,59]
• Pesticides, Solvents, Electromagnetic fields Increased risk, inconclusive results [60]

PD
Traumatic brain injuries, Infections, Inflammations Increased risk [61]

• Metals (Pb, As, Hg, Cd, Al, Fe, Cu, Zn) Increased risk [62,63,64,65]
• Pesticides, Herbicides Increased risk [66,67]
• Head injuries with loss of consciousness Increased risk [66,68]

767
S.H. Tan et al. Biomedicine & Pharmacotherapy 111 (2019) 765–777

In an in-vitro investigation, mutated PARKIN is unable to function as


E3 enzyme and non-functional cellular protein is unable to be tagged
and degraded [86]. Hence, mutation in PARKIN may directly affects the
survival of dopaminergic neuron, leading to PD. Other than that, the
ability of mutated PARKIN to change the solubility of non-functional
protein is found and this encourages accumulation of PARKIN substrate
such as parkin-associated endothelin-like receptor (Pael-R), cell divi-
sion control-related protein (CDCrel 1 and 2a), far upstream sequence
element-binding protein 1 (FBP1), cyclin E and p38 (JTV-1/AIMP2)
[87,88]. This is due to large number of PARKIN substrates are presented
in the brain of PD patient with PARKIN mutation [89]. These accu-
mulated PARKIN substrates alter the normal function of neuronal cell
and result in the development of PD [88].
Evidence also showed that mutation or triplication of Lewy bodies
Fig. 2. The defective UPS increases the accumulation of ubiquitin conjugates,
which is SNCA gene could cause PD [87]. Impairment in clearance of
and eventually tau as well as β-amyloid peptide, both of which are the hall-
SNCA due to mutated SNCA gene results in excessive SNCA accumu-
marks of AD.
lation in the brain [90]. These aggregated SNCA further reduces the
proteasome activity, causing development of PD [87]. Based on these
and the peptidyl glutamyl-like activity of the proteasome [79]. Other evidences, there is a direct connection between abnormal SNCA ag-
than that, E2-25 K/Hip2, an ubiquitin-conjugation enzyme, is reported gregation and dopaminergic neuronal cell death.
to mediate Aβ neurotoxicity due to its enzymatic activity in the in- UCH-L1 which is crucial for the proteasome function, is found
hibition of proteasome [80]. Furthermore, impairment in tau turnover mutated in PD [91]. The ubiquitin hydrolase activity of mutated UCH-
results in AD since tau can be degraded by proteasome [81]. It is sug- L1 is noticeably reduced in an in-vitro study, proving that decrease in
gested that the proteasomal dysfunction in AD occurs when tau-paired polyubiquitin hydrolysis can cause dopaminergic neuronal death [92].
helical filament is prevented from binding with proteasome [82]. Other than this, aging and exogenous stress also encourage aberration
More evidences have shown that defects in UPS may lead to AD. For of UPS and subsequently contribute to PD pathogenesis [93].
instance, normal aging process reduces proteasome activity that may
leads to accumulation of Aβ and tau [83]. Also, accumulation of Aβ and 5. Aberrant alternative splicing events in neurodegenerative
tau will reduce the proteasome activity which then causes even more diseases
accumulation of Aβ and tau. This is a vicious cycle where aggregated
Aβ and tau keeps increasing thus, causing AD to be more severe. Aberrant alternative splicing plays a significant role in neurode-
generative diseases, particularly in AD and PD [94,95]. Primary tran-
4.2. Parkinson’s disease script of pre-mRNA undergoes removal of introns and simultaneous
formations of varying combinations of exon regions, also known as
Growing evidence shows that defective UPS causes PD (Fig. 3). alternative splicing (AS) [95]. AS is one of the most important me-
Under normal condition, cellular proteins undergo a series of reaction chanisms to synthesis a large number of mRNA and protein isoforms
in UPS before degradation [79]. With impairment of UPS, the non- from the low number of human genes [96]. It is a widespread gene
functional intracellular proteins accumulate in the brain, resulting in regulatory process by which exons of primary transcripts (pre-mRNAs)
damaged dopaminergic neurons and causing PD [84]. The enzymes are spliced into different arrangements to produce structurally distinct
responsible for tagging prior non-functional intracellular protein de- mRNA variants [97]. This mechanism of gene product plays a critical
graded by proteasome include ubiquitin activating (E1), ubiquitin role in cell function [98]. It is estimated that more than 75% of genes in
conjugating (E2) and ubiquitin ligating (E3) enzymes [79]. Studies have the human genome are alternatively spliced in CNS [99]. Dysregulation
showed that genetic mutation in PARKIN (E3), SNCA and UCH-L1, a of AS has been linked to the number of human diseases including
member of deubiquitinating enzyme could disrupt UPS and lead to neurodegenerative diseases [100].
aggregation of protein substrate which is harmful to the survival of
neuron [85]. 5.1. Alzheimer’s disease

AD is associated with the dysregulation of several proteins that


stems from aberrant alternative splicing. One such protein is tau, which
is well-known as one of the major contributors to the pathophysiology
of AD. In the human brain, six isoforms of protein tau has been iden-
tified and differs by the inclusion or exclusion of exon 2, 3 and 10
[101]. Disruption of the splicing process involving exon 10 is sufficient
to cause disease of the tau protein [102]. The β-secretase (BACE) and γ-
secretase drives two subsequent enzymatic digestions that forms Aβ.
The exon 3 and 4 of BACE1 have complex splicing profiles and the rich
sequence of G in exon 3 increases the generation of full-length BACE
[103,104]. Familial AD is mostly associated with mutations in genes
encoding presenilin 1 (PS1) and presenilin 2 (PS2) which are integral
membrane proteins. Deletion of exon 9 in PS1 disrupts the structure and
leaves out an important residue, D257 [105]. Exon 4 of PS1 has two 5′
splice sites resulting in two splice variants that differ by four amino
acids, VRSQ, which forms a potential site for protein kinase C (PKC)
Fig. 3. The schematic representation of mutated Parkin, UCH-L1 and SNCA phosphorylation. Rodents only have the longer PS1 isoform, resulting in
genes contribute to defective UPS and the assemble of these non-functional a more readily regulated PS1 protein, thus generating less Aβ [106]. In
intracellular proteins triggers the development of PD. PS2, deletion of exon 5 expresses truncated PS2 isoform, PS2V which

768
S.H. Tan et al. Biomedicine & Pharmacotherapy 111 (2019) 765–777

accumulates rapidly in the hippocampus of patients with sporadic AD. resulting in the inhibition of proteasomal function and formation of
This is due to the overexpression of high-mobility group A protein 1a aggregates in neurons [117].
(HMGA1a), a sequence-specific RNA-binding factor [107]. Another Autosomal recessive juvenile parkinsonism (ARJP) is associated
protein of interest in AD is GFAP (glial fibrillary acidic protein), its with mutations in the PARK2 encoding for PARKIN. PARKIN plays a
variant, GFAP + 1 was found in AD brains due to the expression of two role in the ubiquitin proteasome system and is expressed in many re-
novel splice forms: lacking exon 6 or deletion of 164 nt [108]. Neurons gions of the brain [95]. The change in the expression of the seven
in the hippocampus that failed to degrade aberrant protein result in isoforms derived from PARK2 results in Lewy body (LB) disease [127].
their accumulation, which may lead to neuronal death [109]. In PD, transcript variant 3 (TV3) which lacks of exons 3–5, and tran-
The splicing process influence the gender specific AD progress script variant 12 (TV12) which lacks of exons 2–7, were found to be
[110]. Hypothetically, this may be due to the alterations in the ratio of increased [128]. The pathogenic mechanism of ARJP may be attributed
canonically and alternatively spliced isoforms of the estrogen receptor- to the structural variations of the PARK2 alternatively spliced isoforms.
α (ER-α) [111]. The most common ER-α isoform found in the brain is There was a study reporting that E4SV, an alternatively spliced variant
without exon 7. Complete or partial exclusion of this exon is also as- of parkin that lacks exon 4 was increasingly expressed in sporadic PD
sociated with double, triple, or multiple exon skipping variants [112]. and it has completely lost its enzymatic activity [129].
In AD brain, more wild type ER-α is found rather than the alternatively SNCAIP gene which encodes synphilin-1 generates at least eight
spliced isoforms. Another dominant negative variant, isoforms ex- spliced variants. Two of the variants produces isoforms synphilin-1 A
cluding exon 2, was also found to be reduced in AD brain, particularly and synphilin-1B which are upregulated in PD [82]. Synphilin-1 A lacks
in women. Overexpression of dominant splice variant del. 7 may de- of exon 3 and 4 and includes exon 9 A. Its overexpression may con-
crease the effect of estrogen on cognitive function [113]. Another gene tribute to neurodegeneration via promotion of aggregation and sa-
that may be implicated in AD is receptor for advanced glycation pro- turation of the ubiquitin proteasome system [95,128,130]. Autosomal
ducts (RAGE). Its alternatively spliced form, sRAGE is expressed with a recessive early-onset parkinsonism is also tied to the aberrant alter-
lower level in AD and may disrupt the normal regulation of RAGE native splicing in PTEN-induced putative kinase 1 (PINK1) gene. The
signalling which could lead to abnormal structure of extracellular do- aberrant mRNAs are produced by a 23-bp deletion which disrupts the
mains [114]. 17 A, a non-coding RNA found in intron 3 of the G-pro- acceptor site of exon 7 [131]. In addition, complete deletion of exon 7
tein-coupled receptor 51 gene (GPCR51) regulates the alternative and a novel U1-dependent splice-site mutation in exon 7 was found in a
splicing of the gene, as well as its pre-mRNA maturation. High levels of Spanish family with PD members [132].
17 A was found in AD brain may upregulate the ratio of Aβ production, Other genes minorly associated with PD is protein deglycase DJ-1
and a change in the spliced variant ratio of GPCR which could impair (PARK7), adenosine triphosphatase 13A2 (ATP13A2), phospholipase
normal γ-aminobutyric acid B (GABA B) signalling [115]. A2G6 (PLA2G6) and F-box protein 7 (FBXO7) [133]. Mutations in
PARK7 is linked to autosomal recessive parkinsonism [134,135],
5.2. Parkinson’s disease whereby the alteration in the abundance of PARK7 splice variants are
found in PD brains and could act as potential biomarkers of the disease.
As of now, the aberrant alternative splicing of six genes have been The skipping of exon 13 in ATP13A2 cause the deletion of the third
associated with PD: leucine-rich repeat kinase 2 (LRRK2), parkin 2 transmembrane domain [136] which may contribute to levodopa-re-
(PARK2), synuclein alpha interacting protein (SNCAIP), SNCA, micro- sponsive inherited atypical parkinsonism [137]. Abnormal mRNA
tubule associated protein tau (MAPT), and serine/arginine repetitive splicing of PLA2G6 is caused by a nucleotide mutation in exon 7 of the
matrix 2 (SRRM2) [95]. The first gene identified in the pathophysiology gene. This causes a 4-bp deletion in its transcript, a change in the frame
of PD is SNCA which encodes for SNCA, a protein found in the pre- shift in leukocytes [138]. Parkinsonism pyramidal disease is caused by
synaptic region [116]. SNCA has many functions and studies have mutations in FBX07 in which the invariable splice donor of intron 7 is
shown that the lack of its acidic C-terminus, specifically the exon 5 of removed [133,139,140].
the 3′ region (SNCA112), impairs normal physiological function and
enhances aggregation of protein [117–120]. Dopaminergic neuron cell 6. Impact of mitochondria dysfunction in neurodegenerative
death is promoted by the overexpression of SNCA112 [95]. Another diseases
strong candidate gene for PD is SRRM2 encoding SR repetitive matrix 2
which has two main splice variants that differ at their 3′ region: the full There is overwhelming evidence of impaired mitochondrial function
length SRRM2 and the shorter SRRM2 transcript that lacks exons 12-15. as a causative factor in AD, PD. More recently, evidence has emerged
In PD, SRRM2 switches from the basal transcript level to low levels of for impaired mitochondrial dynamics in AD, PD neurodegenerative
the longer isoform and higher levels of the shorter isoform. It was found diseases. Here, we provide a concise overview of the major findings in
out that the downregulation of the longer isoform is due to the corre- recent years highlighting the importance of mitochondria in AD, PD.
sponding dysregulation of downstream exons and upstream exons
[121]. 6.1. Alzheimer's disease
Tau protein is encoded by MAPT which plays a significant role in
the morphogenesis of axons, polarity of axons, and axonal transport Mitochondria have crucial role in brain function. They provide en-
[122]. Tau mutations that enhance the inclusion of exon 10 by exerting ergy in the form of adenosine triphosphate (ATP) for brain cells and
its effect on three cis-acting regulatory elements, induce overproduction allow communication between brain cells [141]. It has been proposed
of 4R tau which may lead to tau pathology [95,123,124]. that in sporadic AD mitochondrial dysfunction is the primary event that
Mutations in the LRRK2 encoding for dardarin are the most common causes Aβ deposition, synaptic degeneration, and formation of neuro-
cause of apparent sporadic PD and familial late-onset parkinsonism fibrillary tangles [142]. In AD patients, mitochondrial dysfunction is
[125]. LRRK2 acts indirectly by phosphorylating heterogeneous nuclear possible by different aspects including mitochondrial morphology or
ribonucleoproteins (hnRNP) which interact with basal splicing com- mitochondrial dynamic, mitochondrial bioenergetics as well as mi-
plex, causing aberrant splicing variants [117]. The deletion of a 4-bp tochondrial transport (Fig. 4).
intron at the splice donor site of exon 19 has been observed to impair
normal splicing [126]. The reported study showed that LRRK2 6.1.1. Mitochondrial morphology or mitochondrial dynamic
(G2019S) mutant cDNA causes the increased inclusion of exon 10 of Aβ peptide affects mitochondrial fusion and fission cycle. Fission
MAPT gene [119]. High levels of wild type LRRK2 gene influences the cycle of mitochondria allows damaged mitochondria to reduce their
expression of SNCA gene and gives rise to dysfunction in SNCA gene, size so it is easily selected through autophagy process [87]. On the

769
S.H. Tan et al. Biomedicine & Pharmacotherapy 111 (2019) 765–777

neurons showed a notable decrease in anterograde movement of mi-


tochondria [153]. Therefore, mitochondrial fusion and fission defects
(Section 6.1.1) secondarily impair motility, and conversely, transport
defects affect mitochondrial morphology in neurons. However, the
mechanisms underlying this interplay remain to be determined. An
ultrastructural study concluded that there is a loss in normal micro
tubular structure around intracellular Aβ peptide, where this reduction
could lead to impaired mitochondria movement inside neurons [154].

6.2. Parkinson’s disease

PD is described by several characteristics, one of which is mi-


Fig. 4. Disrupted mitochondria especially in terms of their morphology, bioe-
tochondrial dysfunction. Over the last few years, evidence has accu-
nergetics and transport contribute to the risk of AD.
mulated that mitochondrial dysfunction is strongly associated with PD.
Several mutants are known to contribute to abnormal mitochondrial
other hand, fusion cycle allows mitochondrial enzymes and metabolites function, one of which is the mutated SNCA [155]. The accumulation of
to be distributed evenly throughout the mitochondria compartments. insoluble cytoplasmic proteins known as Lewy bodies in neurons is one
Fusion cycle can also slow down the aging process of mitochondria the major characteristics that contribute to PD. SNCA protein is closely
[75]. Dynamin-related protein 1 (Drp1) and mitochondrial fission associated with disruption of mitochondrial function in PD by attaching
protein 1 (Fis1) are involved in fission cycle, while optic atrophy pro- onto mitochondrial membrane and damaging the protein transport
tein (Opa1) and Mitofusins 1 and 2 (Mfn1 and Mfn2) are involved in system [156]. In other words, the destructions caused by SNCA is not
fusion cycle [143]. Aβ peptide can enhance mitochondria fission and only limited to mitochondrial protein transport but also other mi-
reduce mitochondria fusion by affecting the expression of genes [143]. tochondrial functions due to its localization on outer membrane [157].
Therefore, like mitochondrial fusion, fission seems to be required to When the genes in mitochondria undergo mutation, the normal
maintain mitochondrial function in the brain. Trimmer and Borland homeostasis of mitochondria is interfered, which further leads to neu-
(2005) analysed the AD patients brain samples and found the changes ronal toxicity, loss of dopamine neurons and eventually PD. A recent
in mitochondrial morphology [144]. Another study on M17 neuro- study using a transcription factor A, mitochondrial (TFAM) knockout
blastoma cell line with wild-type and Swedish mutant APP (over- mouse model-induced mitochondrial dysfunction has shown depletion
expressed) concluded that cells with Swedish mutant APP showed mi- in dopaminergic neurons [158]. Other than TFAM, two other genes
tochondrial morphology changes from thin and elongated to discovered in recent studies have demonstrated relationship with mi-
fragmented and punctiform mitochondria [145]. When these protective tochondrial dysfunction in familial PD, known as PARKIN and PINK1
mechanisms fail, mitochondrial fission can also promote apoptosis, cell [159] (Fig. 5). These two genes are the critical elements for the quality
death. control of mitochondria [160]. Moreover, there is evidence for a direct
involvement of PINK1 and PARKIN in abnormal mitochondrial dy-
6.1.2. Mitochondrial bioenergetics namics in PD models [161,162]. When mitochondria is damaged, mi-
Electron transport chain (ETC) in mitochondria is responsible for tochondrial depolarization arises when PINK1 selectively accumulate
the production of ATP besides producing antioxidants to prevent neu- on outer membrane of mitochondria (OMM) [163]. PINK1 accumulated
ronal injury [146]. The affected electron transport chain influences the on OMM phosphorylates Parkin and ubiquitin which then leads to the
ATP production and antioxidants expression, however, for synaptic exposure of Parkin ubiquitin ligase activity [164]. The activation of
transmission in brain, high levels of ATP are required [147]. Low level Parkin causes further damage to OMM proteins and stimulates ubi-
of ATP also matches the evidence that energy deficiency in brain is the quitination [165]. At the same time, the affected mitochondria migrate
fundamental characteristic in AD [148]. Low level of ATP also matches to the perinuclear region and undergo formation of mito-aggresomes by
the evidence that energy deficiency in brain is the fundamental char- mitophagy pathway [164].
acteristic in AD [148]. However, in the neurons, PINK1 and PARKIN trigger the degen-
Aβ peptide induces reactive oxygen species (ROS) elevation by eration of the motor adaptor protein Miro, so the movement of the
causing dysfunction in complex I of electron transport system in mi- damaged mitochondria along the axon of neuron may be inhibited
tochondria. When complex I is blocked, amount of ATP reduced, [166]. Eventually, the mitochondria stop at distal ends of neurons and
however amount of ROS increased as higher proportion of oxygen is they are blocked from clearance due to the PINK1/Parkin-mediated
converted into free radicals [149]. As mitochondria are susceptible to pathway [158]. Also, insufficiency of PINK1 proteins is the causative
high ROS levels, they are attacked by ROS which eventually results in factor of Ca2+ accumulation in mitochondria which subsequently leads
mitochondrial membrane potential loss and leads to further ROS in- to the production of ROS and cell apoptosis in PD. The dysfunctional
crease [145,150]. Increased ROS levels and reduced ATP production in mitochondria will lead to decreased membrane potential as well as low
mitochondria ETC pathway leads to AD. This is supported by a study ATP production. The mitochondria may become more susceptible upon
that involved mitochondria preparations from several AD mice models the exposure to toxins. Studies have revealed that mutated PINK1 leads
which revealed a decreased in ATP production and excessive high ROS to recessive type of PD [159].
levels [151]. Furthermore, there are other factors that contribute to mitochon-
drial dysfunction via several different pathways. Both mutated gluco-
6.1.3. Mitochondrial transport cerebrosidase 1 (GBA 1) gene and quinone oxidase products encourage
Movement of mitochondria is important for maintaining normal the production of ROS. The effect of oxidative damage on the mi-
neuronal polarity and synaptic functions such as synaptic potential, tochondrial function is due to the selective loss of mitochondrial com-
synaptic transmission and synaptic plasticity. Obstruction in mi- plex I and alpha-ketoglutarate dehydrogenase complex (KGDHC). This
tochondrial movement can affect communication between neuronal can downregulate the mitochondrial respiratory system and cause in-
cells through synaptic [87,88]. Kinesin is responsible to move mi- trinsic cell death in PD [159].
tochondria in anterograde direction towards the nerve terminal while Mitochondrial dysfunction also occurs when the voltage-dependent
dynein moves mitochondria in retrograde direction towards soma anions channels (VDACs) are blocked [167]. The study proved that
[152]. A study carried out by Calkins et al. in mouse hippocampal homocysteine is a kind of putative neurotoxin that inhibits activity of

770
S.H. Tan et al. Biomedicine & Pharmacotherapy 111 (2019) 765–777

Fig. 5. Illustration of PINK1 and Parkin dysregulation in mitochondria effect on PD.

mitochondrial complex I and leads to loss of dopaminergic neurons as proposed to be associated with the DNA damage, proteins, lipids as well
well as mitochondrial damage [159]. This can reduce the amount of the as activation of apoptosis or autophagy pathways, linking to cell death
important antioxidant, glutathione (GSH) and induce the risk of da- [174,181]. It is suggested that the brain is highly susceptible to oxi-
maged neurons by mitochondria [159,168]. dative stress due to several factors. For instance, ROS formation is
Apart from that, Lon protease, which is found in matrix of mi- catalysed due to high oxygen demand, high polyunsaturated fatty acids
tochondria functions to destroy oxidised and impaired protein. Lon levels as well as high level of redox-active metals such as iron and
Protease dysfunction has been shown to associate with the degradation copper located in the brain cell membranes, contributing to neurode-
of mitochondria [169]. The increased ROS level enhances Lon Protease generative diseases such as AD and PD [180,182].
expression to deal with the damage caused by free radicals. Therefore,
increased expression of Lon Protease along with oxidised protein ac- 7.1. Alzheimer’s disease
cumulation are the indications of neurodegeneration in PD. The study
also revealed that dysfunctional Lon Protease has the ability to inhibit The gradual neuronal loss in AD is mainly contributed by the ag-
proteasome activity which eventually results in mitochondrial impair- gregation of Aβ plaques and tau tangles [183]. It has been proposed
ment [170]. that oxidative stress is associated with Aβ and tau tangles aggregation
via several major mechanisms such as mitochondrial dysfunction,
protein and DNA oxidation, lipid peroxidation as well as proteins ag-
7. Role of oxidative stress and ROS in neurodegenerative diseases gregation (Fig. 6).

ROS are a group of highly reactive oxygen-derived molecules, 7.1.1. Oxidative stress/ROS and mitochondrial dysfunction
characterized by short lifespan due to the unpaired electrons [171,172]. Excessive accumulation of Aβ will increase ROS accumulation and
The examples of ROS include free radicals (superoxide, O2−), non-ra- result in increased oxidative stress. On the other hand, oxidative stress
dicals (hydrogen peroxide, H2O2) and hydroxyl radicals (%OH). They promotes both expression and activity of β- and γ-secretases, so the Aβ
can be generated exogenously and endogenously in human body. production increases [184]. Consequently, they give rise to mitochon-
Meanwhile, their production is normally balanced by antioxidant sys- drial dysfunction [181]. Due to mitochondrial dysfunction, there will
tems in the body. [173]. Also, both environmental toxins and chemicals be a reduction of energy metabolism in brain since ATP production is
are believed to be involved in the ROS production as by-products of inhibited [184,185]. This leads to reduced neuronal gene expression
their respective metabolisms [174]. In human body, ROS are primarily which codes for the subunits of mitochondrial ETC [185]. Meanwhile,
produced via mitochondrial respiratory chain [175,176].
Despite being the by-products of cellular metabolism, optimized
level of ROS is essential to initiate pro-survival pathways as well as
cellular signalling from mitochondria to the rest of the cell [172].
Under normal circumstances, ROS induce the redox sensitive signal
cascades via the redox-buffering capacity of intracellular thiols in-
cluding GSH and thioredoxin (TRX) [177]. As for TRX, it mediates the
protein activity through direct binding to them. Due to their partici-
pation in cell signalling, the high ratios of reduced to oxidized forms are
regulated by their respective enzymes (GSH and TRX reductases) [178].
Consequently, the antioxidant enzymes will act against the cellular
oxidative stress through the reduction of H2O2. As the expression of the
enzymes elevates, they enable the cells to survive through oxidant ex-
posures [179]. However, if there is an imbalance between the anti-
oxidant defence and ROS generation, ROS will accumulate in excess Fig. 6. The connection between ROS, mitochondrial dysfunction, protein, DNA
and thus, oxidative stress happens [180]. The oxidative stress is oxidation, and lipid peroxidation in development of AD.

771
S.H. Tan et al. Biomedicine & Pharmacotherapy 111 (2019) 765–777

mitochondrial dysfunction impairs the Ca2+ homeostasis [186]. The


accumulation of mitochondrial Ca2+ increases ROS generation and
triggers the opening of permeability transition pore (PTP) [181].
Hence, pro-apoptotic molecules translocate from mitochondria towards
apoptosis, a pathway that promotes neural cell death.

7.1.2. Oxidative stress/ROS and oxidation of protein, DNA and lipid


peroxidation
Furthermore, oxidative stress secondary to excess Aβ provokes AD
through the elevated levels of by-products from protein and DNA oxi-
dation as well as lipid peroxidation. In APP transgenic mice study, had
pronounced elevated H2O2 as well as peroxidation of lipid and protein
levels [184]. Also, ROS-mediated protein oxidation links up the protein Fig. 7. The dopamine metabolism, neuroinflammation can induce ROS/oxida-
oxidative damage in AD [187]. On top of that, as DNA bases are highly tive stress level that give rise to PD.
susceptible to oxidative stress, they will undergo oxidative damage
through some ROS-mediated processes [188]. Therefore, the role of
neuroinflammation and SNCA (Fig. 7).
DNA oxidation in AD is validated when increased oxidative damage on
mitochondrial DNA has been shown in aging and in various regions of
AD brain. 7.2.1. Oxidative stress/ROS and enhanced dopamine metabolism
As in lipid peroxidation, their markers such as 4-hydroxynonenal (4- Being an unstable molecule, dopamine carries out auto-oxidation
HNE) and malondialdehyde (MDA) are related to oxidative stress [189]. and generates dopamine quinones and free radicals [200]. In healthy
Lipid peroxidation is a chain reaction where the double bond of un- brain, dopamine levels is mediated by monoamine oxidase-A (MAO-A);
saturated fatty acids is attacked by radicals and formation of highly however, in brain with PD, MAO-B increases and metabolizes dopamine
reactive lipid peroxyl radicals takes places [190]. Due to its ability in predominantly [201]. As a result, H2O2 is formed as the by-product of
protein modification, 4-HNE inhibits glutamate transporters and trig- MAO-B-mediated dopamine metabolism [202]. Thus, the enhanced
gers dysregulation of intracellular signalling of Ca2+, which eventually dopamine metabolism could account for the aggregation of toxic radi-
promotes apoptotic cascade [191,192]. Apart protein modification, 4- cals like ·OH in the brain [203]. Meanwhile, the dopamine quinones
HNE is also cytotoxic to neurons as it impedes the function of neuronal result from dopamine metabolism will then undergo cyclisation process
glucose transporter 3. Hence, it is stated that 4-HNE is a marker and and produce highly reactive amino chrome, which provokes the O2−
toxin to AD [193]. There are studies to advocate this statement by production [204]. As more ROS are being generated, the oxidative
proving that increased level of lipid peroxidation markers has been stress occurs which results in dopaminergic neural cell death and causes
observed in the temporal lobe of AD [189]. Lipid peroxidation is usually PD.
accompanied by poor capability of antioxidant system in AD; the levels
of antioxidants (e.g. Vitamin C & E, uric acid) and antioxidant enzymes 7.2.2. Oxidative stress/ROS and neuroinflammation
(e.g. GSH peroxidase, SOD) show profound reduction in patients with It is proposed that the relationship between oxidative stress and
AD [184,185]. The insufficient antioxidant potential in brain will fa- neuroinflammation contributes to the neurodegeneration in PD [205].
vour oxidative stress and soon this may result in AD [189]. Under normal circumstance, microglia serve as phagocytes to protect
the CNS against microbial pathogens as well as the mediators of
7.1.3. Oxidative stress/ROS and proteins aggregation adaptive and innate immune system [206]. In PD, they become acti-
Generally, AD comprises the accumulation of abnormally-folded tau vated and cause neuronal injury via the production of ROS and pro-
and Aβ proteins in the brain. The accumulation of aberrant proteins inflammatory factors [207]. Activated microglia are crucial sources of
provokes loss or even death of brain neurons. It has been suggested that ROS (e.g. O2−) so they become the culprits for the occurrence of oxi-
oxidative stress is the key factor for protein aggregation in AD. When dative stress in brain with PD. Furthermore, ROS is able to activate the
tau proteins encounter oxidative stress, they will undergo hyperpho- pro-inflammatory pathways, which in turn actuate detrimental en-
sphorylation and lose their biological function [194]. Similarly, when vironment for susceptible dopaminergic neurons [207].
Aβ proteins encounter oxidative stress, the structures of β-sheet and This has been proven when the existence of activated microglia
random coil in the proteins increase and thus, misfolding of Aβ proteins along with increased level of cytokines is found in the SNPC of patients
occurs and leads to Aβ aggregation in the brain [195]. with PD [207,208]. Therefore, neuronal death secondary to activated
Under normal circumstances, proteasome is responsible to eliminate microglia will further promote the activation of microglial cells, de-
oxidized and misfolded proteins [196]. Following aggregation, the veloping a neurotoxic vicious cycle of inflammation perpetuating tissue
aberrant proteins tend to act as the inhibitors of proteasome activity injury [209]. All these processes will then lead to neurodegeneration in
[197]. When proteasome activity is inhibited, the degradation of oxi- PD.
dized and misfolded proteins will be alleviated [76]. Consequently,
aberrant proteins will accumulate in excess and this will further con- 7.2.3. Oxidative stress/ROS and SNCA
tribute to ROS production. The abnormal interaction between protein Apart from dopamine, the discovery of SNCA proteins has provided
accumulation and ROS production provokes apoptotic cascade me- a vital insight into the pathophysiology of oxidative stress-mediated PD.
chanism in AD. Being a natively unfolded neuronal protein, SNCA is responsible to
reuptake synaptic vesicle, store and compartmentalize neuro-
7.2. Parkinson’s disease transmitters (e.g. dopamine) [210]. In patients with PD, SNCA genes
are mutated and multiplicated, so they are more prone to form ag-
Unlike AD, PD is manifested by progressive dopaminergic neurons gregates [211]. Besides that, oxidative stress is able to modify the nu-
loss and diminished dopamine levels in the brain [198]. As the brain clear membrane and translocate SNCA to the nucleus, where formation
needs very high oxygen demand, ROS has become one of the notable of SNCA-histone complexes takes place, so this will lead to SNCA oli-
hallmarks to the loss of dopaminergic neurons in PD [199]. It has been gomerization into insoluble fibrils [212]. These insoluble fibrils will
suggested that ROS-mediated DNA damage in PD is linked to several then induce the deposition of abnormal protein aggregates, known as
pathways, such as enhanced metabolism of dopamine, presence of Lewy bodies [213].

772
S.H. Tan et al. Biomedicine & Pharmacotherapy 111 (2019) 765–777

designs can provide confirmatory evidence that is crucial for efficient


clinical translation.
Pathways will provide ideal vehicles for integrating relevant find-
ings from experimental study. Although treatments are available for
both diseases, the role of treatment is primarily to prevent or delay the
progress of the diseases instead of fully overcoming the diseases. The
challenge in the near future will be to determine effective drugs to
tackle the dysregulated biological pathways for the treatment of neu-
rodegenerative diseases. Drug discovery for AD and PD treatments are
still ongoing and establishing the most desirable design features with
appropriate chemical properties is an interested area of research. A
challenge for future studies is to unravel the molecular nature of these
major pathways interactions for drug research. There is compelling
reason to hope that efforts to high throughput omics based studies will
ultimately lead to new biomarker and therapeutic approaches in AD,
Fig. 8. Conceptual model of candidate pathways contributing to AD, PD neu- PD. Therefore, broader and deeper knowledge on drug aspects are re-
rodegeneration. quired for the target-specific treatments so that lifespans among the
general population can be prolonged in future.
Besides the impact of oxidative stress due to the presence of SNCA
References
fibrils, the mutated SNCA is also related to elevated production of ROS
[214]. Additionally, this notion has been validated when over-
[1] C.E. Cicero, G. Mostile, R. Vasta, V. Rapisarda, S.S. Signorelli, M. Ferrante,
expression of mutant SNCA enhances the risk of dopamine toxicity M. Zappia, A. Nicoletti, Metals and neurodegenerative diseases. A systematic re-
[215], leading to defects in mitochondria and apoptosis, which further view, Environ. Res. 159 (2017) 82–94.
increases their vulnerability towards oxidative stress [216]. These re- [2] R.C. Gardner, K. Yaffe, Epidemiology of mild traumatic brain injury and neuro-
degenerative disease, Mol. Cell. Neurosci. 66 (2015) 75–80.
sults explain that ROS enhances SNCA accumulation, which in turn
[3] G.K. Tofaris, A.H.V. Schapira, Neurodegenerative diseases in the era of targeted
increases ROS, creating a dual pathogenic mechanism leading to neu- therapeutics: how to handle a tangled issue, Mol. Cell. Neurosci. 66 (2015) 1–2.
rodegeneration. [4] D.J. Selkoe, J. Hardy, The amyloid hypothesis of Alzheimer’s disease at 25 years,
EMBO Mol. Med. 8 (2016) 595–608.
In short, through a detailed review, the major pathways common to
[5] R.A. Armstrong, What causes Alzheimer’s disease? Folia Neuropathol. 51 (2013)
AD and PD have been discerned (Fig. 8). These pathways are helpful to 169–188.
discover new targets for further study as well as biomarker and drug [6] M. Rezazadeh, A. Khorrami, T. Yeghaneh, M. Talebi, Genetic factors affecting late-
development. onset Alzheimer’s disease susceptibility, Neuromolecular Med. 18 (2016) 37–49.
[7] M. Hornberger, O. Piguet, Episodic memory in frontotemporal dementia: a critical
review, Brain 135 (2012) 678–692.
[8] S.K. Ravi, B.N. Ramesh, R. Mundugaru, B. Vincent, Multiple pharmacological
8. Conclusions and future directions activities of Caesalpinia crista against aluminium-induced neurodegeneration in
rats: relevance for Alzheimer’s disease, Environ. Toxicol. Pharmacol. 58 (2018)
In conclusion, neurodegenerative diseases such as AD and PD are 202–211.
[9] A.E. Mungenast, S. Siegert, L.H. Tsai, Modeling Alzheimer’s disease with human
usually untreatable as the clinical manifestations showed the perma- induced pluripotent stem (iPS) cells, Mol. Cell. Neurosci. 73 (2016) 13–31.
nent loss of specific neurons in patients. There are several risk factors [10] L. Crews, E. Masliah, Molecular mechanisms of neurodegeneration in Alzheimer’s
contributed to AD and PD, including aging, genetic and environmental disease, Hum. Mol. Genet. Gen. 19 (2010) R12–R20.
[11] L.I. Binder, A.L. Guillozet-Bongaarts, F. Garcia-Sierra, R.W. Berry, Tau, tangles,
factors. Taken together, the selected pathways perturbations are di- and Alzheimer’s disease, Biochim. Biophys. Acta - Mol. Basis Dis. 1739 (2005)
rectly or indirectly generated by several factors involved in PD, AD 216–223.
neurodegenerative diseases. However, our understanding of the me- [12] B.P. Imbimbo, J. Lombard, N. Pomara, Pathophysiology of Alzheimer’s disease,
Neuroimaging Clin. N. Am. 15 (2005) 727–753.
chanisms involved remains rudimentary at best. Through a detailed [13] D.J. Moore, A.B. West, V.L. Dawson, T.M. Dawson, Molecular pathophysiology of
review of pathways and molecular events, numerous pathways common Parkinson’s disease, Annu. Rev. Neurosci. 28 (2005) 57–87.
to AD and PD have been identified, which nominate promising new [14] P. Rizek, N. Kumar, M.S. Jog, An update on the diagnosis and treatment of
Parkinson disease, Cmaj 188 (2016) 1157–1165.
targets for further study as well as biomarker and drug development. [15] C. Politi, C. Ciccacci, G. Novelli, P. Borgiani, Genetics and treatment response in
On top of that, increasing evidence suggests that UPS dysregulation, Parkinson’s disease: an update on pharmacogenetic studies, Neuromol. Med. 20
aberrant alternative splicing, mitochondrial dysfunction and over- (2018) 1–17.
[16] K.A. Jellinger, A.D. Korczyn, Are dementia with Lewy bodies and Parkinson’s
expression of ROS are largely operative in the neurodegenerative dis-
disease dementia the same disease? BMC Med. 16 (2018) 1–16.
eases especially AD and PD. These findings build on established notions [17] A.H.V. Schapira, Glucocerebrosidase and Parkinson disease: recent advances, Mol.
of complex disease aetiology, with multiple processes presumed to in- Cell. Neurosci. 66 (2015) 37–42.
fluence neurodegeneration and clinical outcomes in AD, PD, and related [18] O. Marques, T.F. Outeiro, Alpha-synuclein: from secretion to dysfunction and
death, Cell Death Dis. 3 (2012) e350–7.
disorders. They also advance the understanding of mechanisms likely to [19] H. Sanjari Moghaddam, Z. Valitabar, A. Ashraf-Ganjouei, M. Mojtahed Zadeh,
be crucial in maintaining brain structure and function during normal F. Ghazi Sherbaf, M.H. Aarabi, Cerebrospinal fluid C-reactive protein in
aging. Parkinson’s disease: associations with motor and non-motor symptoms, Neuromol.
Med. 20 (2018) 376–385.
Despite the explanations of the biological pathways in this review, [20] M.A. Stephan, B. Meier, S.W. Zaugg, A. Kaelin-Lang, Motor sequence learning
more studies and researches are still required to support the causal performance in Parkinson’s disease patients depends on the stage of disease, Brain
relationship between the biological pathways and the neurodegenera- Cogn. 75 (2011) 135–140.
[21] E. Niedzielska, I. Smaga, M. Gawlik, A. Moniczewski, P. Stankowicz, J. Pera,
tive diseases as some of the mechanisms remain obscure. These insights M. Filip, Oxidative Stress in neurodegenerative diseases, Mol. Neurobiol. 53
also suggest that biomarker and treatment strategies may require si- (2016) 4094–4125.
multaneous targeting of multiple components, including some specific [22] A.P. Singh, T. Bajaj, D. Gupta, S.B. Singh, A. Chakrawarty, V. Goyal, A.B. Dey,
S. Dey, Serum Mortalin correlated with α-synuclein as serum markers in
to disease stage, in order to assess and modulate neurodegeneration. Parkinson’s disease: a pilot study, Neuromol. Med. 20 (2018) 83–89.
More broadly, pathways and networks can serve as vehicles for in- [23] H. Kabayama, N. Tokushige, M. Takeuchi, M. Kabayama, M. Fukuda,
tegrating findings from diverse studies of neurodegeneration. There are K. Mikoshiba, Parkin promotes proteasomal degradation of synaptotagmin IV by
accelerating polyubiquitination, Mol. Cell. Neurosci. 80 (2017) 89–99.
many active strategies for large scale omics analysis of neurodegen-
[24] E. Noe, K. Marder, K.L. Bell, D.M. Jacobs, J.J. Manly, Y. Stern, Comparison of
erative disease, and findings that converge across these multiple study

773
S.H. Tan et al. Biomedicine & Pharmacotherapy 111 (2019) 765–777

dementia with Lewy bodies to Alzheimer’s disease and Parkinson’s disease with Pharmacol. Toxicol. 54 (2014) 141–164.
dementia, Mov. Disord. 19 (2004) 60–67. [49] A. Goate, M.-C. Chartier-Harlin, M. Mullan, J. Brown, F. Crawford, L. Fidani,
[25] D.W. Dickson, B. Dubois, M. Emre, S. Fahn, J.M. Farmer, D. Galasko, C.G. Goetz, L. Giuffra, A. Haynes, N. Irving, L. James, R. Mant, P. Newton, K. Rooke,
J.H. Growdon, J. Hardy, P. Heutink, K. Kosaka, V.M. Lee, J.B. Leverenz, P. Roques, C. Talbot, M. Pericak-Vance, A. Roses, R. Williamson, M. Rossor,
E. Masliah, I.G. Mckeith, C.W. Olanow, B.M. Ravina, A.B. Singleton, C.M. Tanner, M. Owen, J. Hardy, Segregation of a missense mutation in the amyloid precursor
DLB and PDD boundary issues: diagnosis, treatment, molecular pathology, and protein gene with familial Alzheimer’s disease, Nature 349 (1991) 704–706.
biomarkers, Neurology 68 (2008) 812–819. [50] D.J. Selkoe, M.S. Wolfe, Presenilin: running with scissors in the membrane, Cell
[26] R.B. Postuma, D. Berg, M. Stern, W. Poewe, C.W. Olanow, W. Oertel, J. Obeso, 131 (2007) 215–221.
K. Marek, I. Litvan, A.E. Lang, G. Halliday, C.G. Goetz, T. Gasser, B. Dubois, [51] D.M. Maraganore, M. de Andrade, A. Elbaz, M.J. Farrer, J.P. Ioannidis, R. Krüger,
P. Chan, B.R. Bloem, C.H. Adler, G. Deuschl, MDS clinical diagnostic criteria for W.A. Rocca, N.K. Schneider, T.G. Lesnick, S.J. Lincoln, M.M. Hulihan, J.O. Aasly,
Parkinson’s disease, Mov. Disord. 30 (2015) 1591–1601. T. Ashizawa, M.-C. Chartier-Harlin, H. Checkoway, C. Ferrarese, G. Hadjigeorgiou,
[27] K.W. Park, H.S. Kim, S.M. Cheon, J.K. Cha, S.H. Kim, J.W. Kim, Dementia with N. Hattori, H. Kawakami, J.-C. Lambert, T. Lynch, G.D. Mellick,
Lewy bodies versus Alzheimer’s disease and Parkinson’s disease dementia: a S. Papapetropoulos, A. Parsian, A. Quattrone, O. Riess, E.-K. Tan, C. Van
comparison of cognitive profiles, J. Clin. Neurol. 7 (2011) 19–24. Broeckhoven, for the G.E. of P.D. (GEO-P. Consortium, Collaborative analysis of α-
[28] J. Meireles, J. Massano, Cognitive impairment and dementia in Parkinson’s dis- synuclein gene promoter variability and Parkinson disease, JAMA 296 (2006)
ease: clinical features, diagnosis, and management, Front. Neurol. (May) (2012) 661–670.
1–15. [52] C. Paisán-Ruı́z, S. Jain, E.W. Evans, W.P. Gilks, J. Simón, M. van der Brug, A.L. de
[29] A. Lenka, K.R. Jhunjhunwala, J. Saini, P.K. Pal, Structural and functional neu- Munain, S. Aparicio, A.M. Gil, N. Khan, J. Johnson, J.R. Martinez, D. Nicholl,
roimaging in patients with Parkinson’s disease and visual hallucinations: a critical I.M. Carrera, A.S. Peňa, R. de Silva, A. Lees, J.F. Martı́-Massó, J. Pérez-Tur,
review, Park. Relat. Disord. 21 (2015) 683–691. N.W. Wood, A.B. Singleton, Cloning of the gene containing mutations that cause
[30] A.J. Harding, G.A. Broe, G.M. Halliday, Visual hallucinations in Lewy body disease PARK8-linked Parkinson’s disease, Neuron 44 (2004) 595–600.
relate to Lewy bodies in the temporal lobe, Brain 125 (2002) 391–403. [53] A. Zimprich, S. Biskup, P. Leitner, P. Lichtner, M. Farrer, S. Lincoln, J. Kachergus,
[31] I.G. McKeith, D.W. Dickson, J. Lowe, M. Emre, J.T. O’Brien, H. Feldman, M. Hulihan, R.J. Uitti, D.B. Calne, A.J. Stoessl, R.F. Pfeiffer, N. Patenge,
J. Cummings, J.E. Duda, C. Lippa, E.K. Perry, D. Aarsland, H. Arai, C.G. Ballard, I.C. Carbajal, P. Vieregge, F. Asmus, B. Müller-Myhsok, D.W. Dickson,
B. Boeve, D.J. Burn, D. Costa, T. Del Ser, B. Dubois, D. Galasko, S. Gauthier, T. Meitinger, T.M. Strom, Z.K. Wszolek, T. Gasser, Mutations in LRRK2 cause
C.G. Goetz, E. Gomez-Tortosa, G. Halliday, L.A. Hansen, J. Hardy, T. Iwatsubo, autosomal-dominant parkinsonism with pleomorphic pathology, Neuron 44
R.N. Kalaria, D. Kaufer, R.A. Kenny, A. Korczyn, K. Kosaka, V.M.Y. Lee, A. Lees, (2004) 601–607.
I. Litvan, E. Londos, O.L. Lopez, S. Minoshima, Y. Mizuno, J.A. Molina, [54] J. Mutter, J. Naumann, C. Sadaghiani, R. Schneider, H. Walach, Alzheimer disease:
E.B. Mukaetova-Ladinska, F. Pasquier, R.H. Perry, J.B. Schulz, J.Q. Trojanowski, mercury as pathogenetic factor and apolipoprotein E as a moderator, Neuro
M. Yamada, Diagnosis and management of dementia with Lewy bodies: third re- Endocrinol. Lett. 25 (2004) 331–339.
port of the DLB consortium, Neurology 65 (2005) 1863–1872. [55] M. Santibáñez, F. Bolumar, A.M. García, Occupational risk factors in Alzheimer’s
[32] F. Blanc, R. Mahmoudi, T. Jonveaux, J. Galmiche, G. Chopard, B. Cretin, disease: a review assessing the quality of published epidemiological studies,
C. Demuynck, C. Martin-Hunyadi, N. Philippi, F. Sellal, J.M. Michel, G. Tio, Occup. Environ. Med. 64 (2007) 723–732.
M. Stackfleth, P. Vandel, E. Magnin, J.L. Novella, G. Kaltenbach, A. Benetos, [56] E. Aizenman, P.G. Mastroberardino, Metals and neurodegeneration, Neurobiol.
E.A. Sauleau, Long-term cognitive outcome of Alzheimer’s disease and dementia Dis. 81 (2015) 1–3.
with Lewy bodies: dual disease is worse, Alzheimers Res. Ther. 9 (2017) 1–9. [57] V. Karri, D. Ramos, J.B. Martinez, A. Odena, E. Oliveira, S.L. Coort, C.T. Evelo,
[33] M.G. Kramberger, B. Auestad, S. Garcia-Ptacek, C. Abdelnour, J.G. Olmo, E.C.M. Mariman, M. Schuhmacher, V. Kumar, Differential protein expression of
Z. Walker, A.W. Lemstra, E. Londos, F. Blanc, L. Bonanni, I. McKeith, B. Winblad, hippocampal cells associated with heavy metals (Pb, As, and MeHg) neurotoxicity:
F.J. De Jong, F. Nobili, E. Stefanova, M. Petrova, C. Falup-Pecurariu, I. Rektorova, deepening into the molecular mechanism of neurodegenerative diseases, J.
S. Bostantjopoulou, R. Biundo, D. Weintraub, D. Aarsland, Long-term cognitive Proteomics 187 (2018) 106–125.
decline in dementia with lewy bodies in a large multicenter, international cohort, [58] L. Baum, I.H.S. Chan, S.K.K. Cheung, W.B. Goggins, V. Mok, L. Lam, V. Leung,
J. Alzheimers Dis. 57 (2017) 787–795. E. Hui, C. Ng, J. Woo, H.F.K. Chiu, B.C.Y. Zee, W. Cheng, M.H. Chan, S. Szeto,
[34] R.C. Brown, A.H. Lockwood, B.R. Sonawane, Neurodegenerative diseases: an V. Lui, J. Tsoh, A.I. Bush, C.W.K. Lam, T. Kwok, Serum zinc is decreased in
overview of environmental risk factors, Environ. Health Perspect. 113 (2005) Alzheimer’s disease and serum arsenic correlates positively with cognitive ability,
1250–1256. BioMetals 23 (2010) 173–179.
[35] O. Myhre, H. Utkilen, N. Duale, G. Brunborg, T. Hofer, Metal dyshomeostasis and [59] X. Li, Y. Lv, S. Yu, H. Zhao, L. Yao, The effect of cadmium on Aβ levels in APP/PS1
inflammation in Alzheimer’s and Parkinson’s diseases: possible impact of en- transgenic mice, Exp. Ther. Med. 4 (2012) 125–130.
vironmental exposures, Oxid. Med. Cell. Longev. 2013 (2013). [60] A. Sabarwal, K. Kumar, R.P. Singh, Hazardous effects of chemical pesticides on
[36] T. Finkel, N.J. Holbrook, Oxidants, oxidative stress and the biology of ageing, human health- cancer and other associated disorders, Environ. Toxicol.
Nature 408 (2000) 239–247. Pharmacol. 63 (2018) 103–114.
[37] B.A.I. Payne, P.F. Chinnery, Mitochondrial dysfunction in aging: much progress [61] C. Van Den Heuvel, E. Thornton, R. Vink, Traumatic brain injury and Alzheimer’s
but many unresolved questions, Biochim. Biophys. Acta – Bioenergy 1847 (2015) disease: a review, Prog. Brain Res. 161 (2007) 303–316.
1347–1353. [62] S. Coon, A. Stark, E. Peterson, A. Gloi, G. Kortsha, J. Pounds, D. Chettle, J. Gorell,
[38] R. Macdonald, K. Barnes, C. Hastings, H. Mortiboys, Mitochondrial abnormalities Whole-body lifetime occupational lead exposure and risk of Parkinson’s disease,
in Parkinson’s disease and Alzheimer’s disease: can mitochondria be targeted Environ. Health Perspect. 114 (2006) 1872–1876.
therapeutically? Biochem. Soc. Trans. 46 (2018) 891–909. [63] A.B. Cholanians, A.V. Phan, E.J. Ditzel, T.D. Camenisch, S.S. Lau, T.J. Monks,
[39] L. Wu, P. Rosa-neto, G.R. Hsiung, A.D. Sadovnick, M. Masellis, S.E. Black, J. Jia, Arsenic induces accumulation of α-synuclein: implications for synucleinopathies
S. Gauthier, Early-onset familial Alzheimer’s disease (EOFAD), Can. J. Neurol. Sci. and neurodegeneration, Toxicol. Sci. 153 (2016) 271–281.
39 (2018) 436–445. [64] J.M. Gorell, B.A. Rybicki, C.C. Johnson, E.L. Peterson, Occupational metal ex-
[40] Y.F. Shea, L.W. Chu, A.O.K. Chan, J. Ha, Y. Li, Y.Q. Song, A systematic review of posures and the risk of Parkinson’s disease, Neuroepidemiology 18 (1999)
familial Alzheimer’s disease: differences in presentation of clinical features among 303–308.
three mutated genes and potential ethnic differences, J. Formos. Med. Assoc. 115 [65] Y. Wang, J. Fang, S.S. Leonard, K.M.K. Rao, Cadmium inhibits the electron transfer
(2016) 67–75. chain and induces reactive oxygen species, Free Radic. Biol. Med. 36 (2004)
[41] C.C. Liu, T. Kanekiyo, H. Xu, G. Bu, Apolipoprotein E and Alzheimer disease: risk, 1434–1443.
mechanisms and therapy, Nat. Rev. Neurol. 9 (2013) 106–118. [66] F.D. Dick, Parkinson’s disease and pesticide exposures, Br. Med. Bull. 79–80
[42] H. Checkoway, J.I. Lundin, S.N. Kelada, Neurodegenerative diseases, IARC Sci. (2006) 219–231.
Publ. 1 (2011) 407–419. [67] A. Elbaz, Parkinson’s disease and rural environment, Rev. Prat. 57 (2007) 37–39.
[43] C. Klein, A. Westenberger, Genetics of Parkinson’s disease, Cold Spring Harb. [68] S.M. Goldman, C.M. Tanner, D. Oakes, G.S. Bhudhikanok, A. Gupta,
Perspect. Med. 2 (2012) 363–369. J.W. Langston, Head injury and Parkinson’s disease risk in twins, Ann. Neurol. 60
[44] E. Greggio, M.R. Cookson, Leucine-rich repeat kinase 2 mutations and Parkinson’s (2006) 65–72.
disease: three questions, ASN Neuro 1 (2009) e00002. [69] J. Wong, Altered expression of RNA splicing proteins in Alzheimer’s disease pa-
[45] T. Kitada, S. Asakawa, N. Hattori, H. Matsumine, Y. Yamamura, S. Minoshima, tients: evidence from two microarray studies, Dement. Geriatr. Cogn. Dis. Extra 3
M. Yokochi, Y. Mizuno, N. Shimizu, Mutations in the parkin gene cause autosomal (2013) 74–85.
recessive juvenile parkinsonism, Nature 392 (1998) 605–608. [70] L. Migliore, F. Coppedè, Genetics, environmental factors and the emerging role of
[46] V. Bonifati, P. Rizzu, M.J. Van Baren, O. Schaap, G.J. Breedveld, E. Krieger, epigenetics in neurodegenerative diseases, Mutat. Res. Mol. Mech. Mutagen. 667
M.C.J. Dekker, F. Squitieri, P. Ibanez, M. Joosse, J.W. Van Dongen, N. Vanacore, (2009) 82–97.
J.C. Van Swieten, A. Brice, G. Meco, C.M. Van Duijn, B.A. Oostra, P. Heutink, [71] Y. Tu, C. Chen, J. Pan, J. Xu, Z.-G. Zhou, C.-Y. Wang, The ubiquitin proteasome
Mutations in the DJ-1 gene associated with autosomal recessive early-onset par- pathway (UPP) in the regulation of cell cycle control and DNA damage repair and
kinsonism, Science 299 (2003) 256–259. its implication in tumorigenesis, Int. J. Clin. Exp. Pathol. 5 (2012) 726–738.
[47] E.M. Valente, P.M. Abou-Sleiman, V. Caputo, M.M.K. Muqit, K. Harvey, S. Gispert, [72] S. Oddo, The ubiquitin-proteasome system in Alzheimer’s disease, J. Cell. Mol.
Z. Ali, D. Del Turco, A.R. Bentivoglio, D.G. Healy, A. Albanese, R. Nussbaum, Med. 12 (2008) 363–373.
R. González-Maldonado, T. Deller, S. Salvi, P. Cortelli, W.P. Gilks, D.S. Latchman, [73] D.J. Klionsky, S.D. Emr, Autophagy as a regulated pathway of cellular degrada-
R.J. Harvey, B. Dallapiccola, G. Auburger, N.W. Wood, Hereditary early-onset tion, Science 290 (2000) 1717–1721.
Parkinson’s disease caused by mutations in PINK1, Science 304 (2004) [74] S. Song, S.-Y. Kim, Y.-M. Hong, D.-G. Jo, J.-Y. Lee, S.M. Shim, C.-W. Chung,
1158–1160. S.J. Seo, Y.J. Yoo, J.-Y. Koh, M.C. Lee, A.J. Yates, H. Ichijo, Y.-K. Jung, Essential
[48] S.M. Goldman, Environmental toxins and Parkinson’s disease, Annu. Rev. role of E2-25K/Hip-2 in mediating amyloid-beta neurotoxicity, Mol. Cell 12

774
S.H. Tan et al. Biomedicine & Pharmacotherapy 111 (2019) 765–777

(2003) 553–563. mental retardation, Nat. Genet. 19 (1998) 134–139.


[75] A. Ciechanover, P. Brundin, The ubiquitin proteasome system in neurodegenera- [106] W. Scheper, R. Zwart, F. Baas, Alternative splicing in the N-terminus of
tive diseases: sometimes the chicken, sometimes the egg sponse, and neurode- Alzheimer’s presenilin 1, Neurogenetics 5 (2004) 223–227.
generation, Neuron 40 (2003) 427–446. [107] K. Ohe, A. Mayeda, HMGA1a trapping of U1 snRNP at an authentic 5’ splice site
[76] B.P. Tseng, K.N. Green, J.L. Chan, M. Blurton-jones, M. Frank, Aβ inhibits the induces aberrant exon skipping in sporadic Alzheimer’s disease, Mol. Cell. Biol. 30
proteasome and enhances amyloid and tau accumulation, Neuroscience 29 (2009) (2010) 2220–2228.
1607–1618. [108] E.M. Hol, R.F. Roelofs, E. Moraal, M.A.F. Sonnemans, J.A. Sluijs, E.A. Proper,
[77] S. Oh, H.S. Hong, E. Hwang, H.J. Sim, W. Lee, S.J. Shin, I. Mook-Jung, Amyloid P.N.E. De Graan, D.F. Fischer, F.W. Van Leeuwen, Neuronal expression of GFAP in
peptide attenuates the proteasome activity in neuronal cells, Mech. Ageing Dev. patients with Alzheimer pathology and identification of novel GFAP splice forms,
126 (2005) 1292–1299. Mol. Psychiatry 8 (2003) 786–796.
[78] D.M. Walsh, I. Klyubin, J.V. Fadeeva, W.K. Cullen, R. Anwyl, M.S. Wolfe, [109] F.M. De Vrij, J.A. Sluijs, L. Gregori, D.F. Fischer, W.T. Hermens, D. Goldgaber,
M.J. Rowan, D.J. Selkoe, Naturally secreted oligomers of amyloid β protein po- J. Verhaagen, F.W. Van Leeuwen, E.M. Hol, Mutant ubiquitin expressed in
tently inhibit hippocampal long-term potentiation in vivo, Nature 416 (2002) Alzheimer’s disease causes neuronal death, FASEB J. Off. Publ. Fed. Am. Soc. Exp.
535–539. Biol. 15 (2001) 2680–2688.
[79] K. Gadhave, N. Bolshette, A. Ahire, R. Pardeshi, K. Thakur, C. Trandafir, A. Istrate, [110] K. Andersen, L.J. Launer, M.E. Dewey, L. Letenneur, A. Ott, J.R. Copeland,
S. Ahmed, M. Lahkar, D.F. Muresanu, M. Balea, The ubiquitin proteasomal system: J.F. Dartigues, P. Kragh-Sorensen, M. Baldereschi, C. Brayne, A. Lobo,
a potential target for the management of Alzheimer’s disease, J. Cell. Mol. Med. 20 J.M. Martinez-Lage, T. Stijnen, A. Hofman, Gender differences in the incidence of
(2016) 1392–1407. AD and vascular dementia: the EURODEM Studies. EURODEM Incidence Research
[80] H. Mori, J. Kondo, Y. Ihara, Ubiquitin is a component of paired helical filaments in Group, Neurology 53 (1999) 1992–1997.
Alzheimer’s disease, Science 235 (1987) 1641–1644. [111] C. Wang, F. Zhang, S. Jiang, S.L. Siedlak, L. Shen, G. Perry, X. Wang, B. Tang,
[81] D. Poppek, S. Keck, G. Ermak, T. Jung, A. Stolzing, O. Ullrich, K.J.A. Davies, X. Zhu, Estrogen receptor- α is localized to neurofibrillary tangles in Alzheimer’s
T. Grune, Phosphorylation inhibits turnover of the tau protein by the proteasome: disease, Sci. Rep. 6 (2016) 20352.
influence of RCAN1 and oxidative stress, Biochem. J. 400 (2006) 511–520. [112] C.L. Lorson, E.J. Androphy, An exonic enhancer is required for inclusion of an
[82] S. Keck, R. Nitsch, T. Grune, O. Ullrich, Proteasome inhibition by paired helical essential exon in the SMA-determining gene SMN, Hum. Mol. Genet. 9 (2000)
filament-tau in brains of patients with Alzheimer’s disease, J. Neurochem. 85 22–24.
(2003) 115–122. [113] T.A. Ishunina, D.F. Swaab, Decreased alternative splicing of estrogen receptor-α
[83] R.M. Nisbet, J.-C. Polanco, L.M. Ittner, J. Götz, Tau aggregation and its interplay mRNA in the Alzheimer’s disease brain, Neurobiol. Aging 33 (2012) 286–296.e3.
with amyloid-β, Acta Neuropathol. 129 (2015) 207–220. [114] H. Zhu, Q. Ding, Lower expression level of two RAGE alternative splicing isoforms
[84] Q. Huang, M.E. Figueiredo-Pereira, Ubiquitin/proteasome pathway impairment in in Alzheimer’s disease, Neurosci. Lett. 597 (2015) 66–70.
neurodegeneration: therapeutic implications, Apoptosis 15 (2010) 1292–1311. [115] S. Massone, I. Vassallo, G. Fiorino, M. Castelnuovo, F. Barbieri, R. Borghi,
[85] K.-L. Lim, M. Tan, Role of the ubiquitin proteasome system in Parkinson’s disease, M. Tabaton, M. Robello, E. Gatta, C. Russo, T. Florio, G. Dieci, R. Cancedda,
BMC Biochem. 8 (2007) S13. A. Pagano, 17A, a novel non-coding RNA, regulates GABA B alternative splicing
[86] Y. Zhang, J. Gao, K.K.K. Chung, H. Huang, V.L. Dawson, T.M. Dawson, Parkin and signaling in response to inflammatory stimuli and in Alzheimer disease,
functions as an E2-dependent ubiquitin- protein ligase and promotes the de- Neurobiol. Dis. 41 (2011) 308–317.
gradation of the synaptic vesicle-associated protein, CDCrel-1, Proc. Natl. Acad. [116] M.H. Polymeropoulos, C. Lavedan, E. Leroy, S.E. Ide, A. Dehejia, A. Dutra, B. Pike,
Sci. 97 (2000) 13354–13359. H. Root, J. Rubenstein, R. Boyer, E.S. Stenroos, S. Chandrasekharappa,
[87] M.R. Cookson, P.J. Lockhart, C. McLendon, C. O’Farrell, M. Schlossmacher, A. Athanassiadou, T. Papapetropoulos, W.G. Johnson, A.M. Lazzarini,
M.J. Farrer, RING finger 1 mutations in Parkin produce altered localization of the R.C. Duvoisin, G. Di Iorio, L.I. Golbe, R.L. Nussbaum, Mutation in the alpha-sy-
protein, Hum. Mol. Genet. 12 (2003) 2957–2965. nuclein gene identified in families with Parkinson’s disease, Science 276 (1997)
[88] K.L. Lim, T.M. Lim, Molecular mechanisms of neurodegeneration in Parkinson’s 2045–2047.
disease: clues from Mendelian syndromes, IUBMB Life 55 (2003) 315–322. [117] K. Beyer, Alpha-synuclein structure, posttranslational modification and alternative
[89] J.F. Staropoli, C. McDermott, C. Martinat, B. Schulman, E. Demireva, splicing as aggregation enhancers, Acta Neuropathol. 112 (2006) 237–251.
A. Abeliovich, Parkin is a component of an SCF-like ubiquitin ligase complex and [118] W. Hoyer, D. Cherny, V. Subramaniam, T.M. Jovin, Impact of the acidic C-terminal
protects postmitotic neurons from kainate excitotoxicity, Neuron 37 (2003) region comprising amino acids 109-140 on alpha-synuclein aggregation in vitro,
735–749. Biochemistry 43 (2004) 16233–16242.
[90] M. Sharma, T.C. Su, Systematic mutagenesis of α-synuclein reveals distinct se- [119] H.-J. Lee, C. Choi, S.-J. Lee, Membrane-bound alpha-synuclein has a high ag-
quence requirements for physiological and pathological, J. Neurosci. 32 (2012) gregation propensity and the ability to seed the aggregation of the cytosolic form,
15227–15242. J. Biol. Chem. 277 (2002) 671–678.
[91] M. Ragland, C. Hutter, C. Zabetian, K. Edwards, Association between the ubiquitin [120] M.G. Spillantini, M.L. Schmidt, V.M. Lee, J.Q. Trojanowski, R. Jakes, M. Goedert,
carboxyl-terminal esterase l1 gene (UCHL1) S18Y variant and Parkinson’s disease: Alpha-synuclein in lewy bodies, Nature 388 (1997) 839–840.
a HuGE review and meta-analysis, Am. J. Epidemiol. 170 (2009) 1344–1357. [121] L.A. Shehadeh, K. Yu, L. Wang, A. Guevara, C. Singer, J. Vance,
[92] D.G. Healy, P.M. Abou-Sleiman, N.W. Wood, Genetic causes of Parkinson’s dis- S. Papapetropoulos, SRRM2, a potential blood biomarker revealing high alter-
ease: UCHL-1, Cell Tissue Res. 318 (2004) 189–194. native splicing in Parkinson’s disease, PLoS One 5 (2010) e9104.
[93] R.I. Morimoto, Proteotoxic stress and inducible chaperone networks in neurode- [122] T. Lynch, M. Sano, K.S. Marder, K.L. Bell, N.L. Foster, R.F. Defendini, A.A. Sima,
generative disease and aging, Genes Dev. 22 (2008) 1427–1438. C. Keohane, T.G. Nygaard, S. Fahn, Clinical characteristics of a family with
[94] J.E. Love, E.J. Hayden, T.T. Rohn, Alternative splicing in Alzheimer’s disease, J. chromosome 17-linked disinhibition-dementia-parkinsonism-amyotrophy com-
Parkinsons Dis. Alzheimers Dis. 2 (2015) 6. plex, Neurology 44 (1994) 1878–1884.
[95] R.H. Fu, S.P. Liu, S.J. Huang, H.J. Chen, P.R. Chen, Y.H. Lin, Y.C. Ho, W.L. Chang, [123] M. Hong, V. Zhukareva, V. Vogelsberg-Ragaglia, Z. Wszolek, L. Reed, B.I. Miller,
C.H. Tsai, W.C. Shyu, S.Z. Lin, Aberrant alternative splicing events in Parkinson’s D.H. Geschwind, T.D. Bird, D. McKeel, A. Goate, J.C. Morris, K.C. Wilhelmsen,
disease, Cell Transpl. 22 (2013) 653–661. G.D. Schellenberg, J.Q. Trojanowski, V.M. Lee, Mutation-specific functional im-
[96] E. Daguenet, G. Dujardin, J. Valcarcel, The pathogenicity of splicing defects: pairments in distinct tau isoforms of hereditary FTDP-17, Science 282 (1998)
mechanistic insights into pre-mRNA processing inform novel therapeutic ap- 1914–1917.
proaches, EMBO Rep. 16 (2015) 1640–1655. [124] M. Hutton, C.L. Lendon, P. Rizzu, M. Baker, S. Froelich, H. Houlden, S. Pickering-
[97] B.R. Graveley, Alternative splicing: increasing diversity in the proteomic world, Brown, S. Chakraverty, A. Isaacs, A. Grover, J. Hackett, J. Adamson, S. Lincoln,
Trends Genet. 17 (2001) 100–107. D. Dickson, P. Davies, R.C. Petersen, M. Stevens, E. de Graaff, E. Wauters, J. van
[98] O. Kelemen, P. Convertini, Z. Zhang, Y. Wen, M. Shen, M. Falaleeva, S. Stamm, Baren, M. Hillebrand, M. Joosse, J.M. Kwon, P. Nowotny, L.K. Che, J. Norton,
Function of alternative splicing, Gene 514 (2013) 1–30. J.C. Morris, L.A. Reed, J. Trojanowski, H. Basun, L. Lannfelt, M. Neystat, S. Fahn,
[99] J.M. Johnson, J. Castle, P. Garrett-Engele, Z. Kan, P.M. Loerch, C.D. Armour, F. Dark, T. Tannenberg, P.R. Dodd, N. Hayward, J.B. Kwok, P.R. Schofield,
R. Santos, E.E. Schadt, R. Stoughton, D.D. Shoemaker, Genome-wide survey of A. Andreadis, J. Snowden, D. Craufurd, D. Neary, F. Owen, B.A. Oostra, J. Hardy,
human alternative pre-mrna splicing with exon junction microarrays, Science 302 A. Goate, J. van Swieten, D. Mann, T. Lynch, P. Heutink, Association of missense
(2003) 2141–2144. and 5’-splice-site mutations in tau with the inherited dementia FTDP-17, Nature
[100] N.A. Faustino, T.A. Cooper, N. Andre, Pre-mRNA splicing and human disease, 393 (1998) 702–705.
Genes Dev. 17 (2003) 419–437. [125] M.J. Farrer, Genetics of Parkinson disease: paradigm shifts and future prospects,
[101] A. Andreadis, W.M. Brown, K.S. Kosik, Structure and novel exons of the human. Nat. Rev. Genet. 7 (2006) 306–318.
tau. gene, Biochemistry 31 (1992) 10626–10633. [126] J. Johnson, C. Paisan-Ruiz, G. Lopez, C. Crews, A. Britton, R. Malkani, E.W. Evans,
[102] C.J. Smith, B.H. Anderton, D.R. Davis, J.M. Gallo, Tau isoform expression and A. McInerney-Leo, S. Jain, R.L. Nussbaum, K.D. Foote, R.J. Mandel, A. Crawley,
phosphorylation state during differentiation of cultured neuronal cells, FEBS Lett. S. Reimsnider, H.H. Fernandez, M.S. Okun, K. Gwinn-Hardy, A.B. Singleton,
375 (1995) 243–248. Comprehensive screening of a North American Parkinson’s disease cohort for
[103] L. Chami, F. Checler, BACE1 is at the crossroad of a toxic vicious cycle involving LRRK2 mutation, Neurodegener. Dis. 4 (2007) 386–391.
cellular stress and β-amyloid production in Alzheimer’s disease, Mol. [127] T. Brudek, K. Winge, N.B. Rasmussen, J.M.C. Bahl, J. Tanassi, T.K. Agander,
Neurodegener. 7 (2012) 52. T.M. Hyde, B. Pakkenberg, Altered α-synuclein, parkin, and synphilin isoform
[104] M. Deschênes, B. Chabot, The emerging role of alternative splicing in senescence levels in multiple system atrophy brains, J. Neurochem. 136 (2016) 172–185.
and aging, Aging Cell 16 (2017) 918–933. [128] K. Beyer, M. Domingo-Sàbat, A. Ariza, Molecular pathology of lewy body diseases,
[105] P. D’Adamo, A. Menegon, C. Lo Nigro, M. Grasso, M. Gulisano, F. Tamanini, Int. J. Mol. Sci. 10 (2009) 724–745.
T. Bienvenu, A.K. Gedeon, B. Oostra, S.K. Wu, A. Tandon, F. Valtorta, W.E. Balch, [129] E.K. Tan, H. Shen, J.M.M. Tan, K.L. Lim, S. Fook-Chong, W.P. Hu, M.C. Paterson,
J. Chelly, D. Toniolo, Mutations in GDI1 are responsible for X-linked non-specific V.R. Chandran, K. Yew, C. Tan, Y. Yuen, R. Pavanni, M.C. Wong, K. Puvan,

775
S.H. Tan et al. Biomedicine & Pharmacotherapy 111 (2019) 765–777

Y. Zhao, Differential expression of splice variant and wild-type parkin in sporadic [157] E.M. Rocha, B. De Miranda, L.H. Sanders, Alpha-synuclein: pathology, mi-
Parkinson’s disease, Neurogenetics 6 (2005) 179–184. tochondrial dysfunction and neuroinflammation in Parkinson’s disease, Neurobiol.
[130] A. Eyal, R. Szargel, E. Avraham, E. Liani, J. Haskin, R. Rott, S. Engelender, Dis. 109 (2018) 249–257.
Synphilin-1A: an aggregation-prone isoform of synphilin-1 that causes neuronal [158] L. Hang, J. Thundyil, K.L. Lim, Mitochondrial dysfunction and Parkinson disease: a
death and is present in aggregates from alpha-synucleinopathy patients, Proc. Parkin–AMPK alliance in neuroprotection, Ann. N. Y. Acad. Sci. 1350 (2015)
Natl. Acad. Sci. U. S. A. 103 (2006) 5917–5922. 37–47.
[131] R. Marongiu, F. Brancati, A. Antonini, T. Ialongo, C. Ceccarini, O. Scarciolla, [159] A. Guerrero-Castilla, J. Olivero-Verbel, J. Marrugo-Negrete, Heavy metals in wild
A. Capalbo, R. Benti, G. Pezzoli, B. Dallapiccola, S. Goldwurm, E.M. Valente, house mice from coal-mining areas of Colombia and expression of genes related to
Whole gene deletion and splicing mutations expand the PINK1 genotypic spec- oxidative stress, DNA damage and exposure to metals, Mutat. Res. Genet. Toxicol.
trum, Hum. Mutat. 28 (2007) 98. Environ. Mutagen. 762 (2014) 24–29.
[132] L. Samaranch, O. Lorenzo-Betancor, J.M. Arbelo, I. Ferrer, E. Lorenzo, J. Irigoyen, [160] N. Matsuda, S. Sato, K. Shiba, K. Okatsu, K. Saisho, C.A. Gautier, Y.S. Sou, S. Saiki,
M.A. Pastor, C. Marrero, C. Isla, J. Herrera-Henriquez, P. Pastor, PINK1-linked S. Kawajiri, F. Sato, M. Kimura, M. Komatsu, N. Hattori, K. Tanaka, PINK1 sta-
parkinsonism is associated with Lewy body pathology, Brain 133 (2010) bilized by mitochondrial depolarization recruits Parkin to damaged mitochondria
1128–1142. and activates latent Parkin for mitophagy, J. Cell Biol. 189 (2010) 211–221.
[133] V. La Cognata, V. D’Agata, F. Cavalcanti, S. Cavallaro, Splicing: is there an al- [161] H. Shimura, N. Hattori, S.I. Kubo, Y. Mizuno, S. Asakawa, S. Minoshima,
ternative contribution to Parkinson’s disease? Neurogenetics 16 (2015) 245–263. N. Shimizu, K. Iwai, T. Chiba, K. Tanaka, T. Suzuki, Familial Parkinson disease
[134] P.J. Lockhart, S. Lincoln, M. Hulihan, J. Kachergus, K. Wilkes, G. Bisceglio, gene product, parkin, is a ubiquitin-protein ligase, Nat. Genet. 25 (2000) 302–305.
D.C. Mash, M.J. Farrer, DJ-1 mutations are a rare cause of recessively inherited [162] L. Silvestri, V. Caputo, E. Bellacchio, L. Atorino, B. Dallapiccola, E.M. Valente,
early onset parkinsonism mediated by loss of protein function, J. Med. Genet. 41 G. Casari, Mitochondrial import and enzymatic activity of PINK1 mutants asso-
(2004) e22. ciated to recessive parkinsonism, Hum. Mol. Genet. 14 (2005) 3477–3492.
[135] V. Bonifati, P. Rizzu, F. Squitieri, E. Krieger, N. Vanacore, J.C. van Swieten, [163] S. Sekine, R.J. Youle, PINK1 import regulation; a fine system to convey mi-
A. Brice, C.M. van Duijn, B. Oostra, G. Meco, P. Heutink, DJ-1 (PARK7), a novel tochondrial stress to the cytosol, BMC Biol. 16 (2018) 1–12.
gene for autosomal recessive, early onset parkinsonism, Neurol. Sci. 24 (2003) [164] C. Vives-Bauza, C. Zhou, Y. Huang, M. Cui, R.L.A. de Vries, J. Kim, J. May,
159–160. M.A. Tocilescu, W. Liu, H.S. Ko, J. Magrane, D.J. Moore, V.L. Dawson, R. Grailhe,
[136] J. Ugolino, S. Fang, C. Kubisch, M.J. Monteiro, Mutant Atp13a2 proteins involved T.M. Dawson, C. Li, K. Tieu, S. Przedborski, PINK1-dependent recruitment of
in parkinsonism are degraded by ER-associated degradation and sensitize cells to Parkin to mitochondria in mitophagy, Proc. Natl. Acad. Sci. 107 (2010) 378–383.
ER-stress induced cell death, Hum. Mol. Genet. 20 (2011) 3565–3577. [165] S.R. Yoshii, C. Kishi, N. Ishihara, N. Mizushima, Parkin mediates proteasome-de-
[137] C. Vilarino-Guell, A.I. Soto, S.J. Lincoln, S. Ben Yahmed, M. Kefi, M.G. Heckman, pendent protein degradation and rupture of the outer mitochondrial membrane, J.
M.M. Hulihan, H. Chai, N.N. Diehl, R. Amouri, A. Rajput, D.C. Mash, Biol. Chem. 286 (2011) 19630–19640.
D.W. Dickson, L.T. Middleton, R.A. Gibson, F. Hentati, M.J. Farrer, ATP13A2 [166] X. Wang, D. Winter, G. Ashrafi, J. Schlehe, Y.L. Wong, D. Selkoe, S. Rice, J. Steen,
variability in Parkinson disease, Hum. Mutat. 30 (2009) 406–410. M.J. Lavoie, L. Thomas, PINK1 and Parkin target miro for phosphorylation and
[138] C.-S. Lu, S.-C. Lai, R.-M. Wu, Y.-H. Weng, C.-L. Huang, R.-S. Chen, H.-C. Chang, Y.- degradation to arrest mitochondrial motility, Cell 147 (2011) 893–906.
H. Wu-Chou, T.-H. Yeh, PLA2G6 mutations in PARK14-linked young-onset par- [167] T.M. Dawson, V.L. Dawson, Molecular pathways of neurodegeneration in
kinsonism and sporadic Parkinson’s disease, Am. J. Med. Genet. B Neuropsychiatr. Parkinson’ s disease, Science 302 (2003) 819–822.
Genet. 159B (2012) 183–191. [168] S.K. Jha, N.K. Jha, D. Kumar, R.K. Ambasta, P. Kumar, Linking mitochondrial
[139] A. Di Fonzo, M.C.J. Dekker, P. Montagna, A. Baruzzi, E.H. Yonova, L. Correia dysfunction, metabolic syndrome and stress signaling in neurodegeneration,
Guedes, A. Szczerbinska, T. Zhao, L.O.M. Dubbel-Hulsman, C.H. Wouters, E. de Biochim. Biophys. Acta. Mol. Basis Dis. 1863 (2017) 1132–1146.
Graaff, W.J.G. Oyen, E.J. Simons, G.J. Breedveld, B.A. Oostra, M.W. Horstink, [169] J.K. Ngo, L.C.D. Pomatto, K.J.A. Davies, Upregulation of the mitochondrial Lon
V. Bonifati, FBXO7 mutations cause autosomal recessive, early-onset parkinso- Protease allows adaptation to acute oxidative stress but dysregulation is associated
nian-pyramidal syndrome, Neurology 72 (2009) 240–245. with chronic stress, disease, and aging, Redox Biol. 1 (2013) 258–264.
[140] H. Deng, H. Liang, J. Jankovic, F-Box only protein 7 gene in Parkinsonian-pyr- [170] N. Zahra, I. Kalim, M. Mahmood, N. Naeem, Perilous effects of heavy metals
amidal disease, Arch. Neurol. 70 (2013) 20–24. contamination on human health, J. Anal. Environ. Chem. 18 (2017) 1–17.
[141] A. Johri, M.F. Beal, Mitochondrial dysfunction in neurodegenerative diseases, J. [171] S. Bolisetty, E.A. Jaimes, Mitochondria and reactive oxygen species: physiology
Pharmacol. Exp. Ther. 342 (2012) 619–630. and pathophysiology, Int. J. Mol. Sci. 14 (2013) 6306–6344.
[142] P.I. Moreira, C. Carvalho, X. Zhu, M.A. Smith, G. Perry, Mitochondrial dysfunction [172] D.A. Patten, M. Germain, M.A. Kelly, R.S. Slack, Reactive oxygen species: stuck in
is a trigger of Alzheimer’s disease pathophysiology, Biochim. Biophys. Acta - Mol. the middle of neurodegeneration, J. Alzheimers Dis. 20 (2010) S357–67.
Basis Dis. 1802 (2010) 2–10. [173] A. Phaniendra, D.B. Jestadi, L. Periyasamy, Free radicals: properties, sources,
[143] A.H. Schapira, Mitochondrial dysfunction in neurodegenerative diseases, targets, and their implication in various diseases, Indian J. Clin. Biochem. 30
Neurochem. Res. 33 (2008) 2502–2509. (2015) 11–26.
[144] P.A. Trimmer, M.K. Borland, Differentiated Alzheimer’s disease transmitochon- [174] G.H. Kim, J.E. Kim, S.J. Rhie, S. Yoon, The role of oxidative stress in neurode-
drial cybrid cell lines exhibit reduced organelle movement, Antioxid. Redox generative diseases, Exp. Neurobiol. 24 (2015) 325–340.
Signal. 7 (2005) 1101–1109. [175] X. Li, P. Fang, W.Y. Yang, K. Chan, M. Lavallee, K. Xu, H. Wang, X. Yang,
[145] X. Wang, B. Su, G. Perry, M.A. Smith, X. Zhu, Insights into amyloid-β-induced Endothelial mitochondrial ROS, un-coupled from ATP synthesis, determine both
mitochondrial dysfunction in Alzheimer disease, Free Radic. Biol. Med. 43 (2007) physiological endothelial activation for recruitment of patrolling cells, and pa-
1569–1573. thological recruitment of inflammatory cells, Can. J. Physiol. Pharmacol. 95
[146] M. Lopez Salon, L. Pasquini, M. Besio Moreno, J.M. Pasquini, E. Soto, Relationship (2017) 247–252.
between β-amyloid degradation and the 26S proteasome in neural cells, Exp. [176] X. Li, P. Fang, J. Mai, E.T. Choi, H. Wang, X.F. Yang, Targeting mitochondrial
Neurol. 180 (2003) 131–143. reactive oxygen species as novel therapy for inflammatory diseases and cancers, J.
[147] S. Arun, L. Liu, G. Donmez, Mitochondrial biology and neurological diseases, Curr. Hematol. Oncol. 6 (2013) 19.
Neuropharmacol. 14 (2016) 143–154. [177] V.J. Thannickal, B.L. Fanburg, Reactive oxygen species in cell signaling, Am. J.
[148] J.P. Blass, R.K.-F. Sheu, G.E. Gibson, Inherent abnormalities in energy metabolism Physiol.-Lung C 279 (2000) L1005–L1028.
in Alzheimer disease: interaction with cerebrovascular compromise, Ann. N. Y. [178] J.G. Scandalios, Oxidative stress: molecular perception and transduction of signals
Acad. Sci. 903 (2000) 204–221. triggering antioxidant gene defenses, Braz. J. Med. Biol. Res. 38 (2005) 995–1014.
[149] C. Perier, M. Vila, Mitochondrial biology and Parkinson’s disease, Cold Spring [179] K.J. Davies, K.J.A. Davies, E. Percy, Oxidative stress, antioxidant defenses, and
Harb. Perspect. Med. 2 (2012) a009332. damage removal, repair, and replacement systems, IUBMB Life 50 (2000)
[150] D.C. Wallace, A mitochondrial paradigm of metabolic and degenerative diseases, 279–289.
aging, and cancer: a dawn of evolutionary medicine, Annu. Rev. Genet. 39 (2002) [180] K. Dasuri, L. Zhang, J.N. Keller, Oxidative stress, neurodegeneration, and the
359–407. balance of protein degradation and protein synthesis, Free Radic. Biol. Med. 62
[151] A. Eckert, K. Schmitt, J. Götz, Mitochondrial dysfunction - the beginning of the (2013) 170–185.
end in Alzheimer’s disease? Separate and synergistic modes of tau and amyloid- [181] S. Gandhi, A.Y. Abramov, Mechanism of oxidative stress in neurodegeneration,
toxicity, Alzheimers Res. Ther. 3 (2011) 15. Oxid. Med. Cell. Longev. 2012 (2012) 428010.
[152] A.D. Pilling, D. Horiuchi, C.M. Lively, W.M. Saxton, Kinesin-1 and Dynein are the [182] X. Wang, E.K. Michaelis, Selective neuronal vulnerability to oxidative stress in the
primary motors for fast transport of mitochondria in Drosophila motor axons, Mol. brain, Front. Aging Neurosci. 2 (2010) 1–13.
Biol. Cell 17 (2006) 2057–2068. [183] M.P. Mattson, Pathways towards and away from Alzheimer’s disease, Nature 430
[153] M.J. Calkins, P.H. Reddy, Amyloid beta impairs mitochondrial anterograde (2004) 631–639.
transport and degenerates synapses in Alzheimer’s disease neurons, Biochim. [184] Y. Zhao, B. Zhao, Oxidative stress and the pathogenesis of Alzheimer’s disease,
Biophys. Acta - Mol. Basis Dis. 1812 (2011) 507–513. Oxid. Med. Cell. Longev. 2013 (2013) 316523.
[154] R.H. Takahashi, C.G. Almeida, P.F. Kearney, F. Yu, M.T. Lin, T.A. Milner, [185] X. Wang, W. Wang, L. Li, P. George, L. Hyoung-gon, Z. Xiongwei, Oxidative stress
G.K. Gouras, Oligomerization of Alzheimer’s beta-amyloid within processes and and mitochondrial dysfunction in Alzheimer’s disease, Chin. J. Biochem. Biophys.
synapses of cultured neurons and brain, J. Neurosci. 24 (2004) 3592–3599. 1842 (2015) 1240–1247.
[155] R. Marongiu, B. Spencer, L. Crews, A. Adame, C. Patrick, M. Trejo, B. Dallapiccola, [186] E. Ito, K. Oka, R. Etcheberrigaray, T.J. Nelson, D.L. McPhie, B. Tofel-Grehl,
E.M. Valente, E. Masliah, Mutant Pink1 induces mitochondrial dysfunction in a G.E. Gibson, D.L. Alkon, Internal Ca2+ mobilization is altered in fibroblasts from
neuronal cell model of Parkinson’s disease by disturbing calcium flux, J. patients with Alzheimer disease, Proc. Natl. Acad. Sci. U. S. A. 91 (1994) 534–538.
Neurochem. 108 (2009) 1561–1574. [187] R. Brigelius-Flohe, M. Maiorino, Glutathione peroxidases, Biochim. Biophys. Acta
[156] V.M. Pozo Devoto, T.L. Falzone, Mitochondrial dynamics in Parkinson’s disease: a 1830 (2013) 3289–3303.
role for α-synuclein? Dis. Model. Mech. 10 (2017) 1075–1087. [188] V. Chauhan, A. Chauhan, Oxidative stress in Alzheimer’s disease, Pathophysiology

776
S.H. Tan et al. Biomedicine & Pharmacotherapy 111 (2019) 765–777

13 (2006) 195–208. [202] V. Dias, E. Junn, M.M. Mouradian, The role of oxidative stress in parkinson’s
[189] S. Manoharan, G.J. Guillemin, R.S. Abiramasundari, M.M. Essa, M. Akbar, disease, J. Parkinsons Dis. 3 (2013) 461–491.
M.D. Akbar, The role of reactive oxygen species in the pathogenesis of Alzheimer’s [203] H. Kumar, H.W. Lim, S.V. More, B.W. Kim, S. Koppula, I.S. Kim, D.K. Choi, The
disease, Parkinson’s disease, and Huntington’s disease: a mini review, Oxid. Med. role of free radicals in the aging brain and Parkinson’s disease: convergence and
Cell. Longev. 2016 (2016) 8590578. parallelism, Int. J. Mol. Sci. 13 (2012) 10478–10504.
[190] R.J. Mark, M. a Lovell, W.R. Markesbery, K. Uchida, M.P. Mattson, A role for 4- [204] T.G. Hastings, The role of dopamine oxidation in mitochondrial dysfunction: im-
hydroxynonenal, an aldehydic product of lipid peroxidation, in disruption of ion plications for Parkinson’s disease, J. Bioenerg. Biomembr. 41 (2009) 469–472.
homeostasis and neuronal death induced by amyloid beta-peptide, J. Neurochem. [205] J.M. Taylor, B.S. Main, P.J. Crack, Neuroinflammation and oxidative stress: co-
68 (1997) 255–264. conspirators in the pathology of Parkinson’s disease, Neurochem. Int. 62 (2013)
[191] J.N. Keller, Z. Pang, J.W. Geddes, J.G. Begley, a Germeyer, G. Waeg, M.P. Mattson, 803–819.
Impairment of glucose and glutamate transport and induction of mitochondrial [206] D. Hirayama, T. Iida, H. Nakase, The phagocytic function of macrophage-enforcing
oxidative stress and dysfunction in synaptosomes by amyloid beta-peptide: role of innate immunity and tissue homeostasis, Int. J. Mol. Sci. 19 (2017) pii: E92.
the lipid peroxidation product 4-hydroxynonenal, J. Neurochem. 69 (1997) [207] E.C. Hirsch, T. Breidert, E. Rousselet, S. Hunot, A. Hartmann, P.P. Michel, The role
273–284. of glial reaction and inflammation in Parkinson’s disease, Ann. N. Y. Acad. Sci. 991
[192] M.P. Mattson, S.L. Chan, Neuronal and glial calcium signaling in Alzheimer’s (2003) 214–228.
disease, Cell Calcium 34 (2003) 385–397. [208] A. Gerhard, N. Pavese, G. Hotton, F. Turkheimer, M. Es, A. Hammers, K. Eggert,
[193] L.M. Sayre, D.A. Zelasko, P.L. Harris, G. Perry, R.G. Salomon, M.A. Smith, 4- W. Oertel, R.B. Banati, D.J. Brooks, In vivo imaging of microglial activation with
Hydroxynonenal-derived advanced lipid peroxidation end products are increased [11C](R)-PK11195 PET in idiopathic Parkinson’s disease, Neurobiol. Dis. 21
in Alzheimer’s disease, J. Neurochem. 68 (1997) 2092–2097. (2006) 404–412.
[194] C.-X. Gong, K. Iqbal, Hyperphosphorylation of microtubule-associated protein tau: [209] M.L. Block, L. Zecca, J.-S. Hong, Microglia-mediated neurotoxicity: uncovering the
a promising therapeutic target for Alzheimer disease, Curr. Med. Chem. 15 (2008) molecular mechanisms, Nat. Rev. Neurosci. 8 (2007) 57–69.
2321–2328. [210] L. Yavich, Role of alpha-synuclein in presynaptic dopamine recruitment, J.
[195] G.M. Ashraf, N.H. Greig, T.A. Khan, I. Hassan, S. Tabrez, S. Shakil, I.A. Sheikh, Neurosci. 24 (2004) 11165–11170.
S.K. Zaidi, M. Akram, N.R. Jabir, C.K. Firoz, A. Naeem, I.M. Alhazza, [211] M. Kasten, C. Klein, The many faces of alpha-synuclein mutations, Mov. Disord. 28
G.A. Damanhouri, M.A. Kamal, Protein misfolding and aggregation in Alzheimer’s (2013) 697–701.
disease and type 2 diabetes mellitus, CNS Neurol. Disord. Drug Targets 13 (2014) [212] S. Xu, M. Zhou, S. Yu, Y. Cai, A. Zhang, K. Uéda, P. Chan, Oxidative stress induces
1280–1293. nuclear translocation of C-terminus of α-synuclein in dopaminergic cells, Biochem.
[196] T. Jung, B. Catalgol, T. Grune, The proteasomal system, Mol. Asp. Med. 30 (2009) Biophys. Res. Commun. 342 (2006) 330–335.
191–296. [213] M.G. Spillantini, R.A. Crowther, R. Jakes, M. Hasegawa, M. Goedert, Alpha-sy-
[197] L. Gregori, C. Fuchs, M.E. Figueiredo-Pereira, W.E. Van Nostrand, D. Goldgaber, nuclein in filamentous inclusions of lewy bodies from Parkinson’s disease and
Amyloid β-protein inhibits ubiquitin-dependent protein degradation in vitro, J. dementia with lewy bodies, Proc. Natl. Acad. Sci. U. S. A. 95 (1998) 6469–6473.
Biol. Chem. 270 (1995) 19702–19708. [214] E.V. Mosharov, Alpha-synuclein overexpression increases cytosolic catecholamine
[198] H.E. Moon, S.H. Paek, Mitochondrial dysfunction in Parkinson’s disease, Exp. concentration, J. Neurosci. 26 (2006) 9304–9311.
Neurobiol. 24 (2015) 103–116. [215] S.J. Tabrizi, M. Orth, J.M. Wilkinson, J.W. Taanman, T.T. Warner, J.M. Cooper,
[199] P. Jenner, C.W. Olanow, The pathogenesis of cell death in Parkinson’s disease, A.H. Schapira, Expression of mutant alpha-synuclein causes increased suscept-
Neurology 66 (2006) S24–36. ibility to dopamine toxicity, Hum. Mol. Genet. 9 (2000) 2683–2689.
[200] P. Jenner, J.W. Langston, Explaining ADAGIO: a critical review of the biological [216] M. Orth, S.J. Tabrizi, C. Tomlinson, K. Messmer, L.V.P. Korlipara, A.H.V. Schapira,
basis for the clinical effects of rasagiline, Mov. Disord. 26 (2011) 2316–2323. J.M. Cooper, G209A mutant alpha synuclein expression specifically enhances
[201] M.J. Kumar, J.K. Andersen, Perspectives on MAO-B in aging and neurological dopamine induced oxidative damage, Neurochem. Int. 45 (2004) 669–676.
disease: where do we go from here? Mol. Neurobiol. 30 (2004) 77–89.

777

You might also like