ArticlePDF AvailableLiterature Review

Mechanistic insights into the pathogenesis of neurodegenerative diseases: towards the development of effective therapy

Authors:
  • Conestoga College Ontario Canada

Abstract and Figures

Neurodegeneration is a prevalent and one of the emerging reasons for morbidity, mortality, and cognitive impairment in aging. Dementia is one of such conditions of neurodegeneration, partially manageable, irreversible, and worsens over time. This review is focused on biological and psychosocial risk factors associated with Alzheimer’s and Parkinson’s diseases, highlighting the value of cognitive decline. We further emphasized on current therapeutic strategies from pharmacological and non-pharmacological perspectives focusing on their effects on cognitive impairment, protein aggregation, tau pathology, and improving the quality of life. Deeper mechanistic insights into the multifactorial neurodegeneration could offer the design and development of promising diagnostic and therapeutic strategies.
Role of risk factors—Genetics and aging towards Parkinson’s pathogenesis. Gene PARK1/SNCA encodes the presynaptic and soluble unfolded protein, α- synuclein. In PD, due to mutations and overexpression of its wild type, these abnormally aggregate into filaments, amyloid fibrils called Lewy pathology, becomes insoluble and such toxic misfolded forms contribute to the neuronal death. Their accumulation leads to loss of dopamine neurons impairing synaptic dopamine release and consequently PD pathogenesis. Overexpression of α-synuclein inhibits autophagy and increases the accumulation of aggregate-prone proteins, hence a potent impetus for neurodegeneration. Mutation in the gene PARK2 encoding parkin protein, a ubiquitin E3 ligase inactivates, resulting in neurodegeneration via interference of adverse outcome pathway (AOP) for parkinsonian motor deficits through mitochondrial dysfunction. PARK6 resulting from mutations in PTEN Kinase 1 (PINK1), while PARK7 from mutations in the DJ-1 gene are linked with mitochondrial dysfunction and proteasome-stimulated apoptosis of neurons. Gene PARK8 due to mutations in LRRK2 enhances α-synuclein mobility, its accumulation in dopaminergic neurons and abnormal tau phosphorylation depositions in the brainstem. Progressive mitochondrial dysfunction is a key hallmark of the aging process due to the accumulation of mitochondrial DNA mutations and increased reactive oxygen species production causing oxidative damage, thereby leading to perturbed respiratory chain activity leading to neurodegeneration
… 
This content is subject to copyright. Terms and conditions apply.
Vol.:(0123456789)
1 3
Molecular and Cellular Biochemistry
https://doi.org/10.1007/s11010-021-04120-6
Mechanistic insights intothepathogenesis ofneurodegenerative
diseases: towardsthedevelopment ofeffective therapy
FauziaNazam1· SibhghatullaShaikh2· NaziaNazam3· AbdulazizSaadAlshahrani4· GulamMustafaHasan5· Md.
ImtaiyazHassan6
Received: 19 December 2020 / Accepted: 23 February 2021
© The Author(s), under exclusive licence to Springer Science+Business Media, LLC, part of Springer Nature 2021
Abstract
Neurodegeneration is a prevalent and one of the emerging reasons for morbidity, mortality, and cognitive impairment in
aging. Dementia is one of such conditions of neurodegeneration, partially manageable, irreversible, and worsens over time.
This review is focused on biological and psychosocial risk factors associated with Alzheimer’s and Parkinson’s diseases,
highlighting the value of cognitive decline. We further emphasized on current therapeutic strategies from pharmacological
and non-pharmacological perspectives focusing on their effects on cognitive impairment, protein aggregation, tau pathol-
ogy, and improving the quality of life. Deeper mechanistic insights into the multifactorial neurodegeneration could offer the
design and development of promising diagnostic and therapeutic strategies.
Keywords Alzheimer’s disease; Parkinson’s disease· Therapeutic management· Neurodegenerative diseases·
Neurodegeneration
Introduction
Dementia is a chronic disorder of cognitive decline charac-
terized by deterioration of memory, thinking, behavior, and
normal physical activities [1]. It affects learning, memory,
comprehension, calculation, language, reasoning, and atten-
tion. Yet consciousness remains unaffected and the most
affected ones are elder people [2]. Undoubtedly, with the
progressive aging population of modern society, age-related
disorders have become predominant. Compelling evidence
of the aging population indicates that the number of demen-
tia cases would increase rapidly to over 150 million by 2050,
creating a major public health and social challenge (https ://
www.who.int/news-room/fact-sheet s/detai l/demen tia).
Neurodegenerative diseases (ND) comprise a range of
disorders mainly affecting the neurons in the human brain
[3, 4]. Alzheimer’s disease (AD) is the most familiar basis
of dementia, responsible for an estimated 60–80% of total
cases. Parkinson’s disease (PD) remains the second most
prevalent ND affecting over 6.1 million people [5]. Although
PD is principally a motor disorder, it is consistently associ-
ated with dementia. A significantly growing concern linked
with social and economic burden is imposed particularly by
these ND and their contribution to dementia.
Fauzia Nazam and Sibhghatulla Shaikh have contributed equally
to this work.
* Nazia Nazam
nazianazam@gmail.com
* Md.Imtaiyaz Hassan
mihassan@jmi.ac.in
1 Section ofPsychology, Women’s College, Aligarh Muslim
University, Aligarh, UP202002, India
2 Department ofMedical Biotechnology, Yeungnam
University, Gyeongsan38541, RepublicofKorea
3 Amity Institute ofMolecular Medicine andStem Cell
Research, Amity University, Noida, UttarPradesh201313,
India
4 Department ofMedicine, Najran University, P.O. Box- 1988,
Najran, SaudiArabia
5 Department ofBiochemistry, College ofMedicine,
Prince Sattam Bin Abdulaziz University, P.O. Box173,
Al-Kharj11942, KingdomofSaudiArabia
6 Centre forInterdisciplinary Research inBasic Sciences,
Jamia Millia Islamia, NewDelhi110025, India
Molecular and Cellular Biochemistry
1 3
Tremendous efforts have been employed for the devel-
opment of effective treatments but still, no cure can fully
control, delay, or regress the conditions leading to dementia
[6, 7]. Elucidating the role played by crucial risk factors
in the progression of dementia is an attempt to restrain the
commencement by controlling various risk factors [8, 9].
Thus, identifying probable risk attributes could facilitate in
dropping off the burden in dementia patients.
Thus, the biological risk factors could be regarded as
attributes confirming the heterogeneity of AD or PD as well
as contributing to the etiology and pathogenesis with their
particular mechanisms. The psychosocial risk factors on
the contrary refer to aspects that may affect an individual’s
social and/or psychological response/traits. These biological
and psychosocial factors could be regarded as non-modifia-
ble and modifiable risks, respectively (Fig.1).
In the present review, we intend to elaborate on risk fac-
tors from biological, psychological, and social perspectives.
We further emphasized their management could reduce the
development and progression of NDs.
Non‑modiable biological risk factors
The biological risk factors can enhance the propensity of
dementia arising out of these neurodegenerative diseases and
are appropriately called non-modifiable. The major unmodi-
fiable factors are aging, genetics, and gender. The interplay
of these factors with the psychosocial attributes contributes
to the risk of AD and PD.
Aging
Older age though is not the cause, but it is the greatest
biological risk factor for AD and PD and linked with
dementia. Undoubtedly, with aging, the incidence of
dementia exponentially increases above 65years of age
(5). Although the majority of people with AD are aged 65
and beyond, yet younger individuals may also develop the
disease [10]. Dementia in persons younger than 65years,
arising irrespective of any neurodegenerative disease is
referred to as ‘‘younger-onset’’ or ‘‘early-onset’’ demen-
tia/AD/PD. Although advancing age is one of the greatest
risk factors for AD as well as a challenge to preserve the
cognitive function during the process of aging, it is not
a normal part of aging. With studies suggesting that loss
of neurons is an integral part of the aging brain, it has
become even more challenging. However, a vital ques-
tion is “why aging in some people proceed with normal
cognitive functions while in others there is a deterioration
and disease development?” Perhaps the answer lies in AD
being a multifactorial disease rather than a single cause.
Like AD, PD has the potential to influence well-being
and socio-economic status, and reveal escalating tenden-
cies together with the population longevity. Aging contin-
ues the main risk factor for developing idiopathic PD. In
the population with the age of over 60years, PD affects
more than 1 in 100 individuals, while this 1% rate shows
a whopping 5% prevalence in individuals above 85years
[11]. This underscores the impact of advancing age on
the risk of PD development. Commonly thought of as a
disease of the elderly, yet a small 5% of total cases show
symptoms below 60years of age too of which the majority
is attributed to gene mutations affecting protein metabo-
lism or mitochondrial dysfunction [12]. These hereditary
forms of PD with early-onset may show similar or different
phenotypes or neuropathology compared to sporadic PD.
Fig. 1 Classification of risk
factors for AD and PD. Crucial
risk factors in the development
and progression of dementia can
be “Biological” and “Psycho-
social” risk factors contributing
to pathogenesis, culminating in
dementia. The major biologi-
cal factors are aging, genetics,
and gender, while innumerable
psychosocial risk factors associ-
ated could be broadly catego-
rized into comorbidities and life
events, and lifestyle and mental
health. The interplay of biologi-
cal factors with psychosocial
factors contributes to the risk of
AD and PD
Molecular and Cellular Biochemistry
1 3
Genetics
Genetic is one of the major risk factors that contribute
about 70% risk of AD development [13]. However, genetic
research in PD confirms the involvement of genetic risk
factors, including monogenic forms in 30% of the familial
and 3% to 5% of the sporadic cases [12]. To begin with, we
predominantly focus on the risk genes for AD and PD identi-
fied from genome-wide association studies. This progress in
genetic research has immensely improved our understanding
of pathogenesis and the molecular mechanism of dementia.
The presence of Lewy bodies (LB) is an important fea-
ture of PD [14]. LB comprised ubiquitinated α-synuclein,
synaptic vesicle protein, and parkin. These LBs induce
dysfunction of mitochondria, free radicals release, JNK
pathway-stimulated apoptosis, and microglia-triggered
inflammation in the brain [15]. The main genes reported for
causing PD are α-synuclein (SNCA), parkin, leucine-rich
repeat kinase 2 (LRRK2), PTEN-induced putative kinase
1 (PINK1), and DJ-1 [16]. Among 6 genes with a clear
association with heritable and single gene PD, mutations in
SNCA (PARK1 = 4) and LRRK2 (PARK8) are accounted for
autosomal-dominant PD forms, while mutations in Parkin
(PARK2), PINK1 (PARK6), DJ-1 (PARK7), and ATP13A2
(PARK9) are responsible for autosomal recessive PD forms
[17]. Figure2 illustrates the role of genetics and aging in the
development and progression of AD.
The first “PD gene- PARK1/SNCA” was the gene that
encodes the α-synuclein (presynaptic protein). PARK1-asso-
ciated PD is frequently of early commencement and gener-
ally advances promptly. PARK1/SNCA is the key element
in LB which is mostly phosphorylated at Ser129 leading to
its fibril uptake by neurons and intensifies the development
of PD [14]. α-synuclein restrains BDNF/TrkB signaling
pathway by binding to TrkB receptors, resulting in dopa-
minergic death of neurons [18]. Farrer and colleagues [19]
found a double enhancement of α-synuclein at both mRNA
and protein levels in PD; hence, decreasing the expression
of α-synuclein presents a possible curative approach. The
missense mutation in α-synuclein (A53T) was recognized as
a cause of familial PD [20]. A53T has been seen to inhibit
autophagy in the brain of animal models and lead to synucle-
inopathy [21]. A53T mutations along with A30P and E46K
in the α-synuclein gene cause rare familial PD forms, play-
ing a significant role in the neurodegeneration [22].
Mutation in the second “PD gene- PARK2” and gene
encoding parkin protein, result in an autosomal recessive
form of PD [23]. Parkin is a ubiquitin E3 ligase that mono
and/or polyubiquitinates proteins regulating numerous cel-
lular processes. Loss of parkin’s E3 ligase activity exhibits
a pathogenic role in both inherited and sporadic PD. The
third “PD gene- PARK7” results from mutations in the DJ-1
gene [24]. Inline, the fourth “PD gene PARK6” results from
mutations in PTEN kinase 1 (PINK1). A mitochondrial pro-
tein kinase (PTEN1) is linked with the PD pathogenesis. The
deficiency of PINK1 leads to an autosomal recessive form
of PD. Wild-type PINK1 is important for neuroprotection
against the dysfunction of mitochondria and proteasome-
stimulated apoptosis. However, the mutation in G309D
damages this protective effect, probably by interfering with
adenosine diphosphate binding, and thus reduces the activ-
ity of kinase [25]. PINK1 loss was linked with mitochon-
drial dysfunction causing deficient LETM1 phosphorylation
and damage to the mitochondrial Ca2+ transport [26]. In
addition, an interrupted mitochondrial membrane potential
under stressful conditions has been found in PINK1 mutant-
transfected cells.
The fifth “PD gene-PARK8” along with LRRK2 muta-
tions result in the highest risk of familial PD and causes
the autosomal dominant PD [27]. α-synuclein mobility and
its accumulation are enhanced by LRRK2 G2019S muta-
tion in cultured primary neurons and dopaminergic neu-
rons in substantia nigra pars compacta of PD brain [28].
Tau transmission in mouse brain neurons is promoted by
LRRK2 G2019S mutation, suggesting a need to understand
the tau protein progression and neuropathology in LRRK2-
associated PD [29]. The kinase activities of LRRK2 were
enhanced by mutations in LRRK2; thus, LRRK2 inhibitors
could be employed to block the activity of kinase to contrib-
ute neuroprotection in PD models [30]. The identification
of mutations in the above five PD genes provided deeper
insights into the mechanism of PD pathogenesis.
Amyloid precursor protein (APP) mutation or either of
the presenilin gene mutation (PSEN1 or PSEN2) accounts
for the early-onset AD [31]. The important genetic risk fac-
tor for the progression of late-onset AD is the E4 allele of
apolipoprotein E-APOE4 which is involved in fat metabo-
lism. Besides, CR1, CLU, and PICALM genes have been
recognized as risk factors for early-onset of the AD. In the
brain, APOE is the principal apolipoprotein of the high-
density lipoprotein complex. Pathologically, the E4 allele
of APOE is linked with enhanced deposition of Aβ peptide
in the brain. Although APOE has various functions in the
brain, its role in AD progression relates to its ability to bind
with the Aβ peptide [32].
APP is an integral transmembrane protein expressed in
the brain and metabolized by a series of sequential proteases
together with the γ-secretase complex. Sequential APP pro-
teolysis generates neurotoxic Aβ peptide and its aggregation
within the brain imparts a vital step in AD development
[33]. The common APP mutation (APPswe) changes the
amino acid adjacent to the BACE1 cleavage site, while the
mutations aggregate nearby the γ-secretase cleavage site in
NDs. Mutations in the presenilin (PS1 and PS2) result in
increased Aβ42 –the less soluble toxic contrasted to Aβ40
production [34].
Molecular and Cellular Biochemistry
1 3
Processing of APP by β-secretase pursued by γ-secretase
resulting in the Aβ generation is termed as an amyloido-
genic pathway. The β-secretase first cuts the APP to pro-
duce the C-terminal fragment known as C99 subsequently
γ-secretase cleaves C99 to generate the Aβ peptides [33].
The α-secretase cuts APP in the center of the Aβ domain
and subsequently prevents its production. The member of
a disintegrin and metalloprotease domain (ADAM) family,
such as ADAM9, ADAM10, and ADAM17, has shown to
exhibit α-secretase activity, though the mechanism that con-
trols the activity of α-secretase is still uncertain. However,
it has been established that ADAM10 is the key α-secretase
and cleaves the APP in a non-amyloidogenic way [35]. As
the β- and γ-secretases are involved in Aβ generation, inhibi-
tion of these enzymes have been considered as the potential
target to control cerebral Aβ levels in AD. On the other hand,
α-secretase activation makes a possible therapeutic approach
in the reduction of Aβ [36].
Fig. 2 Role of risk factors—Genetics and Aging towards AD patho-
genesis. Cleavage of transmembrane amyloid precursor protein
(APP) by α-secretase results in the production of peptides required
for normal neuronal function (non-amyloidogenic pathway). How-
ever, the abnormal cleaving by β-secretase initially generates a mem-
brane-associated fragment and later γ-secretase releases neurotoxic
amyloid-beta (Aβ) peptide (amyloidogenic pathway). Aβ accumula-
tion in the extracellular space is influenced by the Aβ-binding mol-
ecules—APOE which tends to aggregate, giving rise to senile plaques
and other insoluble oligomeric protein forms. Aβ-induced elevated
GSK3β activity causes tau protein phosphorylation producing
increased levels of neurofibrillary tangles (NFT). Once Aβ aggregates
into oligomers and fibrils, it can induce cellular toxicity, neuronal
inflammation, and dysfunction, neurotrophic activity leading to AD
pathogenesis. Neurotoxic Aβ accumulation is the result of impaired
Aβ homeostasis, i.e., between its production and clearance. Defective
Aβ clearance contributes to pathological accumulations of cerebral
Aβ. Aging-related increased aggregation of Aβ in the brain function
as a link between aging and AD. It induces tissue inflammation and
organ dysfunction, together with being vital components of the amy-
loidogenic pathway
Molecular and Cellular Biochemistry
1 3
Tau (microtubule-associated protein) is the principal
component in neurofibrillary tangles (NFTs), a major
player of neurodegeneration. It plays important roles in
the assembly and stabilization of microtubules, and helps
in linking the polymers with other cytoskeletal filaments
[3741]. Hyperphosphorylated tau produces the NFTs in
AD [42]. Mutations in the tau gene causes frontotemporal
dementia FTDP-17 which impairs its binding to micro-
tubules moderately and subsequently induces neurode-
generation [43]. Tau causes induction of Aβ-stimulated
toxicity and hence shows a useful approach for treating
AD and related disorders [40, 4446]. Proteolytic cleavage
of tau is appearing to be a potential marker in AD devel-
opment. Tau has several sites for proteolytic cleavage,
which subsequently result in the breakdown and release
of toxic fragments. For instance, tau cleavage by caspases
at Asp421 or Asp13 at its C-terminus produces a truncated
protein that is susceptible to the aggregation and conse-
quent progression of AD [47].
As the studies on AD and PD validate, however, the
genetic mapping and gene-isolation tools by the Human
Genome Project have aided in the gene identification to
unravel their role in the multifaceted forms of NDs. The
emergence of a consensus hypothesis that aggregates of
amyloid-beta and NFT, and α-synuclein are linked with
neurotoxicity in AD and PD respectively could explain the
pathogenesis of the hereditary forms and idiopathic variety
(Figs.2 and 3). Such insights into underlying mechanism of
Fig. 3 Role of risk factors—Genetics and aging towards Parkinson’s
pathogenesis. Gene PARK1/SNCA encodes the presynaptic and
soluble unfolded protein, α- synuclein. In PD, due to mutations and
overexpression of its wild type, these abnormally aggregate into fila-
ments, amyloid fibrils called Lewy pathology, becomes insoluble and
such toxic misfolded forms contribute to the neuronal death. Their
accumulation leads to loss of dopamine neurons impairing synaptic
dopamine release and consequently PD pathogenesis. Overexpression
of α-synuclein inhibits autophagy and increases the accumulation of
aggregate-prone proteins, hence a potent impetus for neurodegenera-
tion. Mutation in the gene PARK2 encoding parkin protein, a ubiq-
uitin E3 ligase inactivates, resulting in neurodegeneration via inter-
ference of adverse outcome pathway (AOP) for parkinsonian motor
deficits through mitochondrial dysfunction. PARK6 resulting from
mutations in PTEN Kinase 1 (PINK1), while PARK7 from mutations
in the DJ-1 gene are linked with mitochondrial dysfunction and pro-
teasome-stimulated apoptosis of neurons. Gene PARK8 due to muta-
tions in LRRK2 enhances α-synuclein mobility, its accumulation in
dopaminergic neurons and abnormal tau phosphorylation depositions
in the brainstem. Progressive mitochondrial dysfunction is a key hall-
mark of the aging process due to the accumulation of mitochondrial
DNA mutations and increased reactive oxygen species production
causing oxidative damage, thereby leading to perturbed respiratory
chain activity leading to neurodegeneration
Molecular and Cellular Biochemistry
1 3
pathogenesis could benefit us to identify novel drug targets
for these diseases.
Gender
The study on the association between gender differences in
AD or PD is still in its infancy. The convergence of non-
modifiable attributes results in creating distinctive sex dif-
ferences risk profiles for AD. The sex-based prevalence in
AD patients is confirmed to be more than 60% for females
itself [48]. APOE4 risk in AD female sufferers is higher
in stark contrast to males, thereby apparent in women het-
erozygous for the APOE4 allele. Surprisingly, males with
homozygosity for APOE4 are seen to be a greater risk in
terms of AD as well as slight cognitive damage. Gender dif-
ferences in PD are evident from recent findings. Incidence
and prevalence data from epidemiological studies in PD con-
firm these figures to be approximately 1.5–twofold more in
men contrasted to women [49]. The delayed symptomatic
development in PD is seen in women on account of greater
physiological striatal dopamine levels, corroborating the epi-
demiological interpretations of poor incidence and higher
onset age in females [50].
The three major biological risk factors discussed above
are largely studied, yet they fail to explain a definite or
specific biological cause of ND. Rightly kept under the
non-modifiable category, the biological factors pose diffi-
culty to intervene. There is a need to focus on developing
interventions that enhance the overall well-being of elderly
people to physical and psycho-social health elements [51].
Since the risk factors could be cause and effect, their pre-
clinical intervention holds the potential to prevent the onset
and development of NDs. Several psychosocial aspects are
unexplored; thus, more attention needs to study the psycho-
social concerns which could be a novel approach towards the
therapeutic management of NDs.
Modiable psychosocial risk factors
Modifiable factors play a vital role in the development of
AD or PD. Numerous psychosocial risk factors are associ-
ated with the onset of dementia which could be discussed
broadly under categories as follows: (i) Comorbidities and
life events, and (ii) Lifestyle and mental health.
Comorbidities andlife events
Preventing the modifiable risk factors might serve as an
alternative approach to fight these debilitating NDs.
Cardiovascular diseases
Acquired factors such as vascular diseases, diabetes, high
blood pressure, and hyperlipidemia are the risk of AD pro-
gression [52]. Neuropathological studies show a massive 6%
to 47% of individuals with dementia to have coexistence of
cerebrovascular disease pathologies [53]. Cardiovascular
disease and AD have many common risk factors. Aβ and
amyloid accumulation in pial and intracerebral arteries lead
to cerebral amyloid angiopathy (CAA) in more than 80% AD
cases [54]. In AD patients with established cerebral amyloid
angiopathy in small arteries and arterioles, atrophy of vascu-
lar smooth muscle cells layer causes the vessel wall to rup-
ture and bleeding within the cerebra in these patients [55].
On the other hand, in PD patients, cardiovascular comor-
bidities are associated with the degeneration of neurons in
the elderly and this risk factor contributes to axial motor
features in PD [56].
Type 2 diabetes
The link between type 2 diabetes and enhanced AD risk has
illustrated in several epidemiological studies suggesting that
Aβ and tau protein phosphorylation are linked with insu-
lin resistance or deficiency and results in the onset of AD
development [57]. Insulin resistance or deficiency increases
the action of β- and γ-secretases and subsequently reduces
amyloid-beta clearance, leading to its deposition in brain tis-
sue. Besides, insulin resistance or deficiency stimulates the
tau protein hyperphosphorylation, resulting in the generation
the neurofibrillary tangles [58]. A contrasting association is
seen when a comparable evaluation is made between dia-
betes and the risk of PD. Meta-analysis from case–control
studies covering 14 studies, including 21,395 PD patients
and 84,579 control subjects, suggests that diabetes may exert
a lower risk despite significant heterogeneity [59].
Hypertension
Hypertension plays an important role in stroke, post-stroke
dementia, and the pathogenesis of vascular dementia. In a
longitudinal study, hypertension was found to increase the
risk of AD progression [60]. Hypertension, particularly
when exists in middle age, adversely affects cognitive func-
tion. Hypertension causes an alteration in vascular walls
which results in ischemia and cerebral hypoxia, thereby
prompting AD development. Cerebral ischemia stimulates
the presenilin expression, able to accumulate the APP and
Aβ and also caused the blood–brain barrier dysfunction [61].
Enhanced cholesterol levels have been considered as a
risk factor for AD development. About 10% higher cho-
lesterol level was found in the AD patients relative to the
healthy individuals [62]. Similarly, in PD, enhanced total
Molecular and Cellular Biochemistry
1 3
cholesterol at baseline was linked with an enhanced risk
[63]. It is interesting to note that cholesterol-lowering
agents reduce α-synuclein level, while hypercholesterolemia
increases the risk of PD.
Environmental exposure
Exposure to pesticides is associated with the risk of PD and
the person with impaired metabolism of pesticides might
be more susceptible [64]. However, pesticide categories are
broad and comprise chemicals having diverse mechanism.
Only some reports have described specific chemicals or
classes of chemicals, such as organochlorines, insecticides,
wood preservatives, herbicides, dieldrin, and paraquat. Liv-
ing in rural areas, agriculture, and drinking water are con-
tributors to cause of PD [65].
Mood
Mood, particularly depression, is a moderate risk factor for
NDs. Studies have identified the chemical abnormality may
be responsible for a depressive mood and dementia. Folate
deficiency or low folic acid was found in depressed mood
and cognitive impairment in AD [66]. Although folic acid
plays a crucial role across all the life-span developmental
stages (neonate, infant, children, and adolescence), but in the
elderly population, its deficiency is reported in old age brain
which increases vulnerability to AD and vascular dementia.
In elderly Latin American, low folate status was associated
with dementia after control for confusing by the impact of
vitamin B-12, depressive symptoms, and other demographic
variables, such as age, sex, and some years of education.
Recent studies have explained the clinical mechanism of
increased folic acid intake and deterioration in symptoms
that reduces neuroinflammatory disturbances which have an
essential role in AD [67]. However, it is too early to con-
clude that a greater intake of folic acid supplements should
be encouraged in the elderly to reduce the risk of cognitive
impairment.
Social disengagement
The exponential increase in social disengagement among the
geriatric population is noteworthy [68]. Social engagement
is the most crucial social factor correlated with dementia,
and a lower social engagement generally leads to greater
chances of disease development [69]. Strikingly, high or
medium social engagement is associated with a lower risk
of dementia compared to low social engagement [70]. An
average hippocampus volume and high risk of dementia in
Japanese–American men at late-life between the ages of
71–92years are assessed by Hazard Ratio. The value of the
Hazard Ratio was found critically significant in later life
with low social engagement [71]. Thus, it could be inferred
that lower social engagement later in life could be linked to
the risk of dementia.
Lifestyle andmental health
Epidemiological studies fortify that various lifestyle factors
are accountable for varying degrees of Alzheimer’s and PD
risk [72]. Lifestyle interventions could improve or maintain
cognitive function in the absence of medications. Also, other
factors that could be modulated socially are diet, physical
activity and mental health, exposure to stress, and sleep
patterns, which are so far considered a major concern to
control cognitive decline and dementia [73]. Interference
in the insulinergic function by stress, inactive lifestyle, and
overconsumption could lower the peptide level and their
cognate receptors in the brain [74].
Diet
Diet is a leading risk factor in PD [74]. For instance, con-
sumption of lipids and fat-rich foods increases oxidative
stress which is associated with PD and AD risk [75]. Dairy
consumption is also associated with the increased risk of
PD, as shown in large prospective cohort studies [76]. Con-
sumption of milk is linked with PD risk as it might contain
organochlorine pesticides and tetrahydroisoquinoline which
can bypass the blood–brain barrier and stimulate Parkinson-
ism. Of the possible PD risk-reducing dietary components,
caffeine, nicotine, and antioxidants are the most examined
ones. Meta-analyses predict PD risk to 30 and 25% lower in
coffee drinkers and caffeine consumers, respectively, indica-
tive of neuroprotection linked with caffeine only and not
to another coffee constituent [77]. On the contrary, good
health and a salutary lifestyle are major factors correlated
with a better cognition and brain structure. Data supporting
a healthy diet as a preventive or delaying factor for AD are
largely based on cardio-metabolic status which maintains
a healthy tissue vasculature and intact insulin sensitivity,
indicates the protective actions of a balanced diet on brain
function.
Personality traits
The available literature indicates an association of personal-
ity traits and clinical markers of dementia. A positive cor-
relation between an increased neuroticism trait disposition
to experience negative effects, low extraversion, high social
introversion, and decreased conscientiousness [78]. Other
personality traits are associated with AD are paranoid, schiz-
oid, schizotypal, borderline, and narcissistic traits [79]. On
similar lines, PD is also accompanied by a characteristic par-
kinsonian personality, characterized by conscientiousness,
Molecular and Cellular Biochemistry
1 3
punctuality, industriousness, and reduced novelty seeking,
in stark contrast to the elderly in the pink [80].
Education
Twin studies on the Swedish population examining the role
of education in the prognosis of AD and dementia demon-
strate a lower education as a risk factor for Alzheimer’s but
not for dementia [81]. However, other researchers have ques-
tioned the latent factor that could plausibly link low educa-
tion and dementia. An empirical investigation was carried
out in people with 5years or less, 6–8years, and those with
9years of formal education. The vulnerability to dementia
calculated by Odd ratio was found significantly higher in
the first group. The reason attributed is that persons with
higher education have more mental occupations and cogni-
tive reservation, thereby delaying the clinical manifestation
of AD and dementia [82]. Like AD, the higher educational
level, conversely, was allied with a lower risk of cognitive
decline. The hypotheses that education predicts the socio-
economic condition of the person and education improves
the reservation of the brain and increase cognitive capacity
proven and accepted. A school study has even shown gender-
specific findings, that is, risk of dementia is increased in
sisters being poorly educated and who did not receive any
vocational training [83]. It has been reported that below the
age of 80years the incidence is higher in males compared
to females but above this age, the incidence is high in males,
which shows an interaction effect on dementia of age and
gender.
Depression
Depression is directly correlated with poor cognition and
functional status in AD patients. A history of depression
sensitizes for enhanced risk of AD development [1, 69]. A
positive correlation between depression and PD develop-
ment has also been reported [84]. Depression could be a
warning for NDs. Hence, more research towards bringing out
the association between AD and depression hold clinical rel-
evance. Altogether, it can be inferred that psychosocial risk
factors sensitize the person to develop the disease towards
the risk imposed by biological factors. Therefore, the iden-
tification of potential risk factors may facilitate a reduction
in the burden of people afflicted by dementia. Preventing
the risk factors might not possibly influence completely the
disease development in the future. However, establishing the
association of modifiable factors with disease progression is
difficult to determine, primarily due to the absence of objec-
tive and sensitive markers.
Therapeutic andmanagement strategies
Both AD and PD equally demand better healthcare in addi-
tion to institutionalization costs [9]. In the subsequent sec-
tion, we explore how prevention and control over these
risk factors could impact ND development. To meet these
demands, numerous strategies have been developed [85, 86].
These options could broadly be categorized as pharmaco-
logical and psychosocial (non-pharmacological) approaches
(Fig.4).
Fig. 4 Potential preventive
strategies for dementia develop-
ment. Numerous strategies for
cognitive impairment treatment
arising out of NDs could be
pharmacological and psychoso-
cial approaches. Their combina-
tions have proven more effective
in retarding disease progression
Molecular and Cellular Biochemistry
1 3
Pharmacological approaches
Improving cognition and slowing the advancing symptoms
is the foremost aim in drug development for both condi-
tions. At present, various drugs received FDA approval since
they are a step ahead in the path of better treatment and due
course may lead to a cure.
Cholinesterase inhibitors (ChEIs)
Several pieces of evidence from neuropathological and
imaging studies have shown substantial cholinergic deficits
in NDs [87]. Deficits in cholinergic activity are associated
with cognitive deficits in either of the conditions which may
be addressed by the design and development of cholinester-
ase inhibitors (ChEIs). These inhibitors are considered safe
and effective in AD and PD patients. ChEI, such as Riv-
astigmine, is successful in cognitive improvement by inhib-
iting cortical AChE and butyrylcholinesterase (BuChE) in
affected brain regions—hippocampus and amygdala [88].
While rivastigmine and galantamine are efficacious in AD
patients with mild to moderate symptoms, donepezil benefits
are extended in alleviating severe AD symptoms [89].
In contrast to AD, a lesser number of clinical trials of
ChEI in PD patients are reported. Yet, studies focused on
ChEIs for PD treatment suggest an improved cognitive func-
tion [90]. Tacrine was the first studied ChEI in Parkinson’s
patients with significant improvements in cognition. It bears
dual inhibition of AChE and BuChE along with modula-
tion effect on the nicotinic receptor. In a placebo-controlled
study, rivastigmine adequately improved dementia con-
nected with PD but was associated with few side effects. The
efficacy together with the safety of AChI-donepezil hydro-
chloride in PD dementia was studied and found that done-
pezil could improve cognition and executive function [91].
N‑Methyl D‑Aspartate (NMDA) receptor antagonist
Although cholinergic drugs increase existing levels of acetyl-
choline in surviving brain cells, they are not successful in pre-
venting neuronal death as well as disease progression. Hence,
evaluating the potential AD and PD treatments and improv-
ing clinical management via different mechanisms is essential.
Substantial evidence favors the role of disrupted glutamate
in the pathophysiology of neurodegenerative disorders [92].
Various glutamate-gated channels, in addition to a group of
G-Protein Coupled Receptors (GPCRs), play crucial roles in
synaptic plasticity and the cellular processes underlying learn-
ing and memory. Among them, glutamate NMDA receptor
antagonists emerged as attractive therapeutic targets [93].
Memantine is one of such antagonists that modulate the flow
of glutamatergic neuronal transmission relying on glutamate
as the main excitatory neurotransmitter [94]. During normal
physiological functions, memantine ineffectively blocks the
low receptor activity levels. While enhanced glutamate con-
centrations is associated with increased activation of NMDA
receptors. Hence, blocks the lethal effects of overactive glu-
tamatergic activity such as compromised synaptic plasticity
and damage to neuron. Memantine—an accepted drug is
focused on the symptomatic treatment of adequate to severe
AD. Although insufficient evidence of the memantine efficacy
exists in PD, yet it is regarded as safe and well tolerated in such
patients [95]. Notably, memantine being is widely accepted in
both AD and PD, its combination with other therapies could
be an effective alternative.
Other treatment options
Altogether, the benefits out of symptomatic treatments
employed for Alzheimer’s patients are modest and do not
alter natural disease progression. Yet, disease-modifying
approaches are under the exploration of which many tar-
get the amyloid cascade. These agents are either aimed at
decreasing Aβ production, reducing aggregation, or enhanc-
ing its clearance. Inhibiting phosphorylation of tau protein
is also promising and is under exploration in animal models
as well as pilot human studies for AD patients. Similarly,
for PD, where the present pharmacological approaches
are mainly symptomatic rather than preventive and with
no cure to date, various dopaminergic (DAergic) drugs or
therapeutic candidates based on the novel mechanisms are
employed [96]. Levodopa and other DAergic drugs are effec-
tive pharmacological options for PD patients (141). DAergic
drugs, when combined with an inhibitor against monoam-
ine oxidase type B, catechol-O-methyl transferase, and cho-
linergic, hold potential for improved motor control or also
improve levodopa-mediated motor complications. Besides,
non-motor symptoms of PD, such as the cognitive deficit,
neuropsychiatric impairment, disturbed sleep pattern, and
dementia, caused by intrinsic PD pathology or drug-medi-
ated side effects, have gained attention. The involvement of
potential risks is highly recommended for modifying disease
progression or even improving clinical symptoms of NDs. A
better understanding of preclinical markers, neuroprotective
strategies are the silver lining for "at-risk" patients before the
clinical onset of the disease. In this context, non-pharmaco-
logical or psychosocial interventions will possibly offer both
symptomatic relief and disease modifications.
Psychosocial (non‑pharmacological)
interventions
Although a decrease in the cognitive impairment features in
NDs is evident from pharmacotherapies, the gain is minimal
and short term. Moreover, the prospective side effects out
Molecular and Cellular Biochemistry
1 3
of pharmacotherapies render the clinical management more
challenging as improving one symptom could end up dete-
riorating the other. Hence, non-pharmacological-based inter-
ventions are required to fight innumerable clinical aspects,
such as cognitive, functional, behavioral, and affective NDs.
Health professionals and scientists are now endorsing non-
pharmacological therapies for healing than just cure. Psy-
chosocial interventions refer to non-pharmacological tech-
niques, including cognitive training, cognitive stimulation
therapy, and cognitive rehabilitation, which aim to improve
cognition impairments.
Cognitive stimulation therapy (CST)
Limited evidence prevails supporting any significant positive
or negative impact of cognitive training (CT) and rehabili-
tation (CR) on cognitive outcomes in both AD and PD. In
comparison to CT and CR technique, the impact of the CS
approach in AD significantly improved [97]. A study based
on randomized controlled group design has shown the com-
bined efficacy of pharmacological and non-pharmacological
CST. The experimental group administered with AChEI
(Donepezil) along with 21h session of CST showed a slow
cognitive decline in comparison with the control group [98].
The impact of cognitive treatment with medication helps in
the improvement of language and cognition and stabilizes
routine activities in the patient with mild AD. Irrespective
of the fact that CS has proven efficacious, most of the studies
neglect to include it in the patients with PD. Many studies
investigated cognitive training in PD patients, yet most of
them deal with the non-demented population [99].
Motor therapy
Dance Movement Therapy (DMT) is one such motor thera-
pies, though underutilized but is preventive in cognitive deg-
radation [100]. DMT involves the use of sound and motor
movement for cognitive stimulation with the perspective
that declines in memory are a part of the aging process, but
mental and physical stimulation may prove preventive. This
therapy is based on the assumption of the body–mind con-
nection and neurophysiological mechanism works behind
DMT as it alters cognition through neurotransmitters. Motor
therapy improves functional abilities, social well-being, and
quality of life in patients with PD [101].
Music therapy
Music therapy has been found to reduce negative symptoms
(behavioral agitation) of AD particularly dementia [102].
Rhythmic sound is found to stimulate neural activity and
improve stages 1, 2, and 3 of AD. It improves memory
and orientation, reduces depression and anxiety, alleviates
delirium, agitation, hallucination, irritability, and language
disorder in the group of moderate AD. Music therapy
improves movement-related symptoms; however, group
music therapy has a substantial impact on the speech qual-
ity and voice range of a person with PD [103].
Dyadic exercise intervention
Studies with randomized control group design have shown a
positive impact of home-based group exercise on the execu-
tive function of patients with AD measured by Clock Draw-
ing Test [104]. The exercise would help increase the vascular
biomarkers of patients with AD. Although non-pharmaco-
logical treatments in AD have been studied substantially
and proven efficacious, limited research are undertaken for
Parkinson’s patients.
Other interventions
The psychosocial interventions combined with pharmaco-
logical interventions have produced better results [105]. A
combination of psychosocial interventions (CST, cognitive
training) and AChEI was found to be more effective than
inhibitor alone. It may well be inferred that pharmacological
treatments offset the symptoms but lack stable and effica-
cious results. Moreover, they are associated with various
side effects and are expensive. This could be overcome to
an appreciable extent by cost-effective non-pharmacological
interventions and their extensive use to improve cognition
or retard progression of cognitive impairment. The direct
impact of this approach would be delayed hospitalization and
a reduction in the costs of national healthcare. This would
improve both well-being of patients as well as caregivers.
Conclusions
Although AD and PD are having a distinct difference in their
clinical and pathological landscapes, both share several com-
mon features. The situation is worsening since no effective
cure for this multifactorial disease is in practice. A detailed
knowledge and better understanding of the risk factors
linked with the onset and advancement of these disorders
are critical for discovering novel targets and new therapeu-
tic approaches. Besides, early diagnosis could reduce the
incidence and pathophysiological advances of NDs. ChEIs
alter the clinical indicators of AD and the therapy results in
a modest but significant improvement. However, in the PD
pharmacological interventions by ChEIs are the first choice
since they have a positive effect on cognition, neuropsychi-
atric symptoms, and functions. Nevertheless, a large number
of randomized trials of ChEIs are needed to evaluate their
precise role. Future investigations should be focused on the
Molecular and Cellular Biochemistry
1 3
design and development of effective treatment options, based
on clinically vital outcomes. Moreover, high-quality rand-
omized control trials are needed to reach a strong conclusion
with powerful evidence.
The pharmacological therapies are successful in improv-
ing the symptoms of dementia and/or slowing the disease
progression. Nevertheless, the biological risk factors lack
explanation to the exact mechanism behind neurodegen-
eration. Focusing on the development of interventions that
improve cognition would be beneficial. It is believed that
the efficacy of treatment could be optimized if interven-
tions are initiated in the early or prodromal disease stage. In
this context, screening and treatment for psychosocial risk
factors would certainly decrease the threat. The available
interventions from pharmacological and psychosocial per-
spective would not only benefit health risk but also benefit
the healthcare cost.
Acknowledgements M.I.H. extends sincere thanks to the National
Medicinal Plants Board, Ministry of AYUSH, Government of India
for financial support (Grant No. Z. 18017/187/CSS/R&D/DL-01/2019-
20-NMPB-IVA). F.N. is thankful for the records of the manuscript
available in Social Science Cyber Library, Aligarh Muslim University,
Aligarh, India.
Author contributions F.N., N.N., and S.S. contributed to conceptual-
ization; F.N., N.N., G.M.H., and A.S.A. were involved in writing—
original draft; G. M. H. and M.I.H contributed to writing—review and
editing. All the authors have read and agreed to the published version
of the manuscript.
Data availability All data generated or analyzed during this study are
included in this published article.
Declarations
Conflict of interest The authors declare no conflict of interest.
References
1. Enache D, Winblad B, Aarsland D (2011) Depression in demen-
tia: epidemiology, mechanisms, and treatment. Curr Opin Psy-
chiatry 24:461–472
2. Kasl-Godley J, Gatz M (2000) Psychosocial interventions for
individuals with dementia: an integration of theory, therapy,
and a clinical understanding of dementia. Clin Psychol Rev
20:755–782
3. Kumar V, Sami N, Kashav T, Islam A, Ahmad F, Hassan MI
(2016) Protein aggregation and neurodegenerative diseases: from
theory to therapy. Eur J Med Chem 124:1105–1120
4. Voet S, Srinivasan S, Lamkanfi M, van Loo G (2019) Inflam-
masomes in neuroinflammatory and neurodegenerative diseases.
EMBO Mol Med 11:e10248
5. Rocca WA (2018) The burden of Parkinson’s disease: a world-
wide perspective. Lancet Neurol 17:928–929
6. Srinageshwar B, Petersen RB, Dunbar GL, Rossignol J
(2020) Prion-like mechanisms in neurodegenerative disease:
implications for Huntington’s disease therapy. Stem Cells Transl
Med 9:559–566
7. Kumar V, Islam A, Hassan MI, Ahmad F (2016) Therapeutic
progress in amyotrophic lateral sclerosis-beginning to learning.
Eur J Med Chem 121:903–917
8. Kumar V, Islam A, Hassan MI, Ahmad F (2016) Delineating the
relationship between amyotrophic lateral sclerosis and fronto-
temporal dementia: sequence and structure-based predictions.
Biochimica et Biophysica Acta Mol Basis Dis 1862:1742–1754
9. Sami N, Rahman S, Kumar V, Zaidi S, Islam A, Ali S, Ahmad
F, Hassan MI (2017) Protein aggregation, misfolding and con-
sequential human neurodegenerative diseases. Int J Neurosci
127:1047–1057
10. As A (2018) 2018 Alzheimer’s disease facts and figures. Alzhei-
mers Dement 14:367–429
11. Wood-Kaczmar A, Gandhi S, Wood N (2006) Understanding
the molecular causes of Parkinson’s disease. Trends Mol Med
12:521–528
12. Klein C, Westenberger A (2012) Genetics of Parkinson’s disease.
Cold Spring Harb Perspect Med 2:a008888
13. Karch CM, Cruchaga C, Goate AM (2014) Alzheimer’s disease
genetics: from the bench to the clinic. Neuron 83:11–26
14. Kalia LV, Lang AE, Hazrati L-N, Fujioka S, Wszolek ZK, Dick-
son DW, Ross OA, Van Deerlin VM, Trojanowski JQ, Hurtig
HI (2015) Clinical correlations with Lewy body pathology in
LRRK2-related Parkinson disease. JAMA Neurol 72:100–105
15. Outeiro TF, Koss DJ, Erskine D, Walker L, Kurzawa-Akanbi
M, Burn D, Donaghy P, Morris C, Taylor J-P, Thomas A (2019)
Dementia with Lewy bodies: an update and outlook. Mol Neu-
rodegener 14:1–18
16. Nalls MA, Pankratz N, Lill CM, Do CB, Hernandez DG, Saad
M, DeStefano AL, Kara E, Bras J, Sharma M (2014) Large-scale
meta-analysis of genome-wide association data identifies six new
risk loci for Parkinson’s disease. Nat Genet 46:989–993
17. Nalls MA, Blauwendraat C, Vallerga CL, Heilbron K, Bandres-
Ciga S, Chang D, Tan M, Kia DA, Noyce AJ, Xue A (2019)
Identification of novel risk loci, causal insights, and heritable
risk for Parkinson’s disease: a meta-analysis of genome-wide
association studies. Lancet Neurol 18:1091–1102
18. Kang SS, Zhang Z, Liu X, Manfredsson FP, Benskey MJ, Cao
X, Xu J, Sun YE, Ye K (2017) TrkB neurotrophic activities are
blocked by α-synuclein, triggering dopaminergic cell death in
Parkinson’s disease. Proc Natl Acad Sci 114:10773–10778
19. Farrer M, Kachergus J, Forno L, Lincoln S, Wang DS, Huli-
han M, Maraganore D, Gwinn-Hardy K, Wszolek Z, Dick-
son D (2004) Comparison of kindreds with parkinsonism and
α-synuclein genomic multiplications. Annal Neurol 55:174–179
20. Polymeropoulos MH, Lavedan C, Leroy E, Ide SE, Dehejia A,
Dutra A, Pike B, Root H, Rubenstein J, Boyer R (1997) Mutation
in the α-synuclein gene identified in families with Parkinson’s
disease. Science 276:2045–2047
21. Pupyshev AB, Korolenko TA, Akopyan AA, Amstislavskaya
TG, Tikhonova MA (2018) Suppression of autophagy in the
brain of transgenic mice with overexpression of A53T-mutant
α-synuclein as an early event at synucleinopathy progression.
Neurosci Lett 672:140–144
22. Tan EK, Skipper LM (2007) Pathogenic mutations in Parkinson
disease. Hum Mutat 28:641–653
23. Fang Y-Q, Mao F, Zhu M-J, Li X-H (2019) Compound heterozy-
gous mutations in PARK2 causing early-onset Parkinson disease:
a case report. Medicine 98:e14228
24. Abou-Sleiman PM, Healy DG, Quinn N, Lees AJ, Wood NW
(2003) The role of pathogenic DJ-1 mutations in Parkinson’s
disease. Ann Neurol 54:283–286
25. Valente EM, Abou-Sleiman PM, Caputo V, Muqit MM, Harvey
K, Gispert S, Ali Z, Del Turco D, Bentivoglio AR, Healy DG
Molecular and Cellular Biochemistry
1 3
(2004) Hereditary early-onset Parkinson’s disease caused by
mutations in PINK1. Science 304:1158–1160
26. Huang E, Qu D, Huang T, Rizzi N, Boonying W, Krolak D, Ciana
P, Woulfe J, Klein C, Slack RS (2017) PINK1-mediated phos-
phorylation of LETM1 regulates mitochondrial calcium transport
and protects neurons against mitochondrial stress. Nat Commun
8:1–11
27. Funayama M, Hasegawa K, Ohta E, Kawashima N, Komiyama
M, Kowa H, Tsuji S, Obata F (2005) An LRRK2 mutation as a
cause for the parkinsonism in the original PARK8 family. Annal
Neurol 57:918–921
28. Volpicelli-Daley LA, Abdelmotilib H, Liu Z, Stoyka L,
Daher JPL, Milnerwood AJ, Unni VK, Hirst WD, Yue Z,
Zhao HT (2016) G2019S-LRRK2 expression augments
α-synuclein sequestration into inclusions in neurons. J Neurosci
36:7415–7427
29. Nguyen APT, Daniel G, Valdés P, Islam MS, Schneider BL,
Moore DJ (2018) G2019S LRRK2 enhances the neuronal trans-
mission of tau in the mouse brain. Hum Mol Genet 27:120–134
30. West AB (2017) Achieving neuroprotection with LRRK2 kinase
inhibitors in Parkinson disease. Exp Neurol 298:236–245
31. Jonsson T, Atwal JK, Steinberg S, Snaedal J, Jonsson PV, Bjorns-
son S, Stefansson H, Sulem P, Gudbjartsson D, Maloney J (2012)
A mutation in APP protects against Alzheimer’s disease and age-
related cognitive decline. Nature 488:96–99
32. Goate A, Hardy J (2012) Twenty years of Alzheimer’s disease-
causing mutations. J Neurochem 120:3–8
33. O’brien RJ and Wong PC, (2011) Amyloid precursor pro-
tein processing and Alzheimer’s disease. Annu Rev Neurosci
34:185–204
34. Kelleher RJ, Shen J (2017) Presenilin-1 mutations and Alzhei-
mer’s disease. Proc Natl Acad Sci 114:629–631
35. Endres K, Fahrenholz F (2010) Upregulation of the α-secretase
ADAM10–risk or reason for hope? FEBS J 277:1585–1596
36. Yan R, Vassar R (2014) Targeting the β secretase BACE1 for
Alzheimer’s disease therapy. Lancet Neurol 13:319–329
37. Shamsi A, Anwar S, Mohammad T, Alajmi MF, Hussain A,
Rehman M, Hasan GM, Islam A, Hassan M (2020) MARK4
inhibited by AChE inhibitors, donepezil and Rivastigmine tar-
trate: insights into Alzheimer’s disease therapy. Biomolecules
10:789
38. Anwar S, Shamsi A, Kar RK, Queen A, Islam A, Ahmad F,
Hassan MI (2020) Structural and biochemical investigation of
MARK4 inhibitory potential of cholic acid: towards therapeutic
implications in neurodegenerative diseases. Int J Biol Macromol
161:596–604
39. Naz F, Shahbaaz M, Bisetty K, Islam A, Ahmad F, Hassan MI
(2015) Designing new kinase inhibitor derivatives as therapeutics
against common complex diseases: structural basis of microtu-
bule affinity-regulating kinase 4 (MARK4) inhibition. Omics A
J Integr Biol 19:700–711
40. Naqvi AAT, Jairajpuri DS, Noman OMA, Hussain A, Islam A,
Ahmad F, Alajmi MF, Hassan MI (2020) Evaluation of pyrazo-
lopyrimidine derivatives as microtubule affinity regulating kinase
4 inhibitors: towards therapeutic management of Alzheimer’s
disease. J Biomol Struct Dyn 38:3892–3907
41. Naz F, Anjum F, Islam A, Ahmad F, Hassan MI (2013) Microtu-
bule affinity-regulating kinase 4: structure, function, and regula-
tion. Cell Biochem Biophys 67:485–499
42. Annadurai N, Agrawal K, Džubák P, Hajdúch M, Das V (2017)
Microtubule affinity-regulating kinases are potential druggable
targets for Alzheimer’s disease. Cell Mol Life Sci 74:4159–4169
43. Hutton M, Lendon CL, Rizzu P, Baker M, Froelich S, Houlden
H, Pickering-Brown S, Chakraverty S, Isaacs A, Grover A (1998)
Association of missense and 5-splice-site mutations in tau with
the inherited dementia FTDP-17. Nature 393:702–705
44. Turab Naqvi AA, Hasan GM, Hassan M (2020) Targeting tau
hyperphosphorylation via kinase inhibition: strategy to address
Alzheimer’s disease. Curr Top Med Chem 20:1059–1073
45. Voura M, Khan P, Thysiadis S, Katsamakas S, Queen A, Hasan
GM, Ali S, Sarli V, Hassan MI (2019) Probing the inhibition
of microtubule affinity regulating kinase 4 by N-substituted
acridones. Sci Rep 9:1–17
46. Naz F, Sami N, Naqvi AT, Islam A, Ahmad F, Imtaiyaz Hassan
M (2017) Evaluation of human microtubule affinity-regulating
kinase 4 inhibitors: fluorescence binding studies, enzyme, and
cell assays. J Biomol Struct Dyn 35:3194–3203
47. Guha S, Fischer S, Johnson GV, Nehrke K (2020) Tauopathy-
associated tau modifications selectively impact neurodegenera-
tion and mitophagy in a novel C. elegans single-copy trans-
genic model. Mol Neurodegener 15:1–16
48. Riedel BC, Thompson PM, Brinton RD (2016) Age, APOE and
sex: triad of risk of Alzheimer’s disease. J Steroid Biochem
Mol Biol 160:134–147
49. Darweesh SK, Koudstaal PJ, Stricker BH, Hofman A, Ikram
MA (2016) Trends in the incidence of Parkinson disease in
the general population: the Rotterdam study. Am J Epidemiol
183:1018–1026
50. Lee A, Gilbert RM (2016) Epidemiology of Parkinson disease.
Neurol Clin 34:955–965
51. Longo VD, Antebi A, Bartke A, Barzilai N, Brown-Borg HM,
Caruso C, Curiel TJ, De Cabo R, Franceschi C, Gems D (2015)
Interventions to slow aging in humans: are we ready? Aging
Cell 14:497–510
52. Mielke MM, Vemuri P, Rocca WA (2014) Clinical epidemiol-
ogy of Alzheimer’s disease: assessing sex and gender differ-
ences. Clin Epidemiol 6:37
53. Liu W, Wong A, Law AC, Mok VC (2015) Cerebrovascular
disease, amyloid plaques, and dementia. Stroke 46:1402–1407
54. De Reuck J, Deramecourt V, Cordonnier C, Leys D, Maurage
C, Pasquier F (2011) The impact of cerebral amyloid angiopa-
thy on the occurrence of cerebrovascular lesions in demented
patients with Alzheimer features: a neuropathological study.
Eur J Neurol 18:913–918
55. Cordonnier C (2011) Brain microbleeds: more evidence, but
still a clinical dilemma. Curr Opin Neurol 24:69–74
56. Günaydın ZY, Özer FF, Karagöz A, Bektaş O, Karataş MB,
Vural A, Bayramoğlu A, Çelik A, Yaman M (2016) Evaluation
of cardiovascular risk in patients with Parkinson disease under
levodopa treatment. JGC 13:75
57. Li X, Song D, Leng SX (2015) Link between type 2 diabetes
and Alzheimer’s disease: from epidemiology to mechanism and
treatment. Clin Interv Aging 10:549
58. Götz J, Ittner L, Lim Y-A (2009) Common features between
diabetes mellitus and Alzheimer’s disease. Cell Mol Life Sci
66:1321–1325
59. Xu Q, Park Y, Huang X, Hollenbeck A, Blair A, Schatzkin
A, Chen H (2011) Diabetes and risk of Parkinson’s disease.
Diabetes Care 34:910–915
60. Chen J, Zhang C, Wu Y, Zhang D (2019) Association between
hypertension and the risk of Parkinson’s disease: a meta-anal-
ysis of analytical studies. Neuroepidemiology 52:181–192
61. Lennon MJ, Makkar SR, Crawford JD, Sachdev PS (2019)
Midlife hypertension and Alzheimer’s disease: a systematic
review and meta-analysis. J Alzheimers Dis 71:307–316
62. Fonseca ACR, Resende R, Oliveira CR, Pereira CM (2010)
Cholesterol and statins in Alzheimer’s disease: current con-
troversies. Exp Neurol 223:282–293
63. Simon KC, Chen H, Schwarzschild M, Ascherio A (2007)
Hypertension, hypercholesterolemia, diabetes, and risk of
Parkinson disease. Neurology 69:1688–1695
Molecular and Cellular Biochemistry
1 3
64. Pezzoli G, Cereda E (2013) Exposure to pesticides or solvents
and risk of Parkinson disease. Neurology 80:2035–2041
65. Klingelhoefer L, Reichmann H (2015) Pathogenesis of Parkin-
son disease—the gut–brain axis and environmental factors. Nat
Rev Neurol 11:625–636
66. Reynolds E (2014) The neurology of folic acid deficiency.
Handb Clin Neurol 120:927–943
67. Shen L, Ji H-F (2015) Associations between homocysteine,
folic acid, vitamin B12 and Alzheimer’s disease: insights from
meta-analyses. J Alzheimers Dis 46:777–790
68. Trinh J, Farrer M (2013) Advances in the genetics of Parkinson
disease. Nat Rev Neurol 9:445–454
69. García-Alberca JM, Cruz B, Lara JP, Garrido V, Gris E, Lara
A, Castilla C (2012) Disengagement coping partially medi-
ates the relationship between caregiver burden and anxiety
and depression in caregivers of people with Alzheimer’s dis-
ease. Results from the MÁLAGA-AD study. J Affect Disord
136:848–856
70. Zhou Z, Wang P, Fang Y (2018) Social engagement and its
change are associated with dementia risk among Chinese older
adults: a longitudinal study. Sci Rep 8:1–7
71. Saczynski JS, Pfeifer LA, Masaki K, Korf ES, Laurin D, White
L, Launer LJ (2006) The effect of social engagement on inci-
dent dementia: the Honolulu-Asia aging study. Am J Epidemiol
163:433–440
72. Topiwala H, Terrera GM, Stirland L, Saunderson K, Russ TC,
Dozier MF, Ritchie CW (2018) Lifestyle and neurodegeneration
in midlife as expressed on functional magnetic resonance imag-
ing: a systematic review. Alzheimer’s Dementia Transl Res Clin
Interven 4:182–194
73. Postuma RB, Iranzo A, Hogl B, Arnulf I, Ferini-Strambi L,
Manni R, Miyamoto T, Oertel W, Dauvilliers Y, Ju YE (2015)
Risk factors for neurodegeneration in idiopathic rapid eye move-
ment sleep behavior disorder: a multicenter study. Ann Neurol
77:830–839
74. Fernandez A, Santi A, Torres Aleman I (2018) Insulin peptides
as mediators of the impact of life style in Alzheimer’s disease.
Brain Plasticity 4:3–15
75. Solfrizzi V, Panza F, Frisardi V, Seripa D, Logroscino G,
Imbimbo BP, Pilotto A (2011) Diet and Alzheimer’s disease risk
factors or prevention: the current evidence. Expert Rev Neurother
11:677–708
76. Hughes KC, Gao X, Kim IY, Wang M, Weisskopf MG, Schwar-
zschild MA, Ascherio A (2017) Intake of dairy foods and risk of
Parkinson disease. Neurology 89:46–52
77. Costa J, Lunet N, Santos C, Santos J, Vaz-Carneiro A (2010) Caf-
feine exposure and the risk of Parkinson’s disease: a systematic
review and meta-analysis of observational studiess. J Alzheimers
Dis 20:S221–S238
78. Terracciano A, An Y, Sutin AR, Thambisetty M, Resnick SM
(2017) Personality change in the preclinical phase of Alzheimer
disease. JAMA Psychiatry 74:1259–1265
79. Terracciano A, Sutin AR, An Y, O’Brien RJ, Ferrucci L, Zon-
derman AB, Resnick SM (2014) Personality and risk of Alzhei-
mer’s disease: new data and meta-analysis. Alzheimers Dement
10:179–186
80. Sieurin J, Gustavsson P, Weibull CE, Feldman AL, Petzinger
GM, Gatz M, Pedersen NL, Wirdefeldt K (2016) Personality
traits and the risk for Parkinson disease: a prospective study. Eur
J Epidemiol 31:169–175
81. Holmer J, Eriksdotter M, Schultzberg M, Pussinen PJ, Buhlin K
(2018) Association between periodontitis and risk of Alzheimer’s
disease, mild cognitive impairment and subjective cognitive
decline: a case–control study. J Clin Periodontol 45:1287–1298
82. Ngandu T, von Strauss E, Helkala E-L, Winblad B, Nissinen
A, Tuomilehto J, Soininen H, Kivipelto M (2007) Education
and dementia: what lies behind the association? Neurology
69:1442–1450
83. Bickel H, Kurz A (2009) Education, occupation, and dementia:
the Bavarian school sisters study. Dement Geriatr Cogn Disord
27:548–556
84. Hsieh SH (2020) Depression in Parkinson’s disease. In: UCLA
85. Sami N, Kumar V, Islam A, Ali S, Ahmad F, Hassan I (2017)
Exploring missense mutations in tyrosine kinases implicated
with neurodegeneration. Mol Neurobiol 54:5085–5106
86. Hrelia P, Sita G, Ziche M, Ristori E, Marino A, Cordaro M,
Molteni R, Spero V, Malaguti M, Morroni F (2020) Common
protective strategies in neurodegenerative disease: focusing on
risk factors to target the cellular redox system. Oxid Med Cell
Longev 2020:1–18
87. Santos TCD, Gomes TM, Pinto BAS, Camara AL, Paes AMdA
(2018) Naturally occurring acetylcholinesterase inhibitors and
their potential use for Alzheimer’s disease therapy. Front Phar-
macol 9:1192
88. Tan C-C, Yu J-T, Wang H-F, Tan M-S, Meng X-F, Wang C,
Jiang T, Zhu X-C, Tan L (2014) Efficacy and safety of donepezil,
galantamine, rivastigmine, and memantine for the treatment of
Alzheimer’s disease: a systematic review and meta-analysis. J
Alzheimers Dis 41:615–631
89. Marucci G, Moruzzi M, Amenta F (2010) Donepezil in the
treatment of Alzheimer’s disease. Diagnosis and management
in dementia. Elsevier, Amsterdam, pp 495–510
90. Svenningsson P, Westman E, Ballard C, Aarsland D (2012) Cog-
nitive impairment in patients with Parkinson’s disease: diagnosis,
biomarkers, and treatment. Lancet Neurol 11:697–707
91. Dubois B, Tolosa E, Katzenschlager R, Emre M, Lees AJ, Schu-
mann G, Pourcher E, Gray J, Thomas G, Swartz J (2012) Done-
pezil in Parkinson’s disease dementia: a randomized, double-
blind efficacy and safety study. Mov Disord 27:1230–1238
92. Emre M, Ford PJ, Bilgiç B, Uç EY (2014) Cognitive impair-
ment and dementia in Parkinson’s disease: practical issues and
management. Mov Disord 29:663–672
93. Vanle B, Olcott W, Jimenez J, Bashmi L, Danovitch I, IsHak WW
(2018) NMDA antagonists for treating the non-motor symptoms
in Parkinson’s disease. Transl Psychiatry 8:1–15
94. Alam S, Lingenfelter KS, Bender AM, Lindsley CW (2017) Clas-
sics in chemical neuroscience: memantine. ACS Chem Neurosci
8:1823–1829
95. Song X, Jensen MØ, Jogini V, Stein RA, Lee C-H, Mchaourab
HS, Shaw DE, Gouaux E (2018) Mechanism of NMDA receptor
channel block by MK-801 and memantine. Nature 556:515–519
96. Dong J, Cui Y, Li S, Le W (2016) Current pharmaceutical treat-
ments and alternative therapies of Parkinson’s disease. Curr Neu-
ropharmacol 14:339–355
97. Woods B, Aguirre E, Spector AE, Orrell M (2012) Cogni-
tive stimulation to improve cognitive functioning in peo-
ple with dementia. Cochrane Database Syst Rev. https ://doi.
org/10.1002/14651 858.CD005 562.pub2
98. Spector A, Woods B, Orrell M (2008) Cognitive stimulation
for the treatment of Alzheimer’s disease. Expert Rev Neurother
8:751–757
99. Angelucci F, Peppe A, Carlesimo GA, Serafini F, Zabberoni S,
Barban F, Shofany J, Caltagirone C, Costa A (2015) A pilot study
on the effect of cognitive training on BDNF serum levels in indi-
viduals with Parkinson’s disease. Front Hum Neurosci 9:130
100. Hackney ME, Bennett CG (2014) Dance therapy for individu-
als with Parkinson’s disease: improving quality of life. Res Rev
Parkinsonism 4:17–25
101. Lihala S, Mitra S, Neogy S, Datta N, Choudhury S, Chatterjee K,
Mondal B, Halder S, Roy A, Sengupta M (2020) Dance move-
ment therapy in rehabilitation of Parkinson’s disease–a feasibility
study. J Bodywork Mov Ther 26:12–17
Molecular and Cellular Biochemistry
1 3
102. Fang R, Ye S, Huangfu J, Calimag DP (2017) Music therapy is
a potential intervention for cognition of Alzheimer’s disease: a
mini-review. Transl Neurodegener 6:1–8
103. Pereira APS, Marinho V, Gupta D, Magalhães F, Ayres C, Teix-
eira S (2019) Music therapy and dance as gait rehabilitation in
patients with Parkinson disease: a review of evidence. J Geriatr
Psychiatry Neurol 32:49–56
104. Lamotte G, Shah RC, Lazarov O, Corcos DM (2017) Exercise
training for persons with Alzheimer’s disease and caregivers: a
review of dyadic exercise interventions. J Mot Behav 49:365–377
105. Duan Y, Lu L, Chen J, Wu C, Liang J, Zheng Y, Wu J, Rong
P, Tang C (2018) Psychosocial interventions for Alzheimer’s
disease cognitive symptoms: a Bayesian network meta-analysis.
BMC Geriatr 18:1–11
Publisher’s Note Springer Nature remains neutral with regard to
jurisdictional claims in published maps and institutional affiliations.
... In the context of the absence of curative treatment for AD at present, research focused on the pathophysiology of this neurodegenerative disease (NDD) is one of the main modalities to enhance the development of more effective therapies in the near future [20]. In this regard, the aim of this article is to offer a detailed overview of the role of Aβ monomers in the pathogenesis of AD, focusing mainly on clinically relevant data obtained from the most important trials conducted in recent years. ...
Article
Full-text available
Amyloid beta peptide is an important biomarker in Alzheimer’s disease, with the amyloidogenic hypothesis as one of the central hypotheses trying to explain this type of dementia. Despite numerous studies, the etiology of Alzheimer’s disease remains incompletely known, as the pathological accumulation of amyloid beta aggregates cannot fully explain the complex clinical picture of the disease. Or, for the development of effective therapies, it is mandatory to understand the roles of amyloid beta at the brain level, from its initial monomeric stage prior to aggregation in the form of senile plaques. In this sense, this review aims to bring new, clinically relevant data on a subject intensely debated in the literature in the last years. In the first part, the amyloidogenic cascade is reviewed and the possible subtypes of amyloid beta are differentiated. In the second part, the roles played by the amyloid beta monomers in physiological and pathological (neurodegenerative) conditions are illustrated based on the most relevant and recent studies published on this topic. Finally, considering the importance of amyloid beta monomers in the pathophysiology of Alzheimer’s disease, new research directions with diagnostic and therapeutic impacts are suggested.
... Misfolding of protein leads to the build-up of debris in the neuronal region, and diseases arising due to this issue are named proteinopathy. According to the etiology, certain NDs, namely transmissible spongiform encephalopathies, Alzheimer's disease (AD), and Parkinson's disease (PD), are classified under proteinopathies [21,22]. Aging is one of the major risk factors, and the two factors that greatly contribute to aging are mitochondrial damage and oxidative stress [23]. ...
Article
Full-text available
Purpose of Review What accounts for the low success rate of the treatment against neurodegenerative disorders (ND) is the inability of the neurons to replenish again after cell death. Neuronal cell death and axon degenerations are common as we age. The etiology of the ND holds multiple reasons, among which the accumulation of metal ions is the common cause. Recent Findings Metals are the necessary elements to maintain homeostasis and participate in the regulation of other metabolic activities. The presence and concentration of specific metals at specific locations are required to perform specific functions. The concentration of the metals should be maintained on a scale to avoid unfavorable actions. Deficiency or accumulation of metals results in neuron dysfunction or death. Issues occurring due to this need to be addressed by appropriate imaging techniques to cure the condition. Proper diagnostic methods to locate the cell death and provide better therapy are still lacking in the case of ND. This is due to the physiological barrier, the blood–brain barrier (BBB), which provides less or no access to the foreign bodies into the brain. The only approach to avail access to the brain through BBB as a theranostic appeal to treat ND is to implement nanoassisted methods. Summary Bioengineered nanostructures can easily gain entry to the brain due to their tiny architectures. Several deep research investigations provide confidence in developing nanomaterials to assist the physician in curing the diseased condition and restoring normal metabolism. We have discussed several pieces of research evidence proving that nanotechnology has a substantial role as a theranostic platform for ND. This manuscript is believed to provide a better understanding of the disease and the recent findings in the field.
... Neurodegenerative diseases, such as Alzheimer's disease patients (AD), Huntington's disease (HD) and Parkinson's disease (PD), are common causes of morbidity, mortality, and cognitive impairment in older adults [1]. The increasing prevalence of neurodegenerative diseases brings excellent social and economic burdens and has become a global problem [2]. ...
Article
Full-text available
Background Neurodegenerative diseases, such as Alzheimer's disease patients (AD), Huntington's disease (HD) and Parkinson’s disease (PD), are common causes of morbidity, mortality, and cognitive impairment in older adults. Objective We aimed to understand the transcriptome characteristics of the cortex of neurodegenerative diseases and to provide an insight into the target genes of differently expressed microRNAs in the occurrence and development of neurodegenerative diseases. Methods The Limma package of R software was used to analyze GSE33000, GSE157239, GSE64977 and GSE72962 datasets to identify the differentially expressed genes (DEGs) and microRNAs in the cortex of neurodegenerative diseases. Bioinformatics methods, such as GO enrichment analysis, KEGG enrichment analysis and gene interaction network analysis, were used to explore the biological functions of DEGs. Weighted gene co-expression network analysis (WGCNA) was used to cluster DEGs into modules. RNA22, miRDB, miRNet 2.0 and TargetScan7 databases were performed to predict the target genes of microRNAs. Results Among 310 Alzheimer's disease (AD) patients, 157 Huntington's disease (HD) patients and 157 non-demented control (Con) individuals, 214 co-DEGs were identified. Those co-DEGs were filtered into 2 different interaction network complexes, representing immune-related genes and synapse-related genes. The WGCNA results identified five modules: yellow, blue, green, turquoise, and brown. Most of the co-DEGs were clustered into the turquoise module and blue module, which respectively regulated synapse-related function and immune-related function. In addition, human microRNA-4433 (hsa-miR-4443), which targets 18 co-DEGs, was the only 1 co-up-regulated microRNA identified in the cortex of neurodegenerative diseases. Conclusion 214 DEGs and 5 modules regulate the immune-related and synapse-related function of the cortex in neurodegenerative diseases. Hsa-miR-4443 targets 18 co-DEGs and may be a potential molecular mechanism in neurodegenerative diseases' occurrence and development.
... With increasing age, the prevalence of PD also increases. This prevalence soars up to 5 % in persons aged 85 or above (Nazam et al., 2021). Familial PD patients under the age of 60 contribute around 5 % of all PD cases similar to sporadic cases due to mutations in multiple genes such as PINK-1, Parkin, DJ-1, and asyn. ...
Article
Aging is a progressive loss of physiological function that increases risk of disease and death. Among the many factors that contribute to human aging, mitochondrial dysfunction has emerged as one of the most prominent features of the aging process. It has been linked to the development of various age-related pathologies, including Parkinson's disease (PD). Mitochondria has a complex quality control system that ensures mitochondrial integrity and function. Perturbations in these mitochondrial mechanisms have long been linked to various age-related neurological disorders. Even though research has shed light on several aspects of the disease pathology, the underlying mechanism of age-related factors responsible for individuals developing this disease is still unknown. This review article aims to discuss the role of mitochondria in the transition from normal brain aging to pathological brain aging, which leads to the progression of PD. We have discussed the emerging evidence on how age-related disruption of mitochondrial quality control mechanisms contributes to the development of PD-related pathophysiology.
Article
Full-text available
This study investigates the association between socioeconomic position (SEP)-in terms of income and education-and mortality from neurodegenerative diseases, that is, dementia, parkinsonism, and motor neuron diseases (MNDs). We calculated age-standardized mortality rates and mortality rate ratios using log linear Poisson regression for different SEP groups, stratified by gender, age-group, and care home residency, utilizing the 2011 Belgian census linked to register data on cause-specific mortality for 2011 to 2016. Mortality was significantly higher in the lowest educational-and income groups. The largest disparities were found in dementia mortality. Income had a strong negative effect on parkinsonism mortality, education a positive effect. We found no significant association between SEP and MND. Our study provides evidence supporting the presence of socioeconomic disparities in mortality due to neurodegeneration. We found a strong negative association between SEP and NDD mortality, which varies between NDD, gender and care home residency. What do we already know about this topic? NDDs are recognized as a significant global public health concern, with documented connections to SEP, underscoring the critical importance of addressing these conditions for comprehensive public health strategies worldwide. How does your research contribute to the field? Our research provides evidence of socioeconomic disparities in mortality due to neurodegeneration, while unveiling distinct patterns for dementia, parkinsonism, and MNDs across gender and care home residency. What are your research's implications toward theory, practice, or policy? The findings underscore the importance of addressing socioeconomic inequalities in healthcare, emphasizing the need for targeted policy interventions, particularly in the context of NDDs.
Article
Iron deposition in the central nervous system (CNS) is one of the causes of neurodegenerative diseases. Human transferrin (hTf) acts as an iron carrier present in the blood plasma, preventing it from contributing to redox reactions. Plant compounds and their derivatives are frequently being used in preventing or delaying Alzheimer's disease (AD). Thymoquinone (TQ), a natural product has gained popularity because of its broad therapeutic applications. TQ is one of the significant phytoconstituent of Nigella sativa. The binding of TQ to hTf was determined by spectroscopic methods and isothermal titration calorimetry. We have observed that TQ strongly binds to hTf with a binding constant (K) of 0.22 × 106 M−1 and forming a stable complex. In addition, isothermal titration calorimetry revealed the spontaneous binding of TQ with hTf. Molecular docking analysis showed key residues of the hTf that were involved in the binding to TQ. We further performed a 250 ns molecular dynamics simulation which deciphered the dynamics and stability of the hTf‐TQ complex. Structure analysis suggested that the binding of TQ doesn't cause any significant alterations in the hTf structure during the course of simulation and a stable complex is formed. Altogether, we have elucidated the mechanism of binding of TQ with hTf, which can be further implicated in the development of a novel strategy for AD therapy.
Article
Full-text available
Salivary Aβ40, Aβ42, t-tau, and p-tau 181 are commonly employed in Alzheimer’s disease (AD) investigations. However, the collection method of these biomarkers can affect their levels. To assess the impact of saliva collection methods on biomarkers in this study, 15 healthy people were employed in the morning with six saliva collection methods. The chosen methods were then applied in 30 AD patients and 30 non-AD controls. The levels of salivary biomarkers were calculated by a specific enzyme-linked immunosorbent assay. The receiver operating characteristic was utilized to assess salivary biomarkers in AD patients. The results demonstrated that the highest levels of salivary Aβ40, Aβ42, t-tau, and p-tau were in different saliva collection methods. The correlations between different saliva biomarkers in the same collection method were different. Salivary Aβ40, Aβ42, t-tau, and p-tau had no significant association. Salivary Aβ42 was higher in AD than in non-AD controls. However, p-tau/t-tau and Aβ42/Aβ40 had some relevance. The area under the curve for four biomarkers combined in AD diagnosis was 92.11%. An alternate saliva collection method (e.g., USS in Aβ40, UPS in Aβ42, t-tau, SSS in p-tau 181) was demonstrated in this study. Moreover, combining numerous biomarkers improves AD diagnosis.
Article
Background Alzheimer´s disease (AD) is the most widespread dementia in the world, followed by vascular dementia. Since AD is a heterogeneous disease that shows several varied phenotypes, it is not easy to make an accurate diagnosis, so it arises when the symptoms are clear and the disease is already very advanced. Therefore, it is important to find out biomarkers for AD early diagnosis that facilitate treatment or slow down the disease. Classic biomarkers are obtained from cerebrospinal fluid and plasma, along with brain imaging by positron emission tomography. Attempts have been made to discover uncommon biomarkers from other body fluids, which are addressed in this update. Objective This update aims to describe recent biomarkers from minimally invasive body fluids for the patients, such as saliva, urine, eye fluid or tears. Methods Biomarkers were determined in patients versus controls by single tandem mass spectrometry, and immunoassays. Metabolites were identified by nuclear magnetic resonance, and microRNAs with genome-wide high-throughput real-time polymerase chain reaction-based platforms. Results Biomarkers from urine, saliva, and eye fluid were described, including peptides/proteins, metabolites, and some microRNAs. The association with AD neuroinflammation and neurodegeneration was analyzed, highlighting the contribution of matrix metalloproteinases, the immune system and microglia, as well as the vascular system. Conclusion Unusual biomarkers have been developed, which distinguish each stage and progression of the disease, and are suitable for the early AD diagnosis. An outstanding relationship of biomarkers with neuroinflammation and neurodegeneration was assessed, clearing up concerns of the etiopathogenesis of AD.
Article
Background Alzheimer's disease (AD) is a progressive neurodegenerative disease of growing interest given that there is cognitive damage and symptom onset acceleration. Therefore, it is important to find AD biomarkers for early diagnosis, disease progression, and discrimination of AD and other diseases. Objective To update the relevance of mass spectrometry for the identification of peptides and proteins involved in AD useful as discriminating biomarkers. Methods Proteomics and peptidomics technologies that show the highest possible specificity and selectivity for AD biomarkers are analyzed, together with the biological fluids used. In addition to positron emission tomography and magnetic resonance imaging, MALDI-TOF mass spectrometry is widely used to identify proteins and peptides involved in AD. The use of protein chips in SELDI technology and electroblotting chips for peptides makes feasible small amounts (L) of samples for analysis. Results Suitable biomarkers are related to AD pathology, such as intracellular neurofibrillary tangles; extraneuronal senile plaques; neuronal and axonal degeneration; inflammation and oxidative stress. Recently, peptides were added to the candidate list, which are not amyloid-b or tau fragments, but are related to coagulation, brain plasticity, and complement/neuroinflammation systems involving the neurovascular unit. Conclusion The progress made in the application of mass spectrometry and recent chip techniques is promising for discriminating between AD, mild cognitive impairment, and matched healthy controls. The application of this technique to blood samples from patients with AD has shown to be less invasive and fast enough to determine the diagnosis, stage of the disease, prognosis, and follow-up of the therapeutic response.
Article
Full-text available
Mutations in the leucine-rich repeat kinase 2 (LRRK2) gene cause late-onset, autosomal dominant Parkinson's disease (PD). LRRK2 mutations typically give rise to Lewy pathology in the brains of PD subjects yet can induce tau-positive neuropathology in some cases. The pathological interaction between LRRK2 and tau remains poorly defined. To explore this interaction in vivo, we crossed a well-characterized human P301S-tau transgenic mouse model of tauopathy with human G2019S-LRRK2 transgenic mice or LRRK2 knockout (KO) mice. We find that endogenous or pathogenic LRRK2 expression has minimal effects on the steady-state levels, solubility and abnormal phosphorylation of human P301S-tau throughout the mouse brain. We next developed a new model of tauopathy by delivering AAV2/6 vectors expressing human P301S-tau to the hippocampal CA1 region of G2019S-LRRK2 transgenic or LRRK2 KO mice. P301S-tau expression induces hippocampal tau pathology and marked degeneration of CA1 pyramidal neurons in mice, however, this occurs independently of endogenous or pathogenic LRRK2 expression. We further developed new AAV2/6 vectors co-expressing human WT-tau and GFP to monitor the neuron-to-neuron transmission of tau within defined hippocampal neuronal circuits. While endogenous LRRK2 is not required for tau transmission, we find that G2019S-LRRK2 markedly enhances the neuron-to-neuron transmission of tau in mice. Our data suggest that mutant tau-induced neuropathology occurs independently of LRRK2 expression in two mouse models of tauopathy but identifies a novel pathogenic role for G2019S-LRRK2 in promoting the neuronal transmission of WT-tau protein. These findings may have important implications for understanding the development of tau neuropathology in LRRK2-linked PD brains.
Article
Full-text available
Background A defining pathological hallmark of the progressive neurodegenerative disorder Alzheimer’s disease (AD) is the accumulation of misfolded tau with abnormal post-translational modifications (PTMs). These include phosphorylation at Threonine 231 (T231) and acetylation at Lysine 274 (K274) and at Lysine 281 (K281). Although tau is recognized to play a central role in pathogenesis of AD, the precise mechanisms by which these abnormal PTMs contribute to the neural toxicity of tau is unclear. Methods Human 0N4R tau (wild type) was expressed in touch receptor neurons of the genetic model organism C. elegans through single-copy gene insertion. Defined mutations were then introduced into the single-copy tau transgene through CRISPR-Cas9 genome editing. These mutations included T231E, to mimic phosphorylation of a commonly observed pathological epitope, and K274/281Q, to mimic disease-associated lysine acetylation – collectively referred as “PTM-mimetics” – as well as a T231A phosphoablation mutant. Stereotypical touch response assays were used to assess behavioral defects in the transgenic strains as a function of age. Genetically-encoded fluorescent biosensors were expressed in touch neurons and used to measure neuronal morphology, mitochondrial morphology, mitophagy, and macro autophagy. Results Unlike existing tau overexpression models, C. elegans single-copy expression of tau did not elicit overt pathological phenotypes at baseline. However, strains expressing disease associated PTM-mimetics (T231E and K274/281Q) exhibited reduced touch sensation and neuronal morphological abnormalities that increased with age. In addition, the PTM-mimetic mutants lacked the ability to engage neuronal mitophagy in response to mitochondrial stress. Conclusions Limiting the expression of tau results in a genetic model where modifications that mimic pathologic tauopathy-associated PTMs contribute to cryptic, stress-inducible phenotypes that evolve with age. These findings and their relationship to mitochondrial stress provides a new perspective into the pathogenic mechanisms underlying AD.
Article
Full-text available
Neurodegenerative disease is an umbrella term for different conditions which primarily affect the neurons in the human brain. In the last century, significant research has been focused on mechanisms and risk factors relevant to the multifaceted etiopathogenesis of neurodegenerative diseases. Currently, neurodegenerative diseases are incurable, and the treatments available only control the symptoms or delay the progression of the disease. This review is aimed at characterizing the complex network of molecular mechanisms underpinning acute and chronic neurodegeneration, focusing on the disturbance in redox homeostasis, as a common mechanism behind five pivotal risk factors: aging, oxidative stress, inflammation, glycation, and vascular injury. Considering the complex multifactorial nature of neurodegenerative diseases, a preventive strategy able to simultaneously target multiple risk factors and disease mechanisms at an early stage is most likely to be effective to slow/halt the progression of neurodegenerative diseases.
Article
Full-text available
Microtubule affinity-regulating kinase (MARK4) plays a key role in Alzheimer’s disease (AD) development as its overexpression is directly linked to increased tau phosphorylation. MARK4 is a potential drug target of AD and is thus its structural features are employed in the development of new therapeutic molecules. Donepezil (DP) and rivastigmine tartrate (RT) are acetylcholinesterase (AChE) inhibitors and are used to treat symptomatic patients of mild to moderate AD. In keeping with the therapeutic implications of DP and RT in AD, we performed binding studies of these drugs with the MARK4. Both DP and RT bound to MARK4 with a binding constant (K) of 107 M−1. The temperature dependency of binding parameters revealed MARK−DP complex to be guided by static mode while MARK−RT complex to be guided by both static and dynamic quenching. Both drugs inhibited MARK4 with IC50 values of 5.3 μM (DP) and 6.74 μM (RT). The evaluation of associated enthalpy change (ΔH) and entropy change (ΔS) implied the complex formation to be driven by hydrogen bonding making it seemingly strong and specific. Isothermal titration calorimetry further advocated a spontaneous binding. In vitro observations were further complemented by the calculation of binding free energy by molecular docking and interactions with the functionally-important residues of the active site pocket of MARK4. This study signifies the implications of AChE inhibitors, RT, and DP in Alzheimer’s therapy targeting MARK4.
Article
Full-text available
Huntington’s disease (HD) is an autosomal dominant neurodegenerative disorder caused by a CAG repeat expansions in the huntingtin gene resulting in the synthesis of a misfolded form of the huntingtin protein (mHTT) which is toxic. The current treatments for HD are only palliative. Some of the potential therapies for HD include gene therapy (using antisense oligonucleotides and clustered regularly interspaced short palindromic repeats-Cas9 system) and stem-cell-based therapies. Various types of stem cell transplants, such as mesenchymal stem cells, neural stem cells, and reprogrammed stem cells, have the potential to either replace the lost neurons or support the existing neurons by releasing trophic factors. Most of the transplants are xenografts and allografts; however, recent reports on HD patients who received grafts suggest that the mHTT aggregates are transferred from the host neurons to the grafted cells as well as to the surrounding areas of the graft by a “prion-like” mechanism. This observation seems to be true for autotransplantation paradigms, as well. This article reviews the different types of stem cells that have been transplanted into HD patients and their therapeutic efficacy, focusing on the transfer of mHTT from the host cells to the graft. Autotransplants of reprogramed stem cells in HD patients are a promising therapeutic option. However, this needs further attention to ensure a better understanding of the transfer of mHTT aggregates following transplantation of the gene-corrected cells back into the patient. Significance statement Stem cell transplantation, along with gene editing using a variety of molecular tools, is one of the most promising strategies that is being investigated by many researchers as a potential treatment for neurodegenerative diseases. Huntington’s disease (HD) is one of the neurodegenerative diseases in which transplantation has been widely studied using different types of innate as well as reprogrammed/modified stem cells as a potential therapy. Transplantation using different types of stem cells, such as mesenchymal stem cells, neural stem cells, embryonic stem cells, and induced pluripotent stem cells accompanied by clustered regularly interspaced short palindromic repeats-Cas9-based gene editing, was performed in laboratory settings, which could have an impact in the clinics in the near future. Though the treatment strategies had encouraging outcomes, one of the major issues identified recently was that mutant huntingtin protein aggregates transfer from the HD cells to the wild-type/transplanted cells in the host brain, by a “prion-like” mechanism. The finding brings into question to what extent these stem cell/gene-corrected cell transplants are a viable option for treating HD.
Chapter
Observations of a loss of cholinergic function in neocortex and hippocampus in Alzheimer's disease (AD) have led to the hypothesis that replacing the cholinergic function may be of therapeutic benefit. Among the different approaches proposed or tested, acetylcholinesterase (AChE) and cholinesterase (ChE) inhibition became the focus of newly developed drugs that obtained the first regulatory recognition for the treatment of AD. Donepezil, licensed for the symptomatic treatment of AD, is the second AChE/ChE inhibitor introduced into the market. This compound displays several effects, including a protection against amyloid β, ischemia, and glutamate toxicity. Moreover, it slows the progression of hippocampal atrophy and upregulates nicotinic acetylcholine receptors. From a clinical point of view, donepezil ameliorates cognitive symptoms and to a lesser extent the behavioral problems occurring in AD. Early and long-lasting treatment with donepezil is more effective than delayed treatment. In spite of all this, the overall clinical benefit of donepezil is small and may not be clinically meaningful.
Article
Introduction Environmental enrichment during physical exercise was found beneficial in neurological disorders. Application of dance in a structured way could effectively enhance the environment of physical rehabilitation. Therefore, dance therapy can be an alternative exercise program with potential benefit in affect, cognition and social integration in various neurological disorders. Objective This pre-post experimental study without control was designed to assess the impact of dance movement therapy on cognition, quality of life and motor symptoms in PD patients. Methods A group of 10 mild-moderate PD patients from Movement Disorders Clinic; I-NK, participated in group sessions for a period of 2 months (twice a week). Each session involved verbal communication followed by warming up movements and concluded with target oriented physical activities, focused on physical symptoms, emotional and cognitive aspects. All the patients were assessed before and after the intervention using Unified Parkinson’s Disease Rating Scale part III (UPDRS part III), Hoehn and Yahr Scale (H and Y), Parkinson’s Disease Questionnaire 39 (PDQ-39) and Montreal Cognitive Assessment (MOCA). Results We observed a change in median MOCA score from 19.00 to 22.00 (p .027). PDQ 39 also showed change in median score from 59.50 to 30.00 (p .027). The change in UPDRS III (0.08) and H and Y (.157) failed to reach significant limit. Conclusion Dance Movement Therapy was found beneficial in overall cognition and quality of life in patients with mild- moderate PD. Studies with larger sample size will assess the long-term safety and effectiveness of this alternative therapy in future.
Article
Microtubule associated protein tau is involved in the tubulin binding leading to microtubule stabilization in neuronal cells which is essential for stabilization of neuron cytoskeleton. The regulation of tau activity is accommodated by several kinases which phosphorylate tau protein on specific sites. In pathological conditions, abnormal activity of tau kinases such as glycogen synthase kinase-3 β (GSK3β), cyclin-dependent kinase 5 (CDK5), c-Jun N-terminal kinases (JNKs), extracellular signal-regulated kinase 1 and 2 (ERK1/2) and microtubule affinity regulating kinase (MARK) leads to tau hyperphosphorylation. Hyperphosphorylation of tau protein leads to aggregation of tau into paired helical filaments like structures which are major constituents of neurofibrillary tangles, ahallmark of Alzheimer’s disease. In this review, we discuss various tau protein kinases and their association with tau hyperphosphorylation. We also discuss various strategies and the advancements made in the area of Alzheimer's disease drug development by designing effective and specific inhibitors for such kinases using traditional in vitro/in vivo methods and state of the art in silico techniques.
Article
Background: Genome-wide association studies (GWAS) in Parkinson's disease have increased the scope of biological knowledge about the disease over the past decade. We aimed to use the largest aggregate of GWAS data to identify novel risk loci and gain further insight into the causes of Parkinson's disease. Methods: We did a meta-analysis of 17 datasets from Parkinson's disease GWAS available from European ancestry samples to nominate novel loci for disease risk. These datasets incorporated all available data. We then used these data to estimate heritable risk and develop predictive models of this heritability. We also used large gene expression and methylation resources to examine possible functional consequences as well as tissue, cell type, and biological pathway enrichments for the identified risk factors. Additionally, we examined shared genetic risk between Parkinson's disease and other phenotypes of interest via genetic correlations followed by Mendelian randomisation. Findings: Between Oct 1, 2017, and Aug 9, 2018, we analysed 7·8 million single nucleotide polymorphisms in 37 688 cases, 18 618 UK Biobank proxy-cases (ie, individuals who do not have Parkinson's disease but have a first degree relative that does), and 1·4 million controls. We identified 90 independent genome-wide significant risk signals across 78 genomic regions, including 38 novel independent risk signals in 37 loci. These 90 variants explained 16–36% of the heritable risk of Parkinson's disease depending on prevalence. Integrating methylation and expression data within a Mendelian randomisation framework identified putatively associated genes at 70 risk signals underlying GWAS loci for follow-up functional studies. Tissue-specific expression enrichment analyses suggested Parkinson's disease loci were heavily brain-enriched, with specific neuronal cell types being implicated from single cell data. We found significant genetic correlations with brain volumes (false discovery rate-adjusted p=0·0035 for intracranial volume, p=0·024 for putamen volume), smoking status (p=0·024), and educational attainment (p=0·038). Mendelian randomisation between cognitive performance and Parkinson's disease risk showed a robust association (p=8·00 × 10−7). Interpretation: These data provide the most comprehensive survey of genetic risk within Parkinson's disease to date, to the best of our knowledge, by revealing many additional Parkinson's disease risk loci, providing a biological context for these risk factors, and showing that a considerable genetic component of this disease remains unidentified. These associations derived from European ancestry datasets will need to be followed-up with more diverse data.