ArticlePDF AvailableLiterature Review

Nucleic Acid-Based Theranostics for Tackling Alzheimer's Disease

Authors:
  • Centre for Molecular Medicine and Innovative Therapeutics - Murdoch University & Perron Institute for Neurological and Translational Science

Abstract and Figures

Nucleic acid-based technologies have received significant interest in recent years as novel theranostic strategies for various diseases. The approval by the United States Food and Drug Administration (FDA) of Nusinersen, an antisense oligonucleotide drug, for the treatment of spinal muscular dystrophy highlights the potential of nucleic acids to treat neurological diseases, including Alzheimer's disease (AD). AD is a devastating neurodegenerative disease characterized by progressive impairment of cognitive function and behavior. It is the most common form of dementia; it affects more than 20% of people over 65 years of age and leads to death 7-15 years after diagnosis. Intervention with novel agents addressing the underlying molecular causes is critical. Here we provide a comprehensive review on recent developments in nucleic acid-based theranostic strategies to diagnose and treat AD.
Content may be subject to copyright.
Theranostics 2017, Vol. 7, Issue 16
http://www.thno.org
3933
T
Th
he
er
ra
an
no
os
st
ti
ic
cs
s
2017; 7(16): 3933-3947. doi: 10.7150/thno.21529
Review
Nucleic Acid-Based Theranostics for Tackling
Alzheimer’s Disease
Madhuri Chakravarthy1, 2, Suxiang Chen1, 2, Peter R. Dodd3, Rakesh N. Veedu1, 2, 3
1. Centre for Comparative Genomics, Murdoch University, Murdoch, Perth, Australia 6150;
2. Perron Institute for Neurological and Translational Science, QEII Medical Centre, Nedlands, Perth, Australia 6005;
3. School of Chemistry and Molecular Biosciences, The University of Queensland, St Lucia, Brisbane, Australia 4072.
Corresponding author: Rakesh N. Veedu, PhD. Centre for Comparative Genomics, Murdoch University, Building 390 Discovery Drive, Perth, Western
Australia, Australia 6150. Email: R.Veedu@murdoch.edu.au
© Ivyspring International Publisher. This is an open access article distributed under the terms of the Creative Commons Attribution (CC BY-NC) license
(https://creativecommons.org/licenses/by-nc/4.0/). See http://ivyspring.com/terms for full terms and conditions.
Received: 2017.06.20; Accepted: 2017.07.28; Published: 2017.09.05
Abstract
Nucleic acid-based technologies have received significant interest in recent years as novel
theranostic strategies for various diseases. The approval by the United States Food and Drug
Administration (FDA) of Nusinersen, an antisense oligonucleotide drug, for the treatment of spinal
muscular dystrophy highlights the potential of nucleic acids to treat neurological diseases, including
Alzheimer’s disease (AD). AD is a devastating neurodegenerative disease characterized by
progressive impairment of cognitive function and behavior. It is the most common form of
dementia; it affects more than 20% of people over 65 years of age and leads to death 715 years
after diagnosis. Intervention with novel agents addressing the underlying molecular causes is
critical. Here we provide a comprehensive review on recent developments in nucleic acid-based
theranostic strategies to diagnose and treat AD.
Key words: nucleic acids; Alzheimer’s disease; amyloid beta peptides; tau peptide; chemically modified
oligonucleotides; nucleic acid therapeutics.
Introduction
Nucleic acid-based technologies typically use
synthetic oligonucleotides ̴8–50 nucleotides in length,
most of which bind to RNA through Watson-Crick
base pairing to alter the expression of the targeted
RNA and protein. Novel chemical modifications and
conjugation strategies have been developed to
improve pharmacokinetics and tissue-specific
delivery. Vitravene, Kynamro, Nusinersen and
Eteplirsen are antisense oligonucleotides (AOs)
approved by the FDA to treat cytomegalovirus
retinitis, familial hypercholesterolemia, spinal
muscular atrophy, and Duchenne muscular
dystrophy respectively [1-3]. The nucleic acid aptamer
drug Macugen was approved for age-related macular
degeneration [4]. These successful clinical translations
demonstrate the potential of nucleic acid-based
technologies and provide scope for developing novel
therapeutics for AD. AD is the most common form of
dementia; it accounts for 70% of cases with that
diagnosis. Globally there are ~47 million current
cases; 7.7 million new cases are added each year [5].
AD is characterized by a progressive loss of memory
and cognitive function [6]. Patients eventually need
24-hour care that places emotional and economic
burdens on the community. There is no cure for AD,
nor any treatment that addresses its underlying
molecular cause [5]. Current treatments use
cholinesterase inhibitors [7] and N-methyl-D-aspartate
receptor (NMDA) antagonists [8] that improve
cognitive function and reduce symptoms temporarily
but do not stop the progression of the disease. The
current approach to diagnosis relies on a combination
of cognitive and clinical assessment, genetic profiling,
and magnetic resonance imaging to measure
anatomical changes in the brain [9], but confirmation
relies on post-mortem neuropathological assessment
Ivyspring
International Publisher
Theranostics 2017, Vol. 7, Issue 16
http://www.thno.org
3934
and misdiagnosis is common [6]. Two hallmarks of
the disease are extracellular amyloid-β (Aβ) plaques
(mainly an agglomeration of peptides) and
intracellular neurofibrillary tangles (hyperpho-
sphorylated tau peptides). In this review we focus on
the potential of nucleic acid therapeutic, diagnostic,
and research strategies that target both Aβ and tau
pathologies to help diagnose and treat AD.
Amyloid β (Aβ) hypothesis
The hypothesis states that there is an
imbalance of toxic peptide production and
clearance [10-12]. The main Aβ species, Aβ
1-40 and
1-42, can aggregate to form fibrils and plaques
[10-12]. Aβ1-40 and 1-42 are produced by the aberrant
splicing of amyloid precursor protein (APP) by β-site
APP cleaving enzyme 1 (BACE1) and γ-secretase
(Figure 1) [11-15]. Mutations in the APP and
Presenilin genes (PSEN1 codes for the catalytic
subunits of γ-secretase) increase Aβ1-42 levels [10-12,
14, 16-18] and lead to early-onset familial AD. Down
syndrome cases have an extra copy of chromosome
23, and hence of the APP gene, and develop Aβ
plaques early in adulthood [19]. Oligomers of
promote synaptic loss, neuronal dysfunction, and cell
death [20, 21]. 1-42 inhibits the maintenance of
hippocampal long-term potentiation, resulting in
altered memory function [10, 22] and reduced
synaptic neurotransmission through NMDA
receptor-mediated signaling [10, 22, 23]. Aβ toxicity
has also been implicated in inflammation [11],
oxidative stress [11, 24], cholinergic transmission [23],
glucose metabolism [25, 26], and cholesterol
metabolism [27].
Tau hypothesis
Microtubule-associated protein tau (tau),
predominantly expressed in neuronal axons, is
involved in microtubule assembly and stability. Tau is
regulated by phosphorylation [28, 29].
Hyperphosphorylation decreases the ability of tau to
bind to microtubules, leading to reduced trafficking,
destabilization of microtubules, and synaptic loss [29,
30] (Figure 2). Abnormal tau can aggregate into paired
helical filaments to form neurofibrillary tangles [31] in
the cytosol and sequester normal tau to inhibit
microtubule assembly [29]. Alternatively, tau
aggregation may be a protective mechanism to stop
hyperphosphorylated tau sequestering normal tau
and inhibit microtubule assembly [29]. Tau
hyperphosphorylation is detrimental in various
neurodegenerative diseases termed “tauopathies” [28,
32]. Hyperphosphorylation of tau correlates with
neurodegeneration and cognitive decline [29, 32].
Other post-translational modifications of tau,
including abnormal glycosylation and reduced
β-linked acylation of N-acetylglucosamine, increase
hyperphosphorylation [29, 33]. Inhibition of the
ubiquitin-proteasome system may also increase the
aggregation of hyperphosphorylated tau [31].
Other hypothesis of AD
Drugs currently approved by the FDA for the
treatment of AD are Donepezil, Rivastigmine,
Galantamine and Memantine (Table 1) [34-37]. These
agents enhance cholinergic and glutamatergic
neurotransmission and improve cognitive function
temporarily. However, they do not slow the
progression of the disease. Oxidative stress [38],
inflammation [39], insulin impairment [40, 41] and
abnormal cholesterol metabolism [27] may also play
roles (Table 1), but will not be considered in depth
here.
Figure 1. Non-amyloidogenic and amyloidogenic pathways in AD neurons. In the amyloidogenic pathway the APP is aberrantly spliced by BACE1 and γ-secretase to
overproduce toxic Aβ species.
Theranostics 2017, Vol. 7, Issue 16
http://www.thno.org
3935
Figure 2. The roles of tau in normal neurons and of hyperphosphorylation in AD neurons that lead to neuronal toxicity.
Table 1. Therapeutic molecules in clinical trials, their targets, and trial outcomes.
Drug molecule
Role/ Target
Trial stage
Results
Donepezil (Pfizer)
Cholinesterase inhibitor
FDA approved- Although they improve the symptoms temporarily these
drugs do not stop the progression of the disease.
Rivastigmine (Novartis)
Cholinesterase inhibitor
Galantamine (Jansen-Cilag)
Cholinesterase inhibitor
Memantine (Lundbeck)
NMDA receptor antagonist
Tramiprosate
Aβ aggregation inhibitor
Phase III
No significant benefit. May promote abnormal tau
aggregation
Colostrinin
Aβ aggregation inhibitor
Phase III
Modest improvements not sustained
Scyllo-inositol
Stabilizes Aβ aggregates and
inhibits toxicity
Phase II
No statistically significant effect. Reduced Aβ in
cerebrospinal fluid
Aβ vaccination
Aβ aggregation inhibitor
Phase II
Halted because patients developed
meningo-encephalitis
Bapineuzumab
Aβ aggregation inhibitor
Phase III
End points not significantly different
Solanezumab
Aβ aggregation inhibitor
Phase III
End points not significantly improved
Anti-amyloid Ab
Aβ aggregation inhibitor
Phase III
No positive primary outcome
Other mAbs
Aβ aggregation inhibitor
Various
No positive outcome
Tarenflurbil
γ-secretase inhibitor
Phase III
No significant improvement
LY450139 (Eli Lilly)
γ-secretase inhibitor
Phase III
Discontinued: no Aβ40/42 reduction
BMS-708163 (B-M Squibb)
γ-secretase inhibitor
Phase II
Terminated due to lack of favorable pharmacodynamics
Verubecestat
BACE1 inhibitor
Phase III
Currently running
Rogiglitazone
BACE1 inhibitor and Type 2
diabetes drug
Phase III
No positive outcome
Pioglitaozone
BACE1 inhibitor and Type 2
diabetes drug
Phase III
No positive outcome
Methyl thionium chloride
Tau aggregation inhibitor
Phase II
Significantly improved cognitive function
Tideglusib
GSK3β
Phase IIb
No positive outcome
Davunetide
Microtubule stabilizer
Phase III
No significant improvement
Antioxidants
ROS
Phase III
No positive outcome
Anti-inflammatories
Inflammation
Phase III
No significant improvement
Intranasal insulin
Insulin impairment
Pilot
Improvement in patients without APOE-ε4 allele
Other anti-diabetics
Insulin impairment
Phase III
Currently running
Statins
Cholesterol metabolism
Phase III
Preliminary results positive; mechanism unknown.
Theranostics 2017, Vol. 7, Issue 16
http://www.thno.org
3936
Current therapeutic molecules and
clinical trials for the treatment of AD
Many disease-modifying therapeutics show
positive outcomes in animal models but
disappointing results in clinical trials (for drug
candidates and ongoing trials see Table 1). Current
strategies have been comprehensively reviewed [42].
Poor outcomes might have arisen because each agent
is targeting a single pathway, whereas AD is a
complex disease and it may be important to aim at
multiple targets [43, 44]. Success in developing a
suitable therapeutic approach is challenging because
the pathogenesis of AD is unknown [45]. Trials might
be affected by factors such as genetics, metabolism,
and diet [46], but there is clearly a need to develop
novel therapeutics for this disease.
Nucleic-acid based molecules for tackling
AD
Unlike conventional small-molecule drugs,
nucleic acid-based therapeutic agents such as AOs,
small-interfering RNAs (siRNAs), microRNA
moieties that target oligonucleotides (antimiRs and
miRNA mimics), and DNAzymes/ribozymes can
regulate the expression of key proteins by selectively
targeting their mRNAs. The outcome is mRNA
cleavage, repair, or steric blockade (Figure 3). The
class of modified nucleic acids called aptamers can
target proteins and inhibit their function (Figure 3).
Nucleic acid-based strategies could be an effective
alternative to drug development for AD because they
can target a range of pathological features.
Figure 3. Nucleic acid-based therapeutic strategies. mRNA: messenger RNA; RNase H: ribonuclease H; siRNA: small interfering RNA; RISC: RNA inducing silencing
complex; AO: antisense oligonucleotide; antimiR: anti-microRNA; miRNA mimic: microRNA mimic
Theranostics 2017, Vol. 7, Issue 16
http://www.thno.org
3937
Figure 4. Examples of chemically-modified nucleotide analogues. 2’-OMe: 2’-O-methyl; 2’-MOE:2’-O-methoxyethyl; 2’-F: 2’-fluoro; 2’-NH2: 2’-amino; FANA:
fluoroarabinonucleotide; LNA: locked nucleic acid; TNA: threose nucleic acid; PNA: peptide nucleic acid; PMO: phosphorodiamidate morpholino oligomer; MNA:
morpholino nucleic acid; HNA: hexitol nucleic acid; CeNA: cyclohexenyl nucleic acid; ANA: anhydrohexitol nucleic acid
Improving the stability and efficacy of nucleic
acid-based therapeutics
Therapeutic oligonucleotides composed of
naturally occurring nucleotides are rapidly degraded
in vivo, which makes them unsuitable for drug
development. To improve their pharmacokinetic
properties, chemically modified nucleotide analogues
with high resistance to nucleases are normally used. A
number of analogues have been developed by
modifying the base or sugar moieties, or the
inter-nucleotide linkages (see Figure 4) [76-78].
Phosphorothioate DNA [79], 2'-O-methyl (2’-OMe)
RNA [80], 2'-fluoro (2’-F) RNA [81],
2'-O-methoxyethyl (2’-MOE) RNA [82], and
phosphorodiamidate morpholino (PMO) [83]
analogues have been successfully utilized in
FDA-approved oligonucleotide drugs. Analogues
such as locked nucleic acids (LNA) [84, 85], peptide
nucleic acids (PNA) [86], tricyclo-DNA (tcDNA) [87],
and cyclohexenyl nucleic acids (CeNA) [88] also show
excellent biophysical properties and offer further
scope for novel oligonucleotide development. These
chemistries can be used to construct fully modified or
mixmer oligonucleotides. Aptamers can be modified
during the selection or post-selection stages to
improve their affinity and bioavailability [77].
Another challenge in the clinical utilization of
unmodified oligonucleotides is rapid renal clearance
from the blood due to their small size that falls under
the renal filtration threshold. To increase their
bioavailability, oligonucleotides can be conjugated
with polyethylene glycol (PEG) to increase their size,
which could also improve their resistance to nucleases
[89]. Several PEGylated drugs have been approved by
the FDA for clinical use [89, 90]. Other strategies
include conjugating the oligonucleotides to albumin,
which has a size of around 7 nm and shows reduced
renal clearance and can therefore increase the
circulation half-life of the oligonucleotides.
Phosphorothioate modified oligonucleotides also
showed reduced renal clearance by binding to plasma
proteins like albumin to avoid glomerular filtration
[91]. Another strategy is the synthesis of neutral
siRNA (masking the negative charge on the
phosphate backbone). Neutral siRNA showed
reduced renal clearance [92].
Recent progress in modified nucleic acids
for AD
Antisense oligonucleotides
A classical nucleic acid approach to controlling
the expression of proteins is to use AOs, short
single-stranded synthetic oligonucleotides, which can
precisely target an mRNA transcript to regulate the
Theranostics 2017, Vol. 7, Issue 16
http://www.thno.org
3938
expression of the protein it codes for. Antisense
mechanisms include RNase H recruitment and
cleavage of mRNA, modulation of splicing in
pre-mRNA, and steric blockade of either mature or
pre-mRNAs (Figure 3). RNase H-mediated cleavage
involves designing a short DNA oligonucleotide that
binds to the target mRNA to form a RNA-DNA
duplex [93]. The duplex is recognized and cleaved by
endogenous RNase H. AOs that modulate pre-mRNA
splicing can be used to repair defective RNA and
eliminate disease-associated splice variants [94].
Many pre-mRNA transcripts are alternatively spliced
to produce different mRNA, and hence protein,
variants [94].
APP
Many groups have designed AOs that target
APP to reduce APP expression. An early study by
Allinquant et al. [95] developed AOs that successfully
blocked rat APP synthesis. Administration of the AOs
showed that APP played a role in axonal and
dendritic growth, and thus in neuronal differentiation
[95]. ISIS Pharmaceuticals (now Ionis Pharma) have
patented (US 2003/0232435 A1) 78 gapmer AOs with
2'-MOE wings and central DNA region. The AOs
target various regions of APP mRNA and inhibit
3982% of APP protein expression [96].
Kumar and colleagues [97] developed
phosphorothioated DNA AOs against sequences that
correspond to the Aβ region of APP (17-42 amino
acids). Administration of the AOs led to improved
cognitive function in senescence-accelerated
mouse-prone 8 (SAMP8) mice. SAMP8 mice have a
natural mutation that leads to APP over-expression,
impaired removal, and loss of memory with
increasing age. The AOs that target the mid-Aβ region
reduced APP levels by 4368% in the amygdala,
septum and hippocampus [97]. The mice showed
improvement in acquisition and retention in the
footshock avoidance paradigm, which reversed their
deficits in learning and memory [97]. AOs that target
the sequences that correspond to the region of APP
coding for the first 1730 amino acids of Aβ were the
subject of intellectual property protection [98]. Banks
and colleagues [99] showed that a radioactively
tagged phosphorothioate DNA AO targeting the
region of APP could transit the blood-brain barrier
(BBB) of mice to enter the cerebrospinal fluid. When a
100-fold higher dose of the AO was injected into the
brain by intracerebroventricular injection it reversed
the learning and memory deficits in SAMP8 mice,
possibly through reduced oxidative stress. Poon et al.
[100] used proteomics to show that lower Aβ levels
result in reduced oxidative stress in brain.
Opazo et al. [101] transfected the AOs described
by Kumar and colleagues [97] into the CTb cell line, a
neuronal line from mice that overexpresses APP, and
the CNh cell line from normal mice. The AOs resulted
in APP knockdown in CTb cells by 36%, 40% and 50%
compared with normal CNh cells after 24 h, 48 h and
72 h respectively [101]. By 72 h after AO transfection,
choline uptake was similar to that in CNh cells and
there was increased choline release in response to
glutamate, nicotine and KCl depolarization, which
reached similar levels to those observed in CNh cells.
The CTb cells come from a Down syndrome mouse
model, which show some learning deficits and
cholinergic dysfunction that are similar to those found
in AD [102]. Similarly, Rojas et al. [103] showed that
APP overexpression reduced the expression and
retrograde transport of nerve growth factor. This
reduced nicotine-induced stimulation of α3β2 nicotinic
acetylcholine receptor and in consequence lowered
intracellular Ca2+ responses in CTb cells. The effects of
APP overexpression were restored close to normal by
treatment with AOs targeting APP expression.
Chauhan and colleagues [104] designed gapmer
AOs with 2'-OMe and DNA nucleotides on a
phosphorothioate backbone that target the β-secretase
cleavage site of APP and found that they reduced
brain 40 and 42 levels in a mouse model of AD.
The AOs were delivered intracerebroventricularly
and showed rapid uptake and retention for 30
minutes. They efficiently crossed cell membranes into
the nuclear and cytoplasmic compartments of
neuronal and non-neuronal cells. Chauhan and Siegel
[105] designed two additional AOs targeting the β-
and γ-secretase site of APP in the Tg2576 mouse
model that expresses APP. The AO targeting the
mutated β-secretase site increased soluble APPα by
43% and decreased soluble 40 and 42 levels by
39%, whereas the AO targeting the γ-secretase site
had no effect. The AO targeting β-secretase also
inhibited acetylcholinesterase activity, increasing
acetylcholine by five-fold in cortex compared with
controls.
Erickson et al. [106] peripherally administered an
APP AO to SAMP8 mice. This resulted in a 30%
increase in APP levels but no change in soluble
levels. The treated mice showed improved memory.
They also showed [107] that AO-mediated APP
knockdown in Tg2576 mouse brains reduced cytokine
expression and improved learning and memory.
Attenuating APP overexpression may improve
learning and memory by reducing inflammation (also
implicated in AD pathology).
BACE1
Yan et al. [108] developed two AOs that target
β-secretase aspartyl protease and found that they
Theranostics 2017, Vol. 7, Issue 16
http://www.thno.org
3939
reduced the release of 40 and Aβ42 by 5080%.
Vassar and colleagues [109] also used AOs that target
β-secretase to reduce Aβ40 and Aβ42 production by
around 30%. These studies showed that β-secretase is
important for the production of 40 and Aβ42 and
highlighted BACE1 as an important target for AD.
Wolfe et al. [110] designed splice-modulating AOs to
target BACE1, since alternatively spliced transcript
variants at exons 2 and 3 do not show β-secretase
activity. The AOs reduced production
significantly in cells without altering total BACE1
mRNA.
Presenilin 1 (PSEN1)
Refolo et al. [111] found that AOs targeting
PSEN1 in a human cell line reduced PSEN1
holoprotein by 80% 12 days after treatment and by
90% after 14 days. This was correlated with a two-fold
increase in 42 levels. Grilli et al. [112] found that
hippocampal primary neurons overexpressing
mutant PSEN1 were vulnerable to excitotoxic and
hypoxia-hypoglycemic damage and increased cell
death. They designed two phosphorothioate AOs
targeting PSEN1 in wild type mice. In contrast to
Refolo et al. [111] they found that lower PSEN1
expression reduced cell death and provided
neuroprotection [112]. Fiorini et al. [113] administered
AOs targeting PSEN1 to aged SAMP8 mice and found
they reduced brain oxidative stress biomarkers. In the
T-maze foot shock avoidance and novel object
recognition tests the mice showed a reversal of
learning and memory deficits.
Tau
Caceres et al. [114] showed that an AO targeting
the 5' end of the tau gene, in the region before the start
codon, showed strong inhibition of neurite elongation
in primary rat neurons. Immunoblotting revealed that
the tau protein level was reduced in AO-treated mice
but not in control mice. The effect of AO treatment on
cognition needs to be assessed. DeVos et al. [115]
screened 80 AOs targeting tau and selected the three
that showed the best knockdown of tau to test in vivo.
The latter reduced tau mRNA levels by more than
75%. The best AO was selected for further testing in
mice; it lowered brain tau mRNA and protein
significantly in a dose-dependent manner. Behavioral
impacts and neurotoxicity were not measured.
Kalbfuss et al. [116] developed splice-modulating AOs
modified with 2'-OMe nucleotides to target the tau
exon 10 splice junctions to reduce exon 10 inclusion.
Exclusion of exon 10 increases the ratio of tau proteins
lacking the microtubule-binding domain. In
consequence, the microtubule cytoskeleton becomes
destabilized as observed in frontotemporal dementia
and parkinsonism.
Peacey et al. [117] designed bipartite AOs that
bound to the hairpin structure at the boundary
between exon 10 and intron 10 of tau to inhibit exon
10 splicing, reversing the effect of disease-causing
mutations in cells. Liu et al. [118] developed a
small-molecule (mitoxantrone) conjugated to a
bipartite AO that binds to the tau RNA hairpin
structure. The conjugate also inhibited exon 10
splicing in cell-free conditions more effectively than
mitoxantrone or the bipartite AO alone, but induced
cytotoxicity. The same group used a PNA-modified
bipartite AO conjugated to mitoxantrone that
inhibited tau splicing but was also cytotoxic [110].
Sud et al. [119] developed PMO AOs to modulate
the splicing of tau and tau expression. The AOs were
designed to target sequences at the donor and
acceptor splice sites, the splicing branch points, and
splicing enhancers and inhibitors to induce exon
skipping. Exons 0, 1, 4, 5, 7, 9, and 10 were targeted.
Exons 1, 4, 5, 7, and 9 are found in all 6 isoforms of tau
while exon 10 is present in only three of the six
isoforms. Of the 31 AOs tested, AO E1.4 targeting the
splice donor site at the exon 1 intron 1 junction
reduced tau mRNA expression by 50%. The other
AOs effective in this region were a combination of
AOs that targeted the splice donor and acceptor sites
and the start codon. AO E5.3 targeted the splice donor
site at the exon 5 intron 5 junction and reduced total
tau mRNA expression by 2946%. It also reduced tau
protein level by 5862%. The resulting transcript was
missing exons 4 to exon 7 using the normal splice
sites. AO E7.7 targeted the splice donor site on exon 7;
it reduced tau mRNA expression by 30% and tau
protein levels by 67%. E5.3 injected into mice in vivo
produced lower tau mRNA levels than in non-injected
regions.
GSK-β
Farr et al. [120] showed that a phosphorothioated
AO that targets GSK-3β decreased GSK-3β protein
levels in the cortex of SAMP8 mice. There were
improvements in learning and memory, reduced
oxidative stress, increased levels of the antioxidant
transcription factor nuclear factor erythroid-2 related
factor 2, and decreased tau phosphorylation.
Acetylcholinesterase (AChE)
Fu et al. [121] found that AOs against human
AChE mRNA reduced AChE activity in an AD mouse
model after 8 h; the effect lasted till 42 h. Lower
enzymic activity was accompanied by improvement
on behavioral tasks, which showed increased memory
retention and improved water maze performance
(shorter swimming time).
Theranostics 2017, Vol. 7, Issue 16
http://www.thno.org
3940
Apolipoprotein E receptor 2 (ApoER2)
ApoER2 may be a primary risk factor for
late-onset AD [122, 123]. Dysregulation of ApoER2
splicing may result in impaired synaptic homeostasis.
Cerebral injection of mice with AOs targeting the
adjacent introns enhanced exon 19 inclusion, an effect
that persisted for up to 6 months [123]. The mice
showed improvement in -induced cognitive
defects. It was postulated that the AOs bind to the
splicing factor SRF1 to reduce its expression and
increase the inclusion of exon 19, thereby increasing
the level of the active form of ApoER2 to enhance
NMDA receptor phosphorylation.
siRNA
These are short synthetic double stranded RNA
oligonucleotides that target complementary mRNA
and silence gene expression through the assembly of
the RNA-induced silencing complex (RISC) (Figure 3)
[93, 124]. Chemical modifications can be introduced
into the siRNA to increase its stability against
nucleases and increase its selectivity for the target
(Figure 4).
APP
Miller and colleagues [125] found that siRNAs
against the Swedish mutant in APP that causes a
familial form of AD silenced the expression of mutant
alleles. The siRNAs were designed to ensure that they
bound specifically to the mutant alleles and not the
wild-type. The mutation was placed in the central
region of the siRNA duplex to achieve high silencing
efficiency.
BACE1
McSwiggen and colleagues [126] patented 325
siRNAs that target BACE (NCBI ID: NM_012104). The
patent covers sequences with various chemically
modified siRNAs that include 2'-deoxy, 2'-F and
2'-OMe pyrimidine and purine nucleotides,
phosphorothioate internucleotide linkages and
inverted deoxyabasic caps. Four of the siRNAs
reduced BACE expression by 4090% at 25 nM
concentration, but there was no data on whether this
altered Aβ40 and 42 expression. Basi et al. [127]
made a siRNA that reduced the BACE1 mRNA level
by 50% and BACE protein level by more than 90%. It
decreased the secretion of peptide without
affecting BACE2 expression, indicating specificity for
BACE1. Kao et al. [128] also designed siRNAs, where
two of the siRNAs reduced BACE1 mRNA by more
than 90% and Aβ production by 3641%. Pretreatment
of neurons with the siRNA increased neuroprotection
against hydrogen peroxide-induced oxidative stress.
Modarresi et al. [129] injected LNA-modified siRNAs
targeting BACE1 antisense transcripts into the third
ventricle of Tg-19959 mice to downregulate BACE1
and BACE1 antisense transcripts, which led to lower
BACE1 protein levels and less Aβ production and
aggregation in the brain. Notably, Cai et al. [130]
showed that siRNAs targeting BACE1 inhibited it in
mice and increased choroidal neovascularization:
BACE1 is also expressed in the neural retina and in in
vitro and in vivo angiogenesis. Although BACE1
inhibition may be therapeutically beneficial in AD, it
may contribute to retinal pathologies and exacerbate
conditions such as age-related macular degeneration.
Heterogeneous nuclear ribonucleoprotein H
A G-rich region in exon 3 of BACE1 may form a
G-quadruplex structure and recruit a splicing
regulator, heterogeneous nuclear ribonucleoprotein
H, that regulates splicing to increase generation of the
BACE1 501 isoform. Fisette et al. [131] reported that
siRNA and short hairpin RNA candidates that target
heterogeneous nuclear ribonucleoprotein H reduced
its expression and thereby decreased BACE1 501
isoform levels and Aβ production.
AntimiRs and miRNA mimics
miRNAs are short non-coding RNAs that
regulate protein expression post-transcriptionally.
miRNA mimics can modulate RNA and protein
expression by acting like their endogenous miRNA
counterparts. AntimiRs can modulate RNA and
protein expression by inhibiting endogenous miRNA
similar to AOs (Figure 3). miRNAs silence gene
expression by translational repression and/or mRNA
degradation [132, 133]. miRNAs are first transcribed
by RNA polymerases II or III to form long primary
miRNA with a 5’ CAP and a poly(A) tail [133-135].
These are then processed in the nucleus into short
70-nucleotide hairpin structures called precursor
miRNAs (pre-miRNA) by the microprocessor
complex [133-135]. The pre-miRNAs are exported to
the cytoplasm by Exportin 5 and processed by Dicer
into double-stranded miRNA duplexes, which are
approximately 22 nucleotides long [133-135].
BACE1
An endogenous non-coding BACE1 antisense
transcript stabilizes the BACE1 transcript and may
upregulate BACE1 in AD cases. BACE1 antisense
binds to BACE1 at the miR-485-5p binding site and
suppresses BACE1 expression. Faghihi et al. [136]
found that LNA-antimiRs that target miR-485-5p
decreased miRNA-induced suppression of BACE1
and increased BACE1 antisense expression. Hebért et
al. [137] showed that miR-29a/b-1 cluster was
significantly reduced in sporadic AD patients and
correlated with increased BACE1 expression and
Theranostics 2017, Vol. 7, Issue 16
http://www.thno.org
3941
generation, and therefore may be potential targets for
miRNA mimics as a therapeutic strategy for AD.
Tau
Mi34a reduces endogenous tau expression at
both the mRNA and protein level in M17D cells by
binding to the 3' UTR region of tau [110], whereas
miR-34c levels are elevated in the hippocampus of AD
patients and mouse AD models [138]. Wolfe et al. [110]
used LNA antimiRs to inhibit miR-34a,
-34b and -34c and found increased tau expression.
Zovolis et al. [138] found that an antimiR that targets
miR-34c rescued learning in mouse models.
Acetyl-CoA acyl transferase
Acetyl-CoA acyl transferase has a role in lipid
metabolism that has been implicated in the
pathogenesis of AD. Murphy et al. [139] inhibited
Acetyl-CoA acyl transferase using an artificial miRNA
to reduce Aβ plaque burden and improve cognition in
a mouse model of AD. The miRNA also reduced
full-length human APP levels.
Brain-derived neurotropic factor
Brain-derived neurotropic factor regulates
synaptic plasticity and memory and is decreased in
AD brains [140-142], while miR-206 suppresses
brain-derived neurotropic factor levels and memory
function in AD mice [143]. Lee et al. [143] injected an
anti-miR candidate AM-206 that targets miR-206 into
the third ventricle of Tg2576 mice. It increased brain
levels of brain-derived neurotropic factor, enhanced
hippocampal synaptic density, neurogenesis, and
memory. Intranasally administered AM-206 also
reached the brain and had similar effects to the
injected AM-206.
DNAzymes/Ribozymes as therapeutic
candidates for AD
A target RNA can be cleaved to reduce its
expression using catalytic oligonucleotides such as
DNAzymes and ribozymes [144]. The arms of these
enzymes hybridise with the target RNA and cleave
their targets through the catalytic loop in the middle.
DNAzymes and ribozymes cleave the phosphodiester
bond at the purine-pyrimidine or purine-purine
junction (Figure 3).
BACE1
Nawrot et al. [145] designed RNA-cleaving
hammerhead ribozymes that downregulated BACE1
mRNA expression by more than 90% in HEK293 and
SH-SY5Y cells and reduced 40 and Aβ42 production
by more than 80%. They also showed that a
DNAzyme with the 1023 catalytic loop reduced
BACE mRNA expression by 70%. However, whether
the reduced BACE mRNA expression leads to
reduced production is unknown and requires
validation.
Nucleic acid aptamers
Aptamers are short single stranded RNA or
DNA oligonucleotides with unique three-dimensional
structure that bind to targets with high affinity and
specificity. Aptamers can be developed against a
variety of targets ranging from small molecules to
complex proteins over whole cells. Aptamers can be
used for therapeutic, diagnostic (biosensors and
molecular imaging), and targeted drug delivery
applications. They are typically selected from large
DNA and RNA oligonucleotide libraries through a
process called Systematic Evolution of Ligands by
EXponential enrichment (SELEX) [146, 147].
Aβ
Ylera et al. [148] were the first to report novel
RNA aptamers that bound to 1-40 fibrils with high
affinity (2948 nM). Bunka et al. [149] made aptamers
against amyloid-like fibrils from β2-microglobulin.
They bound to the target with high affinity, but also
bound to other amyloid fibrils including, but not
confined to, those found in dialysis-related
amyloidosis patients. Rahimi et al. [150] also
developed RNA aptamers against Aβ fibrils, but these
also interacted with other amyloidogenic proteins by
binding to a common β-sheet motif. They bound to
fibrils with ≥15-fold higher sensitivity than
thioflavin-T, suggesting that aptamers might be
diagnostic tools for AD. Takahashi et al. [151] isolated
two RNA aptamers, N2 and E2, that bound to
monomeric Aβ40 with dissociation constants of 21.6
and 10.9 µM respectively. Though the affinities were
quite low for clinical use, enzyme-linked
immunosorbent assay (ELISA) showed that they
could inhibit aggregation efficiently. When
conjugated to AuNP gold nanoparticles, N2 and E2
bound to both Aβ monomers and oligomers. Mathew
et al. [152] showed that the N2 aptamer conjugated to
curcumin-polymer nanoparticles enhanced binding
to, and disaggregated, amyloid plaques, which were
then cleared by phagocytosis. The study targeted
peripheral amyloid as peripheral organs may also
generate amyloid proteins, which have also been
implicated in AD. Targeting peripheral amyloid is
easier due to the challenges in the brain delivery of
aptamers.
Farrar et al. [153] developed a fluorescently
tagged aptamer that bound to Aβ oligomers in both
AD and transgenic mouse brain tissue. The aptamer
may be useful for imaging, which has diagnostic
implications. Similarly, Babu and colleagues [154]
Theranostics 2017, Vol. 7, Issue 16
http://www.thno.org
3942
developed an aptamer complexed with ruthenium
that binds to, and inhibits the formation of,
oligomers. The aptamer-ruthenium interaction
increases luminescence intensity, which is reduced
when the aptamer binds to monomer or
oligomers.
BACE1
Rentmeister et al. [155] made an RNA aptamer
that binds to the short cytoplasmic tail of BACE1. It is
a good research tool to investigate the biological
function of the cytoplasmic tail without interfering
with BACE1 transport and localization. Liang et al.
[156] developed two DNA aptamers, A1 and A4, that
bind to the extracellular domain of BACE1 with high
affinity (Kd 1569 nM) and specificity. They have
similar affinities to the anti-BACE1 antibody. In vitro,
APP Swedish mutant cells treated with A1 showed
lower 40 and 42 levels than control cells. sAPPβ
expression decreased with A1 treatment compared
with untreated controls.
Tau
Kim et al. [157] used recombinant his-tagged
tau40 to select aptamers from an RNA library through
SELEX. 12 rounds of selection produced a tau-1
aptamer, which represented ~76% of identified
aptamers, that reduced the levels of oligomeric tau (by
~94%) in vitro in a dose-dependent manner. However,
it could not de-oligomerize pre-existing tau oligomers
and had no effect on tau degradation. The aptamer
bound to tau protein and inhibited its
oligomerization, unlike control aptamers. Primary
neurons treated with tau-1 aptamer showed less
cytotoxicity than controls but no difference in
membrane integrity or viability; there was little effect
on normal tau function. Primary rat cortical neurons
administered tau oligomers and treated with tau-1
aptamers showed significantly less oligomeric tau
phosphorylation at Ser199/202 but there was no effect
on monomeric tau. Extracellular tau oligomers also
stress neighboring neurons. Administration of tau
oligomers leads to severe neurotoxicity, which was
reduced by tau-1 aptamer treatment. Tau-1 aptamers
can prevent or reverse cytotoxicity mediated by tau
oligomerization both in a non-neuronal cell line and
in primary rat cortical neurons. Unfortunately, the
tau-1 aptamers isolated by Kim et al. bound only to
one of the six isoforms of tau. Therefore, the effects of
tau-1 aptamers observed in mice may not translate
clinically, because six isoforms are prone to
aggregation and implicated in neurodegeneration. To
be successful clinically, the aptamers must be able to
cross the BBB and the neuronal cell membrane, and
disaggregate the neurofibrillary tangles after binding
[158]. Kim et al. [159] reported a DNA
aptamer-antibody sandwiched to the tau-381 isoform
that detected tau in human plasma at femtomolar
concentrations by surface plasmon resonance.
The ubiquitin-proteasome system
Lee et al. [160] developed an aptamer against
USP14, an enzyme that delays protein degradation by
the ubiquitin-proteasome system. Recombinant
USP14 was incubated with a random RNA library for
SELEX. Three aptamers, USP14-1, USP14-2 and
USP-14-3, were identified, all of which bound to
USP14 with high affinity. USP14-3 showed the
strongest inhibition of deubiquitination, which may
be due to its ability to bind both USP14 and UCH37.
UCH37 is a protein that also slows protein
degradation in the proteasome. The aptamers have
yet to be tested in mice for their effect on tau
oligomerization and degeneration.
Prion protein
Mashima et al. [161] isolated aptamers against
bovine prion protein by SELEX that may have
therapeutic potential in prion diseases and AD.
oligomers bind to the prion protein to block long-term
potentiation. Thus, prion protein may mediate Aβ
oligomer-induced synaptic dysfunction.
Brain delivery of nucleic acid molecules
Receptor-mediated endocytosis or nanoparticle
conjugation strategies
The barrier for any successful drug to treat
neurological diseases is its failure to cross the BBB.
Various approaches have been made to overcome this
issue, and in many instances the drug can be
conjugated with other molecules to improve brain
delivery. Lipophilic molecules under 500 Da can cross
the BBB by simple diffusion. Therapeutic nucleic acids
are typically too large to cross the BBB, although
nanotechnology is now starting to overcome this
problem. This was the subject of a comprehensive
review by Kanwar et al. [162]. Nucleic acids can be
transported through the BBB by receptor-mediate
endocytosis when conjugated to molecules such as
transferrin, insulin, leptin, and insulin-like growth
factor 1: these bind to their receptors on the BBB,
which allows them to cross the BBB. Many studies
have used nucleic acids conjugated to molecules that
target the transferrin receptor. Transferrin-conjugated
nanoparticles, or nanoparticles conjugated to
transferrin receptor antibodies, can transport drugs
across the BBB [163, 164]. An aptamer that targeted
the mouse transferrin receptor allowed a lysosomal
enzyme to enter cells via endocytosis; this can be
applied to drug transport [165]. Two aptamers that
Theranostics 2017, Vol. 7, Issue 16
http://www.thno.org
3943
bound to epithelial cell adhesion molecule and
transferrin receptor were fused together; the product
could bind to cancer cells expressing the epithelial cell
adhesion molecule after crossing the BBB by
transferrin-receptor targeting [166].
Cell-penetrating peptide-based delivery systems
The use of cell-penetrating peptide-based
delivery systems is another approach for transporting
nucleic acids across the BBB. These systems generally
contain between 8 and 30 amino acids. Fluorescein
isothiocyanate-labeled cell-penetrating peptides were
conjugated to a morpholino AO targeting the mutant
ataxia telangiectasia gene and were able to cross the
BBB [167]. Heitz et al. [168] reviewed the development
of cell-penetrating peptides.
Intracerebroventricular infusion
Nucleic acids can also be introduced directly into
the cerebrospinal fluid by intracerebroventricular
infusion. The advantage of this over the use of
targeting molecules is that the drugs are delivered at
therapeutic concentrations quickly. However,
intracerebroventricular infusions are highly invasive
and rely on diffusion of the drugs throughout the
ventricular system. The drugs can then enter the
blood stream, because the cerebrospinal fluid turns
over every 45 hours [169]. Pardridge et al. [169]
described the advantages and disadvantages of
different drug delivery methods to the brain.
Intranasal methods are non-invasive and can deliver
nucleic acids directly to the brain. They have been
used successfully to deliver insulin to AD patients
[170]. Various molecules delivered intranasally have
improved cognitive function in a mouse model of AD
and in clinical trials, as summarized in the review by
Hanson et al. [171]. Many reviews describe how
nose-to-brain delivery occurs, the different drugs that
have been successfully delivered this way, and its
potential for treating neurodegenerative diseases such
as AD [172, 173].
Conclusions and Future Perspectives
Nucleic acid-based approaches offer great
promise for developing novel therapeutics for AD, a
complex neurodegenerative disease with several
pathological features. Confounds include genetic
factors, metabolic disorders including high cholesterol
levels, insulin resistance due to impaired glucose
metabolism, and dysfunction in various molecular
pathways. Existing therapies only treat AD
symptoms, not the underlying molecular causes.
Although many drug molecules have shown success
in cell and animal models, this effect often cannot be
replicated in human trials. There is an unmet need for
better theranostic strategies. The drug Nusinersen,
recently approved by the FDA for spinal muscular
atrophy, shows that nucleic acids have potential for
the treatment of neurological diseases, including AD.
Their efficacy in targeting several pathways that
underlie AD highlights their potential to be
developed as novel therapeutics for AD.
Abbreviations
FDA: Food and Drug Administration; AD:
Alzheimer’s disease; AO: antisense oligonucleotides;
NMDA: N-methyl-D- aspartate receptor; Aβ: amyloid
β; APP: amyloid precursor protein; BACE1: β-site
amyloid precursor protein cleaving enzyme 1; PSEN:
Presenilin; tau: microtubule associated protein;
siRNA: small interfering RNA; antimiR:
anti-microRNA; 2’-OMe: 2’-O-methyl; 2’-MOE:
2’-O-methoxyethyl; 2’-F: 2’-fluoro; LNA: locked
nucleic acids; PNA: peptide nucleic acids; PMO:
phosphorodiamidate morpholino; tcDNA:
tricyclo-DNA; CeNA: cyclohexenyl nucleic acid; PEG:
polyethylene glycol; mRNA: messenger RNA;
SAMP8: senescence- accelerated mouse-prone 8; BBB:
blood brain barrier; AChE: acetylcholinesterase;
ApoER2: Apolipoprotein E receptor 2; RISC: RNA
induced silencing complex; SELEX: systemic
evolution of ligands by exponential enrichment.
Acknowledgements
RNV thanks the funding from the McCusker
Charitable Foundation and Perron Institute for
Neurological Diseases and Translational Science.
RNV and MC thank the funding support from Greg
and Dale Higham.
Competing Interests
The authors have declared that no competing
interest exists.
References
1. Mendell JR, Goemans N, Lowes LP, Alfano LN, Berry K, Shao J, et al.
Longitudinal effect of eteplirsen versus historical control on ambulation in
Duchenne muscular dystrophy. Ann Neurol. 2016; 79: 257–271.
2. Crooke ST, Geary RS. Clinical pharmacological properties of mipomersen
(Kynamro), a second generation antisense inhibitor of apolipoprotein B. Br J
Clin Pharmacol. 2013; 76: 269–276.
3. Vitravene Study Group. A randomized controlled clinical trial of intravitreous
fomivirsen for treatment of newly diagnosed peripheral cytomegalovirus
retinitis in patients with AIDS. Am J Ophthalmol. 2002; 133: 467–474.
4. Macugen Diabetic Retinopathy Study Group. A phase II randomized
double-masked trial of pegaptanib, an anti-vascular endothelial growth factor
aptamer, for diabetic macular edema. Ophthalmology. 2005; 112: 1747–1757.
5. Prince M, Comas-Herrera A, Knapp M, Guerchet M, Karagiannidou M. World
Alzheimer Report 2016: Improving healthcare for people living with
dementia. In: Rees G, editor. World Alzheimer Report. London, UK: London
School of Economics and Political Sciences; 2016:140.
6. [No authors listed]. Alzheimers Association Update. Alzheimers Dement
(Amst). 2015; 11: 104–105.
7. Birks JS. Cholinesterase inhibitors for Alzheimer’s disease. Cochrane Database
Syst Rev. 2006.
8. Olivares D, Deshpande VK, Shi Y, Lahiri DK, Greig NH, Rogers JT, et al.
N-methyl D-aspartate (NMDA) receptor antagonists and memantine
Theranostics 2017, Vol. 7, Issue 16
http://www.thno.org
3944
treatment for Alzheimer’s disease, vascular dementia and Parkinson's disease.
Curr Alzheimer Res. 2012; 9: 746–758.
9. Robinson SW, Fernandes M, Husi H. Current advances in systems and
integrative biology. Comput Struct Biotechnol J. 2014; 11: 35–46.
10. Haass C, Selkoe DJ. Soluble protein oligomers in neurodegeneration: lessons
from the Alzheimer’s amyloid β-peptide. Nat Rev Mol Cell Biol. 2007; 8:
101–112.
11. Selkoe DJ. Alzheimer’s disease: genes, proteins, and therapy. Physiol Rev.
2001; 81: 741–766.
12. Blennow K, de Leon MJ, Zetterberg H. Alzheimer’s disease. The Lancet. 2006;
368: 387–403.
13. Haass C, Selkoe DJ. Cellular processing of beta-amyloid precursor protein and
the genesis of amyloid β-peptide. Cell. 1993; 75: 1039–1042.
14. Selkoe DJ. Translating cell biology into therapeutic advances in Alzheimer’s
disease. Nature. 1999; 399: A23–A31.
15. Wilquet V, De Strooper B. Amyloid-β precursor protein processing in
neurodegeneration. Curr Opin Neurol. 2004; 14: 582–588.
16. Selkoe DJ. Normal and abnormal biology of the β-amyloid precursor protein.
Annu Rev Neurosci. 1994; 17: 489–517.
17. Tanzi RE, Bertram L. Twenty years of the Alzheimer’s disease amyloid
hypothesis: a genetic perspective. Cell. 2005; 120: 545–555.
18. Hardy JA, Higgins GA. Alzheimer’s disease: the amyloid cascade hypothesis.
Science. 1992; 256: 184-185.
19. Lahiri DK, Farlow MR, Greig NH, Sambamurti K. Current drug targets for
Alzheimer’s disease treatment. Drug Dev Res. 2002; 56: 267–281.
20. Palop JJ, Mucke L. Amyloid-β-induced neuronal dysfunction in Alzheimer’s
disease: from synapses toward neural networks. Nat Neurosci. 2010; 13:
812–818.
21. Sindi IA, Tannenberg RK, Dodd PR. Role for the neurexin-neuroligin complex
in Alzheimer’s disease. Neurobiol Aging. 2014; 35: 746–756.
22. Ondrejcak T, Klyubin I, Hu N-W, Barry AE, Cullen WK, Rowan MJ.
Alzheimer’s disease amyloid β-protein and synaptic function.
Neuromolecular Med. 2010; 12: 13–26.
23. Danysz W, Parsons CG. Alzheimers disease, βamyloid, glutamate, NMDA
receptors and memantinesearching for the connections. Br J Pharmacol. 2012;
167: 324352.
24. Schelterns P, Feldman H. Treatment of Alzheimer’s disease; current status and
new perspectives. Lancet Neurol. 2003; 2: 539–547.
25. Ferreira ST, Clarke JR, Bomfim TR, De Felice FG. Inflammation, defective
insulin signaling, and neuronal dysfunction in Alzheimer’s disease.
Alzheimers Dement (Amst). 2014; 10: S76–S83.
26. Hoyer S. Glu cose metabolism and ins ulin receptor signal transduction in
Alzheimer disease. Eur J Pharmacol. 2004; 490: 115–125.
27. Maulik M, Westaway D, Jhamandas J, Kar S. Role of cholesterol in APP
metabolism and its significance in Alzheimer’s disease pathogenesis. Mol
Neurobiol. 2013; 47: 37–63.
28. Lee VM, Goedert M, Trojanowski JQ. Neurodegenerative tauopathies. Annu
Rev Neurosci. 2001; 24: 1121–1159.
29. Iqbal K, Alonso AdC, Chen S, Chohan MO, El-Akkad E, Gong C-X, et al. Tau
pathology in Alzheimer disease and other tauopathies. Biochim Biophys Acta
Mol Bas Dis. 2005; 1739: 198–210.
30. Medina M, Avila J. New perspectives on the role of tau in Alzheimer’s disease.
Implications for therapy. Biochem Pharmacol. 2014; 88: 540–547.
31. Hernandez F, Avila J. Tauopathies. Cell Mol Life Sci. 2007; 64: 2219–2233.
32. Brandt R, Hundelt M, Shahani N. Tau alteration and neuronal degeneration in
tauopathies: mechanisms and models. Biochim Biophys Acta Mol Bas Dis.
2005; 1739: 331–354.
33. Hart GW, Slawson C, Ramirez-Correa G, Lagerlof O. Cross talk between
O-GlcNAcylation and phosphorylation: roles in signaling, transcription, and
chronic disease. Annu Rev Biochem. 2011; 80: 825–858.
34. Feldman H, Gauthier S, Hecker J, Vellas B, Emir B, Mastey V, et al. Efficacy of
donepezil on maintenance of activities of daily living in patients with
moderate to severe Alzheimer’s disease and the effect on caregiver burden. J
Am Geriatr Soc. 2003; 51: 737–744.
35. McKeith I, Del Ser T, Spano P, Emre M, Wesnes K, Anand R, et al. Efficacy of
rivastigmine in dementia with Lewy bodies: a randomised, double-blind,
placebo-controlled international study. The Lancet. 2000; 356: 2031–2036.
36. Raskind M, Peskind E, Wessel T, Yuan W, Galantamine USA Study Group.
Galantamine in AD A 6-month randomized, placebo-controlled trial with a
6-month extension. Neurology. 2000; 54: 2261–2268.
37. Tariot PN, Farlow MR, Grossberg GT, Graham SM, McDonald S, Gergel I, et
al. Memantine treatment in patients with moderate to severe Alzheimer
disease already receiving donepezil: a randomized controlled trial. JAMA.
2004; 291: 317–324.
38. Wang X, Wang W, Li L, Perry G, Lee H-g, Zhu X. Oxidative stress and
mitochondrial dysfunction in Alzheimer’s disease. Biochim Biophys Acta Mol
Bas Dis. 2014; 1842: 1240–1247.
39. Pimplikar SW. Neuroinflammation in Alzheimer’s disease: from pathogenesis
to a therapeutic target. J Clin Immunol. 2014; 34: 64–69.
40. Freiherr J, Hallschmid M, Frey II WH, Brünner YF, Chapman CD, Hölscher C,
et al. Intranasal insulin as a treatment for Alzheimer’s disease: a review of
basic research and clinical evidence. CNS Drugs. 2013; 27: 505–514.
41. Ribarič S. The rationale for insulin therapy in Alzheimer’s disease. Molecules.
2016; 21: E689.
42. Galimberti D, Scarpini E. Progress in Alzheimer’s disease. J Neurol. 2012; 259:
201–211.
43. Mangialasche F, Solomon A, Winblad B, Mecocci P, Kivipelto M. Alzheimer’s
disease: clinical trials and drug development. Lancet Neurol. 2010; 9: 702–716.
44. Karran E. Current status of vaccination therapies in Alzheimer’s disease. J
Neurochem. 2012; 123: 647–651.
45. St George-Hyslop PH, Morris JC. Will anti-amyloid therapies work for
Alzheimer’s disease? The Lancet. 2008; 372: 180–182.
46. Ballard C, Gauthier S, Corbett A, Brayne C, Aarsland D, Jones E. Alzheimer’s
disease. The Lancet. 2011; 377: 1019–1031.
47. Santa-Maria I, Hernández F, Del Rio J, Moreno FJ, Avila J. Tramiprosate, a
drug of potential interest for the treatment of Alzheimer’s disease, promotes
an abnormal aggregation of tau. Mol Neurodegener. 2007; 2: 17.
48. Aisen PS, Saumier D, Briand R, Laurin J, Gervais F, Tremblay P, et al. A Phase
II study targeting amyloid-beta with 3APS in mild-to-moderate Alzheimer
disease. Neurology. 2006; 67: 1757–1763.
49. Gervais F, Chalifour R, Garceau D, Kong X, Laurin J, McLaughlin R, et al.
Glycosaminoglycan mimetics: a therapeutic approach to cerebral amyloid
angiopathy. Amyloid. 2001; 8 Suppl 1: 28–35.
50. Bilikiewicz A, Gaus W. Colostrinin (a naturally occurring, proline-rich,
polypeptide mixture) in the treatment of Alzheimer’s disease. J Alzheimers
Dis. 2004; 6: 17–26.
51. Leszek J, Inglot AD, Janusz M, Lisowski J, Krukowska K, Georgiades JA.
Colostrinin: a proline-rich polypeptide (PRP) complex isolated from ovine
colostrum for treatment of Alzheimer's disease. A double-blind,
placebo-controlled study. Archivum Immunologiæ et Therapiæ
Experimentalis. 1999; 47: 377–385.
52. Popik P, Bobula B, Janusz M, Lisowski J, Vetulani J. Colostrinin, a polypeptide
isolated from early milk, facilitates learning and memory in rats. Pharmacol
Biochem Behav. 1999; 64: 183–189.
53. Townsend KP, Praticò D. Novel therapeutic opportunities for Alzheimer’s
disease: focus on nonsteroidal anti-inflammatory drugs. FASEB J. 2005; 19:
1592–1601.
54. Gilman S, Koller M, Black RS, Jenkins L, Griffith SG, Fox NC, et al. Clinical
effects of Aβ immunization (AN1792) in patients with AD in an interrupted
trial. Neurology. 2005; 64: 1553–1562.
55. Salloway S, Sperling R, Fox NC, Blennow K, Klunk W, Raskind M, et al. Two
phase 3 trials of bapineuzumab in mild-to-moderate Alzheimer’s disease. N
Engl J Med. 2014; 370: 322–333.
56. Doody RS, Thomas RG, Farlow M, Iwatsubo T, Vellas B, Joffe S, et al. Phase 3
trials of solanezumab for mild-to-moderate Alzheimer’s disease. N Engl J
Med. 2014; 370: 311–321.
57. Relkin N. Clinical trials of intravenous immunoglobulin for Alzheimer’s
disease. J Clin Immunol. 2014; 34: 74–79.
58. Moreth J, Mavoungou C, Schindowski K. Passive anti-amyloid
immunotherapy in Alzheimer’s disease: What are the most promising targets?
Immunity and Ageing. 2013; 10: 18.
59. McGeer PL, McGeer EG. The amyloid cascade-inflammatory hypothesis of
Alzheimer disease: implications for therapy. Acta Neuropathol. 2013; 126:
479–497.
60. Kukar T, Prescott S, Eriksen JL, Holloway V, Murphy MP, Koo EH, et al.
Chronic administration of R-flurbiprofen attenuates learning impairments in
transgenic amyloid precursor protein mice. BMC Neurosci. 2007; 8: 54.
61. Galasko DR, Graff-Radford N, May S, Hendrix S, Cottrell BA, Sagi SA, et al.
Safety, tolerability, pharmacokinetics, and levels after short-term
administration of R-flurbiprofen in healthy elderly individuals. Alzheimer
Disease and Associated Disorders. 2007; 21: 292–299.
62. Wilcock GK, Black SE, Hendrix SB, Zavitz KH, Swabb EA, Laughlin MA, et al.
Efficacy and safety of tarenflurbil in mild to moderate Alzheimer’s disease: a
randomised phase II trial. The Lancet Neurology. 2008; 7: 483–493.
63. Siemers ER, Quinn JF, Kaye J, Farlow MR, Porsteinsson A, Tariot P, et al.
Effects of a γ-secretase inhibitor in a randomized study of patients with
Alzheimer disease. Neurology. 2006; 66: 602–604.
64. Coric V, Salloway S, van Dyck C, Kerselaers W, Kaplita S, Curtis C, et al. A
phase II study of the γ-secretase inhibitor avagacestat (BMS-708163) in
predementia Alzheimer’s disease. Alzheimers Dement (Amst). 2013; 9: P283.
65. Kennedy ME, Stamford AW, Chen X, Cox K, Cumming JN, Dockendorf MF, et
al. The BACE1 inhibitor verubecestat (MK-8931) reduces CNS β-amyloid in
animal models and in Alzheimer’s disease patients. Science Translational
Medicine. 2016; 8: 363ra150.
66. Miller BW, Willett KC, Desilets AR. Rosiglitazone and pioglitazone for the
treatment of Alzheimer’s disease. Ann Pharmacother. 2011; 45: 1416–1424.
67. Wischik CM, Edwards PC, Lai RY, Roth M, Harrington CR. Selective
inhibition of Alzheimer disease-like tau aggregation by phenothiazines. Proc
Natl Acad Sci USA. 1996; 93: 11213–11218.
68. Martinez A, Gil C, Perez DI. Glycogen synthase kinase 3 inhibitors in the next
horizon for Alzheimer’s disease treatment. Int J Alzheimers Dis. 2011; 2011:
280502.
69. Lovestone S, Boada M, Dubois B, Hüll M, Rinne JO, Huppertz H-J, et al. A
phase II trial of tideglusib in Alzheimer’s disease. J Alzheimers Dis. 2015; 45:
75–88.
70. Mecocci P, Polidori MC. Antioxidant clinical trials in mild cognitive
impairment and Alzheimer’s disease. Biochim Biophys Acta Mol Bas Dis.
2012; 1822: 631–638.
Theranostics 2017, Vol. 7, Issue 16
http://www.thno.org
3945
71. Ray B, Lahiri DK. Neuroinflammation in Alzheimer’s disease: different
molecular targets and potential therapeutic agents including curcumin. Curr
Opin Pharmacol. 2009; 9: 434–444.
72. Wyss-Coray T. Inflammation in Alzhe imer disease: driving force, bystander or
beneficial response? Nat Med. 2006; 12: 1005–1015.
73. Heneka MT, Carson MJ, El Khoury J, Landreth GE, Brosseron F, Feinstein DL,
et al. Neuroinflammation in Alzheimer’s disease. Lancet Neurol. 2015; 14:
388–405.
74. Yarchoan M, Arnold SE. Repurposing diabetes drugs for brain insulin
resistance in Alzheimer disease. Diabetes. 2014; 63: 2253–2261.
75. Casserly I, Topol EJ. Convergence of atherosclerosis and Alzheimer's disease:
inflammation, cholesterol, and misfolded proteins. The Lancet. 2004; 363:
1139–1146.
76. De Mesmaeker A, Haener R, Martin P, Moser HE. Antisense oligonucleotides.
Acc Chem Res. 1995; 28: 366–374.
77. Lipi F, Chen S, Chakravarthy M, Rakesh S, Veedu RN. In vitro evolution of
chemically-modified nucleic acid aptamers: pros and cons, and
comprehensive selection strategies. RNA Biol. 2016; 13: 1232-1245.
78. Chan JH, Lim S, Wong W. Antisense oligonucleotides: from design to
therapeutic application. Clin Exp Pharmacol Physiol. 2006; 33: 533–540.
79. Vu H, Hirschbein BL. Internucleotide phosphite sulfurization with
tetraethylthiuram disulfide. Phosphorothioate oligonucleotide synthesis via
phosphoramidite chemistry. Tetrahedron Lett. 1991; 32: 3005-3008.
80. Majlessi M, Nelson NC, Becker MM. Advantages of 2 -O-methyl
oligoribonucleotide probes for detecting RNA targets. Nucleic Acids Res.
1998; 26: 2224-2229.
81. Kawasaki AM, Casper MD, Freier SM, Lesnik EA, Zounes MC, Cummins LL,
et al. Uniformly modified 2'-deoxy-2'-fluoro-phosphorothioate
oligonucleotides as nuclease-resistant antisense compounds with high affinity
and specificity for RNA targets. J Med Chem. 1993; 36: 831-841.
82. Geary RS, Watanabe TA, Truong L, Freier S, Lesnik EA, Sioufi NB, et al.
Pharmacokinetic properties of 2 -O-(2-methoxyethyl)-modified
oligonucleotide analogs in rats. J Pharmacol Exp Ther. 2001; 296: 890-897.
83. Summerton J, WELLER D. Morpholino antisense oligomers: design,
preparation, and properties. Antisense Nucleic Acid Drug Dev. 1997; 7:
187-195.
84. Veedu RN, Wengel J. Locked nucleic acids: promising nucleic acid analogs for
therapeutic applications. Chem Biodivers. 2010; 7: 536-542.
85. Veedu RN, Wengel J. Locked nucleic acid as a novel class of therapeutic
agents. RNA Biol. 2009; 6: 321-323.
86. Hyrup B, Nielsen PE. Peptide nucleic acids (PNA): synthesis, properties and
potential applications. Bioorg Med Chem. 1996; 4: 5-23.
87. Renneberg D, Leumann CJ. Watson Crick base-pairing properties of
tricyclo-DNA. J Am Chem Soc. 2002; 124: 5993-6002.
88. Wang J, Verbeure B, Luyten I, Lescrinier E, Froeyen M, Hendrix C, et al.
Cyclohexene nucleic acids (CeNA): serum stable oligonucleotides that activate
RNase H and increase duplex stability with complementary RNA. J Am Chem
Soc. 2000; 122: 8595-8602.
89. Ikeda Y, Nagasaki Y. Impacts of PEGylation on the gene and oligonucleotide
delivery system. J Appl Polym Sci. 2014; 131.
90. Knop K, Hoogenboom R, Fischer D, Schubert US. Poly (ethylene glycol) in
drug delivery: pros and cons as well as potential alternatives. Angew Chem
Int Ed. 2010; 49: 6288-6308.
91. White PJ, Anastasopoulos F, Pouton CW, Boyd BJ. Overcoming biological
barriers to in vivo efficacy of antisense oligonucleotides. Expert Rev Mol Med.
2009; 11.
92. Juliano RL. The delivery of therapeutic oligonucleotides. Nucleic Acids Res.
2016; 44: 6518-6548.
93. Bennett CF, Swayze EE. RNA target ing therapeutics: molecular mechanisms of
antisense oligonucleotides as a therapeutic platform. Annu Rev Pharmacol
Toxicol. 2010; 50: 259–293.
94. Kole R, Krainer AR, Altman S. RNA therapeutics: beyond RNA interference
and antisense oligonucleotides. Nat Rev Drug Discov. 2012; 11: 125–140.
95. Allinquant B, Hantraye P, Mailleux P, Moya K, Bouillot C, Prochiantz A.
Downregulation of amyloid precursor protein inhibits neurite outgrowth in
vitro. J Cell Biol. 1995; 128: 919–927.
96. Dobie K. Antisense modulation of amyloid β protein precursor expression. US:
Scios Inc., Fremont, CA (US); 2007.
97. Kumar VB, Farr SA, Flood JF, Kamlesh V, Franko M, Banks WA, et al.
Site-directed antisense oligonucleotide decreases the expression of amyloid
precursor protein and reverses deficits in learning and memory in aged
SAMP8 mice. Peptides. 2000; 21: 1769–1775.
98. Kumar VB. Antisense modulation of amyloid β protein expression. US: St.
Louis University, St. Louis, MO (US); 2001.
99. Banks WA, Farr SA, Butt W, Kumar VB, Franko MW, Morley JE. Delivery
across the blood-brain barrier of antisense directed against amyloid β: reversal
of learning and memory deficits in mice overexpressing amyloid precursor
protein. J Pharmacol Exp Ther. 2001; 297: 1113–1121.
100. Poon HF, Farr SA, Banks WA, Pierce WM, Klein JB, Morley JE, et al. Proteomic
identification of less oxidized brain proteins in aged senescence-accelerated
mice following administration of antisense oligonucleotide directed at the Aβ
region of amyloid precursor protein. Brain Res Mol Brain Res. 2005; 138: 8–16.
101. Opazo P, Saud K, de Saint Pierre M, Cárdenas AM, Allen DD, SeguraAguilar
J, et al. Knockdown of amyloid precursor protein normalizes cholinergic
function in a cell line derived from the cerebral cortex of a trisomy 16 mouse:
an animal model of Down syndrome. J Neurosci Res. 2006; 84: 1303–1310.
102. Saud K, Arriagada C, Cárdenas AM, Shimahara T, Allen DD, Caviedes R, et al.
Neuronal dysfunction in Down syndrome: contribution of neuronal models in
cell culture. Journal of Physiology, Paris. 2006; 99: 201–210.
103. Rojas G, Cárdenas AM, Fernández-Olivares P, Shimahara T, Segura-Aguilar J,
Caviedes R, et al. Effect of the knockdown of amyloid precursor protein on
intracellular calcium increases in a neuronal cell line derived from the cerebral
cortex of a trisomy 16 mouse. Exp Neurol. 2008; 209: 234–242.
104. Chauhan NB. Trafficking of intracerebroventricularly injected antisense
oligonucleotides in the mouse brain. Antisense Nucleic Acid Drug Dev. 2002;
12: 353–357.
105. Chauhan NB, Siegel GJ. Antisense inhibition at the β-secretase-site of
β-amyloid precursor protein reduces cerebral amyloid and acetyl
cholinesterase activity in Tg2576. Neuroscience. 2007; 146: 143–151.
106. Erickson MA, Niehoff ML, Farr SA, Morley JE, Dillman LA, Lynch KM, et al.
Peripheral administration of antisense oligonucleotides targeting the
amyloid-β protein precursor reverses AβPP and LRP-1 overexpression in the
aged SAMP8 mouse brain. J Alzheimers Dis. 2012; 28: 951–960.
107. Erickson M, Farr S, Niehoff M, Morley J, Banks W. Antisense directed against
the amyloid precursor protein reduces cytokine expression in the brain and
improves learning and memory in the Tg2576 mouse model of Alzheimer’s
disease. Brain Behav Immun. 2012; 26: S27.
108. Yan R, Bienkowski MJ, Shuck ME, Miao H, Tory MC, Pauley AM, et al.
Membrane-anchored aspartyl protease with Alzheimer’s disease β-secretase
activity. Nature. 1999; 402: 533–537.
109. Vassar R, Bennett BD, Babu-Khan S, Kahn S, Mendiaz EA, Denis P, et al.
β-Secretase cleavage of Alzheimer’s amyloid precursor protein by the
transmembrane aspartic protease BACE. Science. 1999; 286: 735–741.
110. Wolfe MS. Targeting mRNA for Alzheimer’s and related dementias.
Scientifica. 2014; 2014: 1-13.
111. Refolo LM, Eckman C, Prada CM, Yager D, Sambamurti K, Mehta N, et al.
Antisenseinduced reduction of Presenilin 1 expression selectively increases
the production of amyloid β42 in transfected cells. J Neurochem. 1999; 73:
23832388.
112. Grilli M, Diodato E, Lozza G, Brusa R, Casarini M, Uberti D, et al. Presenilin-1
regulates the neuronal threshold to excitotoxicity both physiologically and
pathologically. Proc Natl Acad Sci USA. 2000; 97: 12822–12827.
113. Fiorini A, Sultana R, Förster S, Perluigi M, Cenini G, Cini C, et al. Antisense
directed against PS-1 gene decreases brain oxidative markers in aged
senescence accelerated mice (SAMP8) and reverses learning and memory
impairment: a proteomics study. Free Radic Biol Med. 2013; 65: 1–14.
114. Caceres A, Kosik KS. Inhibition of neurite polarity by tau antisense
oligonucleotides in primary cerebellar neurons. Nature. 1990; 343: 461–463.
115. DeVos SL, Goncharoff DK, Chen G, Kebodeaux CS, Yamada K, Stewart FR, et
al. Antisense reduction of tau in adult mice protects against seizures. J
Neurosci. 2013; 33: 12887–12897.
116. Kalbfuss B, Mabon SA, Misteli T. Correction of alternative splicing of tau in
frontotemporal dementia and parkinsonism linked to chromosome 17. J Biol
Chem. 2001; 276: 42986–42993.
117. Peacey E, Rodriguez L, Liu Y, Wolfe MS. Targeting a pre-mRNA structure
with bipartite antisense molecules modulates tau alternative splicing. Nucleic
Acids Res. 2012; 40: 9836-9849.
118. Liu Y, Rodriguez L, Wolfe MS. Template-directed synthesis of a small
molecule-antisense conjugate targeting an mRNA structure. Bioorg Chem.
2014; 54: 7–11.
119. Sud R, Geller ET, Schellenberg GD. Antisense-mediated exon skipping
decreases tau protein expression: a potential therapy for tauopathies. Mol Ther
Nucleic Acids. 2014; 3: e180.
120. Farr SA, Ripley JL, Sultana R, Zhang Z, Niehoff ML, Platt TL, et al. Antisense
oligonucleotide against GSK-3β in brain of SAMP8 mice improves learning
and memory and decreases oxidative stress: Involvement of transcription
factor Nrf2 and implications for Alzheimer disease. Free Radic Biol Med. 2014;
67: 387–395.
121. Fu A-L, Zhang X-M, Sun M-J. Antisense inhibition of acetylcholinesterase gene
expression for treating cognition deficit in Alzheimer’s disease model mice.
Brain Res. 2005; 1066: 10–15.
122. Wasser CR, Herz J. Splicing therapeutics for Alzheimer’s disease. EMBO Mol
Med. 2016; 8: 308–310.
123. Hinrich AJ, Jodelka FM, Chang JL, Brutman D, Bruno AM, Briggs CA, et al.
Therapeutic correction of ApoER2 splicing in Alzheimer’s disease mice using
antisense oligonucleotides. EMBO Mol Med. 2016; 8: 328–345.
124. Hannon GJ. RNA interference. Nature. 2002; 418: 244–251.
125. Miller VM, Gouvion CM, Davidson BL, Paulson HL. Targeting Alzheimer’s
disease genes with RNA interference: an efficient strategy for silencing mutant
alleles. Nucleic Acids Res. 2004; 32: 661–668.
126. McSwiggen J. RNA interference mediated treatment of Alzheimer’s disease
using short interfering RNA. Terpstra, Anita, J.; McDonnell Boehnen Hulbert
& Berghoff, Suite 3200, 300 South Wacker Drive, Chicago, IL 60606 (US); 2002.
127. Basi G, Frigon N, Barbour R, Doan T, Gordon G, McConlogue L, et al.
Antagonistic effects of β-site amyloid precursor protein-cleaving enzymes 1
and 2 on β-amyloid peptide production in cells. J Biol Chem. 2003; 278:
31512–31520.
128. Kao S-C, Krichevsky AM, Kosik KS, Tsai L-H. BACE1 suppression by RNA
interference in primary cortical neurons. J Biol Chem. 2004; 279: 1942–1949.
Theranostics 2017, Vol. 7, Issue 16
http://www.thno.org
3946
129. Modarresi F, Faghihi MA, Patel NS, Sahagan BG, Wahlestedt C,
Lopez-Toledano MA. Knockdown of BACE1-AS nonprotein-coding transcript
modulates β-amyloid-related hippocampal neurogenesis. Int J Alzheimers
Dis. 2011; 2011: 929042.
130. Cai J, Qi X, Kociok N, Skosyrski S, Emilio A, Ruan Q, et al. βSecretase
(BACE1) inhibition causes retinal pathology by vascular dysregulation and
accumulation of age pigment. EMBO Mol Med. 2012; 4: 980991.
131. Fisette JF, Montagna DR, Mihailescu MR, Wolfe MS. AGRich element forms a
Gquadruplex and regulates BACE1 mRNA alternative splicing. J Neurochem.
2012; 121: 763–773.
132. Jonas S, Izaurralde E. Towards a molecular understanding of
microRNA-mediated gene silencing. Nat Rev Genet. 2015; 16: 421-433.
133. Ha M, Kim VN. Regulation of microRNA biogenesis. Nat Rev Mol Cell Biol.
2014; 15: 509-524.
134. He L, Hannon GJ. MicroRNAs: small RNAs with a big role in gene regulation.
Nat Rev Genet. 2004; 5: 522-531.
135. Stenvang J, Petri A, Lindow M, Obad S, Kauppinen S. Inhibition of microRNA
function by antimiR oligonucleotides. Silence. 2012; 3: 1.
136. Faghihi MA, Zhang M, Huang J, Modarresi F, Van der Brug MP, Nalls MA, et
al. Evidence for natural antisense transcript-mediated inhibition of microRNA
function. Genome Biol. 2010; 11: R56.
137. Hébert SS, Horré K, Nicolaï L, Papadopoulou AS, Mandemakers W,
Silahtaroglu AN, et al. Loss of microRNA cluster miR-29a/b-1 in sporadic
Alzheimer's disease correlates with increased BACE1/β-secretase expression.
Proc Natl Acad Sci. 2008; 105: 6415-6420.
138. Zovoilis A, Agbemenyah HY, AgisBalboa RC, Stilling RM, Edbauer D, Rao P,
et al. microRNA34c is a novel target to treat dementias. EMBO J. 2011; 30:
42994308.
139. Murphy SR, Chang CC, Dogbevia G, Bryleva EY, Bowen Z, Hasan MT, et al.
ACAT1 knockdown gene therapy decreases amyloid-β in a mouse model of
Alzheimer’s disease. Mol Ther. 2013; 21: 1497–1506.
140. Zuccato C, Cattaneo E. Brain-derived neurotrophic factor in
neurodegenerative diseases. Nat Rev Neurol. 2009; 5: 311-322.
141. Peng S, Wuu J, Mufson EJ, Fahnestock M. Precursor form of brain-derived
neurotrophic factor and mature brain-derived neurotrophic factor are
decreased in the pre-clinical stages of Alzheimer’s disease. J Neurochem. 2005;
93: 1412–1421.
142. Phillips HS, Hains JM, Armanini M, Laramee GR, Johnson SA, Winslow JW.
BDNF mRNA is decreased in the hippocampus of individuals with
Alzheimer’s disease. Neuron. 1991; 7: 695–702.
143. Lee ST, Chu K, Jung KH, Kim JH, Huh JY, Yoon H, et al. miR206 regulates
brainderived neurotrophic factor in Alzheimer disease model. Ann Neurol.
2012; 72: 269–277.
144. Achenbach J, Chiuman W, Cruz R, Li Y. DNAzymes: from creation in vitro to
application in vivo. Curr Pharm Biotechnol. 2004; 5: 321–336.
145. Nawrot B. Targeting BACE with small inhibitory nucleic acids a future for
Alzheimer’s disease therapy? Acta Biochim Pol. 2004; 51: 431–444.
146. Gopinath SCB. Methods developed for SELEX. Anal Bioanal Chem. 2007; 387:
171–182.
147. Stoltenburg R, Reinemann C, Strehlitz B. SELEX a (r) evolutionary method
to generate high-affinity nucleic acid ligands. Biomol Eng. 2007; 24: 381–403.
148. Ylera F, Lurz R, Erdmann VA, Fürste JP. Selection of RNA aptamers to the
Alzheimer’s disease amyloid peptide. Biochem Biophys Res Commun. 2002;
290: 1583–1588.
149. Bunka DH, Mantle BJ, Morten IJ, Tennent GA, Radford SE, Stockley PG.
Production and characterization of RNA aptamers specific for amyloid fibril
epitopes. J Biol Chem. 2007; 282: 34500–34509.
150. Rahimi F, Murakami K, Summers JL, Chen C-HB, Bitan G. RNA aptamers
generated against oligomeric Aβ40 recognize common amyloid aptatopes
with low specificity but high sensitivity. PLoS One. 2009; 4: e7694.
151. Takahashi T, Tada K, Mihara H. RNA aptamers selected against amyloid
β-peptide (Aβ) inhibit the aggregation of Aβ. Mol Biosyst. 2009; 5: 986–991.
152. Mathew A, Aravind A, Brahatheeswaran D, Fukuda T, Nagaoka Y, Hasumura
T, et al. Amyloid-binding aptamer conjugated curcumin–PLGA nanoparticle
for potential use in Alzheimer’s disease. Bionanoscience. 2012; 2: 83–93.
153. Farrar CT, William CM, Hudry E, Hashimoto T, Hyman BT. RNA aptamer
probes as optical imaging agents for the detection of amyloid plaques. PLoS
One. 2014; 9: e89901.
154. Babu E, Mareeswaran PM, Sathish V, Singaravadivel S, Rajagopal S. Sensing
and inhibition of amyloid-β based on the simple luminescent
aptamer–ruthenium complex system. Talanta. 2015; 134: 348–353.
155. Rentmeister A, Bill A, Wahle T, Walter J, Famulok M. RNA aptamers
selectively modulate protein recruitment to the cytoplasmic domain of
β-secretase BACE1 in vitro. RNA. 2006; 12: 1650–1660.
156. Liang H, Shi Y, Kou Z, Peng Y, Chen W, Li X, et al. Inhibition of BACE1
Activity by a DNA Aptamer in an Alzheimer’s Disease Cell Model. PLoS One.
2015; 10: e0140733.
157. Kim JH, Kim E, Choi WH, Lee J, Lee JH, Lee H, et al. Inhibitory RNA aptamers
of tau oligomerization and their neuroprotective roles against proteotoxic
stress. Mol Pharm. 2016; 13: 2039–2048.
158. Tannenberg RK, Lauridsen LH, Kanwar JR, Dodd PR, Veedu RN. Nucleic acid
aptamers as novel class of therapeutics to mitigate Alzheimer’s disease
pathology. Curr Alzheimer Res. 2013; 10: 442–448.
159. Kim S, Wark AW, Lee HJ. Femtomolar detection of tau proteins in undiluted
plasma using surface plasmon resonance. Anal Chem. 2016; 88: 7793–7799.
160. Lee JH, Shin SK, Jiang Y, Choi WH, Hong C, Kim D-E, et al. Facilitated Tau
degradation by USP14 aptamers via enhanced proteasome activity. Sci Rep.
2015; 5: 10757.
161. Mashima T, Matsugami A, Nishikawa F, Nishikawa S, Katahira M. Unique
quadruplex structure and interaction of an RNA aptamer against bovine prion
protein. Nucleic Acids Res. 2009; 37: 6249–6258.
162. Kanwar JR, Sun X, Punj V, Sriramoju B, Mohan RR, Zhou SF, et al.
Nanoparticles in the treatment and diagnosis of neurological disorders:
untamed dragon with fire power to heal. Nanomedicine : Nanotechnology,
Biology, and Medicine. 2012; 8: 399–414.
163. Ulbrich K, Hekmatara T, Herbert E, Kreuter J. Transferrin- and
transferrin-receptor-antibody-modified nanoparticles enable drug delivery
across the bloodbrain barrier (BBB). Eur J Pharm Biopharm. 2009; 71: 251–256.
164. Gan CW, Feng SS. Transferrin-conjugated nanoparticles of
poly(lactide)-D-alpha-tocopheryl polyethylene glycol succinate diblock
copolymer for targeted drug delivery across the blood-brain barrier.
Biomaterials. 2010; 31: 7748–7757.
165. Chen CH, Dellamaggiore KR, Ouellette CP, Sedano CD, Lizadjohry M,
Chernis GA, et al. Aptamer-based endocytosis of a lysosomal enzyme. Proc
Natl Acad Sci USA. 2008; 105: 15908–15913.
166. Macdonald J, Henri J, Goodman L, Xiang D, Duan W, Shigdar S. Development
of a bifunctional aptamer targeting the transferrin receptor and epithelial cell
adhesion molecule (EpCAM) for the treatment of brain cancer metastases.
ACS Chem Neurosci. 2017; 8: 777–784.
167. Du L, Kayali R, Bertoni C, Fike F, Hu H, Iversen PL, et al. Arginine-rich
cell-penetrating peptide dramatically enhances AMO-mediated ATM aberrant
splicing correction and enables delivery to brain and cerebellum. Hum Mol
Genet. 2011; 20: 3151–3160.
168. Heitz F, Morris MC, Divita G. Twenty years of cell-penetrating peptides: from
molecular mechanisms to therapeutics. Br J Pharmacol. 2009; 157: 195–206.
169. Pardridge WM. Drug targeting to the brain. Pharm Res. 2007; 24: 1733–1744.
170. Claxton A, Baker LD, Wilkinson CW, Trittschuh EH, Chapman D, Watson GS,
et al. Sex and ApoE genotype differences in treatment response to two doses of
intranasal insulin in adults with mild cognitive impairment or Alzheimer’s
disease. J Alzheimers Dis. 2013; 35: 789–797.
171. Hanson LR, Frey WH, 2nd. Intranasal delivery bypasses the blood-brain
barrier to target therapeutic agents to the central nervous system and treat
neurodegenerative disease. BMC Neurosci. 2008; 9 Suppl 3: S5.
172. Dhuria SV, Hanson LR, Frey WH, 2nd. Intranasal delivery to the central
nervous system: mechanisms and experimental considerations. J Pharm Sci.
2010; 99: 1654–1673.
173. Lochhead JJ, Thorne RG. Intranasal delivery of biologics to the central nervous
system. Adv Drug Deliv Rev. 2012; 64: 614–628.
Author Biography
Madhuri Chakravarthy
obtained her Bachelor’s (2011)
in Biomedical Science and
Master’s (2014) in Biomedical
Science from the University of
Auckland, New Zealand. She
is currently a PhD student
under the supervision of Dr.
Rakesh N. Veedu at the
Murdoch University, Western Australia. Her research
is centered on developing novel nucleic acid
technologies to tackle neurological diseases.
Suxiang Chen received
his Bachelor degree of
Agriculture in 2010 in South
China Agricultural University.
He then obtained his Master
degree in Biotechnology
(Advanced) and a second
Master degree in Technology
and Innovation Management from the University of
Queensland in 2013 and 2014 respectively. Later, he
worked in Darui Biotechonology Co. Ltd and Gene
Denovo Biotechnology Co. Ltd in Guangzhou, China
as a core technician and technical support team
Theranostics 2017, Vol. 7, Issue 16
http://www.thno.org
3947
member respectively. Since 2016, he has been
pursuing a PhD degree at Murdoch University under
the supervision of Dr. Rakesh N. Veedu. His current
research focuses include the development of novel
RNA targeting therapies for tackling Duchenne
muscular dystrophy and Type-2 Diabetes.
A/Prof. Peter R. Dodd
obtained his PhD from
Imperial College, University
of London, and is currently
the Director of Queensland
Brain Bank at the School of
Chemistry and Molecular
Biosciences, The University of Queensland, Brisbane,
Australia. His lab uses human autopsy tissues to
study amino acid neurotransmission, and has
developed protocols to prepare good-quality mRNA,
miRNA, and proteins for mapping and quantification.
His lab studied phenotype-genotype interactions in
alcoholics and dementia cases, partitioned by sex and
comorbid disease, with key findings including altered
expression of GABA receptors in alcoholics without
comorbid disease and of glutamate receptors in
cirrhotic alcoholics, and altered expression of
glutamate transporters and receptors in Alzheimer
disease. Prof. Dodd’s group was the first to use
microarrays to study the human brain transcriptome.
Overall, his lab researches the pathogenesis of human
neurological diseases.
Rakesh N. Veedu is
currently leading the precision
nucleic acid theranostics group
at Murdoch University and
Perron Institute for Neuro-
logical and Translational
Science. He obtained his PhD
in synthetic organic chemistry in 2006 from The
University of Queensland, Australia under the
supervision of Prof. Curt Wentrup after completing
his MSc from Griffith University, Australia. He then
continued his postdoctoral career under the
supervision of Prof. Jesper Wengel at the Nucleic Acid
Center, University of Southern Denmark in the field
of nucleic acid chemical biology. Later in 2009, he was
appointed as a Research Associate Professor within
the Nucleic Acid Center. He then returned to The
University of Queensland in mid-2010 and established
his independent research career in Functional Nucleic
Acid Theranostics Development. His current research
is focused on developing novel precision nucleic acid
theranostics for tackling solid cancers, neurological
diseases including Alzheimer’s disease and infectious
diseases using nucleic acid aptamers, antisense
oligonucleotides, siRNA, antimiRs, molecular beacons
and DNAzymes.
... Aptamer sequence can be modified to allow tunable recognition of targets. Aptamers have been explored for use as sensing tools for drug monitoring and disease diagnosis, and also for use in therapeutics either through targeted delivery of therapeutic cargo into cells or regulating protein function to influence biological processes [25][26][27]. Although several aptamer-based drugs are now in clinical testing, one, Pegaptanib, is already on the market to treat age-related macular degeneration (AMD). ...
Article
Full-text available
A diverse array of organic and inorganic materials, including nanomaterials, has been extensively employed in multifunctional biomedical applications. These applications encompass drug/gene delivery, tissue engineering, biosensors, photodynamic and photothermal therapy, and combinatorial sciences. Surface and bulk engineering of these materials, by incorporating biomolecules and aptamers, offers several advantages such as decreased cytotoxicity, improved stability, enhanced selectivity/sensitivity toward specific targets, and expanded multifunctional capabilities. In this comprehensive review, we specifically focus on aptamer-modified engineered materials for diverse biomedical applications. We delve into their mechanisms, advantages, and challenges, and provide an in-depth analysis of relevant literature references. This critical evaluation aims to enhance the scientific community's understanding of this field and inspire new ideas for future research endeavors.
... Over the past few years, a large number of theranostic systems have been reported for anti-AD. 3,[15][16][17] Among them, nanosystembased theranostics (nanotheranostics) are the main types in the literature, probably due to the general characteristics of nanosystems, including easy multifunctionalization for achieving diagnostics and therapeutics against various pathological factors of AD simultaneously, and the potential to cross the blood-brain barrier (BBB). 15 Nevertheless, most of the AD nanotheranostics remain at the laboratory level and confront many challenges for clinical translation. ...
Article
Full-text available
Alzheimer's disease (AD) is the most common form of neurodegenerative dementia. As a multifactorial disease, AD involves several etiopathogenic mechanisms, in which multiple pathological factors are interconnected with each other. This complicated and unclear pathogenesis makes AD lack effective diagnosis and treatment. Theranostics, exerting the synergistic effect of diagnostic and therapeutic functions, would provide a promising strategy for exploring AD pathogenesis and developing drugs for combating AD. With the efforts in small drug-like molecules for both diagnosis and treatment of AD, small-molecule-based theranostic agents have attracted significant attention owing to their facile synthesis, high biocompatibility and reproducibility, and easy clearance from the body through the excretion systems. In this review, the small-molecule-based theranostic agents reported in the literature for anti-AD are classified into four groups according to their diagnostic modalities. Their design rationales, chemical structures, and working mechanisms for theranostics are summarized. Finally, the opportunities for small-molecule-based theranostic agents in AD are also proposed.
Article
Full-text available
Background Parkinson's disease (PD) is a degenerative neurological condition marked by the gradual loss of dopaminergic neurons in the substantia nigra pars compacta. The precise etiology of PD remains unclear, but emerging evidence suggests a significant role for disrupted autophagy—a crucial cellular process for maintaining protein and organelle integrity. Methods This review focuses on the role of non‐coding RNAs (ncRNAs) in modulating autophagy in PD. We conducted a comprehensive review of recent studies to explore how ncRNAs influence autophagy and contribute to PD pathophysiology. Special attention was given to the examination of ncRNAs' regulatory impacts in various PD models and patient samples. Results Findings reveal that ncRNAs are pivotal in regulating key processes associated with PD progression, including autophagy, α‐synuclein aggregation, mitochondrial dysfunction, and neuroinflammation. Dysregulation of specific ncRNAs appears to be closely linked to these pathogenic processes. Conclusion ncRNAs hold significant therapeutic potential for addressing autophagy‐related mechanisms in PD. The review highlights innovative therapeutic strategies targeting autophagy‐related ncRNAs and discusses the challenges and prospective directions for developing ncRNA‐based therapies in clinical practice. The insights from this study underline the importance of ncRNAs in the molecular landscape of PD and their potential in novel treatment approaches.
Article
Full-text available
The last decade (2013–2023) has seen unprecedented successes in the clinical translation of therapeutic antisense oligonucleotides (ASOs). Eight such molecules have been granted marketing approval by the United States Food and Drug Administration (US FDA) during the decade, after the first ASO drug, fomivirsen, was approved much earlier, in 1998. Splice-modulating ASOs have also been developed for the therapy of inborn errors of metabolism (IEMs), due to their ability to redirect aberrant splicing caused by mutations, thus recovering the expression of normal transcripts, and correcting the deficiency of functional proteins. The feasibility of treating IEM patients with splice-switching ASOs has been supported by FDA permission (2018) of the first “N-of-1” study of milasen, an investigational ASO drug for Batten disease. Although for IEM, owing to the rarity of individual disease and/or pathogenic mutation, only a low number of patients may be treated by ASOs that specifically suppress the aberrant splicing pattern of mutant precursor mRNA (pre-mRNA), splice-switching ASOs represent superior individualized molecular therapeutics for IEM. In this work, we first summarize the ASO technology with respect to its mechanisms of action, chemical modifications of nucleotides, and rational design of modified oligonucleotides; following that, we precisely provide a review of the current understanding of developing splice-modulating ASO-based therapeutics for IEM. In the concluding section, we suggest potential ways to improve and/or optimize the development of ASOs targeting IEM.
Article
Full-text available
There is a significant need for fast, cost-effective, and highly sensitive protein target detection, particularly in the fields of food, environmental monitoring, and healthcare. The integration of high-affinity aptamers with metal-based nanomaterials has played a crucial role in advancing the development of innovative aptasensors tailored for the precise detection of specific proteins. Aptamers offer several advantages over commonly used molecular recognition methods, such as antibodies. Recently, a variety of metal-based aptasensors have been established. These metallic nanomaterials encompass noble metal nanoparticles, metal oxides, metal-carbon nanotubes, carbon quantum dots, graphene-conjugated metallic nanostructures, as well as their nanocomposites, metal–organic frameworks (MOFs), and MXenes. In general, these materials provide enhanced sensitivity through signal amplification and transduction mechanisms. This review primarily focuses on the advancement of aptasensors based on metallic materials for the highly sensitive detection of protein targets, including enzymes and growth factors. Additionally, it sheds light on the challenges encountered in this field and outlines future prospects. We firmly believe that this review will offer a comprehensive overview and fresh insights into metallic nanomaterials-based aptasensors and their capabilities, paving the way for the development of innovative point-of-care (POC) diagnostic devices.
Article
Full-text available
Alzheimer’s disease (AD) is a complex multifactorial disorder that poses a substantial burden on patients, caregivers, and society. Considering the increased aging population and life expectancy, the incidence of AD will continue to rise in the following decades. However, the molecular pathogenesis of AD remains controversial, superior blood-based biomarker candidates for early diagnosis are still lacking, and effective therapeutics to halt or slow disease progression are urgently needed. As powerful genetic regulators, microRNAs (miRNAs) are receiving increasing attention due to their implications in the initiation, development, and theranostics of various diseases, including AD. In this review, we summarize miRNAs that directly target microtubule-associated protein tau (MAPT), amyloid precursor protein (APP), and β-site APP-cleaving enzyme 1 (BACE1) transcripts and regulate the alternative splicing of tau and APP. We also discuss related kinases, such as glycogen synthase kinase (GSK)-3β, cyclin-dependent kinase 5 (CDK5), and death-associated protein kinase 1 (DAPK1), as well as apolipoprotein E, that are directly targeted by miRNAs to control tau phosphorylation and amyloidogenic APP processing leading to Aβ pathologies. Moreover, there is evidence of miRNA-mediated modulation of inflammation. Furthermore, circulating miRNAs in the serum or plasma of AD patients as noninvasive biomarkers with diagnostic potential are reviewed. In addition, miRNA-based therapeutics optimized with nanocarriers or exosomes as potential options for AD treatment are discussed.
Article
Full-text available
Alzheimer's disease (AD) is a chronic progressive neurodegenerative disease that constitutes the third most common cause of death among elderly individuals. AD patients suffer from behavioral problems regularly, like memory loss, thinking deficits, and cognitive disorders. Currently, the incidence rate of AD has been on the rise worldwide. Therefore, it is imperative to develop effective strategies for the treatment and prevention of AD. The inherent characteristics of low‐dimensional nanomaterials, primarily their smaller size and surface structure, make them well‐suited for fulfilling the essential requirements of therapeutic drug design. Consequently, these nanomaterials offer a highly appealing and alternative approach to treating AD. Here, the low‐dimensional nanomaterials used in the treatment of AD to provide new ideas, methods, and comprehensive perspectives for the management of AD are reviewed. This review will advance the comprehensive understanding of the potential of low‐dimensional nanomaterials in the field of neuro medicine, which also will have a positive impact on future research.
Article
Full-text available
Alzheimer's disease (AD) is a neurodegenerative disease and the most common cause of dementia. Although it is discovered more than 100 years ago and is the subject of many scientific studies, much is yet to be discovered about many aspects of this disease, including the precise biological changes that cause it, the best approach to its early diagnosis, and effective therapeutic interventions to slow or stop the progression. This article briefly reviews the disease progression, risk factors, and pathogenesis of the disease. In addition, the current early diagnostic and therapeutical strategies for AD are presented.
Chapter
Last few decades unveiled the existence of diverse pool of non-coding RNAs, small RNAs, and repetitive elements which can play a pivotal role in maintaining the cellular homeostasis and dysfunction. Time- and tissue-dependent gene regulation by these small RNAs in-vivo made them a new class of pharmacological agents. The therapeutic effects of conventional treatment are fugitive as they are developed to target proteins rather than the actual underlying cause. On the other hand, nucleic acid-based therapeutics can procure long-term effect with much higher efficacy through gene silencing, editing. Thus, nucleic acid-based therapeutics are emerging as one of the most reliable on-target therapy for the treatment of genetically well-defined neurodegenerative diseases like Alzheimer’s disease, Huntington’s disease, Parkinson’s disease, amyotrophic lateral sclerosis, etc. Here, we will discuss the important nucleic acid-based therapies to treat neurodegenerative diseases along with the opportunities and challenges depending on the disease and the drug target. In addition, we will also accentuate the present-day application of these inhibitory small RNAs and antisense oligonucleotides in animal models and clinical trials.KeywordsAntisense oligonucleotidessiRNAAdeno-associated virus vectorsNeurodegenerative disease
Article
Full-text available
Nucleic acid aptamers are single-stranded DNA or RNA oligonucleotide sequences that bind to a specific target molecule with high affinity and specificity through their ability to adopt three-dimensional structure in solution. Aptamers have huge potential as targeted therapeutics, diagnostics, delivery agents and as biosensors. However, aptamers composed of natural nucleotide monomers are quickly degraded in vivo and show poor pharmacodynamics properties. To overcome this, chemically-modified nucleic acid aptamers are developed by incorporating modified nucleotides after or during the selection process by Systematic Evolution of Ligands by EXponential enrichment (SELEX). This review will discuss the development of chemically-modified aptamers and provide the pros and cons, and new insights on in vitro aptamer selection strategies by using chemically-modified nucleic acid libraries.
Article
Full-text available
Alzheimer's disease (AD) is the most common form of dementia, with a prevalence that increases with age. By 2050, the worldwide number of patients with AD is projected to reach more than 140 million. The prominent signs of AD are progressive memory loss, accompanied by a gradual decline in cognitive function and premature death. AD is the clinical manifestation of altered proteostasis. The initiating step of altered proteostasis in most AD patients is not known. The progression of AD is accelerated by several chronic disorders, among which the contribution of diabetes to AD is well understood at the cell biology level. The pathological mechanisms of AD and diabetes interact and tend to reinforce each other, thus accelerating cognitive impairment. At present, only symptomatic interventions are available for treating AD. To optimise symptomatic treatment, a personalised therapy approach has been suggested. Intranasal insulin administration seems to open the possibility for a safe, and at least in the short term, effective symptomatic intervention that delays loss of cognition in AD patients. This review summarizes the interactions of AD and diabetes from the cell biology to the patient level and the clinical results of intranasal insulin treatment of cognitive decline in AD.
Article
Full-text available
The oligonucleotide therapeutics field has seen remarkable progress over the last few years with the approval of the first antisense drug and with promising developments in late stage clinical trials using siRNA or splice switching oligonucleotides. However, effective delivery of oligonucleotides to their intracellular sites of action remains a major issue. This review will describe the biological basis of oligonucleotide delivery including the nature of various tissue barriers and the mechanisms of cellular uptake and intracellular trafficking of oligonucleotides. It will then examine a variety of current approaches for enhancing the delivery of oligonucleotides. This includes molecular scale targeted ligand-oligonucleotide conjugates, lipid- and polymer-based nanoparticles, antibody conjugates and small molecules that improve oligonucleotide delivery. The merits and liabilities of these approaches will be discussed in the context of the underlying basic biology.
Article
Full-text available
Memantine, a partial antagonist of N-methyl-D-aspartate receptor (NMDAR), approved for moderate to severe Alzheimer's disease (AD) treatment within the US and Europe under brand name Namenda (Forest), Axura and Akatinol (Merz), and Ebixa and Abixa (Lundbeck), may have potential in alleviating additional neurological conditions, such as vascular dementia (VD) and Parkinson's disease (PD). In various animal models, memantine has been reported to be a neuroprotective agent that positively impacts both neurodegenerative and vascular processes. While excessive levels of glutamate result in neurotoxicity, in part through the over-activation of NMDARs, memantine-as a partial NMDAR antagonist, blocks the NMDA glutamate receptors to normalize the glutamatergic system and ameliorate cognitive and memory deficits. The key to memantine's therapeutic action lies in its uncompetitive binding to the NMDAR through which low affinity and rapid off-rate kinetics of memantine at the level of the NMDAR-channel preserves the physiological function of the receptor, underpinning memantine's tolerability and low adverse event profile. As the biochemical pathways evoked by NMDAR antagonism also play a role in PD and since no other drug is sufficiently effective to substitute for the first-line treatment of L-dopa despite its side effects, memantine may be useful in PD treatment with possibly fewer side effects. In spite of the relative modest nature of its adverse effects, memantine has been shown to provide only a moderate decrease in clinical deterioration in AD and VD, and hence efforts are being undertaken in the design of new and more potent memantine-based drugs to hopefully provide greater efficacy.
Article
Full-text available
Apolipoprotein E receptor 2 (ApoER2) is an apolipoprotein E receptor involved in long-term potentiation, learning, and memory. Given its role in cognition and its association with the Alzheimer's disease (AD) risk gene, apoE, ApoER2 has been proposed to be involved in AD, though a role for the receptor in the disease is not clear. ApoER2 signaling requires amino acids encoded by alternatively spliced exon 19. Here, we report that the balance of ApoER2 exon 19 splicing is deregulated in postmortem brain tissue from AD patients and in a transgenic mouse model of AD. To test the role of deregulated ApoER2 splicing in AD, we designed an antisense oligonucleotide (ASO) that increases exon 19 splicing. Treatment of AD mice with a single dose of ASO corrected ApoER2 splicing for up to 6 months and improved synaptic function and learning and memory. These results reveal an association between ApoER2 isoform expression and AD, and provide preclinical evidence for the utility of ASOs as a therapeutic approach to mitigate Alzheimer's disease symptoms by improving ApoER2 exon 19 splicing.
Article
Full-text available
The earliest clinical manifestation of Alzheimer's disease (AD) is cognitive impairment caused by synaptic dysfunction. ApoE4, the primary risk factor for late-onset AD, disrupts synaptic homeostasis by impairing synaptic ApoE receptor trafficking. Alternative splicing of ApoE receptor-2 (Apoer2) maintains synaptic homeostasis. In this issue, Hinrich et al (2016) show that Apoer2 splicing is impaired in human AD brains and murine AD models and that restoring normal splicing in the mouse rescues amyloid-induced cognitive defects.
Article
The treatment of brain disorders is greatly hindered by the presence of the blood brain barrier which restricts the overwhelming majority of small molecules from entering the brain. A novel approach in which to overcome this barrier is to target receptor mediated transport mechanisms present on the endothelial cell membranes. Therefore, we fused an aptamer that binds to epithelial cell adhesion molecule-expressing cancer cells to an aptamer targeting the transferrin receptor. This generated a proof of concept bi-functional aptamer that can overcome the blood brain barrier and potentially specifically target brain disorders. The initial fusion of the two sequences enhanced the binding affinity of both aptamers while maintaining specificity. Additionally, mutations were introduced into both binding loops to determine their effect on aptamer specificity. The ability of the aptamer to transcytose the blood brain barrier was then confirmed in vivo following a 1 nanomole injection. This study has shown that through the fusion of two aptamer sequences, a bi-functional aptamer can be generated which has the potential to be developed for the specific treatment of brain disorders.
Article
β-Amyloid (Aβ) peptides are thought to be critically involved in the etiology of Alzheimer’s disease (AD). The aspartyl protease β-site amyloid precursor protein cleaving enzyme 1 (BACE1) is required for the production of Aβ, and BACE1 inhibition is thus an attractive target for the treatment of AD. We show that verubecestat (MK-8931) is a potent, selective, structurally unique BACE1 inhibitor that reduced plasma, cerebrospinal fluid (CSF), and brain concentrations of Aβ40, Aβ42, and sAPPβ (a direct product of BACE1 enzymatic activity) after acute and chronic administration to rats and monkeys. Chronic treatment of rats and monkeys with verubecestat achieved exposures >40-fold higher than those being tested in clinical trials in AD patients yet did not elicit many of the adverse effects previously attributed to BACE inhibition, such as reduced nerve myelination, neurodegeneration, altered glucose homeostasis, or hepatotoxicity. Fur hypopigmentation was observed in rabbits and mice but not in monkeys. Single and multiple doses were generally well tolerated and produced reductions in Aβ40, Aβ42, and sAPPβ in the CSF of both healthy human subjects and AD patients. The human data were fit to an amyloid pathway model that provided insight into the Aβ pools affected by BACE1 inhibition and guided the choice of doses for subsequent clinical trials.
Article
The ability to directly detect tau protein and other neurodegenerative biomarkers in human plasma at clinically relevant concentrations continues to be a significant hurdle for the establishment of diagnostic tests for Alzheimer's disease (AD). In this article, we introduce a new DNA aptamer/antibody sandwich assay pairing and apply it for the detection of human Tau 381 in undiluted plasma at concentrations as low as 10 fM. This was achieved on a multi-channel surface plasmon resonance (SPR) platform with the challenge of working in plasma overcome through the development of a tailored mixed monolayer surface chemistry. In addition, a robust methodology was developed involving various same chip control measurements on reference channels to which the detection signal was normalized. Comparative measurements in plasma between SPR and enzyme-linked immunosorbent assay (ELISA) measurements were also performed to highlight both the 1000-fold performance enhancement of SPR and the ability to measure both spiked and native concentrations that are not achievable with ELISA.
Article
Tau is a cytosolic protein that functions in the assembly and stabilization of axonal microtubule networks. Its oligomerization may be the rate-limiting step of insoluble aggregate formation, which is a neuropathological hallmark of Alzheimer's disease (AD) and a number of other tauopathies. Recent evidence indicates that soluble tau oligomers are the toxic species for tau-mediated pathology during AD progression. Herein, we describe novel RNA aptamers that target human tau and were identified through an in vitro selection process. These aptamers significantly inhibited the oligomerization propensity of tau both in vitro and in cultured cell models of tauopathy without affecting the half-life of tau. Tauopathy model cells treated with the aptamers were less sensitized to proteotoxic stress induced by tau overexpression. Moreover, the tau aptamers significantly alleviated synthetic tau oligomer-mediated neurotoxicity and dendritic spine loss in primary hippocampal neurons. Thus, our study demonstrates that delaying tau assembly with RNA aptamers is an effective strategy for protecting cells under various neurodegenerative stresses originating from pathogenic tau oligomerization.