ArticlePDF Available

Identification and Characterization of DNA Aptamers Specific for Phosphorylation Epitopes of Tau Protein

Authors:

Abstract and Figures

Tau proteins are proteins that stabilize microtubules, but their hyperphosphorylation can result in the formation of protein aggregates and, over time, neurodegeneration. This phenomenon, termed tauopathy, is pathologically involved in several neurodegenerative disorders. DNA aptamers are single-stranded oligonucleotides capable of specific binding to target molecules. Using tau epitopes predisposed for phosphorylation, we identified 6 distinct aptamers that bind to tau at two phosphorylatable epitopes (Thr-231 and Ser-202) and to full-length Tau441 proteins with nanomolar affinity. In addition, several of these aptamers also inhibit tau phosphorylation (IT4, IT5, IT6) and tau oligomerization (IT3, IT4, IT5, IT6). This is the first report to identify tau epitope-specific aptamers. Such tau aptamers can be used to detect tau in biofluids and uncover the mechanism of tauopathy. They can be further developed into novel therapeutic agents in mitigating tauopathy-associated neurodegenerative disorders.
Content may be subject to copyright.
Identification and Characterization of DNA Aptamers Specific for
Phosphorylation Epitopes of Tau Protein
I-Ting Teng,†, Xiaowei Li,†, Hamad Ahmad Yadikar§, Zhihui Yang§, Long Li, Yifan Lyu,
Xiaoshu Pan, Kevin K. Wang*,§,||, and Weihong Tan*,†,‡
Department of Chemistry, Department of Physiology and Functional Genomics, Center for
Research at Bio/Nano Interface, UF Health Cancer Center, UF Genetics Institute and McKnight
Brain Institute, University of Florida, Gainesville, Florida 32611-7200, United States
Molecular Science and Biomedicine Laboratory (MBL), State Key Laboratory for Chemo/Bio-
Sensing and Chemometrics, College of Chemistry and Chemical Engineering, College of Biology,
and Aptamer Engineering Center of Hunan Province, Hunan University, Changsha 410082, China
§Department of Emergency Medicine, Department of Chemistry, Department of Neuroscience,
and Department of Psychiatry, McKnight Brain Institute, University of Florida, Gainesville, Florida
32611, United States
||rain Rehabilitation Research Center (BRRC), Malcom Randall Veterans Affairs Medical Center,
1601 SW Archer Road, Gainesville Florida 32608, United States
Abstract
Tau proteins are proteins that stabilize micro-tubules, but their hyperphosphorylation can result in
the formation of protein aggregates and, over time, neuro-degeneration. This phenomenon, termed
tauopathy, is pathologically involved in several neurodegenerative disorders. DNA aptamers are
single-stranded oligonucleotides capable of specific binding to target molecules. Using tau
epitopes predisposed for phosphorylation, we identified six distinct aptamers that bind to tau at
two phosphorylatable epitopes (Thr-231 and Ser-202) and to full-length Tau441 proteins with
nanomolar affinity. In addition, several of these aptamers also inhibit tau phosphorylation (IT4,
IT5, IT6) and tau oligomerization (IT3, IT4, IT5, IT6). This is the first report to identify tau
epitope-specific aptamers. Such tau aptamers can be used to detect tau in biofluids and uncover the
mechanism of tauopathy. They can be further developed into novel therapeutic agents in mitigating
tauopathy-associated neurodegenerative disorders.
Graphical Abstract
*Corresponding Authors: tan@chem.ufl.edu, kwang@ufl.edu.
These authors contributed equally.
Supporting Information
The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/jacs.8b08645.
Additional experimental details, additional characterization figures and data, peptide sequences, and DNA sequences (PDF)
The authors declare no competing financial interest.
HHS Public Access
Author manuscript
J Am Chem Soc
. Author manuscript; available in PMC 2019 October 31.
Published in final edited form as:
J Am Chem Soc
. 2018 October 31; 140(43): 14314–14323. doi:10.1021/jacs.8b08645.
Author Manuscript Author Manuscript Author Manuscript Author Manuscript
INTRODUCTION
Tau proteins are microtubule-associated proteins known to promote the assembly of
microtubules and to maintain microtubule integrity, which is essential for axonal transport
and morphogenesis.1,2 Normal tau proteins bind with micro-tubules and prevent these track-
like structures from breaking apart, allowing nutrients and molecules to be transported along
the cells. However, tau is also found to be pathologically involved in several neurological
disorders, termed tauopathies, in which aggregations of tau are deposited in brain neurons.3
In the case of Alzheimer’s disease, pathological tau proteins self-assemble into paired
helical filaments, which later aggregate into insoluble neurofibrillary tangles.4 The transport
system for neurons is disrupted along the process, causing nutrients and other essential
supplies to cease moving along the cells. Neurons with tangles and nonfunctioning micro-
tubules consequently undergo apoptosis and eventually cell death. Such a phenomenon is
also observed in a range of other neurodegenerative diseases.5,6 Various forms of insoluble
abnormal tau aggregates are involved in tauopathies, but they share a common composition
of hyperphosphorylated tau.
Aptamers are nucleic acid probes capable of specific binding to defined targets.7 They are
selected through an amplification-evolution process termed systematic evolution of ligands
by exponential enrichment (SELEX).8 Aptamers have been selected against a variety of
targets, including metal ions,9 fluorescent dyes,10 amino acids,11 nucleotides,12 antibiotics,
13 metabolites,14 peptides,15 proteins,16 viruses,17 organelles,18 or even whole cells.19 As
such, aptamers have shown remarkable specificity in discriminating targets from their
analogue counterparts, such as differentiating among homologous proteins differing by only
a few amino acids20 or one single amino acid,21 or even between enantiomers.22 Because
molecular recognition is essential in many biological processes, aptamers can potentially
inhibit the functions of their targets. For instance, binding of thrombin with its aptamers has
been shown to undermine its activity and decrease the rate of blood clotting.23
The most fundamental difference between healthy tau and pathological tau can be ascribed
to the level of phosphor-ylation. Abnormally phosphorylated tau lesion causes death of
neuron cells, resulting in irreversible and progressive neuro-degeneration. Understanding the
origin and mechanism of tauopathy is a key step toward developing a means to delay or even
fight against it. Since pathological tau loses its affinity for microtubules due to an
abnormally high degree of phosphor-ylation, we hypothesized that aptamers binding to
Teng et al. Page 2
J Am Chem Soc
. Author manuscript; available in PMC 2019 October 31.
Author Manuscript Author Manuscript Author Manuscript Author Manuscript
phosphor-ylatable regions on tau protein might be useful in studying the molecular
mechanism(s) underlying tauopathy. It was also anticipated that aptamers binding to the
phosphorylated sites on tau could be exploited to investigate the formation of
hyperphosphorylated tau aggregates and detect the phosphor-ylation level. Therefore, we
herein describe a SELEX process using fragments of phosphorylatable regions from tau
protein and their corresponding phosphorylated forms as targets to search for site-specific
tau aptamers for potential further use as tauopathy-detecting agents and possible therapeutic
agents.
RESULTS AND DISCUSSION
Tau Epitope-Specific Aptamer Discovery and Selection.
The selection began with a library containing 20 nmol of primer-flanked, 66-nucleotide-
long, single-stranded DNA. Four phosphorylatable peptide epitopes from tau and their
corresponding phosphorylated peptides (Table S1) were used as putative targets (Figure 1).
The detailed screening strategy is described in Supporting Information.
The 10 most abundant sequences found in pool #17 and their population percentage in each
of the sequenced pools are listed in Table S2. They are denoted as IT1 through IT10 and
were chemically synthesized and labeled with FAM at the 5’-end, followed by HPLC
purification. Primary binding analysis was examined by flow cytometry. Seven out of the 10
candidates are confirmed aptamers (Figure 2F, K–P). Five of them (IT1, IT4, IT5, IT6, and
IT9) possess strong and specific binding to T231 peptide, while IT3 bypassed the
phosphorylated site and bound to both T231 and T231P peptides. Aptamer IT2, on the other
hand, recognized not only both T231 and T231P, but it also bound to S202. It is noteworthy
that IT8, IT9, and IT10 bore a high level of sequence homology to IT3, IT5, and IT2,
respectively. In fact, the discrepancy only occurs at the 19th nucleotide (Table S3). IT9 and
its predecessor IT5 displayed similar binding strength toward T231, suggesting that the
nucleotide at position 19 for these two similar sequences had no significant impact on their
binding abilities to T231. However, IT8 and IT10 both lost their binding abilities to the
prospective targets found with IT3 and IT2, proving that the binding strength of an aptamer
could be greatly compromised by merely altering one nucleotide within the crucial binding
region. Based on the fact that IT8 and IT10 manifested no binding preference to either of the
peptides, they are likely the byproducts that resulted from PCR amplification with edge
effect.
The initial binding tests were carried out at 4 °C to ensure having the optimal secondary
structures for aptamers. The binding abilities of the selected aptamers were then examined at
room temperature and at 37 °C. None of the aptamers lost binding ability at room
temperature or 37 °C (data not shown), suggesting that these aptamers are suitable for future
in vivo studies.
Sequence Truncation of the Selected Aptamers.
The aptamers identified are full-length sequences evolved from the initial library, which
contains the fixed primer binding regions on both ends to serve the PCR amplification
Teng et al. Page 3
J Am Chem Soc
. Author manuscript; available in PMC 2019 October 31.
Author Manuscript Author Manuscript Author Manuscript Author Manuscript
process. However, some full-length aptamers can be shortened into a minimally functional
sequence without compromising direct interaction with the target.24 Herein we synthesized
such truncated versions of aptamers IT1, IT2, IT3, IT4, IT5, and IT6 based on secondary
structures predicted by Integrated DNA Technologies OligoAnalyzer Tool (Figure 2A–J,
Table S4). Two to three T-bases were also added as a spacer next to fluorescein if the
truncated sequence ended with G-base or GC pair in the stem.
We successfully identified the binding motif from aptamer IT1 as sequence IT1c, which is
less than half the length of its original IT1 sequence. IT2 was at first truncated to IT2a based
on an evident stem–loop motif observed in the predicted secondary structure, and IT2a
demonstrated binding ability equal to IT2 for targets T231, T231P, and S202. Still, an even
more stringent truncation of IT2, candidate IT2b, failed to maintain the properties of the
original aptamer. A less exacting approach was then implemented to further shorten aptamer
IT2a. Sequence IT2c presented binding ability similar to IT2a, but a gradual loss of binding
ability was observed by further truncating the stem of IT2c into IT2d and IT2e, suggesting
the importance of a stable stem for binding of aptamer IT2c to its targets. A less restrictive
truncation of aptamer IT3 into sequence IT3c resulted in partial binding to T231, but it
caused a complete loss of binding to T231P. For IT4 and IT5, the complete sequences are
required for target recognition. Finally, a branched hairpin structure, IT6a, retained selective
binding ability to T231 as aptamer IT6.
Determination of Binding Kinetics/Affinities of the Selected Aptamers.
In this kinetics study, his-tag peptides, as well as his-tag Tau441 protein, were immobilized
on the surface of the biosensor, while aptamer candidates were the analytes in the solution
phase. The binding interaction of analyte to immobilized ligand was measured in real time
as the change in the number of molecules bound to the biosensor that caused a shift in the
interference pattern reflected from sensor surfaces. The kinetic on-rates (
k
on), off-rates
(
k
off), and equilibrium dissociation constants (Kd) measured for all aptamer-target pairs are
summarized in Table S5.
All aptamers reported here displayed high binding affinities toward Tau441 with
K
d values
ranging from 5.5 nM to 68 nM. The observed on-rates (
k
on) ranged between 104 M−1 s−1
and ~106 M−1 s−1. Aptamer IT1 presented the fastest on-rate ((9.3 ± 1.9) × 105 M−1 s−1)
toward Tau441 protein, while the slowest on-rate ((1.067 ± 0.018) × 104 M−1 s−1) was
detected for IT2 binding to Tau441. However, IT2 also demonstrated an extremely slow off-
rate (
k
off) ((5.9 ± 1.2) × 10−5 s−1) for Tau441, exhibiting the lowest
K
d (5.5 ± 1.1 nM) for
Tau441 protein among all aptamers. The lowest
K
d toward peptides, on the other hand, was
observed between IT5 and T231 (5.0 ± 0.3 nM). Meanwhile, IT5 also exhibited the second
lowest
K
d (7.6 ± 0.6 nM) for Tau441 protein among all aptamers, indicating the high binding
affinities of IT5 toward both T231 peptide and Tau441 protein. Although aptamer IT9 differs
by only one base from IT5, as shown in Table S3, it showed appreciably slower
k
on and,
therefore, higher equilibrium dissociation constants compared to IT5.
While IT1 showed similar binding kinetics toward T231 and Tau441, the truncated aptamer
IT1c exhibited a much slower off-rate for Tau441 protein than for T231 peptide. On the
other hand, truncation of IT2 did not affect the association of the aptamers to T231 peptide,
Teng et al. Page 4
J Am Chem Soc
. Author manuscript; available in PMC 2019 October 31.
Author Manuscript Author Manuscript Author Manuscript Author Manuscript
but the shorter aptamer IT2a did display a faster dissociation rate for T231 compared to IT2.
In fact, the dissociation rates with any of the targets were increased with IT2a compared to
the off-rates with IT2. In addition, a faster on-rate for T231P peptide was found with either
IT2 or IT2a in comparison to that of T231. With IT2a, especially, the on-rate toward T231P
was almost 5 times faster than its on-rate toward T231, and this on-rate was also 2.5-fold
faster than the association of IT2 to T231P. However, the association between IT2a and S202
peptide showed an opposite tendency. The on-rate of IT2a for S202 was found to be more
than 6 times slower than its IT2 counterpart. Therefore, the truncation of IT2 appears to
benefit its recognition of T231P, while, at the same time, losing its binding affinity toward
the S202 site. Moreover, while the association of IT2a to Tau441 is almost 50 times faster
than IT2, its off-rate is also 60 times faster than that of IT2. Overall, the binding affinity of
IT2a was weakened by the truncation. Unlike IT1c and IT2a, the shorter version of IT6 did
not affect the binding kinetics and affinity. IT6a still behaved much like IT6. Finally, the on-
rates of IT3 to both T231 and T231P peptides were found to be quite similar, but the off-rate
with T231P was almost 2-fold faster than that with T231.
Specificity of the Selected Tau Aptamers against Tau441 Protein.
The binding specificity of the selected tau aptamers to full-length Tau441 protein was
confirmed by nondenaturing gel electrophoresis after incubating each of the FITC-labeled
aptamers with either target Tau441 protein or the individual nontarget proteins separately.
Each of the tau aptamers (IT1-IT6 and IT9) alone displayed one main band at the lower part
of the gel (Figure 3, lane 1). Some minor upper bands observed could have resulted from
dimeric or multimeric forms of the oligonucleotides. The monomeric aptamers disappeared
in the presence of the target Tau441 protein due to the formation of aptamer-Tau441
complexes (Figure 3, lane 2). No cross-reactivity was observed between the aptamers and
the nontarget proteins, including S100B (S100 calcium-binding protein B) (11 kDa), UCH-
L1 (ubiquitin carboxyl-terminal hydrolase L1) (25 kDa),
α
casein (23 kDa),
β
casein (24
kDa), BSA (bovine serum albumin) (66 kDa), and IgG (immunoglobin G) (150 kDa) (Figure
3, lane 3–8). In particular, S100B25 and UCH-L126 are important references because they
are also brain-associated protein and traumatic brain injury (TBI) biofluid biomarker
proteins.27
Aptamer-Based Sandwich ELISA for Detection of Tau.
To further study the feasibility of using tau aptamers for tau protein detection, a sandwich
enzyme-linked immunosorbent-assay (ELISA) was performed in triplicate. Tau protein was
first captured by DAKO antibody. The biotin-labeled IT4 aptamer was used as the detection
probe to specifically recognize tau protein. The 620 nm absorbance was recorded after
incubating samples with streptavidin-labeled polyHRP enzyme and TMB substrate. As
shown in Figure 4, IT4 aptamer strongly binds to tau protein with a dose-dependent increase
in signal, as opposed to nontarget phosphorylated tau, indicating the possibility of utilizing
IT4 aptamer to specifically quantify the tau level in biological samples.
Inhibitory Effects of Tau Aptamers on Tau Phosphor-ylation.
To analyze the ability of tau aptamers to inhibit tau phosphorylation, tau was phosphorylated
by kinase glycogen synthase kinase-3
β
(GSK3
β
) for 24 h at 37 °C in the presence or
Teng et al. Page 5
J Am Chem Soc
. Author manuscript; available in PMC 2019 October 31.
Author Manuscript Author Manuscript Author Manuscript Author Manuscript
absence of tau aptamers. We used phospho-T231 tau monoclonal antibody (RZ3) to probe
the phosphorylated tau. A faint 68K band was observed with control Tau441 protein (Figure
5A, lane 1), while an intense band showing a molecular shift from 68K to 70K was detected
with the commercially available positive control GSK3
β
prephosphorylated Tau441 (Figure
5A, lane 2). Lane 3 shows the experimentally GSK3
β
phosphorylated Tau441 at both the
monomeric 70K band and oligomeric 150 K p-tau band. The presence of random
oligonucleotide had no effect on the levels of phosphorylated monomeric tau or oligomeric
p-tau (Figure 5, lane 4). IT1 and IT1c showed minor reduction of monomeric p-tau, but they
had no effect on oligomeric p-tau (Figure 5, lanes 5 and 6). The presence of IT2, IT2a and
IT3 somehow promoted phosphorylated tau toward the oligomeric form (Figure 5, lanes 7–
9). The reason behind this result is not fully understood yet, but we suspect this is due to the
fact that these aptamers bind to more than one epitope (Figure 2K,L). Among all aptamers,
IT4 eliminated both monomeric p-tau and oligomeric p-tau most dramatically (Figure 5, lane
10). IT5 also showed some inhibition on monomeric p-tau, but IT6 and IT6a had no
significant effects (Figure 5, lanes 11–13).
Heparin-Induced Aggregation of Tau441 and the Inhibitory Effects of Tau Aptamers.
To examine the inhibitory effects of tau aptamer on the formation of tau oligomers in vitro,
heparin was used as an oligomerization inducer.28 In brief, tau protein was incubated with
either the aptamer or a random sequence for 1 h followed by treatment with heparin for 16 h
to form tau oligomers. The reaction products were then analyzed by SDS-PAGE and
immune-blotted with polyclonal total tau antibody. Tau441 was detected at 68K in
monomeric form (Figure 6A, lane 1). Heparin treatment caused the formation of high
molecular weight (~150 K oligomeric tau) (Figure 6A, lane 2). When random sequence
oligonucleotide was preincubated with tau, the oligomeric assembly of tau induced was
virtually unaffected. In contrast, IT4, IT5, IT6, and the truncated IT6a showed the most
robust inhibition of oligomeric tau formation among the tested aptamers. Truncated IT2a and
IT3 also showed partial inhibition of tau oligomerization. However, the remaining aptamers
did not show any statistically significant effects on tau oligomerization (Figure 6A,B).
In a separate experiment, an oligomerization study was performed with 1, 10, 25, and 50
μ
M
of IT3, followed by probing with a monoclonal tau antibody (DA9). A dose-dependent effect
to reduce the formation of tau oligomers was confirmed (≥90% inhibition of tau oligomers
was observed at 50
μ
M of IT3) (Figure 6C).
Investigation of Tau Binding to Microtubules in the Presence of Aptamers.
A microtubule binding protein spin-down assay was used to investigate the binding of tau to
microtubules in the presence of aptamers. The assay was first validated according to the
manufacturer’s instructions (Figure S7). Briefly, microtubules were collected in the pellet
fraction along with any proteins bound on MTs. Proteins that have no affinity to MTs would
remain in the supernatant. The existence of each protein in the supernatant and pellet
fractions was analyzed through gel electrophoresis and protein staining. We then confirmed
that Tau441 was only found in the supernatant fraction when there were no MTs.
Conversely, in the presence of MTs, Tau441 predominantly bound to MTs and was pulled
down together with MTs during centrifugation (Figure S8A). A much reduced amount of tau
Teng et al. Page 6
J Am Chem Soc
. Author manuscript; available in PMC 2019 October 31.
Author Manuscript Author Manuscript Author Manuscript Author Manuscript
stayed in the supernatant. To study the effects of aptamers to the tau-bound and preformed
MT, the two most promising aptamers from the inhibition study (IT4 and IT5) were added
separately into the mixture of Tau441/MTs for 30 min before centrifugation. The results
indicated that IT4 and IT5 could detach partially of the bound Tau441 from MTs. This
outcome was not unanticipated though, as it has been reported that the T231 site on Tau441,
the target epitope of IT4 and IT5, has a pronounced influence on the binding affinity of tau
to microtubules, although it is distinct from the well-recognized microtubule-binding
domain.29
CONCLUSIONS
Although the exact causes of tauopathy and the related neurodegeneration remain unclear,
evidence has shown that pathological tau always appears in the hyperphosphorylated and
oligomeric form. To provide a specific tool for further investigating the relationship between
tauopathy and tau phosphorylation/oligomerization, we herein propose a peptide-based
selection process to identify site-specific tau-binding aptamers. Even though some previous
studies have found tau-binding DNA sequences30 and a tau-binding RNA aptamer,31 none of
these probes has a confirmed binding region. One essential advantage of using peptides
instead of the whole protein to carry out the aptamer selection for tau is that the binding sites
have been predetermined and are confined to the regions that are pathologically relevant.
Four phosphorylatable regions (T181, S202, T231, S396/S404) from tau protein and the
corresponding phosphorylated epitopes from pathological tau are used here as putative
targets to demonstrate the peptide-based tau aptamer selection method. Aptamer IT3
bypasses phosphorylation on T231 and recognizes both T231 and T231P. Aptamer IT2 could
bind to T231, T231P, and S202, but not S202P. The remaining selected aptamers are highly
specific to the T231 site. Additionally, aptamers IT1, IT2, and IT6 were structurally
minimized into shorter versions (IT1c, IT2c, IT6a) without appreciably compromising their
binding abilities. However, aptamers IT3, IT4, and IT5 were extremely susceptible to any
attempts at sequence truncation. All tau aptamers reported show high affinity and specificity
to tau protein. The dissociation constants of these aptamers against Tau441 protein range
from 5.5 nM to 68 nM.
We then successfully demonstrated the feasibility of using tau aptamers as a detection tool
by performing an aptamer–antibody sandwich ELISA using total tau antibody and IT4
aptamer. Phosphorylated tau was used as a negative control here, and our results showed that
this assay is highly specific to the desired target.
Next, since our tau aptamers preferentially bind to nonphosphorylated tau, we reasoned that
they might prevent tau from being phosphorylated at the T231 site by blocking substrate site
access to protein kinases. Our results did, indeed, confirm this by showing that aptamers IT4
and IT5 strongly inhibited tau phosphorylation by GSK3
β
at T231. In our results, the
GSK3
β
-phosphorylated Tau441 showed a high molecular weight form detected at 150 K by
RZ3 antibody, while this effect was not seen with the purified prephosphorylated Tau441. It
has been reported previously that high concentration of tau in solution causes it to aggregate
when incubated at 37 °C.28 Thus, the likely explanation is that the hyperphosphorylated
Teng et al. Page 7
J Am Chem Soc
. Author manuscript; available in PMC 2019 October 31.
Author Manuscript Author Manuscript Author Manuscript Author Manuscript
Tau441 by GSK3
β
had a higher tendency to oligomerize in solution. Indeed, we confirmed
that IT4, IT5, and IT6 also reduced the levels of oligomeric tau. In parallel, we also
hypothesized that tau aptamers might interfere with tau oligomerization, and such
hypothesis was tested with an oligomerization study. It was observed that aptamers IT3, IT4,
IT5, IT6, and IT6a could partially, or completely, inhibit tau oligomerization in vitro. Taken
together, IT4 and IT5 showed the most robust effects on the reduction of tau phosphorylation
and oligomerization, suggesting that they are therapeutic candidates for future animal-based
models and possible clinical studies. Though we later found that IT4 and IT5 have moderate
influence on the binding affinity of tau to microtubules, this result does not prevent us from
utilizing these aptamers for detection in biofluids. Besides, this result can be attributed to the
fact that the T231 epitope has been identified as a critical site in regulating tau’s ability to
bind microtubules.29 This does not eliminate the possibility of using our platform to find
aptamers that target other epitopes and have no impacts on tau’s binding ability to
microtubules. A separate study reporting an RNA aptamer targeting tau also showed that the
RNA-based tau aptamer could alleviate synthetic tau oligomer-mediated neurotoxicity in
primary rat hippocampal neurons in culture.31
However, it is worth noting that RNA aptamers are not as stable as DNA-based aptamers in
biological conditions. Furthermore, unlike our study, the reported RNA aptamer has
undefined tau binding sites, and its specificity, affinity, and on- and off-rates for tau protein
were not defined.
In summary, our novel findings demonstrate, for the first time, the feasibility of identifying
tau epitope-specific aptamers. When tau DNA aptamers are further optimized, we envision
that they might have multiple utility in terms of uncovering the mechanism of tauopathy or
in diagnosing and treating tauopathy-associated disorders. In terms of diagnosis of
tauopathies, tau and p-tau antibodies exist, but the advantages of using tau aptamers include
their stability, ease of storage, and thus longer shelf life when being used in a point-of-care
device. Aside from tau phosphorylation, tau acetylation has also been identified as a
dominant post-translational modification in tauopathies. A recent study revealed that tau
acetylation at residues K280/K281 impairs tau-mediated microtubule stabilization and
enhances the formation of tau aggregates.32 Hence, our peptide-based aptamer selection
process could also be applied to identify probes that recognize specific tau acetylation sites,
thus providing new tools to investigate the role of tau acetylation in tau pathology, or even
new antagonists to convey therapeutic approaches by binding to critical acetylation sites. In
terms of therapeutic treatments for tauopathy, immunotherapy with tau/p-tau antibodies for
tauopathy is a current research area.33 However, significant shortcomings have been
assigned to immunotherapy. In particular, antibodies are biological drugs, which are
expensive to produce because biological purity, activity qualification, batch-to-batch
consistency, removal of toxic contaminants, and antibody humanization must be achieved.
However, on the basis of the present study, we conclude that tau aptamers can be a practical
alternative to antibodies, as tau aptamer synthesis can be readily scaled up for human trials.
Importantly, nonphosphorylated tau was identified to be a major component of tau aggregate
in mouse and human tauopathy brain.34 Thus, our nonphosphorylated tau-preferring
aptamers are ideally positioned as a therapeutic for tauopathy diseases. In addition, aptamers
have already been used in human as FDA-approved therapeutics.35–37 To realize the
Teng et al. Page 8
J Am Chem Soc
. Author manuscript; available in PMC 2019 October 31.
Author Manuscript Author Manuscript Author Manuscript Author Manuscript
potential of tau aptamers as therapeutics, future studies should focus on further optimizing
aptamer properties (such as reducing their interference of tau–MT interaction), optimizing
their plasma half-life, improving blood–brain barrier penetration, and conducting proof-of
principle studies in animal models of tauopathies.
EXPERIMENTAL SECTION
Buffers and Peptides.
The buffer used to disperse peptides and peptide-beads was Dulbecco’s phosphate-buffered
saline (PBS) with calcium chloride and magnesium chloride (Sigma-Aldrich). The buffer
used to prepare DNA solutions, as well as the buffer for all binding reactions, was PBS
supplemented with additional 5 mM MgCl2. The sequences of the peptides used in this
study are listed in Table S1. All peptides were purchased from GenScript; they all came with
N-terminal acetylation and 2 units of 6-aminocaproic acid (Ahx) linker followed by six
histidine residues (his-tag) at their C-terminus.
Immobilization of Tau/p-Tau Peptides.
Twenty microliters of stock Ni Sepharose high performance beads (~7 × 105 Ni-NTA beads,
GE Healthcare Life Sciences) was equilibrated with 500
μ
L of PBS three times. After quick
spin-down with a minicentrifuge (Fisher Scientific), the supernatant was discarded. The
remaining pellet was suspended in 100
μ
L of PBS containing either 400
μ
g/mL or 800
μ
g/mL of the designated peptide and left overnight at 4 °C on a shaker. The prepared peptide
beads were kept in the peptide solution and left unwashed at 4 °C until use. Beads used in
counter selections or negative controls in binding tests were prepared as mentioned above
but suspended in PBS as opposed to peptide solution.
DNA Library and Primers.
The forward primer and the reverse primer were labeled with FAM and biotin, respectively,
at their 5’-ends. The sequence of the forward primer was 5’-FAM-CAG CAC CGT CAA
CTG AAT-3’; the sequence of the reverse primer was 5’-biotin-ACA TCT CCA TCG CAT
CAC-3’. The DNA library consisted of a randomized 30-nt region flanked by primer binding
sites: 5’-CAG CAC CGT CAA CTG AAT-(N30)-GTG ATG CGA TGG AGA TGT-3’. All
library and primer sequences were purchased from Integrated DNA Technologies and
purified by reverse phase HPLC.
Polymerase Chain Reaction (PCR).
PCR parameters were optimized before the selection process. All PCR mixtures contained
50 mM KCl, 10 mM Tris-HCl (pH 8.3), 1.5 mM MgCl2, 0.2 mM each dNTP, 0.5
μ
M each
primer, and Hot Start Taq DNA polymerase (0.015 units/
μ
L). PCR was performed on a
BioRad C1000 Thermo Cycler, and all PCR reagents were purchased from Takara. The
amplification began with a hot start at 95 °C for 90 s to activate Taq DNA polymerase. Then
each of the repeated amplification cycles was performed at 95 °C for 30 s, 57 °C for 30 s,
and 72 °C for 60 s, followed by a one-time final extension at 72 °C for 3 min.
Teng et al. Page 9
J Am Chem Soc
. Author manuscript; available in PMC 2019 October 31.
Author Manuscript Author Manuscript Author Manuscript Author Manuscript
SELEX Procedures.
The SELEX was performed by binding the DNA library to the peptide-immobilized beads
and eluting DNA survivors from the washed peptide beads, followed by PCR amplification
in a reciprocal manner. In the first round, an initial DNA library consisting of 20 nmol of the
randomized oligonucleo-tides was used as a starting pool. For later rounds, 150 pmol to 20
pmol of samples amplified from previously recovered survivors were used. In each round,
the DNA pool was heated at 95 °C for 3 min, followed by rapid cooling on ice for 5 min
before starting incubation, allowing the DNA sequences to form the most favorable
secondary structures. Beads were washed with 1 mL of PBS three times before use. Counter
selection with bare Ni-NTA beads was introduced in round #6 and was included in all
subsequent rounds. Sandwiched counter selections were used in round #8 and all later
rounds. All incubations were performed at 4 °C. After each cycle of negative incubation,
supernatant was collected, while the undesired sequences bound on Ni-NTA beads were
discarded. For positive incubations, peptide beads were washed with buffer to remove
unbound sequences. The washing stringency was enhanced as the selection progressed to
remove weaker candidates. The washed peptide beads were heated at 95 °C for 7 min,
followed by quick spin-down with a minicentrifuge. The surviving candidates in the
supernatant were collected and were ready for PCR amplification during the first five rounds
when negative selections were not included. For later rounds, the sequences recovered from
peptide beads were treated with counter selection before PCR amplification. The number of
optimized PCR amplification cycles for each round was confirmed with agarose gel
electrophoresis. Streptavidin Sepharose high performance beads (GE Healthcare Life
Sciences) were used to isolate the PCR products from the reaction mixture. The fluorophore-
labeled amplicons were then separated from the biotinylated antisense DNA by eluting with
20 mM NaOH. Finally, the ssDNA was desalted with a NAP-5 column (GE Healthcare Life
Sciences). The entire process is summarized in Table S6.
Flow Cytometric Analysis of Enriched Pools or Aptamer Candidates.
BD Accuri C6 flow cytometry (BD Immunocytometry Systems) was used to monitor the
enrichment of candidates during the selection process and evaluate the binding specificity of
the selected candidates. For each sample, 20
μ
L of peptide beads or bare beads (~1.4 × 105
beads) was washed with 500
μ
L of PBS three times. After removal of the supernatant, beads
were suspended in 80
μ
L of buffer containing either a DNA pool or an aptamer candidate to
be tested at 250 nM for 30 min. When examining the saturation of fluorescent signal with
the selection pools, the DNA concentration used was decreased to 100 nM instead.
Incubations were carried out at 4 °C unless stated otherwise. Afterward, beads were washed
against 500
μ
L of buffer twice and then suspended in 80
μ
L of buffer for flow cytometric
analysis.
Next-Generation Deep Sequencing of DNA Survivors.
Enriched DNA pools from rounds #15–17 were submitted for sequencing. Primers without
FAM or biotin modification were used to amplify the DNA pools to be sequenced. The
amplicons from each pool were then barcoded separately using the TruSeq DNA library
preparation kit (Illumina). The samples were submitted to Illumina next-generation DNA
Teng et al. Page 10
J Am Chem Soc
. Author manuscript; available in PMC 2019 October 31.
Author Manuscript Author Manuscript Author Manuscript Author Manuscript
sequencing at the University of Florida, ICBR Sequencing Core Facility. The sequencing
results were analyzed using in-house software.
Chemical Synthesis and Purification of Aptamer Candidates.
The 10 most abundant sequences from round #17 were synthesized by the standard
phosphoramidite method using a 3400 DNA synthesizer (Applied Biosystems). Reagents for
synthesis were purchased from Glen Research. The synthesized candidates were then
purified by reversed phase HPLC (Varian Prostar), using a C18 column and
triethylammonium acetate (TEAA, 0.1M)/acetonitrile (Fisher Scientific) as the mobile
phase.
Sequence Truncation of the Selected Aptamers.
The selected aptamers were truncated based on their secondary structures, as predicted by
Integrated DNA Technologies OligoAnalyzer Tool. Shorter versions of each of the selected
aptamers were synthesized and purified as described above. Their binding abilities were
evaluated by flow cytometry as described previously for full-length candidates.
Measurement of Binding Kinetics/Affinities of the Selected Aptamers.
The binding affinities and binding kinetics of the selected aptamers were determined on an
Octet RED384 instrument (Pall Fortebio). All measurements were performed on 384-well
plates that were agitated at 1000 rpm at 30 °C. The his-tag peptides were immobilized at a
concentration of 5
μ
g/mL on Ni-NTA sensors for 10 min, followed by rinsing the sensors in
PBS for 10 min. Then the sensors were brought into fresh association buffer (PBS with 5
mM Mg2+) to establish a baseline for another 10 min. For kinetics analysis, the peptide-
immobilized sensors were transferred to the wells containing aptamer dilutions (1
μ
M, 1.5
μ
M, 2
μ
M, 3
μ
M, 4
μ
M) for the association step (10 min) and then moved to the wells with
fresh buffer for the dissociation step (10 min). A reference sensor was used without the
treatment with aptamer solution, but other steps remained the same. When measuring the
binding of aptamers with his-tag Tau441 protein (SignalChem), the experimental layout was
the same as that described above, except that the loading concentration of his-tag Tau441
protein was 1
μ
g/mL supplemented with 0.75 mM imidazole, and the concentrations for
aptamer dilutions were 0.75
μ
M, 1.5
μ
M, 3
μ
M, 6
μ
M, and 12
μ
M. Association rate
constants (
k
on), dissociation rate constants (
k
off), and equilibrium dissociation constants (
K
d)
for each aptamer binding to its target peptide were calculated using the ForteBio data
analysis software 10.0.
Specificity of the Selected Tau Aptamers against Tau441 Protein.
Gel electrophoresis was carried out to confirm the binding specificity between the aptamers
against Tau441 protein. S100B, UCH-L1,
α
casein,
β
casein, BSA, and IgG were used as
nontarget proteins for reference. A 10-
μ
L mixture containing FITC-labeled aptamer at 200
nM and designated protein at 0.05 mg/mL were incubated at 4 °C for 30 min before loading
onto an 8% nondenaturing polyacrylamide gel. Electrophoresis was initially carried out at 70
V for 10 min after introduction of the samples, and then the voltage was increased to 150 V
Teng et al. Page 11
J Am Chem Soc
. Author manuscript; available in PMC 2019 October 31.
Author Manuscript Author Manuscript Author Manuscript Author Manuscript
for 45 min in 1x TBE buffer. Finally, gels were scanned using a Typhoon Imaging System
(Amersham Biosciences).
Aptamer–Antibody Sandwich ELISA.
The coating buffer containing 1
μ
g/mL of total tau antibody (DAKO) in carbonate-
bicarbonate buffer (pH 9.6) was pipetted into a 96-well plate (100
μ
L per well) and left
overnight at 4 °C. The wells were washed three times with TBST (Tris-buffered saline, with
Tween 20, pH 8.0, Sigma) and incubated with StartingBlock Blocking Buffers (Thermo
Scientific) for 1 h at room temperature. After removal of the blocking buffer and washing
three times with TBST, different concentrations of Tau441 protein were added into each well
(100
μ
L). The plate was stored at 4 °C overnight.
On the next day, the plate was washed three times with TBST and then treated with 500 nM
detecting aptamer or randomly scrambled aptamer sequence (RS) in 100
μ
L of PVP buffer
(20 mg/mL polyvinylpyrrolidone and 100
μ
g/mL salmon sperm in TBST). The reaction was
left on a shaker for 1 h. Afterward, the wells were washed with TBST three times and 100
μ
L of 1:10K diluted Pierce Streptavidin Poly-HRP (Thermo Scientific) solution was added
to each well and incubated for 1 h at room temperature. Finally, 100
μ
L of TMB solution
was added to each well, and the 620 nm absorbance was recorded by a plate reader.
Inhibition of Tau441 Phosphorylation by Aptamers.
Purified Tau441 (R-peptide) was used to perform phosphorylation by GSK3
β
Tau441 (1
μ
g)
was first preincubated with the aptamers (50
μ
M) for 1 h before adding the enzyme GSK3
β
(200 ng) to the mixture of kinase assay buffer. The reaction buffer contained 20 mM Tris-
HCl, 50 mM KCl, 10 mM MgCl2, pH 7.5. The assay was initiated by the addition of6 mM
ATP. The reaction was carried out at 37 °C for 16 h on a shaker. Prephosphorylated GSK3
β
Tau441 (400 ng) was used as a positive control. Total reaction volume was 25
μ
L. Five
microliters of the reaction sample was diluted with 15
μ
L of 4X Laemmli sample buffer,
followed by SDS-PAGE and Western blotting with RZ3 antibody.
Assay of in Vitro Oligomerization of Tau441 Induced by Heparin.
To examine the inhibitory effect of tau aptamers on heparin-induced tau oligomerization, we
preincubated different aptamers (50
μ
M) with Tau441 (10
μ
M) for 1 h at room temperature
in the oligomerization buffer (20 mM Tris HCL, pH 7.4, 100 mM NaCl, 1 mM DTT, and 1
mM EDTA), followed by the addition of heparin (5 mg/mL) for 16 h incubation at 37 °C.
The aptamers were prepared in 1X PBS with 5 mM MgCl2. The reaction was stopped by
addition of Laemmli sample buffer. The reaction products were analyzed by SDS-PAGE,
followed by immunoblotting with antitau antibody (DAKO).
SDS–PAGE Electrotransfer and Immunoblot Analysis.
Equal protein amount was prepared for sodium dodecyl sulfate–polyacrylamide gel
electrophoresis (SDS–PAGE) in 8X loading buffer containing 0.25 M Tris (pH 6.8), 2 mM
DTT, 8% SDS, and 0.02% bromophenol blue. Each sample was subjected to SDS-PAGE on
4–20% precast gels and then transferred onto PVDF membranes. The membranes were
blocked in 5% milk for 1 h and then incubated with primary antibodies RZ3 (pT231;
Teng et al. Page 12
J Am Chem Soc
. Author manuscript; available in PMC 2019 October 31.
Author Manuscript Author Manuscript Author Manuscript Author Manuscript
1:1000) or DAKO (total tau; 1:3000) that were detected using goat antimouse or goat
antirabbit IgG conjugated to alkaline phosphatase (Amersham, Piscataway, NJ) for 1 h at
room temperature. Immunoreactive bands were detected by developing with nitro blue
tetrazolium and 5-bromo-4-chloro-3’-indolylphosphate (NBT/BCIP) (KPL). Quantitative
evaluation of protein levels was performed via computer-assisted densitometric scanning
(NIH ImageJ, version 1.6).
Statistical Analysis.
Monomeric p-tau (70K) and oligomeric tau (150 K) were normalized by the levels of
GSK3
β
phosphorylated Tau441 or heparin-induced tau oligomers and shown as percentage.
Data were presented as mean ± SEM for
n
= 3. Statistical analysis was performed with two-
way ANOVA Tukey’s Test. For multiple comparisons, one-way ANOVA, followed by the
Bonferroni’s post hoc test, was performed. *
p
< 0.05, **
p
< 0.01, ***
p
< 0.001 and ****
p
<
0.0001, ns: nonsignificant. GraphPad Prism 7.0 (GraphPad).
Investigation of Tau Binding to Microtubules in the Presence of Aptamers.
The Microtubule Binding Protein Spin-Down Assay Kit (Cytoskeleton Inc., Denver, CO)
was used to investigate the binding of tau to microtubules in the presence of aptamers. To
validate the assay, microtubules (MTs) were first polymerized in the presence of GTP from
tubulin proteins (mixture of
α
- and
β
subunits), and then microtubule-associated protein
fraction (MAP2, 16
μ
g) and BSA protein (7.5
μ
g) were incubated with MTs (8 nM) as
positive and negative protein controls, respectively. Mixtures were centrifuged at 100 000g
at 4 °C for 40 min. The collected supernatant and pellet fractions were then mixed with
Novex 4X LDS sample buffer (Thermo Scientific) and analyzed via a 4–20% SDS-PAGE
gel (Thermo Scientific). The gel was stained by GelCode Blue Safe Protein Stain
(Coomassie blue) (Thermo Scientific) and scanned by Typhoon using Cy5 channel. To study
the interference of tau aptamers in the binding between Tau441 and MTs, Tau411 (5
μ
g) was
incubated with tubulin proteins, which later polymerized into MTs accordingly to the
manufacturer’s instructions as above. In a separate experiment, 20
μ
M FITC-labeled IT4 or
IT5 was added to the mixture of preformed MT in the presence of tau for 30 min of
incubation. The samples were centrifuged and analyzed as described above. The gel was
scanned using both Cy5 and Cy2 channel for protein and aptamer imaging, respectively.
Supplementary Material
Refer to Web version on PubMed Central for supplementary material.
ACKNOWLEDGMENTS
This work was supported in part by National Institute of Health (GM R35 127130) and NSF 1645215 as well as
NSFC grants (NSFC 21521063) (W.H.T.), and by University of Florida College of Medicine funding, Department
of Veterans Affairs (Merit Award Veterans Affairs, (I01 RX001859 A02), Rehabilitation Research and Development
Office, BRRC award no. B9256-C and BRRC Innovation Award no. 0118BRRC-01 (K.K.W).
Teng et al. Page 13
J Am Chem Soc
. Author manuscript; available in PMC 2019 October 31.
Author Manuscript Author Manuscript Author Manuscript Author Manuscript
REFERENCES
(1). Weingarten MD; Lockwood AH; Hwo SY; Kirschner MW A protein factor essential for
microtubule assembly. Proc. Natl. Acad. Sci. U. S. A 1975, 72 (5), 1858–1862. [PubMed:
1057175]
(2). Mandelkow EM; Biernat J; Drewes G; Gustke N; Trinczek B; Mandelkow E Tau domains,
phosphorylation, and interactions with microtubules. Neurobiol. Aging 1995, 16 (3), 355–362.
[PubMed: 7566345]
(3). Sergeant N; Delacourte A; Buee L Tau protein as a differential biomarker of tauopathies. Biochim.
Biophys. Acta, Mol. Basis Dis 2005, 1739 (2–3), 179–197.
(4). Brunden KR; Trojanowski JQ; Lee VMY Advances in tau-focused drug discovery for Alzheimer’s
disease and related tauopathies. Nat. Rev. Drug Discovery 2009, 8 (10), 783–793. [PubMed:
19794442]
(5). Mandelkow E Alzheimer’s disease: The tangled tale of tau. Nature 1999, 402 (6762), 588–589.
[PubMed: 10604460]
(6). Schraen-Maschke S; Sergeant N; Dhaenens CM; Bombois S; Deramecourt V; Caillet-Boudin ML;
Pasquier F; Maurage CA; Sablonniere B; Vanmechelen E; Buee L Tau as a biomarker of
neurodegenerative diseases. Biomarkers Med. 2008, 2 (4), 363–384.
(7). Ellington AD; Szostak JW In vitro selection of RNA molecules that bind specific ligands. Nature
1990, 346 (6287), 818–822. [PubMed: 1697402]
(8). Tuerk C; Gold L Systematic evolution of ligands by exponential enrichment: RNA ligands to
bacteriophage T4 DNA polymerase. Science 1990, 249 (4968), 505–510. [PubMed: 2200121]
(9). Qu H; Csordas AT; Wang J; Oh SS; Eisenstein MS; Soh HT Rapid and label-free strategy to isolate
aptamers for metal ions. ACS Nano 2016, 10 (8), 7558–7565. [PubMed: 27399153]
(10). Wilson C; Szostak JW Isolation of a fluorophore-specific DNA aptamer with weak redox activity.
Chem. Biol 1998, 5 (11), 609–617. [PubMed: 9831529]
(11). Famulok M Molecular recognition of amino acids by RNA-aptamers: An L-citrulline binding
RNA motif and its evolution into an L-arginine binder. J. Am. Chem. Soc 1994, 116 (5), 1698–
1706.
(12). Lauhon CT; Szostak JW RNA aptamers that bind flavin and nicotinamide redox cofactors. J. Am.
Chem. Soc 1995, 117 (4), 1246–1257. [PubMed: 11539282]
(13). Niazi JH; Lee SJ; Gu MB Single-stranded DNA aptamers specific for antibiotics tetracyclines.
Bioorg. Med. Chem 2008, 16 (15), 7245–7253. [PubMed: 18617415]
(14). Paige JS; Nguyen-Duc T; Song W; Jaffrey SR Fluorescence imaging of cellular metabolites with
RNA. Science 2012, 335 (6073), 1194. [PubMed: 22403384]
(15). Ferreira CSM; Matthews CS; Missailidis S DNA aptamers that bind to MUC1 tumour marker:
design and characterization of MUC1-binding single-stranded DNA sptamers. Tumor Biol. 2006,
27 (6), 289–301.
(16). Bock LC; Griffin LC; Latham JA; Vermaas EH; Toole JJ Selection of single-stranded DNA
molecules that bind and inhibit human thrombin. Nature 1992, 355 (6360), 564–566. [PubMed:
1741036]
(17). Balogh Z; Lautner G; Bardóczy V; Komorowska B; Gyurcsányi RE; Mészáros T Selection and
versatile application of virus-specific aptamers. FASEB J. 2010, 24 (11), 4187–4195. [PubMed:
20624933]
(18). Tawaraya Y; Hyodo M; Ara MN; Yamada Y; Harashima H RNA aptamers for targeting
mitochondria using a mitochondria-based SELEX method. Biol. Pharm. Bull 2014, 37 (8), 1411–
1415. [PubMed: 25087963]
(19). Shangguan D; Li Y; Tang Z; Cao ZC; Chen HW; Mallikaratchy P; Sefah K; Yang CJ; Tan W
Aptamers evolved from live cells as effective molecular probes for cancer study. Proc. Natl.
Acad. Sci. U. S. A 2006, 103 (32), 11838–11843. [PubMed: 16873550]
(20). Conrad R; Keranen LM; Ellington AD; Newton AC Isozyme-specific inhibition of protein kinase
C by RNA aptamers. J. Biol. Chem 1994, 269 (51), 32051–32054. [PubMed: 7528207]
Teng et al. Page 14
J Am Chem Soc
. Author manuscript; available in PMC 2019 October 31.
Author Manuscript Author Manuscript Author Manuscript Author Manuscript
(21). Chen L; Rashid F; Shah A; Awan HM; Wu M; Liu A; Wang J; Zhu T; Luo Z; Shan G The
isolation of an RNA aptamer targeting to p53 protein with single amino acid mutation. Proc. Natl.
Acad. Sci. U. S. A 2015, 112 (32), 10002–10007. [PubMed: 26216949]
(22). Shoji A; Kuwahara M; Ozaki H; Sawai H Modified DNA aptamer that binds the (R)-isomer of a
thalidomide derivative with high enantioselectivity. J. Am. Chem. Soc 2007, 129 (5), 1456–1464.
[PubMed: 17263432]
(23). Griffin LC; Tidmarsh GF; Bock LC; Toole JJ; Leung LL In vivo anticoagulant properties of a
novel nucleotide-based thrombin inhibitor and demonstration of regional anticoagulation in
extracorporeal circuits. Blood 1993, 81 (12), 3271–3276. [PubMed: 8507864]
(24). Bing T; Yang X; Mei H; Cao Z; Shangguan D Conservative secondary structure motif of
streptavidin-binding aptamers generated by different laboratories. Bioorg. Med. Chem 2010, 18
(5), 1798–1805. [PubMed: 20153201]
(25). Selinfreund RH; Barger SW; Pledger WJ; Van Eldik LJ Neurotrophic protein S100 beta
stimulates glial cell proliferation. Proc. Natl. Acad. Sci. U. S. A 1991, 88 (9), 3554–3558.
[PubMed: 1902567]
(26). Bishop P; Rocca D; Henley JM Ubiquitin C-terminal hydrolase L1 (UCH-L1): Structure,
distribution and roles in brain function and dysfunction. Biochem. J 2016, 473 (16), 2453–2462.
[PubMed: 27515257]
(27). Wang KK; Yang Z; Zhu T; Shi Y; Rubenstein R; Tyndall JA; Manley GT An update on diagnostic
and prognostic biomarkers for traumatic brain injury. Expert Rev. Mol. Diagn 2018, 18 (2), 165–
180.
(28). Lim S; Haque MM; Kim D; Kim DJ; Kim YK Cell-based models to investigate tau aggregation.
Comput. Struct. Biotechnol J. 2014, 12 (20–21), 7–13. [PubMed: 25505502]
(29). Cho JH; Johnson GV Primed phosphorylation of tau at Thr231 by glycogen synthase kinase
3beta (GSK3beta) plays a critical role in regulating tau’s ability to bind and stabilize
microtubules. J. Neurochem 2004, 88 (2), 349–358. [PubMed: 14690523]
(30). Krylova SM; Musheev M; Nutiu R; Li Y; Lee G; Krylov SN Tau protein binds single-stranded
DNA sequence specifically – the proof obtained in vitro with non-equilibrium capillary
electrophoresis of equilibrium mixtures. FEBS Lett. 2005, 579 (6), 1371–1375. [PubMed:
15733843]
(31). Kim JH; Kim E; Choi WH; Lee J; Lee JH; Lee H; Kim DE; Suh YH; Lee MJ Inhibitory RNA
aptamers of tau oligomerization and their neuroprotective roles against proteotoxic stress. Mol.
Pharmaceutics 2016, 13 (6), 2039–2048.
(32). Trzeciakiewicz H; Tseng JH; Wander CM; Madden V; Tripathy A; Yuan CX; Cohen TJ A dual
pathogenic mechanism links tau acetylation to sporadic tauopathy. Sci. Rep 2017, 7, 44102.
[PubMed: 28287136]
(33). Li C; Götz J Tau-based therapies in neurodegeneration: opportunities and challenges. Nat. Rev.
Drug Discovery 2017, 16 (12), 863–883. [PubMed: 28983098]
(34). Kimura T; Hatsuta H; Masuda-Suzukake M; Hosokawa M; Ishiguro K; Akiyama H; Murayama
S; Hasegawa M; Hisanaga S The abundance of nonphosphorylated tau in mouse and human
tauopathy brains revealed by the use of phos-tag method. Am. J. Pathol 2016, 186 (2), 398–409.
[PubMed: 26687814]
(35). Stein CA; Castanotto D FDA-approved oligonucleotide therapies in 2017. Mol. Ther 2017, 25
(5), 1069–1075. [PubMed: 28366767]
(36). Dassie JP; Giangrande PH Current progress on aptamer-targeted oligonucleotide therapeutics.
Ther. Delivery 2013, 4 (12), 1527–1546.
(37). Röthlisberger P; Gasse C; Hollenstein M Nucleic acid aptamers: emerging applications in
medical imaging, nanotechnology, neurosciences, and drug delivery. Int. J. Mol. Sci 2017, 18
(11), 2430.
Teng et al. Page 15
J Am Chem Soc
. Author manuscript; available in PMC 2019 October 31.
Author Manuscript Author Manuscript Author Manuscript Author Manuscript
Figure 1.
Tau aptamer discovery by SELEX and characterization workflow. Four peptide epitopes
from tau and their corresponding phosphorylated peptides were used as putative targets. The
enrichment was monitored using flow cytometry. The most abundant 10 candidates were
characterized after deep sequencing. The binding affinity/specificity and the inhibitory
effects with the identified aptamers were then verified.
Teng et al. Page 16
J Am Chem Soc
. Author manuscript; available in PMC 2019 October 31.
Author Manuscript Author Manuscript Author Manuscript Author Manuscript
Figure 2.
Primary binding analysis between aptamer candidates and their putative peptide targets. (A-
E) Representative predicted secondary structures of aptamer IT1 and its subsequent
truncated forms, IT1a, IT1b, IT1c, and IT1d. (F and M–P) Full-length sequences IT1, IT4,
IT5, IT6, and IT9 are aptamers that specifically bind to T231 peptide. (F–J) The binding
abilities of truncated aptamers IT1a, IT1b, and IT1c to T231 epitope are not seriously
compromised when compared to the full-length aptamer IT1. However, binding strength is
lost when further truncating the stem of IT1c into IT1d. (K) Sequence IT2 demonstrates a
unique binding profile on three of the peptides (T231, T231P, and S202). (L) Sequence IT3
bypasses the phosphorylated site and recognizes both T231 and T231P.
Teng et al. Page 17
J Am Chem Soc
. Author manuscript; available in PMC 2019 October 31.
Author Manuscript Author Manuscript Author Manuscript Author Manuscript
Figure 3.
Binding specificity of the tau aptamers toward tau protein. Gel electrophoresis of each
FITC-labeled tau aptamer, (A) IT1, (B) IT2, (C) IT3, (D) IT4, (E) IT5, (F) IT6, and (G) IT9,
after incubation with the protein of interest. Aptamers form binding complexes with tau and
therefore show a smaller migration in the presence of tau. S100B, UCH-L1,
α
casein,
β
casein, BSA, and IgG are nontarget reference proteins and have no retention effect on the
migration of aptamers under electrophoresis.
Teng et al. Page 18
J Am Chem Soc
. Author manuscript; available in PMC 2019 October 31.
Author Manuscript Author Manuscript Author Manuscript Author Manuscript
Figure 4.
Aptamer-antibody sandwich ELISA for detection of tau. Both tau and phosphorylated tau
can be captured by total tau antibody. However, aptamer IT4 only detects the presence of
nonphosphorylated T231 residue. The scrambled random sequence (RS) and phosphorylated
tau protein are included in this test as negative controls. All the sequences are labeled with
biotin, which later reacts with streptavidin-labeled polyHRP enzyme and TMB substrate to
reveal the colorimetric reaction products.
Teng et al. Page 19
J Am Chem Soc
. Author manuscript; available in PMC 2019 October 31.
Author Manuscript Author Manuscript Author Manuscript Author Manuscript
Figure 5.
Inhibitory effects of tau-binding aptamers on phosphorylation of Tau441 in vitro. (A) In
vitro tau phosphorylation and oligomerization assay was performed by incubating Tau441 (1
μ
g) with aptamers (50
μ
M) for 1 h, followed by incubation with GSK3
β
(200 ng) for 16 h.
Samples were analyzed by SDS-PAGE, followed by Western blotting with phospho-tau
antibody (RZ3). (B) Quantification of monomeric p-tau (70K) and (C) oligomeric p-tau (150
K) bands of tau protein. The monomeric p-tau (70K) and oligomeric tau (150 K) bands were
normalized by the levels of GSK3
β
-phosphorylated Tau441 and are shown as percentage.
*
p
< 0.05, **
p
< 0.01, ***
p
< 0.001 and ****
p
< 0.0001 (
n
= 3, two-way ANOVA).
Teng et al. Page 20
J Am Chem Soc
. Author manuscript; available in PMC 2019 October 31.
Author Manuscript Author Manuscript Author Manuscript Author Manuscript
Figure 6.
In vitro inhibitory effects of aptamers on oligomerization of Tau441. (A) In vitro tau
oligomerization assay was performed using Tau441 (10
μ
M), heparin (5 mg/mL) and
aptamers (50
μ
M) for 16 h. Samples were analyzed by SDS-PAGE, followed by Western
blotting with total tau antibody (DAKO). (B) Quantification of oligomeric tau (150 K). (C)
In vitro dose-response effect of IT3 on heparin-induced tau oligomerization assay as
detected by DA9 antibody. The reaction was performed in the presence of IT3 aptamer (1,
10, 25, 50
μ
M) for 24 h. The oligomeric tau bands (150 K) were normalized by the level of
heparin-induced tau oligomers only and are shown as percentage. *
p
< 0.05, **
p
< 0.01,
***
p
< 0.001 and ****
p
< 0.0001 (
n
= 3, two-way ANOVA).
Teng et al. Page 21
J Am Chem Soc
. Author manuscript; available in PMC 2019 October 31.
Author Manuscript Author Manuscript Author Manuscript Author Manuscript
... To target tauopathy-associated neurodegenerative diseases, we utilized the laboratory in vitro evolution (LIVE) method to further evolve aptamers that bind Tau or Tau epitopes in our group [18] or in other laboratories. [19] Previously selected DNA or RNA Tau aptamers exhibited high affinities against the target but did not guarantee optimal inhibition or modulation of Tau oligomerization/ aggregation. ...
... The selection was then performed with the preselected and enriched Tau binding DNA pool from the third round of GE-Selection as starting library. To generate Tau aggregates in cell-free conditions, we first attempted to use arachidonic acid (AA) [22] and heparin [18] to induce tau aggregation. AA was found to be a robust inducer of Tau oligomerization/aggregation since it promoted Tau aggregation in a dose-dependent manner (50-100 μM), as shown by denaturing SDS-PAGE ( Figure 2C). ...
... After the successful truncation and modification of the BW1 aptamer to obtain the BW1c aptamer with higher affinity, we asked whether the resulting aptamer would retain its inhibitory function which had previously guided the functional selection process. Therefore, we investigated the inhibitory ability of BW1c, together with BW1 and a previously discovered Tau aptamer, IT2a (from our previous work based on Tauepitope binding Selection [18] ), to compare the inhibitory ability of aptamers developed from binding-criteria-based selection vs. functional selection. Thioflavin-T (ThT [24] ) intercalates directly with β-sheet structures in self-assembled, multimeric proteins. ...
Article
Full-text available
Pathological hyperphosphorylation and aggregation of microtubule‐associated Tau protein contribute to Alzheimer's Disease (AD) and other related tauopathies. Currently, no cure exists for Alzheimer's Disease. Aptamers offer significant potential as next‐generation therapeutics in biotechnology and the treatment of neurological disorders. Traditional aptamer selection methods for Tau protein focus on binding affinity rather than interference with pathological Tau. In this study, we developed a new selection strategy to enrich DNA aptamers that bind to surviving monomeric Tau protein under conditions that would typically promote Tau aggregation. Employing this approach, we identified a set of aptamer candidates. Notably, BW1c demonstrates a high binding affinity (Kd = 6.6 nM) to Tau protein and effectively inhibits arachidonic acid (AA)‐induced Tau protein oligomerization and aggregation. Additionally, it inhibits GSK3β‐mediated Tau hyperphosphorylation in cell‐free systems and okadaic acid‐mediated Tau hyperphosphorylation in cellular milieu. Lastly, retro‐orbital injection of BW1c tau aptamer shows the ability to cross the blood brain barrier and gain access to neuronal cell body. Through further refinement and development, these Tau aptamers may pave the way for a first‐in‐class neurotherapeutic to mitigate tauopathy‐associated neurodegenerative disorders.
... Alumina polishing powder (Al 2 O 3 ) was purchased from Wuhan GOSL Technology Co., Ltd. Anhydrous ethanol, P-tau231 aptamer (sequence 5′-HS-(CH 2 ) 6 -CAG CAC CGT CAA CTG AAT GGG GAG TGG TGG GGC GGG GGC CGG ATC CGT GAT GCG ATG GAG ATGT-3′ [32,33], synthesized by Sangon Biotech, in Shanghai), and Tau-441-GSK3β-phosphorylated (was calibrated as P-tau 231) were obtained from Signalchem. Bovine serum albumin (BSA), fibrinogen, cancer antigen 19-9 (CA19-9), and γ-globulin were sourced from Aladdin Reagent Shanghai Co., Ltd. ...
Article
Full-text available
The instant screening of patients with a tendency towards developing Alzheimer’s disease (AD) is significant for providing preventive measures and treatment. However, the current imaging-based technology cannot meet the requirements in the early stage. Developing biosensor-based liquid biopsy technology could be overcoming this bottleneck problem. Herein, we developed a simple, low-cost, and sensitive electrochemical aptamer biosensor for detecting phosphorylated tau protein threonine 231 (P-tau231), the earliest and one of the most efficacious abnormally elevated biomarkers of AD. Gold nanoparticles (AuNPs) were electrochemically synthesized on a glassy carbon electrode as the transducer, exhibiting excellent conductivity, and were applied to amplify the electrochemical signal. A nucleic acid aptamer was designed as the receptor to capture the P-tau231 protein, specifically through the formation of an aptamer-antigen complex. The proposed biosensor showed excellent sensitivity in detecting P-tau 231, with a broad linear detection range from 10 to 10⁷ pg/mL and a limit of detection (LOD) of 2.31 pg/mL. The recoveries of the biosensor in human serum ranged from 97.59 to 103.26%, demonstrating that the biosensor could be used in complex practical samples. In addition, the results showed that the developed biosensor has good repeatability, reproducibility, and stability, which provides a novel method for the early screening of AD. Graphical Abstract
... The ssDNA library contained 40 random nucleotides flanked by two primer binding sequences referring to a previous study. 49 The sequences of primers for aptamer amplifying were forward, 5'-CAGCACCGTCAACTGAAT-3', and reverse, 5'-ACATCTCCATCGCATCAC-3'. ...
Article
Full-text available
Immune checkpoint blockade has become an effective approach to reverse the immune tolerance of tumor cells. Indoleamine 2,3-dioxygenase 1 (IDO1) is frequently upregulated in many types of cancers and contributes to the establishment of an immunosuppressive cancer microenvironment, which has been thought to be a potential target for cancer therapy. However, the development of IDO1 inhibitors for clinical application is still limited. Here, we isolated a DNA aptamer with a strong affinity and inhibitory activity against IDO1, designated as IDO-APT. By conjugating with nanoparticles, in situ injection of IDO-APT to CT26 tumor-bearing mice significantly suppresses the activity of regulatory T cells and promotes the function of CD8+ T cells, leading to tumor suppression and prolonged survival. Therefore, this functional IDO1-specific aptamer with potent anti-tumor effects may serve as a potential therapeutic strategy in cancer immunotherapy. Our data provide an alternative way to target IDO1 in addition to small molecule inhibitors.
... Despite the high performance of aptamers, only a handful of studies have been conducted using aptamers as bio-receptors for the PEC detection of AD biomarkers. To date, a few aptamer sequences recognizing AD biomarkers (such as Aβ 40 peptides, Aβ oligomers and tau proteins) have been identified [55][56][57]. Further studies are needed to develop additional aptamer sequences specific for AD biomarkers including neurofilament light (NfL) proteins. ...
Article
Alzheimer's disease (AD) is the most prevalent cause of dementia, affecting one in nine people aged 65 years and older worldwide. Pathological hallmarks include the accumulation of senile plaques and neurofibrillary tangles, which are associated with changes in the levels of AD biomarkers, such as amyloid peptides and tau proteins. As the neuropathological process of AD commences decades before the onset of cognitive symptoms, an accurate assessment of AD biomarkers is critical for the early identification of at-risk AD patients. Photoelectrochemical (PEC) bioanalysis is a promising sensing methodology that exhibits superior sensing performance owing to the complete separation of the excitation source and detection signal. Numerous efforts have been made to develop PEC analytical platforms that target AD biomarkers. This review provides an overview of the recent advances in PEC sensing of AD biomarkers in body fluids. The major components of PEC sensing platforms (such as photoactive transducers, affinity agents, and targeted AD biomarkers) and PEC signaling strategies for detecting AD biomarkers are outlined. In addition, we summarize the current issues and strategies for advancing PEC sensing techniques to meet the ever-increasing demand for the early diagnosis of AD.
Article
Full-text available
Pathological hyperphosphorylation and aggregation of microtubule‐associated Tau protein contribute to Alzheimer's Disease (AD) and other related tauopathies. Currently, no cure exists for Alzheimer's Disease. Aptamers offer significant potential as next‐generation therapeutics in biotechnology and the treatment of neurological disorders. Traditional aptamer selection methods for Tau protein focus on binding affinity rather than interference with pathological Tau. In this study, we developed a new selection strategy to enrich DNA aptamers that bind to surviving monomeric Tau protein under conditions that would typically promote Tau aggregation. Employing this approach, we identified a set of aptamer candidates. Notably, BW1c demonstrates a high binding affinity (Kd = 6.6 nM) to Tau protein and effectively inhibits arachidonic acid (AA)‐induced Tau protein oligomerization and aggregation. Additionally, it inhibits GSK3β‐mediated Tau hyperphosphorylation in cell‐free systems and okadaic acid‐mediated Tau hyperphosphorylation in cellular milieu. Lastly, retro‐orbital injection of BW1c tau aptamer shows the ability to cross the blood brain barrier and gain access to neuronal cell body. Through further refinement and development, these Tau aptamers may pave the way for a first‐in‐class neurotherapeutic to mitigate tauopathy‐associated neurodegenerative disorders.
Article
Full-text available
Aptamers developed using in vitro Systematic Evolution of Ligands by Exponential Enrichment (SELEX) technology are single-stranded nucleic acids 10-100 nucleotides in length. Their targets, often with specificity and high affinity, range from ions and small molecules to proteins and 26 other biological molecules, as well as to larger systems, including cells, tissues, and animals. Ap- 27tamers often rival conventional antibodies with improved performance, due to aptamers’ unique biophysical and biochemical properties, including small size, synthetic accessibility, facile modify- 29cation, low production cost, and low immunogenicity. Therefore, there is sustained interest in engi- 30neering and adapting aptamers for many applications, including diagnostics and therapeutics. Recently, aptamers have shown promise as early diagnostic biomarkers and in precision medicine for neurodegenerative and neurological diseases. Here, we critically review neuro-targeting aptamers and their potential applications in neuroscience research, neuro-diagnostics, and neuromedicine. We also discuss challenges that must be overcome, including delivery across the blood-brain barrier, increased affinity, and improved in vivo stability and in vivo pharmacokinetic properties.
Article
DNA walkers have been widely explored and applied as biosensor elements to detect disease-related biomarkers. Traditional interface-anchored DNA walkers typically have a fixed swing arm range and an orientation of the preset track, which might complicate the design of a sensor system and limit its application in more scenes. We propose a simple electrochemical aptasensor to accurately detect Alzheimer's disease (AD) based on a nicking enzyme-powered DNA walker. In this method, bifunctional magnetic nanoparticles are used to identify and capture Aβ oligomers (AβO) and Tau and release the DNA walker. As the DNA walker moves freely on the surface of the electrode, the nicking enzymes circularly cleave and release the two signal substrate chains, significantly amplifying the signal. It has been demonstrated that the constructed sensor can sensitively detect AβO and Tau, and the combined analysis of dual markers improves the accuracy of AD diagnosis. Furthermore, this method can distinguish normal individuals from AD patients in real cerebrospinal fluid samples. The excellent performance of this biosensor makes it promising for clinical applications in diagnosing AD patients and prognosis assessment.
Article
Full-text available
Aptamers associated with cancer targeting therapy are commonly focused on cell membrane proteins; however, the study of intracellular, particularly, nuclear proteins is limited. The nuclear phosphatase PAC1 has been reported to be a potential T cell-related immunotherapeutic target. Here, we identified an aptamer, designated as PA5, with high affinity and specificity for PAC1 through the systematic evolution of ligands by exponential enrichment (SELEX) procedure. We then developed a dual-module aptamer PAC1-AS consisting of a cell-internalizing module and a targeting module, which can recognize PAC1 in the nucleus under physiological conditions. This modularized aptamer raises the possibility of manipulating endosomes and provides insights into the exploration and development of an efficient cancer immunotherapy approach.
Article
Full-text available
Introduction: Traumatic brain injury (TBI) is a major worldwide neurological disorder of epidemic proportions. To date, there are still no FDA-approved therapies to treat any forms of TBI. Encouragingly, there are emerging data showing that biofluid-based TBI biomarker tests have the potential to diagnose the presence of TBI of different severities including concussion, and to predict outcome. Area covered: The authors provide an update on the current knowledge of TBI biomarkers, including protein biomarkers for neuronal cell body injury (UCH-L1, NSE), astroglial injury (GFAP, S100B), neuronal cell death (αII-spectrin breakdown products), axonal injury (NF proteins), white matter injury (MBP), post-injury neurodegeneration (total Tau and phospho-Tau), post-injury autoimmune response (brain antigen-targeting autoantibodies), and other emerging non-protein biomarkers. The authors discuss biomarker evidence in TBI diagnosis, outcome prognosis and possible identification of post-TBI neurodegernative diseases (e.g. chronic traumatic encephalopathy and Alzheimer’s disease), and as theranostic tools in pre-clinical and clinical settings. Expert Commentary: A spectrum of biomarkers is now at or near the stage of formal clinical validation of their diagnostic and prognostic utilities in the management of TBI of varied severities including concussions. TBI biomarkers could serve as a theranostic tool in facilitating drug development and treatment monitoring.
Article
Full-text available
Recent progresses in organic chemistry and molecular biology have allowed the emergence of numerous new applications of nucleic acids that markedly deviate from their natural functions. Particularly, DNA and RNA molecules—coined aptamers—can be brought to bind to specific targets with high affinity and selectivity. While aptamers are mainly applied as biosensors, diagnostic agents, tools in proteomics and biotechnology, and as targeted therapeutics, these chemical antibodies slowly begin to be used in other fields. Herein, we review recent progress on the use of aptamers in the construction of smart DNA origami objects and MRI and PET imaging agents. We also describe advances in the use of aptamers in the field of neurosciences (with a particular emphasis on the treatment of neurodegenerative diseases) and as drug delivery systems. Lastly, the use of chemical modifications, modified nucleoside triphosphate particularly, to enhance the binding and stability of aptamers is highlighted.
Article
Full-text available
Oligonucleotides (oligos) have been under clinical development for approximately the past 30 years, beginning with antisense oligonucleotides (ASOs) and apatmers and followed about 15 years ago by siRNAs. During that lengthy period of time, numerous clinical trials have been performed and thousands of trial participants accrued onto studies. Of all the molecules evaluated as of January 2017, the regulatory authorities assessed that six provided clear clinical benefit in rigorously controlled trials. The story of these six is given in this review.
Article
Full-text available
Tau acetylation has recently emerged as a dominant post-translational modification (PTM) in Alzheimer’s disease (AD) and related tauopathies. Mass spectrometry studies indicate that tau acetylation sites cluster within the microtubule (MT)-binding region (MTBR), suggesting acetylation could regulate both normal and pathological tau functions. Here, we combined biochemical and cell-based approaches to uncover a dual pathogenic mechanism mediated by tau acetylation. We show that acetylation specifically at residues K280/K281 impairs tau-mediated MT stabilization, and enhances the formation of fibrillar tau aggregates, highlighting both loss and gain of tau function. Full-length acetylation-mimic tau showed increased propensity to undergo seed-dependent aggregation, revealing a potential role for tau acetylation in the propagation of tau pathology. We also demonstrate that methylene blue, a reported tau aggregation inhibitor, modulates tau acetylation, a novel mechanism of action for this class of compounds. Our study identifies a potential “two-hit” mechanism in which tau acetylation disengages tau from MTs and also promotes tau aggregation. Thus, therapeutic approaches to limit tau K280/K281 acetylation could simultaneously restore MT stability and ameliorate tau pathology in AD and related tauopathies.
Article
Full-text available
Ubiquitin C-terminal hydrolase L1 (UCH-L1) is an extremely abundant protein in the brain where, remarkably, it is estimated to make up 1–5% of total neuronal protein. Although it comprises only 223 amino acids it has one of the most complicated 3D knotted structures yet discovered. Beyond its expression in neurons UCH-L1 has only very limited expression in other healthy tissues but it is highly expressed in several forms of cancer. Although UCH-L1 is classed as a deubiquitinating enzyme (DUB) the direct functions of UCH-L1 remain enigmatic and a wide array of alternative functions has been proposed. UCH-L1 is not essential for neuronal development but it is absolutely required for the maintenance of axonal integrity and UCH-L1 dysfunction is implicated in neurodegenerative disease. Here we review the properties of UCH-L1, and how understanding its complex structure can provide new insights into its roles in neuronal function and pathology.
Article
Using a novel in vitro selection/amplification technique, we have recently identified a new class of thrombin inhibitors based on single- stranded DNA oligonucleotides. One oligonucleotide, GGTTGGTGTGGTTGG (thrombin, aptamer), showed potent anticoagulant activity in vitro. We have initiated pharmacologic studies in cynomolgus monkeys to study the thrombin aptamer's in vivo anticoagulant properties. Upon infusion of the thrombin aptamer, anticoagulation was rapidly achieved, with a plateau reached within 10 minutes. There was a linear dose-response relationship between thrombin aptamer infusion rate and prolongation of plasma prothrombin time. Ten minutes after the infusion was stopped, no prolongation of prothrombin time was observed, indicating that the thrombin aptamer has an extremely short in vivo half-life, estimated to be 108 +/- 14 seconds. In addition, inhibition of thrombin-induced platelet aggregation in platelet-rich plasma was observed ex vivo without an effect on collagen-induced aggregation, indicating that the inhibition was specific for thrombin and not due to a nonspecific inhibitory effect on platelets. To exploit the short in vivo half-life of the thrombin aptamer, its ability to achieve regional anticoagulation in an extracorporeal hemofiltration circuit in sheep was tested. Doubling of the prothrombin time in the circuit was observed, whereas the systemic prothrombin time was minimally prolonged. We conclude that the thrombin aptamer is a potent anticoagulant in vivo, and specifically inhibits thrombin-induced platelet aggregation ex vivo. The rapid onset of action and short half- life in vivo suggest that the thrombin aptamer may be useful in anticoagulation with extracorporeal circuits and may have distinct advantages in certain acute clinical settings.
Article
Aggregates of the microtubule-associated protein tau are a defining feature of several neurodegenerative diseases that are collectively known as tauopathies, and constitute one of the hallmark lesions of Alzheimer disease (AD). Given the lack of efficacy to date of amyloid-β-targeted therapies for AD, interest is growing in tau as a potential alternative target. Several drug candidates, which are now in clinical trials, aim to reduce tau levels or to prevent the aggregation or pathological post-translation modifications of this protein. In this Review, we discuss preclinical and clinical studies in light of an increased understanding of the physiological and pathological roles of tau, advances in animal models of tauopathy, the identification of novel targets and the availability of novel tracers to track tau.
Article
Generating aptamers that bind to specific metal ions is challenging because existing aptamer discovery methods typically require chemical labels or modifications that can alter the structure and properties of the ions. In this work, we report an aptamer discovery method that enables us to generate high-quality structure-switching aptamers (SSAs) that undergo a conformational change upon binding a metal ion target, without the requirement of labels or chemical modifications. Our method is more efficient than conventional selection methods because it enables direct measurement of target binding via fluorescence-activated cell sorting (FACS), isolating only the desired aptamers with the highest affinity. Using this strategy, we obtained a highly specific DNA SSA with ∼30-fold higher affinity than the best aptamer for Hg(2+) in the literature. We also discovered DNA aptamers that bind to Cu(2+) with excellent affinity and specificity. Both aptamers were obtained within four rounds of screening, demonstrating the efficiency of our aptamer discovery method. Given the growing availability of FACS, we believe our method offers a general strategy for discovering high-quality aptamers for other ions and small-molecule targets in an efficient and reproducible manner.
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.
Article
Tauopathies are neurodegenerative diseases characterized by aggregates of hyperphosphorylated tau. Previous studies have identified many disease-related phosphorylation sites on tau. However, it is not understood how tau is hyperphosphorylated and what extent these sites are phosphorylated in both diseased and normal brains. Most previous studies have used phospho-specific antibodies to analyze tau phosphorylation. These results are useful but do not provide information about nonphosphorylated tau. Here, we applied the method of Phos-tag SDS-PAGE, in which phosphorylated tau was separated from nonphosphorylated tau in vivo. Among heterogeneously phosphorylated tau species in adult mouse brains, the nonphosphorylated 0N4R isoform was detected most abundantly. In contrast, perinatal tau and tau in cold water-stressed mice were all phosphorylated with a similar extent of phosphorylation. In normal elderly human brains, nonphosphorylated 0N3R and 0N4R tau were most abundant. A slightly higher phosphorylation of tau, which may represent the early step of hyperphosphorylation, was increased in Alzheimer disease patients at Braak stage V. Tau with this phosphorylation state was pelleted by centrifugation, and sarkosyl-soluble tau in either Alzheimer disease or corticobasal degeneration brains showed phosphorylation profiles similar to tau in normal human brain, suggesting that hyperphosphorylation occurs in aggregated tau. These results indicate that tau molecules are present in multiple phosphorylation states in vivo, and nonphosphorylated forms are highly expressed among them.