ArticlePDF Available

Quantitative Proteomic Analysis of the Hippocampus in the 5XFAD Mouse Model at Early Stages of Alzheimer's Disease Pathology

Authors:

Abstract and Figures

Alzheimer's disease (AD) is characterized by progressive memory loss accompanied by synaptic and neuronal degeneration. Although research has shown that substantial neurodegeneration occurs even during the early stages of AD, the detailed mechanisms of AD pathogenesis are largely unknown because of difficulties in diagnosis and limitations of the analytical methods. The 5XFAD mouse model harbors five early-onset familial AD (FAD) mutations and displays substantial amyloid plaques and neurodegeneration. Here, we use quantitative mass spectrometry to identify proteome-wide changes in the 5XFAD mouse hippocampus during the early stages of AD pathology. A subset of the results was validated with immunoblotting. We found that the 5XFAD mice display higher expression of ApoE, ApoJ (clusterin), and nicastrin, three important proteins in AD that are known to participate in amyloid-β processing and clearance, as well as the neurological damage/glial marker protein GFAP and other proteins. A large subset of the proteins that were up- or downregulated in 5XFAD brains have been implicated in neurological disorders and cardiovascular disease, suggesting an association between cardiovascular disease and AD. Common upstream regulator analysis of upregulated proteins suggested that the XBP1, NRF2, and p53 transcriptional pathways were activated, as was IGF-1R signaling. Protein interactome analysis revealed an interconnected network of regulated proteins, with two major sub-networks centered on AβPP processing membrane complexes and mitochondrial proteins. Together with a recent study on the transcriptome of 5XFAD mice, our study allows a comprehensive understanding of the molecular events occurring in 5XFAD mice during the early stages of AD pathology.
Content may be subject to copyright.
Journal of Alzheimer’s Disease 36 (2013) 321–334
DOI 10.3233/JAD-130311
IOS Press
321
Quantitative Proteomic Analysis of the
Hippocampus in the 5XFAD Mouse Model at
Early Stages of Alzheimer’s Disease
Pathology
Ingie Honga,1, Taewook Kangb,c,1, YongCheol Yooc,1, Royun Parkd,1, Junuk Leea, Sukwon Leea,
Jeongyeon Kima, Boemjong Songa, Se-Young Kimc, Minho Moond,KiNaYun
c, Jin Young Kimc,
Inhee Mook-Jungd,, Young Mok Parkb,c,and Sukwoo Choia,
aSchool of Biological Sciences, College of Natural Sciences, Seoul National University, Seoul, Korea
bGraduate School of Analytical Science and Technology (GRAST), Chungnam National University, Daejeon, Korea
cMass Spectrometry Research Center, Korea Basic Science Institute, Ochang, Korea
dDepartment of Biochemistry and Molecular Biology, Seoul National University College of Medicine, Seoul, Korea
Handling Associate Editor: Bhumsoo Kim
Accepted 18 March 2013
Abstract. Alzheimer’s disease (AD) is characterized by progressive memory loss accompanied by synaptic and neuronal
degeneration. Although research has shown that substantial neurodegeneration occurs even during the early stages of AD, the
detailed mechanisms of AD pathogenesis are largely unknown because of difficulties in diagnosis and limitations of the analytical
methods. The 5XFAD mouse model harbors five early-onset familial AD (FAD) mutations and displays substantial amyloid
plaques and neurodegeneration. Here, we use quantitative mass spectrometry to identify proteome-wide changes in the 5XFAD
mouse hippocampus during the early stages of AD pathology. A subset of the results was validated with immunoblotting. We
found that the 5XFAD mice display higher expression of ApoE, ApoJ (clusterin), and nicastrin, three important proteins in AD
that are known to participate in amyloid-processing and clearance, as well as the neurological damage/glial marker protein
GFAP and other proteins. A large subset of the proteins that were up- or downregulated in 5XFAD brains have been implicated in
neurological disorders and cardiovascular disease, suggesting an association between cardiovascular disease and AD. Common
upstream regulator analysis of upregulated proteins suggested that the XBP1, NRF2, and p53 transcriptional pathways were
activated, as was IGF-1R signaling. Protein interactome analysis revealed an interconnected network of regulated proteins, with
two major sub-networks centered on APP processing membrane complexes and mitochondrial proteins. Together with a recent
study on the transcriptome of 5XFAD mice, our study allows a comprehensive understanding of the molecular events occurring
in 5XFAD mice during the early stages of AD pathology.
Keywords: Alzheimer’s disease, amyloid-, memory, mild cognitive impairment, PS1, proteomics, transgenic mice
1These authors contributed equally to this work.
Correspondence to: Sukwoo Choi, Ph.D., School of Biological
Sciences, College of Natural Sciences, Seoul National Uni-
versity, Seoul 151-742, Korea. E-mail: sukwoo12@snu.ac.kr.;
Young Mok Park, Ph.D., Mass Spectrometry Research Center,
Korea Basic Science Institute, Ochang 363-883, Korea. E-mail:
ympark@kbsi.re.kr.; Inhee Mook-Jung, Ph.D., Department of Bio-
chemistry and Biomedical Sciences, Seoul National University
College of Medicine, 28 Yungun-dong, Jongno-gu, Seoul 110–799,
Korea. E-mail: inhee@snu.ac.kr.
ISSN 1387-2877/13/$27.50 © 2013 – IOS Press and the authors. All rights reserved
322 I. Hong et al. / Quantitative Proteomic Analysis of 5XFAD Mice
INTRODUCTION
Alzheimer’s disease (AD) is the most common form
of irreversible dementia in elderly people, tragically
affecting more than 26 million people worldwide [1].
Clinically, the disease is characterized by progres-
sive memory loss and a decline in cognitive abilities
(such as language, attention, and executive function).
A definitive diagnosis requires postmortem assess-
ment of the two histological hallmarks of AD: neuritic
amyloid-(A) plaques and neurofibrillary tangles
(NFTs) [2, 3]. Several symptomatic treatments exist
for AD; however, no disease-modifying therapies are
currently available [4–6]. Interestingly, the onset of AD
is preceded by a period of mild cognitive impairment
(MCI) during which individuals display only mild
disturbances in cognitive functions such as episodic
memory [7]. Aaccumulation and aggregation, along
with various synaptic deficits, are widely found in
MCI, but no clinical method is available for identi-
fying or treating prodromal AD [8]. A recent study
even shows that Adeposition may start as early
as 15 years before symptoms begin [9]. Disease-
modifying treatments for AD, when they become
available, will thus need to intervene early in the course
of the disease, before neurodegeneration is too severe
and thus irreversible, which underscores the impor-
tance of understanding the early pathogenesis of AD
[6].
Ample genetic and biochemical data support the
hypothesis that Aaccumulation and aggregation in
the brain are early and important events in the patho-
genesis of AD [10, 11]. Mutations associated with
early-onset familial AD (FAD) are dominantly inher-
ited and are found in the amyloid-protein precursor
(APP) gene itself or in the presenilin 1 (PS1) and
PS2 genes [3], the products of which, together with
nicastrin, Aph1, and Pen-2, are essential components
of the -secretase protein complex [12]. The 4 allele
of the apolipoprotein E (APOE) gene has been iden-
tified as the major risk factor for the more common
late-onset AD [13]. The presence of an APOE 4 allele
increases the risk of AD by approximately fourfold,
presumably by diminishing protection or augmenting
toxicity compared to other alleles [14].
Neuroinflammation is another hallmark of AD, and
activated astrocytes and microglia have been found sur-
rounding amyloid plaques [15, 16]. Due to the complex
nature of AD, it is challenging to determine whether
a given pathological structure or biochemical change
drives the disease, is a neutral bystander, or simply
represents an unsuccessful defense mechanism.
The development of quantitative proteomics has
greatly accelerated our understanding of large-scale
protein networks and the identification of impor-
tant biomarkers [17–19]. Because processes such
as alternative splicing, protein processing, and post-
translational modification are central to generating
the full complexity of life, mass-spectrometry-based
proteomics offers an unprecedented level of detail in
the biochemical assessment of biological processes.
Recent developments in chemical peptide labeling with
isobaric tags such as TMT and iTRAQ have enabled
comparison of the expression levels of thousands of
proteins across complex samples. AD pathology has
also been studied using proteomic approaches, mostly
with cerebrospinal fluid [20] or postmortem patient
samples [21]. However, there are substantial difficul-
ties with interpreting data from postmortem tissue
and from individuals of diverse genetic and patho-
logical backgrounds. For instance, the analysis of
postmortem brain tissue cannot discriminate between
changes that are specifically involved in AD versus
those that are simply a consequence of neuronal degen-
eration. Here, we employ TMT tagging to uncover the
proteome-wide changes in 5XFAD transgenic mice, a
mouse model that co-expresses five FAD mutations in
APP and PS1 [22] and develops early A42 accu-
mulation (1.5 months), amyloid deposition and gliosis
(2 months), neuron loss (2 months) [23], synapse
degeneration (4 months), spatial learning deficits (4–6
months), and increased p25 levels (9–12 months) [22,
24]. We focus on the early phase of pathology, when
synapse and memory deficits are first detected [22],
to elucidate mechanisms that are involved in the onset
and development of AD. This study broadly comple-
ments our recent study on RNA expression in 5XFAD
mice, which used next-generation RNA-Seq technol-
ogy [25].
MATERIALS AND METHODS
Experimental animals
Transgenic AD model mice with five famil-
ial AD mutations (5XFAD) were purchased from
Jackson Laboratories (strain: B6SJL-Tg [APP-
SwFlLon,PS1*M146L*L286V] 6799Vas/J, stock no.
006554) and maintained by crossing hemizygous
transgenic mice with B6SJL F1 mice [22]. Founder
transgenic mice were identified by polymerase chain
reaction (PCR), and non-transgenic littermate mice
served as controls. All mice were housed in groups
I. Hong et al. / Quantitative Proteomic Analysis of 5XFAD Mice 323
of three to five per cage, with food and water avail-
able ad libitum under standardized environmental
conditions.
5XFAD mice express both mutant human amyloid-
protein precursor (APP695) with the Swedish muta-
tion (K670N, M671L), Florida mutation (I716V), and
London mutation (V717I) and human presenilin 1
harboring two FAD mutations (M146L and L286V).
Both transgenes are expressed under the control of the
mouse Thy1 promoter to induce overexpression in the
brain [22]. These mice exhibit AD-related pathology
earlier than other animal models, and amyloid depo-
sition starts in the deep cortex and subiculum at 2
months of age. Synaptic marker proteins decrease at
4–9 months, and memory deficits are detected from 4–6
months of age [26, 27]. Importantly, amyloid plaques
and neuronal death are not detected in the hippocampus
at 4 months. Because we focused on the early events
of AD pathology, before severe neurodegeneration has
occurred, all animals in this study were killed at 4
months of age.
Animal treatment and maintenance were performed
in accordance with the Principles of Laboratory Ani-
mal Care (NIH publication No. 85–23, revised 1985)
and the Animal Care and Use Guidelines of Seoul
National University, Seoul, Korea. All efforts were
made to minimize animal suffering and to reduce the
number of mice used.
Crude synaptosome fractionation and sample
preparation
Mice were anesthetized and decapitated to extract
brain tissue. Three biological replicates from three dif-
ferent mice were generated for both the 5XFAD and
wild-type littermate groups. Hippocampal brain tissue
samples were homogenized in ice-cold homogeniza-
tion buffer containing 10 mM Tris (pH 7.6), 320 mM
sucrose, 5 mM NaF, 1 mM Na3VO4, 1 mM EDTA,
1 mM EGTA,and protease/phosphatase inhibitor cock-
tail. The homogenates were centrifuged at 1,000×
gat 4C for 10 min to remove nuclei and large
debris. The resultant supernatant was further cen-
trifuged at 10,000×gat 4C for 30 min to obtain a
crude synaptosomal fraction, which was lysed in mod-
ified RIPA buffer containing 50 mM Tris (pH 7.6),
150 mM NaCl, 5 mM NaF, 1 mM Na3VO4, 0.5% Tri-
ton X-100, 0.5% sodium deoxycholate, 0.1% SDS,
and protease/phosphatase inhibitor cocktail. Samples
were sonicated and spun down at 15,000 ×gat4
C
for 15 min. Protein content was measured by a bicin-
choninic acid assay (Thermo Scientific) following the
manufacturer’s protocol, and confirmed by SDS-PAGE
and silver staining.
TMT labeling
Samples were subsequently tagged with tan-
dem mass tags for quantitative mass spectrometry
(TMTsixplex™ Isobaric Mass Tagging Kit, Thermo
Scientific). Briefly, 100g was taken from each
sample and was reduced with 500 mM tris(2-
carboxyethyl)phosphine (TCEP) at 55C for 1 h and
then alkylated with 300 mM iodoacetamide (IAA)
at 37C in the dark for 30 min. The samples were
desalted using a 10,000 MW-cutoff membrane filter
and dissolved in 100 mM triethylammonium bicarbon-
ate (TEAB) buffer to a final concentration of 1 g/L.
Sequencing-grade trypsin (Promega, Madison, WI,
USA) was added at 1:100 (w/w) into proteins in TEAB
buffer and incubated overnight at 37C. Three sam-
ples of both transgenic 5XFAD mice and littermates
were individually labeled using TMT-126, 128, 130
(5XFAD mice), and TMT-127, 129, 131 (wild-type
littermates) following the manufacturer’s instructions.
Aqueous hydroxylamine solution (5% w/v) was added
to quench the reaction. The six samples were then
combined, speed-vacuum dried, and then dissolved in
50 L of water containing 0.1% formic acid for LC-
MS/MS analysis.
2D-LC-MS/MS
The TMT-labeled samples were analyzed using
a 2D-LC-MS/MS system consisting of a nanoAC-
QUITY UltraPerformance LC System (Waters, USA)
and an LTQ Orbitrap Elite mass spectrometer (Thermo
Scientific, USA) equipped with a nano-electrospray
source. A detailed description of 2D-LC-MS/MS anal-
ysis can be found in the literature [28, 29]. Briefly,
a strong cation exchange (5 m, 3 cm) column was
placed just before the C18 trap column (id 180 m,
length 20 mm, and particle size 5 m; Waters). Pep-
tide solutions were loaded in 5 L aliquots for each
run. Peptides were displaced from the strong cation
exchange phase to the C18 phase by a salt gradient that
was introduced through an autosampler loop and then
desalted for 10 min at a flow rate of 4 L/min. Then, the
trapped peptides were separated on a 200 mm home-
made microcapillary column consisting of C18 (Aqua;
particle size 3 m) packed into 100 m silica tubing
with an orifice id of 5 m.
An eleven-step salt gradient was performed using
3L of 0, 25, 50, 100, 250, and 500 mM ACN (0.1%
324 I. Hong et al. / Quantitative Proteomic Analysis of 5XFAD Mice
formic acid in water) and 4, 5, 9, and an additional
9L at 500 mM ACN (0.1% formic acid in 30% ACN).
The mobile phases, A and B, were composed of 0 and
100% acetonitrile, respectively, and each contained
0.1% formic acid. The LC gradient began with 5%
B for 1 min and was ramped to 20% B over 5 min, to
45% B over 90 min, to 95% B over 1 min, and remained
at 95% B over 13 min and was then ramped to 5%
B for another 5 min. The column was re-equilibrated
with 5% B for 15 min before the next run. The voltage
applied to produce an electrospray was 2.0 kV. Dur-
ing the chromatographic separation, the LTQ Orbitrap
Elite was operated in a data-dependent mode. The MS
data were acquired using the following parameters:
four data-dependent CID-high energy collision disso-
ciation (CID-HCD) dual MS/MS scans per full scan;
CID scans were acquired in LTQ with two-microscan
averaging; full scans and HCD scans were acquired in
Orbitrap at resolutions of 60,000 and 15,000, respec-
tively, with two-microscan averaging; 35% normalized
collision energy in CID and 45% normalized collision
energy in HCD; ±1 Da isolation window. Previously
fragmented ions were excluded for 60 s. In CID-HCD
dual scans, each selected parent ion was first frag-
mented by CID and then by HCD.
Protein identification, quantification, and
statistical analysis
The resultant MS/MS spectra were analyzed
against the latest murine IPI database (IPI.MOUSE.
7.26.2012). Protein identification, quantification, and
analysis were performed with Integrated Proteomics
Pipeline - IP2 (Integrated Proteomics Applications)
using ProLuCID, DTASelect2 and Census. The rate
of decoy hits in the combined forward and reversed
database was less than 1% of the forward hits. Pro-
LuCID [27] was used to identify the peptides with
the following parameters: a precursor mass error tol-
erance of 25 ppm and a fragment ion mass error of
600 ppm. The enzyme was specified as trypsin, and
three potential missed cleavages were allowed. TMT
modification at the N-terminus and lysine residues and
carbamidomethylation at cysteine residues were cho-
sen as static modifications. Oxidation at methionine
was chosen as the variable modification.
The CID and HCD tandem MS spectra from the
same precursor ion are often combined by software
to allow better peptide identification and quantifica-
tion [30]. We used in-house software in which reporter
ions from the HCD spectrum were inserted into the
CID spectrum with the same precursor ion at the pre-
vious scan. Reporter ions were extracted from small
windows (±20 ppm) around their expected m/z in the
HCD spectrum. The output data files were filtered
and sorted to compose the protein list using DTAS-
elect [31] with two or more peptide assignments for
a protein identification. Quantitative analysis was con-
ducted using Census [32], and the intensity at a reporter
ion channel for a protein was calculated as the average
of this reporter ion’s intensities from all constituent
peptides from the identified protein. The resulting
ratios were logarithmized (base = 2) to achieve a nor-
mal distribution. The median and standard deviation
were calculated, and ratio values were corrected for the
median to account for variability among different pairs
[33]. Ratios were averaged, and proteins with ratio val-
ues beyond p< 0.01 in normal distribution were defined
as significantly regulated. To further assess the indi-
vidual statistical significance of the expression level
change in each protein, one-sample t-tests were used.
Gene annotation and upstream regulator analysis
Gene Ontology annotation enrichment analysis was
performed using the DAVID Bioinformatics Resource
(v6.7) developed by NIAID, at the National Insti-
tutes of Health [34]. DAVID analysis enabled the
enrichment of functional-related gene groups and cel-
lular compartments. Ingenuity Pathway Analysis (IPA;
Ingenuity Systems, http://www.ingenuity.com) was
used to functionally annotate genes implicated in dis-
ease and search for common upstream regulators.
Network analysis for the regulated proteins
The proteins of our study were searched against the
STRING database version 9 [35] for protein-protein
interactions. Only interactions between the proteins
belonging to the increased or decreased groups were
selected. STRING defines a metric called “confidence
score” to define interaction confidence; we selected all
interactions for our regulated dataset that had a con-
fidence score 0.7 (high confidence). The resulting
interactome had 16 nodes and 30 interactions, which
we call the 5XFAD interaction network (FADIN)
(Fig. 5).
Immunoblotting
Immunochemistry was carried out to quantify spe-
cific proteins present in the samples and to validate
I. Hong et al. / Quantitative Proteomic Analysis of 5XFAD Mice 325
TMT analysis. Samples that were saved for valida-
tion after TMT analysis were loaded and separated on
polyacrylamide gels and transferred onto PVDF mem-
branes, blocked with 10% (w/v) milk, and incubated
with primary antibody for 2 h at room temperature.
The following primary antibodies were diluted in
Tris-buffered saline with 0.1% Tween-20 (TBST)
with 3% (w/v) BSA and 0.001% sodium azide:
anti-nicastrin (1:1,000; MAB5556, Millipore), anti-
GFAP (1:1,000; 18-0063, Invitrogen) antibodies were
used. Goat anti-mouse HRP or goat anti-rabbit HRP
secondary antibodies were applied for 1 h at room
temperature before staining by chemiluminescence.
The optical density for each band was quantified with
ImageJ (NIH). Quantification was performed in the
linear range of band intensity to ensure accurate quan-
tification.
RESULTS
Quantification of proteomic changes in the
hippocampus of 5XFAD mice
The hippocampi of 4-month-old 5XFAD mice and
their wild-type littermates were dissected, and pro-
teins were extracted. After TMT-labeling, the samples
were mixed, and the ratios of individual proteins were
examined by quantitative proteomics. A total of 1481
proteins were identified and quantified from 10,766
unique peptides. Tables 1 and 2 show the proteins
whose levels were strongly increased or decreased in
5XFAD mice compared to wild-type littermates. The
overexpressed protein APP showed a strong increase,
as did other well-known AD-related proteins such as
the apolipoproteins ApoE and clusterin (ApoJ), the
-secretase protein complex subcomponent nicastrin
(these three proteins are known to contribute to Apro-
cessing and clearance [36–38]), and the glial marker
protein GFAP (69 increased proteins in total). The lev-
els of proteins including Shroom2, centaurin alpha 1
(Adap1, Arf-GAP with dual PH domain-containing
protein 1), Opcml, Mapre2 (APC protein-binding
EB1 gene family homologue), and CB1 cannabinoid
receptor-interacting protein 1 (Cnrip1) were strongly
decreased in 5XFAD mice (28 decreased proteins in
total). The MS/MS spectra for two representative pro-
teins, ApoE and GFAP, are shown in Fig. 1A and B,
where the level of 5XFAD-labeled mass tags (126, 128,
130) is significantly higher than in control littermates
(127, 129, 131).
Validation of AβPP overexpression in 5XFAD
transgenic mice
Because 5XFAD mice overexpress a human mutant
form of APP (K670N, M671L, I716V, V717I), we
were able to use the resolution of mass spectrometry
to discern mouse native APP. A trypsinized peptide
unique to mouse native APP showed similar quantifi-
cation levels in 5XFAD mice and wild-type littermates
(p= 0.1111, unpaired t-test), whereas total APP was
overexpressed by approximately threefold in 5XFAD
mice (p< 0.0023), in good agreement with our prior
RNA expression study [25]. Unfortunately, the other
overexpressed protein PS1 was not quantified in our
experiments. These results verify the overexpression of
mutant APP in 5XFADmice and support the accuracy
of our quantitative analysis.
Validation of quantitative analysis with western
blot
We next validated the quantification of our
TMT analysis in a small subset of proteins, using
immunoblotting. Western blots for GFAP and nicas-
trin verified strong increases in protein expression,
correlating well with the TMT analysis (Fig. 3A, B)
and further validating our quantitative analysis with
orthogonal methodology.
Enrichment analysis of increased and decreased
proteins in 5XFAD mice
We used DAVID [34] with Gene Ontology to
functionally categorize the proteins regulated in
5XFAD mice by biological process, cellular compo-
nent, and molecular function (Fig. 4). The results
for biological process showed strong enrichment
of oxidative stress-related proteins, along with ion
homeostasis, protein folding, nerve impulse transmis-
sion, and axon ensheathment-related proteins. Cellular
component analysis showed strong enrichment of
endoplasmic reticulum (ER), pigment granule, vesi-
cle, membrane-enclosed lumen, and mitochondrial
proteins. Molecular function annotation revealed that
protein disulfide isomerase (PDI) activity, which is
mostly localized to the ER, is enriched in the regulated
proteins of 5XFAD mice, along with oxidoreduc-
tase activity, dehydrogenase activity, cofactor binding,
unfolded protein binding, myelin sheath constituents,
and copper ion binding.
326 I. Hong et al. / Quantitative Proteomic Analysis of 5XFAD Mice
Table 1
Upregulated proteins in the 5XFAD mouse hippocampus
IPI GENE Average No of IPI GENE Average No of
SYMBOL fold change Peptides SYMBOL fold change Peptides
(TG/WT) (TG/WT)
IPI00230494 App 3.3922 IPI00124291 Vps45 1.20 2
IPI00323571 Apoe 3.2918 IPI00263013 Plp1 1.20 90
IPI00123059 Scg5 1.74 3 IPI00758024 Prdx6 1.20 32
IPI00320420 Clu 1.735 IPI00134058 Erp44 1.20 2
IPI00130758 Itm2b 1.682 IPI00316989 Epdr1 1.204
IPI00649033 Gfap 1.6323 IPI00395177 Abr 1.19 6
IPI00121514 Stip1 1.37 16 IPI00312018 Mlec 1.192
IPI00133218 Arl8b 1.362 IPI00128826 Cacng8 1.19 8
IPI00230145 Fth1 1.35 3 IPI00853902 Mllt4 1.192
IPI00404551 Ctsd 1.32 18 IPI00123342 Hyou1 1.19 4
IPI00752486 Cst3 1.28 9 IPI00135686 Ppib 1.19 7
IPI00853896 Tmsb4x 1.279 IPI01027185 Txn2 1.18 32
IPI00112129 Gatm 1.27 2 IPI00130589 Sod1 1.18 5
IPI00885558 Pdia3 1.26 6 IPI00137227 Rab2a 1.1813
IPI00944803 Prpsap2 1.26 2 IPI00113052 Tsfm 1.18 3
IPI00857249 Clstn3 1.26 5 IPI00222496 Pdia6 1.18 10
IPI00129526 Hsp90b1 1.24 32 IPI00830254 Ppp1r12b 1.18 2
IPI00130624 Pld3 1.24 12 IPI00828469 Lap3 1.18 8
IPI00118674 Ncstn 1.242 IPI00221454 Prdx6-ps1 1.17 17
IPI00989331 Wnk4 1.24 6 IPI00399449 Nsfl1c 1.1716
IPI00845833 Ppp3r1 1.23 50 IPI00113386 Ethe1 1.17 2
IPI00474959 Rcn2 1.23 8 IPI00775948 Rpl7 1.177
IPI00230108 Pdia3 1.2338 IPI00652358 Arsb 1.176
IPI00129178 Oat 1.2310 IPI00399958 Calu 1.17 4
IPI00458048 Sorbs1 1.22 3 IPI00320241 Dnajb11 1.17 2
IPI00123639 Calr 1.21 11 IPI00337975 Scn8a 1.16 2
IPI00223594 Mbp 1.21 138 IPI00125509 Fxc1 1.163
IPI00118832 Erp29 1.21 9 IPI00776047 Aifm1 1.16 5
IPI00124115 S100a13 1.20 6 IPI00223377 Mbp 1.16 139
IPI00464317 Gls 1.20 46 IPI00130640 Hrsp12 1.16 17
IPI00271951 Pdia4 1.20 4 IPI00331549 Dhrs1 1.1611
IPI00230737 Aqp4 1.203 IPI00762452 Idh1 1.16 11
IPI00329953 Stx7 1.203 IPI00170212 Cox16 1.16 2
IPI00109108 Stt3a 1.203 IPI00153107 Blmh 1.15 5
IPI00133522 P4hb 1.20 11
Average fold change was calculated by averaging three ratio values of 5XFAD mice/wild-type littermates. Asterisks indicate proteins with ratios
that were significantly different from 1 (one-sample t-test, p< 0.05).
Disease-related functional annotation of proteins
with increased and decreased levels in 5XFAD
mice
The functional annotation and upstream regula-
tor analysis were performed using Ingenuity Pathway
Analysis. The significance of the canonical pathway
and functional annotation was determined with p-
values (as an index of the confidence in the overlap)
and/or ratios (as an index of the amount of over-
lap). Functional annotation indicated that 28 of the
97 regulated proteins in 5XFAD mice had previously
been implicated in neurological disorders. The impli-
cated neurological disorders included not only AD
but also Huntington’s disease, Parkinson’s disease,
and amyotrophic lateral sclerosis, suggesting that var-
ious neurodegenerative mechanisms are activated in
5XFAD mice (Table 3). Another large subset of the
regulated proteins (14 molecules) was related to car-
diovascular disease (CVD), verifying our previous
RNA-Seq results [25] and supporting the hypothesis
that there is a strong association between CVD and
AD [39].
Upstream regulator analysis of proteins
upregulated in 5XFAD mice
To identify upstream signaling pathways that may
underlie the increased expression of proteins in
5XFAD mice, we used the upstream regulator anal-
ysis in IPA. Three transcriptional regulators (XBP1,
NFE2L2, TP53) and one transmembrane receptor
signaling pathway (IGF-1R) showed significant acti-
vation z-scores, suggesting that the activation of these
I. Hong et al. / Quantitative Proteomic Analysis of 5XFAD Mice 327
Table 2
Down-regulated proteins in the 5XFAD mouse hippocampus
IPI GENE Average No of IPI GENE Average No of
SYMBOL Fold Change Peptides SYMBOL Fold Change Peptides
(TG/WT) (TG/WT)
IPI00848690 Ly6h 0.87 8 IPI00857911 Opcml 0.8419
IPI00308976 Me3 0.87 5 IPI00874497 Pde1a 0.835
IPI00122975 Gpm6b 0.86 6 IPI00620222 Myo1c 0.83 3
IPI00857239 Atp6v0c 0.86 5 IPI00131695 Alb 0.82 3
IPI00115977 Me2 0.86 3 IPI00890092 Aldh7a1 0.82 4
IPI00311175 Tuba8 0.85407 IPI00323166 Nnt 0.81 5
IPI00407339 Hist1h4 0.85 9 IPI00128857 Me1 0.81 4
IPI00756510 Smap2 0.85 2 IPI00988950 Hbb-b1 0.81 14
IPI00108330 Cnrip1 0.855 IPI00378768 Adap1 0.80 3
IPI00880852 Vamp2 0.85 51 IPI00121443 Cox6a1 0.80 3
IPI00474073 Mapre2 0.854 IPI00463074 Shroom2 0.795
IPI00408059 Ppp2r5c 0.85 2 IPI00387416 Ubqln2 0.79 3
IPI00126551 Diras2 0.85 2 IPI00469114 Hba-a 0.78 2
IPI00463489 Opcml 0.8424 IPI00626782 Arfgef1 0.69 2
Average fold change was calculated by averaging three ratio values of 5XFAD mice/wild-type littermates. Asterisks indicate proteins with ratios
that were significantly different from 1 (one-sample t-test, p< 0.05). No of peptides shows number of independent spectra used for quantification.
Table 3
5XFAD-regulated molecules implicated in neurological disorders
Functional annotation p-value # of Molecules Molecules
Movement Disorders 2.76E-05 16 AIFM1,APOE,App,AQP4,CLU,CTSD,ETHE1,FTH1,
GFAP,MBP,PLP1,PPP3R1,PRDX6,SCN8A,SOD1,
TMSB10/TMSB4X
Neuromuscular disease 1.23E-05 15 AIFM1,APOE,AQP4,CLU,CST3,DNAJB11,FTH1,
GFAP,MBP,PDIA3,PPP3R1,PRDX6,SCN8A,SOD1,
TMSB10/TMSB4X
Progressive motor neuropathy 2.40E-07 12 AIFM1,APOE,AQP4,CLU,CST3,DNAJB11,FTH1,
GFAP,MBP,PDIA3,SCN8A,SOD1
Disorder of basal ganglia 2.91E-04 12 AIFM1,APOE,AQP4,CLU,FTH1,GFAP,MBP,PPP3R1,
PRDX6,SCN8A,SOD1,TMSB10/TMSB4X
Dementia 2.66E-05 11 AIFM1,APOE,BLMH,CLU,CST3,CTSD,GFAP,GLS,
ITM2B,NCSTN,SOD1
Tauopathy 1.78E-04 10 AIFM1,APOE,BLMH,CLU,CST3,CTSD,GFAP,
NCSTN,SCN8A,SOD1
Neurological signs 1.14E-03 10 AIFM1,APOE,App,AQP4,CLU,GFAP,PPP3R1,
PRDX6,SCN8A,TMSB10/TMSB4X
alzheimer’s disease 5.55E-04 9 AIFM1,APOE,BLMH,CLU,CST3,CTSD,GFAP,
NCSTN,SOD1
huntington’s Disease 2.57E-03 9 AIFM1,APOE,AQP4,CLU,GFAP,PPP3R1,
PRDX6,SCN8A,TMSB10/TMSB4X
Cerebrovascular dysfunction 7.75E-06 7 AIFM1,APOE,AQP4,CLU,CST3,HYOU1,ITM2B
Parkinson’s disease 4.26E-05 7 AIFM1,APOE,AQP4,FTH1,GFAP,MBP,SOD1
Gliosis 4.91E-06 6 ABR,APOE,App,ITM2B,PLP1,SOD1
Amyotrophic lateral sclerosis 3.01E-05 6 AIFM1,APOE,CLU,GFAP,SCN8A,SOD1
Inflammatory demyelinating disease 6.04E-05 6 AQP4,CST3,DNAJB11,MBP,PDIA3,PLP1
Deposition of amyloid fibrils 1.31E-06 4 APOE,App,CLU,ITM2B
Astrocytosis 1.08E-04 4 APOE,App,PLP1,SOD1
Ischemia of brain 2.51E-04 4 AQP4,CLU,CST3,HYOU1
Damage of brain 4.37E-04 4 AQP4,CST3,HYOU1,SOD1
Multiple Sclerosis 4.35E-03 4 CST3,DNAJB11,MBP,PDIA3
Ataxia 6.98E-03 4 App,CTSD,PLP1,SCN8A
Neurodegeneration 8.33E-03 4 APOE,GFAP,PLP1,SOD1
Sporadic amyotrophic lateral sclerosis 2.67E-05 3 CLU,GFAP,SOD1
Cerebral amyloid angiopathy 3.27E-05 3 APOE,CST3,ITM2B
Ischemic injury of brain 2.06E-04 3 AQP4,CST3,HYOU1
Hereditary CNS demyelinating disease 3.04E-04 3 APOE,GFAP,PLP1
Cognitive impairment 2.42E-03 3 APOE,App,CST3
Paralysis 5.23E-03 3 MBP,PLP1,SOD1
328 I. Hong et al. / Quantitative Proteomic Analysis of 5XFAD Mice
A
B
I. Hong et al. / Quantitative Proteomic Analysis of 5XFAD Mice 329
Fig. 2. Validation of APP expression in 5XFAD mice. A) Using a single peptide unique to mouse native APP (TEEISEVK), the expression
level of endogenous APP was analyzed. Note the similar expression levels in wild-type (WT) and transgenic (TG) mice. B) Total APP
expression levels (based on multiple peptides) including both native mouse APP and over-expressed human mutant APP were compared. TG
mice overexpressed APP by approximately threefold.
Table 4
Activated upstream regulator analysis of 5XFAD mice
Upstream Molecule Predicted Activation p-value Target molecules
Regulator Type Activation State z-score of overlap in dataset
XBP1 transcriptional regulator Activated 2.919 3.88E-11 CALR,DNAJB11,ERP29,ERP44,HSP90B1,HYOU1,
PDIA3,PDIA4,PDIA6,PPIB,SOD1
NFE2L2 transcriptional regulator Activated 2.938 2.50E-10 CTSD,DNAJB11,ERP29,FTH1,HSP90B1,OAT,PDIA3,
PDIA4,PDIA6,PPIB,S100A13,SOD1,STIP1
TP53 transcriptional regulator Activated 2.766 7.13E-06 APOE,App,CALU,CLU,CTSD,GATM,IDH1,OAT,
P4HB,PDIA6,PRDX6,SOD1,SORBS1,
STIP1,TMSB10/TMSB4X
IGF1R transmembrane receptor Activated 2.236 5.10E-04 AQP4,CALU,CLU,FTH1,RPL7
pathways contributes to the increased protein expres-
sion in 5XFAD mice (Table 4). Notably, NFE2L2 is
the key component of the NRF2-mediated oxidative
stress response, which is known to be neuroprotective
in AD. Further, TP53 (tumor protein p53) is known to
be upregulated in AD [40], and insulin/IGF-1 signaling
is distorted in AD [41–43].
Analysis of the 5XFAD interaction network
To find protein-protein interaction networks among
regulated proteins, we employed STRING (Search
Tool for the Retrieval of Interacting Genes/Proteins)
analysis [35] for the 87 proteins that were upreg-
ulated (69) or downregulated (28) in 5XFAD mice
Fig. 1. The CID and HCD MS/MS spectra of two representative peptides. The peaks in the CID reveal precursor peptide sequence based on
fragmentation patterns, and the HCD MS/MS spectrum detects reporter ions of m/z 126 to 131, which are used for quantification of the precursor
peptide. A) Tandem MS spectra of peptide LQAEIFQAR from ApoE. HCD shows strong increases in transgenic samples (labeled 126, 128, 130)
compared to wild-type (127, 129, 131). B) Tandem MS spectra of peptide LALDIEIATYR from GFAP, which was also increased in transgenic
samples.
330 I. Hong et al. / Quantitative Proteomic Analysis of 5XFAD Mice
Fig. 3. Validation of quantitative analysis with western blotting. A)
Western blot results for GFAP and nicastrin. B) TMT analysis and
western blotting show quantification of GFAP and nicastrin in good
agreement.
(Tables 1 and 2). In the global STRING-generated
protein-protein network that we named the 5XFAD
interaction network (FADIN, Fig. 5A), several com-
plexes and cellular functions formed prominent, tightly
connected clusters divided by cellular localization. As
shown in Fig. 5B, one network is centered on APP
and its processing machinery in the cell membrane.
Another network of mainly ER-localized proteins is
shown in Fig. 5C.
DISCUSSION
Here, we used quantitative mass spectrometry to
identify proteome-wide changes in the 5XFAD mouse
hippocampus at early stages of AD pathology and val-
idated a subset of the results using immunoblotting.
We found that the FAD mutant APP and PS1 that are
overexpressed in 5XFAD mice induce increases in the
expression of ApoE, ApoJ (clusterin), and nicastrin,
three important proteins in AD that are known to par-
ticipate in Aprocessing and clearance, as well as in
the expression of other proteins, including the neu-
rological damage/glial marker protein GFAP. A large
subset of the proteins that were strongly upregulated or
downregulated in 5XFAD brains have been implicated
in neurological disorders such as AD, Parkinson’s
disease, Huntington’s disease, and amyotrophic lat-
eral sclerosis, suggesting a substantial overlap of the
pathological mechanisms of these conditions. The
second-largest subset of regulated proteins was related
to CVD, supporting the hypothesis that there is a
strong association between CVD and AD [39]. Grow-
ing literature indicates that CVD and CVD risk factors
such as hypertension, high LDL cholesterol, low HDL
cholesterol, and diabetes are associated with increased
risk of AD, and together with our results suggest that
the pathology of both diseases might share common
molecular mechanisms.
Common upstream regulator analysis of increased
proteins suggested that the NRF2 and p53 tran-
scriptional pathways were activated, as was IGF-1
signaling. Considerable data points to a potential role
of the insulin/IGF-1 pathway in the pathogenesis of
AD [42]. In particular, AD patients show changes in
insulin/IGF-1 levels as well as insulin receptor/IGF-
1 receptor levels, and their response to insulin is
defective [43]. Our finding that targets of IGF-1R are
upregulated supports the overactivation of this path-
way in AD, which is still a matter of debate. Targets of
the XBP1 pathway were also upregulated, suggesting
the activation of the unfolded protein response [44].
Protein interactome analysis revealed an intercon-
nected network of regulated proteins (FADIN), with
two major sub-networks centered on APP process-
ing membrane complexes and mitochondrial proteins,
respectively. Together with a recent study on the
transcriptome of 5XFAD mice, our study provides a
comprehensive understanding of the molecular events
occurring in 5XFAD mice during the early stages of
AD pathology.
Our results point to strong homeostatic and neu-
roprotective responses in the brains of 5XFAD mice
after injury. ApoE and ApoJ (clusterin), which are
the main chaperones of Aand are known to coop-
eratively suppress Alevels [36, 37], had strongly
increased expression in 5XFAD brains. The increased
expression of the -secretase protein complex compo-
nent nicastrin in 5XFAD mice may have occurred in
response to the overexpressed substrate APP or to
match the overexpressed PS1, another component of
the -secretase complex. Although the other subunits
I. Hong et al. / Quantitative Proteomic Analysis of 5XFAD Mice 331
Fig. 4. DAVID Gene Ontology enrichment analysis of proteins increased and decreased in 5XFAD mice. A) Enrichment analysis by biological
process. B) Enrichment analysis by cellular component C) Enrichment analysis by molecular function.
of the -secretase protein complex were not quanti-
fied in mass spectrometry, it is possible that nicastrin,
together with Aph-1 and Pen-2, are co-regulated with
the overexpressed PS1 in a coordinated fashion to
maintain stoichiometry of the functional enzyme com-
plex [45]. The most notable cellular processes among
the proteins that were up- or downregulated in 5XFAD
mice were redox homeostasis and oxidative stress
response, and many proteins were of mitochondrial
origin, suggesting substantial oxidative stress and acti-
vation of defense mechanisms. Moreover, the activated
transcription regulator Nrf2 is the primary cellular
defense against the cytotoxic effects of oxidative stress.
Several mitochondrial molecules important in oxida-
tive stress, such as Nnt [46] and Cox6A1 [47], are
decreased in 5XFAD mice, possibly contributing to
mitochondrial dysfunction. A previous 2D-gel pro-
teomics study with the Tg2576 mouse model also
found several proteins of the respiratory chain reg-
ulated in synaptosomal/mitochondrial fractions [48],
further supporting the role of mitochondrial abnor-
malities in AD. Activated microglia and astrocytes
are abundant in the proximity of neuritic plaques, and
gliosis is known to contribute to the progression of
AD. Although plaques are not readily detected in the
hippocampus of 5XFAD mice at the age we sampled,
proteins known to be expressed by glial cells, such
as ApoE, GFAP, Plp1, and SOD1, were increased in
5XFAD samples, suggesting that gliosis may partially
precede plaque formation.
The 5XFAD mice used in this study overexpress
mutant APP, which leads to global increases in A.
Aexposure is known to block long-term potentia-
tion and exaggerate or induce long-term depression
in the brain [49–51]. Because the reversal of memory-
related long-term potentiation can account for memory
weakening [26, 52–57], we hypothesize that proteins
that were upregulated or downregulated in the present
results may contribute to the process of synapse and
memory impairment. Functional studies targeting pro-
teins that were regulated in this study but have not
been thoroughly explored previously may lead to the
discovery of novel molecular roles in AD and memory.
Because the samples analyzed for this study were
from identical genetic backgrounds except for famil-
ial AD transgenes, our results possess considerable
332 I. Hong et al. / Quantitative Proteomic Analysis of 5XFAD Mice
AB
C
Fig. 5. The 5XFAD interaction network (FADIN) shows an intricate protein-protein interaction network among regulated proteins in 5XFAD
mice (A). The color of label shows nature of regulation (red-increase; blue-decrease), and more strongly regulated proteins are presented with
larger labels. Two prominent sub-networks are evident, divided by cellular localization. B) One network is centered on APP and its processing
machinery in the cell membrane. C) Another network of mainly ER-localized proteins.
advantages over human samples in discerning the
causal pathology of AD. Nevertheless, it is still diffi-
cult to determine whether a given biochemical change
drives the disease or represents a defensive response,
as well as which individual molecular interactions are
key to pathology. Our study will provide a foundation
for further studies exploring the detailed and causal
relationships of AD pathology.
ACKNOWLEDGMENTS
This work was supported by National Research
Foundation of Korea (NRF) grants (2011-0019226 and
2011-0018209) and by a Korea Basic Science Institute
grant (T33615 to YMP, KHK, and SWC). I.H. was
supported by NRF grant # 2012R1A6A3A01019438
and I. M-J was supported by SNUH Research grant #
03-2012-0120.
Authors’ disclosures available online (http://www.j-
alz.com/disclosures/view.php?id=1723).
REFERENCES
[1] Brookmeyer R, Johnson E, Ziegler-Graham K, Arrighi HM
(2007) Forecasting the global burden of Alzheimer’s disease.
Alzheimers Dement 3, 186-191.
[2] Dubois B, Feldman HH, Jacova C, Dekosky ST, Barberger-
Gateau P, Cummings J, Delacourte A, Galasko D, Gauthier
S, Jicha G, Meguro K, O’Brien J, Pasquier F, Robert P,
Rossor M, Salloway S, Stern Y, Visser PJ, Scheltens P (2007)
Research criteria for the diagnosis of Alzheimer’s disease:
Revising the NINCDS-ADRDA criteria. Lancet Neurol 6,
734-746.
[3] Ittner LM, Gotz J (2011) Amyloid-beta and tau–a toxic pas
de deux in Alzheimer’s disease. Nat Rev Neurosci 12, 65-72.
[4] Citron M (2010) Alzheimer’s disease: Strategies for disease
modification. Nat Rev Drug Discov 9, 387-398.
[5] Selkoe DJ (2011) Resolving controversies on the path to
Alzheimer’s therapeutics. Nat Med 17, 1060-1065.
[6] Gandy S, Dekosky ST (2013) Toward the treatment and pre-
vention of Alzheimer’s disease: Rational strategies and recent
progress. Annu Rev Med 64, 367-383.
[7] Petersen RC (2003) Mild cognitive impairment clinical trials.
Nat Rev Drug Discov 2, 646-653.
[8] Blennow K, Hampel H, Weiner M, Zetterberg H (2010) Cere-
brospinal fluid and plasma biomarkers in Alzheimer disease.
Nat Rev Neurol 6, 131-144.
[9] Bateman RJ, Xiong C, Benzinger TL, Fagan AM, Goate A,
Fox NC, Marcus DS, Cairns NJ, Xie X, Blazey TM, Holtz-
man DM, Santacruz A, Buckles V, Oliver A, Moulder K,
Aisen PS, Ghetti B, Klunk WE, McDade E, Martins RN,
Masters CL, Mayeux R, Ringman JM, Rossor MN, Schofield
PR, Sperling RA, Salloway S, Morris JC, Dominantly Inher-
ited Alzheimer N (2012) Clinical and biomarker changes in
dominantly inherited Alzheimer’s disease. N Engl J Med 367,
795-804.
[10] Hardy J, Selkoe DJ (2002) The amyloid hypothesis of
Alzheimer’s disease: Progress and problems on the road to
therapeutics. Science 297, 353-356.
I. Hong et al. / Quantitative Proteomic Analysis of 5XFAD Mice 333
[11] Straub JE, Thirumalai D (2011) Toward a molecular theory of
early and late events in monomer to amyloid fibril formation.
Annu Rev Phys Chem 62, 437-463.
[12] DeStrooper B, Annaert W (2010) Novel research horizons for
presenilins and gamma-secretases in cell biology and disease.
Annu Rev Cell Dev Biol 26, 235-260.
[13] Corder EH, Saunders AM, Strittmatter WJ, Schmechel DE,
Gaskell PC, Small GW, Roses AD, Haines JL, Pericak-Vance
MA (1993) Gene dose of apolipoprotein E type 4 allele and
the risk of Alzheimer’s disease in late onset families. Science
261, 921-923.
[14] Bu G (2009) Apolipoprotein E and its receptors in
Alzheimer’sdisease: Pathways, pathogenesis and therapy. Nat
Rev Neurosci 10, 333-344.
[15] Griffin WS, Sheng JG, Roberts GW, Mrak RE (1995)
Interleukin-1 expression in different plaque types in
Alzheimer’s disease: Significance in plaque evolution. J Neu-
ropathol Exp Neurol 54, 276-281.
[16] Sheng JG, Mrak RE, Griffin WS (1997) Glial-neuronal inter-
actions in Alzheimer disease: Progressive association of
IL-1alpha+ microglia and S100beta+astrocytes with neurofib-
rillary tangle stages. J Neuropathol Exp Neurol 56, 285-290.
[17] Bantscheff M, Schirle M, Sweetman G, Rick J, Kuster B
(2007) Quantitative mass spectrometry in proteomics: A crit-
ical review. Anal Bioanal Chem 389, 1017-1031.
[18] Domon B, Aebersold R (2010) Options and considera-
tions when selecting a quantitative proteomics strategy. Nat
Biotechnol 28, 710-721.
[19] Cox J, Mann M (2011) Quantitative, high-resolution pro-
teomics for data-driven systems biology. Annu Rev Biochem
80, 273-299.
[20] Zhang J, Goodlett DR, Quinn JF, Peskind E, Kaye JA, Zhou
Y, Pan C, Yi E, Eng J, Wang Q, Aebersold RH, Montine TJ
(2005) Quantitative proteomics of cerebrospinal fluid from
patients with Alzheimer disease. J Alzheimers Dis 7, 125-133.
discussion 173-180
[21] Schonberger SJ, Edgar PF, Kydd R, Faull RL, Cooper GJ
(2001) Proteomic analysis of the brain in Alzheimer’s dis-
ease: Molecular phenotype of a complex disease process.
Proteomics 1, 1519-1528.
[22] Oakley H, Cole SL, Logan S, Maus E, Shao P, Craft J,
Guillozet-Bongaarts A, Ohno M, Disterhoft J, Van Eldik L,
Berry R, Vassar R (2006) Intraneuronal beta-amyloid aggre-
gates, neurodegeneration, and neuron loss in transgenic mice
with five familial Alzheimer’s disease mutations: Potential
factors in amyloid plaque formation. J Neurosci 26, 10129-
10140.
[23] Moon M, Hong HS, Nam DW, Baik SH, Song H, Kook
SY, Kim YS, Lee J, Mook-Jung I (2012) Intracellular
amyloid-beta accumulation in calcium-binding protein-
deficient neurons leads to amyloid-beta plaque formation in
animal model of Alzheimer’s disease. J Alzheimers Dis 29,
615-628.
[24] Kimura R, Ohno M (2009) Impairments in remote memory
stabilization precede hippocampal synaptic and cognitive fail-
ures in 5XFAD Alzheimer mouse model. Neurobiol Dis 33,
229-235.
[25] KimKH, Moon M, Yu SB, Mook-Jung I, Kim JI (2012) RNA-
Seq analysis of frontal cortex and cerebellum from 5XFAD
mice at early stage of disease pathology. J Alzheimers Dis 29,
793-808.
[26] Kim J, Lee S, Park K, Hong I, Song B, Son G, Park H,
Kim WR, Park E, Choe HK, Kim H, Lee C, Sun W, Kim
K, Shin KS, Choi S (2007) Amygdala depotentiation and fear
extinction. Proc Natl Acad SciUSA104, 20955-20960.
[27] Xu T, Venable JD, Park SK, Cociorva D, Lu B, Liao L,
Wohlschlegel J, Hewel J, Yates JR (2006) ProLuCID, a fast
and sensitive tandem mass spectra-based protein identifica-
tion program. Mol Cell Proteomics 5, S174-S174.
[28] Washburn MP, Wolters D, Yates JR, 3rd (2001) Large-scale
analysis of the yeast proteome by multidimensional protein
identification technology. Nat Biotechnol 19, 242-247.
[29] Byun K, Young Kim J, Bayarsaikhan E, Kim D, Jeong GB,
Yun KN, Kyeong Min H, Kim SU, Yoo JS, Lee B (2012)
Quantitative proteomic analysis reveals that lipopolysac-
charide induces mitogen-activated protein kinase-dependent
activation in human microglial cells. Electophoresis 33,
3756-3763.
[30] Li Z, Adams RM, Chourey K, Hurst GB, Hettich RL, Pan
C (2012) Systematic comparison of label-free, metabolic
labeling, and isobaric chemical labeling for quantitative
proteomics on LTQ Orbitrap Velos. J Proteome Res 11, 1582-
1590.
[31] Cociorva D, D LT, Yates JR (2007) Validation of tandem mass
spectrometry database search results using DTASelect. Curr
Protoc Bioinformatics Chapter 13, Unit 13 14.
[32] ParkSK, Venable JD, Xu T, Liao L, YatesJR (2006) A tool for
quantitative analysis of high-throughput mass spectrometry
data. Mol Cell Proteomics 5, S199-S199.
[33] Raso C, Cosentino C, Gaspari M, Malara N, Han X,
McClatchy D, Park SK, Renne M, Vadala N, Prati U, Cuda G,
Mollace V, Amato F, Yates JR, 3rd (2012) Characterization of
breast cancer interstitial fluids by TmT labeling, LTQ-orbitrap
velos mass spectrometry, and pathway analysis. J Proteome
Res 11, 3199-3210.
[34] Huang DW, Sherman BT, Lempicki RA (2009) Systematic
and integrative analysis of largegene lists using DAVIDbioin-
formatics resources. Nat Protoc 4, 44-57.
[35] Szklarczyk D, Franceschini A, Kuhn M, Simonovic M, Roth
A, Minguez P, Doerks T, Stark M, Muller J, Bork P, Jensen LJ,
von Mering C (2011) The STRING database in 2011: Func-
tional interaction networks of proteins, globally integrated and
scored. Nucleic Acids Res 39, D561-D568.
[36] DeMattos RB, Cirrito JR, Parsadanian M, May PC, O’Dell
MA, Taylor JW, Harmony JA, Aronow BJ, Bales KR, Paul
SM, Holtzman DM (2004) ApoE and clusterin cooperatively
suppress Abeta levels and deposition: Evidence that ApoE
regulates extracellular Abeta metabolism in vivo.Neuron 41,
193-202.
[37] JiangQ, Lee CY, Mandrekar S, WilkinsonB, Cramer P, Zelcer
N, Mann K, Lamb B, Willson TM, Collins JL, Richardson JC,
Smith JD, Comery TA, Riddell D, Holtzman DM, Tontonoz
P, Landreth GE (2008) ApoE promotes the proteolytic degra-
dation of Abeta. Neuron 58, 681-693.
[38] Yu G, Nishimura M, Arawaka S, Levitan D, Zhang L, Tan-
don A, Song YQ, Rogaeva E, Chen F, Kawarai T, Supala
A, Levesque L, Yu H, Yang DS, Holmes E, Milman P,
Liang Y, Zhang DM, Xu DH, Sato C, Rogaev E, Smith M,
Janus C, Zhang Y, Aebersold R, Farrer LS, Sorbi S, Bruni
A, Fraser P, St George-Hyslop P (2000) Nicastrin modu-
lates presenilin-mediated notch/glp-1 signal transduction and
betaAPP processing. Nature 407, 48-54.
[39] Stampfer MJ (2006) Cardiovascular disease and Alzheimer’s
disease: Common links. J Intern Med 260, 211-223.
[40] Hooper C, Meimaridou E, Tavassoli M, Melino G, Lovestone
S, Killick R (2007) p53 is upregulated in Alzheimer’s disease
and induces tau phosphorylation in HEK293a cells. Neurosci
Lett 418, 34-37.
[41] Vardy ER, Rice PJ, Bowie PC, Holmes JD, Grant PJ, Hooper
NM (2007) Increased circulating insulin-like growth factor-1
334 I. Hong et al. / Quantitative Proteomic Analysis of 5XFAD Mice
in late-onset Alzheimer’s disease. J Alzheimers Dis 12, 285-
290.
[42] Freude S, Schilbach K, Schubert M (2009) The role of IGF-1
receptor and insulin receptor signaling for the pathogene-
sis of Alzheimer’s disease: From model organisms to human
disease. Curr Alzheimer Res 6, 213-223.
[43] Moloney AM, Griffin RJ, Timmons S, O’Connor R, Ravid R,
O’Neill C (2010) Defects in IGF-1 receptor, insulin receptor
and IRS-1/2 in Alzheimer’s disease indicate possible resis-
tance to IGF-1 and insulin signalling. Neurobiol Aging 31,
224-243.
[44] Iwakoshi NN, Lee AH, Vallabhajosyula P, Otipoby KL,
Rajewsky K, Glimcher LH (2003) Plasma cell differentiation
and the unfolded protein response intersect at the transcription
factor XBP-1. Nat Immunol 4, 321-329.
[45] Sato T, Diehl TS, Narayanan S, Funamoto S, Ihara Y, De
Strooper B, Steiner H, Haass C, Wolfe MS (2007) Active
gamma-secretase complexes contain only one of each com-
ponent. J Biol Chem 282, 33985-33993.
[46] Huang TT, Naeemuddin M, Elchuri S, Yamaguchi M, Kozy
HM, Carlson EJ, Epstein CJ (2006) Genetic modifiers of the
phenotype of mice deficient in mitochondrial superoxide dis-
mutase. Hum Mol Genet 15, 1187-1194.
[47] Grant CM, MacIver FH, Dawes IW (1997) Mitochondrial
function is required for resistance to oxidative stress in the
yeast Saccharomyces cerevisiae. FEBS Lett 410, 219-222.
[48] Gillardon F, Rist W, Kussmaul L, Vogel J, Berg M, Danzer K,
Kraut N, Hengerer B (2007) Proteomic and functional alter-
ations in brain mitochondria from Tg2576 mice occur before
amyloid plaque deposition. Proteomics 7, 605-616.
[49] Walsh DM, Klyubin I, Fadeeva JV, Cullen WK, Anwyl R,
Wolfe MS, Rowan MJ, Selkoe DJ (2002) Naturally secreted
oligomers of amyloid beta protein potently inhibit hippocam-
pal long-term potentiation in vivo.Nature 416, 535-539.
[50] Hsieh H, Boehm J, Sato C, Iwatsubo T, Tomita T, Sisodia S,
Malinow R (2006) AMPAR removal underlies Abeta-induced
synaptic depression and dendritic spine loss. Neuron 52, 831-
843.
[51] Haass C, Selkoe DJ (2007) Soluble protein oligomers in
neurodegeneration: Lessons from the Alzheimer’s amyloid
beta-peptide. Nat Rev Mol Cell Biol 8, 101-112.
[52] Hong I, Kim J, Lee J, Park S, Song B, Kim J, An B, Park K,
Lee HW, Lee S, Kim H, Park SH, Eom KD, Lee S, Choi S
(2011) Reversible plasticity of fear memory-encoding amyg-
dala synaptic circuits even after fear memory consolidation.
PLoS One 6, e24260.
[53] HongI, Kim J, Song B, Park K, Shin K, Eom KD, Han PL, Lee
S, Choi S (2012) Fear conditioning occludes late-phase long-
term potentiation at thalamic input synapses onto the lateral
amygdala in rat brain slices. Neurosci Lett 506, 121-125.
[54] Hong I, Song B, Lee S, Kim J, Choi S (2009) Extinction of
cued fear memory involves a distinct form of depotentiation
at cortical input synapses onto the lateral amygdala. Eur J
Neurosci 30, 2089-2099.
[55] Pascoli V, Turiault M, Luscher C (2012) Reversal of cocaine-
evoked synaptic potentiation resets drug-induced adaptive
behaviour. Nature 481, 71-75.
[56] Hong I, Kim J, Song B, Park S, Lee J, Kim J, An B, Lee S,
Choi S (2011) Modulation of fear memory by retrieval and
extinction: A clue for memory deconsolidation. Rev Neurosci
22, 205-229.
[57] Kim J, Song B, Hong I, Lee J, Park S, Eom JY, Lee CJ,
Lee S, Choi S (2010) Reactivation of fear memory renders
consolidated amygdala synapses labile. J Neurosci 30, 9631-
9640.
... These mice express both human APP and PS1 with five FAD mutations in neurons, leading to overproduction of Aβ42 peptides, which progressively accumulate in intraneuronal deposits and extracellular amyloid plaques [5] in mouse brains. Strikingly, Aβ42 aggregation in 5xFAD mouse brains is accompanied by activated neuroinflammation, loss of synapse functions and neurodegeneration, which are all well-known pathobiological features of AD patient brains [20,[22][23][24]. In a recent proteomics study, convincing data were obtained that proteome changes in brains of 5xFAD tg mice and AD patients are cross-correlated [25], supporting the hypothesis that this disease model mimics key aspects of AD pathogenesis. ...
Article
Full-text available
Background Alzheimer’s disease (AD) is characterized by the intra- and extracellular accumulation of amyloid-β (Aβ) peptides. How Aβ aggregates perturb the proteome in brains of patients and AD transgenic mouse models, remains largely unclear. State-of-the-art mass spectrometry (MS) methods can comprehensively detect proteomic alterations, providing relevant insights unobtainable with transcriptomics investigations. Analyses of the relationship between progressive Aβ aggregation and protein abundance changes in brains of 5xFAD transgenic mice have not been reported previously. Methods We quantified progressive Aβ aggregation in hippocampus and cortex of 5xFAD mice and controls with immunohistochemistry and membrane filter assays. Protein changes in different mouse tissues were analyzed by MS-based proteomics using label-free quantification; resulting MS data were processed using an established pipeline. Results were contrasted with existing proteomic data sets from postmortem AD patient brains. Finally, abundance changes in the candidate marker Arl8b were validated in cerebrospinal fluid (CSF) from AD patients and controls using ELISAs. Results Experiments revealed faster accumulation of Aβ42 peptides in hippocampus than in cortex of 5xFAD mice, with more protein abundance changes in hippocampus, indicating that Aβ42 aggregate deposition is associated with brain region-specific proteome perturbations. Generating time-resolved data sets, we defined Aβ aggregate-correlated and anticorrelated proteome changes, a fraction of which was conserved in postmortem AD patient brain tissue, suggesting that proteome changes in 5xFAD mice mimic disease-relevant changes in human AD. We detected a positive correlation between Aβ42 aggregate deposition in the hippocampus of 5xFAD mice and the abundance of the lysosome-associated small GTPase Arl8b, which accumulated together with axonal lysosomal membranes in close proximity of extracellular Aβ plaques in 5xFAD brains. Abnormal aggregation of Arl8b was observed in human AD brain tissue. Arl8b protein levels were significantly increased in CSF of AD patients. Conclusions We report a comprehensive biochemical and proteomic investigation of hippocampal and cortical brain tissue derived from 5xFAD transgenic mice, providing a valuable resource to the neuroscientific community. We identified Arl8b, with significant abundance changes in 5xFAD and AD patient brains. Arl8b might enable the measurement of progressive lysosome accumulation in AD patients and have clinical utility as a candidate biomarker.
... Additionally, the increased expression of CB2 and GPR55 receptors positively correlated with a rise in neuroinflammatory markers [27]. However, although the 5xFAD animals display strong neuroinflammation (astrogliosis and microgliosis) in response to amyloid deposition [26,28], these mice are overexpressing the amyloid precursor protein (APP). This results in an overproduction of other APP fragments in addition to Aβ peptides [29]. ...
Article
Full-text available
Background: The endocannabinoid system (ECS) and associated lipid transmitter-based signaling systems play an important role in modulating brain neuroinflammation. ECS is affected in neurodegenerative disorders, such as Alzheimer's disease (AD). Here we have evaluated the non-psychotropic endocannabinoid receptor type 2 (CB2) and lysophosphatidylinositol G-protein-coupled receptor 55 (GPR55) localization and expression during Aβ-pathology progression. Methods: Hippocampal gene expression of CB2 and GPR55 was explored by qPCR analysis, and brain distribution was evaluated by immunofluorescence in the wild type (WT) and APP knock-in AppNL-G-F AD mouse model. Furthermore, the effects of Aβ42 on CB2 and GPR55 expression were assessed in primary cell cultures. Results: CB2 and GPR55 mRNA levels were significantly upregulated in AppNL-G-F mice at 6 and 12 months of age, compared to WT. CB2 was highly expressed in the microglia and astrocytes surrounding the Aβ plaques. Differently, GPR55 staining was mainly detected in neurons and microglia but not in astrocytes. In vitro, Aβ42 treatment enhanced CB2 receptor expression mainly in astrocytes and microglia cells, whereas GPR55 expression was enhanced primarily in neurons. Conclusions: These data show that Aβ pathology progression, particularly Aβ42, plays a crucial role in increasing the expression of CB2 and GPR55 receptors, supporting CB2 and GPR55 implications in AD.
... Oxidative stress is associated with damaged Aβ peptide and contributes to AD progression [38]. Furthermore, oxidative stress was shown to be one of the biological processes associated with differentially expressed proteins in 5XFAD and WT mice at the age of 4 months [39]. We previously reported that a higher redox ratio correlates with higher reactive oxygen species (ROS), generated by oxidative insults [25,26,28], and higher oxidative stress in inflamed rhesus macaques brains [29]. ...
Article
Full-text available
A substantial decline in nicotinamide adenine dinucleotide (NAD) has been reported in brain tissue homogenates or neurons isolated from Alzheimer’s disease (AD) models. NAD, together with flavin adenine dinucleotide (FAD), critically supports energy metabolism and maintains mitochondrial redox homeostasis. Optical redox imaging (ORI) of the intrinsic fluorescence of reduced NAD (NADH) and oxidized FAD yields cellular redox and metabolic information and provides biomarkers for a variety of pathological conditions. However, its utility in AD has not been characterized at the tissue level. We performed ex vivo ORI of freshly dissected hippocampi from a well-characterized AD mouse model with five familial Alzheimer’s disease mutations (5XFAD) and wild type (WT) control littermates at various ages. We found (1) a significant increase in the redox ratio with age in the hippocampi of both the WT control and the 5XFAD model, with a more prominent redox shift in the AD hippocampi; (2) a higher NADH in the 5XFAD versus WT hippocampi at the pre-symptomatic age of 2 months; and (3) a negative correlation between NADH and Aβ42 level, a positive correlation between Fp and Aβ42 level, and a positive correlation between redox ratio and Aβ42 level in the AD hippocampi. These findings suggest that the ORI can be further optimized to conveniently study the metabolism of freshly dissected brain tissues in animal models and identify early AD biomarkers.
... Thus, here, we have explored the effects induced by a chronic low dose treatment of a combination of the sEHi (TPPU), with either the reversible AChEi (6-Cl-THA) or a pseudoirreversible AChEi (Riv), in two different mouse models of AD, SAMP8 and 5XFAD mice. As previously mentioned, SAMP8 is a model of LOAD, which is suitable for studying age-related cognitive impairment associated with a decline of the cholinergic system [25,26] and neuroinflammation [27], whereas 5XFAD is a widely used model of early-onset familial AD (EOFAD) [28]. ...
Article
Full-text available
Alzheimer’s disease (AD) is a progressive neurological disorder with multifactorial and heterogeneous causes. AD involves several etiopathogenic mechanisms such as aberrant protein accumulation, neurotransmitter deficits, synaptic dysfunction and neuroinflammation, which lead to cognitive decline. Unfortunately, the currently available anti-AD drugs only alleviate the symptoms temporarily and provide a limited therapeutic effect. Thus, new therapeutic strategies, including multitarget approaches, are urgently needed. It has been demonstrated that a co-treatment of acetylcholinesterase (AChE) inhibitor with other neuroprotective agents has beneficial effects on cognition. Here, we have assessed the neuroprotective effects of chronic dual treatment with a soluble epoxide hydrolase (sEH) inhibitor (TPPU) and an AChE inhibitor (6-chlorotacrine or rivastigmine) in in vivo studies. Interestingly, we have found beneficial effects after chronic low-dose co-treatment with TPPU and 6-chlorotacrine in the senescence-accelerated mouse prone 8 (SAMP8) mouse model as well as with TPPU and rivastigmine co-treatment in the 5XFAD mouse model, in comparison with the corresponding monotherapy treatments. In the SAMP8 model, no substantial improvements in synaptic plasticity markers were found, but the co-treatment of TPPU and 6-chlorotacrine led to a significantly reduced gene expression of neuroinflammatory markers, such as interleukin 6 (Il-6), triggering receptor expressed on myeloid cell 2 (Trem2) and glial fibrillary acidic protein (Gfap). In 5XFAD mice, chronic low-dose co-treatment of TPPU and rivastigmine led to enhanced protein levels of synaptic plasticity markers, such as the phospho-cAMP response element-binding protein (p-CREB) ratio, brain-derived neurotrophic factor (BDNF) and postsynaptic density protein 95 (PSD95), and also to a reduction in neuroinflammatory gene expression. Collectively, these results support the neuroprotectant role of chronic low-dose co-treatment strategy with sEH and AChE inhibitors in AD mouse models, opening new avenues for effective AD treatment.
... This large panel includes proteins for typical phenotype markers of major brain cell types, AD hallmark pathways, neuron functions, enzymes, canonical signaling pathways (e.g., insulin signaling, calcium signaling, JAK/STAT signaling, autophagy signaling and apoptosis), and transcriptional factors. They were selected from Mouse Brain of Allen Brain Atlas, AlzPathway and publications related with 5xFAD mouse model (Supplementary Data 2) [40][41][42] . All the antibodies targeting the relevant proteins were conjugated with their corresponding complementary oligos, and the whole set was divided into 4 panels for 4 staining rounds of CycMIST experiments (Supplementary Data 1). ...
Article
Full-text available
Despite the fact that proteins carry out nearly all cellular functions and mark the differences of cells, the existing single-cell tools can only analyze dozens of proteins, a scale far from full characterization of cells and tissue yet. Herein, we present a single-cell cyclic multiplex in situ tagging (CycMIST) technology that affords the comprehensive functional proteome profiling of single cells. We demonstrate the technology by detecting 182 proteins that include surface markers, neuron function proteins, neurodegeneration markers, signaling pathway proteins, and transcription factors. Further studies on cells derived from the 5XFAD mice, an Alzheimer’s Disease (AD) model, validate the utility of our technology and reveal the deep heterogeneity of brain cells. Through comparison with control mouse cells, we have identified differentially expressed proteins in AD pathology. Our technology could offer new insights into cell machinery and thus may advance many fields including drug discovery, molecular diagnostics, and clinical studies. Current single-cell tools are limited by the number of proteins they can analyse. Here the authors report a single-cell cyclic multiplex in situ tagging (CycMIST) method for functional proteome profiling of single cells, allowing multiple rounds of multiplexing of the same single cells on a microchip.
... Our mapping of the regional 5xFAD proteome revealed few changes in the early phases of amyloid pathology. However, several of our findings are in accordance with previous reports, including elevated GFAP, APP, Clusterin (Clu), and Cathepsin-D (Ctsd) expression in the 5xFAD hippocampus [35,36]. Our finding of elevated hippocampal expression of proteins related to the complement system and lysosomal function has also been observed in human AD brain samples [37,38]. ...
Article
Full-text available
Alzheimer’s disease (AD) is an unremitting neurodegenerative disorder characterized by cerebral amyloid-β (Aβ) accumulation and gradual decline in cognitive function. Changes in brain energy metabolism arise in the preclinical phase of AD, suggesting an important metabolic component of early AD pathology. Neurons and astrocytes function in close metabolic collaboration, which is essential for the recycling of neurotransmitters in the synapse. However, this crucial metabolic interplay during the early stages of AD development has not been sufficiently investigated. Here, we provide an integrative analysis of cellular metabolism during the early stages of Aβ accumulation in the cerebral cortex and hippocampus of the 5xFAD mouse model of AD. Our electrophysiological examination revealed an increase in spontaneous excitatory signaling in the 5xFAD hippocampus. This hyperactive neuronal phenotype coincided with decreased hippocampal tricarboxylic acid (TCA) cycle metabolism mapped by stable ¹³ C isotope tracing. Particularly, reduced astrocyte TCA cycle activity and decreased glutamine synthesis led to hampered neuronal GABA synthesis in the 5xFAD hippocampus. In contrast, the cerebral cortex of 5xFAD mice displayed an elevated capacity for oxidative glucose metabolism, which may suggest a metabolic compensation in this brain region. We found limited changes when we explored the brain proteome and metabolome of the 5xFAD mice, supporting that the functional metabolic disturbances between neurons and astrocytes are early primary events in AD pathology. In addition, synaptic mitochondrial and glycolytic function was selectively impaired in the 5xFAD hippocampus, whereas non-synaptic mitochondrial function was maintained. These findings were supported by ultrastructural analyses demonstrating disruptions in mitochondrial morphology, particularly in the 5xFAD hippocampus. Collectively, our study reveals complex regional and cell-specific metabolic adaptations in the early stages of amyloid pathology, which may be fundamental for the progressing synaptic dysfunctions in AD.
... Poznato je da glioza doprinosi napredovanju Alchajmerove bolesti (8,9). Hong i saradnici su uz pomoć kvantitativne proteomske analize, rađene u hipokampusu transgenih 5xFAD miševa starosti 4 meseca i u odgovarajućoj kontrolnoj grupi pokazali porast proteina koji se eksprimiraju u glija ćelijama i učestvuju u patogenezi Alchajmerove bolesti (16). Iako se amiloidni plakovi ne otkrivaju lako u hipokampusu miševa ovog uzrasta, među ostalim proteinima nađena je i povećana ekspresija GFAP-a, što sugeriše da glioza može delimično i da prethodi formiranju plaka. ...
Article
Full-text available
Introduction: Alzheimer's disease is the most common neurodegenerative disorder, characterized by the formation of amyloid plaques and the neurofibrillary tangles in the brain of an ill person, leading to neuronal damage and loss. Activation of astrocytes and astrogliosis occurs along with this process. Due to ethical limitations in working with human tissue, numerous transgenic animal models have been developed to study the pathogenesis of these processes. Early Ab deposition is observed in the cortex and the hippocampus. Aim: This study aimed to determine the difference in the presence of GFAP positive cells in the hippocampus between transgenic 5xFAD mice aged 36 weeks and their corresponding controls. Material and Methods: The 5xFAD mice model of Alzheimer's disease was used, characterized by early formation of amyloid plaques but without the presence of neurofibrillar tangles. Transgenic and control animals were sacrificed at 36 weeks of age. The visualization of GFAP-positive cells in the hippocampus of their brains was done by using immunohistochemistry and antibody for glial fibrillary acidic protein - GFAP, the major marker of astrocytes. Quantification of immuno-reactivity was done by using the Icy software system. Results: There was a statistically significant difference in the expression of GFAP in the dentate gyrus and the granular zone of the hippocampus between the transgenic and control group at 36 weeks of age, while the significant change in the CA1-3 regions was not observed between investigated groups. Conclusion: Obtained results confirm the involvement of astrogliosis in the pathophysiology of Alzheimer's disease and indicate an earlier occurrence of astrogliosis in the dentate gyrus and granular zone, in relation to other regions of the hippocampus, in the 36-week-old 5xFAD mice.
Article
Given continued failure of BACE1 inhibitor programs at symptomatic and prodromal stages of Alzheimer’s disease (AD), clinical trials need to target the earlier preclinical stage. However, trial design is complex in this population with negative diagnosis of classical hippocampal amnesia on standard memory tests. Besides recent advances in brain imaging, electroencephalogram, and fluid-based biomarkers, new cognitive markers should be established for earlier diagnosis that can optimize recruitment to BACE1 inhibitor trials in presymptomatic AD. Notably, accelerated long-term forgetting (ALF) is emerging as a sensitive cognitive measure that can discriminate between asymptomatic individuals with high risks for developing AD and healthy controls. ALF is a form of declarative memory impairment characterized by increased forgetting rates over longer delays (days to months) despite normal storage within the standard delays of testing (20–60 min). Therefore, ALF may represent a harbinger of preclinical dementia and the impairment of systems memory consolidation, during which memory traces temporarily stored in the hippocampus become gradually integrated into cortical networks. This review provides an overview of the utility of ALF in a rational design of next-generation BACE1 inhibitor trials in preclinical AD. I explore potential mechanisms underlying ALF and relevant early-stage biomarkers useful for BACE1 inhibitor evaluation, including synaptic protein alterations, astrocytic dysregulation and neuron hyperactivity in the hippocampal-cortical network. Furthermore, given the physiological role of the isoform BACE2 as an AD-suppressor gene, I also discuss the possible association between the poor selectivity of BACE1 inhibitors and their side effects (e.g., cognitive worsening) in prior clinical trials.
Article
Full-text available
Human studies consistently identify bioenergetic maladaptations in brains upon aging and neurodegenerative disorders of aging (NDAs), such as Alzheimer's disease, Parkinson's disease, Huntington's disease, and Amyotrophic lateral sclerosis. Glucose is the major brain fuel and glucose hypometabolism has been observed in brain regions vulnerable to aging and NDAs. Many neurodegenerative susceptible regions are in the topological central hub of the brain connectome, linked by densely interconnected long-range axons. Axons, key components of the connectome, have high metabolic needs to support neurotransmission and other essential activities. Long-range axons are particularly vulnerable to injury, neurotoxin exposure, protein stress, lysosomal dysfunction, etc. Axonopathy is often an early sign of neurodegeneration. Recent studies ascribe axonal maintenance failures to local bioenergetic dysregulation. With this review, we aim to stimulate research in exploring metabolically oriented neuroprotection strategies to enhance or normalize bioenergetics in NDA models. Here we start by summarizing evidence from human patients and animal models to reveal the correlation between glucose hypometabolism and connectomic disintegration upon aging/NDAs. To encourage mechanistic investigations on how axonal bioenergetic dysregulation occurs during aging/NDAs, we first review the current literature on axonal bioenergetics in distinct axonal subdomains: axon initial segments, myelinated axonal segments, and axonal arbors harboring pre-synaptic boutons. In each subdomain, we focus on the organization, activity-dependent regulation of the bioenergetic system, and external glial support. Second, we review the mechanisms regulating axonal nicotinamide adenine dinucleotide (NAD+) homeostasis, an essential molecule for energy metabolism processes, including NAD+ biosynthetic, recycling, and consuming pathways. Third, we highlight the innate metabolic vulnerability of the brain connectome and discuss its perturbation during aging and NDAs. As axonal bioenergetic deficits are developing into NDAs, especially in asymptomatic phase, they are likely exaggerated further by impaired NAD+ homeostasis, the high energetic cost of neural network hyperactivity, and glial pathology. Future research in interrogating the causal relationship between metabolic vulnerability, axonopathy, amyloid/tau pathology, and cognitive decline will provide fundamental knowledge for developing therapeutic interventions.
Article
Full-text available
DAVID bioinformatics resources consists of an integrated biological knowledgebase and analytic tools aimed at systematically extracting biological meaning from large gene/protein lists. This protocol explains how to use DAVID, a high-throughput and integrated data-mining environment, to analyze gene lists derived from high-throughput genomic experiments. The procedure first requires uploading a gene list containing any number of common gene identifiers followed by analysis using one or more text and pathway-mining tools such as gene functional classification, functional annotation chart or clustering and functional annotation table. By following this protocol, investigators are able to gain an in-depth understanding of the biological themes in lists of genes that are enriched in genome-scale studies.
Data
Full-text available
Insulin like growth factor-1 receptor (IGF-1R) and insulin receptor (IR) signalling control vital growth, survival and metabolic functions in the brain. Here we describe specific and significant alterations in IGF-1R, IR, and their key substrate adaptor proteins IRS-1 and IRS-2 in Alzheimer's disease (AD). Western immunoblot analysis detected increased IGF-1R levels, and decreased levels of IGF-1-binding protein-2 (IGFBP-2), a major IGF-1-binding protein, in AD temporal cortex. Increased IGF-1R was observed surrounding and within amyloid-(A)-containing plaques, also evident in an animal model of AD, and in astrocytes in AD. However, despite the overall increase in IGF-1R levels, a significantly lower number of neurons expressed IGF-1R in AD, and IGF-1R was aberrantly distributed in AD neurons especially evident in those with neurofibrillary tangles (NFTs). IR protein levels were similar in AD and control cases, however, the IR was concentrated intracellularly in AD neurons, unlike its distribution throughout the neuronal cell soma and in dendrites in control brain. Significant decreases in IRS-1 and IRS-2 levels were identified in AD neurons, in association with increased levels of inactivated phospho Ser312 IRS-1 and phospho Ser616 IRS-1, where increased levels of these phosphoserine epitopes colocalised strongly with NFTs. Our results show that IGF-1R and IR signalling is compromised in AD neurons and suggest that neurons that degenerate in AD may be resistant to IGF-1R/IR signalling.
Article
Full-text available
Background: Our goal was to forecast the global burden of Alzheimer’s disease and evaluate the potential impact of interventions that delay disease onset or progression. Methods: A stochastic, multistate model was used in conjunction with United Nations worldwide population forecasts and data from epidemiological studies of the risks of Alzheimer’s disease. Results: In 2006, the worldwide prevalence of Alzheimer’s disease was 26.6 million. By 2050, the prevalence will quadruple, by which time 1 in 85 persons worldwide will be living with the disease. We estimate about 43% of prevalent cases need a high level of care, equivalent to that of a nursing home. If interventions could delay both disease onset and progression by a modest 1 year, there would be nearly 9.2 million fewer cases of the disease in 2050, with nearly the entire decline attributable to decreases in persons needing a high level of care. Conclusions: We face a looming global epidemic of Alzheimer’s disease as the world’s population ages. Modest advances in therapeutic and preventive strategies that lead to even small delays in the onset and progression of Alzheimer’s disease can significantly reduce the global burden of this disease.
Presentation
01-02-02 The goal was to forecast the global burden of Alzheimer’s disease and evaluate the potential impact of interventions that delay disease onset or progression.
Article
The apolipoprotein E type 4 allele (APOE-epsilon 4) is genetically associated with the common late onset familial and sporadic forms of Alzheimer's disease (AD). Risk for AD increased from 20% to 90% and mean age at onset decreased from 84 to 68 years with increasing number of APOE-epsilon 4 alleles in 42 families with late onset AD. Thus APOE-epsilon 4 gene dose is a major risk factor for late onset AD and, in these families, homozygosity for APOE-epsilon 4 was virtually sufficient to cause AD by age 80.
Article
Apolipoprotein E (apoE) and clusterin can influence structure, toxicity, and accumulation of the amyloid-β (Aβ) peptide in brain. Both molecules may also be involved in Aβ metabolism prior to its deposition. To assess this possibility, we compared PDAPP transgenic mice that develop age-dependent Aβ accumulation in the absence of apoE or clusterin as well as in the absence of both proteins. apoE−/− and clusterin−/− mice accumulated similar Aβ levels but much less fibrillar Aβ. In contrast, apoE−/−/clusterin−/− mice had both earlier onset and markedly increased Aβ and amyloid deposition. Both apoE−/− and apoE−/−/clusterin−/− mice had elevated CSF and brain interstitial fluid Aβ, as well as significant differences in the elimination half-life of interstitial fluid Aβ measured by in vivo microdialysis. These findings demonstrate additive effects of apoE and clusterin on influencing Aβ deposition and that apoE plays an important role in regulating extracellular CNS Aβ metabolism independent of Aβ synthesis.