ArticlePDF AvailableLiterature Review

BACE1 inhibitors: Current status and future directions in treating Alzheimer's disease

Authors:

Abstract

Alzheimer's disease (AD) is an irreversible, progressive neurodegenerative brain disorder with no current cure. One of the important therapeutic approaches of AD is the inhibition of β‐site APP cleaving enzyme‐1 (BACE1), which is involved in the rate‐limiting step of the cleavage process of the amyloid precursor protein (APP) leading to the generation of the neurotoxic amyloid β (Aβ) protein after the γ‐secretase completes its function. The produced insoluble Aβ aggregates lead to plaques deposition and neurodegeneration. BACE1 is, therefore, one of the attractive targets for the treatment of AD. This approach led to the development of potent BACE1 inhibitors, many of which were advanced to late stages in clinical trials. Nonetheless, the high failure rate of lead drug candidates targeting BACE1 brought to the forefront the need for finding new targets to uncover the mystery behind AD. In this review, we aim to discuss the most promising classes of BACE1 inhibitors with a description and analysis of their pharmacodynamic and pharmacokinetic parameters, with more focus on the lead drug candidates that reached late stages of clinical trials, such as MK8931, AZD‐3293, JNJ‐54861911, E2609, and CNP520. In addition, the manuscript discusses the safety concerns and insignificant physiological effects, which were highlighted for the most successful BACE1 inhibitors. Furthermore, the review demonstrates with increasing evidence that despite tremendous efforts and promising results conceived with BACE1 inhibitors, the latest studies suggest that their clinical use for treating Alzheimer's disease should be reconsidered. Finally, the review sheds light on alternative therapeutic options for targeting AD.
Med Res Rev. 2019;1-46. wileyonlinelibrary.com/journal/med © 2019 Wiley Periodicals, Inc.
|
1
Received: 18 February 2019
|
Revised: 22 May 2019
|
Accepted: 13 June 2019
DOI: 10.1002/med.21622
REVIEW ARTICLE
BACE1 inhibitors: Current status and future
directions in treating Alzheimer's disease
Nour M. MoussaPacha
1
|
Shifaa M. Abdin
1
|
Hany A. Omar
1,2,3
|
Hasan Alniss
1,2
|
Taleb H. AlTel
1,2
1
Sharjah Institute for Medical Research,
University of Sharjah, Sharjah, United Arab
Emirates
2
College of Pharmacy and College of
Medicine, University of Sharjah, Sharjah,
United Arab Emirates
3
Faculty of Pharmacy, BeniSuef University,
BeniSuef, Egypt
Correspondence
Taleb H. AlTel, Sharjah Institute for Medical
Research, University of Sharjah, P.O. Box
27272 Sharjah, UAE.
Email: taltal@sharjah.ac.ae
Funding information
University of Sharjah, Grant/Award Number:
15011101007 and 15011101002
Abstract
Alzheimer's disease (AD) is an irreversible, progressive
neurodegenerative brain disorder with no current cure.
One of the important therapeutic approaches of AD is the
inhibition of βsite APP cleaving enzyme1 (BACE1), which is
involved in the ratelimiting step of the cleavage process
of the amyloid precursor protein (APP) leading to the
generation of the neurotoxic amyloid β(Aβ) protein after
the γsecretase completes its function. The produced
insoluble Aβaggregates lead to plaques deposition and
neurodegeneration. BACE1 is, therefore, one of the
attractive targets for the treatment of AD. This approach
led to the development of potent BACE1 inhibitors, many of
Abbreviations: AChE, Acetylcholinesterase; AD, Alzheimer's disease; AD3, Alzheimer's disease gene 3; AD4, Alzheimer's disease gene 4; Ala, Alanine
amino acid; ALS, Amyotrophic lateral sclerosis; APH1, Anterior pharynxdefective 1; APOE, Apolipoprotein E; APP, Amyloid precursor protein; APPsα,
Soluble amyloid precursor protein αfragment; APPsβ, Soluble amyloid precursor protein βfragment; Arg, Arginine amino acid; ARIA, Amyloidrelated
imaging abnormalities; ARIAE, Amyloidrelated imaging abnormalities attributed to edema or effusion; Asn, Aspargine amino acid; Asp, Aspartic acid;
ATP, Adenosine triphosphate; Aβ, Amyloid βprotein; Aβ40, Amyloid βprotein with 40 amino acids; Aβ42, Amyloid βprotein with 42 amino acids; Aβo,
Soluble Aβoligomer; BACE1, Betasite amyloid precursor protein cleaving enzyme1; BACE2, Betasite amyloid precursor protein cleaving enzyme2;
BBB, Bloodbrain barrier; CatD, Cathepsin D; CatE, Cathepsin E; CB1, Cannabinoid type 1 receptor; CB2, Cannabinoid type 2 receptor; cHEA, Cyclic
hydroxyethylamine; CML, Chronic myeloid leukemia; CNS, Central nervous system; CSF, Cerebrospinal fluid; Cys199, Cysteine 199 amino acid; DMF, 5,7
Dimethoxyflavone; ECS, Endocannabinoid system; FDA, Food and Drug Administration; FRET, Fluorescence resonance energy transfer; Glu, Glutamic
acid; Gln, Glutamine amino acid; Gly, Glycine amino acid; GSK3, Glycogen synthase kinase3; GSK3β, Glycogen synthase kinase3β; HEA,
Hydroxyethylamine; HIV, Human immunodeficiency Virus; HTS, High throughput screening; IC50, 50% Inhibitory concentration; Ig, Immunoglobulin; Ile,
Isoleucine amino acid; Ki, Inhibitory constant; KP, Kaempferia parviflora;LCMS/MS., Liquid chromatography mass spectrometry; Leu, Leucine amino acid;
mAbs, Monoclonal antibodies; MRI, Magnetic resonance imaging; MTLs, Multitarget ligands; MTDLs, Multitargetdirected ligands; MWM, Morris water
maze; Nct, Nicastrin; NFT, Neurofibrillary tangles; NMDA, Nmethyl Daspartate; NRG1, Neuregulin1; Pgp, Pglycoprotein; Ptau, Phosphorylated tau;
PEN2, Presenilin enhancer 2; PK, Pharmacokinetics; PMF, 3,5,7,3,4′‐Pentamethoxyflavone; PrPC, Cellular prion protein; PS1, Presenillin 1; PS2,
Presenillin 2; RIPK1, Receptorinteracting serine/threonineprotein kinase 1; SAR, Structure activity relationship; Ser, Serine amino acid; SEZ6, Seizure
protein 6; SFKs, Src family of nonreceptor tyrosine kinases; SOD1, Superoxide dismutase 1; SRC, Sarcomafamily kinases; SRIS, Systemic inflammatory
response syndrome; TGN, TransGolgi network; THC, Tetrahydrocannabinol; Thr, Threonine amino acid; TMF, 5,7,4′‐Trimethoxyflavone; Trp, Tryptophan
amino acid; Tyr, Tyrosine amino acid; Val, Valine amino acid.
Nour M. MoussaPacha and Shifaa M. Abdin contributed equally to this study.
which were advanced to late stages in clinical trials.
Nonetheless, the high failure rate of lead drug candidates
targeting BACE1 brought to the forefront the need for
finding new targets to uncover the mystery behind AD. In
this review, we aim to discuss the most promising classes of
BACE1 inhibitors with a description and analysis of their
pharmacodynamic and pharmacokinetic parameters, with
more focus on the lead drug candidates that reached late
stages of clinical trials, such as MK8931, AZD3293,
JNJ54861911, E2609, and CNP520. In addition, the
manuscript discusses the safety concerns and insignificant
physiological effects, which were highlighted for the most
successful BACE1 inhibitors. Furthermore, the review
demonstrates with increasing evidence that despite
tremendous efforts and promising results conceived with
BACE1 inhibitors, the latest studies suggest that their
clinical use for treating Alzheimer's disease should be
reconsidered. Finally, the review sheds light on alternative
therapeutic options for targeting AD.
KEYWORDS
Alzheimer's disease, amyloid hypothesis, amyloidβ, BACE1
inhibitors, Fyn, GSK3β,βsecretase
1
|
INTRODUCTION
Dementia is a syndrome that encompasses a variety of symptoms with gradual progression, which hinders early
diagnosis of the disease. Despite the progressive development of the symptoms, not all individuals share the
same rate of disease progression. Dementia can be exhibited as a manifestation of multiple diseases. However,
the dominant form of dementia is known to be Alzheimer's disease (AD).
1
AD is a progressive neurodegenerative
disease characterized by memory and neuronal loss, difficulties in speaking, problemsolving, and other cognitive
skills, along with changes in the mood and behavior, which interfere with the person's daily performance.
2,3
These
symptoms result from the neuronal cell damage that is responsible for the cognitive function in the brain. This
damage may extend to neurons in different parts of the brain. Thus, additional adverse symptoms may appear on
the affected individual. Living with this disease can be very disabling, especially, that AD could be ultimately fatal. It
is believed that Alzheimer's is a complex disease where several factors contribute to its development. The
uttermost risk factors and causes for AD include age, lifestyle, environmental factors, family history, and bearing
the Apolipoprotein E (APOE)ε4 gene.
4
Moreover, genetic mutations are one of the causative factors of AD, as AD
genes were reported to be expressed on different chromosomes causing subsequent mutation in the expressed
gene product.
2
For instance, AD gene 3 (AD3) was found to be located on chromosome 14 leading to mutations on
Presenilin 1 gene. In parallel, AD gene 4 (AD4) is considered as a candidate for chromosome 1, which would cause a
mutation of Presenilin 2 gene.
2
In addition, one of the identified genetic mutations associated with AD is the
presence of AD gene 1 on chromosome 21 causing abnormal changes in the amyloid precursor protein (APP).
2
2
|
MOUSSAPACHA ET AL.
These genetic changes are affecting the function of vital gene products in the brain leading ultimately to the
abnormality of AD, specifically familial AD, which is characterized as an early onset form of AD.
2
It is worth mentioning that dementia is considered the 6th leading cause of death in the United States.
5
In 2015,
around 47.47 million people worldwide were diagnosed with dementia, and the number of people living with
dementia increases with the population ages.
6
It is expected to have these numbers triple from 50 million to
152 million by 2050.
6
Moreover, AD interferes with numerous aspects of human life including social, economic,
physical, and psychological aspects, which create a tremendous economic burden. Hence, the global socioeconomic
cost of dementia in 2015 was around US$ 818 billion.
6
The fact that such a lifethreatening disease was not yet
battled with potent diseasemodifying drugs made AD on the top agenda of many researchers and scientists, with
the hope to win the battle of AD treatment. Currently, there are several ongoing clinical trials for potential new
medications for AD (Table 1). However, the latest news from many of these clinical trials indicated a high failure
rate of many lead drug candidates, specifically the ones targeting βsite APP cleaving enzyme 1 (BACE1).
7-9
Moreover, the elegant contributions from Cummings et al highlighted several small molecules BACE1 inhibitors
that are currently in clinical trials for AD treatment. Around 26 candidates were introduced in clinical trials,
including agents that target the amyloidβ(Aβ) protein through BACE1 inhibition and others as immunotherapies.
10
In this article, we report the latest research results on AD, by covering the most recent molecules developed across
multiple classes of BACE1 inhibitors and focusing on the fate of the developed BACE1 inhibitors that advanced into
clinical trials. In addition, the latest developments in the field that aim to provide promising diseasemodifying
TABLE 1 BACE1 Inhibitors in different stages of clinical trials
Compd. name/code and structure Class Phase NCT number Status Sponsor Ref.
Anti
amyloid,
BACE1
inhibitor
Phase II/
III
clinical
trials
NCT02245737 Ongoing AstraZeneca, Eli
Lilly & Co.
7,8
Anti
amyloid,
BACE
inhibitor
Phase III
clinical
trials
NCT01739348 Terminated Merck
7,8
Anti
amyloid,
BACE
inhibitor
Phase III
clinical
trials
NCT03036280 Ongoing Eisai, Biogen
7,8
Anti
amyloid,
BACE
inhibitor
Phase II/
III
clinical
trials
NCT02406027 Terminated Janssen
7,8
CNP520 BACE1
inhibitor
Phase II/
III
clinical
trials
NCT02576639 Ongoing Novartis, Amgen
9
Abbreviation: BACE1, βsite APP cleaving enzyme 1.
MOUSSAPACHA ET AL.
|
3
drugs, such as immunotherapies and targeting routes other than the amyloid cascade hypothesis, are also
discussed.
2
|
ALZHEIMER'S DISEASE PATHOLOGY AND TARGETS
The main neuropathic hallmarks of AD are characterized as two lesions in the brain, namely, the extracellular
amyloid plaques and the intracellular neurofibrillary tangles (NFT) (Figure 1).
11
They appear initially in the
hippocampus that is responsible for the consolidation of information flow from shortmemory to longmemory and
extend to the cortical gray matter killing cells in the brain and compromising their functions.
12
Neurofibrillary tangles are insoluble bundles of fibers, comprised of the intracellular aggregation of
hyperphosphorylated tau (ptau) protein, which is responsible for microtubules stabilization. The presence of
these tangles is not limited to AD disease as they can also be found in other neurodegenerative disorders like Kuf's
disease and subacute sclerosing panencephalitis.
3
On the other hand, the neuritic plaques, or senile plaques, are
considered as a characteristic presenilin feature of AD, and they are visualized as spherical lesions surrounded by
an array of abnormal dendrites and axons, which results from the extracellular aggregation of amyloidβ(Aβ)
protein following the sequential cleavage of APP.
2
In the early steps of the disease pathogenesis, the accumulation
of Aβproteins is of significant concern as it leads to the formation of NFT, and has a neurotoxic and
neuroinflammatory effects that result in synaptic loss and neuronal cell death.
13
APP is a large type 1 membrane
protein present in many tissues throughout the body and it is mainly expressed in the brain and the kidneys.
13
There are three enzymes involved in the processing of this protein as illustrated in Figure 2, α,β, and
γsecretases.
14,15
The cleavage of APP by each of these three enzymes result in different end products. One of
these end products, which is the insoluble Aβprotein, is considered the leading cause of AD abnormality (Figure
2).
16
Despite the fact that AD is significantly correlated to the accumulation of the insoluble Aβplaques,
nonetheless, recent studies reported that the soluble form of amyloidβ,Aβo oligomers also participate greatly in
these neurodegenerative conditions by acting as neurotoxin causing severe synaptic damage.
17,18
In addition, it was
FIGURE 1 Pathological difference between neurons of the healthy brain and AD patient's brain. Amyloid
plaque accumulation is a key feature of AD. Amyloid plaque has been detected between the neurons in AD brain,
along with the formation of abnormal neurofibrillary tangles within the neuronal cells. Both lesions have a toxic and
degenerative effect on the neurons, and overall brain atrophy is observed. AD, Alzheimer's disease [Color figure
can be viewed at wileyonlinelibrary.com]
4
|
MOUSSAPACHA ET AL.
reported that these oligomers can induce their neurotoxic response by glial cells involvement.
18
In vivo studies to
further explore the cytotoxic effect of Aβoligomers were conducted. In one study, Aβoligomers were injected to
rats hippocampus, which resulted in several neurological defects such as loss of synaptic transmission, a decline in
cognitive function along with cell death upon the aggregation of Aβoligomers.
19
In fact, numerous reports confirm the toxic role of Aβoligomers in AD pathology, where they emphasize on the
fact that a true distinction should be made on whether Aβplaques or the soluble oligomers are the true targets to
peruse for alleviating AD hallmarks.
20
This finding was enforced from the results of a study comparing the effects of
fibrillar vs oligomeric forms of Aβ. The results of this study indicated that rats injected with Aβoligomers exhibited
severe neurological deficits with greater neurodegeneration and inflammation compared with rats injected with a
fibrillar form of Aβ.
21
Further evidence on the harmful role of Aβoligomers has emerged from the need to design
therapeutic strategies that eradicate Aβoligomers before the progression and development of Aβplaques, else the
therapy would not yield beneficial outcome.
22
Henceforth, it is evident that the accumulation of Aβoligomers plays
a vital role in AD progression, where AD severity could be linked to the degree of Aβaccumulation, which can
manifest its neurotoxic effect by a wide variety of mechanisms including apoptosis promotion, synaptic loss, pore
formation, and loss of membrane potential.
23
Other findings indicated that, when the substrate APP, is bound to αsecretase in the nonamyloidogenic
pathway, it produces an ectodomain, which is a soluble amyloid precursor protein αfragment (APPsα)leaving
behind Cterminus fragment bound to the membrane (C83), which will be converted by γsecretase to a
protein fragment called P3.
15
Henceforth, the enzymatic action of αsecretase is not leading to the formation
and accumulation of the insoluble Aβprotein, which rules out the need for considering this pathway as an
FIGURE 2 APP metabolism by secretase enzymes. APP is a large type 1 membrane protein that is subjected to
three enzymes, αβ, and γsecretases, which generate different end products. When αsecretase competes with
βsecretase for the APP substrate, a soluble ectodomain called APPsαis produced leaving behind a Cterminus
fragment bound to the membrane (C83) that will be processed to a protein fragment called P3 by the effect of
γsecretase. Accordingly, αsecretase action is not leading to the formation of any insoluble Aβprotein and this
pathway is known as the nonamyloidogenic pathway. On the other side, the amyloidogenic pathway starts from the
cleavage at the Nterminus of APP by βsecretase (BACE1), which leads to the production of two fragments, the
soluble ectodomain APPsβand the membranebound Cterminus (C99). C99 is then processed by γsecretase and
generate two types of Aβproteins, Aβ40 and Aβ42. Aβ42 is a key player in AD pathogenesis. AD, Alzheimer's
disease; APP, amyloid precursor protein [Color figure can be viewed at wileyonlinelibrary.com]
MOUSSAPACHA ET AL.
|
5
option to target of AD. On the contrary, it was found that the biosynthesis of Aβstarts with the cleavage at
the Nterminus of the protein by βsite APP cleaving enzyme 1 (BACE1). This leads to the production of two
fragments, the ectodomain that is a soluble amyloid precursor protein βfragment (APPsβ)andthe
membranebound Cterminus (C99). C99 is then processed by γsecretase complex which is constituted of
four transmembrane proteins: presenilin (PS1 or PS2), nicastrin (Nct), anterior pharynxdefective
phenotype (APH1), and presenilin enhancer 2 (PEN2).
15
The interaction of C99 with γsecretase leads to
the production of two types of Aβproteins, Aβ40 and Aβ42.
24
Aβ42 is a critical player in AD pathogenesis
according to the latest evidence, which has entitled this etiology as the amyloid cascade hypothesis.
25
Interestingly, apart from the amyloid hypothesis, there are several Aβindependent phenomena's that can
induce and contribute to AD severe hallmarks. For instance, it was reported that the accumulation
of C99 could foster great neurological deficits, as it wasobservedinaninvivotransgenicmicemodelwhere
C99 accumulation was induced without Aβdisposition or tau phosphorylation. These mice exhibited an early
apathy like behavior and synaptic plasticity alteration, which can be traced to be solely dependent
on C99 accumulation apart from any abnormal Aβaccumulation.
26
In addition, the Lauritzen group,
27
revealed that C99 accumulation can be directly associated with endosomalautophagiclysosomal
dysfunction, which is one of the famous features of AD. In this report, it was demonstrated that in two in
vivo transgenic mice models that have an accumulation of C99, indicated that this fragment is the main driver
of endosomalautophagiclysosomal dysfunction.
27
Moreover, Lauritzen et al,
28
discovered important
amount of evidence to support the role of C99 accumulation and oligomerization with early neurotoxicity
and onset of AD, as these reports elucidate that this fragment can mediate its toxicity by causing lysosomal
autophagic dysfunction, brain alteration, and memoryrelated behavioral changes.
28
Additional important
findings reported by the Nixon group,
29
indicated that the endosomalautophagic lysosomal dysfunction
induced by C99 accumulation could be one of the key drivers to AD abnormality. In fact, it was reported that
thereisacloseassociationbetweenβamyloidogenesis and endosomalautophagic lysosomal dysfunction,
where both pathological aspects stem from the common genetic basis and cooperate to produce the
multiple features of AD.
29
Together these reports indicated the complexity of AD pathology and
confirmed the toxic effect of C99, which could underlie some of the earlystage anatomical hallmarks of
Alzheimer's disease etiology.
Studies suggested that Aβ42 aggregates are responsible for the formation of senile plaques, which is believed to
initiate a neurotoxic cascade that will end up with the clinical manifestations of Alzheimer's disease.
30
Thus, targeting AD
might be achieved by inhibiting the Aβproduction, promoting its clearance, inhibiting its aggregation, or restricting the
toxic effects of the abnormal accumulation of Aβdeposits. These suggested approaches are considered attractive anti
amyloid strategies for tackling AD (Figure 3).
30,31
However, blocking the formation of the aggregates is generally believed
to be technically challenging to achieve. Moreover, the mechanism of Aβclearance has not been fully understood yet, and
up to date, there is no viable tactic to promote Aβclearance. These challenges have therefore pushed the research wheel
over the past 9 years to focus mainly on blocking the Aβgeneration. This goal can be attained by blocking the enzymes
involved in the production of Aβprotein. Thus, the aim of slowing the AD progression can result from blocking Aβ
generation when targeting both βand γsecretases.
31
Inhibition of γsecretase aimed to knock out the two most important proteins (PS1 and PS2) in the γsecretase
complex, which are the integral membrane proteins found in the endoplasmic reticulum and Golgi apparatus. These
proteins control the Notch signaling pathway which is accountable for cell differentiation and proliferation during
embryonic development.
32
Despite the fact that it was very appealing for many researchers to develop several
inhibitors for γsecretase, the results of its inhibition were not satisfactory. Transgenic PS1 knockout mice were
unhealthy, not fertile and had lagging subventricular areas and cortical dysplasia.
32
Thus, blocking γsecretase is more likely to result in adverse effects due to the vital biological function of PS1.
The undesirable results obtained from γsecretase inhibition drove the attention toward the inhibition of
βSecretase as an attractive target to reverse or stop the disease progression.
6
|
MOUSSAPACHA ET AL.
βsecretase is an aspartyl protease that cleaves the APP in the lumina and is believed to be the ratedetermining
step in the Aβgeneration. BACE1 inhibition provides multiple advantages, among which is the prevention of Aβ
formation at an early stage of APP processing. Moreover, BACE1 knockout homozygote mice exhibited a total loss
of Aβproduction without any significant side effect.
32
Moreover, many in vivo studies confirmed that βsecretase
play a vital role in AD progression and the use of BACE1 inhibitors improved the neurological function and
surpassed the neurological defects imposed in mice of AD model.
33
For example, in a recent study utilizing NB360
as BACE1 inhibitor in mice with familial Alzheimer disease, the administration of this agent resulted in the
enhancement of neuronal activity and the improvement of memory.
34
These effects were attributed to the ability
of NB360 to reduce the formation of Aβoligomers. When researchers reintroduced soluble Aβoligomers to these
mice, the positive implication of NB360 was reverted and the mice showed again the typical symptoms of AD.
34
Therefore, these in vivo studies revealed evidence on the implication of BACE1 inhibitors as a valid approach to
tackle AD and that Aβoligomers are having a downstream effect on AD progression.
35
Based on many reports, even BACE1 inhibitors have the ability to reduce the production and to lower the levels
of Aβ, it is still not clear if these effects would translate into a treatmentof AD disease. These results are
connected to the failure of many BACE1 inhibitors in clinical trials which reflects on the controversy between the
highly effective BACE1 inhibitionmediated treatments of AD in animal models vs the failure in clinical trials.
FIGURE 3 Amyloid cascade hypothesis and therapeutic approaches. Since Aβ42 aggregates are believed to be
the key player in the formation of senile plaques that will initiate a neurotoxic cascade resulting in the appearance
of AD clinical manifestations, targeting AD may be attained by inhibiting the Aβproduction, inhibiting its
aggregation, promoting its clearance, or restricting the toxic effects of the abnormal accumulation of Aβdeposits.
All of these suggested approaches are considered attractive antiamyloid strategies for tackling AD. However,
researchers have faced several challenges in most of these approaches, pushing the research wheel mainly on
blocking the Aβgeneration by blocking the enzymes involved in the production of Aβprotein, both βand
γsecretases. AD, Alzheimer's disease [Color figure can be viewed at wileyonlinelibrary.com]
MOUSSAPACHA ET AL.
|
7
One possible explanation for such an incongruity might be related to the difference in the genetic background
between mice and humans. Such a genetic difference between the two species was fruitfully discussed before.
36,37
However, the persistent negative results of BACE1 inhibitors in clinical trials should not be overlooked since these
results contain important lessons to the future. While some understood these findings as a judgment of amyloid
cascade hypothesis (ACH) validity, however, such a conclusion appears to be premature; since the ACH remains the
most recognized theory when compared with other alternative interpretations of AD pathology. At this junction,
the dark picture of clinical trials results using BACE1 inhibitors and perhaps other Aβtargeted therapies does not
indicate that such treatments will be in place in the near future. Furthermore, due to the complexity of AD, BACE1
inhibition alone is insufficient for clinical improvement via the reduction of Aβ. This approach may have
to be combined with other strategies like Aβand/or tau clearance for diseasemodifying effects. In summary, with
these unsuccessful clinical trials in hand, many lessons to the future can be framed as follows: (a) familial
Alzheimer's disease (FAD) and sporadic Alzheimer's disease (SAD) should be contemplated as two markedly
different diseases as far as the mechanisms of amyloid production are concerned. (b) Human trials of BACE1
inhibitors should be conducted separately, with two independent familial AD and sporadic AD cohorts. (c) Targeting
other aspects of Alzheimer's disease other than βsecretase of the Aβprecursor protein might have comparable
impacts in familial and sporadic AD lesions.
Another potential method to target AD is the inhibition of tau aggregation by small molecules to avoid
the development of tau lesions, however, unfortunately, this approach lacks a complete understanding of the
pathology of tau.
38
Taken together these findings paved the way for the design and synthesis of BACE1 inhibitors and in the
following sections, the most promising motifs are discussed.
3
|
BACE1 AS A POTENTIAL TARGET FOR THE TREATMENT OF
ALZHEIMER'S DISEASE
3.1
|
Structure and properties of BACE1
BACE1 has taken a central stage in targeting AD and several companies have developed BACE1 inhibitors to tackle
this devastating illness and many of these inhibitors reached advanced stages in clinical trials. Inhibition of BACE1
is therefore considered one of the major methodologies to address AD.
13
However, many obstacles need to be overcome to design effective and physiologically significant BACE1
inhibitor with minimal offtarget effects. BACE1 is a structurally challenging protein that possesses structural
similarities with other aspartyl proteases,
39
a family that includes many enzymes distributed in different parts of
the human body such as BACE2, pepsin, renin, cathepsin D (CatD), and cathepsin E (CatE). Thus, achieving
selectivity in BACE1 inhibition without affecting other proteases is crucial for developing effective BACE1
inhibitors and eliminating the offtarget side effects.
39,40
Furthermore, an additional challenge in the discovery of BACE1 inhibitor is related to the size of BACE1 active
site, which consists of catalytic aspartic acid residues, flap, and 10S loop, which is known to be of a relatively large
size (Figure 4), therefore, having a small molecule to occupy this large active site represents a grand challenge.
31
In addition, the issue of the ability of these compounds to pass the bloodbrain barrier (BBB), is another
challenge.
41
Furthermore, many of the developed BACE1 inhibitors were prone to efflux by Pglycoprotein (Pgp),
which complicates the process of drug entry to the brain, even when brain permeability is attained via BBB
penetration. Pgp efflux is, therefore another limiting factor of these inhibitors.
41
Despite the above potential hurdles in the design of BACE1 inhibitors, many laboratories succeeded in
developing selective, potent, and bioavailable inhibitors. To date, many of these compounds have shown promising
clinical effects, however, none have passed the final stages of clinical trials yet to receive final FDA approval and
reach the market.
8
8
|
MOUSSAPACHA ET AL.
Before starting to delineate the chemistry and structureactivity relationships (SAR) of BACE1 inhibitors, it is
worth discussing some structural features of the enzyme. BACE1 belongs to type I transmembrane aspartyl
proteases family and consists of 501 amino acids (Figure 5).
42
It has an Nterminal,aCterminal, and an
interdomain, which connects the Nterminal and Cterminal domains.
Most of the structural characteristics of BACE1 were exploited in the discovery of the earliest BACE1 inhibitors
OM992, an eightresidue transitionstate inhibitor, that showed a potent inhibitory activity for BACE1 with
K
i
= 1.6 nM. The interaction of the OM992 eight residues with the BACE1 active site is illustrated in Figure 6.
43,44
Twentyeight of the amino acid residues composing the active site of BACE1 were recognized as ligandbinding
site and all of them within 5 Å from the ligand.
45
The ligandbinding site is presented as a surface including the
subpockets S1, S2, S3, S4, S1,S2,S3, and S4.
46
Moreover, optimal BACE1 activity is achieved at acidic pH, as
BACE1 is located within acidic intracellular compartments including transGolgi network (TGN) and endosomes
where it cleaves its substrates leading to the formation of Aβprotein fragments. The optimal enzyme activity under
acidic conditions suggests that a BACE1 inhibitor having a basic amine residue with a pK
a
of ca. 6.0 or more would
exhibit better binding affinity.
41
The catalytic dyad, Asp32, and Asp228 are located in the ligandbinding sites and
are crucial for the proteolytic activity of the enzyme.
47
Binding of a ligand to these two amino acids enhances its
binding affinity and in turn, its potency.
The βsecretase active site is shielded by a βhairpin loop, which is a large part of the binding site, between
Val67 and Glu77 in the Nterminal lobe. It is usually known as the flapand considered the most flexible part of
the active site that is believed to control the access of the substrate to the active site by its conformational
changes. When the active site is in the inactive form, the flap tends to be in its open conformation. However, the
presence of a substrate or inhibitor stabilizes the flap in its closed form.
47
An important residue in the flap is the Tyr71 that takes a conformation complementary to the nature and shape
of the ligand when bound to the binding site. Changing the flap position relative to the catalytic aspartic acid
residues affords a mean for the substrate or the ligand to diffuse in and out of the active
site. Furthermore, it is also documented that Tyr71 flexibility plays a fundamental role in defining
BACE1 confirmation as opened or closed.
47
The ability of Tyr71 hydroxyl group to form a hydrogen bond with
FIGURE 4 The binding pocket of BACE1 contains three important parts, A, the catalytic aspartic acid residues
that are crucial for the proteolytic activity of BACE1, B, the flap that is the most flexible part of the binding site and
controls the access of the substrate, and C, the 10 seconds loop that is located near the S3 pocket [Color figure can
be viewed at wileyonlinelibrary.com]
MOUSSAPACHA ET AL.
|
9
the NH of the Trp76 side chain permits the Tyr71 to physically separate the S1 and S2subpockets, resulting in a
closed BACE1 confirmation. While the open BACE1 conformation is observed when the flap moves away from the
catalytic Asp, which in turn abolishes the formation of a hydrogen bond between the residues Tyr71 and Trp76, in
that case, Tyr71 will not be appropriately positioned between the S1 and S2subpockets. In other words, the
existence of a hydrogen bond between the residues Tyr71 and Trp76 implies that the conformation of BACE1 as
closed, and the absence of this Hbond describe it as open.
47,48
Molecular dynamics simulation studies proved that
the open and closed forms are freely accessible at room temperature, which implicates the conformational
flexibility of Tyr71 and the active site flap in BACE1. This unique structure and composition of this enzyme have
permitted the design of a variety of inhibitors.
41
FIGURE 5 Structural characteristics of BACE1 binding site. BACE1 is a type I transmembrane aspartyl protease
consists of 501 different amino acids with a Nterminal,aCterminal, and an interdomain. BACE1, betasite
amyloid precursor protein cleaving enzyme1 [Color figure can be viewed at wileyonlinelibrary.com]
10
|
MOUSSAPACHA ET AL.
Another essential feature of βsecretase active site is the 10 seconds loop which is positioned near Ser10, in the
S3 pocket (Figure 4). When the 10 seconds loop adapts an open conformation, this allows for maximum binding of
substrate to the S3 pocket. The three entities, catalytic aspartic acid residues, βhairpin loop, and 10 seconds loop
together form a binding pocket for βsecretase's substrate or inhibitor. When a ligand binds, this triggers the flap to
close which stimulates further interaction with the 10 seconds loop, generating a stable complex structure.
49
Another feature in the BACE1 structure is the presence of water molecules that is known as the special pocket,
which is not usually occupied by the ligand. The water molecule that is allocated in the middle of the catalytic site is
believed to have an essential role in the hydrolysis of peptide bonds.
41
The latest hypothesis related to the BACE1
mechanism of action suggested that one of the catalytic aspartyl residues acts as an acid/base sink. Kinetic studies
demonstrated that the basic pK
a
value for the Asp residue is 3.5 and that for the acidic residue feature a pK
a
value
of 5.2. Moreover, the best in vitro catalytic activity of the enzyme was obtained when the pH approximates to 4.5.
41
For complete activation, the enzyme acquires mono protonation in the active form (Figure 7).
3
The four oxygen
atoms of Asp32 and Asp228 of BACE1 active site are not equal owing to their different chemical environment,
FIGURE 6 The binding interactions of OM992 with the active site of BACE1. OM992 is the first BACE1
inhibitor that binds to eight amino acid residues in different subpockets. BACE1, betasite amyloid precursor
protein cleaving enzyme1
FIGURE 7 Mechanism of action of the catalytic aspartic acid residues. Water molecules have a vital role in facilitating
the hydrolysis of peptide bonds by the aspartic acid residues [Color figure can be viewed at wileyonlinelibrary.com]
MOUSSAPACHA ET AL.
|
11
suggesting a possible set of distinct protonated states. Qualitative analysis of this state suggests that only the
monoand dideprotonated species are relevant for the design of potent inhibitor.
41
With these in mind, the effective inhibition of BACE1 requires an inhibitor that possesses a higher affinity to
the binding site of the enzyme when compared to the indigenous substrate. This form of competitive inhibition can
be achieved by maximizing the number of binding interactions between BACE1 and the inhibitor, with focus on
binding to the catalytic amino acid dyad.
50
The elongated active site of BACE1 allows for accommodating up to
11 amino acids of substrates. The 11 subpockets have a broad amino acid tolerance, but many central ones (such as
P1 and P1) are hydrophobic in nature and accommodate hydrophobic side chains. This characteristic can be
advantageous in the development of BACE1 inhibitors with enhanced lipophilicity to improve their membrane
permeability and penetration of the BBB.
50
3.2
|
Classes of BACE1 inhibitors
3.2.1 |Peptidomimetic BACE1 inhibitors
Compounds in this class feature the presence of an amide bond or one of its isosteres.
51
Many transitionstate
bioisosters have been used as the main element for the design of aspartyl protease inhibitors including
hydroxyethyleneisostere, hydroxyethylamine, isophthalamidebased inhibitors, and others like reduced
amide, statins, and norstatinsas well as macrocyclicbased inhibitors.
52
Hydroxyethylene isosteric inhibitors are the first class of BACE1 inhibitors to be developed. Designing a
hydroxyethylenebased inhibitor with an oxazolylmethyl substituent at P3 position such as compound 1(Table 2),
led to a very potent peptidomimetic BACE1 inhibitor (K
i
= 0.12 nM), with high selectivity of more than 3800fold
over BACE2 and 2500fold over CatD.
53
Other structural modifications were performed to develop inhibitors with
LeuAla isoster and isopthalamide moieties at P2 position.
54
Thus, incorporation of a hydrophobic group to this
skeleton at P3 position led to the formation of compound 2, where the (R)methyl benzylamide showed better
potency than benzylamide or (S)methyl benzylamide. Although this inhibitor possesses a slight selectivity against
BACE2 and CatD (28 and 37fold respectively), it possessed an excellent BACE1 IC
50
of 39 nM.
54
However, the
results of in vivo studies were encouraging, as the intraperitoneal injection of 8 mg/kg of compound 2resulted in a
30% reduction of Aβ40 brain levels in transgenic mice after 4 hours.
54
In other work, a series of hydoxyethylene
based compounds were synthesized with the aim to discover the potential interactions with BACE1 S1 and S3
binding subpockets. The most potent compound among the 30 compounds developed in this series was compound
3that exhibited a remarkable IC
50
value of 69 nM.
55
This compound possesses bulkier substituents on the
hydroxyethelene skeleton compared with other compounds in the same series, which indicated that the binding site
of the enzyme at P1 and P3 dispositions could accommodate such large moieties. Nonetheless, this promising
activity was not supported by the cell binding assay results that showed a very low percentage of inhibition (ca. 3%)
of Aβ. This weak activity might be due to the polar amide residues contained in this compound and its high
molecular weight which might be limiting factors that affected its cell permeability.
55
Other important work from Chang et al,
56
focused also on the design of potent BACE1 inhibitors based on the
hydroxyethylamine (HEA). For example, compound 4with a 3methoxybenzyl group at P1, and phenylalanine side
chain at P1showed a subnanomolar IC
50
of 1.0 nM.
56,57
Furthermore, compound 4revealed a 39fold selectivity
for BACE1 over BACE2 and 23fold selectivity over CatD. Administration of 8 mg/kg of inhibitor 4in Tg2576 mice
showed a 65% reduction of Aβ40 levels. Other in vivo cognitive studies were performed on transgenic mice, where
the compound was given in a dose of 33.4 μg/day to four groups of different ages and different treatment periods.
The results were promising and revealed that the cognitive function in younger AD mice can be rescued by a partial
reduction in Aβformation and neuritic plaque. The same study indicated no detectable signs of toxicity or
accumulation of uncleared APP.
56
Another important hydroxyethylene bioisoster is compound 5, which was developed by the Beswick group. This
sixmembered ring, sultam derivative 5, exhibited a high BACE1 IC
50
and cellular Aβactivity
58
and a 44fold
12
|
MOUSSAPACHA ET AL.
TABLE 2 Peptidomimetic BACE1 inhibitors profile (IC
50
, selectivity, efflux ratio, and bioavailability)
Hydroxyethylenebased Inhibitors BACE1 IC
50
BACE1 selectivity
Pgp efflux ratio Bioavailability Ref.Over CatD Over BACE2
K
i
= 0.12 nM > 2500 fold > 3800 fold ‐‐
39
K
i
= 1.1 nM 37fold 28fold ‐‐
40
69 nM ‐‐
41
Hydroxyethylaminebased inhibitors
1nM 23fold 39fold ‐‐
42,43
(Continues)
MOUSSAPACHA ET AL.
|
13
TABLE 2 (Continued)
Hydroxyethylenebased Inhibitors BACE1 IC
50
BACE1 selectivity
Pgp efflux ratio Bioavailability Ref.Over CatD Over BACE2
4 nM 663fold 44fold Low
44
55 nM 7.6fold 3
45
2 nM 100fold 2
45
Isophthalamidebased inhibitors
15 nM ‐‐Low
46,48
18 nM 204fold ‐‐
25
2.5 nM 60fold ‐‐ ‐
51
(Continues)
14
|
MOUSSAPACHA ET AL.
TABLE 2 (Continued)
Hydroxyethylenebased Inhibitors BACE1 IC
50
BACE1 selectivity
Pgp efflux ratio Bioavailability Ref.Over CatD Over BACE2
0.32 nM > 1000 fold ‐‐ ‐
51
0.4 nM > 250 000 fold 42fold Low
52
Note: Data is not available.
Abbreviation: BACE1, betasite amyloid precursor protein cleaving enzyme1.
MOUSSAPACHA ET AL.
|
15
selectivity for BACE1 over BACE2. The remarkable potency of this compound was believed to be due to the
presence of metaethylamine residue resident in the S3 subpocket, hence enhancing the binding affinity in the
active site by interacting with the catalytic dyad. Unfortunately, compound 5showed low oral bioavailability in
transgenic mice studies. This was revealed following oral administration at a dose of 250 mg/kg two times daily that
led to only a 23% decrease in Aβ42 levels in the brain of the diseased mouse. Interestingly, in the same study, it was
also found that the administration of compound 5with Pgp inhibitor reduced the Aβ42 levels by ca. 55%.
58
An elegant contribution from Rueeger et al is the design and synthesis of a novel class of cyclic hydroxyethylamine
(cHEA) BACE1 inhibitors employing the de novo design strategy. These compounds, this class of compounds should be
mentioned here, were found to possess good BBB permeability.
59
In this study, it was concluded that such a
framework might offer appropriate attachment vectors for direct extensions into the BACE1 subpockets. Thus, guided
by structurebased optimization, a series of cHEABACE1 inhibitors containing 4hydroxybenzyl arms were developed
and showed greater BBB penetration and less Pgp efflux compared with their acyclic HEA congeners. For example,
compound 6exhibited a BACE1 IC
50
of 55 nM with an efflux ratio of3. On the basis of these results, further structural
modifications were pursued to improve the potency, selectivity and metabolic stability, to deliver a highly potent
alkoxysubstituted 4aminofluorobenzyl cHEA inhibitors, such as compound 7with a subnanomolar potency against
BACE1 (IC
50
= 2.0 nM) and more than 100fold selectivity for BACE1 over catD.
59
Isophthalamidebased inhibitors are another transitionstate isosteres that have been developed as BACE1
inhibitors with increased BBB penetration while preserving the potency achieved by the HEA inhibitors. These
compounds have shown remarkable selectivity for BACE1 over other aspartic proteases.
60,61
For example, inhibitor
8exhibited considerable in vitro activity against BACE1 (IC
50
= 15 nM). Molecular modeling studies of
isophthalamide scaffolds revealed the formation of essential Hbonding with the catalytic aspartates (Asp32 and
Asp228) by the hydroxyethylamine substituent, for example, 8. While the cyclopropyl moiety was found to be
directed toward the S1subpocket of the enzyme active site. Furthermore, the righthand amide arm of the same
compound was found to be crucial appendage for the inhibition of BACE1 by forming two essential hydrogen bonds
with the enzyme's binding site (the amide NH with the Gly230 carbonyl group and the amide carbonyl with the NH
group of Gln73).
62
Unfortunately, compound 8showed poor BBB penetration, which encouraged other researchers
FIGURE 8 Cocrystal structure of compound 20 bound to BACE1. The Xray structure showed that the
3azaxanthene core interacts with Trp76 by crucial hydrogen bonds, while other hydrogen bonds are formed
between the oxygen of the dihydropyran and Tyr198, and between the pyridyl nitrogen and Ser229. BACE1,
betasite amyloid precursor protein cleaving enzyme1 [Color figure can be viewed at wileyonlinelibrary.com]
16
|
MOUSSAPACHA ET AL.
to design and develop new peptidomimetic analogs based on isophthalic acid scaffolds with the aim to enhance BBB
permeability. For instance, AlTel and coworkers used compound 8as a lead to design and synthesize new motifs
derived from isophthalic acid with enhanced pharmacodynamic properties as potential BACE1 inhibitors,
employing different structurebased drug design strategies. The traditional medicinal chemistry strategies followed
in this study, such as rigidification, contraction, and bioisostere replacement led to the development of BACE1
inhibitors with high potencies.
39,63
Scheme 1 illustrates the main strategies utilized to design and develop a series
of BACE1 inhibitors with improved affinity and inhibition activity.
64
The most potent compound of this series
(compound 9), exhibited a BACE1 IC
50
of 18 nM, with enhanced selectivity toward BACE1 over BACE2.
39
FIGURE 9 Role of RIPK1 inhibition in inflammation and necroptosis. RIPK1 inhibition led to the disruption of
toxic activity of the resultant complexes which in turn reduced the inflammatory mediators and amyloid burden,
and improved memory function. RIPK1, receptorinteracting serine/threonineprotein kinase 1 [Color figure can be
viewed at wileyonlinelibrary.com]
N
H
HN
N
O
S
31
FIGURE 10 Structure of RIPK1 inhibitor. RIPK1, receptorinteracting serine/threonineprotein kinase 1
MOUSSAPACHA ET AL.
|
17
Another important initiative from Bjorklund et al who developed promising isophthalamide analogs, for
example, compound 10. Initially, compound 10 with an IC
50
of 2.5 nM and K
i
of 1 nM was prepared, however, the
low selectivity toward CatD (K
i
= 59 nM) was the main drawback of this inhibitor. Henceforth, the focus was shifted
toward the main difference between the S1subpockets of BACE1 and that of CatD to optimize the potency and
improve the selectivity over CatD. In this context, SAR studies showed that the S1pocket of BACE1 could
accommodate larger groups and form polar interactions, contrary to CatD S1subpocket which can only engage in
hydrophobic interactions with the ligand. Based on this observation, the introduction of a small polar group like
methoxyl at P1, provided the most potent inhibitor in this series, compound 11, which showed an increase in
potency (IC
50 =
0.3 nM) and highly improved selectivity against CatD of more than 1000 folds.
65
Related to the same group also developed compound 12 which is isophthalamidebased inhibitors bearing a
1,3,4oxadiazole ester moiety. This compound was found to have an excellent potency against BACE1 with an IC
50
of 0.4 nM, in which both of the P2 and P3 arms were adjusted to increase the efficacy and cellular permeability.
However, it revealed moderated selectivity (42fold) against BACE2 and improved selectivity (more than 250,000
fold) against CatD.
66
Intraperitoneal administration of inhibitor 12 at a 100 mg/kg dose in mice resulted in a
moderate decline of Aβ
40
concentration (26%) levels after 4 hours, with a brain exposure of 1.8 μM. Interestingly,
time and dosedependent studies, using this compound, in rhesus monkeys displayed a 65% reduction of plasma
Aβ40 levels after 4 hours postinjection, a return of Aβ40 levels after 8 hours, and full recovery after 24 hours.
N
N
NH
2
O
HN
O
O
32
N
N
S
O
O
33
S
O
Cl Cl
34
R
O
X
35
FIGURE 11 Structure of GSK3βinhibitors
FIGURE 12 Aβo Role in Fyn kinase activation. Aβo are soluble Aβaggregates with synaptotoxic effect. They
interact with PrPC on the neuronal cell membrane activating a downstream signaling cascade via the nonreceptor
tyrosine kinase Fyn and generating cellular damage. Thus, studies suggested Fyn inhibition as a therapeutic
approach to suppress and minimize neuronal damage [Color figure can be viewed at wileyonlinelibrary.com]
18
|
MOUSSAPACHA ET AL.
Unfortunately, due to the extensive hepatic metabolism by CYP3A4 enzyme, the oral bioavailability of this
compound was found to be poor with less than 1% bioavailability upon 10 mg/kg oral dosing.
66
In conclusion, the last decade or so has witnessed the synthesis of many peptidomimeticbased inhibitors,
particularly HEA inhibitors, but unfortunately, many of these peptidomimetics showed poor oral bioavailability,
short in vivo halflife, limited BBB penetration, and low reduction of Aβlevels, which ultimately placed these
compounds in the back chamber of the moving train.
55
Although the isophthalamide class showed very promising
results, for example, compound 11, with BACE1 IC
50
of 0.32 nM, and more than 1000fold selectivity for BACE1
over CatD, however, none of these motifs made it to clinical trials. Overall, despite the tremendous efforts that
were made to design peptidomimeticbased inhibitors, the majority of these inhibitors did not exhibit good
pharmacological properties. Nevertheless, these motifs furnished the way toward the design of other classes of
inhibitors with improved pharmacokinetic and pharmacodynamic parameters. The focus was therefore shifted
toward the next generations of BACE1 inhibitors, namely the nonpeptidomimetic inhibitors.
52
3.2.2 |Nonpeptidomimetic BACE1 inhibitors
The development of nonpeptidomimetic BACE1 inhibitors has been facilitated by the utilization of highthroughput
screening (HTS) campaigns followed by the SAR development of the best hits. These inhibitors offered various
advantages over the peptidomimetic inhibitors, including improved metabolic stability. Moreover, nonpeptidomi-
metic inhibitors are smaller in size, which had improved BBB permeability and decreased Pgp efflux ratio. Many
different classes of nonpeptidomimetic inhibitors were developed. A salient feature of these inhibitors is the
presence of an aminoazine residue embedded in their scaffold. The different classes of these inhibitors could be
seen in compounds that contain acyl guanidine,2aminopyridine, aminoimidazole, amino/iminohydantoin,
aminothiazoline and aminoxazoline, dihydroquinazoline, aminoquinoline, and aminopyrrolidinebased Inhibitors
(Table 3).
66
The acyl guanidine scaffold attracted the attention of many research groups due to their highly distinctive
features compared to other nonpeptidomimetic BACE1 inhibitors. They showed high BACE1 potency with
remarkable pharmacokinetic properties including very low Pgp efflux, which allowed these compounds to reach
advanced stages in clinical trials (Table 1).
42,67
It has also been found that, adding a pyrrolidine moiety to
O
O
Cl H
N
NN
ON
N
O
O
36
H
N
O
Cl
N
S
N
H
N
N
N
N
HO
37
FIGURE 13 Structure of Fyn kinase inhibitors
O
O
N
O
H
H
N
O
OOH
H
H
38 39 40
FIGURE 14 Structure of cannabinoid type 2 receptor agonists
MOUSSAPACHA ET AL.
|
19
guanidinecontaining compounds, such as compound 13, which is substituted at position 4 with Nacyl pyrrolidine
group, resulted in the development of one of the most potent compounds compared with similar derivatives in the
same series with BACE1 K
i
value of 0.21 μM. Molecular modeling studies of this compound revealed a unique U
shaped conformation that directs the aryl group toward S2 subpocket of BACE1.
68
Furthermore, structureactivity relationship (SAR) studies of these inhibitors revealed that the addition of an
ether linkage on the aryl group was highly tolerated, which encouraged researchers to pursue macrocyclization.
These initiatives were undertaken as a rigidification strategy to enhance the pharmacokinetic profile of these
motifs. Such a strategy resulted in an improvement of both in vitro and in vivo potency, with a significant reduction
in Aβformation. In the same series, compound 14 was the most active, with BACE1 K
i
value of 0.0032 μM and low
Pgp efflux ratio of 2.3. Interestingly, the rigidification strategy led to macrocyclic derivatives that exhibited a great
selectivity over other acyclic derivatives in which compound 14 showed 700fold selectivity over CatD.
68
It is worth to mention that one of the highly effective compounds containing the guanidine scaffold is compound
15: MK8931 (verubecestat/Merck), which is the first BACE1 inhibitor to reach phase III clinical trials. This
compound has remarkable physiochemical properties, including improved oral bioavailability, cellular penetration,
and brain permeability. Results of in vitro studies showed that verubecestat exhibited potent activity against
BACE1 enzyme with an IC
50
of 2.2 nM. This could be explained from its binding modes in which strong hydrogen
bonds between the amidine moiety of verubecestat and the BACE1 catalytic dyad were obvious in its cocrystal
structure. Moreover, verubecestat showed high selectivity for BACE1 when studied on other aspartyl proteases,
with more than 45 000fold selectivity over CatD, among others. Furthermore, the in vivo data support its efficacy
for the management of AD, as the administration of a single oral dose of 10 or 30 mg/kg in rats, significantly
reduced the cerebrospinal fluid (CSF) and cortical Aβ40.
69
Unfortunately, all of these promising findings were
overthrown by the latest results announced in May 2018, which showed the failure of verubecestat to reduce the
cognitive decline in 1958 enrolled patients, along with manifested adverse events.
70
Another important guanidinebased derivative that reached phase III clinical trials is compound 16: E2609
(elenbecestat/Eisai, Biogen). This bicyclic aminodihydrothiazine substituted compound is similar in structure to the
previously mentioned compound 15. It showed a K
i
value of 27 nM and highly reduced the Aβlevels of the CSF and
SCHEME 1 Various design strategies utilized to synthesize a series of potent BACE1 inhibitors based on
isophthalic acid motif. Rigidification, contraction, and bioisostere replacement strategies have been used to develop
inhibitors with improved binding affinity
20
|
MOUSSAPACHA ET AL.
TABLE 3 Nonpeptidomimetic BACE1 inhibitors profile (IC
50
, selectivity, efflux ratio, and bioavailability)
Guanidine and it's analogs BACE1 IC
50
BACE1 Selectivity Pgp
efflux
ratio Bioavailability Ref.Over CatD
Over
BACE2
N
N
H
N
Ph
O
NH
2
Cl
N
H
O
Cl
13
0.21 μM‐‐Low
54
N
N
H
HN
Cl
Cl
O
NH
O
O
H
2
N
14
0.0032 μM 700fold Low
(2.3)
54
S
N
N
N
H
O
N
F
FNH
2
O
O
15
2.2 nM > 45 000 fold Low High
55
K
i
=27nM ‐‐
8
N
S
N
HN
N
O
F
H
2
N
17
‐‐ ‐
8
Amino/iminohydantoinbased Inhibitors
N
Cl
N
N
H
HN
O
18
K
i
=21nM 350 Low Good
57
N
N
H
HN
O
N
19
K
i
= 5.4 nM 7500 Low Excellent
57
Aminothiazoline and aminooxazolinebased
Inhibitors
ON
F
O
NF
N
O
H
2
N
20
0.3 nM ‐‐2.2
58
N
O
HN
O
N
Me
CF
3
NC
F
H
2
N
21 12 nM ‐‐1.9 68%
59
Note: Data is not available.
Abbreviation: BACE1, betasite amyloid precursor protein cleaving enzyme1.
MOUSSAPACHA ET AL.
|
21
plasma in rodents. Another aminoazine derivative is compound 17, known as JNJ54861911 (atabecestat/Janssen).
This aminodihydrothiazinecontaining compound proved to be very successful in reducing Aβlevels, which reached
up to 95% reduction after dosing in healthy volunteers. However, this compound raised safety concerns in phase II/
III clinical trials, which led to the discontinuation of these trials.
8
A promising chloropyrdine analog with
iminohydantoin scaffold (eg, inhibitor 18) was successfully developed. It showed good selectivity for BACE1
(350 fold over CatD) and potent BACE1 inhibitory activity (K
i=
21 nM). Further modifications aimed to improve the
ligandbinding affinity in BACE1 subpockets, which was achieved by replacing the chlorosubstituent with a
propenyl group, in which the linearity of the substituent helped to form stronger interactions with the S3 pocket.
This enhancement in binding affinity was reflected on the potency, which was increased five folds more than that of
a similar chlorinated analog (inhibitor 17) with a K
i
value of 5.4 nM for the newly developed inhibitor 19.
Interestingly, the propenyl substituent improved the selectivity over CatD up to 7500 folds. On the other hand, the
oral administration of compound 19 demonstrated excellent bioavailability, which has been shown to be higher
than that found for compound 18. These modifications have therefore resulted in a more potent compound with a
better pharmacokinetic profile.
71
Other classes of amino azines inhibitors are compounds containing the aminooxazoline and xanthene cores,
which showed a remarkable inhibition potency against BACE1. However, these compounds were found to
exhibit high Pgp efflux ratio.
72
Functionalization of the xanthene core led to compound 20 possessing a
3aza2fluoroxanthene moiety reported by Cheng et al. Inhibitor 20 is a highly effective BACE1 inhibitor that
significantly reduced the brain and CSF Aβlevels in both rats and nonhuman species.
72
Xray costructural studies of
inhibitor 20 bound to BACE1 showed that the 3azaxanthene core formed strong hydrogen bonding interactions
with Trp76. Other hydrogen bonds formed between the oxygen of the dihydropyran and Tyr198. The pyridyl
nitrogen formed an extra hydrogen bond with Ser229 (Figure 8).
72
Another expansion of these amino azines scaffolds was developed by including the substituted aminoxazines,
which showed potent activity against BACE1. For example, compound 21 possesses an IC
50
of 12 nM and cellular
IC
50
of 2 nM.
73
It also showed a dosedependent reduction in Aβ40/42 levels in the brain of mice and revealed a
good oral bioavailability in rats (68%).
73
In Summary, many classes in this family were developed as potent BACE1 inhibitors, from which, compound 15:
MK8931, with an IC
50
value of 2.2 nM, is among the guanidine derivatives, which reached phase III clinical trials
and showed optimum properties including potency, bioavailability, and selectivity. It is noteworthy to mention that
many inhibitors from the aminothiazoline and aminoxazoline family showed potent activities against BACE1 with
IC
50
s in the nanomolar and subnanomolar range. Despite the success in producing potent inhibitors of BACE1 from
this family, the literature lacks detailed data regarding their selectivity profile. Furthermore, it is worthy to note
that the most potent BACE1 inhibitors in this class, which reached clinical trials, were recently revoked due to lack
of clinical efficacy.
8,70
3.2.3 |Naturallyoccurring BACE1 inhibitors
Natural products estate plays an essential role in the discovery of medicinal agents. The last decades have
witnessed the approval of many drugs of natural origin.
74
Moreover, management of complicated diseases such as
AD may require natural materials that contain various natural compounds with targeting multiple proteins rather
than a singletarget synthetic compound.
75
Therefore, phytochemicals enjoy high 3Dcontent and diverse molecular
scaffolds that can serve the clinical needs, besides their safety profile compared to their synthetic counterparts may
offer another advantage.
76
During the development of BACE1 inhibitors of natural products origin, many studies directed the attention
toward naturally occurring polyphenolic compounds, like flavonoids, particularly polymethoxyflavones extracted
from black ginger (Kaempferia parviflora [KP]). KP plant was primarily known to have antioxidant, anticarcinogenic,
and antifatigue properties.
77
However, its extract consists mainly of three major components, namely
22
|
MOUSSAPACHA ET AL.
polymethoxyflavones, 5,7dimethoxyflavone (DMF), 5,7,4′‐trimethoxyflavone (TMF), and 3,5,7,3,4′‐pentamethox-
yflavone (PMF), which showed considerable BACE1 inhibitory activity, with an insignificant effect on αsecretase
and other serine proteases. Youn et al carried out an in silico study utilizing a human BACE1 with nonpeptidyl
polymethoxyflavones, to assess their mode of binding.
78
These results demonstrated the strong ability of KP
polymethoxyflavones to inhibit BACE1 in a dosedependent manner and revealed the Hbond interactions between
the polymethoxyflavones and BACE1. For example, TMF (compound 22) exhibited the most potent BACE1
inhibitory activity with an IC
50
of 36.9 μM, followed by DMF (compound 23) (IC
50 =
49.5 μM) and PMF (compound
24) showed the least potency with an IC
50
of 59.8 μM.
78
To rationalize these findings from the chemical structure
point of view, suggest that BACE1 inhibition resulted from the common methoxyl groups at C5 and C7 in ring A.
Furthermore, a methoxyl group at C4in ring B would enhance the activity, while addition of methoxyl groups at C3
and C3reduced the BACE1 activity (TMF > DMF > PMF). Interestingly, these compounds were reported to be
noncompetitive BACE1 inhibitors, since they interact with the enzyme by binding to the allosteric sites. In contrast,
synthetic peptidomimetic BACE1 inhibitors bind directly to the active site, therefore, they are competitive
inhibitors.
79
A supportive study was done by Mekjaruskul et al in rats, which indicated that DMF, TMF, and PMF
natural products were detected in the rat brain, indicating their ability to cross the BBB.
80
Other triflavonoids derivatives, isolated from Selaginella doederleinii (Selaginellaceae) plant, have been
investigated and showed promising BACE1 inhibitory activity. S. doederleinii is commonly used in Chinese herbal
medicine as antiinflammatory, anticancer, and cardioprotective agent.
81
Zou et al have successfully isolated
several triflalvonoids, for example, selagin triflavonoids 25, which possesses trimeric scaffolds isolated from S.
doederleinii herb. An in vitro study using a fluorescence resonance energy transfer (FRET) technique displayed that
compound 25 was the most potent BACE1 inhibitor with an IC
50
of 0.75 μM, while compound 26 was the weakest
inhibitor (IC
50 =
46.99 μM). Moreover, SAR study suggested that a selagin triflavonoid with a naringenin unit
(compound 25) rather than an apigenin unit (compound 26) is preferred, and have enhanced activity.
81
Another study was performed on Leea indica plant by Hosen et al, where 40 molecules were isolated from
different parts of the plant.
82
L.indica is a huge shrub and its leaves were used in folk medicine as antispasmodic,
anticancer, and antidiarrheal remedy.
83
Virtual screening studies on the isolated compounds have identified two
triterpenes molecules, ursolic acid (compound 27) and lupeol (compound 28) which exhibited higher BACE1 binding
affinity. SAR studies revealed that the binding of these triterpenes in the BACE1 active site is driven by the
formation of hydrogen bonds between ursolic acid and the Asn233 and Thr232 residues, with weak contribution
from hydrophobic interactions, whereas lupeol interacts with Gly11 residue by hydrogen bonding and forms much
stronger hydrophobic interactions with the active site. Furthermore, although both compounds formed
hydrophobic interactions with Tyr71 residue at the flap region, however, none of them was able to interact with
the catalytic aspartic acid residues (Asp32 and Asp228). Theoretical calculations of the PK values have shown that
lupeol could cross the BBB more efficiently than ursolic acid.
82
Moreover, other studies indicated that lupeol have
neuroprotective property.
84
It has also been reported that lupeol inhibits BACE1 with an IC
50
value of 5.12 μM.
82
In another study conducted by Zhu et al,
85
a library of marine natural products with diverse molecular scaffolds
were studied and the results showed antiAD effects with two steroidal extracts isolated from the U.unicinctus.
86
Using the FRET approach, the EC
50
values of these compounds when tested against BACE1, have identified
hecogenin (compound 29) cholest4en3one to possess a value of 116.3 μM, while that of cholest4en3one
(compound 30) was 390.6 μM. In comparison with the more potent synthetic BACE1 inhibitors, the naturally
derived hecogenin and cholest4en3one have lower molecular weights that would allow them to effectively
penetrate the BBB. Furthermore, they are derived from edible marine sources, which gives an indication of their
safety profile.
85
The characteristic features of the mentioned naturally derived products that were extracted from various
plants and organisms are summarized in Table 4. Despite their low activities against BACE1, preliminary findings
suggest that their antiAD effect could offer beneficial advantages over the synthetic leads in terms of their safety
profile, as one of the major drawbacks of the BACE1 inhibitors entered clinical trials. To date, most of the studies
MOUSSAPACHA ET AL.
|
23
TABLE 4 Naturallyderived BACE1 inhibitors profile
Compound Natural Source Chemical classification BACE1 IC
50
Ref.
O
OCH
3
OCH
3
H
3
CO
O
22 (TMF)
Kaempferia parviflora (black ginger) Flavonoids (polymethoxyflavones) 36.9 μM
64
O
OCH
3
H
3
CO
O
23 (DMF)
Kaempferia parviflora (black ginger) Flavonoids (polymethoxyflavones) 49.5 μM
64
O
OCH
3
H
3
CO
O
OCH
3
OCH
3
24 (PMF)
Kaempferia parviflora (black ginger) Flavonoids (polymethoxyflavones) 59.8 μM
64
O
O
O
O
OH
O
OHOH O
OH
OH
OH
O
OH
OH
25 Selaginella doederleinii Flavonoids (triflavonoids) 0.75 μM
67
Selaginella doederleinii Flavonoids (triflavonoids) 46.99 μM
67
HO
H
H
H
O
OH
27 (Ursolic acid)
Leea indica Triterpenes
68
(Continues)
24
|
MOUSSAPACHA ET AL.
TABLE 4 (Continued)
Compound Natural Source Chemical classification BACE1 IC
50
Ref.
HO H
H
H
28 (Lupeol)
Leea indica Triterpenes 5.12 μM
68
HO
H
H
H
O
O
H
H
H
O
H
29 (Hecogenin) Urechis unicinctus Steroids EC
50
= 116.3 μM
71
Urechis unicinctus Steroids EC
50
= 390.6 μM
71
Note: Data is not available.
Abbreviation: BACE1, betasite amyloid precursor protein cleaving enzyme1.
MOUSSAPACHA ET AL.
|
25
conducted on natural compounds were in vitro assays only; therefore, advanced in vivo studies are needed to
address several questions including their mechanism of action and their clinical effect.
4
|
FAILURE OF BACE1 INHIBITORS: LESSONS TO THE FUTURE
For many years now, targeting the amyloid hypothesis to tackle AD has been the cornerstone for many drug
developers. The extensive research has resulted in the development of many potent BACE1 inhibitors; some of
them reached different phases of clinical trials. Having a group of BACE1 inhibitors in advanced clinical stages
was by itself a remarkable achievement, and raised the level of researchersexpectations toward the belief that
a possible solution for AD riddle can be achieved by BACE1 inhibitors which may offer a breakthrough therapy
of this mystery.
42
However, the latest results of some of the BACE1 inhibitors in clinical trials brought
opposing findings that set back the hopes on this target. Among the promising compounds that were developed
by Merck pharmaceutical company and believed to be the major hope in the battle against AD, is the MK8931
(verubecestat), which belongs to the nonpeptidomimetics guanidine family of BACE1 inhibitors. Despite the
remarkable results for its potency (IC
50
= 2.2 nM), selectivity, pharmacokinetic and physiochemical properties,
Merck has recently announced discontinuing phase III clinical trial conducted on 1958 patients with mildto
moderate Alzheimer's disease.
70
The main reason behind this failure is the lack of efficacy, as patients on
verubecestat did not show improvement of their cognitive function compared with the placebo group despite
the ability of verubecestat to reduce Aβlevels in the brain and cerebrospinal fluid.
70
Moreover, another
pertaining cause is the experienced adverse effects associated with verubecestat use, such as futility, rashes,
falls and injuries.
70
Furthermore, it was suggested that the failure of verubecestat to achieve the desired goal
could be attributed to the fact that reducing the level of Aβproduction is of no clinical benefit after the
patients are diagnosed with dementia, as the accumulation of Aβhappens years before symptoms of dementia
appear in AD patients. Another possible explanation for the failure of BACE1 inhibitors in reducing AD
progression might be due to the complexity of the genetic factors responsible for AD disease progression and
the inability of the amyloid hypothesis alone to explain the progression of the disease.
70
One of the famous
examples of BACE1 inhibitors that attracted the attention of many researchers is the failure of JNJ54861911
(atabecestat) to show significant clinical efficacy. This aminodihydrothiazinecontaining derivative was tested
on a group of asymptomatic patients with high risk for AD, in a multicentered trial which enrolled 1650
individuals.
30
Initially, the results were encouraging as atabecestat was capable of achieving a 95% reduction in
Aβproduction after dosing in healthy volunteers. However, on May 17, 2018, it was decided to hold the trial
due to an abnormal elevation in liver enzymes among 600 of the participants.
30
These observations for the
progress of BACE1 inhibitors across different clinical trials clearly indicate that safety was the major barrier
against the success of many BACE1 inhibitors. For instance, two compounds developed by Eli Lilly, LY2811376
and LY2886721 were halted from further progress in clinical trials due to signs of liver injury. Moreover,
BI1181181 developed by Boehringer Ingelheim is a compound that succeeded in reducing Aβbrain levels and
passed twophase I clinical trials. However, once again safety was the main reason that ruled out BI1181181
from the market and forced Boehringer Ingelheim to terminate its development in July 2015. Furthermore, due
to unrevealed causes, AZD3839 from AstraZeneca, and RG7129 from Roche were also terminated in clinical
trials.
30
Henceforth, from the stated examples it is observed that many promising BACE1 inhibitors shared the same
outcome when it comes to clinical application, whereby they failed to reach the soughtafter goal in managing AD.
Therefore, it becomes indispensable to carefully examine what are the factors behind this failure and to seek
answers and solutions to whether the failure lies in utilizing BACE1 as a target for AD or there are missing
elements that need to be considered when designing BACE1 inhibitor. Herein, we report some of the possible
hypothesis in the following sections.
26
|
MOUSSAPACHA ET AL.
4.1
|
Structurebased failure
One of the many factors to procure potent BACE1 inhibition stems from optimizing the interaction
between BACE1 and its ligands; however, this interaction is of unique nature, as it can be described as a
dynamic interaction, whereby several structural features need to be carefully assessed. Hence, taking a
closer insight toward the dynamics of the ligandBACE1 interaction via different tools would be of the
essence and can provide solutions to enhance the design of BACE1 inhibitors while explaining the failure of
the present inhibitors.
For instance, a study by Gueto et al was performed to serve this goal where they have measured the
residueligand interaction energies of 112 amino acid in BACE1 active site using the PM7 semiempirical
technique, with different inhibitors from the hydroxyethylamine family that gave rise to many powerful
BACE1 Inhibitors of favorable pharmacokinetics, yet they did not provide clinical value.
87
The study revealed
that hydroxyethylamine inhibitors have six anchor points when binding to BACE1; Asp93, Asp289, Thr292,
Thr293, Asn294, and Arg296.
87
The binding to these anchor points accounts for 45% of the total BACE1
inhibitor interaction. Thus, designing inhibitors that are capable of forming those vital anchor point
interactions with BACE1 and maintaining this interaction over time is crucial to yield the best possible
inhibitory activity. It was observed that the protonation status of hydroxyethyl amines could be shifted with
time resulting in loss of essential interactions with BACE1 residues shifting the flap confirmation from the
active (closed) back to the open form.
88
Therefore, this shift in the flap confirmation can explain how some
HEAs derivatives would fail to inhibit the BACE1 activity in real time. Furthermore, the quantitative analysis
performed by Geuto et al for the enzyme and the ligandenzyme distances illustrated that the structural
modifications of the ligand head and tail would make a fundamental difference and affect the flap closure. All
of these factors are to be considered when designing a successful HEAs BACE1 inhibitor.
The protonation status of the catalytic dyad residues Asp32 and Asp228 in the presence of the inhibitor can
also be a detrimental factor to a potent BACE1 inhibition, by augmenting and optimizing the ligandenzyme
interaction. It is well perceived that the best activity of the enzyme is achieved in an acidic environment with a pH
of 4.5, which provides the optimal medium for the acidbase interaction between the catalytic dyad residues with
the ligand.
89
However, the protonation status of the dyad while the inhibitor is in place is not well established.
Kocak et al investigated all the possible protonation status of the dyad amino acids in the presence of potent acyl
guanidinebased inhibitor, with the aim to determine the most stable and favorable protonation status.
90
The
results of molecular dynamic simulations showed that when the ligandcontaining NH
2
warhead is protonated (at
pH 4.5), the dideprotonated status of the dyad is the most preferred, and is 14 kcal/mole more stable than all other
presumed options.
90
A similar positive finding was confirmed when the ligand was unprotonated. Thus, one should
consider the protonation status, when designing BACE1 inhibitors, since it might be a pivotal player to stabilize
their interaction with the enzyme.
90
Another important milestone that must be carefully considered when engineering a structure for BACE1
inhibition is the selectivity toward BACE1 over other homologs proteases like BACE2, and CatD as the lack of
specificity can result in significant adverse effects. Despite the structural similarity between, for instance, BACE1
and BACE2 as the two proteases share 64% homology in the amino acid sequence, however, their biological
functions are different.
91
Therefore, it is pivotal to explore the key elements that govern BACE1 selectivity.
Thorough investigations including alignment, molecular dynamic, and docking studies were conducted on four
selective BACE1 inhibitors.
92
These studies showed that the main factor for selectivity is the hindrance that can be
easily fitted and occupied in BACE1 active site but that is not possible in the small catalytic cavity of BACE2,
whereas the large active site of CatD does not offer optimum number of interactions between the ligand and the
enzyme.
92
In addition, the formation of strong electrostatic bonds with Asp32, Asp288 in BACE1 cleft along with
multiple hydrogen bonds, and the Van der Waals interactions, correctly orient the inhibitor in the BACE1 active
site, as it has a different shape when compared to similar proteases.
92
MOUSSAPACHA ET AL.
|
27
4.2
|
The selectivity of BACE1 inhibitors
Tackling the amyloid cascade hypothesis via the inhibition of BACE1,
93
have recently resulted in reports raising
safety concerns and adverse side effects.
70
These alarming effects can be better explained by exploring the
molecular mechanism through which BACE1 processes the amyloidogenic substrate APP. It was reported that
BACE1 is not entirely selective toward targeting solely the amyloidogenic substrate, in fact, it was even noticed
that BACE1 has a greater affinity toward the cleavage of the nonamyloidogenic substrates like neuregulin1
(NRG1).
93
NRG1 has various vital roles in developmental processes, among which is neuronal myelination, in
addition, BACE1 is needed to activate NRG1.
94,95
Hence, BACE1 inhibitor will not only abolish BACE1 role in
processing APP, which is needed to tackle AD, but it will also prevent the cleavage of the nonamyloid substrates
that are required for their function. At the molecular level, it was established that the cleavage of APP by BACE1
happens in an early endosome, where both the substrate and the enzyme undergo endocytosis.
93
On the other
hand, it was reported that BACE1 processing of the nonamyloidogenic substrates is not endocytosisdependent.
93
Therefore, this difference in the subcellular compartmentalization forms a fundamental factor that can be exploited
to create selective BACE1 inhibitors, which target only the endosome containing BACE1 and its amyloidogenic
substrate APP, while leaving BACE1 processing of the nonamyloid substrates undisturbed, to avoid unwanted
adverse effects. In fact, such tactics have been utilized by designing a sterol linked BACE1 inhibitor,
96
where the
sterol linkage can guide the molecule to the lipophilic endosome compartment sparing the free BACE1 from
cleaving NRG1.
97
Sterol linked inhibitor was more efficient in reducing Aβproduction and targeting BACE1 activity
compared with the free inhibitor in cell culture.
96
Moreover, this success was confirmed in vivo by demonstrating
improved safety profile, where treating transgenic Drosophila model with sterol linked BACE1 inhibitor improved
survival rates of the Drosophila larvae compared with control.
96
In addition, the sterol linked inhibitor efficacy was
highlighted in an in vivo study performed on APPs/PSDE9 mice which were injected with either solvent, free
inhibitor, or sterol linked inhibitor to the hippocampus. The sterol linked BACE1 inhibitortreated group showed
the most efficient Aβreduction.
98
Such findings might pave the way toward adopting new strategies to minimize
the side effects associated with BACE1 inhibition.
4.3
|
BACE1 inhibitors and gender variations: differences in brain metabolites
AD occurs more in females than males, where twothirds of patients with AD are women.
99
The exact reason for
this gender variation is not well understood and rather controversial. Reports indicated the decline in the
neuroprotective steroidal hormones like estrogen. However, clinical data in support of its relation to AD remains
obscure.
100
Moreover, wellknown crosstalk between AD genetics and gender might also explain the higher
femalessusceptibility to the disease, where APOEε4 allele attributes to a higher risk for AD in females compared
with males.
101
In addition, pathological brain changes occur at an earlier age in females compared with males during
the period of the disease progression. A study with transgenic APP mutant mice showed that female mice begin to
show a higher rate of Aβaccumulation at an earlier age (69 months) compared with males.
102
In vivo studies
indicated that metabolic profile is also prone to genderspecific regulation, where it is found in APP/PS1 mice
model that the degree of brain metabolites alteration is more significant in females compared with males.
103
Another study conducted by Pan et al,
104
confirmed the genderbased metabolic variation in vivo, on a different
mice strain, where they have used PLB4 mice, a specific strain that has been prone to human BACE1 knock in to
mimic the amyloidogenic AD pathology in humans.
104
In this study, female mice were grouped into two categories
on bases of age, then the PLB4 female mice and the wild type group were prone to metabolic profiling using LCMS/
MS.
104
It was then concluded that the older female PLB4 mice group had the most alteration in their brain
metabolic profile including 21 metabolites among which are; lysophosphatidylcholines, creatinine, sphingomyelin,
and leucine.
104
Thus, this study highlights that overexpression or upregulation of BACE1 interferes with the brain
metabolites regulation in a genderspecific manner, which brings to the forefront the importance and the need to
28
|
MOUSSAPACHA ET AL.
conduct specific analysis on the function and the fate of those altered metabolites upon BACE1 knock in and
inhibition. Indeed, a vital link might be revealed between gender, BACE1, and brain metabolite profiles that can
guide and specify the criteria where BACE1 inhibition would yield the best results.
4.4
|
BACE1inhibitors effects on the synapse structure and functions
The accumulation of Aβplaques in the synapse can result in synaptic impairment and contribute to AD
abnormality.
105
Nonetheless, it was also noted that BACE1 cleavage of APP could induce synaptic dysfunction,
independent from Aβaccumulation.
98
Hence, it is important to observe the effect of APP on synapse function and
how it can be regulated by BACE1 activity. In a study performed by Nigam et al,
106
immunoblotting demonstrated
that reduction in BACE1 activity could result in the accumulation of fulllength APP in the synapse of BACE1/
mice.
106
This intriguing observation offers a possible explanation to the ascending reports on safety concerns and
unwanted side effects arising from the use of BACE1 inhibitors in clinical trials, which resulted in halting such
trials.
106
In addition, a fundamental concept was highlighted in this study, in that the accumulation of fulllength
APP was only induced by complete BACE1 inhibition, whereas, partial inhibition did not have the same effect.
106
Therefore, controlling the dose of BACE1 inhibitor is a determinant player to combat AD successfully. It is
suggested that giving the lowest dose that would be effective in partially reducing BACE1 activity while still
conserving the synaptic function might be the ideal approach. Likewise, it was illustrated by an in vivo study that
managing the dose of BACE1 is crucial to avoid unwanted synaptic damage.
107
According to this report, the
administration of a high dose of BACE1 inhibitor affected synaptic function and plasticity in mice.
107
Nonetheless,
another hypothesis was brought to the forefront to correlate BACE1 inhibition to synaptic dysfunction other than
the impact of BACE1 on APP processing. Zhu et al
108
suggested that BACE1 inhibition induce synaptic impairment
via seizure protein 6 (SEZ6), which is one of the neuronal transmembrane proteins that are selectively cleaved by
BACE1 to produce the soluble SEZ6 and Cterminal fragment.
108
Based on previous studies, it was reported that
knockout of SEZ6 in mice would result in physiological features that mimic those occurring in patients with AD,
such as, memory impairment, and dendritic spinal density reduction.
109
Therefore, Zhu et al tried in their work to
explore whether BACE1 inhibition is causing synaptic dysfunction due to perturbation in SEZ6 function. Thus, it
was demonstrated that prolonged BACE1 inhibition using potent inhibitor like NB360 was only causing reversible
spinal density reduction in wild type mice (control) while such effect was not seen in SEZ6 knockout mice group.
108
Moreover, synaptic impairment caused by BACE1 inhibition was prevented by knocking out SEZ6 from matured
neurons.
108
The results of the Zhu group have therefore confirmed that indeed SEZ6 is associated with synaptic
structural and functional impairment induced by BACE1 inhibition, and this change is reversible and dose
related.
108
Since synaptic side effects are found to be dose dependent, this raised the question of whether the
dose reduction of BACE1 inhibitors would be sufficient to manage AD. In fact, according to a study conducted on
BACE1 + /mice, 40% reduction in Aβplaques was achieved with 50% inhibition of BACE1 without causing
synaptic adverse effects.
110
Thus, these preliminary findings suggest that partial BACE1 inhibition would still be
effective on AD patients who did not reach advanced stages with Aβplaques saturation.
5
|
ALTERNATIVE INTERVENTIONS TO PREVENT AβACCUMULATION
BY IMMUNOTHERAPY
As mentioned previously, AD is a neurodegenerative disorder that is characterized by the accumulation of
misfolded protein aggregates (Aβdeposits) in which they eventually lead to synaptic toxicity and neuronal loss.
111
Therefore, the mainstream in the development of AD treatment is directed against the Aβpeptide. One of the most
attractive and currently tested antiAβapproaches is immunotherapy, particularly, passive immunotherapy which
involves the direct administration of external antibodies. Both active and passive immunization resulted in
MOUSSAPACHA ET AL.
|
29
increased Aβdeposits clearance in humans and AD transgenic mice, but unfortunately, several agents have failed to
improve cognitive functions and resulted in serious adverse events.
112
Moreover, initial clinical trials of active
vaccination have reported cases of meningoencephalitis and studies were therefore terminated,
113
whereas
preclinical trials of passive immunization have displayed promising results.
114
Consequently, the safety concerns of
active immunization have shifted the attention toward devoting efforts to passive immunotherapy.
112
The approach of targeting Aβthrough passive immunotherapy attracted increasing attention. This strategy is
based on a mechanism called peripheral sink hypothesis, where the exogenous antibodies recognize, clear, and
reduce the concentration of Aβin the periphery through capturing these secreted Aβmolecules in blood
circulation. Thus, disrupting the Aβconcentration between the CNS and plasma and this concentration gradient
was found to promote the release of Aβfrom the brain.
115
In addition, these therapeutic antibodies have different
epitopes and vary in their detection and binding affinity to several Aβspecies (monomers, oligomers, protofibrils,
and fibrils).
116
AD was the cornerstone of immunotherapy research over the past 15 years.
111
Different monoclonal
antibodies (mAbs) have been designed and engineered to recognize different Aβconformations, and introduced
into clinical trials. However, the results from testing therapeutic mAbs were mystifying, since some have failed
and showed a lack of efficacy and have many side effects. Others have boosted the experience in improving and
developing better candidates,
112
beside others that showed enhanced Aβclearance.
117
According to many
reports, the stage of initiating the intervention in clinical trials is very critical, since targeting Aβin early stages of
AD or prophylactically may exhibit better impact than in late stages of AD.
118
Table 5, summarizes many
of these engineered antibodies including bapineuzumab, solanezumab, ponezumab, gantenerumab, aducanumab,
and crenezumab.
Bapineuzumab (AAB001, Pfizer/Janssen Pharmaceuticals) is an immunoglobulin (Ig) G1 Aβtargeting mAb,
developed from the humanization of murine mAb 3D6. Bapineuzumab stabilized by hydrogen bonds with five
residues at the Nterminus of Aβ, and it is selectively recognized and clear both fibrillary and soluble Aβspecies.
112
It is the first mAbs that enters clinical trials after the discontinuation of AN1792 (the first active immunotherapy
for AD) trial.
113
Treatment with different doses of bapineuzumab was tested in participants with mild to moderate
AD to asses PK, safety, and tolerability profile. Based on these observations, bapineuzumab successfully reached
phase III clinical trial.
119
However, all bapineuzumab studies after concluding the results were terminated in August
2012,
112
due to the evidence of low efficacy and increased the occurrence of magnetic resonance imaging (MRI)
abnormalities, known as amyloidrelated imaging abnormalities attributed to edema or effusion (ARIAE) with high
doses of bapineuzumab and APOEε4gene carriers.
120,121
Another firstgeneration Aβpassive immunotherapy is solanezumab (LY2062430, Eli Lilly), an IgG1 mAb that is
humanized from the parent murine antibody m266. Solanezumab targets the residues 16 to 26 at the middle region
of Aβpeptide and has the ability to recognize the soluble monomers but not the fibrillary Aβspecies. Preclinical
studies in transgenic PDAPP mice (overexpress human amyloid precursor protein V717F) proved the ability of mAb
m266 to increase the clearance of soluble Aβspecies without binding the amyloid plaques.
112
Subsequently,
solanezumab exhibited an excellent safety profile in phase I and II trials, but unfortunately, no improvement in
cognitive functions was observed.
111
In phase III study, some clinical improvement resulted after months of
treatment in mild patients with AD, however, solanezumab failed to meet the primary outcomes and studies have
recently discontinued.
122
Nonetheless, the good safety profile of solanezumab was the main courage to currently
run prevention trials.
123
Ponezumab (PF04360365, Pfizer) also belongs to the firstgeneration agents, a humanized IgG2 antibody that
binds the residues 30 to 40 at the Cterminus of Aβ40. Comparing it with the other IgG1 mAbs, IgG2 have lower
tendency to stimulate the immune effector function.
124
Ponezumab entered phase I studies and displayed evidence
of safety and tolerability in mild to moderate patients with AD, without signs of ARIA.
112
Yet, the lack of clinical
efficacy was the reason to terminate the studies after phase II trial.
125
Consequently, studies over second generation antiAβmAbs have taken place due to the disappointing
observations from the firstgeneration mAbs. Gantenerumab (RG1450/RO4909832, HoffmanLa Roche), is the first
30
|
MOUSSAPACHA ET AL.
TABLE 5 AntiAβmonoclonal antibodies in clinical trials for AD
Antibody name/code mAb type Epitope Aβspecies Clinical status Company
Refer-
ence
Bapineuzumab (AAB001) Humanized IgG1 Soluble and fibrillar forms Nterminus (residues 15) Phase III
terminated
Pfizer/Janssen
105
Solanezumab (LY2062430) Humanized IgG1 Soluble monomers Mid domain (residues 1626) Phase III
terminated
Eli Lilly
108
Ponezumab (PF04360365) Humanized IgG2 Soluble and aggregated
forms
Cteminus (residues 3040) Phase IIa
terminated
Pfizer
111
Gantenerumab (RG1450,
RO4909832)
Human IgG1 Fibrillar forms Nterminus (residues 312) and mid
domain (residues 1827)
Phase III ongoing HoffmannLa
Roche
114
Aducanumab (BIIB037) Human IgG1 Soluble oligomers and
insoluble fibrils
Nterminus (residues 29) Phase III ongoing Biogen Idec
115
Crenezumab (MABT5102A) Humanized IgG4 Monomers, oligomers, fibrils Mid domain (residues 1324) Phase III ongoing Genentech
118
Abbreviation: AD, Alzheimer's disease.
MOUSSAPACHA ET AL.
|
31
fully human IgG1 antibody that targets the fibrillary Aβform and uniquely binds the epitope that involves both, the
Nterminal (312) and midregion (1827) residues. Hence, a folded peptide where the mid domain overlaps with
the Nterminus is preferred.
112
Preclinical studies in transgenic mice confirmed that gantenerumab extensively
reduced Aβdeposits through mediated microglial phagocytosis and blocking the formation of new plaques, without
affecting the systemic Aβlevels.
126
Gantenerumab underwent phase I and II clinical trials in patients with
prodromal AD and was relatively safe, but small group experienced signs of ARIA in a dose and APOEε4 phenotype
dependent manner.
127
Up to date, a phase III randomized trial of gantenerumab in prodromal AD is still ongoing.
128
Another fully human IgG1 mAb is aducanumab (BIIB037, Biogen), that selectively binds the soluble Aβ
aggregates (oligomers) and insoluble fibrils. It targets the residues 2 to 9 at the Nterminus of Aβpeptide.
123
Preclinical studies in transgenic mice demonstrated the ability of aducanumab analog to cross the BBB, recognize
its Aβtarget and increase the clearance of Aβplaques.
129
Accordingly, aducanumab completed earlier phases of
clinical trials and is currently in phase III studies to measure its efficacy in slowing AD progression and improving
memory loss.
129
Lastly, Crenezumab (MABT5102A, Genentech) is distinctly derived from an IgG4 backbone, thus stimulation of
Fc gamma receptor is diminished. It preferably binds the middleregions of various conformations of Aβat the
residues 13 to 24, however, the affinity for capturing oligomers is 10 times higher than for monomers.
130
Moreover, crenezumab and solanezumab recognize the same epitope, although they have a dissimilar binding
pattern. A random coil structure resulted from the binding of crenezumab to Aβbetween residues 21 and 24, while
solanezumabAβ(residues 2126) complex produces an alphahelical structure.
131
Currently, crenezumab is in
phase III trial for prodromal to patients with mild AD.
132
6
|
FUTURE DIRECTIONS FOR ALZHEIMER'S DISEASE TREATMENT
Alzheimer's disease (AD) is the most devastating neurodegenerative disorder and its pathogeneses are manifested
by many cellular and molecular damage mechanisms, which is attributed to multiple genetic factors. Consequently,
this necessitates the need for multitarget drugs, which might be critical in halting the progression of this complex
disease.
133
Besides, the failure of several clinical trials targeting BACE1 and other targets necessitate the need for
developing novel therapeutic agents with new mechanisms independent of the amyloid hypothesis.
134
Therefore, in
the following sections, important strategies followed to tackle the progression of this devastating disease will be
briefly discussed.
6.1
|
Inhibitors of receptorinteracting serine/threonineprotein kinase 1
One of the theories to tackle AD is to prohibit the conversion of microglia to the pathogenic phenotype, which
causes the progression, and development of a series of neurodegenerative disorders including AD.
135
It was
observed that microglia could play a dual role, initially; it helps in clearing out the cellular toxins like the βamyloid
plaques and hyperphosphorylated tau protein. However, when the microglia become chronically inflamed it will
start acting in an opposite manner by expelling large quantities of toxins causing neurons damage and death.
135
Henceforth, the idea was to suppress the inflammation of microglia by inhibiting receptorinteracting serine/
threonineprotein kinase 1 (RIPK1). Several studies supported this hypothesis, by showing that the pharmacological
and genetic RIPK1 inhibition results in the reduction of inflammatory mediators and amyloid burden, and improved
memory function (Figure 9). Moreover, these studies approved the role of microglia in promoting the degradation
of Aβin vitro.
136
One of the early compounds discovered for this purpose is necrostatin1(31), (Nec1; a wellknown RIPIK1
inhibitor (K
D=
3.1 nM)) that has been used in a number of studies to examine RIPK1 function (Figure 10). This
compound was discovered by phenotypic screening for necrotic cell death inhibitors. SAR studies showed that
32
|
MOUSSAPACHA ET AL.
necrostatin1 places the activation segment in the kinase domain of the receptor in the inactive unphosphorylated
confirmation by allosteric inhibition.
137
The compound comprises two main motifs, indoleamine, and thiohydantoin.
Structural alterations in the thiohydantoin group drastically affect the activity. The removal of the thiohydantoin
methyl abolished the inhibitory activity of the compound. On the other hand, converting the thiohydantoin to
hydantoin enhanced the activity.
138
Selectivity of RIPK1 inhibitors is the main concern as the biological effect can
be attributed to binding to other targets since necrostatin1 can bind to two other kinase receptors.
139
Moreover,
necrostatin1 showed paradoxical finding in terms of the appropriate dosing for administration in disease models of
systemic inflammatory response syndrome (SRIS) which suggests another barrier for its utilization in disease
management.
139
Perhaps these barriers explain the lack of any application of this compound in AD management.
DNL747 is another example of RIPK1 inhibitors that reached phase I clinical trials for the treatment of AD,
marking it as the firstinhuman utilization of RIPK1 inhibitors.
140
Besides the inhibition of microglia inflammation using RIPK1 inhibitors as a strategy to tackle AD, recent work
by Pluvinage et al revealed the presence of Bcell receptor CD22 on aged microglial cells that promote the
antiphagocytic effect of microglia. Therefore, as an alternative approach, the inhibition of CD22 could switch on the
phagocytic activity of glial cells to clear out the neurotoxins and protein aggregates that are known to be
responsible for AD hallmarks.
141
6.2
|
Inhibitors of glycogen synthase kinase 3β
Glycogen synthase kinase3 (GSK3) is an enzyme that has been validated as a target for many diseases due to its
vital role and involvement in multiple biological processes including cell division, apoptosis, and insulin production.
GSK3 has three isozymes, one of them is glycogen synthase kinase3β(GSK3β) which is highly abundant in the
brain and is considered one of the driving factors for AD.
142,143
Malfunction of GSK3βwas found to be the reason
for the hyperphosphorylation and accumulation of tau protein. In addition, it mediates the production of Aβ, thus
contributing to the development of AD hallmarks.
144
This isozyme is therefore considered one of the attractive
targets on the battle against AD cure and many small molecules were designed to inhibit this protein. These
inhibitors can be grouped into different classes encompassing: irreversible inhibitors, allosteric inhibitors, peptide
like inhibitors, metalion inhibitors, adenosine triphosphate (ATP) competitive inhibitors, and nonATP competitive
inhibitors.
145
However, inhibiting GSK3 features many challenges, both GSK3 isoforms GSK3αand GSK3βbear
very similar ATP binding sites, therefore it is quite difficult to attain selective inhibitor that can differentiate
between both isoforms. Furthermore, to selectively regulate the activity of GSK3β, the inhibitor must interact
specifically with its catalytic triad, Arg96, Arg180, and Lys205. Most of the known GSK3βinhibitors have in
common, low molecular weight and flat heterocyclic structure.
146
Many of the GSK3βpotent inhibitors were
designed to compete with ATP active site and were derived from pyrrolopyridinone scaffold. Many of these
inhibitors possess IC50s in the nanomolar range and were capable of reducing tau protein phosphorylation,
e,g, pyridyl pyridine derivative, compound 32, Figure 11, with an IC
50
of 4.4 nM.
147
On the other hand, the nonATP competitive inhibitors have demonstrated success against GSK3β. In this
context, a series of inhibitors was developed having a thiadiazolidinone moiety as the core scaffold. For example,
compound 33 (ie, tideglusib) belongs to this class and is currently in phase II clinical trials for the treatment of mild
to moderate AD.
148,149
These type of ligands are known to act by preventing the indigenous substrate from
accessing the proper orientation of the enzyme binding site.
150
Another class of inhibitors is the irreversible GSK3βinhibitors that are characterized by having the αhalo
methyl group, which is known to bind to Cys199 in the ATP binding site causing an alteration in the active
conformation. In fact, the αcarbonyl thienyl (compound 34), and its phenyl derivatives (compound 35) showed
potent inhibitory activity and are now in phase II clinical trials.
145
Other important approaches to inhibit GSK3βbased on multitarget ligands (MLTs) strategy were considered.
Among others is the dual inhibition of BACE1 and glycogen synthase kinase 3β(GSK3β). These motifs contain a
MOUSSAPACHA ET AL.
|
33
triazinone scaffold as a suitable ligand to concurrently bind to the aspartic acid residues of BACE1 binding site as
well as the ATP site of GSK3β.
151
Perhaps the challenging aspects of GSK3 inhibitors design can be circumvented by utilizing other strategies to
block this enzyme. For example, insulin has been extensively examined as one of the suggested novel approaches to
manage AD,
152
where alterations in brain insulin metabolism were considered as one of the underlying causative
factors for this disease.
152
Insulin and GSK3 enzyme both regulate glycogen metabolism but in opposite ways.
GSK3 is a kinase that keeps glycogen synthase inactive via phosphorylation, whereas insulin tends to activate
glycogen synthetase by dephosphorylation on the same sites.
153
Therefore, in that manner insulin can act to block
GSK3 thereby tackling AD. In fact, intranasal insulin is currently in phase III for AD (NCT01767909).
153
In summary, GSK3βinhibitors can provide a hopeful approach to tackle AD, since some of the agents in this
class are still in clinical development.
154
Nonetheless, particular concerns raised since these compounds might
induce hypoglycemia and tumorigenesis.
154
Moreover, the need to achieve a proper balance between lipophilicity
to cross BBB and hydrophilicity for oral absorption posed a major concern in their development.
154
Currently,
GSK3βnoncompetitive inhibitors appear to be the most promising and safest candidates for clinical use due to
their selectivity and potency.
154
6.3
|
FYN kinase inhibition
Identification of specific signaling pathways responsible for Aβproduction enabled the discovery of promising
interventions that target AD hallmarks.
155
An elegant finding described in 1995 suggested that a soluble rather
than aggregated Aβshowed increased toxicity in neuronal cell cultures.
156
Additional work on the synthesis of Aβ,
identified that the cause of the synaptotoxic material in neuronal cultures was due to Aβoligomers (Aβo), which are
soluble Aβaggregates varying in size from dimers to large molecular weight species.
24,155
The Aβo was found to
interact with the cellular prion protein (PrPC) on the surface of the neuron activating a downstream signaling
pathway converging on the nonreceptor tyrosine kinase Fyn that triggers cellular damage.
24,157
Thus, one of the
proposed therapeutic options that moved beyond the ongoing efforts to remove the soluble assemblies of Aβ(Aβo)
is Fyn inhibition. Fyn is one of the Src family of nonreceptor tyrosine kinases (SFKs), which also includes Src, Lck,
Hck, Blk, Lyn, Fgr, Yes, and Yrk.
158
Pertaining to this hypothesis, transgenic mice models had shown reduced
cognitive functions and enhanced AD phenotype when the function of Fyn increased. However, ablation of Fyn
function has ameliorated AD phenotype.
159
The suggested pathological signaling cascade via SFKs is summarized in
Figure 12, where the extracellular binding domain of Aβo activates the gatekeeper PrPC and Fyn plays a crucial
role in this cascade.
160
Moreover, Fyn has been uniquely linked to the two major pathological hallmarks of AD, because it is not only
activating by Aβvia PrPC, but also it interacts with tau.
155
It has been reported that Fyn is involved in Tau
phosphorylation, which results in synapse damage, behavioral deficits and electroencephalographic deformities in
APP transgenic mice.
161
Thus, many groups attempted to design inhibitors of Fyn protein. Compound 36 (Figure 13), saracatinib
(AZD0530), is one of the orally bioavailable SFK inhibitors with high potency for Src and Fyn, that is currently in an
ongoing phase IIa multicenter study for potential treatment of AD.
159
It inhibits Fyn at the subnanomolar
concentration (IC
50
=810 nM) with excellent pharmacokinetic properties including brain exposure.
159,162
Compound 37, dasatinib (BMS354825), is another selective and potent SFK inhibitor with a Fyn IC
50
of 0.2 nM.
162
Furthermore, it has revealed an improvement in the cognition function in transgenic AD mice.
Although, compound 37 might be a promising drug candidate for AD, however, there are no current studies
on patients with AD.
155
It is worth to mention that Fyn is a challenging target with significant homology with other members of SFKs. In
addition, Fyn is implicated in a wide variety of physiological processes that are essential for normal activity, that is,
34
|
MOUSSAPACHA ET AL.
the role of Fyn in synaptic function and plasticity. Therefore Fyn inhibition might lead to the development of
unintended adverse effects.
163
6.4
|
Cannabinoid type 2 receptor agonists
Endocannabinoid system (ECS) is an essential network of lipid particles and receptors, which contributes to the
regulatory processes of many physiological functions throughout the human body. The two main endocannabinoid
receptors are the cannabinoid type 1 (CB1) and cannabinoid type 2 (CB2) receptors, which belong to the family of
Gproteincoupled receptors.
164
CB1 receptors are placed within the central nervous system (CNS) and in
peripheral tissues.
165
They regulate several important brain functions like emotion, memory, cognition, and pain
perception, via the alteration of the excitatory and inhibitory neurotransmission, whereas CB2 receptors are
expressed outside the CNS and can modulate the immune system.
166
Moreover, recent findings have proved the
presence of CB2 receptors in other tissues, that is, CNS.
167
During the last decade, an accumulated body of evidence demonstrated that targeting CB2 receptors might
serve as a promising therapeutic approach for Alzheimer's and other neurodegenerative diseases, as they are
present in dendritic cells, neuronal cells, and microglia.
168
Furthermore, the stimulation of CB2 receptors is
generallyknowntohaveanantiinflammatory effect or may enhance tissue damage in some conditions.
169
It
should be mentioned that the expression levels of CB2 receptors are dictated by Aβ42 levels and plaque
accumulation, suggesting the stimulation of CB2 receptor expression by these pathogenic events.
170
Thus, the
strong induction of CB2 receptors in the diseased microglia increases the therapeutic benefits, as it would
facilitate selective activation in damaged tissues.
166
Activated microglia generates cytokines, such as TNFα,
and inflammatory mediators, along with neurons and astrocytes during neuroinflammation in most of
the neurodegenerative diseases including AD.
171
However, data suggested that Aβaccumulation, and
neuroinflammation and inflammatory mediators release usually coexist, however, the causes are not well
understood.
172
CB2 receptors on the surface of microglia cells inhibit microgliamediated neurotoxicity, besides
their indirect role in the regulation of Aβlevels in the brain and the enhancement of Aβclearance.
166
Several
studies have confirmed the antiinflammatory effects and the modulation of Aβlevels by CB2 agonists in the in
vivo transgenic mice and in vitro experiments. In this context, the discovery of the selective agonist for CB2,
compound 38 (Figure 14): JWH133, with K
i
of 677 nM for CB1 and K
i
of 3.4 nM for CB2,
173
has shown to
improve the cognitive performance, reduce the microglial activation, and response to Aβ, and decrease the
proinflammatory cytokines in APP/PS1 mice.
174
Another CB1/CB2 receptor agonist, compound 39: WIN55,2122, showed an increase in the memory function
of Aβinduced hippocampal neurodegeneration in adult rats,
175
as well as, a reduction in the release of
proinflammatory cytokines in microglia culture disclosed to the toxic Aβpeptide.
176
Furthermore, compounds 38
and 39, both facilitate the microglial migration, that in turn, induces the phagocytosis of aggregated Aβ.
176
Other studies reported the therapeutic effect of the Δ9tetrahydrocannabinol (THC), compound 40,
177
a major
ingredient of Cannabis sativa, in preventing the hallmark characteristics of AD.
178
THC is one of the most potent
CB1 agonists, with an EC
50
less than 50 nM.
179
It enhances Aβdegradation and reduces the intraneuronal Aβ
aggregation. Further research and investigations are still needed to pursue the use of cannabinoids to tackle
dementia in a clinical trial setting. Evidence from preclinical studies suggested that there is a potential for such a
route to provide effective management of AD by targeting the ECS.
180
Therefore, drugs repositioning that is used for other neurodegenerative diseases might give way to surprising
outcome for the management of AD, such as the case of the FDA approved riluzole for amyotrophic lateral sclerosis
(ALS),
181
that is currently under investigation for halting the progression of AD, and the drug is now in phase II
clinical trial entitled: T2 Protect AD(NCT03605667). The awaited results of this investigation might offer
additional insight for the potential treatment of AD by utilizing glutamate modulators like riluzole.
181
MOUSSAPACHA ET AL.
|
35
6.5
|
Taudirected potential therapy
One of the AD hallmarks is the neurofibrillary tangles produced as a result of hyperphosphorylated tau protein.
182
Therefore, AD is always correlated with the accumulation of hyperphosphorylated tau proteins that assemble
giving rise to filaments and oligomers which then induce neuronal and glial cells degradation.
182
Consequently, this
accumulation of tau proteins can act as a driver for neurodegenerative conditions known as tauopathies,and one
of the AD consequences is due to this phenomena.
183
Therefore, seeking therapeutic alternatives for AD that
target tau proteins has recently captured researchersinterest. In this context, there are many ongoing
investigations targeting the accumulation of tau proteins, among others, is the immunotherapy approach that
indicated promising results reported about the clinical trials progress of many antibodies and vaccines for the
potential treatment of AD.
184,185
In addition, numerous preclinical studies were performed to reduce tau production by targeting tau gene expression
via the antisense oligonucleotidebased approach. The latter is currently in phase I clinical trials.
186
However, it is worth
mentioning that attempts to reduce tau production bear some consequences as tau proteins are key molecules in
numerous neuronal functions such as microtubules assembly and stability, and axonal transport. Furthermore, it has been
reportedthattaudeletioninmiceresultedinbraininsulin resistance, iron accumulation, and cognitive defects.
187,188
Another strategy in targeting tau accumulation is the antiaggregation approach that targets tau proteins phosphorylation
and oligomerization. This strategy utilizes inhibitors of GSK3βor activators of phosphatase enzyme that dephosphorylate
tau proteins.
189
Perhaps one of the success stories of the antiaggregation strategy is the discovery of compound LMTM
(TRx0237) that advanced to phase III clinical trials.
190
At this junction, the complexity of AD pathogenesis, led scientists to adopt the multitargetdirected ligands (MTDLs)
approach, in which more than one AD causative path is targeted. In this regard, many examples employing this route were
reported, among others is the design of dual GSK3βand acetylcholinesterase inhibitors.
191
In addition, MTDLs might
ameliorate the performance of BACE1 inhibitors and synergize the potential of the AD treatment. Henceforth, many
researchers have dedicated their efforts toward the design of ligands that inhibit BACE1 along with other targets like
GSK3β.
192
Another example is dual BACE1/acetylcholinesterase (AChE) inhibitors represented by the rheinhuprine
hybrid compounds reported by Viayna et al.
193
Despite the fact that the idea of MTDLs is rather captivating and
promising, however, to date none of the MTDLs advanced to clinical trials stages. This might be due to the challenges
encountered when utilizing such a strategy due to the complexity in the design of drugs that modulate the function of
more than one target as well as the complications reported from the in vivo studies.
133
7
|
CONCLUSIONS
Understanding Alzheimer's disease (AD) pathogenesis is rather challenging since the exact biological machinery
that causes AD is not well understood. The current AD treatment options provide only symptomatic management
of the disease and do not reverse the disease progression or the associated neuronal damage.
194
Henceforth,
substantial efforts were made to design and develop new therapeutic agents to halt the progression and manage
the disease. Many of these agents were developed based on the amyloid cascade hypothesis, which offered a
wealth of promising compounds that target the downstream events of Alzheimer disease. Among the most
important proteins is the βsecretase (BACE1), which is involved in the ratelimiting step of the βamyloid protein
accumulation, therefore, the discovery of BACE1 inhibitors attracted the utmost attention from researchers
worldwide.
13
Synthetic BACE1 inhibitors belong to two major classes, either peptidomimetics or non
peptidomimetics. The comparison between these two classes would highly favor the nonpeptidomimetics because
peptidomimeticbased inhibitors suffered from large size, poor oral bioavailability, short halflife, weak metabolic
stability, and low BBB penetration.
52
In addition, libraries of inhibitors were isolated from natural sources such as
plants and organisms and showed relatively good BACE1 inhibitory activity.
36
|
MOUSSAPACHA ET AL.
Despite the tremendous efforts and the significant expenditure on the development of BACE1 inhibitors that
reached different phases in clinical trials; unfortunately, the latest results reported the failure of many of these
inhibitors, while others are still in ongoing studies and under the investigation with a high possibility to fail as well.
There are several obstacles that must be overcome to allow the design of effective and selective BACE1 inhibitors
to pass advanced clinical trials.
31
Safety concerns have led to the termination of many BACE1 inhibitors in clinical
trials, which have been shown to suffer from offtarget activity, offsite toxicity or lack of significant physiological
effect in humans, for instance, LY2811376 and LY2886721 from Eli Lilly, AZD3839 from AstraZeneca, and
RG7129 from Roche were terminated during clinical trials.
30
Perhaps the failure of these agents could be
attributed to other reasons as AD is a multifactorial abnormality where various hypotheses contributed to its
understanding. Administration of BACE1 inhibitors by patients whom their abnormality was not induced by the
amyloid hypothesis alone will not deliver a clinically efficient therapeutic option for the success of BACE1
inhibitors. Furthermore, the high failure rate of BACE1 inhibitors might be attributed to the nature of BACE1 active
site, which is relatively large and structurally similar to many other aspartyl protease enzymes distributed in
different parts of the human body; therefore, having a small molecule, for example, BACE1 inhibitor, to selectively
occupy such a relatively large active site represents a major challenge.
31
The fact that the structures of all BACE1
inhibitors that were developed so far are flat molecules with high aromatic and SP
2
contents, might explain the
reason behind the lack of selectivity which is responsible for the observed offtarget activities and toxicity effects
of these inhibitors. This necessitates the need for the design of diverse molecular libraries with high 3Dcontent to
crosstalk specifically with the high 3Dcontent complementary in BACE1. This conclusion is supported by several
examples of successful drug discoveries, which indicated that increasing the SP
3
contents of a ligand to better suit
the complementary binding regions of an active site correlates favorably with selectivity and the success rate in
clinical trials.
195-197
Moreover, recent reports illustrated that BACE1 inhibitors caused structural and functional
synaptic impairment that accounts for safety concerns.
106
Additionally, emerging evidence is exposing gender as a
factor to be considered when designing therapeutic option for AD,
198
which requires further investigations to
determine how this element can better shape AD therapy.
104
Since none of the BACE1 inhibitors succeeded so far to offer a clinical benefit that slow or reverse the
progression of AD, this might lead to the conclusion that BACE1 inhibition is a despairing approach to tackle AD.
The failure of many clinical trials, which were based on the amyloid hypothesis, encouraged researchers to consider
new therapeutic agents with mechanisms independent of this hypothesis. Among the new approaches that have
been attempted are the RIPK1, GSK3β, Fyn kinase inhibition, and CB2 receptor agonists. Moreover, extensive
efforts were directed toward antiAβimmunotherapy, particularly, the passive immunotherapy that involves
the use of exogenous mAbs. Several engineered mAbs were developed, tested in animals and humans and reached
different stages in clinical trials. Some have failed the clinical trials while others are still in an ongoing study.
112
The continuous search for optimum routes and techniques with significant physiological effects to halt and/or
reverse the progression of AD is underway and the near future may carry hopes for patients suffering from this
mysterious disease. A solution for this complicated disease might exist within the new proposed strategies to unpin
this puzzle since the tremendous efforts spent on developing BACE1 inhibitors did not so far show clinical value.
Lastly, with the fourth industrial revolution is exponentially accelerating toward solving many unmet diseases
facing the mankind using multidimensional approaches, perhaps, the next breakthrough of the 21 century would be
the discovery of a multibillion $ drug that ends the mystery of AD.
ACKNOWLEDGEMENTS
This study was supported by generous grants from the Research Funding Department, University of Sharjah, UAE
(15011101007 and 15011101002). We are thankful to Dania Al Ramahi, and Sandi Yaakoub for their help in
gathering some literature data.
MOUSSAPACHA ET AL.
|
37
CONFLICT OF INTERESTS
The authors declare that they have no competing interests.
ORCID
Hany A. Omar http://orcid.org/0000-0002-4670-8149
Hasan Alniss http://orcid.org/0000-0001-8639-9531
Taleb H. AlTel http://orcid.org/0000-0003-4914-9677
REFERENCES
1. Wolfe MS. Secretase targets for Alzheimer's disease: identification and therapeutic potential. J Med Chem.
2001;44(13):20392060.
2. Selkoe DJ. Alzheimer's diseasegenotypes, phenotype, and treatments. Science. 1997;275(5300):630631.
3. Ghosh AK, Osswald HL. BACE1 (betasecretase) inhibitors for the treatment of Alzheimer's disease. Chem Soc Rev.
2014;43(19):67656813.
4. Foraker J, Millard SP, Leong L, et al. The APOE gene is differentially methylated in Alzheimer's disease. J Alzheimers
Dis. 2015;48(3):745755.
5. Alzheimer's Association. Alzheimer's disease facts and figures. Alzheimers Dement. 2018;14(3):367429. 2018.
6. Prince M, Guerchet M, Prina M. The epidemiology and impact of dementia: current state and future trends. Geneva: World
Health Organization; 2015.
7. Mullard A. BACE inhibitor bust in Alzheimer trial. Nat Rev Drug Discov. 2017;16(3):155155.
8. Piton M, Hirtz C, Desmetz C, et al. Alzheimer's disease: advances in drug development. JAlzheimersDis. 2018;65(1):313.
9. Haapasalo A, Hiltunen M. A report from the 8th Kuopio Alzheimer symposium. Neurodegener Dis Manag.
2018;8(5):289299.
10. Cummings J, Lee G, Ritter A, et al. Alzheimer's disease drug development pipeline: 2018. Alzheimers Dement (N Y).
2018;4:195214.
11. Abner EL, Neltner JH, Jicha GA, et al. Diffuse amyloidbeta plaques, neurofibrillary tangles, and the impact of APOE in
elderly persons' brains lacking neuritic amyloid plaques. J Alzheimers Dis. 2018;64(4):13071324.
12. Paasila PJ, Davies DS, Kril JJ, et al. The relationship between the morphological subtypes of microglia and Alzheimer's
disease neuropathology. Brain Pathol. 2019. https://doi.org/10.1111/bpa.12717
13. Vassar R, Kuhn PH, Haass C, et al. Function, therapeutic potential and cell biology of BACE proteases: current status
and future prospects. J Neurochem. 2014;130(1):428.
14. Cole SL, Vassar R. The Alzheimer's disease betasecretase enzyme, BACE1. Mol Neurodegener. 2007;2:22.
15. Ohno M. Alzheimer's therapy targeting the betasecretase enzyme BACE1: Benefits and potential limitations from
the perspective of animal model studies. Brain Res Bull. 2016;126(Pt 2):183198.
16. Ashford JW. The dichotomy of Alzheimer's disease pathology: amyloidbeta and Tau. JAlzheimersDis. 2019;68(1):7783.
17. DeFelice FG, Vieira MN, Saraiva LM, et al. Targeting the neurotoxic species in Alzheimer's disease: inhibitors of Abeta
oligomerization. FASEB J. 2004;18(12):13661372.
18. Ferreira ST, Lourenco MV, Oliveira MM, et al. Soluble amyloidbeta oligomers as synaptotoxins leading to cognitive
impairment in Alzheimer's disease. Front Cell Neurosci. 2015;9:191.
19. Stéphan A, Laroche S, Davis S. Generation of aggregated βamyloid in the rat hippocampus impairs synaptic
transmission and plasticity and causes memory deficits. J Neurosci. 2001;21(15):57035714.
20. Näslund J, Haroutunian V, Mohs R, et al. Correlation between elevated levels of amyloid βpeptide in the brain and
cognitive decline. JAMA. 2000;283(12):15711577.
21. He Y, Zheng MM, Ma Y, et al. Soluble oligomers and fibrillar species of amyloid betapeptide differentially
affect cognitive functions and hippocampal inflammatory response. Biochem Biophys Res Commun. 2012;429
(34):125130.
22. Dobrowolska Zakaria JA, Vassar RJ. A promising, novel, and unique BACE1 inhibitor emerges in the quest to prevent
Alzheimer's disease. EMBO Mol Med. 2018;10(11):e9717.
23. Reiss AB, Arain HA, Stecker MM, et al. Amyloid toxicity in Alzheimer's disease. Rev Neurosci. 2018;29(6):613627.
24. Lauren J, Gimbel DA, Nygaard HB, et al. Cellular prion protein mediates impairment of synaptic plasticity by amyloid
beta oligomers. Nature. 2009;457(7233):11281132.
38
|
MOUSSAPACHA ET AL.
25. Hardy J. The discovery of Alzheimercausing mutations in the APP gene and the formulation of the "amyloid cascade
hypothesis". FEBS J. 2017;284(7):10401044.
26. Bourgeois A, Lauritzen I, Lorivel T, et al. Intraneuronal accumulation of C99 contributes to synaptic
alterations, apathylike behavior, and spatial learning deficits in 3xTgAD and 2xTgAD mice. Neurobiol Aging. 2018;
71:2131.
27. Lauritzen I, PardossiPiquard R, Bourgeois A, et al. Intraneuronal aggregation of the betaCTF fragment of APP (C99)
induces Abetaindependent lysosomalautophagic pathology. Acta Neuropathol. 2016;132(2):257276.
28. Lauritzen I, PardossiPiquard R, Bourgeois A, et al. Does intraneuronal accumulation of carboxyl terminal fragments of the
amyloid precursor protein trigger early neurotoxicity in Alzheimer's disease. Curr Alzheimer Res. 2019;16:453457.
29. Nixon RA. Amyloid precursor protein and endosomallysosomal dysfunction in Alzheimer's disease: inseparable
partners in a multifactorial disease. FASEB J. 2017;31(7):27292743.
30. Yan R. Stepping closer to treating Alzheimer's disease patients with BACE1 inhibitor drugs. Transl Neurodegener.
2016;5(1):13.
31. Citron M. Emerging Alzheimer's disease therapies: inhibition of βsecretase. Neurobiol Aging. 2002;23(6):10171022.
32. Haass C. Take fiveBACE and the gammasecretase quartet conduct Alzheimer's amyloid betapeptide generation.
EMBO J. 2004;23(3):483488.
33. Das B, Yan R. Role of BACE1 in Alzheimer's synaptic function. Transl Neurodegener. 2017;6:23.
34. Neumann U, Rueeger H, Machauer R, et al. A novel BACE inhibitor NB360 shows a superior pharmacological
profile and robust reduction of amyloidbeta and neuroinflammation in APP transgenic mice. Mol Neurodegener.
2015;10:44.
35. Keskin AD, Kekus M, Adelsberger H, et al. BACE inhibitiondependent repair of Alzheimer's pathophysiology. Proc
Natl Acad Sci U S A. 2017;114(32):86318636.
36. Volloch V, Rits S. Results of Beta SecretaseInhibitor Clinical Trials Support Amyloid Precursor ProteinIndependent
Generation of Beta Amyloid in Sporadic Alzheimers Disease. Med Sci. 2018;6(45), https://doi.org/10.3390/
medsci6020045
37. Moussa CE. Betasecretase inhibitors in phase I and phase II clinical trials for Alzheimer's disease. Expert Opin Investig
Drugs. 2017;26(10):11311136.
38. Himmelstein DS, Ward SM, Lancia JK, et al. Tau as a therapeutic target in neurodegenerative disease. Pharmacol Ther.
2012;136(1):822.
39. AlTel TH, Semreen MH, AlQawasmeh RA, et al. Design, synthesis, and qualitative structureactivity evaluations of
novel betasecretase inhibitors as potential Alzheimer's drug leads. J Med Chem. 2011;54(24):83738385.
40. Coimbra JRM, Marques DFF, Baptista SJ, et al. Highlights in BACE1 inhibitors for Alzheimer's disease treatment.
Front Chem. 2018;6:178.
41. Yuan J, Venkatraman S, Zheng Y, et al. Structurebased design of betasite APP cleaving enzyme 1 (BACE1) inhibitors
for the treatment of Alzheimer's disease. J Med Chem. 2013;56(11):41564180.
42. Vassar R. BACE1 inhibitor drugs in clinical trials for Alzheimer's disease. Alzheimers Res Ther. 2014;6(9):89.
43. Hu B, Xiong B, Qiu BY, et al. Construction of a small peptide library related to inhibitor OM992 and its structure
activity relationship to betasecretase. Acta Pharmacol Sin. 2006;27(12):15861593.
44. Hong L, Koelsch G, Lin X, et al. Structure of the protease domain of memapsin 2 (betasecretase) complexed with
inhibitor. Science. 2000;290(5489):150153.
45. Mirsafian H, Mat Ripen A, Merican AF, et al. Amino acid sequence and structural comparison of BACE1 and BACE2
using evolutionary trace method. Sci World J. 2014;2014:482463482466.
46. Rombouts FJR, Alexander R, Cleiren E, et al. Fragment binding to betasecretase 1 without catalytic aspartate
interactions identified via orthogonal screening approaches. ACS Omega. 2017;2(2):685697.
47. Xu Y, Li MJ, Greenblatt H, et al. Flexibility of the flap in the active site of BACE1 as revealed by crystal structures and
molecular dynamics simulations. Acta Crystallogr D Biol Crystallogr. 2012;68(Pt 1):1325.
48. Barman A, Prabhakar R. Computational insights into substrate and site specificities, catalytic mechanism, and
protonation states of the catalytic Asp Dyad of beta secretase. Scientifica (Cairo). 2014;2014:598728.
49. Patel S, Vuillard L, Cleasby A, et al. Apo and inhibitor complex structures of BACE (betasecretase). J Mol Biol.
2004;343(2):407416.
50. Menting KW, Claassen JA. betasecretase inhibitor; a promising novel therapeutic drug in Alzheimer's disease. Front
Aging Neurosci. 2014;6:165.
51. Dash C, Kulkarni A, Dunn B, et al. Aspartic peptidase inhibitors: implications in drug development. Crit Rev Biochem
Mol Biol. 2003;38(2):89119.
52. Luo X, Yan R. Inhibition of BACE1 for therapeutic usein Alzheimer's disease. Int J Clin Exp Pathol. 2010;3(6):618628.
MOUSSAPACHA ET AL.
|
39
53. Ghosh AK, Kumaragurubaran N, Hong L, et al. Design, synthesis and Xray structure of proteinligand
complexes: important insight into selectivity of memapsin 2 (betasecretase) inhibitors. J Am Chem Soc. 2006;
128(16):53105311.
54. Ghosh AK, Kumaragurubaran N, Hong L, et al. Design, synthesis, and Xray structure of potent memapsin 2
(betasecretase) inhibitors with isophthalamide derivatives as the P2P3ligands. J Med Chem. 2007;50(10):
23992407.
55. Sandgren V, Back M, Kvarnstrom I, et al. Design and synthesis of hydroxyethylenebased BACE1 inhibitors
incorporating extended P1 substituents. Open Med Chem J. 2013;7:115.
56. Chang WP, Huang X, Downs D, et al. Betasecretase inhibitor GRL8234 rescues agerelated cognitive decline in APP
transgenic mice. FASEB J. 2011;25(2):775784.
57. Ghosh AK, Kumaragurubaran N, Hong L, et al. Potent memapsin 2 (betasecretase) inhibitors: design, synthesis,
proteinligand Xray structure, and in vivo evaluation. Bioorg Med Chem Lett. 2008;18(3):10311036.
58. Beswick P, Charrier N, Clarke B, et al. BACE1 inhibitors part 3: identification of hydroxy ethylamines (HEAs) with
nanomolar potency in cells. Bioorg Med Chem Lett. 2008;18(3):10221026.
59. Rueeger H, Lueoend R, Machauer R, et al. Discovery of cyclic sulfoxide hydroxyethylamines as potent and selective
betasite APPcleaving enzyme 1 (BACE1) inhibitors: structure based design and in vivo reduction of amyloid beta
peptides. Bioorg Med Chem Lett. 2013;23(19):53005306.
60. Stachel SJ, Coburn CA, Steele TG, et al. Structurebased design of potent and selective cellpermeable inhibitors of
human betasecretase (BACE1). J Med Chem. 2004;47(26):64476450.
61. Kortum SW, Benson TE, Bienkowski MJ, et al. Potent and selective isophthalamide S2 hydroxyethylamine inhibitors
of BACE1. Bioorg Med Chem Lett. 2007;17(12):33783383.
62. RazzaghiAsl N, Ebadi A, Edraki N, et al. Ab initio modeling of a potent isophthalamidebased BACE1 inhibitor: amino
acid decomposition analysis. Med Chem Res. 2012;22(7):32593269.
63. AlTel TH, AlQawasmeh RA, Schmidt MF, et al. Rational design and synthesis of potent dibenzazepine motifs as beta
secretase inhibitors. J Med Chem. 2009;52(20):64846488.
64. Tarazi H, Odeh RA, AlQawasmeh R, et al. Design, synthesis and SAR analysis of potent BACE1 inhibitors: Possible
lead drug candidates for Alzheimer's disease. Eur J Med Chem. 2017;125:12131224.
65. Bjorklund C, Oscarson S, Benkestock K, et al. Design and synthesis of potent and selective BACE1 inhibitors. J Med
Chem. 2010;53(4):14581464.
66. Ghosh AK, Cárdenas EL, Osswald HL. The design, development, and evaluation of BACE1 inhibitors for the treatment
of Alzheimer's disease. In: Wolfe M., ed. Alzheimer's Disease II Topics in Medicinal Chemistry24. Berlin: Springer;
2016:2785.
67. Ghosh AK, Gemma S, Tang J. betaSecretase as a therapeutic target for Alzheimer's disease. Neurotherapeutics.
2008;5(3):399408.
68. Boy KM, Guernon JM, Wu YJ, et al. Macrocyclic prolinyl acyl guanidines as inhibitors of betasecretase (BACE). Bioorg
Med Chem Lett. 2015;25(22):50405047.
69. Kennedy ME, Stamford AW, Chen X, et al. The BACE1 inhibitor verubecestat (MK8931) reduces CNS betaamyloid
in animal models and in Alzheimer's disease patients. Sci Transl Med. 2016;8(363):363ra150.
70. Egan MF, Kost J, Tariot PN, et al. Randomized trial of verubecestat for mildtomoderate Alzheimer's disease. N Engl J
Med. 2018;378(18):16911703.
71. Cumming JN, Smith EM, Wang L, et al. Structure based design of iminohydantoin BACE1 inhibitors: identification of
an orally available, centrally active BACE1 inhibitor. Bioorg Med Chem Lett. 2012;22(7):24442449.
72. Cheng Y, Brown J, Judd TC, et al. An orally available BACE1 inhibitor that affords robust CNS abeta reduction
without cardiovascular liabilities. ACS Med Chem Lett. 2015;6(2):210215.
73. Hilpert H, Guba W, Woltering TJ, et al. betaSecretase (BACE1) inhibitors with high in vivo efficacy suitable for
clinical evaluation in Alzheimer's disease. J Med Chem. 2013;56(10):39803995.
74. Newman DJ, Cragg GM. Natural products as sources of new Drugs from 1981 to 2014. JNatProd. 2016;79(3):629661.
75. Koynova R, Tenchov B. Natural product formulations for the prevention and treatment of Alzheimer's disease: a
patent review. Recent Pat Drug Deliv Formul. 2018;12(1):2339.
76. Hussain G, Rasul A, Anwar H, et al. Role of plant derived alkaloids and their mechanism in neurodegenerative
disorders. Int J Biol Sci. 2018;14(3):341357.
77. Kobayashi S, Kato T, Azuma T, et al. Antiallergenic activity of polymethoxyflavones from Kaempferia parviflora. J
Funct Foods. 2015;13:100107.
78. Youn K, Lee J, Ho CT, et al. Discovery of polymethoxyflavones from black ginger ( Kaempferia parviflora) as potential
βsecretase (BACE1) inhibitors. J Funct Foods. 2016;20:567574.
79. Williams P, Sorribas A, Liang Z. New methods to explore marine resources for Alzheimer's therapeutics. Curr
Alzheimer Res. 2010;7(3):210213.
40
|
MOUSSAPACHA ET AL.
80. Mekjaruskul C, Jay M, Sripanidkulchai B. Modulatory effects of Kaempferia parviflora extract on mouse hepatic
cytochrome P450 enzymes. J Ethnopharmacol. 2012;141(3):831839.
81. Zou Z, Xu P, Zhang G, et al. Selagintriflavonoids with BACE1 inhibitory activity from the fern Selaginella doederleinii.
Phytochemistry. 2017;134:114121.
82. Hosen SMZ, Rubayed M, Dash R, et al. Prospecting and structural insight into the binding of novel plantderived
molecules of leea indica as inhibitors of BACE1. Curr Pharm Des. 2018;24(33):39723979.
83. Wong YH, Abdul Kadir H, Ling SK. Bioassayguided isolation of cytotoxic cycloartane triterpenoid glycosides from the
traditionally used medicinal plant Leea indica.Evid Based Complement Alternat Med. 2012;2012:164689.
84. Kaundal M, Akhtar M, Deshmukh R. Lupeol isolated from Betula alnoides ameliorates amyloid beta induced neuronal
damage via targeting various pathological events and alteration in neurotransmitter levels in rat's brain. Neurol
Neurosci. 2017;08(03):195.
85. Zhu YZ, Liu JW, Wang X, et al. AntiBACE1 and antimicrobial activities of steroidal compounds isolated from marine
Urechis unicinctus.Mar Drugs. 2018;16(3):94.
86. Schinke C, Martins T, Queiroz SCN, et al. Antibacterial compounds from marine bacteria, 20102015. J Nat Prod.
2017;80(4):12151228.
87. GuetoTettay C, MartinezConsuegra A, Zuchniarz J, et al. A PM7 dynamic residueligand interactions energy
landscape of the BACE1 inhibitory pathway by hydroxyethylamine compounds. Part I: the flap closure process. J Mol
Graph Model. 2017;76:274288.
88. GuetoTettay C, Drosos JC, VivasReyes R. Quantum mechanics study of the hydroxyethylaminesBACE1 active site
interaction energies. J Comput Aided Mol Des. 2011;25(6):583597.
89. Shimizu H, Tosaki A, Kaneko K, et al. Crystal structure of an active form of BACE1, an enzyme responsible for
amyloid beta protein production. Mol Cell Biol. 2008;28(11):36633671.
90. Kocak A, Erol I, Yildiz M, et al. Computational insights into the protonation states of catalytic dyad in BACE1acyl
guanidine based inhibitor complex. J Mol Graph Model. 2016;70:226235.
91. AbdulHay SO, Sahara T, McBride M, et al. Identification of BACE2 as an avid ssamyloiddegrading protease. Mol
Neurodegener. 2012;7(1):46.
92. HernandezRodriguez M, CorreaBasurto J, Gutierrez A, et al. Asp32 and Asp228 determine the selective
inhibition of BACE1 as shown by docking and molecular dynamics simulations. Eur J Med Chem. 2016;124:
11421154.
93. Ben Halima S, Mishra S, Raja KMP, et al. Specific inhibition of betasecretase processing of the alzheimer disease
amyloid precursor protein. Cell Rep. 2016;14(9):21272141.
94. Hu X, Hicks CW, He W, et al. Bace1 modulates myelination in the central and peripheral nervous system. Nat
Neurosci. 2006;9(12):15201525.
95. Fleck D, Garratt AN, Haass C, et al. BACE1 dependent neuregulin processing: review. Curr Alzheimer Res.
2012;9(2):178183.
96. Rajendran L, Schneider A, Schlechtingen G, et al. Efficient inhibition of the Alzheimer's disease betasecretase by
membrane targeting. Science. 2008;320(5875):520523.
97. Ehehalt R, Keller P, Haass C, et al. Amyloidogenic processing of the Alzheimer betaamyloid precursor protein
depends on lipid rafts. J Cell Biol. 2003;160(1):113123.
98. Gouras GK, Tampellini D, Takahashi RH, et al. Intraneuronal betaamyloid accumulation and synapse pathology in
Alzheimer's disease. Acta Neuropathol. 2010;119(5):523541.
99. Brookmeyer R, Evans DA, Hebert L, et al. National estimates of the prevalence of Alzheimer's disease in the United
States. Alzheimers Dement. 2011;7(1):6173.
100. Grimm A, Lim YA, MensahNyagan AG, et al. Alzheimer's disease, oestrogen and mitochondria: an ambiguous
relationship. Mol Neurobiol. 2012;46(1):151160.
101. Altmann A, Tian L, Henderson VW, et al. Sex modifies the APOErelated risk of developing Alzheimer disease. Ann
Neurol. 2014;75(4):563573.
102. Zhao L, Mao Z, Woody SK, et al. Sex differences in metabolic aging of the brain: insights into female susceptibility to
Alzheimer's disease. Neurobiol Aging. 2016;42:6979.
103. Trushina E, Nemutlu E, Zhang S, et al. Defects in mitochondrial dynamics and metabolomic signatures of evolving
energetic stress in mouse models of familial Alzheimer's disease. PLoS One. 2012;7(2):e32737.
104. Pan X, Green BD. Temporal effects of neuronspecific betasecretase 1 (BACE1) knockin on the mouse brain
metabolome: implications for Alzheimer's disease. Neuroscience. 2019;397:138146.
105. Haass C, Kaether C, Thinakaran G, et al. Trafficking and proteolytic processing of APP. Cold Spring Harb Perspect Med.
2012;2(5):a006270a006270.
106. Nigam SM, Xu S, Ackermann F, et al. Endogenous APP accumulates in synapses after BACE1 inhibition. Neurosci Res.
2016;109:915.
MOUSSAPACHA ET AL.
|
41
107. Filser S, Ovsepian SV, Masana M, et al. Pharmacological inhibition of BACE1 impairs synaptic plasticity and cognitive
functions. Biol Psychiatry. 2015;77(8):729739.
108. Zhu K, Xiang X, Filser S, et al. Betasite amyloid precursor protein cleaving enzyme 1 inhibition impairs synaptic
plasticity via seizure protein 6. Biol Psychiatry. 2018;83(5):428437.
109. Gunnersen JM, Kim MH, Fuller SJ, et al. Sez6 proteins affect dendritic arborization patterns and excitability of
cortical pyramidal neurons. Neuron. 2007;56(4):621639.
110. Devi L, Ohno M. Effects of BACE1 haploinsufficiency on APP processing and Abeta concentrations in male and female
5XFAD Alzheimer mice at different disease stages. Neuroscience. 2015;307:128137.
111. MontoliuGaya L, Villegas S. Immunotherapy for neurodegenerative diseases: the Alzheimer's disease paradigm. Curr
Opin Chem Eng. 2018;19:5967.
112. vanDyck CH. Antiamyloidbeta monoclonal antibodies for Alzheimer's disease: pitfalls and promise. Biol Psychiatry.
2018;83(4):311319.
113. Gilman S, Koller M, Black RS, et al. Clinical effects of Abeta immunization (AN1792) in patients with AD in an
interrupted trial. Neurology. 2005;64(9):15531562.
114. Bard F, Cannon C, Barbour R, et al. Peripherally administered antibodies against amyloid betapeptide
enter the central nervous system and reduce pathology in a mouse model of Alzheimer disease. Nat Med. 2000;
6(8):916919.
115. Zhang Y, Lee DH. Sink hypothesis and therapeutic strategies for attenuating Abeta levels. Neuroscientist.
2011;17(2):163173.
116. Cehlar O, Skrabana R, Revajova V, et al. Structural aspects of Alzheimer's disease immunotherapy targeted against
amyloidbeta peptide. Bratisl Lek Listy. 2018;119(4):201204.
117. Crane A, Brubaker WD, Johansson JU, et al. Peripheral complement interactions with amyloid beta
peptide in Alzheimer's disease: 2. Relationship to amyloid beta immunotherapy. Alzheimers Dement. 2018;14(2):
243252.
118. Sigurdsson EM. Immunotherapy targeting pathological tau protein in Alzheimer's disease and related tauopathies. J
Alzheimers Dis. 2008;15(2):157168.
119. Sumner IL, Edwards RA, Asuni AA, et al. Antibody engineering for optimized immunotherapy in Alzheimer's disease.
Front Neurosci. 2018;12:254.
120. Lu M, Brashear HR. Pharmacokinetics, pharmacodynamics, and safety of subcutaneous bapineuzumab: a single
ascendingdose study in patients with mild to moderate Alzheimer disease. Clin Pharmacol Drug Dev. 2019;8(3):326335.
121. Vandenberghe R, Rinne JO, Boada M, et al. Bapineuzumab for mild to moderate Alzheimer's disease in two global,
randomized, phase 3 trials. Alzheimers Res Ther. 2016;8(1):18.
122. Honig LS, Vellas B, Woodward M, et al. Trial of solanezumab for mild dementia due to Alzheimer's disease. N Engl J
Med. 2018;378(4):321330.
123. Dolan PJ, Zago W. Passive immunotherapy in Alzheimer's disease. In: Dorszewska Jolanta, Kozubski Wojciech, eds.
Alzheimer's Disease The 21st Century Challenge. London: IntechOpen; 2018:130151.
124. Landen JW, Zhao Q, Cohen S, et al. Safety and pharmacology of a single intravenous dose of ponezumab in subjects
with mildtomoderate Alzheimer disease: a phase I, randomized, placebocontrolled, doubleblind, doseescalation
study. Clin Neuropharmacol. 2013;36(1):1423.
125. Landen JW, Andreasen N, Cronenberger CL, et al. Ponezumab in mildtomoderate Alzheimer's disease: randomized
phase II PETPIB study. Alzheimers Dement (N Y). 2017;3(3):393401.
126. Bohrmann B, Baumann K, Benz J, et al. Gantenerumab: a novel human antiAbeta antibody demonstrates sustained
cerebral amyloidbeta binding and elicits cellmediated removal of human amyloidbeta. J Alzheimers Dis.
2012;28(1):4969.
127. Ostrowitzki S, Deptula D, Thurfjell L, et al. Mechanism of amyloid removal in patients with Alzheimer disease treated
with gantenerumab. Arch Neurol. 2012;69(2):198207.
128. Ostrowitzki S, Lasser RA, Dorflinger E, et al. A phase III randomized trial of gantenerumab in prodromal Alzheimer's
disease. Alzheimers Res Ther. 2017;9(1):95.
129. Sevigny J, Chiao P, Bussiere T, et al. The antibody aducanumab reduces Abeta plaques in Alzheimer's disease. Nature.
2016;537(7618):5056.
130. Adolfsson O, Pihlgren M, Toni N, et al. An effectorreduced antibetaamyloid (Abeta) antibody with unique
abeta binding properties promotes neuroprotection and glial engulfment of Abeta. J Neurosci. 2012;32(28):
96779689.
131. Ultsch M, Li B, Maurer T, et al. Structure of crenezumab complex with abeta shows loss of betahairpin. Sci Rep.
2016;6:39374.
132. Lin H, Ostrowitzki S, Sink KM, et al. Baseline characterics from a phase 3 trial of crenezumab in prodromal to mild
Alzheimer's disease (Cread). Alzheimers Dement. 2018;14(7):P217.
42
|
MOUSSAPACHA ET AL.
133. Prati F, Bottegoni G, Bolognesi ML, et al. BACE1 inhibitors: from recent singletarget molecules to multitarget
compounds for Alzheimer's disease. J Med Chem. 2018;61(3):619637.
134. Kametani F, Hasegawa M. Reconsideration of amyloid hypothesis and tau hypothesis in Alzheimer's disease. Front
Neurosci. 2018;12:25.
135. Mullard A. Microgliatargeted candidates push the Alzheimer drug envelope. Nat Rev Drug Discov. 2018;17(5):303305.
136. Ofengeim D, Mazzitelli S, Ito Y, et al. RIPK1 mediates a diseaseassociated microglial response in Alzheimer's disease.
Proc Natl Acad Sci U S A. 2017;114(41):E8788E8797.
137. Degterev A, Hitomi J, Germscheid M, et al. Identification of RIP1 kinase as a specific cellular target of necrostatins.
Nat Chem Biol. 2008;4(5):313321.
138. Teng X, Degterev A, Jagtap P, et al. Structureactivity relationship study of novel necroptosis inhibitors. Bioorg Med
Chem Lett. 2005;15(22):50395044.
139. Takahashi N, Duprez L, Grootjans S, et al. Necrostatin1 analogues: critical issues on the specificity, activity and in
vivo use in experimental disease models. Cell Death Dis. 2012;3(11):e437e437.
140. Mullard A. BACE failures lower AD expectations, again. Nat Rev Drug Discov. 2018;17(6):385385.
141. Pluvinage JV, Haney MS, Smith BAH, et al. CD22 blockade restores homeostatic microglial phagocytosis in ageing
brains. Nature. 2019;568(7751):187192.
142. Hussein RM, Mohamed WR, Omar HA. A neuroprotective role of kaempferol against chlorpyrifosinduced
oxidative stress and memory deficits in rats via GSK3betaNrf2 signaling pathway. Pestic Biochem Physiol.
2018;152:2937.
143. Bhat RV, Budd Haeberlein SL, Avila J. Glycogen synthase kinase 3: a drug target for CNS therapies. J Neurochem.
2004;89(6):13131317.
144. Hooper C, Killick R, Lovestone S. The GSK3 hypothesis of Alzheimer's disease. J Neurochem. 2008;104(6):14331439.
145. Maqbool M, Mobashir M, Hoda N. Pivotal role of glycogen synthase kinase3: A therapeutic target for Alzheimer's
disease. Eur J Med Chem. 2016;107:6381.
146. Babu AP, Chitti S, Rajesh B, et al. In silico based ligand design and docking studies of GSK3βinhibitors. ChemBio
Inform J. 2010;10:112.
147. Sivaprakasam P, Han X, Civiello RL, et al. Discovery of new acylaminopyridines as GSK3 inhibitors by a structure
guided indepth exploration of chemical space around a pyrrolopyridinone core. Bioorg Med Chem Lett.
2015;25(9):18561863.
148. Lao K, Ji N, Zhang X, et al. Drug development for Alzheimer's disease: review. J Drug Target. 2018;27:110.
149. Mathuram TL, Reece LM, Cherian KM. GSK3 Inhibitors: a doubleedged sword? an update on tideglusib. Drug Res
(Stuttg). 2018;68(8):436443.
150. Martinez A, Alonso M, Castro A, et al. First nonATP competitive glycogen synthase kinase 3 β(GSK3β) inhibitors:
thiadiazolidinones (TDZD) as potential drugs for the treatment of Alzheimer's disease. J Med Chem. 2002;45(6):12921299.
151. Rampa A, Gobbi S, Concetta Di Martino RM, et al. Dual BACE1/GSK3beta inhibitors to combat Alzheimer's disease:
a focused review. Curr Top Med Chem. 2017;17(31):33613369.
152. Morris JK, Burns JM. Insulin: an emerging treatment for Alzheimer's disease dementia? Curr Neurol Neurosci Rep.
2012;12(5):520527.
153. Patel S, Doble B, Woodgett JR. Glycogen synthase kinase3 in insulin and Wnt signalling: a doubleedged sword?
Biochem Soc Trans. 2004;32(Pt 5):803808.
154. EldarFinkelman H, Martinez A. GSK3 inhibitors: preclinical and clinical focus on CNS. Front Mol Neurosci. 2011;4:32.
155. Nygaard HB. Targeting Fyn kinase in Alzheimer's disease. Biol Psychiatry. 2018;83(4):369376.
156. Oda T, Wals P, Osterburg HH, et al. Clusterin (apoJ) alters the aggregation of amyloid βpeptide (Aβ142) and Forms
Slowly Sedimenting Aβcomplexes that cause oxidative stress. Exp Neurol. 1995;136(1):2231.
157. Um JW, Nygaard HB, Heiss JK, et al. Alzheimer amyloidbeta oligomer bound to postsynaptic prion protein activates
Fyn to impair neurons. Nat Neurosci. 2012;15(9):12271235.
158. Boggon TJ, Eck MJ. Structure and regulation of Src family kinases. Oncogene. 2004;23(48):79187927.
159. Kaufman AC, Salazar SV, Haas LT, et al. Fyn inhibition rescues established memory and synapse loss in Alzheimer
mice. Ann Neurol. 2015;77(6):953971.
160. Ochs K, MalagaTrillo E. Common themes in PrP signaling: the Src remains the same. Front Cell Dev Biol. 2014;2:63.
161. Roberson ED, Halabisky B, Yoo JW, et al. Amyloidbeta/Fyninduced synaptic, network, and cognitive impairments
depend on tau levels in multiple mouse models of Alzheimer's disease. J Neurosci. 2011;31(2):700711.
162. Schenone S, Brullo C, Musumeci F, et al. Fyn kinase in brain diseases and cancer: the search for inhibitors. Curr Med
Chem. 2011;18(19):29212942.
163. Nygaard HB, vanDyck CH, Strittmatter SM. Fyn kinase inhibition as a novel therapy for Alzheimer's disease.
Alzheimers Res Ther. 2014;6(1):8.
MOUSSAPACHA ET AL.
|
43
164. Navarro G, BorrotoEscuela D, Angelats E, et al. Receptorheteromer mediated regulation of endocannabinoid
signaling in activated microglia. Role of CB1 and CB2 receptors and relevance for Alzheimer's disease and levodopa
induced dyskinesia. Brain Behav Immun. 2018;67:139151.
165. Hu SS, Mackie K. Distribution of the endocannabinoid system in the central nervous system. Handb Exp Pharmacol.
2015;231:5993.
166. Aso E, Ferrer I. CB2 cannabinoid receptor as potential target against Alzheimer's disease. Front Neurosci. 2016;10:243.
167. Atwood BK, Mackie K. CB2: a cannabinoid receptor with an identity crisis. Br J Pharmacol. 2010;160(3):467479.
168. Del Cerro P, Alquezar C, Bartolome F, et al. Activation of the cannabinoid type 2 receptor by a novel indazole
derivative normalizes the survival pattern of lymphoblasts from patients with lateonset Alzheimer's disease. CNS
Drugs. 2018;32(6):579591.
169. Pacher P, Mechoulam R. Is lipid signaling through cannabinoid 2 receptors part of a protective system? Prog Lipid Res.
2011;50(2):193211.
170. Solas M, Francis PT, Franco R, et al. CB2 receptor and amyloid pathology in frontal cortex of Alzheimer's disease
patients. Neurobiol Aging. 2013;34(3):805808.
171. Wes PD, Sayed FA, Bard F, et al. Targeting microglia for the treatment of Alzheimer's disease. GLIA.
2016;64(10):17101732.
172. Schmole AC, Lundt R, Toporowski G, et al. Cannabinoid receptor 2deficiency ameliorates disease symptoms in a
mouse model with Alzheimer's diseaselike pathology. J Alzheimers Dis. 2018;64(2):379392.
173. Huffman JW, Bushell SM, Miller JRA, et al. 1Methoxy,1deoxy11hydroxyand 11Hydroxy1methoxyΔ8
tetrahydrocannabinols: new selective ligands for the CB2 receptor. Bioorg Med Chem. 2002;10(12):41194129.
174. Aso E, Juves S, Maldonado R, et al. CB2 cannabinoid receptor agonist ameliorates Alzheimerlike phenotype in
AbetaPP/PS1 mice. J Alzheimers Dis. 2013;35(4):847858.
175. Fakhfouri G, Ahmadiani A, Rahimian R, et al. WIN552122 attenuates amyloidbetainduced neuroinflamma-
tion in rats through activation of cannabinoid receptors and PPARgamma pathway. Neuropharmacology.
2012;63(4):653666.
176. MartinMoreno AM, Reigada D, Ramirez BG, et al. Cannabidiol and other cannabinoids reduce microglial activation in
vitro and in vivo: relevance to Alzheimer's disease. Mol Pharmacol. 2011;79(6):964973.
177. Lotsch J, WeyerMenkhoff I, Tegeder I. Current evidence of cannabinoidbased analgesia obtained in preclinical and
human experimental settings. Eur J Pain. 2018;22(3):471484.
178. Cao C, Li Y, Liu H, et al. The potential therapeutic effects of THC on Alzheimer's disease. J Alzheimers Dis.
2014;42(3):973984.
179. Currais A, Quehenberger O, MA A, et al. Amyloid proteotoxicity initiates an inflammatory response blocked by
cannabinoids. NPJ Aging Mech Dis. 2016;2:16012.
180. Sherman C, Ruthirakuhan M, Vieira D, et al. Cannabinoids for the treatment of neuropsychiatric symptoms, pain and
weight loss in dementia. Curr Opin Psychiatry. 2018;31(2):140146.
181. Hunsberger HC, Weitzner DS, Rudy CC, et al. Riluzole rescues glutamate alterations, cognitive deficits, and tau
pathology associated with P301L tau expression. J Neurochem. 2015;135(2):381394.
182. Ferrer I, LopezGonzalez I, Carmona M, et al. Glial and neuronal tau pathology in tauopathies: characterization of
diseasespecific phenotypes and tau pathology progression. J Neuropathol Exp Neurol. 2014;73(1):8197.
183. Spillantini MG, Goedert M. Tau pathology and neurodegeneration. Lancet Neurol. 2013;12(6):609622.
184. Congdon EE, Sigurdsson EM. Tautargeting therapies for Alzheimer disease. Nat Rev Neurol. 2018;14(7):399415.
185. Jadhav S, Avila J, Scholl M, et al. A walk through tau therapeutic strategies. Acta Neuropathol Commun. 2019;7(1):22.
186. Mignon L, Kordasiewicz H, Lane R, et al. Design of the FirstinHuman Study of IONISMAPTRx, a Taulowering
antisense oligonucleotide, in patients with Alzheimer disease (S2.006). Neurology. 2018;90(15 Suppl):S2.006.
187. Lei P, Ayton S, Finkelstein DI, et al. Tau deficiency induces parkinsonism with dementia by impairing APPmediated
iron export. Nat Med. 2012;18(2):291295.
188. Marciniak E, Leboucher A, Caron E, et al. Tau deletion promotes brain insulin resistance. J Exp Med.
2017;214(8):22572269.
189. Li C, Gotz J. Taubased therapies in neurodegeneration: opportunities and challenges. Nat Rev Drug Discov.
2017;16(12):863883.
190. Schwab K, Frahm S, Horsley D, et al. A protein aggregation inhibitor, leucomethylthioninium bis
(hydromethanesulfonate), decreases alphasynuclein inclusions in a transgenic mouse model of synucleino-
pathy. Front Mol Neurosci. 2017;10:447.
191. Jiang XY, Chen TK, Zhou JT, et al. Dual GSK3beta/AChE inhibitors as a new strategy for multitargeting anti
Alzheimer's disease drug discovery. ACS Med Chem Lett. 2018;9(3):171176.
192. Prati F, DeSimone A, Armirotti A, et al. 3,4Dihydro1,3,5triazin2(1H)ones as the first dual BACE1/GSK3beta
fragment hits against Alzheimer's disease. ACS Chem Neurosci. 2015;6(10):16651682.
44
|
MOUSSAPACHA ET AL.
193. Viayna E, Sola I, Bartolini M, et al. Synthesis and multitarget biological profiling of a novel family of rhein derivatives
as diseasemodifying antiAlzheimer agents. J Med Chem. 2014;57(6):25492567.
194. Alzheimer's Association. Alzheimer's disease facts and figures. Alzheimers Dement. 2017;13(4):325373. 2017.
195. Nie F, Kunciw DL, Wilcke D, et al. A multidimensional diversityoriented synthesis strategy for structurally diverse
and complex macrocycles. Angew Chem Int Ed Engl. 2016;55(37):1113911143.
196. Cui J, Hao J, Ulanovskaya OA, et al. Creation and manipulation of common functional groups en route to a skeletally
diverse chemical library. Proc Natl Acad Sci U S A. 2011;108(17):67636768.
197. Schaduangrat N, Prachayasittikul V, Choomwattana S, et al. Multidisciplinary approaches for targeting the secretase
protein family as a therapeutic route for Alzheimer's disease. Med Res Rev. 2019. https://doi.org/10.1002/med.21563
198. Bachurin SO, Bovina EV, Ustyugov AA. Drugs in clinical trials for Alzheimer's disease: the major trends. Med Res Rev.
2017;37(5):11861225.
AUTHOR BIOGRAPHIES
Nour M. MoussaPacha earned a bachelor's degree in pharmacy with highest honors in the academic year
20172018 from University of Sharjah, United Arab Emirates. In October 2018, she started training program in
Dubai Health Authority to get the Professional Licensing in pharmacy.
Shifaa M. Abdin earned a bachelor's degree of pharmacy, in June 2018, from the College of Pharmacy,
University of Sharjah, United Arab Emirates, as 1st achiever (excellent with honor). She is currently enrolled for
M.Sc. degree in Molecular Medicine and Transitional Research at College of Medicine, University of Sharjah,
expected to graduate in June 2020. In addition, she currently works as a research assistant at Sharjah Institute
for Medical Research.
Hany A. Omar Dr. Omar got his Ph.D. in molecular pharmacology and therapeutics at The Ohio State
University, Columbus, Ohio, USA (2010). After postdoctoral studies at The Ohio State University (2011), he led
cancer cell molecular biology research in different places. He is now an Associate Professor of Molecular
Pharmacology and Therapeutics, University of Sharjah, United Arab Emirates & BeniSuef University, BeniSuef,
Egypt. He has been a regular reviewer or ad hoc reviewer for several medical journals, different funding agency
and several journal editorial board members and senior editors. Dr. Omar has published over 65 scientific
reports, including many in the most respected professional journals. Current positions: Associate Professor of
Molecular Pharmacology and Therapeutics, Faculty of Pharmacy, Cairo University (BSU Campus), Egypt and
College of Pharmacy, University of Sharjah, UAE. ViceDean and Associate Professor, College of Pharmacy,
University of Sharjah, Sharjah, UAE (2017 to present).
Hasan Alniss received his Ph.D. in Medicinal Chemistry (2011) from the University of StrathclydeUK. In 2011,
he was appointed Assistant Professor of Medicinal Chemistry at AnNajah National University (Palestine). In
2013, he was granted the Distinguished Scholar Award and joined the University of Toronto as a Visiting
Professor in the Leslie Dan Faculty of Pharmacy, where he worked with Prof. Robert MacGregor on a project to
characterize the factors that stabilize the Gquadruplex structures of nucleic acids. In 2015, Dr. Alniss took up
his current position at the University of Sharjah (UAE) as an Assistant Professor of Medicinal Chemistry. In
terms of research, his primary interest is cancer drug discovery and the biophysical characterization of nucleic
acid structures and their complexes with drugs.
Taleb ALTel received his BS in Chemistry and Chemical Technology in 1987, MS in Natural Product Chemistry
in 1990 from the Department of Chemistry, Jordan University and Pennsylvania State University, and
PhD in 1995 from Tuebingen UniversityGermany under Professor Wolfgang Voelter followed by an NIH
Postdoctoral Fellowship under the supervision of Professor Scott Sieburth at State University of New York,
Stony Brook, NY. In 2003 he was appointed as a visiting associate professor at Duke Chemistry Department,
North Carolina. AlTel joined Transtech PharmaUSA, a Drug Discovery and Development Company from 2004
MOUSSAPACHA ET AL.
|
45
to 2007 as principal scientist and promoted to a Team Leader. In 2007, he joined the college of
pharmacyUniversity of Sharjah in the United Arab Emirates as associate professor of organic medicinal
chemistry then promoted to Professor in 2013. AlTel coauthored over 80 invited lectures, presentations,
keynote Speaker and a coauthor of more than 75 publications in leading journals; inventor on four USpatents
and patent applications. AlTel received many awards including the German Award for the Exchange of Scholars
(DAAD); Abdel Hameed Shoman Prize for Young Arab Researchers (1999); Fulbright Award (2003/2004). Since
2014, AlTel holds the director position of the Research Institute of Medical and Health Sciences, University of
Sharjah, Sharjah, UAE.
How to cite this article: MoussaPacha NM, Abdin SM, Omar HA, Alniss H, AlTel TH. BACE1 inhibitors:
current status and future directions in treating Alzheimer's disease. Med Res Rev. 2019;146.
https://doi.org/10.1002/med.21622
46
|
MOUSSAPACHA ET AL.
... Intranasal delivery routes can bypass the BBB, offering a direct path to the CNS [150]. Also investigated are vaccines considering Aβ aggregates as antigens [151], Aβ aggregation inhibition [152], β-secretase inhibition [153], γ-secretase modulation [154], γ-secretase inhibition [155], intravenous immunoglobulin (IVIG) administration [156], and most widely introducing monoclonal anti-Aβ antibodies [157]. However, none of these strategies have met primary end goals in large-scale clinical trials. ...
Article
Full-text available
Antibodies that can selectively remove rogue proteins in the brain are an obvious choice to treat neurodegenerative disorders (NDs), but after decades of efforts, only two antibodies to treat Alzheimer’s disease are approved, dozens are in the testing phase, and one was withdrawn, and the other halted, likely due to efficacy issues. However, these outcomes should have been evident since these antibodies cannot enter the brain sufficiently due to the blood–brain barrier (BBB) protectant. However, all products can be rejuvenated by binding them with transferrin, preferably as smaller fragments. This model can be tested quickly and at a low cost and should be applied to bapineuzumab, solanezumab, crenezumab, gantenerumab, aducanumab, lecanemab, donanemab, cinpanemab, and gantenerumab, and their fragments. This paper demonstrates that conjugating with transferrin does not alter the binding to brain proteins such as amyloid-β (Aβ) and α-synuclein. We also present a selection of conjugate designs that will allow cleavage upon entering the brain to prevent their exocytosis while keeping the fragments connected to enable optimal binding to proteins. The identified products can be readily tested and returned to patients with the lowest regulatory cost and delays. These engineered antibodies can be manufactured by recombinant engineering, preferably by mRNA technology, as a more affordable solution to meet the dire need to treat neurodegenerative disorders effectively.
... The main pathological features of AD are β-amyloid peptide (Aβ) plaques, neurofibrillary tangles, and neuronal loss in the brain [27,43]. The deposits of extracellular fragments from an erroneous cleavage of the amyloid precursor protein (APP) promoted by β-secretase (BACE-1) and ɣ-secretase in the C99 APP domain is a pivotal event in the Aβ hypothesis [34,51]. The Aβ 1-42 chain has been described as a more cytotoxic isoform due to two insoluble amino acids in its structure [28,55]. ...
Article
Full-text available
The amyloid-beta (Aβ) aggregation in Alzheimer’s disease (AD) triggers neuroinflammation, and neurodegeneration, which lead to cognitive deficits along with other neuropsychiatric symptoms, including depression and anxiety. G protein-coupled receptor 35 (GPR35) is expressed in the brain and is involved in metabolic stresses. However, the role of GPR35 in AD pathogenesis remains unknown. Herein, pharmacological blockade, shRNA-mediated knockdown or knockout of GPR35 was performed to investigate the role and mechanisms of GPR35 in Aβ1–42-induced cognitive impairment and emotional alterations in mice. A series of behavioral, histopathological, and biochemical tests were performed in mice. Our results showed that hippocampal GPR35 expression was significantly increased in Aβ1–42-induced and APP/PS1 AD mouse models. Pharmacological blockade or knockdown of GPR35 ameliorated cognitive impairment and emotional alterations induced by Aβ1–42 in mice. We also found that blockade or knockdown of GPR35 decreased the accumulation of Aβ, and improved neuroinflammation, cholinergic system deficiency, and neuronal apoptosis via the RhoA/ROCK2 pathway in Aβ1–42-treaed mice. However, activation of GPR35 aggravates Aβ1–42-induced cognitive deficits and emotional alterations in mice. In addition, genetic deletion of GPR35 protects against the Aβ1–42-induced cognitive deficits and emotional alterations in mice. Moreover, GPR35 could bind to TLR4. These results indicate that GPR35 participates in the pathogenesis of cognitive deficits and emotional alterations induced by Aβ1–42 in mice, suggesting that GPR35 could be a potential therapeutic target for AD.
... The accumulation of amyloid beta peptides gets deposited in the brain as plaques, and the loss of cholinergic neurons leads to AD pathogenesis (Rajput et al., 2020b). Inhibition of cholinesterases (butyrylcholinesterase (BChE) and acetylcholinesterase (AChE)) that degrade neurotransmitters is crucial (Moussa-Pacha et al., 2020;Rajput et al., 2020a). Phytochemicals from mango and its byproducts have been reported to have inhibitory potential for activities against neurotransmitters, acetylcholine, and degrading enzymes, AChE and BchE . ...
... Aβ peptides are formed after the sequential cleavage of the amyloid precursor protein (APP), catalyzed by β-site APP-cleaving enzyme 1 (BACE1, β-secretase) and γ-secretase [40,41]. Preclinical studies on the development of drugs targeting BACE1 and γ-secretase have been conducted [42,43]. Our results revealed that the levels of Aβ, PHF-tau, NFT-like structures, and cellular apoptosis were significantly attenuated by the treatment with BACE1 inhibitor IV (BSI), a BACE 1 inhibitor, and compound E (CE), a γ-secretase. ...
Article
Full-text available
Background Cerebral organoids (COs) are the most advanced in vitro models that resemble the human brain. The use of COs as a model for Alzheimer’s disease (AD), as well as other brain diseases, has recently gained attention. This study aimed to develop a human AD CO model using normal human pluripotent stem cells (hPSCs) that recapitulates the pathological phenotypes of AD and to determine the usefulness of this model for drug screening. Methods We established AD hPSC lines from normal hPSCs by introducing genes that harbor familial AD mutations, and the COs were generated using these hPSC lines. The pathological features of AD, including extensive amyloid-β (Aβ) accumulation, tauopathy, and neurodegeneration, were analyzed using enzyme-linked immunosorbent assay, Amylo-Glo staining, thioflavin-S staining, immunohistochemistry, Bielschowsky’s staining, and western blot analysis. Results The AD COs exhibited extensive Aβ accumulation. The levels of paired helical filament tau and neurofibrillary tangle-like silver deposits were highly increased in the AD COs. The number of cells immunoreactive for cleaved caspase-3 was significantly increased in the AD COs. In addition, treatment of AD COs with BACE1 inhibitor IV, a β-secretase inhibitor, and compound E, a γ-secretase inhibitor, significantly attenuated the AD pathological features. Conclusion Our model effectively recapitulates AD pathology. Hence, it is a valuable platform for understanding the mechanisms underlying AD pathogenesis and can be used to test the efficacy of anti-AD drugs.
Article
Full-text available
Amyloid protein-β (Aβ) concentrations are increased in the brain in both early onset and late onset Alzheimer’s disease (AD). In early onset AD, cerebral Aβ production is increased and its clearance is decreased, while increased Aβ burden in late onset AD is due to impaired clearance. Aβ has been the focus of AD therapeutics since development of the amyloid hypothesis, but efforts to slow AD progression by lowering brain Aβ failed until phase 3 trials with the monoclonal antibodies lecanemab and donanemab. In addition to promoting phagocytic clearance of Aβ, antibodies lower cerebral Aβ by efflux of Aβ-antibody complexes across the capillary endothelia, dissolving Aβ aggregates, and a “peripheral sink” mechanism. Although the blood-brain barrier is the main route by which soluble Aβ leaves the brain (facilitated by low-density lipoprotein receptor-related protein-1 and ATP-binding cassette sub-family B member 1), Aβ can also be removed via the blood-cerebrospinal fluid barrier, glymphatic drainage, and intramural periarterial drainage. This review discusses experimental approaches to increase cerebral Aβ efflux via these mechanisms, clinical applications of these approaches, and findings in clinical trials with these approaches in patients with AD or mild cognitive impairment. Based on negative findings in clinical trials with previous approaches targeting monomeric Aβ, increasing the cerebral efflux of soluble Aβ is unlikely to slow AD progression if used as monotherapy. But if used as an adjunct to treatment with lecanemab or donanemab, this approach might allow greater slowing of AD progression than treatment with either antibody alone.
Article
JOURNAL/nrgr/04.03/01300535-202502000-00030/figure1/v/2024-06-06T062529Z/r/image-tiff In patients with Alzheimer’s disease, gamma-glutamyl transferase 5 (GGT5) expression has been observed to be downregulated in cerebrovascular endothelial cells. However, the functional role of GGT5 in the development of Alzheimer’s disease remains unclear. This study aimed to explore the effect of GGT5 on cognitive function and brain pathology in an APP/PS1 mouse model of Alzheimer’s disease, as well as the underlying mechanism. We observed a significant reduction in GGT5 expression in two in vitro models of Alzheimer’s disease (Aβ 1 – 42 –treated hCMEC/D3 and bEnd.3 cells), as well as in the APP/PS1 mouse model. Additionally, injection of APP/PS1 mice with an adeno-associated virus encoding GGT5 enhanced hippocampal synaptic plasticity and mitigated cognitive deficits. Interestingly, increasing GGT5 expression in cerebrovascular endothelial cells reduced levels of both soluble and insoluble amyloid-β in the brains of APP/PS1 mice. This effect may be attributable to inhibition of the expression of β-site APP cleaving enzyme 1, which is mediated by nuclear factor-kappa B. Our findings demonstrate that GGT5 expression in cerebrovascular endothelial cells is inversely associated with Alzheimer’s disease pathogenesis, and that GGT5 upregulation mitigates cognitive deficits in APP/PS1 mice. These findings suggest that GGT5 expression in cerebrovascular endothelial cells is a potential therapeutic target and biomarker for Alzheimer’s disease.
Article
Full-text available
Clearance of amyloid-beta (Aβ) from the brain is impaired in both early-onset and late-onset Alzheimer’s disease (AD). Mechanisms for clearing cerebral Aβ include proteolytic degradation, antibody-mediated clearance, blood brain barrier and blood cerebrospinal fluid barrier efflux, glymphatic drainage, and perivascular drainage. ATP-binding cassette (ABC) transporters are membrane efflux pumps driven by ATP hydrolysis. Their functions include maintenance of brain homeostasis by removing toxic peptides and compounds, and transport of bioactive molecules including cholesterol. Some ABC transporters contribute to lowering of cerebral Aβ. Mechanisms suggested for ABC transporter-mediated lowering of brain Aβ, in addition to exporting of Aβ across the blood brain and blood cerebrospinal fluid barriers, include apolipoprotein E lipidation, microglial activation, decreased amyloidogenic processing of amyloid precursor protein, and restricting the entrance of Aβ into the brain. The ABC transporter superfamily in humans includes 49 proteins, eight of which have been suggested to reduce cerebral Aβ levels. This review discusses experimental approaches for increasing the expression of these ABC transporters, clinical applications of these approaches, changes in the expression and/or activity of these transporters in AD and transgenic mouse models of AD, and findings in the few clinical trials which have examined the effects of these approaches in patients with AD or mild cognitive impairment. The possibility that therapeutic upregulation of ABC transporters which promote clearance of cerebral Aβ may slow the clinical progression of AD merits further consideration.
Article
Full-text available
Alzheimer’s disease (AD) is the most common type of dementia, an irreversible neurodegenerative disease that worsens over time. It frequently affects the elderly population and, therefore, its prevalence is soaring in countries with a high proportion of the aging population. The incidence of AD is likely to increase further with increasing life expectancy throughout the world. There is still no sight of therapies that could stop or slow the progression of AD despite the significant expansion in our understanding of the disease at molecular, cellular, and system levels in the last 50 years. The only medications approved by the Food and Drug Administration (FDA), for the treatment of AD are acetylcholinesterase inhibitors (donepezil, galantamine, tacrine, and rivastigmine) and glutamate receptor antagonist (memantine). Apart from acetylcholinesterase inhibition, another important therapeutic target for treating AD is the inhibition of β‐site APP cleaving enzyme‐1 (BACE1), which is involved in the proteolysis of the amyloid precursor protein (APP) leading to the formation of neurotoxic amyloid β (Aβ) protein. Extensive research following this approach has led to the rationally designed synthetic, and potent BACE1 inhibitors, several of which are in clinical trials. However, the high failure rate of the BACE1 inhibitors in clinical trials necessitates finding a different source of BACE1 inhibitors. In this review, we discuss the most promising phytocompounds and plant extracts showing potent BACE1 inhibitory activity. Future research directions are suggested and recommendations are made to expand the use of medicinal plants and their formulations to prevent, mitigate and treat AD.
Article
Full-text available
Microglia maintain homeostasis in the central nervous system through phagocytic clearance of protein aggregates and cellular debris. This function deteriorates during ageing and neurodegenerative disease, concomitant with cognitive decline. However, the mechanisms of impaired microglial homeostatic function and the cognitive effects of restoring this function remain unknown. We combined CRISPR–Cas9 knockout screens with RNA sequencing analysis to discover age-related genetic modifiers of microglial phagocytosis. These screens identified CD22, a canonical B cell receptor, as a negative regulator of phagocytosis that is upregulated on aged microglia. CD22 mediates the anti-phagocytic effect of α2,6-linked sialic acid, and inhibition of CD22 promotes the clearance of myelin debris, amyloid-β oligomers and α-synuclein fibrils in vivo. Long-term central nervous system delivery of an antibody that blocks CD22 function reprograms microglia towards a homeostatic transcriptional state and improves cognitive function in aged mice. These findings elucidate a mechanism of age-related microglial impairment and a strategy to restore homeostasis in the ageing brain.
Article
Full-text available
Microglial associations with both the major Alzheimer’s disease (AD) pathognomonic entities, β‐amyloid‐positive plaques and tau‐positive neurofibrillary tangles, have been noted in previous investigations of both human tissue and mouse models. However, the precise nature of their role in the pathogenesis of AD is debated; the major working hypothesis is that pro‐inflammatory activities of activated microglia contribute to disease progression. In contrast, others have proposed that microglial dystrophy with a loss of physiological and neuroprotective activities promotes neurodegeneration. This immunohistochemical study sought to gain clarity in this area by quantifying the morphological subtypes of microglia in the mildly‐affected primary visual cortex (PVC), the moderately affected superior frontal cortex (SFC) and the severely affected inferior temporal cortex (ITC) of 8 AD cases and 15 age and gender‐matched, non‐demented controls with ranging AD‐type pathology. AD cases had increased β‐amyloid and tau levels compared to controls in all regions. Neuronal loss was observed in the SFC and ITC, and was associated with atrophy in the latter. A major feature of the ITC in AD was a decrease in ramified (healthy) microglia with image analysis confirming reductions in arborized area and skeletal complexity. Activated microglia were not associated with AD but were increased in non‐demented controls with greater AD‐type pathology. Microglial clusters were occasionally associated with β‐amyloid‐ and tau‐positive plaques but represented less than 2% of the total microglial population. Dystrophic microglia were not associated with AD, but were inversely correlated with brain pH suggesting that agonal events were responsible for this morphological subtype. Overall these novel findings suggest that there is an early microglial reaction to AD‐type pathology but a loss of healthy microglia is the prominent feature in severely affected regions of the AD brain.
Article
Full-text available
In this issue, an article by Tiepolt et al. shows that PET scanning using [11C]PiB can demonstrate both cerebral blood flow (CBF) changes and amyloid-β (Aβ) deposition in patients with mild cognitive dysfunction or mild dementia of Alzheimer's disease (AD). The CBF changes can be determined because the early scan counts (1-9 minutes) reflect the flow of the radiotracer in the blood passing through the brain, while the Aβ levels are measured by later scan counts (40-70 minutes) after the radiotracer has been cleared from regions to which the radiotracer did not bind. Thus, two different diagnostic measures are obtained with a single injection. Unexpectedly, the mild patients with Aβ positivity had scan data with only a weak relationship to memory, while the relationships to executive function and language function were relatively strong. This divergence of findings from studies of severely impaired patients highlights the importance of determining how AD pathology affects the brain. A possibility suggested in this commentary is that Aβ deposits occur early in AD and specifically in critical areas of the neocortex affected only later by the neurofibrillary pathology indicating a different role of the amyloid-β protein precursor (AβPP) in the development of those neocortical regions, and a separate component of AD pathology may selectively impact functions of these neocortical regions. The effects of adverse AβPP metabolism in the medial temporal and brainstem regions occur later possibly because of different developmental issues, and the later, different pathology is clearly more cognitively and socially devastating.
Article
Full-text available
Tau neuronal and glial pathologies drive the clinical presentation of Alzheimer’s disease and related human tauopathies. There is a growing body of evidence indicating that pathological tau species can travel from cell to cell and spread the pathology through the brain. Throughout the last decade, physiological and pathological tau have become attractive targets for AD therapies. Several therapeutic approaches have been proposed, including the inhibition of protein kinases or protein-3-O-(N-acetyl-beta-D-glucosaminyl)-L-serine/threonine Nacetylglucosaminyl hydrolase, the inhibition of tau aggregation, active and passive immunotherapies, and tau silencing by antisense oligonucleotides. New tau therapeutics, across the board, have demonstrated the ability to prevent or reduce tau lesions and improve either cognitive or motor impairment in a variety of animal models developing neurofibrillary pathology. The most advanced strategy for the treatment of human tauopathies remains immunotherapy, which has already reached the clinical stage of drug development. Tau vaccines or humanised antibodies target a variety of tau species either in the intracellular or extracellular spaces. Some of them recognise the amino-terminus or carboxy-terminus, while others display binding abilities to the proline-rich area or microtubule binding domains. The main therapeutic foci in existing clinical trials are on Alzheimer’s disease, progressive supranuclear palsy and non-fluent primary progressive aphasia. Tau therapy offers a new hope for the treatment of many fatal brain disorders. First efficacy data from clinical trials will be available by the end of this decade.
Article
Full-text available
Background: Alzheimer disease (AD) can be considered as the most common age related neurodegenerative disorder and also an important cause of death in elderly patients. A number of studies showed the correlation of this disease pathology with BACE1 inhibitor and it is also evident that BACE1 inhibitor can function as a very potent strategy in treating AD. Methods: In this present study, we aimed to prospect for novel plant-derived BACE1 inhibitors from Leea indica and to realise structural basis of their interactions and mechanisms using combined molecular docking and molecular dynamics based approaches. An extensive library of Leea indica plant derived molecule was compiled and computationally screened for inhibitory action against BACE1 by using virtual screening approaches. Furthermore, induced fit docking and classical molecular dynamics along with steered molecular dynamics simulations were employed to get insight of the binding mechanisms. Results: Two triterpenoids, ursolic acid and lupeol were identified through virtual screening; wherein, lupeol showed better binding free energy in MM/GBSA, MM/PBSA and MM/GBVI approaches. Furthermore classical and steered dynamics revealed the favourable hydrophobic interactions between the lupeol and the residues of flap or catalytic dyadof BACE1; however, ursolic acid showed disfavorable interactions with the BACE1. Conclusion: This study therefore unveiled the potent BACE1 inhibitor from a manually curated dataset of Leea indica molecules, which may provide a novel dimension of designing novel BACE1 inhibitors for AD therapy.
Article
Background: Alzheimer's disease (AD) is associated with extracellular accumulation and aggregation of amyloid β (Aβ) peptides ultimately seeding in senile plaques. Recent data show that their direct precursor C99 (βCTF) also accumulates in AD-affected brain as well as in AD-like mouse models. C99 is consistently detected much earlier than Aβ, suggesting that this metabolite could be an early contributor to AD pathology. C99 accumulates principally within endolysosomal and autophagic structures and its accumulation was described as both a consequence and one of the causes of endolysosomalautophagic pathology, the occurrence of which has been documented as an early defect in AD. C99 was also accompanied by C99-derived C83 (αCTF) accumulation occurring within the same intracellular organelles. Both these CTFs were found to dimerize leading to the generation of higher molecular weight CTFs, which were immunohistochemically characterized in situ by means of aggregate-specific antibodies. Discussion: Here, we discuss studies demonstrating a direct link between the accumulation of C99 and C99-derived APP-CTFs and early neurotoxicity. We discuss the role of C99 in endosomal-lysosomalautophagic dysfunction, neuroinflammation, early brain network alterations and synaptic dysfunction as well as in memory-related behavioral alterations, in triple transgenic mice as well as in newly developed AD animal models. Conclusion: This review summarizes current evidence suggesting a potential role of the β -secretasederived APP C-terminal fragment C99 in Alzheimer's disease etiology.
Article
The continual increase of the aging population worldwide renders Alzheimer's disease (AD) a global prime concern. Several attempts have been focused on understanding the intricate complexity of the disease's development along with the on‐ andgoing search for novel therapeutic strategies. Incapability of existing AD drugs to effectively modulate the pathogenesis or to delay the progression of the disease leads to a shift in the paradigm of AD drug discovery. Efforts aimed at identifying AD drugs have mostly focused on the development of disease‐modifying agents in which effects are believed to be long lasting. Of particular note, the secretase enzymes, a group of proteases responsible for the metabolism of the β‐amyloid precursor protein (βAPP) and β‐amyloid (Aβ) peptides production, have been underlined for their promising therapeutic potential. This review article attempts to comprehensively cover aspects related to the identification and use of drugs targeting the secretase enzymes. Particularly, the roles of secretases in the pathogenesis of AD and their therapeutic modulation are provided herein. Moreover, an overview of the drug development process and the contribution of computational (in silico) approaches for facilitating successful drug discovery are also highlighted along with examples of relevant computational works. Promising chemical scaffolds, inhibitors, and modulators against each class of secretases are also summarized herein. Additionally, multitarget secretase modulators are also taken into consideration in light of the current growing interest in the polypharmacology of complex diseases. Finally, challenging issues and future outlook relevant to the discovery of drugs targeting secretases are also discussed.
Article
Beta secretase 1 (BACE1) is an enzyme involved in the pathogenesis of Alzheimer's disease (AD). PLB4 mice are a neuron-specific human BACE1 knock-in mouse model characterized by the accumulation of extracellular Aβ and an AD-like phenotype. In this investigation brain hemispheres from ‘young’ (4–6 months) and ‘old’ (8 months) female PLB4 mice and age-matched wild-type littermates underwent targeted LC-MS/MS metabolomic profiling. Powdered lyophilized brain tissue was extracted in ethanol:PBS 85%:15% (v/v)) and a total of 187 metabolites were quantified using a targeted metabolomics methodology. Multivariate statistical analysis produced models distinguished PLB4 from wild type (WT) mice regardless of their age group. Univariate analysis (t-test) found that more brain metabolites were perturbed in ‘old’ PLB4 mice than ‘young’. Carnosine and 8 phosphatidylcholine species were significantly decreased (p < 0.05) in ‘young’ PLB4 mouse brain. In ‘old’ PLB4 mice a total of 21 metabolites were perturbed including: leucine, creatinine, putrescine and species of acylcarnitines, lysophosphatidylcholines, phosphatidylcholines and sphingomyelin. Within the PLB4 genotype there were a range of age-dependent increases in metabolites. This study indicates that gender-specific responses occur in models of AD-like pathology, but importantly, when changes in PLB4 mice (where Aβ oligomers predominate) are compared with APP/PS1 mice (where Aβ plaques predominate) there are consistent and also divergent effects on the brain metabolome.