3

Amyloid β-Targeted Inhibitory Peptides for Alzheimer’s Disease: Current State and Future Perspectives

Safura Jokar1 • Saeedeh Khazaei2 • Xaniar Esmailzadeh Gameshgoli3 • Mona Khafaji4 • Behnam Yarani5 • Mohammad Sharifzadeh6 • Davood Beiki7 • Omid Bavi8

1Department of Nuclear Pharmacy, Faculty of Pharmacy, Tehran University of Medical Sciences, Tehran, Iran; 2Department of Pharmaceutical Biomaterials, Faculty of Pharmacy, Tehran University of Medical Sciences, Tehran, Iran; 3Colloids and Polymers Physics Group, Particle Physics Department, Universidade de Santiago de Compostela, Santiago de Compostela, Spain; 4Institute for Nanoscience and Nanotechnology, Sharif University of Technology, Tehran, Iran; 5Division of Nanobiotechnology, Department of Life Sciences Engineering, Faculty of New Sciences and Technologies, University of Tehran, Tehran, Iran; 6Department of Toxicology and Pharmacology, Faculty of Pharmacy; Toxicology and Poisoning Research Centre, Tehran University of Medical Sciences, Tehran, Iran; 7Research Center for Nuclear Medicine, Tehran University of Medical Sciences, Tehran, Iran; 8Department of Mechanical and Aerospace Engineering, Shiraz University of Technology, Shiraz, Iran

Abstract: Alzheimer’s disease is the most common irreversible neurodegenerative disorder. To date, there is no cure for Alzheimer’s disease. While multiple pathological mechanisms have been proposed for the onset and progression of Alzheimer’s disease, the hypothesis that attracted much attention is the amyloid hypothesis. The senile plaques that accumulate in the brain of Alzheimer’s disease patients are predominantly composed of beta amyloid (Aβ). Aβ deposition in the brain is thought to occur years before the emergence of clinical symptoms. The overproduction, aggregation, and fibrillation of Aβ, combined with reduced clearance, eventually lead to amyloid plaque formation and subsequent neurotoxicity. Hence, inhibition of Aβ aggregation and the promotion of Aβ clearance have been actively explored as therapeutic strategies for Alzheimer’s disease. This chapter provides an overview of the current knowledge on one such strategy, Aβ-targeted inhibitory peptides.

Keywords: Aβ aggregation in Alzheimer’s disease; biopanning; inhibitory peptides for Alzheimer’s disease; peptide–nanostructure conjugates; peptidomimetics

Authors for correspondence: Omid Bavi, Department of Mechanical and Aerospace Engineering, Shiraz University of Technology, Shiraz, Iran. Email: o.bavi@sutech.ac.ir; Davood Beiki, Research Center for Nuclear Medicine, Tehran University of Medical Sciences, Shariati Hospital, North Kargar Ave. 1411713135, Tehran, Iran. Email: beikidav@tums.ac.ir

Doi: https://doi.org/10.36255/exonpublications.alzheimersdisease.2020.ch3

In: Alzheimer’s Disease: Drug Discovery. Huang X (Editor). Exon Publications, Brisbane, Australia. ISBN: 978-0-6450017-0-9; Doi: https://doi.org/10.36255/exonpublications.alzheimersdisease.2020

Copyright: The Authors.

Licence: This open access article is licenced under Creative Commons Attribution-NonCommercial 4.0 International (CC BY-NC 4.0) https://creativecommons.org/licenses/by-nc/4.0/

INTRODUCTION

Alzheimer’s disease is an age-dependent disorder that is the fifth leading cause of death in people aged 65 years and older. It is estimated that over 50 million people worldwide suffer from Alzheimer’s disease, and this figure is set to increase to 152 million by 2050 with a financial burden of 1.1 trillion US dollars by 2050 (13). Several hypotheses, including the amyloid, cholinergic (4), and Tau protein hypotheses have been proposed to explain the pathophysiology and etiology of Alzheimer’s disease (5). Because of the presence of Aβ in the brain tissue, cerebrospinal fluid, and plasma, the amyloid cascade hypothesis is the most widely accepted. The amyloid cascade hypothesis states that neurodegeneration in Alzheimer’s disease is the result of amyloid plaque and neurofibrillary tangle formations (6, 7). The overproduction, clearance failure, aggregation, and fibrillation of Aβ eventually leads to amyloid plaque formation. These factors also contribute to neuroinflammation and cell death. Aβ deposition in the brain is likely to be the first pathological incident that occurs years before the emergence of clinical symptoms. Aβ is produced through the proteolytic cleavage of the amyloid precursor protein (APP), a transmembrane glycoprotein, which is made up of a cytoplasmic domain with 55 amino acids and a long extracellular domain with 590–680 amino acids (8). APP cleavage by the proteases β- and γ-secretases produce Aβ fragments of varying size depending on the cleavage site (9), of which Aβ40 (about 90%) and Aβ42 (about 5–10%) are the most prevalent (Figure 1). Aβ42 is more toxic than Aβ40. After production, the Aβ peptides aggregate to form amyloid deposits. There are different aggregation forms such as low molecular weight oligomers, protofibrils, as well as mature fibrils that eventually come together to form amyloid deposits in the brain parenchyma and cerebrovascular spaces (10, 11).

Fig 1

Figure 1 Amyloid-β fibrillation and neuronal damage. First, amyloid precursor protein is cleaved by β and γ secretases, respectively. A peptide fragment of 39-42 amino acid is formed depending on the site of cleavage. After cleavage, Aβ monomers start to self-assemble to form soluble toxic aggregates, and finally into insoluble fibrils, which subsequently cause synaptic dysfunction and neuronal death.

Therefore, inhibition of Aβ aggregation and the promotion of Aβ clearance have been investigated as therapeutic strategies for Alzheimer’s disease. Some of these strategies include the use of metal chelators (12), peptides (13), organic molecules (14), and biomolecules (15, 16). Peptides are considered a better option than small molecule-based compounds because of their high affinity for Aβ and low toxicity (17). Although natural amino acid-based peptides are effective inhibitors of Aβ aggregation, they are prone to faster enzymatic degradation and show a tendency for self-assembly into fibrils during administration (15). To overcome these problems, modified peptides have been generated (18) with D-amino acids, retro-inverso cyclization, fluorination, as well as N-methylation of the ester bond (19). With this knowledge, peptides could be potential candidates for inhibiting Aβ conformational transitions, self-assembly, and toxicity against neurons, and promotion of the pathways of the nontoxic fibrillation and early diagnosis of Alzheimer’s disease (20). This chapter provides an overview of the therapeutic potential peptides as Aβ aggregation inhibitors.

PEPTIDIC INHIBITORS

Luhrs and co-workers first experimentally described the structure of the Aβ42 fibril (Figure 2, right) (21). At least four specific structural sites for interaction have been identified on the Aβ fibril (22): (i), hydrophobic regions of Ala30–Val36, and Leu17–Ala21 residues from the C and N-terminal β-sheets respectively; (ii), hydrophilic part using electrostatic interactions between Asp23 and Lys28 residues; (iii), central cleft in the interior of the U-shaped turn; and (iv), Glu22 ladder between the side chains of the Glu22 residues of the adjacent β-strands (Figure 2, bottom right). The formation of the salt-bridge between Asp23 and Lys28 is an essential β-sheet conformation stabilizer. Moreover, it might stimulate the oligomerization of Aβ via stabilizing the Val24–Asn27 turn (23). The hydrophobic residue of Met35 in the C-terminus domain could support fibril stability by hydrophobic interactions. The Met35 binding site can potentially inhibit protein-protein interactions and prevent amyloid fibril formation (24). These sites are probably critical regions in the initiation of Aβ nucleation, conformational transition promotion, and fibril formation. The residues 16KLVFFA21 (Figure 2) of the central hydrophobic core (CHC) region is a critical nucleation site, or self-recognition sequence. The Ile41 and Ala42 residues can modulate Aβ42 oligomer formation (25) by interacting with the N and C-terminus of Aβ42 (26). Figure 2, left, shows the binding sites of the most common inhibitory peptides, such as LPFFD and KVLFF, on the Aβ40 fibril structure.

Fig 2

Figure 2 The structure of Aβ. Right, schematic of the Aβ42 amyloid fibril. Left, binding sites of inhibitory peptides of LPFFD and KVLFF on Aβ40 fibril.

Peptide inhibitors are generally divided into Aβ-based peptide inhibitors and non-Aβ-based peptide inhibitors. A list of select Aβ inhibitory peptides are presented in Table 1.

TABLE 1 A select list of inhibitory peptides against Aβ peptide aggregation

No
Name/sequence of inhibitor peptide
Configuration
Type
Therapeutic findings
Ref.
1 RR (RYYAAFFARR) L-enantiomer Rational design based on hydrophobic core and salt bridge region
  • Inhibition of Aβ40 fibrillation and decrease of Aβ40 induced toxicity (in vitro)
(27)
2 D-(PGKLVYA) and D(KKLVFFARRRRA) D-enantiomer CHC-derived sequence
  • Inhibition of Aβ aggregation
  • Increase life time in an animal model of expressing Aβ42
(28)
3 CP-2 (lLwHsK) Cyclic d,l-α-peptide Screening of the randomly library members (500)
  • Nontoxic
  • Stabilize small Aβ oligomers
  • Disassemble formed Aβ fibrils
(29)
4 D3D3 (D-RPRTRLHTHRNRRPRvTRLHTHRNR) D-enantiomer Not derived from Aβ sequence
  • More inhibitory effect against Aβ42 oligomerization than D3 peptide
  • Formation of nontoxic amorphous instead of toxic oligomers (In vivo)
(30)
5 RD2 (DPTLHTHNRRRR-NH2) D-enantiomer
  • Not derived from Aβ sequence
  • Good binding affinity to Aβ
  • Reduce Aβ fibrillation formation
  • High stability in mouse plasma
  • High bioavailability (In vivo)
(31, 32)
6 AOEP2 (FDYKAEFMPWDT) D-enantiomer
  • A mimotope of the Aβ oligomer
  • Selected by phage display
  • Targeting of all forms of Aβ
  • The considrable reduction of TNF-α
(33)
7 LPYFD-amide L-enantiomer CHC-derived sequence
  • Neuroprotective
  • Decrease tau aggregation (In vivo)
(34)
8 Ac-LPFFN-NH2 L-enantiomer CHC-derived sequence
  • Inhibit of Aβ40 aggregation
  • Stabilize the α-helical conformation of Aβ40
  • Prolong the fibril formation
(35)
9 D-4F peptide D-enantiomer Not derived from Aβ sequence
  • Inhibit Aβ deposition
  • Improve cognitive function
(36)
10 Diazirine-equipped cyclo-KLVF(b-Ph)F cyclic CHC-derived sequence
  • Inhibit Aβ aggregation
  • Prevent Aβ42 induced toxicity
(37)
11 Attached flavin to an Aβ-binding peptide cyclic CHC-derived sequence
  • Inhibition of Aβ aggregation
  • Decrease of the aggregation potency and Aβ induced neurotoxicity
(38)
12 ZAb3 affibody L-enantiomer Selected by phage display technique
  • Inhibit Aβ aggregation
  • Disassemble preformed oligomers
(39)
13 H102 (HKQLPFFEED) L-enantiomer CHC-derived sequence
  • Inhibit Aβ42 fibrillization
(40)
14 Th-NT, Th-CT, and Th-SC Trehalose conjugated peptides, L-enantiomer CHC-derived sequence
  • Inhibit fibril formation
(41)
15 Fc-KLVFF Ferrocene- conjugated peptide, L-enantiomer CHC-derived sequence
  • Inhibit Aβ aggregation
(42)
16 N-methylated Aβ(3237) L-enantiomer C-terminal fragment
  • Inhibits Aβ aggregation
  • Improve Drosophila longevity
  • Increase locomotion
(43)
17 D-Trp-Aib D-enantiomer Rational design
  • Inhibits Aβ oligomerization
  • Decrease cerebral amyloid deposits in transgenic model
  • Improves cognitive activity (in vivo)
(44)
18 GABA-FPLIAIMA D-enantiomer C-terminal fragment
  • Inhibit and reduce the Aβ aggregation
(22)
19 Tyr(Allyl-RCM)-Xaa-Gly(Allyl-RCM) and ‘Gly(Allyl-RCM)-Xaa-Tyr(Allyl-RCM) cyclic CHC-derived sequence
  • Inhibit significantly the Aβ42 aggregation
(45)
20 D(KLVFW)- aminobutyric acid (Aib) D-enantiomer CHC-derived sequence
  • High specific interaction with Aβ42 monomers and oligomers
  • Remarkable inhibition of A Aβ42 fibril formation
  • No cytotoxic effects
(46)

Aβ-based peptide inhibitors

These are based on the structure of the C-terminal fragments (CTFs) and the CHC sequences of the Aβ peptide. They bind to the Aβ peptide at specific sites and prevent its assembly into amyloid fibrils. Peptides consisting of D-enantiomeric amino acids exhibit greater stability against proteases and show a higher binding affinity for Aβ compared with their L-enantiomeric counterparts. Moreover, D-peptides inhibit Aβ aggregation in animal models (28). Retro-inverso peptides are a special class of modified peptides that contain D-amino acids and reversed NH and CO groups in the peptide bonds. These peptides could keep the same spatial position in the side chain of the residues and preserve the desirable 3D structure compared to unchanged L-peptides (47). They also displayed advantages in terms of Aβ aggregation inhibition, higher proteolytic stability, lower self-assembly, and better blood-brain-barrier (BBB) permeability when compared with L-peptides in an animal model (48, 49). Fluorinated hydrophobic valine or phenylalanine in the LVFFA-based peptides can considerably delay the formation of Aβ aggregation. Fluorinated amino acids can also inhibit Aβ aggregation (50). Modification of amide functional groups with a methyl group is another strategy in the development of new inhibitors. N-methylated amide groups could enhance the peptide’s solubility in aqueous solutions and decrease Aβ-induced toxicity. Cyclic peptides have a higher inhibitory activity than acyclic derivatives (51). Because of their high enzymatic resistance, they are degraded slowly. Residues of lysine and glutamic acid have been known to be effective stabilizing and enhancing agents of Aβ fibrillation due to their ability to improve surface tension. In contrast, arginine residues have been reported as aggregation inhibitors or destabilizers (chaotropes) (52).

The 16KLVFF20-based peptides play a crucial role in disrupting Aβ aggregation by binding to full-length Aβ peptides and preventing fibril formation (53, 54). Ac-LVFFARK-NH2 (LK7), designed by adding arginine and lysine to KLVFF, induced a dose-dependent inhibition on Aβ42 fibrillation; however, it was cytotoxic due to high self-assembling properties (55). When conjugated with poly (lactic-co-glycolic acid) nanoparticles (NPs), the LK7-PLGA-NPs complex resulted in the elimination of the LK7 self-assembly feature while inhibiting Aβ42 fibrillation (55). Binding β-cyclodextrin to LK7 (56) improved LK7 peptide solubility, inhibited its tendency to self-aggregate, improved its binding to Aβ, and inhibited Aβ aggregation. Head-to-tail cyclization of LK7 peptide also resulted in a decrease in self-assembly of the LK7, an increase in binding affinity to the Aβ40 peptide, and proteolytic stability in serum. This derivative also can stabilize the Aβ40 secondary structure and inhibit Aβ40 –mediated cytotoxicity. Another derivative of LK7 peptide is Ac-LVFFARKHH-NH2 (LK7-HH), in which LK7 has been conjugated to the HH ligand as a chelator for reducing reactive oxygen species (ROS) production and capturing free and complexed ions of Cu2+ (57). This chelator also improved the anti-aggregate effects of LK7 on Aβ peptide and reduced its self-aggregation properties.

Proline and aspartic acid were exchanged for valine and alanine, respectively, in KVLFFA (58, 59). The derived peptide, referred to as 5-mer iAβ5 with sequence LPFFD, inhibited Aβ aggregation, neurotoxicity, and reduced plaque load (58). Due to the lack of a proton on the secondary substituted nitrogen in the peptide bond of proline residue, it could inhibit the formation of the intramolecular hydrogen bonds into fibrils. Since these small peptides are prone to faster enzymatic degradation and have reduced BBB permeability in vivo, iAβ5p was modified by N-methylation between Pro and Phe residues to improve its stability (60). The results from in vitro and in vivo studies showed that it has the same inhibitory activity as the parental iAβ5 peptide against amyloid fibril formation and neurotoxicity but with improved protease resistance. Also, molecular dynamics simulations show that this peptide has more durable binding and enhanced activity against Aβ40 aggregation in comparison to the iAβ5 peptide. In a similar study, the RIVFF sequence was produced by residue mutations of lysine16 (K) to arginine (R) and leucine17 (L) to isoleucine (I) on the KLVFF segment (61). The results indicated that this peptide could self-aggregate into β-sheet structures by reducing the surface tension of water and at higher concentrations (>250 μM) enhanced the Aβ-induced cytotoxicity.

The peptide D-GRKKRRQRRR-GGGG-DVEFRH (Aβ1−6 A2V-TAT) was investigated in vivo (62, 63). It was generated by modifying the N-terminal fragment of 1DAEFRH6 through mutation of alanine in position 2 to valine and conjugating with the HIV protein transduction domain GRKKRRQRRR (TAT). The resulting peptide showed strong anti-amyloidogenic effects in vitro and Aβ aggregation inhibition in mouse models of Alzheimer’s disease (64). The KLVFWAK motif was designed based on the 16KLVFFAE22 sequence with mutations introduced at phenylalanine and glutamic acid residues to tryptophan and lysine respectively to enhance solubility and disrupt self-assembly via electrostatic repulsion. Results showed that the designed motif could only target the C-terminal region of Aβ oligomers. The designed motif exhibited a lower self-aggregation tendency in comparison to other KLVFF-related sequences. Moreover, it demonstrated a higher binding affinity to Aβ aggregates and fibrils than monomers (65).

RGKLVFFGR (OR1) and RGKLVFFGR-NH (OR2) are retro-inverso peptides (66), designed by the addition of arginine (R) and glycine (G) to the KLVFF sequence. They exhibit high solubility and stability against enzymes. However, only the OR2 peptide showed inhibitory effects on Aβ oligomer formation and cytotoxicity. OR2 was modified to HN-rGklvffGr-Ac (RI-OR2) by acetylation of the C-terminal residue (49). The result illustrated that the peptide has a high resistance to proteolysis, while maintaining the same inhibitory activity in vivo. In a follow-up study, the RI-OR2 peptide was attached to the TAT peptide to improve its permeability into cells and the BBB (48). The results showed the peptide was able to decrease Aβ aggregation, plaque levels, and oxidative damages as well as increase the number of young neurons in the brain.

31IIGLMVGGVVIA42 and 39VVIA42 sequences were designed based on the C-terminal domain of Aβ42 (67). The 39VVIA42 sequences could interact with Aβ42 monomers and smaller oligomers at several sites, specifically at the N-terminal domain. At micromolar concentrations, the VVIA-NH2 peptide inhibited Aβ42 aggregation, exhibited less toxicity, and protected synaptic activity. However, these effects were not observed for the acetylated Ac-VVIA sequence (68). The non-acetylated VVIA-NH2 sequence particularly interacts with the C-terminal domain while the Ac-VVIA peptide has a dispersed binding distribution (68). The Ac-32IGLMVG37-NH2 sequence, a hexapeptide from the C-terminal fragment, has been shown to have a moderate efficacy with less toxicity (69).

O-acyl isopeptide and NMe-b-Ala26 (70) were derived from the full-length Aβ sequence with modification of an ester bond at the Gly25-Ser26 moiety and an N-methyl amide-β-Ala26, respectively. O-acyl isopeptide inhibited Aβ42 fibrillation at equimolar concentrations through an inhibitory mechanism distinct from any other peptidic inhibitors reported previously. Also, this derivative was more soluble than Aβ42 peptides and rapidly decomposed to Aβ42 monomers under physiological conditions through an O-to-N acyl rearrangement reaction whereas NMe-b-Ala26 showed higher chemical stability at physiological conditions.

Non-Aβ-based peptide inhibitors

Carnosine, a natural imidazole-containing dipeptide is a metal ion chelator (71). It inhibits the fibrillation and toxicity of amyloidogenic species such as glycated α-Crystallin, Aβ peptide, and prions. This peptide also inhibits the intramolecular salt bridge formation, which is vital to the stability and elongation of fibrils (71). Peptide D1, QSHYRHISPAQV (72), is another non-Aβ peptide that reduces Aβ aggregation and Aβ-associated cytotoxicity at high concentrations. N-methylated proprietary peptides such as D-NH2 (SEN304) and SEN1576 can inhibit Aβ-associated toxicity in vivo (73). Furthermore, SEN304 is a more potent inhibitor than customized versions of the KLVFF peptide. These peptides could interfere with the nucleation of Aβ, convert them into non-toxic forms, and eliminate toxic oligomers.

PEPTIDE LIBRARY SCREENING

There are many screening approaches to identify target-specific ligands (74, 75). Phage display is one such efficient high-throughput screening method that allows the screening of a wide variety of peptide libraries to identify specific peptide sequences against the desired target (76, 77). Wang et al. synthesized a linear peptide with sequence PYRWQLWWHNWS selected based on the screening of a randomized 12-mer peptide library against the target Aβ1–10 sequence (78). After screening, specific phages were selected and their binding affinity to Aβ1–10 was evaluated by real-time biomolecular interaction analysis. This peptide could specifically bind to Aβ1–10, inhibit the aggregation of Aβ into plaques, and reduce Aβ1–42 induced-apoptosis. Furthermore, it illustrated a protective effect against Aβ1–42-induced memory and learning impairments in animal models (59).

Larbanoix et al. utilized the phage display method to discover a linear hexapeptide against Aβ1–42 aggregation (79). Two of the selected clones, Pep1: LIAIMA and Pep2: IFALMG, corresponding to fragment 31IIGLMV36 from Aβ1–42 peptide, demonstrated the highest binding affinities to Aβ1–42 with Kd values in the micromolar range. Their specific interactions with Aβ1–42 plaques were identified by immunohistochemistry on harvested brain tissue from an animal model of Alzheimer’s disease. The peptides did not induce any toxicity in neurons in vitro. Moreover, the thioflavin T aggregation assay indicated that the designed peptides could suppress the amyloid fibril formation.

In 2010, a random heptapeptide library (XX–P–XXXX) on T7 phage was reported by Kawasaki et al. (80). The library was designed based on the LPFFD sequence XX-P-XXXX, where P is proline, and X is any amino acid. After the fifth-round of biopanning against Aβ1–42 soluble oligomers, eight new peptides containing arginine residues were obtained. The peptide with the strongest affinity to Aβ (RGPRGRV) suppressed the formation of 37–48 kDa oligomers and maintained the monomeric form of Aβ1–42 for up to 24 h. In follow-up studies, to assess the effect of the peptide length on the inhibition of soluble oligomers formation, random libraries containing 3-residue and 4-residue peptides were prepared by phage display and evaluated. The results demonstrated that the 3-residue peptides could not significantly inhibit oligomers formation because of their smaller size. In contrast, the 4-residue peptide with the RFRK sequence inhibited the soluble oligomer formation like the heptapeptide (RGPRGRV). It also showed a slight decrease on Aβ fibrillation (81), similar to the inhibitory activity of the N-Methylated Peptide (SEN304), against Aβ42 aggregation (25, 82, 83). Tsuji-Ueno et al. utilized the all-steps-all-combinations (ASAC) method to explore Aβ42-binding peptide aptamers. The identified peptides from the primary and secondary libraries showed a weak binding affinity to Aβ42 (Kd values in the μ m range) (84). To further improve the peptide aptamers, Gautam et al. applied the mRNA display technique and paired-peptide library method. The library was assembled by a random shuffling method on selected peptide blocks taken from the formed primary and secondary peptide libraries by Tsuji-Ueno et al. (84). They reported two peptides with high binding affinity to Aβ42 (Kd in the nM range) which significantly inhibited the Aβ42 aggregation (85). The improved peptide aptamers, P84 (CGILDPIPWGGSGGSCGILDPIPW) and P131 (GCPCIGIIGGSGGSDCSSDLTPS), where GGSGGS is the linker sequence, demonstrated a higher binding affinity for the Aβ42 peptide (Kd values in the nanomolar range) compared to the primary and secondary Aβ42-binding peptides (86). The results showed that both peptides could inhibit the Aβ42 aggregation and result in the reduction of the cytotoxic effects of Aβ42 fibrils and Aβ42 oligomers in PC12 cells; P84 showed better efficacy than P131 on the cell line.

Groen and co-workers employed mirror-image phage display to identify selective and high-affinity D-peptide ligands for Aβ1–42. The D-enantiomer Aβ1–42 was used as a target for selection from a randomized 12- amino acid peptide library with more than 1 billion different peptides. After six rounds of biopanning, they identified a specific D-enantiomeric peptide, RPRTRLHTHRNR, called D3 (73). The D3 ligand inhibited Aβ aggregation, and dissolved pre-formed Aβ fibrils. Additionally, D3 ligand could disaggregate pre-existing amyloid plaques in the brain and result in an increase in the amount of Aβ monomeric form, which has high clearance from the brain (87). FITC fluorescence data demonstrated that Aβ–D3 clearance might have been associated with pericytes, which have a major role in the clearance of different Aβ40/42 species (88, 89). Glial fibrillary acidic protein (GFAP) staining of astrocytes and CD11b staining of microglia in brain sections revealed that the D3 significantly decreased the amount of plaque-related inflammation markers (active astrocytes and microglia) around the Aβ plaques in comparison to the untreated animals. In addition to the anti-inflammatory properties, this peptide ligand could drastically reduce the Aβ plaque load in brain tissue of transgenic APP-PSD mice after a 30-day treatment with administration of 9 mg D3 per day per mouse. Computational simulation studies demonstrated strong electrostatic interactions between the arginine-rich D3 and negatively charged groups of Aβ nonamer; D3 binding to Aβ nonamer could change the topology of the Aβ oligomers by inducing a twist in them and consequently promote the formation of Aβ nonfibrillar aggregations (73, 90, 91).

Luo et al. applied peptoid chemistry, N-substituted glycine oligomers as a class of peptidomimetics, to develop and improve selective high-affinity ligands for Aβ42 (92). They constructed an on-bead peptoid library of 38,416 unique peptoids. After screening for Aβ42-selective peptoid ligands, the IAM1 ligand and its dimeric form were selected and further evaluated. IAM1 peptide showed about 10-fold more affinity for Aβ42-binding than for Aβ40, and inhibited Aβ42 aggregation in vitro. The dimeric derivative (IAM1)2 demonstrated a 7.4-fold higher affinity for Aβ42 (60 nM) than the monomeric form. Moreover, (IAM1)2 demonstrated neuroprotective effects on primary hippocampal neurons against Aβ42-induced toxicity.

Due to the considerable similarities between the self-assembly of cyclic d,l-α-peptides and amyloid structures, it is possible such peptides can bind to Aβ non-toxic forms and stabilize them (29). Richman et al. described the cyclic peptide CP-2, cyclo-[l–J–w– H–s–K]s (J denotes l-norleucine), by screening a 6-residue library of head-to-tail cyclic d,l-α-peptides consisting of residues Lys, Glu, Ser, Leu, Trp, and His using a one-bead-one-peptide combinatorial approach (29, 93). The selected peptide strongly interacted with Aβ40/Aβ42 and prevented their assembly, entirely disassembled Aβ40 fibrils, and protected PC12 cells against Aβ40/Aβ42-induced toxicity, without having any toxic effects of its own. NMR spectroscopy revealed that the CP-2 peptide, in a self-assembled form, interacted with monomeric and low-oligomeric structures of Aβ40 and induced weak α-helix structures during the initial stage of Aβ40 aggregation and subsequently promoted the conformational transition shift from a more toxic antiparallel β-sheet conformation to the less toxic parallel β-sheet.

In another study, Acerra et al. utilized an intracellular protein-fragment complementation assay (PCA) methodology for the screening of selective high-affinity peptides to Aβ (94). The Aβ25–35 sequence, known to self-assemble into toxic fibrils (95), was inserted into one half of the murine dihydrofolate reductase enzyme as a target, and the Aβ29–35 sequence-based peptide was inserted on the other half of the enzyme (96). After the screening of primary and secondary libraries, two new targeting peptides L2P1, FSKATSN, and L2P2, PVKATTA were selected. These peptides shared no homology with the starting template Aβ29–35. The results showed that all selected peptides could bind Aβ42, inhibit fibril formation, and disaggregate pre-formed fibrils. To further improve the metabolic stability of selected peptides from primary and secondary libraries, their retro-inverso (RI) analogs were evaluated (86). All RI peptide ligands, such as KAR-R1, L2P1a-RI, L2P1b-RI, L2P2a-RI, and L2P2b-RI, inhibited Aβ fibrillation and disaggregated pre-formed fibrils, and reduced Aβ42-induced toxicity in PC-12 cells.

THE CURRENT STATE AND FUTURE DIRECTIONS OF Aβ INHIBITORY PEPTIDES IN ALZHEIMER’S DISEASE

A wide range of peptide-based inhibitors has been evaluated in cellular and animal models as new therapeutic compounds for inhibition of Aβ aggregation. While experimental studies generated promising results, only a few of these inhibitory peptides have been successful enough to enter clinical trials. NAP or Davunetide peptide with NAPVSIPQ sequence, derived from the activity-dependent neuroprotective protein (ADNP), was reported in 2003 by Gozes et al. (97). NAP was able to inhibit Aβ aggregation, disassemble pre-formed fibrils, and protect the neuronal cells from Aβ induced toxicity. Though NAP demonstrated benefits in phase II clinical trials for mild cognitive impairment, it failed in a phase III trial (98100). PPI-1019 peptide (APAN), with a sequence of D-(H-[(Me-L)-VFFL]NH2), is an N-methylated peptide which is derived from the D-enantiomeric Cholyl-LVFFA-NH2 that could inhibit Aβ aggregation and the induce toxicity in experimental studies (101). The phases I and II clinical trials of APAN was completed in patients with mild-moderate Alzheimer’s disease in 2005 (NCT00100282, NCT00100334), but the outcome of this study is still unknown (https://clinicaltrials.gov/, last assessed 28 October 2020). Other reported inhibitory peptides including, D3 (102), D-Trp-Aib−OH (44), D-4F (36), TAT-R1-OR2 (48), NL-RI-OR2-TAT90 (103), and R1-OR2 (49) have shown considerable efficacy in preclinical trials, but they have not yet entered clinical trials.

The reality is that, to date, there is no cure for Alzheimer’s disease. Only optimism remains. Therefore, it is necessary to discover potential peptides for testing in clinical trials. The current inhibitory peptides have certain limitations such as poor BBB permeability and high cytotoxicity. To overcome these problems and further improve the inhibitory activity, a number of studies have focused on peptide–nanostructure conjugates (PNCs) approach that provides an opportunity to increase the capabilities of both these classes of materials (55, 104, 105). Nanostructures could be considered as a potential vehicle to overcome poor BBB permeability and bring hope for neurodegenerative diseases therapy due to their size and various surface modifications (106). As an example, multivalent inhibitors can be developed against Aβ aggregation by decorating gold nanoparticles with VVIA and LPFFD (107). The PNCs approach gives a fascinating insight into the fields of diagnosis and treatment, and provides new opportunities for the design of high-performance peptides (108, 109).

CONCLUSION

Despite a better understanding of the pathogenic mechanisms of Alzheimer’s disease, finding efficient therapeutic compounds to prevent or halt the progression of Alzheimer’s disease continues to be a challenge. Aβ aggregation inhibition-based approaches are being developed with the aim to stop disease progression. While the reported inhibitory peptides have considerable advantages over other compounds, and experimental evidence has been encouraging, bench-to-bedside has not yet become a reality. Therefore, adequate knowledge of binding interactions of these peptides with their biological targets, the ligand-target complex, is required to design more accurate therapeutic biomolecules. Peptide inhibitors have unique properties, particularly, high selectivity, low accumulation in tissues, low side-effects and toxicity, and different chemical and biological synthesis routes when compared with other compounds. As researchers continue to focus on rational design, characterization, optimization, and interaction between the inhibitor and the Aβ peptide complex, more peptide inhibitors are expected to succeed in clinical trials.

Conflict of Interest: The authors declare no potential conflicts of interest with respect to research, authorship and/or publication of this chapter.

Copyright and Permission Statement: We confirm that the materials included in this chapter do not violate copyright laws. Where relevant, appropriate permissions have been obtained from the original copyright holder(s). All original sources have been appropriately acknowledged and/or referenced.

REFERENCES

  1. 2009 Alzheimer’s disease facts and figures. Alzheimer’s & Dementia. 2009;5(3):234–70. https://doi.org/10.1016/j.jalz.2009.03.001
  2. 2018 Alzheimer’s disease facts and figures. Alzheimer’s Dementia. 2018;14(3):367–429. https://doi.org/10.1016/j.jalz.2018.02.001
  3. Prince M, Comas-Herrera A, Knapp M, Guerchet M, Karagiannidou M. World Alzheimer report 2016: improving healthcare for people living with dementia: coverage, quality and costs now and in the future. Alzheimer’s Disease International (ADI); London. 2016.
  4. Bartus RT, Emerich DF. Cholinergic markers in Alzheimer disease. JAMA. 1999;282(23):2208–9. https://doi.org/10.1001/jama.282.23.2208
  5. Mandelkow E-M, Mandelkow E. Tau in Alzheimer’s disease. Trends Cell Biol. 1998;8(11):425–7. https://doi.org/10.1016/S0962-8924(98)01368-3
  6. Finder VH, Glockshuber R. Amyloid-β aggregation. Neurodegen Dis. 2007;4(1):13–27. https://doi.org/10.1159/000100355
  7. Hardy J, Selkoe DJ. The amyloid hypothesis of Alzheimer’s disease: Progress and problems on the road to therapeutics. Science. 2002;297(5580):353–6. https://doi.org/10.1126/science.1072994
  8. Singh M, Singh S, Prasad S, Gambhir I. Nanotechnology in medicine and antibacterial effect of silver nanoparticles. Dig J Nanomater Biostruct. 2008;3(3):115–22.
  9. Haass C, Selkoe DJ. Cellular processing of β-amyloid precursor protein and the genesis of amyloid β-peptide. Cell. 1993;75(6):1039–42. https://doi.org/10.1016/0092-8674(93)90312-E
  10. Hall D, Edskes H. A model of amyloid’s role in disease based on fibril fracture. Biophys Chem. 2009;145(1):17–28. https://doi.org/10.1016/j.bpc.2009.08.004
  11. Hall D, Edskes H. Computational modeling of the relationship between amyloid and disease. Biophys Rev. 2012;4(3):205–22. https://doi.org/10.1007/s12551-012-0091-x
  12. Liu H, Dong X, Liu F, Zheng J, Sun Y. Iminodiacetic acid-conjugated nanoparticles as a bifunctional modulator against Zn2+-mediated amyloid β-protein aggregation and cytotoxicity. J Colloid Interf Sci. 2017;505:973–82. https://doi.org/10.1016/j.jcis.2017.06.093
  13. Wang W, Han Y, Fan Y, Wang Y. Effects of gold nanospheres and nanocubes on amyloid-β peptide fibrillation. Langmuir. 2019;35(6):2334–42. https://doi.org/10.1021/acs.langmuir.8b04006
  14. Brahmachari S, Paul A, Segal D, Gazit E. Inhibition of amyloid oligomerization into different supramolecular architectures by small molecules: Mechanistic insights and design rules. Fut Med Chem. 2017;9(8):797–810. https://doi.org/10.4155/fmc-2017-0026
  15. Jokar S, Khazaei S, Behnammanesh H, Shamloo A, Erfani M, Beiki D, et al. Recent advances in the design and applications of amyloid-β peptide aggregation inhibitors for Alzheimer’s disease therapy. Biophys Rev. 2019;11:1–25. https://doi.org/10.1007/s12551-019-00606-2
  16. Rahimi F, Li H, Sinha S, Bitan G. Modulators of amyloid β-protein (Aβ) self-assembly. Developing Therapeutics for Alzheimer’s Disease. Elsevier; Boston. 2016. p. 97–191. https://doi.org/10.1016/B978-0-12-802173-6.00006-X
  17. Danho W, Swistok J, Khan W, Chu X-J, Cheung A, Fry D, et al. Opportunities and challenges of developing peptide drugs in the pharmaceutical industry. Peptides for Youth. Springer; New York. 2009. p. 467–9. https://doi.org/10.1007/978-0-387-73657-0_201
  18. Bruno BJ, Miller GD, Lim CS. Basics and recent advances in peptide and protein drug delivery. Therap Deliv. 2013;4(11):1443–67. https://doi.org/10.4155/tde.13.104
  19. Goyal D, Shuaib S, Mann S, Goyal B. Rationally designed peptides and peptidomimetics as inhibitors of amyloid-β (Aβ) aggregation: Potential therapeutics of Alzheimer’s disease. ACS Combinator Sci. 2017;19(2):55–80. https://doi.org/10.1021/acscombsci.6b00116
  20. Jokar S, Behnammanesh H, Erfani M, Sharifzadeh M, Gholami M, Sabzevari O, et al. Synthesis, biological evaluation and preclinical study of a novel 99mTc-peptide: A targeting probe of amyloid-β plaques as a possible diagnostic agent for Alzheimer’s disease. Bioorgan Chem. 2020;99:103857. https://doi.org/10.1016/j.bioorg.2020.103857
  21. Lührs T, Ritter C, Adrian M, Riek-Loher D, Bohrmann B, Döbeli H, et al. 3D structure of Alzheimer’s amyloid-β (1–42) fibrils. Proc Natl Acad Sci. 2005;102(48):17342–7. https://doi.org/10.1073/pnas.0506723102
  22. Jokar S, Erfani M, Bavi O, Khazaei S, Sharifzadeh M, Hajiramezanali M, et al. Design of peptide-based inhibitor agent against amyloid-β aggregation: Molecular docking, synthesis and in vitro evaluation. Bioorgan Chem. 2020;102:104050. https://doi.org/10.1016/j.bioorg.2020.104050
  23. Reddy G, Straub JE, Thirumalai D. Influence of preformed asp23− lys28 salt bridge on the conformational fluctuations of monomers and dimers of Aβ peptides with implications for rates of fibril formation. J Phys Chem B. 2009;113(4):1162–72. https://doi.org/10.1021/jp808914c
  24. Friedemann M, Helk E, Tiiman A, Zovo K, Palumaa P, Tõugu V. Effect of methionine-35 oxidation on the aggregation of amyloid-β peptide. Biochem Biophys Rep. 2015;3:94–9. https://doi.org/10.1016/j.bbrep.2015.07.017
  25. Li H, Du Z, Lopes DH, Fradinger EA, Wang C, Bitan G. C-terminal tetrapeptides inhibit Aβ42-induced neurotoxicity primarily through specific interaction at the N-terminus of Aβ42. J Med Chem. 2011;54(24):8451–60. https://doi.org/10.1021/jm200982p
  26. Urbanc B, Cruz L, Yun S, Buldyrev SV, Bitan G, Teplow DB, et al. In silico study of amyloid β-protein folding and oligomerization. Proc Natl Acad Sci. 2004;101(50):17345–50. https://doi.org/10.1073/pnas.0408153101
  27. Liu J, Wang W, Zhang Q, Zhang S, Yuan Z. Study on the efficiency and interaction mechanism of a decapeptide inhibitor of β-amyloid aggregation. Biomacromolecules. 2014;15(3):931–9. https://doi.org/10.1021/bm401795e
  28. Jagota S, Rajadas J. Synthesis of d-amino acid peptides and their effect on beta-amyloid aggregation and toxicity in transgenic Caenorhabditis elegans. Med Chem Res. 2013;22(8):3991–4000. https://doi.org/10.1007/s00044-012-0386-2
  29. Richman M, Wilk S, Chemerovski M, Wärmländer SK, Wahlström A, Gräslund A, et al. In vitro and mechanistic studies of an antiamyloidogenic self-assembled cyclic d, l-α-peptide architecture. J Am Chem Soc. 2013;135(9):3474–84. https://doi.org/10.1021/ja310064v
  30. Brener O, Dunkelmann T, Gremer L, Van Groen T, Mirecka EA, Kadish I, et al. QIAD assay for quantitating a compound’s efficacy in elimination of toxic Aβ oligomers. Sci Rep. 2015;5:13222. https://doi.org/10.1038/srep13222
  31. Leithold LH, Jiang N, Post J, Niemietz N, Schartmann E, Ziehm T, et al. Pharmacokinetic properties of tandem d-peptides designed for treatment of Alzheimer’s disease. Eur J Pharmaceut Sci. 2016;89:31–8. https://doi.org/10.1016/j.ejps.2016.04.016
  32. Klein AN, Ziehm T, Tusche M, Buitenhuis J, Bartnik D, Boeddrich A, et al. Optimization of the all-D peptide D3 for Aβ oligomer elimination. PLoS One. 2016;11(4):e0153035. https://doi.org/10.1371/journal.pone.0153035
  33. Zhang YX, Wang SW, Lu S, Zhang LX, Liu DQ, Ji M, et al. A mimotope of Aβ oligomers may also behave as a β-sheet inhibitor. FEBS Lett. 2017;591(21):3615–24. https://doi.org/10.1002/1873-3468.12871
  34. Granic I, Masman MF, Nijholt IM, Naude PJ, de Haan A, Borbély E, et al. LPYFDa neutralizes amyloid-β-induced memory impairment and toxicity. J Alzheimers Dis. 2010;19(3):991–1005. https://doi.org/10.3233/JAD-2010-1297
  35. Minicozzi V, Chiaraluce R, Consalvi V, Giordano C, Narcisi C, Punzi P, et al. Computational and experimental studies on β-sheet breakers targeting Aβ1–40 fibrils. J Biol Chem. 2014;289(16):11242–52. https://doi.org/10.1074/jbc.M113.537472
  36. Handattu SP, Garber DW, Monroe CE, van Groen T, Kadish I, Nayyar G, et al. Oral apolipoprotein AI mimetic peptide improves cognitive function and reduces amyloid burden in a mouse model of Alzheimer’s disease. Neurobiol Dis. 2009;34(3):525–34. https://doi.org/10.1016/j.nbd.2009.03.007
  37. Kino R, Araya T, Arai T, Sohma Y, Kanai M. Covalent modifier-type aggregation inhibitor of amyloid-β based on a cyclo-KLVFF motif. Bioorgan Med Chem Lett. 2015;25(15):2972–5. https://doi.org/10.1016/j.bmcl.2015.05.027
  38. Sohma Y. Medicinal chemistry focusing on aggregation of amyloid-β. Chem Pharmaceut Bull. 2016;64(1):1–7. https://doi.org/10.1248/cpb.c15-00742
  39. Luheshi LM, Hoyer W, de Barros TP, van Dijk Härd I, Brorsson A-C, Macao B, et al. Sequestration of the Aβ peptide prevents toxicity and promotes degradation in vivo. PLoS Biol. 2010;8(3):e1000334.
  40. Lin L-X, Bo X-Y, Tan Y-Z, Sun F-X, Song M, Zhao J, et al. Feasibility of β-sheet breaker peptide-H102 treatment for Alzheimer’s disease based on β-amyloid hypothesis. PLoS One. 2014;9(11):e112052. https://doi.org/10.1371/journal.pone.0112052
  41. Sinopoli A, Giuffrida A, Tomasello MF, Giuffrida ML, Leone M, Attanasio F, et al. Ac-LPFFD-Th: A trehalose-conjugated peptidomimetic as a strong suppressor of amyloid-β oligomer formation and cytotoxicity. ChemBioChem. 2016;17(16):1541–9. https://doi.org/10.1002/cbic.201600243
  42. Wei C-W, Peng Y, Zhang L, Huang Q, Cheng M, Liu Y-N, et al. Synthesis and evaluation of ferrocenoyl pentapeptide (Fc-KLVFF) as an inhibitor of Alzheimer’s Aβ1–42 fibril formation in vitro. Bioorgan Med Chem Lett. 2011;21(19):5818–21. https://doi.org/10.1016/j.bmcl.2011.07.111
  43. Pratim Bose P, Chatterjee U, Nerelius C, Govender T, Norstrom T, Gogoll A, et al. Poly-N-methylated amyloid β-peptide (Aβ) C-terminal fragments reduce Aβ toxicity in vitro and in Drosophila melanogaster. J Med Chem. 2009;52(24):8002–9. https://doi.org/10.1021/jm901092h
  44. Frydman-Marom A, Rechter M, Shefler I, Bram Y, Shalev DE, Gazit E. Cognitive-performance recovery of Alzheimer’s disease model mice by modulation of early soluble amyloidal assemblies. Angewandte Chem. 2009;121(11):2015–20. https://doi.org/10.1002/ange.200802123
  45. Tanaka F, Shibata K, Monobe Y, Akagi K-I, Masuda Y. Design and synthesis of β-strand-fixed peptides inhibiting aggregation of amyloid β-protein. Bioorgan Med Chem. 2020;28(18):115676. https://doi.org/10.1016/j.bmc.2020.115676
  46. Horsley JR, Jovcevski B, Wegener KL, Yu J, Pukala TL, Abell AD. Rationally designed peptide-based inhibitor of Aβ42 fibril formation and toxicity: A potential therapeutic strategy for Alzheimer’s disease. Biochem J. 2020;477(11):2039–54. https://doi.org/10.1042/BCJ20200290
  47. Chorev M, Goodman M. Recent developments in retro peptides and proteins – An ongoing topochemical exploration. Trends Biotechnol. 1995;13(10):438–45. https://doi.org/10.1016/S0167-7799(00)88999-4
  48. Parthsarathy V, McClean PL, Hölscher C, Taylor M, Tinker C, Jones G, et al. A novel retro-inverso peptide inhibitor reduces amyloid deposition, oxidation and inflammation and stimulates neurogenesis in the APPswe/PS1ΔE9 mouse model of Alzheimer’s disease. PLoS One. 2013;8(1):e54769. https://doi.org/10.1371/journal.pone.0054769
  49. Taylor M, Moore S, Mayes J, Parkin E, Beeg M, Canovi M, et al. Development of a proteolytically stable retro-inverso peptide inhibitor of β-amyloid oligomerization as a potential novel treatment for Alzheimer’s disease. Biochemistry. 2010;49(15):3261–72. https://doi.org/10.1021/bi100144m
  50. Loureiro JA, Crespo R, Börner H, Martins PM, Rocha FA, Coelho M, et al. Fluorinated beta-sheet breaker peptides. J Mater Chem B. 2014;2(16):2259–64. https://doi.org/10.1039/C3TB21483D
  51. Luo J, Abrahams JP. Cyclic peptides as inhibitors of amyloid fibrillation. Chem A Eur J. 2014;20(9):2410–19. https://doi.org/10.1002/chem.201304253
  52. Han X, He G. Toward a rational design to regulate β-amyloid fibrillation for alzheimer’s disease treatment. ACS Chem Neurosci. 2018;9(2):198–210. https://doi.org/10.1021/acschemneuro.7b00477
  53. Tjernberg LO, Näslund J, Lindqvist F, Johansson J, Karlström AR, Thyberg J, et al. Arrest of-amyloid fibril formation by a pentapeptide ligand. J Biol Chem. 1996;271(15):8545–8. https://doi.org/10.1074/jbc.271.15.8545
  54. Matsunaga Y, Fujii A, Awasthi A, Yokotani J, Takakura T, Yamada T. Eight-residue Aβ peptides inhibit the aggregation and enzymatic activity of Aβ42. Regul Peptides. 2004;120(1–3):227–36. https://doi.org/10.1016/j.regpep.2004.03.013
  55. Xiong N, Dong X-Y, Zheng J, Liu F-F, Sun Y. Design of LVFFARK and LVFFARK-functionalized nanoparticles for inhibiting amyloid β-protein fibrillation and cytotoxicity. ACS Appl Mater Interf. 2015;7(10):5650–62. https://doi.org/10.1021/acsami.5b00915
  56. Zhang H, Dong X, Liu F, Zheng J, Sun Y. Ac-LVFFARK-NH2 conjugation to β-cyclodextrin exhibits significantly enhanced performance on inhibiting amyloid β-protein fibrillogenesis and cytotoxicity. Biophys Chem. 2018;235:40–7. https://doi.org/10.1016/j.bpc.2018.02.002
  57. Zhang H, Zhang C, Dong XY, Zheng J, Sun Y. Design of nonapeptide LVFFARKHH: A bifunctional agent against Cu2+-mediated amyloid β-protein aggregation and cytotoxicity. J Mol Recogn. 2018;31(6):e2697. https://doi.org/10.1002/jmr.2697
  58. Soto C, Sigurdsson EM, Morelli L, Kumar RA, Castaño EM, Frangione B. β-sheet breaker peptides inhibit fibrillogenesis in a rat brain model of amyloidosis: Implications for Alzheimer’s therapy. Nat Med. 1998;4(7):822–6. https://doi.org/10.1038/nm0798-822
  59. Wood SJ, Wetzel R, Martin JD, Hurle MR. Prolines and aamyloidogenicity in fragments of the Alzheimer’s peptide. beta./A4. Biochemistry. 1995;34(3):724–30. https://doi.org/10.1021/bi00003a003
  60. Cruz M, Tusell J, Grillo-Bosch D, Albericio F, Serratosa J, Rabanal F, et al. Inhibition of β-amyloid toxicity by short peptides containing N-methyl amino acids. J Peptide Res. 2004;63(3):324–8. https://doi.org/10.1111/j.1399-3011.2004.00156.x
  61. Ramaswamy K, Kumaraswamy P, Sethuraman S, Krishnan UM. Self-assembly characteristics of a structural analogue of Tjernberg peptide. RSC Adv. 2014;4(32):16517–23. https://doi.org/10.1039/C3RA47754A
  62. Cimini S, Sclip A, Mancini S, Colombo L, Messa M, Cagnotto A, et al. The cell-permeable Aβ1-6A2VTAT (D) peptide reverts synaptopathy induced by Aβ1-42wt. Neurobiol Dis. 2016;89:101–11. https://doi.org/10.1016/j.nbd.2015.12.013
  63. Di Fede G, Catania M, Morbin M, Rossi G, Suardi S, Mazzoleni G, et al. A recessive mutation in the APP gene with dominant-negative effect on amyloidogenesis. Science. 2009;323(5920):1473–7. https://doi.org/10.1126/science.1168979
  64. Di Fede G, Catania M, Maderna E, Morbin M, Moda F, Colombo L, et al. Tackling amyloidogenesis in Alzheimer’s disease with A2V variants of Amyloid-β. Sci Rep. 2016;6:20949. https://doi.org/10.1038/srep20949
  65. Aoraha E, Candreva J, Kim JR. Engineering of a peptide probe for β-amyloid aggregates. Mol BioSyst. 2015;11(8):2281–9. https://doi.org/10.1039/C5MB00280J
  66. Austen BM, Paleologou KE, Ali SA, Qureshi MM, Allsop D, El-Agnaf OM. Designing peptide inhibitors for oligomerization and toxicity of Alzheimer’s β-amyloid peptide. Biochemistry. 2008;47(7):1984–92. https://doi.org/10.1021/bi701415b
  67. Fradinger EA, Monien BH, Urbanc B, Lomakin A, Tan M, Li H, et al. C-terminal peptides coassemble into Aβ42 oligomers and protect neurons against Aβ42-induced neurotoxicity. Proc Natl Acad Sci. 2008;105(37):14175–80. https://doi.org/10.1073/pnas.0807163105
  68. Zheng X, Wu C, Liu D, Li H, Bitan G, Shea J-E, et al. Mechanism of C-terminal fragments of amyloid β-protein as Aβ inhibitors: Do C-terminal interactions play a key role in their inhibitory activity? J Phys Chem B. 2016;120(8):1615–23. https://doi.org/10.1021/acs.jpcb.5b08177
  69. Bansal S, Maurya IK, Yadav N, Thota CK, Kumar V, Tikoo K, et al. C-terminal fragment, Aβ32–37, analogues protect against aβ aggregation-induced toxicity. ACS Chem Neurosci. 2016;7(5):615–23. https://doi.org/10.1021/acschemneuro.6b00006
  70. Kawashima H, Sohma Y, Nakanishi T, Kitamura H, Mukai H, Yamashita M, et al. A new class of aggregation inhibitor of amyloid-β peptide based on an O-acyl isopeptide. Bioorgan Med Chem. 2013;21(21):6323–7. https://doi.org/10.1016/j.bmc.2013.08.062
  71. Aloisi A, Barca A, Romano A, Guerrieri S, Storelli C, Rinaldi R, et al. Anti-aggregating effect of the naturally occurring dipeptide carnosine on aβ1-42 fibril formation. PLoS One. 2013;8(7):e68159. https://doi.org/10.1371/journal.pone.0068159
  72. Wiesehan K, Buder K, Linke RP, Patt S, Stoldt M, Unger E, et al. Selection of D-amino-acid peptides that bind to Alzheimer’s disease amyloid peptide Aβ1–42 by mirror image phage display. Chembiochem. 2003;4(8):748–53. https://doi.org/10.1002/cbic.200300631
  73. Van Groen T, Wiesehan K, Funke SA, Kadish I, Nagel-Steger L, Willbold D. Reduction of Alzheimer’s disease amyloid plaque load in transgenic mice by D3, A D-enantiomeric peptide identified by mirror image phage display. ChemMedChem. 2008;3(12):1848. https://doi.org/10.1002/cmdc.200800273
  74. Biyani M, Nishigaki K, Biyani M. Biomolecular display technology: A new tool for drug discovery. Animal Biotechnology: Elsevier; 2014. p. 369–84. https://doi.org/10.1016/B978-0-12-416002-6.00020-1
  75. Zwick MB, Shen J, Scott JK. Phage-displayed peptide libraries. Curr Opin Biotechnol. 1998;9(4):427–36. https://doi.org/10.1016/S0958-1669(98)80017-7
  76. Kang CK, Jayasinha V, Martin PT. Identification of peptides that specifically bind Aβ1–40 amyloid in vitro and amyloid plaques in Alzheimer’s disease brain using phage display. Neurobiol Dis. 2003;14(1):146–56. https://doi.org/10.1016/S0969-9961(03)00105-0
  77. Smith GP. Filamentous fusion phage: Novel expression vectors that display cloned antigens on the virion surface. Science. 1985;228(4705):1315–17. https://doi.org/10.1126/science.4001944
  78. Wang F, Zhou X-L, Yang Q-G, Xu W-H, Wang F, Chen Y-P, et al. A peptide that binds specifically to the β-amyloid of Alzheimer’s disease: Selection and assessment of anti-β-amyloid neurotoxic effects. PLoS One. 2011;6(11):e27649. https://doi.org/10.1371/journal.pone.0027649
  79. Larbanoix L, Burtea C, Ansciaux E, Laurent S, Mahieu I, Vander Elst L, et al. Design and evaluation of a 6-mer amyloid-beta protein derived phage display library for molecular targeting of amyloid plaques in Alzheimer’s disease: Comparison with two cyclic heptapeptides derived from a randomized phage display library. Peptides. 2011;32(6):1232–43. https://doi.org/10.1016/j.peptides.2011.04.026
  80. Kawasaki T, Onodera K, Kamijo S. Selection of peptide inhibitors of soluble Aβ1-42 oligomer formation by phage display. Biosci Biotechnol Biochem. 2010;74;2214–19. https://doi.org/10.1271/bbb.100388
  81. Kawasaki T, Onodera K, Kamijo S. Identification of novel short peptide inhibitors of soluble 37/48 kDa oligomers of amyloid β42. Biosci Biotechnol Biochem. 2011;75(8):1496–501. https://doi.org/10.1271/bbb.110198
  82. Kokkoni N, Stott K, Amijee H, Mason JM, Doig AJ. N-methylated peptide inhibitors of β-amyloid aggregation and toxicity. Optimization of the inhibitor structure. Biochemistry. 2006;45(32):9906–18. https://doi.org/10.1021/bi060837s
  83. Zheng X, Liu D, Klärner F-G, Schrader T, Bitan G, Bowers MT. Amyloid β-protein assembly: The effect of molecular tweezers CLR01 and CLR03. J Phys Chem B. 2015;119(14):4831–41. https://doi.org/10.1021/acs.jpcb.5b00692
  84. Tsuji-Ueno S, Komatsu M, Iguchi K, Takahashi M, Yoshino S, Suzuki M, et al. Novel high-affinity Aβ-binding peptides identified by an advanced in vitro evolution, progressive library method. Protein Peptide Lett. 2011;18(6):642–50. https://doi.org/10.2174/092986611795222678
  85. Ghimire Gautam S, Komatsu M, Nishigaki K. Strong inhibition of beta-amyloid peptide aggregation realized by two-steps evolved peptides. Chem Biol Drug Design. 2015;85(3):356–68. https://doi.org/10.1111/cbdd.12400
  86. Acerra N, Kad NM, Griffith DA, Ott S, Crowther DC, Mason JM. Retro-inversal of intracellular selected β-amyloid-interacting peptides: Implications for a novel Alzheimer’s disease treatment. Biochemistry. 2014;53(13):2101–11. https://doi.org/10.1021/bi5001257
  87. Tanzi R, Moir R, Wagner S. Clearance of Alzheimer’s Aβ peptide: The many roads to perdition. Neuron. 2004;43(5):605–8. https://doi.org/10.1016/S0896-6273(04)00533-1
  88. Sagare AP, Bell RD, Zhao Z, Ma Q, Winkler EA, Ramanathan A, et al. Pericyte loss influences Alzheimer-like neurodegeneration in mice. Nat Commun. 2013;4(1):1–14. https://doi.org/10.1038/ncomms3932
  89. Ma Q, Zhao Z, Sagare AP, Wu Y, Wang M, Owens NC, et al. Blood-brain barrier-associated pericytes internalize and clear aggregated amyloid-β42 by LRP1-dependent apolipoprotein E isoform-specific mechanism. Mol Neurodegen. 2018;13(1):1–13. https://doi.org/10.1186/s13024-018-0286-0
  90. Funke SA, van Groen T, Kadish I, Bartnik D, Nagel-Steger L, Brener O, et al. Oral treatment with the D-enantiomeric peptide D3 improves the pathology and behavior of Alzheimer’s disease transgenic mice. ACS Chem Neurosci. 2010;1(9):639–48. https://doi.org/10.1021/cn100057j
  91. Funke SA, Willbold D. Mirror image phage display – A method to generate D-peptide ligands for use in diagnostic or therapeutical applications. Mol BioSyst. 2009;5(8):783–6. https://doi.org/10.1039/b904138a
  92. Luo Y, Vali S, Sun S, Chen X, Liang X, Drozhzhina T, et al. Aβ42-binding peptoids as amyloid aggregation inhibitors and detection ligands. ACS Chem Neurosci. 2013;4(6):952–62. https://doi.org/10.1021/cn400011f
  93. Lam KS, Salmon SE, Hersh EM, Hruby VJ, Kazmierski WM, Knapp RJ. A new type of synthetic peptide library for identifying ligand-binding activity. Nature. 1991;354(6348):82–4. https://doi.org/10.1038/354082a0
  94. Remy I, Campbell-Valois F, Michnick SW. Detection of protein–protein interactions using a simple survival protein-fragment complementation assay based on the enzyme dihydrofolate reductase. Nature Protocols. 2007;2(9):2120. https://doi.org/10.1038/nprot.2007.266
  95. Pike CJ, Walencewicz-Wasserman AJ, Kosmoski J, Cribbs DH, Glabe CG, Cotman CW. Structure-activity analyses of β-amyloid peptides: Contributions of the β25–35 region to aggregation and neurotoxicity. J Neurochem. 1995;64(1):253–65. https://doi.org/10.1046/j.1471-4159.1995.64010253.x
  96. Acerra N, Kad NM, Mason JM. Combining intracellular selection with protein-fragment complementation to derive Aβ interacting peptides. Protein Eng Design Select. 2013;26(7):463–70. https://doi.org/10.1093/protein/gzt021
  97. Ashur-Fabian O, Segal-Ruder Y, Skutelsky E, Brenneman DE, Steingart RA, Giladi E, et al. The neuroprotective peptide NAP inhibits the aggregation of the beta-amyloid peptide. Peptides. 2003;24(9):1413–23. https://doi.org/10.1016/j.peptides.2003.08.005
  98. Boxer AL, Lang AE, Grossman M, Knopman DS, Miller BL, Schneider LS, et al. Davunetide in patients with progressive supranuclear palsy: A randomised, double-blind, placebo-controlled phase 2/3 trial. Lancet Neurol. 2014;13(7):676–85. https://doi.org/10.1016/S1474-4422(14)70088-2
  99. Matsuoka Y, Jouroukhin Y, Gray AJ, Ma L, Hirata-Fukae C, Li H-F, et al. A neuronal microtubule-interacting agent, NAPVSIPQ, reduces tau pathology and enhances cognitive function in a mouse model of Alzheimer’s disease. J Pharmacol Exp Therapeut. 2008;325(1):146–53. https://doi.org/10.1124/jpet.107.130526
  100. Gozes I, Morimoto BH, Tiong J, Fox A, Sutherland K, Dangoor D, et al. NAP: Research and development of a peptide derived from activity-dependent neuroprotective protein (ADNP). CNS Drug Rev. 2005;11(4):353–68. https://doi.org/10.1111/j.1527-3458.2005.tb00053.x
  101. Findeis MA. Peptide inhibitors of beta amyloid aggregation. Curr Topics Med Chem. 2002;2(4):417–23. https://doi.org/10.2174/1568026024607508
  102. Jiang N, Leithold LH, Post J, Ziehm T, Mauler J, Gremer L, et al. Preclinical pharmacokinetic studies of the tritium labelled D-enantiomeric peptide D3 developed for the treatment of Alzheimer s disease. PLoS One. 2015;10(6):e0128553. https://doi.org/10.1371/journal.pone.0128553
  103. Gregori M, Taylor M, Salvati E, Re F, Mancini S, Balducci C, et al. Retro-inverso peptide inhibitor nanoparticles as potent inhibitors of aggregation of the Alzheimer’s Aβ peptide. Nanomed Nanotechnol Biol Med. 2017;13(2):723–32. https://doi.org/10.1016/j.nano.2016.10.006
  104. Yang J, Liu W, Sun Y, Dong X. LVFFARK-PEG-stabilized black phosphorus nanosheets potently inhibit amyloid-β fibrillogenesis. Langmuir. 2020;36(7):1804–12. https://doi.org/10.1021/acs.langmuir.9b03612
  105. Hao S, Li X, Han A, Yang Y, Fang G, Liu J, et al. CLVFFA-functionalized gold nanoclusters inhibit aβ40 fibrillation, fibrils’ prolongation, and mature fibrils’ disaggregation. ACS Chem Neurosci. 2019;10(11):4633–42. https://doi.org/10.1021/acschemneuro.9b00469
  106. Niu X, Chen J, Gao J. Nanocarriers as a powerful vehicle to overcome blood-brain barrier in treating neurodegenerative diseases: Focus on recent advances. Asian J Pharmaceut Sci. 2019;14(5):480–96. https://doi.org/10.1016/j.ajps.2018.09.005
  107. Xiong N, Zhao Y, Dong X, Zheng J, Sun Y. Design of a molecular hybrid of dual peptide inhibitors coupled on AuNPs for enhanced inhibition of amyloid β-protein aggregation and cytotoxicity. Small. 2017;13(13):1601666. https://doi.org/10.1002/smll.201601666
  108. Jeong W-J, Bu J, Kubiatowicz LJ, Chen SS, Kim Y, Hong S. Peptide–nanoparticle conjugates: A next generation of diagnostic and therapeutic platforms? Nano Convergence. 2018;5(1):38. https://doi.org/10.1186/s40580-018-0170-1
  109. Khafaji M, Zamani M, Golizadeh M, Bavi O. Inorganic nanomaterials for chemo/photothermal therapy: A promising horizon on effective cancer treatment. Biophys Rev. 2019;11(3):335–52. https://doi.org/10.1007/s12551-019-00532-3