Abstract
Alzheimer’s disease (AD) is a progressive neurodegenerative disorder with devastating effects. The greatest risk factor to develop AD is age. Today, only symptomatic therapies are available. Additionally, AD can be diagnosed with certainty only post mortem, whereas the diagnosis “probable AD” can be established earliest when severe clinical symptoms appear. Specific neuropathological changes like neurofibrillary tangles and amyloid plaques define AD. Amyloid plaques are mainly composed of the amyloid-β peptide (Aβ). Several lines of evidence suggest that the progressive concentration and subsequent aggregation and accumulation of Aβ play a fundamental role in the disease progress. Therefore, substances which bind to Aβ and influence aggregation thereof are of great interest. An enormous number of organic substances for therapeutic purposes are described. This review focuses on peptides developed for diagnosis and therapy of AD and discusses the pre- and disadvantages of peptide drugs.
Keywords: Alzheimer’s disease, amyloid-β, therapy, diagnosis, peptides.
INTRODUCTION
Alzheimer’s disease (AD) is a devastating neurodegenerative disorder and the most common cause of dementia. The clinical characteristics are difficulties with memory, apathy and depression, impaired judgment, disorientation, confusion and other. The greatest risk factor for AD is age. In 2007, AD affected 27 million people world-wide with steadily increasing tendency. By 2050, the prevalence is estimated to quadruple, thereby raising significant economic problems, not to mention the suffer of each affected individual [1].
The pathological hallmarks of AD are the presence of neurofibrillary tangles and amyloid deposits in the brain of the patient, as already defined by Alois Alzheimer in 1907 [2]. Neurofibrillary tangles are aggregates of paired helical filament composed of the abnormally phosphorylated and β-folded tau protein. Tau is a hydrophilic microtubule binding protein which is expressed in six human isoforms of 352 to 441 amino acid residues [3-5].
Aβ is the major component of the amyloid plaques. It consists of 39 to 43 amino acid residues. Aβ, especially Aβ1-42, is prone to aggregation and undergoes formation from monomers to oligomers, larger intermediate forms like protofibrils, and, insoluble fibrils and plaques [6]. Aβ is derived from the amyloid precursor protein (APP) by sequential activities of the β- and γ-secretases [7-9]. As originally suggested by the amyloid cascade hypothesis, it appears likely that Aβ peptides and their aggregated forms initiate cellular events leading to the pathologic effects of AD. According to a previous version of the amyloid cascade hypothesis, fibrillar forms of Aβ, deposited in amyloid plaques, have been thought to be responsible for neuronal dysfunction [7-11]. More recent studies support that diffusable Aβ oligomers including protofibrils, prefibrillar aggregates and so called Aβ-derived diffusible ligands (ADDLs), are the major toxic species during disease development and progression [12-14].
Currently, only palliative therapies for AD are available. Acetylcholine inhibitors like Donepezil, Galantamine and the NMDA receptor antagonist Memantine have been approved for clinical use as treatment of cognitive symptoms.
Although it is still controversial if Aβ is the causative agent of AD, inhibition of Aβ production and aggregation are often addressed for therapy development. As a consequence, the majority of AD therapeutic research has been focused on the Aβ peptide. Less effort has been directed towards the development and validation of tau-targeted therapeutic compounds. A number of tau fibril formation inhibitors, derived from multiple chemical classes, have been identified, as reviewed elsewhere [15]. To date, only one tau fibrillization inhibitor, the phenothiazine methylene blue [16], has entered in vivo evaluation. Promising first results have been presented at the ICAD meeting in 2008, but up to now have not been published in a peer-reviewed journal. Currently, only one peptide compound addressing tau pathology is known. Davunetide (DAP) is an eight amino acid peptide derived from the activity-dependent neuroprotective protein ADNP. It decreases tau phosphorylation and Aβ levels in tau transgenic mice and 3 x transgenic (tg)-AD mice. The intranasal formulation AL-108 is currently in clinical development [17-21].
Therapy approaches targeting Aβ include reduction of Aβ production by inhibitors or modulators of the β- or γ-secretases, Aβ immunotherapy, and inhibition or modulation of Aβ polymerization [22, 23]. Examples for the latter are scyllo-inositol [24, 25], amino-propane sulfonic acid (Tramiprosate) [26], PBT-1 [27], polyphenol (-) epigallocatechin-3-gallate (EGCG) [28, 29], oligomeric acylated aminopyrazoles [30] and several more.
Peptides, which are specified as (linear) molecules consisting of two or more (<100) amino acids residues, are today reasonable alternatives to chemical pharmaceuticals. They are key regulators of biological functions and offer high biological activity associated with high specificity and low toxicity. The peptide market is growing fast due to an increased number of therapeutic targets, improved delivery methodologies, the establishment of large biological and synthetic peptide libraries, and high throughput screening or selection. Today, 67 therapeutic peptides are on the market, 150 in clinical phases and more than 400 in the pre-clinics. In spite of this progress, the development of peptide drugs can be severely hampered by their short half-life in vivo. In general, peptides are rapidly degraded by proteases, and their nature implies problems for administration and delivery, especially to the brain. These problems can at least partially be overcome as peptide chemistry permits a variety of methods for peptide modification or the use of D-enantiomeric amino acid residues [31-33].
A variety of small peptides that inhibit aggregation of Aβ and reduce its toxic effects were already described and a fraction of them shown to be effective in AD rodent animal models. Additionally, Aβ binding peptides, developed for a suitable use in in vivo imaging methods and possibly useful for early diagnosis of AD, were described. Both types of peptides, designed for different applications, are reviewed in this article. All peptides discussed in this article and some more are listed in Tables 1 and 2A to 2F, but although we scanned the literature exhaustively, this list does not claim to be complete.
Table 1.
Name | Sequence | Description | D/L | Results | Reference |
---|---|---|---|---|---|
none |
DWGKGGRWRLWPGASGKTEA and PGRSPFTGKKLFNQEFSQDQ |
Selected by phage display | L | Binds amyloid form of Aβ40, labels amyloid plaques in AD brains slices, discussed as carrier protein for plaque treatment and in vivo imaging | Kang et al., 2003 [40] |
D1/ ACI-80 | ACI-80: QSHYRHISPAQV | Selected by mirror image phage display | D | Aggregate specific, stains human Aβ plaques selectively, stains plaques in mice ex vivo | Wiesehan et al.,
2003 [46] Van Groen et al., 2009 [47] |
none |
C-IPLPFYN-C / C-FRHMTEQ-C |
Selected by phage display | L | Kd for Aβ42 in micromolar range, specific interaction with plaques on brain sections (immunochemistry), encouraging preliminary MRI in vivo study in mice after opening of bbb by mannitol | Larbanoix et al., 2008 [48] |
Pep1 Pep2 |
LIAIMA IFALMG |
Selected by phage display, library based on Aβ sequence | L | Kd for Aβ42 in micromolar range, specific interaction with plaques in human brain sections (immunochemistry), inhibit Aβ aggregation | Larbanoix et al., 2011 [49] |
bbb: blood brain barrier. D/L: describes peptide conformation.
Table 2A.
Name | Sequence | Description | D/L | Results | Reference |
---|---|---|---|---|---|
Aβ(16-20) | KLVFF | Based on Aβ sequence | L | Prevents Aβ fibrilization, identification of key amyloidogenic region | Tjernberg et al., 1996, 1997 [50, 51] |
Aβ- (15-22) Aβ- (16-23) Aβ- (17-24) | Include KLVFF | Based on Aβ sequence | L | Inhibit Aβ aggregation in vitro | Matsunaga et al., 2004 [52] |
K4 | KLVFF-dendrimer | Aβ sequence conjugate | L | Inhibitory effect of compound is potentiated in comparison to monomer | Chafekar et al., 2007 [53] |
none | KLFVV retro-inverso version linked to branched hexameric PEG | Aβ sequence conjugate | D | Inhibitory effect on Aβ aggregation is potentiated | Zhang et al., 2003 [54] |
OR1, OR2 |
RGKLVFFGR or RGKLVFFGR-NH2 |
Based on Aβ-(16-20) region | L/D | Inhibit Aβ oligomerization, aggregation and toxicity. Retro-inverso-peptide resistant to proteolysis and more active | Austen et al., 2008; Matharu et al., 2010 [55, 57]; Taylor et al., 2010 [56] |
Aβ- (31-35) | RIIGL | Based on Aβ sequence IIGL | L | Inhibits Aβ aggregation in vitro | Fülöp et al., 2004 [61] |
none |
Aβ(x-42) X=28-39 |
C-terminal Aβ sequence derived peptides | L | Stabilize Aβ in non-toxic oligomers, inhibit Aβ neurotoxicity | Fradinger et al., 2008 [62] |
Aβ12-28P |
Aβ(12-28) V18P mutation, end-protected |
Peptides block Aβ-Apo E4 interaction by competing for the binding site | D | Inhibition of aggregation in vitro, reduction of plaque formation in tg mice | Sadowski et al., 2004 [66] |
D/L: describes peptide conformation. PEG: poly (ethylene glycol). Aβ-derived peptide sequences are written in bold. Tg: transgenic
Table 2F.
Name | Sequence | Description | D/L | Results | Reference |
---|---|---|---|---|---|
DP1 – DP8 | 6 residues from: Ala, Ile, Val, Ser, Thr, Gly |
“Decoy” peptides of combinatorial library | D | Eliminate calcium effect of Aβ1-42 | Blanchard et al., 97, 2000 [106, 107] |
None | Several with 4 groups of consensus sequences: | Out of combinatorial library | L | Inhibit Aβ aggregation | Schwarzman et al., 2005 [108] |
++H+(H/+) | |||||
or | |||||
++XX+ | |||||
or | |||||
(D/E)LVH | |||||
or | |||||
+LVLF | |||||
None | 12 mers | Phage Display with two libraries. | L | Ligands bind targets in different aggregation states, partly affect aggregation | Orner et al., 2006 [109] |
Parent molecule: KLVFFKKKKKK | |||||
Target: Monomeric and fibrillar Aβ. | |||||
ANA1 | 6 or 15 mers | Novel selection involving phage display and counterselection against rat Aβ | L | Inhibit Aβ toxicity | Taddei et al., 2008 [110] |
ANA2 | e.g. TNPNRRNRTPQMLKR (ANA1) | ||||
ANA3 | |||||
1A, 1B, 2 | MSNKGASIGLMAGDVDIADSHA or | Combinatorial library and selection based on enzyme EGFP | L | Inhibitory effect on Aβ aggregation, de-fibrillization | Bain et al., 2009 [111] |
MSNKGASNALMAGDGDIADSHS or | |||||
MQKLDVVAEDAGSNK | |||||
D3 | RPRTRLHTHRNR | Mirror image phage display | D | Modulates Aβ oligomerization, effective in tg mouse model after oral application | Van Groen et al., 2008; Funke et al., 2010 [112, 113] |
JM169 | Trimer-TEG-D3 | Hybrid compound D3 and nonpeptidic β-sheet breaker | D | In vitro more effective than trimer or D3 | Müller-Schiffmann et al., 2010 [115] |
1-8 | e.g. RAPMGGR or | Panning of heptapeptide library against soluble Aβ1-42 | L | Suppress Aβ1-42 37/48 kDa oligomer formation, keep monomer stable | Kawasaki et al., 2010 [128] |
RRPVVGR | |||||
GN-peptide | GNLLTLD | Peptide homologous to sequence from apo A-I, selected by phage display | L | Modulates aggregation, prevents hippocampal neuronal cultures from Aβ induced degeneration | Paula-Lima et al., 2009 [119] |
D-4F | Ac-DWFKAFYDKVAEKFKEAF-NH2 | Apo A-I mimetic peptide | D | Inhibits Aβ deposition and improver cognitive performance of AD tg mice after oral application | [120] |
D/L: describes peptide conformation. TEG: triethyleneglycol. +: positively charged amino acids. H: hydrophobic amino acid. X: hydrophobic or polar amino acid
PEPTIDES DEVELOPED FOR AD DIAGNOSIS BY IN VIVO IMAGING METHODS (SEE TABLE 1)
Today, the specificity of AD diagnosis can already be improved using glucose metabolism sensing positron emission tomography (PET) experiments [34] and perfusion single photon emission computed tomography (SPECT) [35]. The appearance of amyloid plaques probably occurs many years before cognitive symptoms appear [36, 37]. Therefore, in vivo detection and quantification of amyloid species in the brains of patients during the course of the disease, for early diagnosis and the evaluation of the effects of AD-therapies, is an emerging field in AD research. The best characterization of amyloid plaque load in the brain can be expected from imaging approaches using amyloid ligands as contrast agents. To date, many Aβ binding contrast substances failed due to intolerable unspecific binding or poor distribution in the brains of animals. Only a few PET ligands have been applied to clinical studies (for a review, see ref. [38]). The most prominent and best studied is Pittsburgh compound B (PIB) [39], a benzothiazole derivative binding fibrillar Aβ. Novel Aβ binding substances, suitable for in vivo imaging, are urgently needed. Small Aβ binding peptides with favorable drug properties could easily be coupled to radionucleides or other markers for imaging of amyloid plaques in living AD patients.
Phage display technologies allow the identification of peptide ligands for a given target molecule out of a huge library of different peptides expressed on the surface of bacteriophages. Presentation of the peptide library on the surface of bacteriophages ("phage display") as a fusion of peptide and a phage coat protein allows the physical link between the presented peptide and the DNA sequence coding for its amino acid sequence. Diversity of the peptides/proteins can be introduced by combinatorial mutagenesis of the fusion gene. Extremely large numbers of different peptides can be constructed, replicated, selected and amplified in a process called "biopanning". In 2003, Kang et al. have employed phage display selection to identify two 20-amino acid peptides specifically binding to the amyloid form of Aβ1-40, but not to monomeric Aβ. One of the peptides (amino acid sequence DWGKGGRWRLWPGASGKTEA) could be produced recombinantly in E. coli as a fusion protein with thioredoxin, as well as the chemically synthesized version. The recombinant thiopeptide bound Aβ1-40 amyloid with a Kd of 60 nM, determined by ELISA. Both versions specifically stained amyloid plaques in brain tissue slices of AD patients. The authors discussed the molecules as potential probes for in vivo imaging as well as potential carrier molecules to deliver other therapeutic molecules like antioxidants, chelators, and plaque degrading compounds to the desired location of action [40].
To select for an Aβ-binding D-enantiomeric peptide, which specifically binds to fibrillar Aβ species and plaques, aggregated D-enantiomeric Aβ was used as a target in a mirror phage. Mirror phage display allows the use of phage display to identify peptides that consist solely of D-amino acids. D-enatiomeric peptides are highly resistant to proteases, which can dramatically increase serum and saliva half-life. Additionally, D-peptides can be absorbed systemically after oral administration. D-peptide immunogenicity is reported to be reduced in comparison to L-peptides [41-43]. In the selection process, the Aβ1-42 D-enantiomer was used as a target for selection of peptides displayed on the surface of M13 bacteriophages for those that bind best to D-Aβ1-42. For reasons of symmetry, the D-enantiomeric form of the selected 12-mer peptide will also bind to the native L-form of Aβ1-42 [44]. The most representative peptide in the selection procedure, denoted D1, was demonstrated to bind Aβ with an affinity in the submicromolar range. Employing surface plasmon resonance, binding to Aβ oligomers and fibrils, but not to monomers could be demonstrated. D1 stained amyloid plaques in the brain tissue sections derived from AD patients, whereas other, non-Aβ amyloidogenic deposits, were not stained [45, 46]. D1 was also tested for its in vivo binding characteristics in APP/PS1 transgenic mice. Upon direct injection into the brain, D1 bound very specifically to Aβ1-42, staining all dense deposits in the brain but not diffuse plaques, which contain mainly of Aβ1-40 and are not AD specific [47]. This demonstrated that D1 might be suitable for further development into a molecular probe to monitor Aβ1-42 plaque load in the living brain.
In 2008, Larbanoix et al. selected Aβ1-42 binding peptides using a random disulfide constrained heptapeptide phage display library. Two clones (see sequences in Table 1) were enriched. The Kd-values for the phage clones, on which several peptide copies are displayed, were in the picomolar range. After peptide synthesis including biotinylation, the binding affinities dropped to the micromolar range. Nevertheless, preliminary in vivo studies in transgenic mice, in which the functionalized peptides were used as contrast agents, showed high contrast effects in APPV717I/PS1A246E transgenic mice, but not in wildtype controls. For the experiments, however, the blood brain barrier (bbb) had to be permeabilized artificially with 25 % mannitol [48]. Very recently, Larbanoix et al. designed another linear hexapeptidic phage display library based on the Aβ1-42 amino acid sequence and selected against aggregated Aβ1-42 as a target. Two of 26 selected clones, presenting highest binding affinities to Aβ1-42, were translated to synthetic peptides with biotine label (Pep1: LIAIMA and Pep2: IFALMG, corresponding Aβ fragment IIGLMV31-36) and presented lower Kd values (still in the micromolar range) as the peptides described in the first article. Pep1 and Pep2 were highly hydrophobic and are expected to pass the bbb very well. The peptides did not show any sign of toxicity in cell culture. The specific interaction of both peptides with amyloid plaques in human brain tissue was demonstrated by immunohistochemistry [49].
THERAPEUTIC PEPTIDES (SEE TABLES 2A-D)
Table 2D.
Name | Sequence | Description | D/L | Results | Reference |
---|---|---|---|---|---|
none | N-methylated GSNKGAIIGLM | First methylated β-sheet breakers | L | Prevent Aβ aggregation in vitro, inhibit cell toxicity | Hughes et al., 2000 [93, 94] |
Aβ16-22m Aβ16-20m |
KLVFF based e.g. KNmeLVNMEFFNMeAE | Methylated β-sheet breaker, N-methyl groups in alternating positions | L | Prevent Aβ fibril forming, disassemble fibrils in vitro | Gordon et al., 2001, 2002 [78, 95] |
inL (all L) in D (all D) inrD (retro-inverso) |
LKLVFF based, N-methyl-20F | Methylated β-sheet breaker, single N-methyl-amino acids | L/D | Reduces Aβ toxicity in cell culture | Cruz et al., 2004; Grillo-Bosch et al., 2009 [96, 97] |
PPI-1019 “Apan” | Methyl-LVFFL | Methylated β-sheet breaker | D | Completed phase I and II human clinical trials | Jhee et al., 2003 [98, 99] |
e.g. SEN304 | e.g. D-chGly-D-Tyr-D-chGly-D-chGly-D-mLeu |
“meptide”, methylated β-sheet breaker | D | Highly active inhibitors of Aβ aggregation and toxicity | Kokkoni et al., 2006 [100] |
D/L: describes peptide conformation. chGly: cyclohexylglycine; mLeu: N-methylleucine; NMe: N-methylated amino acids. Aβ-derived peptide sequences are written in bold.
Aβ-sequence Derived Peptides (see Table 2A)
In 1996, Tjernberg and coworkers searched for an Aβ ligand to interfere with Aβ-self interaction and polymerization. The strategy was to identify binding sequences within Aβ and, based on their primary structures, to synthesize Aβ derived peptide ligands. The short Aβ fragment KLVFF (Aβ16-20) was identified to bind full length Aβ and to prevent the fibrillization thereof. Alanine substitution experiments revealed that amino acids Lys16, Leu17 and Phe20 were critical for Aβ interference [50]. A molecular modeling study suggested that the association of full length Aβ and the KLVFF peptide lead to the formation of atypical antiparallel β-sheet structures stabilized by Lys16, Leu17 and Phe20 [51].
The effects of KLVFF containing or derived synthetic peptides was further approved in a variety of studies, e.g. by the one of Matsunaga et al. [52], see Table 2A. Additionally, it was proposed that conjugates, bearing several copies of the KLVFF sequence or the retro-inverso version thereof, linked to dendrimers or to branched poly(ethylene glycol) moieties, possess superior affinity and efficiency [53, 54]. In 2008, Austen et al. designed KLVFF derived compounds to address very early aggregation intermediates of Aβ, i.e. Aβ oligomers. The idea was to add water soluble amino acids residues to KLVFF, thereby generating the peptides OR1 (RGKLVFFGR) and OR2 (RGKLVFFGR-amid). Both peptides inhibited Aβ fibrillogenesis, whereas OR2 additionally inhibited oligomer formation and Aβ toxicity on SY5Y cells, supporting the idea that particularly oligomers are responsible for the cytotoxic effects of Aβ [55]. Unlike OR2, the retro-inverso D-enantiomeric version (RI-OR2) of the peptide was highly resistant to proteolysis and stable in human plasma and brain extracts [56]. Additionally, it was reported that retro-inversions of OR1 and OR2 increase the inhibitory effects of the peptides [57].
In 2004, Fülop et al. developed an Aβ aggregation inhibitor based on the Aβ31-34 sequence IIGL, which also plays a fundamental role in Aβ aggregation and cytotoxicity [58-60]. As in the study described above, the strategy was to link a solubilizing amino acid residue to the original sequence RIIGL. In contrast to propionyl-IIGL, another derivative of the same sequence, PIIGL did not self-aggregate and was not toxic to cells in culture. RIIGL inhibited the formation of Aβ fibrils and reduced Aβ cytotoxicity [61].
In 2008, Fradinger et al. prepared a series of Aβ C-terminal fragments (Aβx-42; x = 28-39). The authors of the article tested the hypothesis that C-terminal peptides of Aβ should possess high affinity to full length Aβ and might disrupt oligomer formation, as the C-terminus is supposed to be a key region controlling Aβ aggregation. Cell viability assays identified Aβ31-42 and 39-42 to be the most effective inhibitors of Aβ induced cell toxicity. The peptides additionally prevented the disturbance of synaptic activity by Aβ oligomers. To investigate the in vitro mechanism of action, dynamic light scattering, photo-induced cross-linking and discrete molecular dynamics were applied and gave clear hints that the peptides inhibit Aβ induced toxicity by stabilizing Aβ in non-toxic oligomers [62].
The inheritance of the apolipoprotein (apo) E4 allele has been identified as a major genetic risk factor for sporadic AD [63]. All Apo isoforms are discovered to act as pathological chaperones and propagate Aβ fibril formation, with apo E4 being the most efficient isoform [64, 65]. Sadowski et al. investigated weather blocking of the interaction between apo E4 and Aβ can have therapeutic effects [66]. In earlier studies, Ma et al. demonstrated that the synthetic peptide Aβ12-28 can be used as inhibitor of apo E4-Aβ interaction, inhibiting Aβ fibril formation in vitro [67]. Sadowski et al. modified the Aβ12-28 sequence: substitution of valine at position 18 to proline rendered the peptide non-amyloigogenic and untoxic whereas the affinity of the peptide to apo E4 was not affected. The use of D-amino acids and end-protection increased the serum half-life of the peptide substantially. Aβ12-28P blocked Aβ-apoE4 interaction and reduced Aβ fibrillogenesis and toxicity in vitro. The peptide was bbb-permeable and inhibited Aβ deposition in AD transgenic mice [66].
β-sheet Breaking Peptides (see Tables 2B-2E)
Table 2B.
Name | Sequence | Description | D/L | Results | Reference |
---|---|---|---|---|---|
nn | KLVFF based, with chain of charged amino acids as disruption motif | β-sheet breaker | L | Enhance fibrillization of Aβ oligomers and therefore reduce Aβ toxicity. Effectiveness of inhibitor is dependent on its surface tension modifying properties | Ghanta et al., 1996; Pallitto et al., 1999; Lowe et al., 2001; Kim et al., 2004; Moss et al., 2003; Kim et al., 2003; Gibson et al., 2005 [69-75] |
Aβ16-20e | KLVFF with ester sub-stitution | No hydrogen bond can be formed | L |
Inhibit Aβ aggregation, disassembles
fibrils. Expected to hydrolyze rapidly in vivo |
Gordon et al., 2003 [77, 78] |
AMY-1 AMY-2 | KLVFF based | β-sheet breaker α,α-disubstituted amino acids | L | Inhibition of fibrillization, globular aggregates are formed | Etienne et al., 2006 [79] |
P1, P2 | KLVF-ΔA-I-ΔA and KF- ΔA- ΔA- ΔA-F | Disruption of aggregation by different local confirmation | L | Inhibit Aβ aggregation | Rangachari et al., 2009 [81] |
D/L: describes peptide conformation. ΔAla: α, β-dehydroalanine. Aβ-derived peptide sequences are written in bold. Table
Table 2E.
Name | Sequence | Description | D/L | Results | Reference |
---|---|---|---|---|---|
e.g. PPI-368 PPI-457 |
e.g. Cholyl-LVFFA | β-sheet
breaker Aβ-binding sequences with cholyl-bulky group |
L/D | Aβ specific, inhibit Aβ aggregation potently, reduces Aβ cell toxicity | Findeis et al., 1999, 2001, 2002 [98, 101, 102] |
Aβ-(38-42) | GVVIA, RVVIA | Based on Aβ sequence, amidated at C-termius | L | Inhibit Aβ aggregation and toxicity | Hetényi et al., 2002 [132] |
Trp-Aib | β-sheet breaker | D | Inhibition of Aβ oligomer formation in vitro, effective in AD tg mice after oral application | Frydman-Marom et al., 2009 [103] | |
none | EIVY-rest | β-sheet breaker see text in subchapter “KLVFF derived β-sheet breakers” | L | Modulate Aβ aggregation, depending of
solvent disruptive amino acid sequence. Discussed in this article with KLVFF-based β-sheet breakers |
Sun et al., 2009 [76] |
D/L: describes peptide conformation. AIB: α-aminoisobutyric acid; tg: transgenic. Aβ-derived peptide sequences are written in bold.
The first β-sheet breaker peptides were reported by Soto et al. in 1996 [68]. In general, the term β-sheet breaker describes compounds, containing an Aβ recognition or binding motif which provides specificity, combined with an Aβ oligomer or fibril disrupting motif which can consist of charged amino acids, prolines, methylated amino acids, cholyl-groups and others. By far the most Aβ aggregation inhibiting peptides are β-sheet breaker, demonstrating the effectiveness of these compounds. In most cases, the Aβ recognition domain is based on the Aβ sequence.
β-sheet Breaking Peptides Based on the KLVFF Sequence (see Table 2B)
Starting in 1996, Ghanta et al. designed Aβ binding hybrid peptides based on the KLVFF-binding sequence in addition to a disruption domain consisting of a chain of charged amino acids like KKKKKK or RRRRRR. Interestingly, some of the hybrid peptides accelerated Aβ aggregation, but reduced Aβ toxicity. Presumably, the peptides speeded up the association of potentially toxic Aβ oligomers into less toxic Aβ fibrils as demonstrated by a range of biochemical and biophysical methods. The ability of the compounds to increase solvent tension was a very strong predictor on the effect on Aβ aggregation. Chains of charged amino acids without Aβ recognition motif, used in the assays as control, did not exert comparable strong effects [69-75]. In a similar study, Sun et al. investigated hybrid peptides composed of the critical Aβ binding domain EIVY and solvent disruptive sequences poly E, K or R. The hybrid peptides EIVY-EEEE and EIVY-KKKK enhanced Aβ fibrillization, whereas EIVY-RRRR inhibited Aβ aggregation and altered the morphology of Aβ fibrils to amorphous aggregates. The ability of the hybrid peptides to interfere with Aβ aggregation was also discussed in the context of their abilities to change the surface-tension in solutions [76], in Table 2E.
In 2003, Gordon et al. replaced the amide bonds of Aβ16-20 with ester bonds in an alternating fashion. The ester peptide Aβ16-20e was monomeric under solvation conditions, inhibited Aβ aggregation and disassembled existing fibrils. Aβ16-20e, could, however, build dimers. These data demonstrated that interference with backbone hydrogen bonding is therapeutically attractive [77, 78].
Other β-sheet breakers based on the KLVFF sequence are AMY-1 and AMY-2. Both contain alpha, alpha-disubstituted amino acids at alternating positions and arrest Aβ fibril growth. Instead, large globular aggregates are formed [79]. Rangachari et al. investigated the α,β-dehydroalanine (ΔAla) containing peptides P1 (KLVF-ΔA-IΔA) and P2 (KF-ΔA-ΔA-ΔA-F). The design of P2 was based on the experience that peptides containing the FxxxxF motif bind to the groove composed of the GxxxG motif (amino acids G33-G37) in Aβ fibrils [80] α,β-dehydro-amino acid residues are known to be strong inducers of specific peptide conformations. Additionally, they increased resistance to proteolysis in vivo. Both peptides under investigation inhibited Aβ aggregation [81].
Proline Based β-sheet Breaker (see Table 2C)
Table 2C.
Name | Sequence | Description | D/L | Results | Reference |
---|---|---|---|---|---|
iAβ5 | Ac-LPFFD-amid | Proline β-sheet breaker | L/D | Aβ fibril inhibition and de-fibrillization in vitro. Reduction of plaque load and Aβ induced pathological processes in rat model and in AD tg mice. Improvement of rat spatial memory impairments | Soto et al., 1996; Soto et al., 1998; Poduslo et al., 1998; Sigurdsson et al., 2000; Permanne et al., 2002; Chacon et al., 2004 [68, 83-86, 135] |
iAβ5 | LPFFD-derivatives | Proline β-sheet breaker | L | Methylation of amide nitrogen increased in vitro and in vivo stability while maintaining iAβ5 activity in vitro | Adessi et al., 2003 [131] |
iAβ5-PEG | LPFFD-PEG | Proline β-sheet breaker | L | Biological activity of iAβ5 (in vivo) is not affected by PEG. Aimed to improve pharmacological properties, especially in vivo degradation | Rocha et al., 2009 [136] |
none | LPYFD | Proline β-sheet breaker | L | Decreased neurite degeneration, tau aggregation and cell viability reduction induced by Aβ | Datki et al., 2004 [87] |
LPYFDa | LPYFDamid | Proline β-sheet breaker, amidated | L | Protects neurons in vitro and in vivo after intraperitonial administration to rat models | Szegedi et al., 2005; Juhász et al., 2009 [88, 89] |
D/L: describes peptide conformation. Tg: transgenic. PEG: poly (ethylene glycol).
One peptide compound which was extensively investigated in vitro, as well as in animal models, was introduced 1996 by Soto et al [68]. The authors developed a peptide partially homologous to the central hydrophobic region of Aβ (amino acids 17-21: LVFFA), containing the amino acid proline to prevent the formation of β-sheet structure and to inhibit Aβ amyloid formation. Proline has special characteristics and is a well-known β-sheet blocker [82]. To increase the solubility of the peptide compound, charged amino acids were added to the ends. The so called “inhibitor of fibrillogenesis” (iAβ1: RDLPFFPVPID) did not aggregate itself, inhibited amyloid formation and disassembled existing fibrils in vitro. The relative dissociation constant of iAβ1 was determined by fluorescence spectroscopy to be approximately 80 nM. In order to enhance bbb permeability and to reduce generation of immune responses in vivo, the peptide was shortened, and the derivatives iAβ3 (seven amino acids) and iAβ5 (five amino acids) were proven to be similar good or even better Aβ aggregation inhibitors in vitro than the basic compound, compared e.g. in ThT assays. The D-enantiomeric version of iAβ1 was as effective as the L-version, and more resistant to proteases [68]. In later studies, iAβ5 was shown to inhibit Aβ cytotoxicity in cell culture. It also reduced Aβ fibrillogenesis in a rat brain model of amyloidogenesis. Male Fischer-334 rats were injected with freshly solubilized Aβ1-42 directly into the amygdala and the animals were sacrificed 8 days later. Co-injection of iAβ5 in 20 % molar excess lead to significantly reduced plaque deposition [83]. In addition, it was shown that iAβ5 induced disassembly of fibrils, reduced Aβ induced histopathological changes like neuronal shrinkage and the extent of interleukin-1β positive microglia cells surrounding Aβ deposits [84]. The chronic intraperitoneal administration of the peptide to the rat model described above lead to a significant improvement of spatial learning acquisition in Morris water maze tests and working memory tests. To perform these studies, iAβ5 was end protected by N-terminal acetylation and C-terminal amidylation [85], as the non-protected peptide was unstable in blood and proteolytically degraded very rapidly [86]. In 2002, two different AD transgenic mouse models (double transgenics overexpressing human APP with London mutation V717I and human PS1 with A246E mutation, and single transgenics only overexpressing human APP V717I) were used to demonstrate that intraperitoneal injected, end-protected iAβ5 reduced amyloid plaque formation, neuronal cell death and brain inflammatory processes. Pharmacokinetic studies in mice and rats demonstrated good stability of the end-protected peptide as well as high capability of the peptide to cross the bbb. The mechanism of peptide action remained unclear, but administration of large doses of the peptide did not lead to antibody production in the treatment and evaluation period [86].
In 2004, Datki and Coworkers investigated the effects of a LPFFD based peptide, amino acid sequence LPYFD, on the Aβ induced changes on neuroblastoma cells. Neurite degeneration and tau aggregation were significantly decreased. Additionally, Aβ induced cell toxicity was reduced [87]. The pentapeptide LPYFD-amid protected neurons against Aβ toxicity in vitro and in vivo. Intraperitoneally administered LPYFD-amid crossed the blood brain barrier in rats at least to a certain extent, and protected against synaptotoxic effects of Aβ up to 3.5 h after i.p. injection [88, 89].
Methylated β-sheet Breaking Peptides (see Table 2D)
Several teams have studied the effects of N-methyl amino acid incorporation into peptides. For the resulting compounds, the term “meptides” was established. The first methylated β-sheet breaker was described by Hughes et al. After it has been suggested that Aβ25-25, amino acid sequence GSNKGAIIGLM, resembles a biologically active region of Aβ that forms large β-sheet fibrils and is highly cytotoxic [90-92], the authors used it as a full length model for Aβ1-42 and synthesized six N-methylated derivatives to prove that those could prevent Aβ wildtype aggregation and cytotoxicity. As the derivatives were homologous to Aβ, they were expected to bind the wildtype form and to prevent further addition of Aβ monomers. It was assumed that N-methylation could block hydrogen bonding at the outer edge of the assembling amyloid, disrupting peptide-peptide interactions that promote Aβ fibrillization. As expected, N-methylated peptide variants had significant influence on Aβ aggregation, as investigated using a variety of biophysical methods. Notably N-methyl-Gly-33 was shown to be amenable to inhibit Aβ aggregation and cytotoxicity. The localization of the N-methyl group was very critical as some of the other peptides did not prevent Aβ aggregation, but altered fibril morphology [93, 94].
In 2001, Gordon et al. described the synthesis and biochemical characterization of rationally designed “meptides” based on the KLVFF sequence, containing N-methyl amino acids in alternating positions of the sequence. One of the compounds, termed Aβ16-22m (NH2-K(Me-L)V(Me-F)F(Me-A)E-CONH2) was shown to be highly soluble in aqueous media and monomeric in buffer solution. It inhibited Aβ fibrillization and disassembled preformed Aβ fibrils in vitro. Inhibition was sequence specific and dependent of N-methylization. Protease resistance of the methylated peptide was increased in comparison to the unmethylated Aβ16-22 peptides [78]. The Aβ16-20m peptide, a truncated version of the peptide described above, was synthesized in order to eliminate the charged Glu residue, providing the inhibitor with a net positive charge. Aβ16-20m was effective to inhibit Aβ polymerization and to dissassemble preformed fibrils. In addition, it was highly water soluble despite its composition of hydrophobic amino acids. The peptide passed spontaneously model phospholipid bilayers and cell membranes, suggesting promising pharmacological properties [95].
In 2004, Cruz et al. developed the peptide “inL”, based on the KLVFF Aβ recognition element. The authors added an additional Lys to the N-terminus, in order to increase solubility, and an N-methyl-20F in order to block Aβ aggregation [96]. The peptide inhibited Aβ toxicity in cell culture very efficiently. In 2009, the corresponding D-peptide “inD”, as well as the retro-inverso peptide “inrD”, was investigated. Both D-enantiomeric peptides were more resistant against protease degradation, as expected. The retro-inverso peptide was shown to be a more effective inhibitor of Aβ aggregation than the other two peptide versions [97].
The peptidic inhibitor PPI-1019, also known as Apan, is derived from the D-enantiomeric Cholyl-LVFFA-NH2 (see chapter other β-sheet breaking peptides). The cholyl-group was replaced by a methyl-group and the C-terminal residue was changed from D-alanine to D-leucine. PPI-1019 completed phase I and II clinical trials and was found to be safe, well tolerated and amenable to cross bbb. After peptide administration, levels of Aβ1-40 in the CSF increased, which might be discussed as a sign for Aβ clearance out of the brain [98, 99].
In 2006, Kokkoni et al. optimized “meptides” based on the KLVFF sequence in a large approach based on five peptide libraries. Peptide length, methylation sites, end-blocking, side chain identity and chirality were varied. The most interesting compound, judged by Aβ fibrillization and cell cytotoxicity inhibition activity, was D-[(chGly)-(Tyr)-(chGly)-(chGly)-(mLeu)]-NH2 and rules could be stated to predict peptide performance. Ideal inhibitors should be D-peptides, possess a free N- but an amidated C-terminus and residues one to four should be large, branched hydrophobic side chains. Only one methylated amino acid was essential [100].
Other β-sheet Breaking Peptides (see Table 2E)
In 1999, Findeis et al. started the development of new β-sheet breaking peptides derived from 15-residue Aβ peptides. The anti-Aβ fibrillization activity of the peptides was enhanced by modification of their amino terminus, where different organic reagents were attached. In subsequent libraries, the size of the inhibitors as well as the amino acid sequences were optimized, and finally, the lead compound cholyl-LVFFA-OH, designated PPI-368, was identified. PPI-368, the acyl-D-amino acid analogue PPI-433 and the amide analogue PPI-457 inhibited Aβ polymerization potently. The latter two were stable in monkey CSF for at least 24 h. Unfortunately, hepatic first –pass elimination was reported for all compounds after biodistribution studies were performed [101, 102].
In 2009, Fryman-Marom et al. introduced the D-enantiomeric β-sheet breaker NH2-D-Trp-Aib-OH, which combined an indole and α-aminobutyric acid (Aib) [103]. Aib has been shown to induce helical conformations and to disrupt β-sheet structures, the Ramachandran plot indicating that Aib even has the potential to be an better β-sheet breaker than proline [104]. NH2-D-Trp-Aib-OH specifically interacted with (untoxic) low molecular weight Aβ oligomers and inhibited their growth to larger, celltoxic Aβ forms in vitro. The compound was stable, safe, orally bioavailable, and crossed the bbb in the range of 4 to 8%, depending on the route of administration. It reduced the amount of plaques in the brains of AD transgenic mice and improved their cognitive performance [103].
Peptides Selected in Combinatorial Libraries
Combinatorial peptide libraries offer a powerful technique to select highly specific peptide ligands for different pharmacological interesting targets [44, 105].
Already in 1997, Blanchard and colleaques selected so-called “decoy peptides” using a combinatorial library of approximately 43000 individual sequences composed of the D-amino acids Ala, Ile, Val, Ser, Thr and Gly. The 6 residue peptides were chosen by their ability to complex with tagged Aβ25-35 peptide and had β-sheet forming potential, associating with Aβ and blocking aggregation. Some of the selected peptides abolished the calcium influx, caused by aggregated Aβ25-35 or Aβ1-42 in cell culture [106, 107].
Schwarzman and al. selected inhibitors of amyloid formation by screening of a FliTrx random peptide library. 12 residue peptides were displayed on the surface of E. coli by fusion to a flaggelar protein. Synthetic Aβ1-42 was used as a target molecule in five rounds of biopanning. Four groups of Aβ binding peptides were isolated (see Table 2F), two of which were enriched by positively charged amino acids. Most clones were shown to bind monomeric Aβ, but exhibited very low binding to fibrillar Aβ 1-42 [108].
Orner et al. identified peptides that bind to Aβ in either monomeric or fibrillar state using phage display approaches with monomeric or fibrillar Aβ as targets, respectively. Two libraries were designed, guided by the group’s previous studies on the KLVFFK6 peptide (see above). The first library displayed sequences with the PoPoPoKLVFFPoPoPo motif, where Po indicates a residue with polar side chain. The second library contained sequences with XXXKLpLpArArPoPoPoPo motif, where X is any amino acid, and Lp and Ar indicate residues with lipophilic and aromatic side chains. Peptides with selectivity for monomeric versus fibrillar Aβ could be identified and most of the selected peptides bound to Aβ10-35 with higher affinity than the parent peptide Ac-KLVFFKKKKK-OH. Peptides selected for Aβ monomer binding did not affect aggregation, whereas peptides selected to bind fibrillar Aβ increased the aggregation of Aβ dramatically, altering the morphology of the resulting aggregate. This effect was clearly correlated with affinity of the peptides to the N-terminal part of Aβ [109].
Taddei et al. reported on an approach with the aim to inhibit the catalytic production of H2O2 by Aβ, which is dependent on Aβ’s superoxide dismutase (SOD)-like activity. A phage display procedure with 6- and 15mer peptide libraries was applied to select for peptides that target the active site of human Aβ’s SOD-like activity, in order to prevent its interaction with redox-active metal ions. As the SOD-like activity site is not present in rat Aβ, a counterselection step with rat Aβ was employed in the phage display procedure. Aβ1-40 or Aβ1-42 were used as targets. 25 peptides which bound to Aβ were identified, and two of the three most enriched peptides, named amyloid neutralizing agents (ANA) 1 to 3, were shown to significantly reduce Aβ’s SOD-like activity in cell culture. A 15-mer peptide additionally reduced Aβ toxicity in cell culture and seemed to be comparably potent as the known Aβ metal-mediated redox activity inhibitor Clioquinol [110].
A very sophisticated system was used by Baine et al. to select for peptides that inhibit Aβ aggregation in two combinatorially diverse peptide libraries. The goal was to select peptides which bind the two hydrophobic patches of Aβ and block aggregation by highly charged and polar aspartatic acid residues. Aβ1-42 was genetically fused to EGFP. When expressed in E. coli, aggregation of Aβ inhibits the correct folding of EGFP and therefore its fluorescent properties. In the selection process, randomized peptides were co-expressed with Aβ-EGFP. Peptides which were resistant to degradation by cellular proteases and inhibited Aβ aggregation permitted EGFP to be folded properly. Colonies with the brightest fluorescence were chosen for further characterization. Three candidate peptides were selected and characterized, being capable to inhibit Aβ aggregation. One of them even disaggregated preformed Aβ fibrils [111].
Recently a highly specific D-enantiomeric ligand for Aβ has been identified using a mirror image phage display approach with a huge randomized 12-mer peptide library (> 1 billion different peptides). The dominant peptide sequence RPRTRLHTHRNR was obtained, referred to as D3. D3 modulated Aβ aggregation and inhibited Aβ toxicity in cell culture. In vitro data clearly demonstrated that D3 is able to precipitate toxic Aβ oligomers into large, high-molecular-weight, nontoxic, ThT negative, nonamyloidogenic amorphous aggregates that fail to act as seeds in Aβ fibril formation assays. D3 did not increase the concentration of monomeric Aβ. Computational simulations of an Aβ nonamer in the presence and absence of D3 proved strong interactions between the arginine-rich D3 and negatively charged groups of Aβ, which were expected to compensate the charge on the Aβ surface and reduce solubility and promote the aggregation of Aβ. Moreover, D3 binding also showed effects on the topology of the Aβ oligomers, which induced a large twist and facilitated the formation of nonfibrillar aggregates [112, 113]. Van Groen et al. demonstrated the usage of FITC-labelled D3 for both in vitro and in vivo staining of Aβ-1-42 in the brains of transgenic AD-model mice [47]. Additionally, D3 was proven to have notable bbb permeability in an in vitro bbb cell culture model which further demonstrated the therapeutic potential of D3 [114]. Most recently, oral treatments of Aβ transgenic mice with D3 yielded significant cognitive improvement, reduction of plaque load and plaque-related inflammation [112]. In 2010, Müller-Schiffmann et al. reported on the D3 hybrid compound JM169, which combined the D-enantiomeric peptide with a β-sheet breaking compound via a linker substance. The authors demonstrated that the hybrid compound was more efficient in vitro than the sum of its components and had novel properties [115].
In 2009, Paula-Lima et al. used a phage display approach to select peptides binding to the aggregated form of Aβ. One of the identified heptapeptides with the amino acid sequence GNLLTLD (designated GN peptide) was detected to be homologous to the N-terminal domain of mammalian apolipoprotein A-I. Apo A-I, the major protein component of high-density lipoprotein (HDL), has a central role in reverse cholesterol transport [116, 117] and anti-oxidant as well as anti-inflammatory properties [118]. It was shown that purified human apoA-I and Aβ formed complexes. The interaction of apo A-I also rendered the morphology of amyloid aggregates [119]. Likewise in 2009, Handattu et al. evaluated the apo A-I mimetic peptide D-4F, synthesized from D-amino acids and co-administered with pravastatin, as a treatment for AD transgenic mice [120]. D-4F was developed based on the presence of lipid-associating amphipathic α-helices in apo A-I and possessed the ability to avidly bind lipids [121, 122]. Most studies of D-4F were focused on its potential role in atherosclerosis management [123, 124]. Several studies suggested that AD may have an inflammatory component similar to atherosclerosis that is associated with very small vessels such as arterioles [125-127]. In the study, groups of male mice were treated with D-4F and pravastatin via the drinking water. In comparison to the controls, the treated group showed significantly increased cognitive behavior, reduced plaque deposition and reduced inflammatory responses [120].
Very recently, Kawasaki et al. constructed a random library to obtain peptide inhibitors specific to inhibit formation of soluble 37/48 kDa Aβ1-42 oligomers. The random library was based on the LPFFD sequence: XX-P-XXX, where X means any amino acid. Novel peptides containing arginine residues were enriched while panning for soluble Aβ1-42. Selected ligands with the strongest affinity to Aβ contained three arginine residues and suppressed formation of 37/48 kDa oligomers and kept the monomeric form of Aβ even after 24 hours of incubation [128].
CONCLUSION
During the past years, several peptide inhibitors of Aβ aggregation have been investigated for their applicability as new therapeutic lead compounds. In conclusion, it must be stated that only very few, iAβ5 [84-86], Aβ12-28P [66], LPYFDa [88, 89], trp-Aib [103], D-4F [120] and D3 [47, 112, 113] were proven to be effective in rodent mouse models. Only one compound, PPI-1019, is tested in clinical trials. Despite the high diversity of peptides, combined with their simplicity, high specificity, low toxicity and high biological activity [31, 33], peptides are susceptible to proteolytic degradation and in general do not circulate for more than a few minutes in blood [129]. Additionally, peptides in general do not cross membranes very well [130].
Several strategies were already applied to overcome high protease susceptibility of peptides and to improve bbb permeability. Different chemical modifications, including incorporation of conformationally constrained amino acids, or modifications of the peptide backbone, have been performed. For example, Adessi et al. introduced a methyl-group at the nitrogen atom of one amid bond in iAβ5, resulting in a 10-fold higher in vivo life-time in comparison to the unmodified iAβ5 [131]. Moreover, end-protection is commonly used to shield peptides against proteolytic degradation as well as to increase bbb permeability [66, 89, 132]. Another promising strategy to improve the peptide stability is the use of D-enantiomeric amino acids which are considered to be rather protease resistant in vivo and in addition often less immunogenic than the respective L-peptides [42, 66, 133]. Additionally, D-peptides can be taken up systematically after oral administration [134].
It was assumed for a long time that Aβ deposited in extracellular amyloid fibrils and plaques were the major pathogenic species in AD. However, over the past decade, accumulating evidence suggests that Aβ oligomers are the toxic moiety responsible for synaptic dysfunction and neuronal cell loss [12]. At the moment it is unclear whether peptide inhibitors should be targeted to monomeric Aβ, oligomeric Aβ or Aβ fibrils deposited in plaques. It is also unclear if Aβ1-40 or Aβ1-42 should be addressed and weather compounds need to cross the bbb in order to be effective. More work is necessary to elucidate the nature of the most synaptotoxic Aβ species during AD development and progression. The ongoing research on peptides which target distinct Aβ species, as well as the investigation of their influence on Aβ aggregation and toxicity, will provide further understanding of the molecular mechanisms involved in AD.
ACKNOWLEDGEMENTS
D.W. is supported by the Technology-Transfer Fonds of Forschungszentrum Jülich. D.W. and S.A.F. are supported by Deutsche Forschungsgemeinschaft (Research Training Group 1033).
REFERENCES
- 1. Brookmeyer R, Johnson E, Ziegler-Graham K, Arrighi HM. Forecasting the global burden of Alzheimer's disease. Alzheimers & Dementia. 2007;3(3 ):186–91. doi: 10.1016/j.jalz.2007.04.381. [DOI] [PubMed] [Google Scholar]
- 2. Alzheimer A. Über eine eigenartige Erkrankung der Hirnrinde. All Z psychait. 1907;64:146–8. doi: 10.1002/ca.980080612. [DOI] [PubMed] [Google Scholar]
- 3. Goedert M, Spillantini MG, Jakes R, Rutherford D, Crowther RA. Multiple isoforms of human microtubule-associated protein tau: sequences and localization in neurofibrillary tangles of Alzheimer's disease. Neuron. 1989;3(4 ):519–26. doi: 10.1016/0896-6273(89)90210-9. [DOI] [PubMed] [Google Scholar]
- 4. Leroy K, Bretteville A, Schindowski K, Gilissen E, Authelet M, De Decker R, et al. Early axonopathy preceding neurofibrillary tangles in mutant tau transgenic mice. Am J Pathol. 2007;171(13 ):976–92. doi: 10.2353/ajpath.2007.070345. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5. Vieira MN, Forny-Germano L, Saraiva LM, Sebollela A, Martinez AM, Houzel JC, et al. Soluble oligomers from a non-disease related protein mimic Abeta-induced tau hyperphosphorylation and neurodegeneration. J Neurochem. 2007;103(2 ):736–48. doi: 10.1111/j.1471-4159.2007.04809.x. [DOI] [PubMed] [Google Scholar]
- 6. Finder VH, Glockshuber R. Amyloid-beta aggregation. Neurodegener Dis. 2007;4(1 ):13–27. doi: 10.1159/000100355. [DOI] [PubMed] [Google Scholar]
- 7. Haass C, Selkoe DJ. Cellular processing of beta-amyloid precursor protein and the genesis of amyloid beta-peptide. Cell. 1993;75(6 ):1039–42. doi: 10.1016/0092-8674(93)90312-e. [DOI] [PubMed] [Google Scholar]
- 8. Kang J, Lemaire HG, Unterbeck A, Salbaum JM, Masters CL, Grzeschik KH, et al. The precursor of Alzheimer's disease amyloid A4 protein resembles a cell-surface receptor. Nature. 1987; 325(6106 ):733–6. doi: 10.1038/325733a0. [DOI] [PubMed] [Google Scholar]
- 9. Weidemann A, Konig G, Bunke D, Fischer P, Salbaum JM, Masters CL, et al. Identification, biogenesis, and localization of precursors of Alzheimer's disease A4 amyloid protein. Cell. 1989;57(1 ):115–26. doi: 10.1016/0092-8674(89)90177-3. [DOI] [PubMed] [Google Scholar]
- 10. Hardy JA, Higgins GA. Alzheimer's disease: the amyloid cascade hypothesis. Science. 1992;256(5054 ):184–5. doi: 10.1126/science.1566067. [DOI] [PubMed] [Google Scholar]
- 11. Selkoe DJ. The molecular pathology of Alzheimer's disease. Neuron. 1991;6(4 ):487–98. doi: 10.1016/0896-6273(91)90052-2. [DOI] [PubMed] [Google Scholar]
- 12. Haass C, Selkoe DJ. Soluble protein oligomers in neurodegeneration: lessons from the Alzheimer's amyloid beta-peptide. Nat Rev Mol Cell Biol. 2007;8(2 ):101–12. doi: 10.1038/nrm2101. [DOI] [PubMed] [Google Scholar]
- 13. Shankar GM, Bloodgood BL, Townsend M, Walsh DM, Selkoe DJ, Sabatini BL. Natural oligomers of the Alzheimer amyloid-beta protein induce reversible synapse loss by modulating an NMDA-type glutamate receptor-dependent signaling pathway. J Neurosci. . 2007;27(11 ):2866–75. doi: 10.1523/JNEUROSCI.4970-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14. Shankar GM, Li S, Mehta TH, Garcia-Munoz A, Shepardson NE, Smith I, et al. Amyloid-beta protein dimers isolated directly from Alzheimer's brains impair synaptic plasticity and memory. Nat Med. 2008;14(18 ):837–42. doi: 10.1038/nm1782. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15. Ballatore C, Brunden KR, Trojanowski JQ, Lee VM, Smith AB, 3rd, Huryn D. Modulation of protein-protein interactions as a therapeutic strategy for the treatment of neurodegenerative tauopathies. Curr Top Med Chem. 2011;11(3 ):317–30. doi: 10.2174/156802611794072605. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16. Wischik CM, Edwards PC, Lai RY, Roth M, Harrington CR. Selective inhibition of Alzheimer disease-like tau aggregation by phenothiazines. Proc Natl Acad Sci U S A. 1996;93(20 ):11213–8. doi: 10.1073/pnas.93.20.11213. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17. Bassan M, Zamostiano R, Davidson A, Pinhasov A, Giladi E, Perl O, et al. Complete sequence of a novel protein containing a femtomolar-activity-dependent neuroprotective peptide. J Neurochem. 1999;72(3 ):1283–93. doi: 10.1046/j.1471-4159.1999.0721283.x. [DOI] [PubMed] [Google Scholar]
- 18. Gozes I. Tau pathology and future therapeutics. Curr Alzheimer Res. 2010;7(8 ):685–96. doi: 10.2174/156720510793611628. [DOI] [PubMed] [Google Scholar]
- 19. Matsuoka Y, Jouroukhin Y, Gray AJ, Ma L, Hirata-Fukae C, Li HF, 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 Ther. 2008;325(1 ):146–53. doi: 10.1124/jpet.107.130526. [DOI] [PubMed] [Google Scholar]
- 20. Shiryaev N, Jouroukhin Y, Giladi E, Polyzoidou E, Grigoriadis NC, Rosenmann H, et al. NAP protects memory, increases soluble tau and reduces tau hyperphosphorylation in a tauopathy model. Neurobiol Dis. 2009;34(2 ):381–8. doi: 10.1016/j.nbd.2009.02.011. [DOI] [PubMed] [Google Scholar]
- 21. Stewart AJ, Fox A, Morimoto BH, Gozes I. Looking for novel ways to treat the hallmarks of Alzheimer's disease. Expert Opin Investig Drugs. 2007;16(8 ):1183–96. doi: 10.1517/13543784.16.8.1183. [DOI] [PubMed] [Google Scholar]
- 22. Jakob-Roetne R, Jacobsen H. Alzheimer's disease: from pathology to therapeutic approaches. Angew Chem Int Ed Engl. 2009;48(17 ):3030–59. doi: 10.1002/anie.200802808. [DOI] [PubMed] [Google Scholar]
- 23. Nerelius C, Johansson J, Sandegren A. Amyloid beta-peptide aggregation. What does it result in and how can it be prevented? Front Biosci. 2009;14:1716–29. doi: 10.2741/3335. [DOI] [PubMed] [Google Scholar]
- 24. McLaurin J, Golomb R, Jurewicz A, Antel JP, Fraser PE. Inositol stereoisomers stabilize an oligomeric aggregate of Alzheimer amyloid beta peptide and inhibit abeta -induced toxicity. J Biol Chem Jun 16. 2000;275(24 ):18495–502. doi: 10.1074/jbc.M906994199. [DOI] [PubMed] [Google Scholar]
- 25. Townsend M, Cleary JP, Mehta T, Hofmeister J, Lesne S, O'Hare E, et al. Orally available compound prevents deficits in memory caused by the Alzheimer amyloid-beta oligomers. Ann Neurol . 2006;60(6 ):668–76. doi: 10.1002/ana.21051. [DOI] [PubMed] [Google Scholar]
- 26. Gervais F, Paquette J, Morissette C, Krzywkowski P, Yu M, Azzi M, et al. Targeting soluble Abeta peptide with Tramiprosate for the treatment of brain amyloidosis. Neurobiol Aging. 2007;28(4 ):537–47. doi: 10.1016/j.neurobiolaging.2006.02.015. [DOI] [PubMed] [Google Scholar]
- 27. Price KA, Crouch PJ, White AR. Therapeutic treatment of Alzheimer's disease using metal complexing agents. Recent Pat CNS Drug Discov. 2007;2(3 ):180–7. doi: 10.2174/157488907782411774. [DOI] [PubMed] [Google Scholar]
- 28. Bieschke J, Russ J, Friedrich RP, Ehrnhoefer DE, Wobst H, Neugebauer K, et al. EGCG remodels mature {alpha}-synuclein and amyloid-{beta} fibrils and reduces cellular toxicity. Proc Natl Acad Sci U S A. 2010. [DOI] [PMC free article] [PubMed]
- 29. Ehrnhoefer DE, Bieschke J, Boeddrich A, Herbst M, Masino L, Lurz R, et al. EGCG redirects amyloidogenic polypeptides into unstructured, off-pathway oligomers. Nat Struct Mol Biol. 2008; 15(6 ):558–66. doi: 10.1038/nsmb.1437. [DOI] [PubMed] [Google Scholar]
- 30. Nagel-Steger L, Demeler B, Meyer-Zaika W, Hochdorffer K, Schrader T, Willbold D. Modulation of aggregate size- and shape-distributions of the amyloid-beta peptide by a designed beta-sheet breaker. Eur Biophys J. 2010;39(3 ):415–22. doi: 10.1007/s00249-009-0416-2. [DOI] [PubMed] [Google Scholar]
- 31. Danho W, Swistok J, Khan W, Chu XJ, Cheung A, Fry D, et al. Opportunities and challenges of developing peptide drugs in the pharmaceutical industry. Adv Exp Med Biol. 2009;611:467–9. doi: 10.1007/978-0-387-73657-0_201. [DOI] [PubMed] [Google Scholar]
- 32. Estrada LD, Soto C. Disrupting beta-amyloid aggregation for Alzheimer disease treatment. Curr Top Med Chem. 2007;7(1 ):115–26. doi: 10.2174/156802607779318262. [DOI] [PubMed] [Google Scholar]
- 33. Lien S, Lowman HB. Therapeutic peptides. Trends Biotechnol. 2003;21(12 ):556–62. doi: 10.1016/j.tibtech.2003.10.005. [DOI] [PubMed] [Google Scholar]
- 34. Silverman DH. Brain 18F-FDG PET in the diagnosis of neurodegenerative dementias: comparison with perfusion SPECT and with clinical evaluations lacking nuclear imaging. J Nucl Med. 2004;45(4 ):594–607. [PubMed] [Google Scholar]
- 35. Dougall NJ, Bruggink S, Ebmeier KP. Systematic review of the diagnostic accuracy of 99mTc-HMPAO-SPECT in dementia. Am J Geriatr Psychiatry. 2004;12(6 ):554–70. doi: 10.1176/appi.ajgp.12.6.554. [DOI] [PubMed] [Google Scholar]
- 36. Braak H, Braak E. Frequency of stages of Alzheimer-related lesions in different age categories. Neurobiol Aging. 1997;18(4 ):351–7. doi: 10.1016/s0197-4580(97)00056-0. [DOI] [PubMed] [Google Scholar]
- 37. Thal DR, Rub U, Orantes M, Braak H. Phases of A beta-deposition in the human brain and its relevance for the development of AD. Neurology. 2002;58(12 ):1791–800. doi: 10.1212/wnl.58.12.1791. [DOI] [PubMed] [Google Scholar]
- 38. Nordberg A. Amyloid imaging in early detection of Alzheimer's disease. Neurodegener Dis. 2010;7(1-3 ):136–8. doi: 10.1159/000289223. [DOI] [PubMed] [Google Scholar]
- 39. Klunk WE, Jacob RF, Mason RP. Quantifying amyloid beta-peptide (Abeta) aggregation using the Congo red-Abeta (CR-abeta) spectrophotometric assay. Anal Biochem. 1999;266(1 ):66–76. doi: 10.1006/abio.1998.2933. [DOI] [PubMed] [Google Scholar]
- 40. Kang CK, Jayasinha V, Martin PT. Identification of peptides that specifically bind Abeta1-40 amyloid in vitro and amyloid plaques in Alzheimer's disease brain using phage display. Neurobiol Dis. . 2003;14(1 ):146–56. doi: 10.1016/s0969-9961(03)00105-0. [DOI] [PubMed] [Google Scholar]
- 41. 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. doi: 10.1039/b904138a. [DOI] [PubMed] [Google Scholar]
- 42. Schumacher TN, Mayr LM, Minor DL, Jr, Milhollen MA, Burgess MW, Kim PS. Identification of D-peptide ligands through mirror-image phage display. Science. 1996;271(5257 ):1854–7. doi: 10.1126/science.271.5257.1854. [DOI] [PubMed] [Google Scholar]
- 43. Wiesehan K, Willbold D. Mirror-image phage display: aiming at the mirror. Chembiochem: a European journal of chemical biology. [Review] 2003;4(9 ):811–5. doi: 10.1002/cbic.200300570. [DOI] [PubMed] [Google Scholar]
- 44. Gallop MA, Barrett RW, Dower WJ, Fodor SP, Gordon EM. Applications of combinatorial technologies to drug discovery. 1. Background and peptide combinatorial libraries. J Med Chem . 1994;37(9 ):1233–51. doi: 10.1021/jm00035a001. [DOI] [PubMed] [Google Scholar]
- 45. Bartnik D, Funke SA, Andrei-Selmer LC, Bacher M, Dodel R, Willbold D. Differently selected D-enantiomeric peptides act on different Abeta species. Rejuvenation research. 2010;13(2-3 ):202–5. doi: 10.1089/rej.2009.0924. [DOI] [PubMed] [Google Scholar]
- 46. 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 abeta1-42 by mirror image phage display. Chembiochem. 2003;4(8 ):748–53. doi: 10.1002/cbic.200300631. [DOI] [PubMed] [Google Scholar]
- 47. van Groen T, Kadish I, Wiesehan K, Funke SA, Willbold D. In vitro and in vivo staining characteristics of small, fluorescent, Abeta42-binding D-enantiomeric peptides in transgenic AD mouse models. ChemMedChem. 2009;4(2 ):276–82. doi: 10.1002/cmdc.200800289. [DOI] [PubMed] [Google Scholar]
- 48. Larbanoix L, Burtea C, Laurent S, Van Leuven F, Toubeau G, Vander Elst L, et al. Potential amyloid plaque-specific peptides for the diagnosis of Alzheimer's disease. Neurobiol Aging. 2010; 31(10 ):1679–89. doi: 10.1016/j.neurobiolaging.2008.09.021. [DOI] [PubMed] [Google Scholar]
- 49. 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;32:1232–43. doi: 10.1016/j.peptides.2011.04.026. [DOI] [PubMed] [Google Scholar]
- 50. Tjernberg LO, Naslund J, Lindqvist F, Johansson J, Karlstrom AR, Thyberg J, et al. Arrest of beta-amyloid fibril formation by a pentapeptide ligand. J Biol Chem. 1996;271(15 ):8545–8. doi: 10.1074/jbc.271.15.8545. [DOI] [PubMed] [Google Scholar]
- 51. Tjernberg LO, Lilliehook C, Callaway DJ, Naslund J, Hahne S, Thyberg J, et al. Controlling amyloid beta-peptide fibril formation with protease-stable ligands. J Biol Chem. 1997;272(19 ):12601–5. doi: 10.1074/jbc.272.19.12601. [DOI] [PubMed] [Google Scholar]
- 52. Matsunaga Y, Fujii A, Awasthi A, Yokotani J, Takakura T, Yamada T. Eight-residue Abeta peptides inhibit the aggregation and enzymatic activity of Abeta42. Regul Pept. 2004;120(1-3 ):227–36. doi: 10.1016/j.regpep.2004.03.013. [DOI] [PubMed] [Google Scholar]
- 53. Chafekar SM, Malda H, Merkx M, Meijer EW, Viertl D, Lashuel HA, et al. Branched KLVFF tetramers strongly potentiate inhibition of beta-amyloid aggregation. Chembiochem. 2007;8(15 ):1857–64. doi: 10.1002/cbic.200700338. [DOI] [PubMed] [Google Scholar]
- 54. Zhang G, Leibowitz MJ, Sinko PJ, Stein S. Multiple-peptide conjugates for binding beta-amyloid plaques of Alzheimer's disease. Bioconjug Chem. 2003;14(1 ):86–92. doi: 10.1021/bc025526i. [DOI] [PubMed] [Google Scholar]
- 55. Austen BM, Paleologou KE, Ali SA, Qureshi MM, Allsop D, El- Agnaf OM. Designing peptide inhibitors for oligomerization and toxicity of Alzheimer's beta-amyloid peptide. Biochemistry. 2008; 47(7 ):1984–92. doi: 10.1021/bi701415b. [DOI] [PubMed] [Google Scholar]
- 56. Taylor M, Moore S, Mayes J, Parkin E, Beeg M, Canovi M, et al. Development of a proteolytically stable retro-inverso peptide inhibitor of beta-amyloid oligomerization as a potential novel treatment for Alzheimer's disease. Biochemistry. 2010;49(15 ):3261–72. doi: 10.1021/bi100144m. [DOI] [PubMed] [Google Scholar]
- 57. Matharu B, El-Agnaf O, Razvi A, Austen BM. Development of retro-inverso peptides as anti-aggregation drugs for beta-amyloid in Alzheimer's disease. Peptides. 2010;31(10 ):1866–72. doi: 10.1016/j.peptides.2010.06.033. [DOI] [PubMed] [Google Scholar]
- 58. Bond JP, Deverin SP, Inouye H, el-Agnaf OM, Teeter MM, Kirschner DA. Assemblies of Alzheimer's peptides A beta 25-35 and A beta 31-35: reverse-turn conformation and side-chain interactions revealed by X-ray diffraction. J Struct Biol. 2003; 141(2 ):156–70. doi: 10.1016/s1047-8477(02)00625-1. [DOI] [PubMed] [Google Scholar]
- 59. Liu R, McAllister C, Lyubchenko Y, Sierks MR. Residues 17-20 and 30-35 of beta-amyloid play critical roles in aggregation. J Neurosci Res. 2004;75(2 ):162–71. doi: 10.1002/jnr.10859. [DOI] [PubMed] [Google Scholar]
- 60. Yan XZ, Qiao JT, Dou Y, Qiao ZD. Beta-amyloid peptide fragment 31-35 induces apoptosis in cultured cortical neurons. Neuroscience . 1999;92(1 ):177–84. doi: 10.1016/s0306-4522(98)00727-1. [DOI] [PubMed] [Google Scholar]
- 61. Fulop L, Zarandi M, Datki Z, Soos K, Penke B. Beta-amyloid-derived pentapeptide RIIGLa inhibits Abeta(1-42) aggregation and toxicity. Biochem Biophys Res Commun. 2004;324(1 ):64–9. doi: 10.1016/j.bbrc.2004.09.024. [DOI] [PubMed] [Google Scholar]
- 62. Fradinger EA, Monien BH, Urbanc B, Lomakin A, Tan M, Li H, et al. C-terminal peptides coassemble into Abeta42 oligomers and protect neurons against Abeta42-induced neurotoxicity. Proc Natl Acad Sci U S A. 2008;105(37 ):14175–80. doi: 10.1073/pnas.0807163105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63. Schmechel DE, Saunders AM, Strittmatter WJ, Crain BJ, Hulette CM, Joo SH, et al. Increased amyloid beta-peptide deposition in cerebral cortex as a consequence of apolipoprotein E genotype in late-onset alzheimer disease. Proc Natl Acad Sci U S A. 1993; 90(20 ):9649–53. doi: 10.1073/pnas.90.20.9649. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64. Golabek AA, Soto C, Vogel T, Wisniewski T. The interaction between apolipoprotein E and Alzheimer's amyloid beta-peptide is dependent on beta-peptide conformation. J Biol Chem. 1996; 271(18 ):10602–6. doi: 10.1074/jbc.271.18.10602. [DOI] [PubMed] [Google Scholar]
- 65. Wisniewski T, Castano EM, Golabek A, Vogel T, Frangione B. Acceleration of Alzheimer's fibril formation by apolipoprotein E in vitro. Am J Pathol. 1994;145(5 ):1030–5. [PMC free article] [PubMed] [Google Scholar]
- 66. Sadowski M, Pankiewicz J, Scholtzova H, Ripellino JA, Li Y, Schmidt SD, et al. A synthetic peptide blocking the apolipoprotein E/beta-amyloid binding mitigates beta-amyloid toxicity and fibril formation in vitro and reduces beta-amyloid plaques in transgenic mice. Am J Pathol. 2004;165(3 ):937–48. doi: 10.1016/s0002-9440(10)63355-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67. Ma J, Brewer HB, Jr, Potter H. Alzheimer A beta neurotoxicity: promotion by antichymotrypsin, ApoE4; inhibition by A betarelated peptides. Neurobiol Aging. 1996;17(5 ):773–80. doi: 10.1016/0197-4580(96)00112-1. [DOI] [PubMed] [Google Scholar]
- 68. Soto C, Kindy MS, Baumann M, Frangione B. Inhibition of Alzheimer's amyloidosis by peptides that prevent beta-sheet conformation. Biochem Biophys Res Commun. 1996;226(3 ):672–80. doi: 10.1006/bbrc.1996.1413. [DOI] [PubMed] [Google Scholar]
- 69. Ghanta J, Shen CL, Kiessling LL, Murphy RM. A strategy for designing inhibitors of beta-amyloid toxicity. J Biol Chem. 1996;271(47 ):29525–8. doi: 10.1074/jbc.271.47.29525. [DOI] [PubMed] [Google Scholar]
- 70. Gibson TJ, Murphy RM. Design of peptidyl compounds that affect beta-amyloid aggregation: importance of surface tension and context. Biochemistry. 2005;44(24 ):8898–907. doi: 10.1021/bi050225s. [DOI] [PubMed] [Google Scholar]
- 71. Kim JR, Gibson TJ, Murphy RM. Targeted control of kinetics of beta-amyloid self-association by surface tension-modifying peptides. J Biol Chem. 2003;278(42 ):40730–5. doi: 10.1074/jbc.M305466200. [DOI] [PubMed] [Google Scholar]
- 72. Kim JR, Murphy RM. Mechanism of accelerated assembly of beta-amyloid filaments into fibrils by KLVFFK(6) Biophys J. 2004;86(5 ):3194–203. doi: 10.1016/S0006-3495(04)74367-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73. Lowe TL, Strzelec A, Kiessling LL, Murphy RM. Structure-function relationships for inhibitors of beta-amyloid toxicity containing the recognition sequence KLVFF. Biochemistry. 2001; 40(26 ):7882–9. doi: 10.1021/bi002734u. [DOI] [PubMed] [Google Scholar]
- 74. Moss MA, Nichols MR, Reed DK, Hoh JH, Rosenberry TL. The peptide KLVFF-K(6) promotes beta-amyloid(1-40) protofibril growth by association but does not alter protofibril effects on cellular reduction of 3-(4,5-dimethylthiazol-2-yl)-2 5- diphenyltetrazolium bromide (MTT) Mol Pharmacol. 2003;64(5 ): 1160–8. doi: 10.1124/mol.64.5.1160. [DOI] [PubMed] [Google Scholar]
- 75. Pallitto MM, Ghanta J, Heinzelman P, Kiessling LL, Murphy RM. Recognition sequence design for peptidyl modulators of beta-amyloid aggregation and toxicity. Biochemistry. 1999;38(12 ):3570–8. doi: 10.1021/bi982119e. [DOI] [PubMed] [Google Scholar]
- 76. Sun X, Wu WH, Liu Q, Chen MS, Yu YP, Ma Y, et al. Hybrid peptides attenuate cytotoxicity of beta-amyloid by inhibiting its oligomerization: implication from solvent effects. Peptides. 2009;30(7 ):1282–7. doi: 10.1016/j.peptides.2009.04.012. [DOI] [PubMed] [Google Scholar]
- 77. Gordon DJ, Meredith SC. Probing the role of backbone hydrogen bonding in beta-amyloid fibrils with inhibitor peptides containing ester bonds at alternate positions. Biochemistry. 2003;42(2 ):475–85. doi: 10.1021/bi0259857. [DOI] [PubMed] [Google Scholar]
- 78. Gordon DJ, Sciarretta KL, Meredith SC. Inhibition of beta-amyloid( 40) fibrillogenesis and disassembly of beta-amyloid(40) fibrils by short beta-amyloid congeners containing N-methyl amino acids at alternate residues. Biochemistry. 2001;40(28 ):8237–45. doi: 10.1021/bi002416v. [DOI] [PubMed] [Google Scholar]
- 79. Etienne MA, Aucoin JP, Fu Y, McCarley RL, Hammer RP. Stoichiometric inhibition of amyloid beta-protein aggregation with peptides containing alternating alpha,alpha-disubstituted amino acids. J Am Chem Soc. 2006;128(11 ):3522–3. doi: 10.1021/ja0600678. [DOI] [PubMed] [Google Scholar]
- 80. Sato T, Kienlen-Campard P, Ahmed M, Liu W, Li H, Elliott JI, et al. Inhibitors of amyloid toxicity based on beta-sheet packing of Abeta40 and Abeta42. Biochemistry. 2006;45(17 ):5503–16. doi: 10.1021/bi052485f. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81. Rangachari V, Davey ZS, Healy B, Moore BD, Sonoda LK, Cusack B, et al. Rationally designed dehydroalanine (DeltaAla)-containing peptides inhibit amyloid-beta (Abeta) peptide aggregation. Biopolymers. 2009;91(6 ):456–65. doi: 10.1002/bip.21151. [DOI] [PubMed] [Google Scholar]
- 82. Wood SJ, Wetzel R, Martin JD, Hurle MR. Prolines and amyloidogenicity in fragments of the Alzheimer's peptide beta/A4. Biochemistry. 1995;34(3 ):724–30. doi: 10.1021/bi00003a003. [DOI] [PubMed] [Google Scholar]
- 83. Soto C, Sigurdsson EM, Morelli L, Kumar RA, Castano EM, Frangione B. Beta-sheet breaker peptides inhibit fibrillogenesis in a rat brain model of amyloidosis: implications for Alzheimer's therapy. Nat Med. 1998;4(7 ):822–6. doi: 10.1038/nm0798-822. [DOI] [PubMed] [Google Scholar]
- 84. Sigurdsson EM, Permanne B, Soto C, Wisniewski T, Frangione B. In vivo reversal of amyloid-beta lesions in rat brain. J Neuropathol Exp Neurol. 2000;59(1 ):11–7. doi: 10.1093/jnen/59.1.11. [DOI] [PubMed] [Google Scholar]
- 85. Chacon MA, Barria MI, Soto C, Inestrosa NC. Beta-sheet breaker peptide prevents Abeta-induced spatial memory impairments with partial reduction of amyloid deposits. Mol Psychiatry. 2004;9(10 ):953–61. doi: 10.1038/sj.mp.4001516. [DOI] [PubMed] [Google Scholar]
- 86. Permanne B, Adessi C, Saborio GP, Fraga S, Frossard MJ, Van Dorpe J, et al. Reduction of amyloid load and cerebral damage in a transgenic mouse model of Alzheimer's disease by treatment with a beta-sheet breaker peptide. FASEB J. 2002;16(8 ):860–2. doi: 10.1096/fj.01-0841fje. [DOI] [PubMed] [Google Scholar]
- 87. Datki Z, Papp R, Zadori D, Soos K, Fulop L, Juhasz A, et al. In vitro model of neurotoxicity of Abeta 1-42 and neuroprotection by a pentapeptide: irreversible events during the first hour. Neurobiol Dis. 2004;17(3 ):507–15. doi: 10.1016/j.nbd.2004.08.007. [DOI] [PubMed] [Google Scholar]
- 88. Juhasz G, Marki A, Vass G, Fulop L, Budai D, Penke B, et al. An intraperitoneally administered pentapeptide protects against Abeta (1-42) induced neuronal excitation in vivo. J Alzheimers Dis. 2009;16(1 ):189–96. doi: 10.3233/JAD-2009-0947. [DOI] [PubMed] [Google Scholar]
- 89. Szegedi V, Fulop L, Farkas T, Rozsa E, Robotka H, Kis Z, et al. Pentapeptides derived from Abeta 1-42 protect neurons from the modulatory effect of Abeta fibrils--an in vitro and in vivo electrophysiological study. Neurobiol Dis. 2005;18(3 ):499–508. doi: 10.1016/j.nbd.2004.12.008. [DOI] [PubMed] [Google Scholar]
- 90. Pike CJ, Walencewicz-Wasserman AJ, Kosmoski J, Cribbs DH, Glabe CG, Cotman CW. Structure-activity analyses of beta-amyloid peptides: contributions of the beta 25-35 region to aggregation and neurotoxicity. J Neurochem. 1995;64(1 ):253–65. doi: 10.1046/j.1471-4159.1995.64010253.x. [DOI] [PubMed] [Google Scholar]
- 91. Shearman MS, Ragan CI, Iversen LL. Inhibition of PC12 cell redox activity is a specific, early indicator of the mechanism of beta-amyloid- mediated cell death. Proc Natl Acad Sci U S A. 1994; 91(4 ):1470–4. doi: 10.1073/pnas.91.4.1470. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92. Terzi E, Holzemann G, Seelig J. Reversible random coil-beta-sheet transition of the Alzheimer beta-amyloid fragment (25-35) Biochemistry. 1994;33(6 ):1345–50. doi: 10.1021/bi00172a009. [DOI] [PubMed] [Google Scholar]
- 93. Doig AJ, Hughes E, Burke RM, Su TJ, Heenan RK, Lu J. Inhibition of toxicity and protofibril formation in the amyloid-beta peptide beta(25-35) using N-methylated derivatives. Biochem Soc Trans. 2002;30(4 ):537–42. doi: 10.1042/bst0300537. [DOI] [PubMed] [Google Scholar]
- 94. Hughes E, Burke RM, Doig AJ. Inhibition of toxicity in the beta-amyloid peptide fragment beta -(25-35) using N-methylated derivatives: a general strategy to prevent amyloid formation. J Biol Chem. 2000;275(33 ):25109–15. doi: 10.1074/jbc.M003554200. [DOI] [PubMed] [Google Scholar]
- 95. Gordon DJ, Tappe R, Meredith SC. Design and characterization of a membrane permeable N-methyl amino acid-containing peptide that inhibits Abeta1-40 fibrillogenesis. J Pept Res. 2002;60(1 ):37–55. doi: 10.1034/j.1399-3011.2002.11002.x. [DOI] [PubMed] [Google Scholar]
- 96. Cruz M, Tusell JM, Grillo-Bosch D, Albericio F, Serratosa J, Rabanal F, et al. Inhibition of beta-amyloid toxicity by short peptides containing N-methyl amino acids. J Pept Res. 2004;63(3 ):324–8. doi: 10.1111/j.1399-3011.2004.00156.x. [DOI] [PubMed] [Google Scholar]
- 97. Grillo-Bosch D, Carulla N, Cruz M, Sanchez L, Pujol-Pina R, Madurga S, et al. Retro-enantio N-methylated peptides as beta-amyloid aggregation inhibitors. ChemMedChem. 2009;4(9 ):1488–94. doi: 10.1002/cmdc.200900191. [DOI] [PubMed] [Google Scholar]
- 98. Findeis MA. Peptide inhibitors of beta amyloid aggregation. Curr Top Med Chem. 2002;2(4 ):417–23. doi: 10.2174/1568026024607508. [DOI] [PubMed] [Google Scholar]
- 99.Jhee. Single dose escalation study of PPI-1019, an A-beta aggregation Inhibitor. New Clinical drug evaluation unit. Boca Raaton, FL: 2003. [Google Scholar]
- 100. Kokkoni N, Stott K, Amijee H, Mason JM, Doig AJ. N-Methylated peptide inhibitors of beta-amyloid aggregation and toxicity. Optimization of the inhibitor structure. Biochemistry. 2006;45(32 ):9906–18. doi: 10.1021/bi060837s. [DOI] [PubMed] [Google Scholar]
- 101. Findeis MA, Lee JJ, Kelley M, Wakefield JD, Zhang MH, Chin J, et al. Characterization of cholyl-leu-val-phe-phe-ala-OH as an inhibitor of amyloid beta-peptide polymerization. Amyloid. 2001;8(4 ):231–41. doi: 10.3109/13506120108993819. [DOI] [PubMed] [Google Scholar]
- 102. Findeis MA, Musso GM, Arico-Muendel CC, Benjamin HW, Hundal AM, Lee JJ, et al. Modified-peptide inhibitors of amyloid beta-peptide polymerization. Biochemistry. 1999;38(21 ):6791–800. doi: 10.1021/bi982824n. [DOI] [PubMed] [Google Scholar]
- 103. 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. Angew Chem Int Ed Engl. 2009;48(11 ):1981–6. doi: 10.1002/anie.200802123. [DOI] [PubMed] [Google Scholar]
- 104. Gilead S, Gazit E. Inhibition of amyloid fibril formation by peptide analogues modified with alpha-aminoisobutyric acid. Angew Chem Int Ed Engl. 2004;43(31 ):4041–4. doi: 10.1002/anie.200353565. [DOI] [PubMed] [Google Scholar]
- 105. Smith GP, Scott JK. Libraries of peptides and proteins displayed on filamentous phage. Methods Enzymol. 1993;217:228–57. doi: 10.1016/0076-6879(93)17065-d. [DOI] [PubMed] [Google Scholar]
- 106. Blanchard BJ, Hiniker AE, Lu CC, Margolin Y, Yu AS, Ingram VM. Elimination of Amyloid beta Neurotoxicity. J Alzheimers Dis . 2000;2(2 ):137–49. doi: 10.3233/jad-2000-2214. [DOI] [PubMed] [Google Scholar]
- 107. Blanchard BJ, Konopka G, Russell M, Ingram VM. Mechanism and prevention of neurotoxicity caused by beta-amyloid peptides: relation to Alzheimer's disease. Brain Res. 1997;776(1-2 ):40–50. doi: 10.1016/s0006-8993(97)01003-2. [DOI] [PubMed] [Google Scholar]
- 108. Schwarzman AL, Tsiper M, Gregori L, Goldgaber D, Frakowiak J, Mazur-Kolecka B, et al. Selection of peptides binding to the amyloid b-protein reveals potential inhibitors of amyloid formation. Amyloid. 2005;12(4 ):199–209. doi: 10.1080/13506120500350762. [DOI] [PubMed] [Google Scholar]
- 109. Orner BP, Liu L, Murphy RM, Kiessling LL. Phage display affords peptides that modulate beta-amyloid aggregation. J Am Chem Soc. 2006;128(36 ):11882–9. doi: 10.1021/ja0619861. [DOI] [PubMed] [Google Scholar]
- 110. Taddei K, Laws SM, Verdile G, Munns S, D'Costa K, Harvey AR, et al. Novel phage peptides attenuate beta amyloid-42 catalysed hydrogen peroxide production and associated neurotoxicity. Neurobiol Aging. 2008;31(2 ):203–14. doi: 10.1016/j.neurobiolaging.2008.03.023. [DOI] [PubMed] [Google Scholar]
- 111. Baine M, Georgie DS, Shiferraw EZ, Nguyen TP, Nogaj LA, Moffet DA. Inhibition of Abeta42 aggregation using peptides selected from combinatorial libraries. J Pept Sci. 2009;15(8 ):499–503. doi: 10.1002/psc.1150. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112. 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. doi: 10.1021/cn100057j. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113. 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 ):1848–52. doi: 10.1002/cmdc.200800273. [DOI] [PubMed] [Google Scholar]
- 114. Liu H, Funke SA, Willbold D. Transport of Alzheimer disease amyloid-beta-binding D-amino acid peptides across an in vitro blood-brain barrier model. Rejuvenation Research. 2010;13(2-3 ):210–3. doi: 10.1089/rej.2009.0926. [DOI] [PubMed] [Google Scholar]
- 115. Muller-Schiffmann A, Marz-Berberich J, Andreyeva A, Ronicke R, Bartnik D, Brener O, et al. Combining independent drug classes into superior, synergistically acting hybrid molecules. Angew Chem Int Ed Engl. 2010;49(46 ):8743–6. doi: 10.1002/anie.201004437. [DOI] [PubMed] [Google Scholar]
- 116. Glomset JA. The plasma lecithins: cholesterol acyltransferase reaction. J Lipid Res. 1968;9(2 ):155–67. [PubMed] [Google Scholar]
- 117. Fielding PE, Nagao K, Hakamata H, Chimini G, Fielding CJ. A two-step mechanism for free cholesterol and phospholipid efflux from human vascular cells to apolipoprotein A-1. Biochemistry . 2000;39(46 ):14113–20. doi: 10.1021/bi0004192. [DOI] [PubMed] [Google Scholar]
- 118. Navab M, Anantharamaiah GM, Fogelman AM. The role of high-density lipoprotein in inflammation. Trends Cardiovasc Med. 2005;15(4 ):158–61. doi: 10.1016/j.tcm.2005.05.008. [DOI] [PubMed] [Google Scholar]
- 119. Paula-Lima AC, Tricerri MA, Brito-Moreira J, Bomfim TR, Oliveira FF, Magdesian MH, et al. Human apolipoprotein A-I binds amyloid-beta and prevents Abeta-induced neurotoxicity. Int J Biochem Cell Biol. 2009;41(6 ):1361–70. doi: 10.1016/j.biocel.2008.12.003. [DOI] [PubMed] [Google Scholar]
- 120. Handattu SP, Garber DW, Monroe CE, van Groen T, Kadish I, Nayyar G, et al. Oral apolipoprotein A-I mimetic peptide improves cognitive function and reduces amyloid burden in a mouse model of Alzheimer's disease. Neurobiol Dis. 2009;34(3 ):525–34. doi: 10.1016/j.nbd.2009.03.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121. Segrest JP, Jones MK, De Loof H, Brouillette CG, Venkatachalapathi YV, Anantharamaiah GM. The amphipathic helix in the exchangeable apolipoproteins: a review of secondary structure and function. J Lipid Res. 1992;33(2 ):141–66. [PubMed] [Google Scholar]
- 122. Anantharamaiah GM, Mishra VK, Garber DW, Datta G, Handattu SP, Palgunachari MN, et al. Structural requirements for antioxidative and anti-inflammatory properties of apolipoprotein A-I mimetic peptides. J Lipid Res. 2007;48(9 ):1915–23. doi: 10.1194/jlr.R700010-JLR200. [DOI] [PubMed] [Google Scholar]
- 123. Navab M, Anantharamaiah GM, Reddy ST, Fogelman AM. Apolipoprotein A-I mimetic peptides and their role in atherosclerosis prevention. Nat Clin Pract Cardiovasc Med. 2006;3(10 ):540–7. doi: 10.1038/ncpcardio0661. [DOI] [PubMed] [Google Scholar]
- 124. Shah PK, Chyu KY. Apolipoprotein A-I mimetic peptides: potential role in atherosclerosis management. Trends Cardiovasc Med. [Review] 2005;15(8 ):291–6. doi: 10.1016/j.tcm.2005.09.003. [DOI] [PubMed] [Google Scholar]
- 125. Li L, Cao D, Garber DW, Kim H, Fukuchi K. Association of aortic atherosclerosis with cerebral beta-amyloidosis and learning deficits in a mouse model of Alzheimer's disease. The American journal of pathology. 2003;163(6 ):2155–64. doi: 10.1016/s0002-9440(10)63572-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126. Kunjathoor VV, Tseng AA, Medeiros LA, Khan T, Moore KJ. beta- Amyloid promotes accumulation of lipid peroxides by inhibiting CD36-mediated clearance of oxidized lipoproteins. J Neuroinflammation. 2004;1(1 ):23. doi: 10.1186/1742-2094-1-23. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127. Cao D, Fukuchi K, Wan H, Kim H, Li L. Lack of LDL receptor aggravates learning deficits and amyloid deposits in Alzheimer transgenic mice. Neurobiology of aging. 2006;27(11 ):1632–43. doi: 10.1016/j.neurobiolaging.2005.09.011. [DOI] [PubMed] [Google Scholar]
- 128. Kawasaki T, Onodera K, Kamijo S. Selection of peptide inhibitors of soluble Abeta(1-42) oligomer formation by phage display. Bioscience, biotechnology and biochemistry. 2010;74(11 ):2214–9. doi: 10.1271/bbb.100388. [DOI] [PubMed] [Google Scholar]
- 129. Fauchere JL, Thurieau C. Evaluation of the stability of peptides and pseudopeptides as a tool in peptide drug design. Advances in Drug Research. 1992;23:127–59. [Google Scholar]
- 130. Adessi C, Soto C. Converting a peptide into a drug: strategies to improve stability and bioavailability. Curr Med Chem. 2002;9(9 ):963–78. doi: 10.2174/0929867024606731. [DOI] [PubMed] [Google Scholar]
- 131. Adessi C, Frossard MJ, Boissard C, Fraga S, Bieler S, Ruckle T, et al. Pharmacological profiles of peptide drug candidates for the treatment of Alzheimer's disease. J Biol Chem. 2003;278(16 ):13905–11. doi: 10.1074/jbc.M211976200. [DOI] [PubMed] [Google Scholar]
- 132. Hetenyi C, Szabo Z, Klement E, Datki Z, Kortvelyesi T, Zarandi M, et al. Pentapeptide amides interfere with the aggregation of beta-amyloid peptide of Alzheimer's disease. Biochem Biophys Res Commun. 2002;292(4 ):931–6. doi: 10.1006/bbrc.2002.6745. [DOI] [PubMed] [Google Scholar]
- 133. Chalifour RJ, McLaughlin RW, Lavoie L, Morissette C, Tremblay N, Boule M, et al. Stereoselective interactions of peptide inhibitors with the beta-amyloid peptide. J Biol Chem. 2003;278(37 ):34874–81. doi: 10.1074/jbc.M212694200. [DOI] [PubMed] [Google Scholar]
- 134. Pappenheimer JR, Karnovsky ML, Maggio JE. Absorption and excretion of undegradable peptides: role of lipid solubility and net charge. J Pharmacol Exp Ther. 1997;280(1 ):292–300. [PubMed] [Google Scholar]
- 135. Poduslo JF, Curran GL, Kumar A, Frangione B, Soto C. Beta-sheet breaker peptide inhibitor of Alzheimer's amyloidogenesis with increased blood-brain barrier permeability and resistance to proteolytic degradation in plasma. J Neurobiol. 1999;39(3 ):371–82. [PubMed] [Google Scholar]
- 136. Rocha S, Cardoso I, Borner H, Pereira MC, Saraiva MJ, Coelho M. Design and biological activity of beta-sheet breaker peptide conjugates. Biochem Biophys Res Commun. 2009;380(2 ):397–401. doi: 10.1016/j.bbrc.2009.01.090. [DOI] [PubMed] [Google Scholar]