Next Article in Journal
LncLocation: Efficient Subcellular Location Prediction of Long Non-Coding RNA-Based Multi-Source Heterogeneous Feature Fusion
Next Article in Special Issue
NMR Characterization of Angiogenin Variants and tRNAAla Products Impacting Aberrant Protein Oligomerization
Previous Article in Journal
Identification of the Lineage Markers and Inhibition of DAB2 in In Vitro Fertilized Porcine Embryos
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Soluble Prion Peptide 107–120 Protects Neuroblastoma SH-SY5Y Cells against Oligomers Associated with Alzheimer’s Disease

by
Elham Rezvani Boroujeni
1,2,
Seyed Masoud Hosseini
1,*,
Giulia Fani
2,
Cristina Cecchi
2 and
Fabrizio Chiti
2,*
1
Department of Microbiology and Microbial Biotechnology, Faculty of Life Science and Biotechnology, Shahid Beheshti University, Tehran 1983969411, Iran
2
Department of Experimental and Clinical Biomedical Sciences, University of Florence, Viale G.B Morgagni 50, 50134 Florence, Italy
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2020, 21(19), 7273; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms21197273
Submission received: 6 August 2020 / Revised: 27 September 2020 / Accepted: 28 September 2020 / Published: 1 October 2020
(This article belongs to the Special Issue Protein Oligomerization)

Abstract

:
Alzheimer’s disease (AD) is the most prevalent form of dementia and soluble amyloid β (Aβ) oligomers are thought to play a critical role in AD pathogenesis. Cellular prion protein (PrPC) is a high-affinity receptor for Aβ oligomers and mediates some of their toxic effects. The N-terminal region of PrPC can interact with Aβ, particularly the region encompassing residues 95–110. In this study, we identified a soluble and unstructured prion-derived peptide (PrP107–120) that is external to this region of the sequence and was found to successfully reduce the mitochondrial impairment, intracellular ROS generation and cytosolic Ca2+ uptake induced by oligomeric Aβ42 ADDLs in neuroblastoma SH-SY5Y cells. PrP107–120 was also found to rescue SH-SY5Y cells from Aβ42 ADDL internalization. The peptide did not change the structure and aggregation pathway of Aβ42 ADDLs, did not show co-localization with Aβ42 ADDLs in the cells and showed a partial colocalization with the endogenous cellular PrPC. As a sequence region that is not involved in Aβ binding but in PrP self-recognition, the peptide was suggested to protect against the toxicity of Aβ42 oligomers by interfering with cellular PrPC and/or activating a signaling that protected the cells. These results strongly suggest that PrP107–120 has therapeutic potential for AD.

1. Introduction

Alzheimer’s Disease International (ADI) and the American Alzheimer’s Association (AA) have estimated that over than 50 million people are living with dementia, and that Alzheimer’s disease (AD) is the most common cause and may account for 60–70% of dementia cases [1]. According to this, these numbers will increase dramatically, and this disease will considerably challenge the world healthcare system in the future.
Evidence indicates that AD is an aging-related disease, and frequency in individuals aged 85 or older is higher (one in three) compared to age 65 (one in nine) [1,2]. AD involves a loss of memory, cognitive impairment and behavioral instability [3]. Both extracellular neuritic plaques mainly composed of the amyloid beta (Aβ) peptide and intraneuronal neurofibrillary tangles containing the hyperphosphorylated tau protein are important histopathological hallmarks associated with AD [4,5].
Although accumulation of senile plaques formed by Aβ is a major histopathological trait of AD, it is widely accepted that soluble Aβ oligomers, forming as intermediate species in the process of neuritic plaque formation or released from the plaques, can effectively play a key role in neuronal dysfunction and impair synaptic structure and function [6,7]. It is well-known that such oligomers can inhibit long-term potentiation (LTP)—a correlate of synaptic plasticity [8,9]—as well as activating expression of the complement system [10,11] and general neurotoxic [6,12].
Several studies have reported that the cell membrane protein PrPC mediates the abnormal effects of Aβ oligomers, particularly the oligomer-induced inhibition of LTP [13,14,15]. It was also reported that memory deficits in AD transgenic mice require the presence of PrPC [16], and that loss of synaptic markers, axonal degeneration and early death in transgenic mice are fully dependent on PrPC [17]. Other reports have shown that co-expression of PrPC and Aβ reduces longevity in Drosophila melanogaster, and that expression of Aβ individually cannot create pathogenic phenotypes [18]. In spite of the large body of evidence that PrPC mediates the aberrant effects of Aβ oligomers, other studies have indicated that it is not a necessary component for the toxicity cascade induced by Aβ oligomers, as Aβ-induced LTP inhibition or memory impairment has also been reported to occur independently of the overexpression or ablation of PrPC in the transgenic mice [19,20,21]. This has led to a controversy that remains unresolved [22].
PrPC is a membrane protein that is found in both neurons and glial cells [23] and has been proposed to have a role in cellular signaling, copper homeostasis, cell adhesion and even neuroprotection [24,25]. The mature form of human PrPC consists of a single polypeptide chain of 208 amino acid residues (residues 23–230). The N-terminal region of the protein (residue 23–120) is unstructured, while the C-terminal region (residues 121–230) is largely structured, consisting of three α-helices (residues 144–154, 173–194, 200–228) and two short anti-parallel β-strands (residues 128–131, 161–164) [26]. The protein is anchored to the cell membrane via a glycosyl-phosphatidyl-inositol (GPI) anchor encompassing residues 231–253, which is excised in the mature form [26,27,28]. PrPC function is poorly understood, although roles have been suggested in synaptic transmission, long-term memory, circadian rhythms, T cell function, hematopoietic stem cell renewal, copper-binding, apoptosis and oxidative stress homeostasis, among others [29]. The scrapie form (PrPSc) is derived from PrPC [30,31,32]. The central event is the conversion from an α-helix rich structure to a form with a high β-sheet content. Accumulation of this infectious misfolded form of prion protein (PrPSc) can cause a group of neurodegenerative diseases affecting both human and animals [30,31,32].
A large number of data have shown that PrPC has a high binding affinity for Aβ42 oligomers [33,34]. Even studies showing that PrPC-expressing and PrPC knock-out mice were equally susceptible to Aβ42 oligomer-induced cognitive impairment recognized that the oligomers interacted with PrPC with high affinity [19]. The first evidence of binding between Aβ42 oligomers and PrPC dates back to 2009, when it was also found that the N-terminal region of PrPC, particularly the 95–110 residues, was involved in the binding [35]. One year later, Chen and co-workers showed that both N-terminal residues 23–27 and 92–110 were critically important for binding [36]. The importance of these two sequence regions was then confirmed by later reports [13,14,37,38,39,40,41,42].
Unlike the many reports showing a role of the N-terminal region of PrPC in the binding of Aβ oligomers, none of the studies reported so far have highlighted a role of the region encompassing residues 106–126 in Aβ42 oligomer binding [34]. By contrast, this PrP segment was found to be mainly responsible for prion aggregation [43,44,45,46,47,48,49]. The AGAAAAGA palindromic sequence 113–120, in particular, is necessary for PrPC-PrPSc interaction [44,45], and the two glycine residues at positions 114 and 119 have been suggested as particularly important for fibril formation [50].
As a segment involved in PrP self-recognition, but not in Aβ-PrPC complex formation, we reasoned that a short peptide encompassing this region of the sequence might inhibit Aβ42 oligomer toxicity. The hypothesis for this idea is that such a peptide might leave the Aβ oligomers unbound and unaltered, while engaging in interactions with PrPC. Our designed peptide encompassing residues 107–120 (PrP107–120) was found to be very soluble in physiological conditions, most probably because it lacks the highly hydrophobic region VVGGLG (residues 121–126) of the PrPC fragment 106–126 responsible for prion aggregation [44,45,49]. Importantly, the peptide was found to prevent the generic toxic effects of Aβ42 oligomers on neuroblastoma SH-SY5Y cells in absence of Aβ42 oligomer binding and structural reorganization, but in presence of partial binding between PrP107–120 peptide and endogenous cellular PrPC. Since this peptide is not aggregation-prone and has beneficial effects against Aβ-induced toxicity, it is of potentially great interest for rationalizing the pathogenesis of AD and routes to its prevention, as well as in setting up therapeutic strategies for the treatment of AD.

2. Results

2.1. Freshly Dissolved PrP107–120 is Monomeric

We calculated the theoretical hydrodynamic radius (Rh) for a peptide with the size of PrP107–120 (14 amino acid residues) in an unfolded state, using the equation described in Section 4, and previously published in [51]. According to this equation, the Rh value was found to be 0.99 ± 0.56 nm. Dynamic light scattering (DLS) shows that PrP107–120 dissolved in water has a hydrodynamic diameter (Dh) of 2.11 ± 0.68 nm, corresponding to a Rh of 1.05 ± 0.34 nm (Figure 1), which is in good agreement with that estimated theoretically. The DLS distribution also showed very large PrP107–120 aggregates at 70–7000 nm, but these are quantitatively irrelevant, as light scattering intensity is known to scale with the sixth power of the size. These results reveal that PrP107–120 dissolved in water is unfolded and predominantly non-aggregated.

2.2. PrP107–120 Remains Monomeric and Unstructured under Different Conditions

We then investigated whether different conditions that are generally favorable for protein aggregation have the ability to promote fibrillation for the designed peptide. The different conditions are listed in Section 4 and include various peptide concentrations, salt concentrations, pH values and co-solvents. As a representative example, we show the results obtained at 1.0 mg/mL peptide in 20 mM phosphate buffer, 200 mM Na2SO4, with a pH 7.0, at 37 °C (Figure 2). The DLS distribution showed a Dh of ~1 nm immediately after incubation (0 h) under these conditions, which appears dominant in the population if we take into account that the intensity of scattered light scales with the sixth power of the diameter (Figure 2A). This peak is still clearly visible even after 5 days incubation under these conditions (Figure 2A). Aggregates increased in population after 5 days, yet they appear to be quantitatively irrelevant if we consider the relationship between scattered light intensity and diameter mentioned above. The Thioflavin T (ThT) fluorescence did not increase following the addition of the peptide after 5 days incubation under these conditions relative to the blank containing only ThT (Figure 2B), and the far-UV circular dichroism (CD) spectrum also remained unchanged with a single negative peak at ~198 nm, which is typical of highly disordered states (Figure 2C). Hence, we did not observe a ThT-positive structure or significant changes in size and CD spectrum even after 5 days incubation. Similar results were obtained in all conditions tested (data not shown), indicating that PrP107–120 is soluble and stable under all the conditions studied here that are, by contrast, potentially favorable for amyloid fibril formation.

2.3. PrP107-120 Reduces Aβ42 Cytotoxicity on SH-SY5Y Cells

In order to analyze whether PrP107-120 can rescue the cellular dysfunction induced by Aβ42 oligomers, we used amyloid-derived diffusible ligands (ADDLs) formed from Aβ42 peptide according to a well-established protocol [52]. ADDLs were chosen as representative Aβ42 oligomers because they are widely used [52,53,54,55], and their morphology and purity are routinely verified [56,57]. These have been found to be toxic and increase intracellular Ca2+ and reactive oxygen species (ROS) levels in cultured cells [53,58], and have been found in post-mortem AD brains using both polyclonal and monoclonal conformation-sensitive antibodies specific to ADDLs [54,59]. To this aim, we analyzed the effects of Aβ42 ADDLs with a final concentration of 3-µM monomer equivalents (m.e.) on the metabolic activities of human SH-SY5Y cells. This immortalized neuronal cell model is mostly used for AD research, as human cholinergic neurons are difficult to obtain and maintain and unsuitable for routine experiments.
The metabolic activity of the cells was analyzed using the 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) assay, which is a widely used indicator of mitochondrial reduction capacity. The ability of SH-SY5Y cells to reduce MTT significantly decreased to 66.6 ± 4.5% following treatment for 24 h with Aβ42 ADDLs at 3 µM m.e. (Figure 3). By contrast, no detectable change was observed in cells treated with PrP107–120 and monomeric Aβ42 at final concentrations of 0.75 and 3 µM, respectively (Figure 3). The addition of the PrP107–120 (0.75 µM) to Aβ42 ADDLs (3 µM) significantly increased cell viability to 79.7 ± 4.3% compared with the cells treated with Aβ42 ADDLs (3 µM) in the absence of the peptide (Figure 3). In another experiment, PrP107–120 at a concentration of 0.75 µM was added to monomeric Aβ42 peptide at a concentration of 3 µM, and the resulting sample was maintained under conditions favorable for ADDL formation before addition to the cells (Aβ42 + PrP107–120). In this case, we found a cell viability of 86.4 ± 2.5%, again indicating protection by the PrP107–120 peptide. These results suggest that PrP107–120 can act as an inhibitor of Aβ42 ADDL oligomer toxicity.

2.4. PrP107-120 Reduces Ca2+ Influx Induced by Aβ42 Oligomers

An early cellular insult caused by Aβ42 ADDLs, and Aβ42 oligomers more generally, when added to the cellular medium, is the permeabilization of the plasma membrane with a rapid influx of calcium ions (Ca2+) from the extracellular space to the cytosol [53,58,60]. To investigate whether PrP107–120 can prevent the increase of Ca2+ levels mediated by the Aβ42 ADDL oligomers, we monitored the influx of Ca2+ in SH-SY5Y cells treated with Aβ42 ADDLs in presence and absence of PrP107–120, at the same concentrations used for the MTT assay. The quantification of the intracellular Ca2+-derived fluorescence in confocal microscopy images shows that the treatment of the cells with Aβ42 ADDLs caused a significant increase in intracellular Ca2+ up to 246 ± 21% compared with untreated cells, taken as 100% (Figure 4). By contrast, the cellular exposure to Aβ42 + PrP107–120, Aβ42 ADDLs + PrP107–120 and PrP107–120 alone triggered a minor increase of intracellular Ca2+-derived fluorescence (i.e., to 116 ± 13%, 132 ± 7% and 130 ± 11% respectively) that was significantly lower than that observed in cells treated with Aβ42 ADDLs (Figure 4). This result suggests a protective role of PrP107–120 in the Ca2+ influx mediated by the Aβ42 oligomers in neuronal cells. Since the concentrations of Ca2+ and PrP107–120 in the cell medium were 2 mM and 0.75 µM, respectively, it can be ruled out that the inhibition of ADDL-induced Ca2+ influx mediated by the prion peptide arises from peptide-Ca2+ binding, as the peptide is over three orders of magnitude sub-stoichiometric.

2.5. PrP107-120 Reduces Intracellular Reactive Oxygen Species (ROS) Induced by Aβ42 Oligomers

Another effect caused early by Aβ42 ADDLs when added to the extracellular medium of cells is the increase of ROS levels in the cytosol [58,61]. We observed an increase of ROS-derived fluorescence up to 190 ± 19% in SH-SY5Y cells treated with Aβ42 ADDLs compared with untreated cells, and no significant change in cells treated with PrP107–120 alone (Figure 5). By contrast, the treatment with Aβ42 ADDLs + PrP107–120 and Aβ42 + PrP107–120 did not cause any significant change in ROS-derived fluorescence levels compared with untreated cells, and levels significantly lower than those observed after treatment with Aβ42 ADDLs (Figure 5). Therefore, PrP107–120 can effectively protect SH-SY5Y cells against the oxidative stress induced by Aβ42 ADDLs.

2.6. PrP107-120 Reduces the Toxicity of Other Model Oligomers

In order to assess whether the protective role of PrP107–120 observed with ADDLs was specific to this peptide and oligomer system or was more generally exerted against misfolded protein oligomers, we also tested whether PrP107–120 has the ability to decrease the cytotoxicity of another type of misfolded oligomer, those formed by the model protein HypF-N named type A and found to have effects similar to those of Aβ42 [62,63,64,65]. We observed an increase of ROS-derived fluorescence in SH-SY5Y cells upon treatment with type A HypF-N oligomers up to 214 ± 39% (Figure 6). Exposure to HypF-N oligomers + PrP107–120 and HypF-N preincubated with PrP107–120 under conditions promoting type A HypF-N oligomer formation showed non-significant changes in intracellular ROS-derived fluorescence (93 ± 2% and 124 ± 19%, respectively).

2.7. PrP107-120 Reduces Aβ42 ADDLs Internalization in SH-SY5Y Cells

In order to study the effects of PrP107-120 on the cellular internalization of Aβ42 ADDLs in SH-SY5Y cells, we performed immunostaining experiments using the 6E10 specific antibody against Aβ42 and wheat germ agglutinin to stain Aβ42 and the cell membrane, respectively. Cells were exposed for 1 h to the various protein samples described above and added to the cell medium. Thus, for the intracellular Aβ42-derived fluorescence (green) in the SH-SY5Y cells, we observed more than 78% reduction in cells treated with Aβ42 ADDLs + PrP107–120, and more than 74% reduction in cells treated with Aβ42 + PrP107–120 in terms of intracellular ADDLs levels with respect to cells treated with Aβ42 ADDLs taken as 100% (Figure 7). These results indicate that PrP107–120 is able to significantly reduce Aβ42 ADDL internalization in SH-SY5Y cells when added either before or after the Aβ42 oligomerization process.

2.8. PrP107-120 Does Not Change the Structure or Aggregation State of Aβ42 ADDLs

To determine whether PrP107–120 can modify the structure of Aβ42 ADDLs into non-toxic Aβ42 oligomers or cause a change in their aggregation state (either fibrils, large aggregates or monomers), we carried out a number of tests using dot-blot, ThT fluorescence, far-UV CD and ANS fluorescence on ADDLs in the presence and absence of PrP107–120, using the same samples used for cell toxicity.
The presence of Aβ42 ADDLs was monitored by a dot-blot immunoassay using the conformation-sensitive antibody 19.3 specific for Aβ42 ADDLs [66] and the monoclonal antibody 6E10, which is able to bind all types of Aβ42 species. For Aβ42 ADDLs, Aβ42 + PrP107–120 and Aβ42 ADDLs + PrP107–120 we observed a recognition for both antibodies, whereas the PrP107–120 spot did not show any cross-reaction (Figure 8A), suggesting that PrP107–120 cannot change the structure or oligomerization state of ADDLs.
42 ADDLs did not bind ThT and did not increase its fluorescence, behavior that was not found to be affected by the PrP107–120 peptide (Figure 8B) and suggests that the peptide was not able to change the structure of the ADDLs into a stable and ThT-positive β-sheet structure. Moreover, the CD spectra observed for the Aβ42 + PrP107–120 and Aβ42 ADDLs + PrP107–120 were found to be similar to those resulting from the sum of the spectra of Aβ42 ADDLs alone and PrP107–120 alone, indicating that the PrP107–120 peptide did not significantly change the secondary structure of Aβ42 ADDLs (Figure 8C). Finally, the spectrum of ANS in the presence of Aβ42 ADDLs did not change if the ADDLs were formed in the presence of the peptide (Aβ42 + PrP107–120) or pre-incubated in the presence of the peptide after their formation (Aβ42 ADDLs + PrP107–120), suggesting again that the PrP107–120 peptide was not able to change the Aβ42 ADDL structure (Figure 8D).

2.9. PrP107-120 Does Not Colocalise with Aβ42 ADDLs but Partially Colocalises with Cellular PrPC in SH-SY5Y Cells

The inability of PrP107-120 to change the structure of Aβ42 ADDLs, yet its ability to decrease ADDL toxicity, raised two possible hypotheses: (i) PrP107-120 binds to ADDLs and shields their hydrophobic patches that are responsible for toxicity, or (ii) it interacts with the cells and renders them less vulnerable to ADDL toxicity. To verify which hypothesis was correct, the interaction of PrP107–120 and Aβ42 was studied in cell cultures. Confocal microscopy images showed an absence of co-localization between Aβ42 ADDLs (3 µM m.e.) and PrP107-120, both with ratios of 4:1 (0.75 µM PrP107–120) and 1:1 (3 µM PrP107–120), suggesting an absence of interaction between the two peptides (Figure 9). By contrast, PrP107–120 (0.75 µM) was found to partially colocalize with endogenous cellular PrPC (Figure 10, white arrows). These results also showed a significant, albeit weak, expression of PrPC in our SH-SY5Y cell system, which is in agreement with the levels of expression of the Human Protein Atlas (Figure 10).

3. Discussion

The results obtained here show that the synthetic peptide PrP107–120 is soluble and unstructured in solution and can significantly protect neuroblastoma SH-SY5Y cells against a representative form of oligomeric Aβ42, namely, ADDLs. In particular, in the presence of the peptide, Aβ42 oligomers caused a remarkably lower Ca2+ influx from the cellular medium to the cytosol, a remarkably lower increase of cellular ROS and a significantly lower alteration of mitochondrial metabolic activity. Furthermore, they had a remarkably lower ability to enter the cytosol across the cell membrane. The absence of interaction of the PrP107–120 peptide with ADDLs, and the partial colocalization of the peptide with endogenous cellular PrPC, indicates that the peptide protects the cells against Aβ42 oligomer toxicity, at least in part, by interfering with cellular PrPC.
Previous studies have suggested that the cell surface protein PrPC mediates the toxicity of Aβ42 oligomers by binding to them and affecting synaptic plasticity and other neuronal functions [13,14,15,16,17,18], although a general consensus on this point has not yet been found [19,20,21]. One PrPC region of the sequence thought to be involved in Aβ42 binding encompasses approximately residues 95–110 [13,14,33,34,35,37,38,39,40,41,42]. An antibody raised against residues 93–109 of PrPC (anti-PrPC93–109 GD11) and a synthetic peptide corresponding to residues 98–107 (PrP98–107) were found to inhibit the toxicity of the Aβ42 ADDLs to organotypic hippocampal slices, whereas a control PrP213–230 peptide did not have any such effects [37]. It was proposed that the antibody and the PrP98–107 peptide exert their effect by binding to PrPC and the diffusible Aβ42 oligomers, respectively, thus preventing their interaction in both cases [37].
In this study we have used a different perspective: Rather than using a synthetic PrPC peptide binding to Aβ42, we have used a synthetic PrPC peptide involved in PrP self-recognition (PrP107–120), with the goal of achieving a similar biological result with a different mechanism, namely, binding to membrane-anchored PrPC and impeding the binding of the latter to Aβ42 oligomers. We found that PrP107–120 significantly reduced the toxicity of Aβ42 oligomers with or without pre-incubation of the peptide with Aβ42 during the process of ADDL oligomer formation. Biophysical analyses and a dot-blot immunoassay excluded an alteration of the Aβ42 ADDL structure or aggregation state in the presence of the peptide, ruling out the hypothesis that the reduction of Aβ42 oligomer toxicity and internalization mediated by PrP107-120 is due to this reason. Moreover, no PrP107–120 colocalization with Aβ42 was found in the cell cultures using confocal microscopy even at high concentrations of PrP107–120, indicating a lack of any significant interaction between the two peptide species. In contrast, a partial colocalization between PrP107-120 and endogenous PrPC was found, indicating that the peptide can bind to the cellular prion protein. Furthermore, akin to previous observations, the role of PrPC in mediating oligomer toxicity is not restricted to Aβ42 oligomers, but also to other β-sheet rich proteins [67], and our results indicate a similar effect of PrP107–120 on Aβ42 oligomers and type-A HypF-N oligomers, used here as a positive control of toxic oligomeric species. However, we cannot completely exclude that PrP107–120 stimulates a well-defined intracellular signaling [25] and causes less vulnerability and greater resistance against Aβ42 oligomer toxicity independently of its interaction with membrane-anchored PrPC.
Overall, although the results obtained here in vitro on a cell culture line need to be validated in vivo on animal models, the present study shows the potential therapeutic value of peptides corresponding to the region of the sequence of PrPC involved in self-recognition, or other molecules potentially mimicking the same sequence trait, for the treatment of AD. This could open new avenues to the identification of the mechanism of interaction of soluble prion-derived peptides and Aβ42 oligomers, as well as to the mechanism through which PrPC mediates the toxic effects of Aβ42 oligomers.

4. Materials and Methods

4.1. PrP107–120 Preparation

The synthetic human prion protein fragment spanning residues 107–120 (PrP107–120) with the sequence Ac-TNMKHMAGAAAAGA (purity by HPLC > 95%) was purchased from Biomatik (Wilmington, DE, USA). The <5% impurities are mainly peptides with similar sequences, as are often found in solid-state peptide synthesis [68]. The counterions of peptide preparations (trifluoroacetate and guanidinium) are not supposed to be cell protectors and did not interfere with our analysis. Metal ions or other agents were not present. The lyophilized peptide was stored at −20 °C, and for each experiment 1 mg of peptide was dissolved in 1 mL of water.

4.2. Preparation of Aβ42 ADDLs, Aβ42 ADDLs + PrP107–120 and Aβ42 + PrP107–120 Samples

Lyophilized synthetic Aβ42 in a trifluoroacetate salt (Bachem, Bubendorf, Switzerland) was dissolved in pure hexafluoro-2-isopropanol (HFIP) to 1 mM. For each experiment, the solvent was evaporated using gentle nitrogen flow and the peptide was reconstituted in 2% (v/v) dimethyl sulfoxide (DMSO) and F12 Ham medium to a final concentration of 100 µM. Aβ-derived diffusible ligands (ADDLs) were prepared after 24 h incubation at 4 °C, as previously reported [52], and checked for their distinctive characteristics using atomic force microscopy, Western blotting and dot-blot, as previously described [56,57]. In this study, we used the following samples: (A) 100 µM m.e. of Aβ42 ADDLs; (B) 25 µM PrP107–120 (final concentration) added to preformed 100 µM m.e. of Aβ42 ADDLs at a 1:4 molar ratio, incubated at 4 °C for 2 h (Aβ42 ADDLs + PrP107–120); (C) 25 µM PrP107–120 (final concentration) added to 100 µM Aβ42 at a 1:4 molar ratio before Aβ42 aggregation, maintained at 4 °C for 24 h under the same conditions used to form Aβ42 ADDLs (Aβ42 + PrP107–120); and (D) 25 µM PrP107–120. All four samples were in 2% (v/v) DMSO and F-12 Ham medium and were diluted 33-fold before each experiment into cellular medium without DMSO to final concentrations of 3 and 0.75 µM (m.e.) for Aβ42 and PrP107–120, respectively.

4.3. Preparation of HypF-N Oligomers

HypF-N was purified as described previously [62] and stored at −80 °C. Before each experiment, the protein sample was thawed, centrifuged at 13,000 rpm (17950× g) for 10 min, and the concentration was measured at 280 nm. The sample was then diluted to 48 µM in 12% (v/v) trifluoroethanol, 50 mM acetate buffer and 2 mM dithiothreitol, at pH 5.5, as this condition is known to promote aggregation of HypF-N and proteins of the same structural family [62,69]. After 4 h at 25 °C, the sample was centrifuged at 12,000 rpm (15300× g) for 15 min and the pellet was dried with a gentle nitrogen flow and resuspended to 12 µM m.e. in cellular medium with or without PrP107–120, to a final concentration of 3 µM (4:1 molar ratio).

4.4. Bicinchoninic Acid (BCA) Assay

Stock bovine serum albumin (BSA) standard (Sigma-Aldrich, Saint Louis, MO, USA) was prepared at a 2-mg/mL final concentration in water. Eight BSA samples with concentrations ranging from 0 to 2000 µg/mL were prepared by dilution for the standard curve. The bicinchoninic acid (BCA) working reagent was prepared by mixing 10 mL of solution A (Bicinchoninic Acid solution, Sigma-Aldrich) with 200 µL of solution B (4% w/v CuSO4·5H2O in water). The blank and protein samples were mixed with the resulting BCA working reagent at a 1:8 ratio; thus, 25 µL of each peptide or BSA or blank samples were mixed with 200 µL of BCA working reagent and were added to 96-microplate wells. The plate was incubated at 60 °C for 15 min and, after 5 min cooling to room temperature, the absorbance of all wells was measured at 562 nm using an ultrafast BioTek Synergy H1 plate reader (Winooski, VT, USA). All absorbance values were blank subtracted. Peptide concentration was calculated by interpolation using a standard curve (absorbance versus protein concentration) obtained with the eight BSA samples.

4.5. Fibrillation of PrP107–120

The aggregation kinetics of PrP107–120 was investigated under different solution conditions at 37 °C: (A) 0.5 mg/mL peptide in 20 mM phosphate buffer, 200 mM NaCl, pH 7.0; (B) 0.5 mg/mL peptide in 20 mM HCO3, 200 mM NaCl, pH 10.5; (C) 0.5 mg/mL peptide in 20 mM phosphate buffer, 200 mM NaCl, 10% (v/v) trifluoroethanol, pH 7.0; (D) 1.0 mg/mL peptide in 20 mM acetate buffer, 200 mM NaCl, pH 4.0; (E) 1.0 mg/mL peptide in 20 mM HCO3, 200 mM NaCl, pH 10.5; (F) 1.0 mg/mL peptide in 20 mM phosphate buffer, 500 mM NaCl, pH 7.0; (G) 1.0 mg/mL peptide in 20 mM phosphate buffer, 15% (v/v) methanol, pH 7.0; (H) 1.0 mg/mL peptide in 20 mM phosphate buffer, 200 mM Na2SO4, pH 7.0. Fibril formation was monitored by the ThT fluorescence assay, DLS and CD spectroscopy for all conditions.

4.6. Dynamic Light Scattering

In a first experiment, to determine the size of monomeric PrP107–120, 1 mg of peptide was dissolved in 1 mL of water. The sample was filtered through a 0.22-µm filter and size distribution analysis was performed at 25 °C by a Zetasizer Nano S DLS device from Malvern Panalytical (Malvern, Worcestertshire, UK) thermostated with a Peltier system and checked for its reliability with polystyrene latex beads with known hydrodynamic diameter. A 10-mm plastic cell with a reduced volume was used. The refractive index and viscosity were 1.333 and 0.88 cP. The cell position and attenuator index were set automatically. The theoretical Rh for a peptide was calculated by the following equation:
Rh (Å) = (2.21 ± 1.07) N 0.57 ± 0.02
where N is the number of amino acid residues [46]. The theoretical Dh was twice the Rh value. In a second experiment, to investigate the effect of different conditions on peptide aggregation, measurements were carried out using 1 mg of peptide dissolved in 1 mL of selected buffers and following the size distribution over time for several days at 37 °C, using the same technical apparatus and cell described above. The refractive index and viscosity were set according to the various conditions.

4.7. ThT Fluorescence

ThT (Sigma-Aldrich) was dissolved to 25 µM, in 25 mM phosphate buffer, with a pH of 6.0, then filtered by a 0.45-µm filter. For a given peptide sample, 60 µL aliquots were added to 440 µL of resulting solution. A 2 × 10 mm optical path-length cuvette was used, and fluorescence spectra were recorded at 37 °C using a PerkinElmer LS 55 fluorimeter (Waltham, MA, USA) equipped with a thermostated cell holder attached to a Thermo Haake C25P water bath (Karlsrube, Germany) with an excitation wavelength of 440 nm and an emission wavelength range of 460–600 nm.

4.8. Circular Dichroism (CD) Spectroscopy

To determine the PrP107-120 polymerization kinetics in the presence of different conditions, 0.2 mg/mL of PrP107-120 peptide was used. To investigate the effect of the PrP107–120 peptide on Aβ42 oligomeric ADDLs, three samples were prepared, each 300 µL, including two separate samples of 22.2 µM Aβ42 ADDLs and one sample containing both 22.2 µM Aβ42 and 5.55 µM PrP107–120 at a 4:1 molar ratio. For removing DMSO, dialysis was performed using Spectra/Por 3 dialysis kits (MWCO 3.5 kDa, Spectrum Labs/Thermo Fisher Scientific, Waltham, MA, USA) and 10 mM phosphate buffer, with a pH of 6.0, for 3 h at 4 °C. For the preparation of Aβ42 ADDLs + PrP107–120 sample, PrP107–120 was added at a final concentration of 5.55 µM after dialysis and was incubated for 2 h at 4 °C. Far-UV CD spectra were measured at 25 °C between 190–260 nm with a 1-nm spectral step size and 50-nm/min scanning rate on a JASCO J-810 spectropolarimeter (Tokyo, Japan) equipped with thermostated cell-holder attached to a Thermo Haake C25P water bath (Karlsruhe, Germany). A 1-mm thermostated quartz cuvette was used for all CD spectra. Spectra were blank-subtracted and normalized to mean residue ellipticity.

4.9. Cell Culture

Adherent human neuroblastoma SH-SY5Y cells (A.T.C.C, Manassas, VA, USA) were cultured in Dulbecco’s Modified Eagle’s Medium (DMEM, Sigma-Aldrich), F-12 Ham with 25 mM n-(2-hydroxyethyl)piperazine-n’-(2-ethanesulfonic) acid (HEPES) and NaHCO3, supplemented with 10% fetal bovine serum (FBS), 2 mM glutamine and 1% penicillin/streptomycin and maintained at 37 °C and 5% CO2. When the cells reached 90% confluence, they were split using 0.25% trypsin-EDTA to a maximum of 20 passages.

4.10. MTT Assay

SH-SY5Y cells were plated at a density of 15 × 103 cells per well in a 96-well plate. After 24 h at 37 °C in a 5% CO2 atmosphere, cells were incubated for 24 h with different samples (Aβ42 ADDLs, Aβ42 + PrP107–120, Aβ42 ADDLs + PrP107–120 and PrP107–120, all pre-dissolved and pre-incubated in 2% (v/v) DMSO and F-12 Ham medium, as described above). Final concentrations of Aβ42 and PrP107–120 in DMEM were 3 and 0.75 µM (m.e.), respectively, with a 4:1 molar ratio. Cells were incubated with 0.5 mg/mL of MTT solution in Roswell Park Memorial Institute (RPMI) medium for 3 h at 37 °C, then with lysis buffer (20% SDS, 50% n,n-dimethylformamide, pH 4.7) for 1 h at 37 °C. The optical density was measured at 590 nm by a microplate reader (BioTek, Winooski, VT, USA).

4.11. Measurement of Intracellular Ca2+ Levels

SH-SY5Y cells were seeded on a glass coverslip in a six-well plate at a density of 40 × 103 cells per well. After 24 h, 600 µL of various samples (Aβ42 ADDLs, Aβ42 + PrP107–120, Aβ42 ADDLs + PrP107–120 and PrP107–120, all pre-dissolved and pre-incubated in 2% (v/v) DMSO and F-12 Ham medium, as described above) were added to cells for 1 h at 37 °C. Final concentrations of Aβ42 and PrP107-120 in DMEM were 3 and 0.75 µM (m.e.), respectively. After washing with phosphate buffered saline (PBS), cells were loaded with a 4-µM Fluo-4 AM probe (Invitrogen/Thermo Fisher Scientific, Waltham, MA, USA) for 10 min at 37 °C. Imaging was performed after excitation at 488 nm with a TCS SP8 confocal scanning microscopy system (Leica Microsystems, Mannheim, Germany), using a Leica Plan Apo 63x oil immersion objective, taking a series of 1-µm-thick optical sections (1024 × 1024) through the cell depth for each sample and projecting them as a single composite image by superimposition. A minimum of four images were captured for each sample and four replicates were used for each condition. Images were analyzed using Image J software (NIH, Bethesda, MD, USA).

4.12. Measurement of Intracellular Reactive Oxygen Species (ROS)

To detect intracellular accumulations of ROS, SH-SY5Y cells were cultured for 24 h on glass coverslips in a 6-well plate at a density of 40 × 103 cells per well. The medium was then replaced with 600 µL of various samples (Aβ42 ADDLs, Aβ42 + PrP107–120, Aβ42 ADDLs + PrP107–120 and PrP107–120, all pre-dissolved and pre-incubated in 2% (v/v) DMSO and F-12 Ham medium, as described above) with final concentrations of 3 and 0.75 µM (m.e.) for Aβ42 and PrP107–120 in DMEM, respectively. After 45 min, 5 µM of 2,7-dichlorodihydrofluorecein diacetate probe (CM-H2DFDA, Thermo Fisher Scientific, Waltham, MA, USA) was added for 15 min at 37 °C. Finally, cells were washed twice in PBS and then fixed in 2% (w/v) paraformaldehyde for 10 min at room temperature. Cell image acquisition was performed using the TCS SP8 confocal system described in Section 4.11. All measurements were performed in triplicates and ROS levels were calculated with Image J software (NIH). The ROS detected with this probe included mainly hydrogen peroxide [70].

4.13. Immunofluorescence Staining

SH-SY5Y cells were seeded in glass coverslips for 24 h in a 6-well plate at a density of 40 × 103 cells per well, then exposed to 600 µL of various samples (Aβ42 ADDLs, Aβ42 + PrP107–120, Aβ42 ADDLs + PrP107–120 and PrP107–120, all pre-dissolved and pre-incubated in 2% (v/v) DMSO and F-12 Ham medium, as described above) with final concentrations of 3 and 0.75 µM (m.e.) for Aβ42 and PrP107-120 in DMEM, respectively. After 1 h at 37 °C, cells were washed with PBS and stained with 1:1000 diluted Alexa Fluor 633-conjugated wheat germ agglutinin (Life Technologies/Thermo Fisher Scientific, Waltham, MA, USA) for 15 min. Cells were rinsed again and fixed with 2% (w/v) paraformaldehyde for 10 min at room temperature. After cell membrane permeabilization with 0.5% BSA in PBS + 0.5% Triton X-100, cells were incubated with 1:800 diluted mouse monoclonal 6E10 antibody (BioLegend, San Diego, CA, USA) for 1 h in 37 °C. After washing three times with PBS, cells were incubated with 1:1000 diluted Alexa Fluor 488-conjugated anti-mouse secondary antibody (Life Technologies/Thermo Fisher Scientific, Waltham, MA, USA) for 1 h. All experiments were repeated three times and representative images of confocal microscopy are presented. Fluorescence was quantified with Image J software (NIH).

4.14. Dot Blot

To determine the effect of PrP107–120 on ADDLs structure, 2 µL of each sample (Aβ42 ADDLs, Aβ42 + PrP107–120, Aβ42 ADDLs + PrP107–120 and PrP107–120, all dissolved in 2% (v/v) DMSO and F-12 Ham medium, as described above) were transferred onto a nitrocellulose membrane and allowed to dry for 15 min. For blocking the membrane, 1% (w/v) BSA in Tris-buffered saline and 0.1% Tween 20 (TBST) was used. After 1 h, the membrane was probed with 1:800 diluted mouse monoclonal antibody 6E10 (BioLegend, San Diego, CA, USA) or with 1:500 human antibody specific to ADDLs (clone 19.3, Creative Biolabs, Shirley, NY, USA) at 4 °C overnight. The day after, the membrane was washed three times and subsequently incubated for 1 h at room temperature with 1:3000 peroxidase-conjugated anti-mouse secondary antibody (Abcam, Cambridge, MA, USA) or 1:1000 peroxidase-conjugated anti-human secondary antibody (EMD Millipore, Temecula, CA, USA). Imaging was performed using an Amersham imager 600 (Cytiva, Washington DC, MD, USA).

4.15. ANS Binding Assay

The ANS solution was prepared by dissolving 30 mg of 8-anilino-1-naphthalenesulfonic acid (ANS, Sigma-Aldrich) in 10 mL of 25-µM phosphate buffer, with a pH of 6.0. After 20 min shaking, the solution was filtered by a 0.45-µm filter. ANS concentration was measured by optical absorbance at 375 nm and diluted in the same buffer to 55 µM. For each experiment, 450 µL of ANS solution was mixed with 50 µL of each sample (Aβ42 ADDLs, Aβ42 + PrP107–120, Aβ42 ADDLs + PrP107–120 and PrP107–120, all dissolved in 2% (v/v) DMSO and F-12 Ham medium, as described above) or buffer as a blank. A 2- × 10-mm optical path-length was used and fluorescence spectra were recorded at 25 °C using the same PerkinElmer LS 55 fluorimeter described above, with excitation at 380 nm and emission at 400–650 nm.

4.16. Analysis of ADDLs Co-Localization with PrP107-120

BODIPY TMR-X NHS Ester (Thermo Fisher Scientific, Waltham, MA, USA) was dissolved in DMSO. PrP107–120 was dissolved in 0.1 M NaHCO3 buffer, with a pH of 7.0. The two solutions were diluted at room temperature with continuous shaking for 1 h in the latter buffer at final concentrations of 3 mM peptide and 0.3 mM dye. Cells were cultured on glass coverslips in a 6-well plate at a density of 40 × 103 cells per well. After 24 h, cells were treated for 1 h with 600 µL of various samples (Aβ42 ADDLs, Aβ42 + labelled PrP107–120, Aβ42 ADDLs + labelled PrP107–120 and labelled PrP107–120, all pre-dissolved and pre-incubated in 2% (v/v) DMSO and F-12 Ham medium, as described above) with final concentrations of 3 and 0.75 µM (m.e.) for Aβ42 and labelled PrP107–120 in DMEM, respectively. After rinsing with PBS twice, 1:100 diluted fluorescent dye Hoechest (Immunochemistry Technologies, Bloomington, MN, USA) was added for 10 min at 37 °C. Following four washing steps with PBS, cells were fixed using 2% paraformaldehyde. Permeabilization of cells and Aβ42 staining were performed as described above. In another experiment, the concentration of labelled PrP107–120 was increased to 100 µM, and a 1:1 ratio for labelled PrP107–120 to Aβ42 was used in the initial pre-treatment in 2% (v/v) DMSO and F-12 Ham medium to prepare the Aβ42 + labelled PrP107–120 sample.

4.17. Analysis of PrP107-120 Co-Localization with PrPC

PrP107-120 was labelled with BODIPY TMR-X NHS Ester, as described in Section 4.16. SH-SY5Y cells were cultured on a glass coverslip in a 6-well plate for 24 h at a density of 40 × 103 cells per well. After washing with PBS, cells were treated with 600 µL of labelled peptide at a final concentration of 0.75 µM for 1 h. After washing with PBS twice, 2% paraformaldehyde was added for 10 min. After fixation, cells were incubated with 1:250 diluted mouse monoclonal PrP (5B2) antibody (Santa Cruz Biotechnology, Santa Cruz, CA, USA) for 1 h at 37 °C. After washing with PBS, cells were incubated with 1:1000 diluted Alexa Fluor 488-conjugated anti-mouse secondary antibody (Life Technologies/Thermo Fisher Scientific, Waltham, MA, USA). Imaging was performed after excitation at 488 and 561 nm to detect the cellular prion protein and the PrP107–120 peptide, respectively, using the TCS SP8 confocal system described in Section 4.11.

4.18. Statistical Analysis

All data are presented as means ± S.E.M. (standard error of the mean). The difference between groups was analyzed using a Student’s t-test. The single (*/#), double (**/##) and triple (***/###) symbols refer to p values < 0.05, < 0.01 and < 0.001, respectively.

Author Contributions

Conceptualization, S.M.H. and F.C.; Data curation, E.R.B. and G.F.; Formal analysis, E.R.B. and G.F.; Funding acquisition, C.C.; Investigation, E.R.B.; Methodology, C.C. and F.C.; Resources, C.C. and F.C.; Supervision, F.C.; Validation, E.R.B.; Visualization, E.R.B. and C.C.; Writing–original draft, E.R.B.; Writing–review & editing, S.M.H., G.F., C.C. and F.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was founded by Regione Toscana (FAS-Salute 2014, project SUPREMAL) and the Progetto Dipartimento di Eccellenza “Gender Medicine”.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ADAlzheimer’s disease
PrPCCellular prion protein
Amyloid beta
LTPLong-term potentiation
GPIGlycosyl-phosphatidyl-inositol
RhHydrodynamic radius
DhHydrodynamic diameter
ADDLsAmyloid-derived diffusible ligands
DMSODimethyl sulfoxide
HFIPHexafluoro-2-isopropanol
ThTThioflavin T
CDCircular dichroism
DLSDynamic light scattering
DMEMDulbecco’s Modified Eagle’s Medium
FBSFetal bovine serum
MTT3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide
ROSReactive oxygen species
RPMIRoswell Park Memorial Institute
BSABovine serum albumin
BCABicinchoninic acid

References

  1. Hebert, L.E.; Weuve, J.; Scherr, P.A.; Evans, D.A. Alzheimer disease in the United States (2010–2050) estimated using the 2010 census. Neurology 2013, 80, 1778–1783. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Bienias, J.L.; Beckett, L.A.; Bennett, D.A.; Wilson, R.S.; Evans, D.A. Design of the Chicago health and aging project (CHAP). J. Alzheimer’s Dis. 2003, 5, 349–355. [Google Scholar] [CrossRef] [PubMed]
  3. Kelley, B.J.; Petersen, R.C. Alzheimer’s disease and mild cognitive impairment. Neurol. Clin. 2007, 25, 577–609. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Goedert, M. Tau protein and the neurofibrillary pathology of Alzheimer’s disease. Trends Neurosci. 1993, 16, 460–465. [Google Scholar] [CrossRef]
  5. Huang, H.C.; Jiang, Z.F. Accumulated amyloid-β peptide and hyperphosphorylated tau protein: Relationship and links in Alzheimer’s disease. J. Alzheimer’s Dis. 2009, 16, 15–27. [Google Scholar] [CrossRef] [PubMed]
  6. Benilova, I.; Karran, E.; De Strooper, B. The toxic Aβ oligomer and Alzheimer’s disease: An emperor in need of clothes. Nat. Neurosci. 2012, 15, 349. [Google Scholar] [CrossRef]
  7. Selkoe, D.J.; Hardy, J. The amyloid hypothesis of Alzheimer’s disease at 25 years. EMBO Mol. Med. 2016, 8, 595–608. [Google Scholar] [CrossRef]
  8. Hu, N.W.; Smith, I.M.; Walsh, D.M.; Rowan, M.J. Soluble amyloid-β peptides potently disrupt hippocampal synaptic plasticity in the absence of cerebrovascular dysfunction in vivo. Brain 2008, 131, 2414–2424. [Google Scholar] [CrossRef]
  9. Li, S.; Jin, M.; Koeglsperger, T.; Shepardson, N.E.; Shankar, G.M.; Selkoe, D.J. Soluble Aβ oligomers inhibit long-term potentiation through a mechanism involving excessive activation of extrasynaptic NR2B-containing NMDA receptors. J. Neurosci. 2011, 31, 6627–6638. [Google Scholar] [CrossRef]
  10. Fischer, B.; Schmoll, H.; Platt, D.; Popa-Wagner, A.; Riederer, P.; Bauer, J. Complement C1q and C3 mRNA expression in the frontal cortex of Alzheimer’s patients. J. Mol. Med. 1995, 73, 465–471. [Google Scholar] [CrossRef]
  11. Yasojima, K.; Schwab, C.; McGeer, E.G.; McGeer, P.L. Up-regulated production and activation of the complement system in Alzheimer’s disease brain. Am. J. Pathol. 1999, 154, 927–936. [Google Scholar] [CrossRef]
  12. Larson, M.E.; Lesné, S.E. Soluble Aβ oligomer production and toxicity. J. Neurochem. 2012, 120, 125–139. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Barry, A.E.; Klyubin, I.; Mc Donald, J.M.; Mably, A.J.; Farrell, M.A.; Scott, M.; Walsh, D.M.; Rowan, M.J. Alzheimer’s disease brain-derived amyloid-β-mediated inhibition of LTP in vivo is prevented by immunotargeting cellular prion protein. J. Neurosci. 2011, 31, 7259–7263. [Google Scholar] [CrossRef] [PubMed]
  14. Freir, D.B.; Nicoll, A.J.; Klyubin, I.; Panico, S.; Mc Donald, J.M.; Risse, E.; Asante, E.A.; Farrow, M.A.; Sessions, R.B.; Saibil, H.R.; et al. Interaction between prion protein and toxic amyloid β assemblies can be therapeutically targeted at multiple sites. Nat. Commun. 2011, 2336. [Google Scholar] [CrossRef]
  15. Klyubin, I.; Nicoll, A.J.; Khalili-Shirazi, A.; Farmer, M.; Canning, S.; Mably, A.; Linehan, J.; Brown, A.; Wakeling, M.; Brandner, S.; et al. Peripheral administration of a humanized anti-PrP antibody blocks Alzheimer’s disease Aβ synaptotoxicity. J. Neurosci. 2014, 34, 6140–6145. [Google Scholar] [CrossRef] [Green Version]
  16. Um, J.W.; Strittmatter, S.M. Amyloid-β induced signaling by cellular prion protein and Fyn kinase in Alzheimer disease. Prion 2013, 7, 37–41. [Google Scholar] [CrossRef] [Green Version]
  17. Gimbel, D.A.; Nygaard, H.B.; Coffey, E.E.; Gunther, E.C.; Laurén, J.; Gimbel, Z.A.; Strittmatter, S.M. Memory impairment in transgenic Alzheimer mice requires cellular prion protein. J. Neurosci. 2010, 30, 6367–6374. [Google Scholar] [CrossRef]
  18. Younan, N.D.; Chen, K.F.; Rose, R.S.; Crowther, D.C.; Viles, J.H. Prion protein stabilizes amyloid-β (Aβ) oligomers and enhances Aβ neurotoxicity in a Drosophila model of Alzheimer’s disease. J. Biol. Chem. 2018, 293, 13090–13099. [Google Scholar] [CrossRef] [Green Version]
  19. Balducci, C.; Beeg, M.; Stravalaci, M.; Bastone, A.; Sclip, A.; Biasini, E.; Tapella, L.; Colombo, L.; Manzoni, C.; Borsello, T.; et al. Synthetic amyloid-β oligomers impair long-term memory independently of cellular prion protein. Proc. Natl. Acad. Sci. USA 2010, 107, 2295–2300. [Google Scholar] [CrossRef] [Green Version]
  20. Calella, A.M.; Farinelli, M.; Nuvolone, M.; Mirante, O.; Moos, R.; Falsig, J.; Mansuy, I.M.; Aguzzi, A. Prion protein and Aβ-related synaptic toxicity impairment. EMBO Mol. Med. 2010, 2, 306–314. [Google Scholar] [CrossRef] [Green Version]
  21. Kessels, H.W.; Nguyen, L.N.; Nabavi, S.; Malinow, R. The prion protein as a receptor for amyloid-β. Nature 2010, 466, E3–E4. [Google Scholar] [CrossRef] [Green Version]
  22. Li, S.; Selkoe, D.J. A mechanistic hypothesis for the impairment of synaptic plasticity by soluble Aβ oligomers from Alzheimer’s brain. J. Neurochem. 2020. [Google Scholar] [CrossRef] [PubMed]
  23. Bruce, M.E.; McBride, P.A.; Farquhar, C.F. Precise targeting of the pathology of the sialoglycoprotein, PrP, and vacuolar degeneration in mouse scrapie. Neurosci. Lett. 1989, 102, 1–6. [Google Scholar] [CrossRef]
  24. Roucou, X.; Gains, M.; LeBlanc, A.C. Neuroprotective functions of prion protein. J. Neurosci. Res. 2004, 75, 153–161. [Google Scholar] [CrossRef]
  25. Westergard, L.; Christensen, H.M.; Harris, D.A. The cellular prion protein (PrPC): Its physiological function and role in disease. Biochim. Biophys. Acta (BBA)-Mol. Basis Dis. 2007, 1772, 629–644. [Google Scholar] [CrossRef] [Green Version]
  26. Zahn, R.; Liu, A.; Lührs, T.; Riek, R.; von Schroetter, C.; García, F.L.; Billeter, M.; Calzolai, L.; Wider, G.; Wüthrich, K. NMR solution structure of the human prion protein. Proc. Natl. Acad. Sci. USA 2000, 97, 145–150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Zuegg, J.; Gready, J.E. Molecular dynamics simulation of human prion protein including both N-linked oligosaccharides and the GPI anchor. Glycobiology 2000, 10, 959–974. [Google Scholar] [CrossRef] [PubMed]
  28. Acevedo-Morantes, C.Y.; Wille, H. The structure of human prions: From biology to structural models—Considerations and pitfalls. Viruses 2014, 6, 3875–3892. [Google Scholar] [CrossRef]
  29. Zomosa-Signoret, V.; Arnaud, J.D.; Fontes, P.; Alvarez-Martinez, M.T.; Liautard, J.P. Physiological role of the cellular prion protein. Vet. Res. 2008, 39, 1–16. [Google Scholar] [CrossRef] [Green Version]
  30. Pan, K.M.; Baldwin, M.; Nguyen, J.; Gasset, M.; Serban, A.N.A.; Groth, D.; Mehlhorn, I.; Huang, Z.; Fletterick, R.J.; Cohen, F.E. Conversion of alpha-helices into beta-sheets features in the formation of the scrapie prion proteins. Proc. Natl. Acad. Sci. USA 1990, 90, 10962–10966. [Google Scholar] [CrossRef] [Green Version]
  31. Prusiner, S.B. Prions. Proc. Natl. Acad. Sci. USA 1998, 95, 13363–13383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Collinge, J. Prion diseases of humans and animals: Their causes and molecular basis. Annu. Rev. Neurosci. 2001, 24, 519–550. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Purro, S.A.; Nicoll, A.J.; Collinge, J. Prion protein as a toxic acceptor of amyloid-β oligomers. Biol. Psychiatry 2018, 83, 358–368. [Google Scholar] [CrossRef] [PubMed]
  34. Zhang, Y.; Zhao, Y.; Zhang, L.; Yu, W.; Wang, Y.; Chang, W. Cellular prion protein as a receptor of toxic amyloid-β42 oligomers is important for Alzheimer’s disease. Front. Cell. Neurosci. 2019, 13, 339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Laurén, J.; Gimbel, D.A.; Nygaard, H.B.; Gilbert, J.W.; Strittmatter, S.M. Cellular prion protein mediates impairment of synaptic plasticity by amyloid-β oligomers. Nature 2009, 457, 1128–1132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Chen, S.; Yadav, S.P.; Surewicz, W.K. Interaction between human prion protein and Amyloid-β (Aβ) oligomers role of N-terminal residues. J. Biol. Chem. 2010, 285, 26377–26383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Kudo, W.; Lee, H.P.; Zou, W.Q.; Wang, X.; Perry, G.; Zhu, X.; Smith, M.A.; Petersen, R.B.; Lee, H.G. Cellular prion protein is essential for oligomeric amyloid-β-induced neuronal cell death. Hum. Mol. Genet. 2012, 21, 1138–1144. [Google Scholar] [CrossRef] [Green Version]
  38. Larson, M.; Sherman, M.A.; Amar, F.; Nuvolone, M.; Schneider, J.A.; Bennett, D.A.; Aguzzi, A.; Lesné, S.E. The complex PrPc-Fyn couples human oligomeric Aβ with pathological Tau changes in Alzheimer’s disease. J. Neurosci. 2012, 32, 16857–16871. [Google Scholar] [CrossRef] [Green Version]
  39. Cox, T.O.; Gunther, E.C.; Brody, A.H.; Chiasseu, M.T.; Stoner, A.; Smith, L.M.; Haas, L.T.; Hammersley, J.; Rees, G.; Dosanjh, B.; et al. Anti-PrPC antibody rescues cognition and synapses in transgenic alzheimer mice. Ann. Clin. Transl. Neurol. 2019, 6, 554–574. [Google Scholar] [CrossRef] [Green Version]
  40. Fluharty, B.R.; Biasini, E.; Stravalaci, M.; Sclip, A.; Diomede, L.; Balducci, C.; La Vitola, P.; Messa, M.; Colombo, L.; Forloni, G.; et al. An N-terminal fragment of the prion protein binds to amyloid-β oligomers and inhibits their neurotoxicity in vivo. J. Biol. Chem. 2013, 288, 7857–7866. [Google Scholar] [CrossRef] [Green Version]
  41. Dohler, F.; Sepulveda-Falla, D.; Krasemann, S.; Altmeppen, H.; Schlüter, H.; Hildebrand, D.; Zerr, I.; Matschke, J.; Glatzel, M. High molecular mass assemblies of amyloid-β oligomers bind prion protein in patients with Alzheimer’s disease. Brain 2014, 137, 873–886. [Google Scholar] [CrossRef] [PubMed]
  42. Kostylev, M.A.; Kaufman, A.C.; Nygaard, H.B.; Patel, P.; Haas, L.T.; Gunther, E.C.; Vortmeyer, A.; Strittmatter, S.M. Prion-protein-interacting amyloid-β oligomers of high molecular weight are tightly correlated with memory impairment in multiple Alzheimer mouse models. J. Biol. Chem. 2015, 290, 17415–17438. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Gu, Y.; Fujioka, H.; Mishra, R.S.; Li, R.; Singh, N. Prion peptide 106–126 modulates the aggregation of cellular prion protein and induces the synthesis of potentially neurotoxic transmembrane PrP. J. Biol. Chem. 2002, 277, 2275–2286. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Kuwata, K.; Matumoto, T.; Cheng, H.; Nagayama, K.; James, T.L.; Roder, H. NMR-detected hydrogen exchange and molecular dynamics simulations provide structural insight into fibril formation of prion protein fragment 106–126. Proc. Natl. Acad. Sci. USA 2003, 100, 14790–14795. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Norstrom, E.M.; Mastrianni, J.A. The AGAAAAGA palindrome in PrP is required to generate a productive PrPSc-PrPC complex that leads to prion propagation. J. Biol. Chem. 2005, 280, 27236–27243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Bergström, A.L.; Chabry, J.; Bastholm, L.; Heegaard, P.M. Oxidation reduces the fibrillation but not the neurotoxicity of the prion peptide PrP106–126. Biochim. Biophys. Acta (BBA)-Proteins Proteom. 2007, 1774, 1118–1127. [Google Scholar]
  47. Walsh, P.; Simonetti, K.; Sharpe, S. Core structure of amyloid fibrils formed by residues 106–126 of the human prion protein. Structure 2009, 17, 417–426. [Google Scholar] [CrossRef] [Green Version]
  48. Walsh, P.; Neudecker, P.; Sharpe, S. Structural Properties and Dynamic Behavior of Nonfibrillar Oligomers Formed by PrP (106−126). J. Am. Chem. Soc. 2010, 132, 7684–7695. [Google Scholar] [CrossRef]
  49. Sengupta, I.; Udgaonkar, J.B. Structural mechanisms of oligomer and amyloid fibril formation by the prion protein. Chem. Commun. 2018, 54, 6230–6242. [Google Scholar] [CrossRef]
  50. Florio, T.; Paludi, D.; Villa, V.; Principe, D.R.; Corsaro, A.; Millo, E.; Damonte, G.; D’arrigo, C.; Russo, C.; Schettini, G.; et al. Contribution of two conserved glycine residues to fibrillogenesis of the 106–126 prion protein fragment. Evidence that a soluble variant of the 106–126 peptide is neurotoxic. J. Neurochem. 2003, 85, 62–72. [Google Scholar] [CrossRef]
  51. Wilkins, D.K.; Grimshaw, S.B.; Receveur, V.; Dobson, C.M.; Jones, J.A.; Smith, L.J. Hydrodynamic radii of native and denatured proteins measured by pulse field gradient NMR techniques. Biochemistry 1999, 38, 16424–16431. [Google Scholar] [CrossRef] [PubMed]
  52. Lambert, M.P.; Viola, K.L.; Chromy, B.A.; Chang, L.; Morgan, T.E.; Yu, J.; Venton, D.L.; Krafft, G.A.; Finch, C.E.; Klein, W.L. Vaccination with soluble Aβ oligomers generates toxicity-neutralizing antibodies. J. Neurochem. 2001, 79, 595–605. [Google Scholar] [CrossRef] [PubMed]
  53. Cascella, R.; Evangelisti, E.; Bigi, A.; Becatti, M.; Fiorillo, C.; Stefani, M.; Chiti, F.; Cecchi, C. Soluble oligomers require a ganglioside to trigger neuronal calcium overload. J. Alzheimer’s Dis. 2017, 60, 923–938. [Google Scholar] [CrossRef]
  54. Gong, Y.; Chang, L.; Viola, K.L.; Lacor, P.N.; Lambert, M.P.; Finch, C.E.; Krafft, G.A.; Klein, W.L. Alzheimer’s disease-affected brain: Presence of oligomeric Aβ ligands (ADDLs) suggests a molecular basis for reversible memory loss. Proc. Natl. Acad. Sci. USA 2003, 100, 10417–10422. [Google Scholar]
  55. Ghadami, S.A.; Chia, S.; Ruggeri, F.S.; Meisl, G.; Bemporad, F.; Habchi, J.; Cascella, R.; Dobson, C.M.; Vendruscolo, M.; Knowles, T.P.; et al. Transthyretin inhibits primary and secondary nucleations of amyloid-β peptide aggregation and reduces the toxicity of its oligomers. Biomacromolecules 2020, 21, 1112–1125. [Google Scholar] [CrossRef]
  56. Limbocker, R.; Chia, S.; Ruggeri, F.S.; Perni, M.; Cascella, R.; Heller, G.T.; Meisl, G.; Mannini, B.; Habchi, J.; Michaels, T.C.; et al. Trodusquemine enhances Aβ 42 aggregation but suppresses its toxicity by displacing oligomers from cell membranes. Nat. Commun. 2019, 10, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Banchelli, M.; Cascella, R.; D’Andrea, C.; Cabaj, L.; Osticioli, I.; Ciofini, D.; Li, M.S.; Skupień, K.; de Angelis, M.; Siano, S.; et al. Nanoscopic insights into the surface conformation of neurotoxic amyloid β oligomers. RSC Adv. 2020, 10, 21907–21913. [Google Scholar] [CrossRef]
  58. De Felice, F.G.; Velasco, P.T.; Lambert, M.P.; Viola, K.; Fernandez, S.J.; Ferreira, S.T.; Klein, W.L. Aβ oligomers induce neuronal oxidative stress through an N-methyl-D-aspartate receptor-dependent mechanism that is blocked by the Alzheimer drug memantine. J. Biol. Chem. 2007, 282, 11590–11601. [Google Scholar] [CrossRef] [Green Version]
  59. Lacor, P.N.; Buniel, M.C.; Furlow, P.W.; Clemente, A.S.; Velasco, P.T.; Wood, M.; Viola, K.L.; Klein, W.L. Aβ oligomer-induced aberrations in synapse composition, shape, and density provide a molecular basis for loss of connectivity in Alzheimer’s disease. J. Neurosci. 2007, 27, 796–807. [Google Scholar] [CrossRef]
  60. Demuro, A.; Mina, E.; Kayed, R.; Milton, S.C.; Parker, I.; Glabe, C.G. Calcium dysregulation and membrane disruption as a ubiquitous neurotoxic mechanism of soluble amyloid oligomers. J. Biol. Chem. 2005, 280, 17294–17300. [Google Scholar] [CrossRef] [Green Version]
  61. Cenini, G.; Cecchi, C.; Pensalfini, A.; Bonini, S.A.; Ferrari-Toninelli, G.; Liguri, G.; Memo, M.; Uberti, D. Generation of reactive oxygen species by beta amyloid fibrils and oligomers involves different intra/extracellular pathways. Amino Acids 2010, 38, 1101–1106. [Google Scholar] [CrossRef] [Green Version]
  62. Campioni, S.; Mannini, B.; Zampagni, M.; Pensalfini, A.; Parrini, C.; Evangelisti, E.; Relini, A.; Stefani, M.; Dobson, C.M.; Cecchi, C.; et al. A causative link between the structure of aberrant protein oligomers and their toxicity. Nat. Chem. Biol. 2010, 6, 140–147. [Google Scholar] [CrossRef]
  63. Zampagni, M.; Cascella, R.; Casamenti, F.; Grossi, C.; Evangelisti, E.; Wright, D.; Becatti, M.; Liguri, G.; Mannini, B.; Campioni, S.; et al. A comparison of the biochemical modifications caused by toxic and non-toxic protein oligomers in cells. J. Cell. Mol. Med. 2011, 15, 2106–2116. [Google Scholar] [CrossRef] [PubMed]
  64. Evangelisti, E.; Cecchi, C.; Cascella, R.; Sgromo, C.; Becatti, M.; Dobson, C.M.; Chiti, F.; Stefani, M. Membrane lipid composition and its physicochemical properties define cell vulnerability to aberrant protein oligomers. J. Cell Sci. 2012, 125, 2416–2427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Cascella, R.; Conti, S.; Tatini, F.; Evangelisti, E.; Scartabelli, T.; Casamenti, F.; Wilson, M.R.; Chiti, F.; Cecchi, C. Extracellular chaperones prevent Aβ42-induced toxicity in rat brains. Biochim. Biophys. Acta (BBA)-Mol. Basis Dis. 2013, 1832, 1217–1226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Savage, M.; Shughrue, P.; Wolfe, A.; McCampbell, A. Method for Detection of Amyloid Beta Oligomers in a Fluid Sample and Uses Thereof. U.S. Patent Application No. 13/544,554, 28 2013. [Google Scholar]
  67. Resenberger, U.K.; Harmeier, A.; Woerner, A.C.; Goodman, J.L.; Müller, V.; Krishnan, R.; Vabulas, R.M.; Kretzschmar, H.A.; Lindquist, S.; Hartl, F.U.; et al. The cellular prion protein mediates neurotoxic signalling of β-sheet-rich conformers independent of prion replication. EMBO J. 2011, 30, 2057–2070. [Google Scholar] [CrossRef]
  68. Picotti, P.; De Franceschi, G.; Frare, E.; Spolaore, B.; Zambonin, M.; Chiti, F.; de Laureto, P.P.; Fontana, A. Amyloid fibril formation and disaggregation of fragment 1-29 of apomyoglobin: Insights into the effect of pH on protein fibrillogenesis. J. Mol. Biol. 2007, 367, 1237–1245. [Google Scholar] [CrossRef]
  69. De Simone, A.; Dhulesia, A.; Soldi, G.; Vendruscolo, M.; Hsu, S.T.D.; Chiti, F.; Dobson, C.M. Experimental free energy surfaces reveal the mechanisms of maintenance of protein solubility. Proc. Natl. Acad. Sci. USA 2011, 108, 21057–21062. [Google Scholar] [CrossRef] [Green Version]
  70. Yang, C.; Jiang, L.; Zhang, H.; Shimoda, L.A.; DeBerardinis, R.J.; Semenza, G.L. Analysis of hypoxia-induced metabolic reprogramming. In Methods in Enzymology; Academic Press: Cambridge, MA, USA, 2014; Volume 542, pp. 425–455. [Google Scholar]
Figure 1. Size distribution of PrP107–120. Size distribution by light scattering intensity obtained with dynamic light scattering (DLS) for PrP107–120 dissolved in plain water at 25 °C. Peptide concentration was 1 mg/mL. The large aggregates at 70–7000 nm are quantitatively irrelevant, as light scattering intensity scales with the sixth power of the size.
Figure 1. Size distribution of PrP107–120. Size distribution by light scattering intensity obtained with dynamic light scattering (DLS) for PrP107–120 dissolved in plain water at 25 °C. Peptide concentration was 1 mg/mL. The large aggregates at 70–7000 nm are quantitatively irrelevant, as light scattering intensity scales with the sixth power of the size.
Ijms 21 07273 g001
Figure 2. Effect of 20 mM phosphate buffer, 200 mM Na2SO4, pH 7.0, 37 °C on PrP107–120 (1.0 mg/mL) aggregation. (A) Size distributions by light scattering intensity of the peptide sample obtained with DLS at t = 0 h (left) and 120 h (right) under the conditions described above. (B) Thioflavin T (ThT) fluorescence spectra with buffer (blank) and PrP107–120 sample after 0 and 120 h. Peptide sample was incubated as described above. ThT assay was carried out at 22 µM ThT, 0.12 mg/mL PrP107–120 (final concentrations in the cuvette), pH 6.0, 37 °C. (C) Far-UV circular dichroism (CD) spectra of PrP107–120 incubated for 0 and 120 h. Peptide sample was incubated as described above and diluted to 0.2 mg/mL in the same buffer before spectrum acquisition at 25 °C.
Figure 2. Effect of 20 mM phosphate buffer, 200 mM Na2SO4, pH 7.0, 37 °C on PrP107–120 (1.0 mg/mL) aggregation. (A) Size distributions by light scattering intensity of the peptide sample obtained with DLS at t = 0 h (left) and 120 h (right) under the conditions described above. (B) Thioflavin T (ThT) fluorescence spectra with buffer (blank) and PrP107–120 sample after 0 and 120 h. Peptide sample was incubated as described above. ThT assay was carried out at 22 µM ThT, 0.12 mg/mL PrP107–120 (final concentrations in the cuvette), pH 6.0, 37 °C. (C) Far-UV circular dichroism (CD) spectra of PrP107–120 incubated for 0 and 120 h. Peptide sample was incubated as described above and diluted to 0.2 mg/mL in the same buffer before spectrum acquisition at 25 °C.
Ijms 21 07273 g002
Figure 3. Cell viability assay in the presence of Aβ42 oligomers and PrP107–120. 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) reduction capacity of SH-SY5Y cells following 24 h treatment with Aβ42 amyloid-derived diffusible ligands (ADDLs), monomeric Aβ42, PrP107–120, Aβ42 ADDLs + PrP107–120 and Aβ42 and PrP107–120 pre-incubated under conditions promoting ADDL formation prior to addition to the cell medium. All samples were initially in 2% (v/v) dimethyl sulfoxide (DMSO) and F-12 Ham medium at concentrations of 100 and 25 µM (m.e.) for Aβ42 and PrP107–120, respectively, and were diluted 33-fold before each experiment into cellular medium without DMSO to final concentrations of 3 and 0.75 µM (m.e.) for Aβ42 and PrP107–120, respectively. The data shown are mean values ± SEM of five independent experiments. The triple asterisks (***) refer to p values < 0.001 relative to the untreated cells. The single (#) and triple (###) symbols refer to p values < 0.05 and < 0.001, respectively, relative to Aβ42 ADDLs.
Figure 3. Cell viability assay in the presence of Aβ42 oligomers and PrP107–120. 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) reduction capacity of SH-SY5Y cells following 24 h treatment with Aβ42 amyloid-derived diffusible ligands (ADDLs), monomeric Aβ42, PrP107–120, Aβ42 ADDLs + PrP107–120 and Aβ42 and PrP107–120 pre-incubated under conditions promoting ADDL formation prior to addition to the cell medium. All samples were initially in 2% (v/v) dimethyl sulfoxide (DMSO) and F-12 Ham medium at concentrations of 100 and 25 µM (m.e.) for Aβ42 and PrP107–120, respectively, and were diluted 33-fold before each experiment into cellular medium without DMSO to final concentrations of 3 and 0.75 µM (m.e.) for Aβ42 and PrP107–120, respectively. The data shown are mean values ± SEM of five independent experiments. The triple asterisks (***) refer to p values < 0.001 relative to the untreated cells. The single (#) and triple (###) symbols refer to p values < 0.05 and < 0.001, respectively, relative to Aβ42 ADDLs.
Ijms 21 07273 g003
Figure 4. Analysis of intracellular Ca2+ levels of SH-SY5Y cells treated with Aβ42 oligomers and PrP107–120. (A) Representative scanning confocal microscopy images of intracellular free Ca2+ levels in SH-SY5Y cells loaded with Fluo-4 AM probe. The cells were treated for 1 h with Aβ42 ADDLs, PrP107–120, Aβ42 ADDLs + PrP107–120 and Aβ42 + PrP107–120 pre-incubated under conditions promoting ADDL formation prior to addition to the cell medium. All samples were initially in 2% (v/v) DMSO and F-12 Ham medium at concentrations of 100 and 25 µM (m.e.) for Aβ42 and PrP107–120, respectively, and were diluted 33-fold before each experiment into cellular medium without DMSO to final concentrations of 3 and 0.75 µM (m.e.) for Aβ42 and PrP107–120, respectively. Scale bar = 15 µm. (B) Semi-quantitative analysis of intracellular Ca2+ derived fluorescence. Experimental errors are S.E.M. (n = 4). The triple (***) asterisks refer to p values < 0.001 relative to the untreated cells. The triple (###) symbols refer to p values < 0.001 relative to Aβ42 ADDLs.
Figure 4. Analysis of intracellular Ca2+ levels of SH-SY5Y cells treated with Aβ42 oligomers and PrP107–120. (A) Representative scanning confocal microscopy images of intracellular free Ca2+ levels in SH-SY5Y cells loaded with Fluo-4 AM probe. The cells were treated for 1 h with Aβ42 ADDLs, PrP107–120, Aβ42 ADDLs + PrP107–120 and Aβ42 + PrP107–120 pre-incubated under conditions promoting ADDL formation prior to addition to the cell medium. All samples were initially in 2% (v/v) DMSO and F-12 Ham medium at concentrations of 100 and 25 µM (m.e.) for Aβ42 and PrP107–120, respectively, and were diluted 33-fold before each experiment into cellular medium without DMSO to final concentrations of 3 and 0.75 µM (m.e.) for Aβ42 and PrP107–120, respectively. Scale bar = 15 µm. (B) Semi-quantitative analysis of intracellular Ca2+ derived fluorescence. Experimental errors are S.E.M. (n = 4). The triple (***) asterisks refer to p values < 0.001 relative to the untreated cells. The triple (###) symbols refer to p values < 0.001 relative to Aβ42 ADDLs.
Ijms 21 07273 g004
Figure 5. Analysis of intracellular reactive oxygen species (ROS) levels of SH-SY5Y cells treated with Aβ42 oligomers and PrP107–120. (A) Representative scanning confocal microscopy images of intracellular free ROS levels in SH-SY5Y cells loaded with CM-H2DCFDA. The cells were treated with the same samples described in the Figure 4 legend. Scale bar = 15 µm. (B) Semi-quantitative analysis of intracellular ROS-derived fluorescence. Experimental errors are S.E.M. (n = 3). The triple (***) asterisks refer to p values < 0.001 relative to the untreated cells. The triple (###) symbols refer to p values < 0.001 relative to Aβ42 ADDLs.
Figure 5. Analysis of intracellular reactive oxygen species (ROS) levels of SH-SY5Y cells treated with Aβ42 oligomers and PrP107–120. (A) Representative scanning confocal microscopy images of intracellular free ROS levels in SH-SY5Y cells loaded with CM-H2DCFDA. The cells were treated with the same samples described in the Figure 4 legend. Scale bar = 15 µm. (B) Semi-quantitative analysis of intracellular ROS-derived fluorescence. Experimental errors are S.E.M. (n = 3). The triple (***) asterisks refer to p values < 0.001 relative to the untreated cells. The triple (###) symbols refer to p values < 0.001 relative to Aβ42 ADDLs.
Ijms 21 07273 g005
Figure 6. Analysis of intracellular ROS levels of SH-SY5Y cells treated with type A HypF-N oligomers and PrP107–120. (A) Representative scanning confocal microscopy images of intracellular free ROS levels in SH-SY5Y cells loaded with CM-H2DCFDA. The cells were treated for 1 h with HypF-N oligomers, PrP107–120, HypF-N oligomers + PrP107-120 and a sample containing HypF-N and PrP107–120 pre-incubated under conditions promoting HypF-N oligomer formation prior to addition to the cell medium. Final HypF-N and PrP107–120 concentrations were 12 and 3 µM (m.e.), respectively. Scale bar = 15 µm. (B) Semi-quantitative analysis of intracellular ROS-derived fluorescence. Experimental errors are S.E.M. (n = 3). The double (**) asterisks refer to p values < 0.01 relative to the untreated cells. The double (##) and single (#) symbols refer to p values < 0.01 and < 0.05, respectively, relative to HypF-N oligomers.
Figure 6. Analysis of intracellular ROS levels of SH-SY5Y cells treated with type A HypF-N oligomers and PrP107–120. (A) Representative scanning confocal microscopy images of intracellular free ROS levels in SH-SY5Y cells loaded with CM-H2DCFDA. The cells were treated for 1 h with HypF-N oligomers, PrP107–120, HypF-N oligomers + PrP107-120 and a sample containing HypF-N and PrP107–120 pre-incubated under conditions promoting HypF-N oligomer formation prior to addition to the cell medium. Final HypF-N and PrP107–120 concentrations were 12 and 3 µM (m.e.), respectively. Scale bar = 15 µm. (B) Semi-quantitative analysis of intracellular ROS-derived fluorescence. Experimental errors are S.E.M. (n = 3). The double (**) asterisks refer to p values < 0.01 relative to the untreated cells. The double (##) and single (#) symbols refer to p values < 0.01 and < 0.05, respectively, relative to HypF-N oligomers.
Ijms 21 07273 g006
Figure 7. Effect of PrP107–120 on Aβ42 ADDL internalization in SH-SY5Y cells. (A) Representative scanning confocal microscopy images of SH-SY5Y cells treated with the indicated samples and showing Aβ42 ADDLs. The cells were treated with the same samples described in the Figure 4 legend. The cellular membrane was stained with wheat germ agglutinin (red fluorescence) and Aβ42 ADDLs were labelled with mouse monoclonal primary antibody 6E10 and anti-mouse secondary antibody (green fluorescence). Scale bar = 10 µm. (B) Semi-quantitative analysis of intracellular Aβ42 ADDL-derived fluorescence. Experimental errors are S.E.M. (n = 3). The triple (***) asterisks refer to p values < 0.001 relative to the untreated cells. The triple (###) symbols refer to p values < 0.001 relative to Aβ42 ADDLs.
Figure 7. Effect of PrP107–120 on Aβ42 ADDL internalization in SH-SY5Y cells. (A) Representative scanning confocal microscopy images of SH-SY5Y cells treated with the indicated samples and showing Aβ42 ADDLs. The cells were treated with the same samples described in the Figure 4 legend. The cellular membrane was stained with wheat germ agglutinin (red fluorescence) and Aβ42 ADDLs were labelled with mouse monoclonal primary antibody 6E10 and anti-mouse secondary antibody (green fluorescence). Scale bar = 10 µm. (B) Semi-quantitative analysis of intracellular Aβ42 ADDL-derived fluorescence. Experimental errors are S.E.M. (n = 3). The triple (***) asterisks refer to p values < 0.001 relative to the untreated cells. The triple (###) symbols refer to p values < 0.001 relative to Aβ42 ADDLs.
Ijms 21 07273 g007
Figure 8. Effect of PrP107–120 on Aβ42 ADDL structure. (A) Dot-blot immunoassay for Aβ42 ADDLs, Aβ42 + PrP107–120, Aβ42 ADDLs + PrP107–120 and PrP107–120 samples. All samples were initially in 2% (v/v) DMSO and F-12 Ham medium at concentrations of 100 and 25 µM (m.e.) for Aβ42 and PrP107–120, respectively. They were then spotted in two different nitrocellulose membranes, probed with the conformation-sensitive anti-ADDL antibody 19.3 (first line) and with conformation-insensitive anti-Aβ42 antibody 6E10 (second line). (B) ThT fluorescence assay for free ThT (blue), Aβ42 ADDLs (black), Aβ42 + PrP107–120 (red) and Aβ42 ADDLs + PrP107–120 (green). Samples were incubated as described above. ThT assay was carried out at 22 µM ThT (final concentration in the cuvette), pH 6.0, 37 °C. (C) Far-UV CD spectra for Aβ42 ADDLs (black), Aβ42 + PrP107–120 (red) and Aβ42 ADDLs + PrP107–120 (green) and PrP107–120 (purple). Samples were incubated as described above and diluted before spectrum acquisition to final concentrations of 22.2 µM Aβ42 and 5.55 µM PrP107–120, 25 °C. (D) ANS fluorescence spectra for free ANS (55 µM final concentration) and the same samples indicated in panel C. Fluorescence spectra were recorded at 25 °C.
Figure 8. Effect of PrP107–120 on Aβ42 ADDL structure. (A) Dot-blot immunoassay for Aβ42 ADDLs, Aβ42 + PrP107–120, Aβ42 ADDLs + PrP107–120 and PrP107–120 samples. All samples were initially in 2% (v/v) DMSO and F-12 Ham medium at concentrations of 100 and 25 µM (m.e.) for Aβ42 and PrP107–120, respectively. They were then spotted in two different nitrocellulose membranes, probed with the conformation-sensitive anti-ADDL antibody 19.3 (first line) and with conformation-insensitive anti-Aβ42 antibody 6E10 (second line). (B) ThT fluorescence assay for free ThT (blue), Aβ42 ADDLs (black), Aβ42 + PrP107–120 (red) and Aβ42 ADDLs + PrP107–120 (green). Samples were incubated as described above. ThT assay was carried out at 22 µM ThT (final concentration in the cuvette), pH 6.0, 37 °C. (C) Far-UV CD spectra for Aβ42 ADDLs (black), Aβ42 + PrP107–120 (red) and Aβ42 ADDLs + PrP107–120 (green) and PrP107–120 (purple). Samples were incubated as described above and diluted before spectrum acquisition to final concentrations of 22.2 µM Aβ42 and 5.55 µM PrP107–120, 25 °C. (D) ANS fluorescence spectra for free ANS (55 µM final concentration) and the same samples indicated in panel C. Fluorescence spectra were recorded at 25 °C.
Ijms 21 07273 g008
Figure 9. Absence of co-localization of Aβ42 ADDLs with PrP107–120. Cells were treated for 1 h with the same samples described in the Figure 4 legend and indicated on top of the images (added to the cell medium). PrP107–120 was labelled with BODIPY TMR-X NHS Ester (red fluorescence), Aβ42 ADDLs with mouse monoclonal primary antibody 6E10 and a secondary antibody (green fluorescence) and the nuclei with the Hoechst dye (blue florescence). Scale bar = 10 µm.
Figure 9. Absence of co-localization of Aβ42 ADDLs with PrP107–120. Cells were treated for 1 h with the same samples described in the Figure 4 legend and indicated on top of the images (added to the cell medium). PrP107–120 was labelled with BODIPY TMR-X NHS Ester (red fluorescence), Aβ42 ADDLs with mouse monoclonal primary antibody 6E10 and a secondary antibody (green fluorescence) and the nuclei with the Hoechst dye (blue florescence). Scale bar = 10 µm.
Ijms 21 07273 g009
Figure 10. Partial co-localization of PrP107-120 with endogenous cellular PrPC. Cells were treated for 1 h with the PrP107–120 sample described in the Figure 9 legend and indicated on top of the images (added to the cell medium). PrP107–120 was labelled with BODIPY TMR-X NHS Ester (red fluorescence) and cellular PrPC was labelled with mouse monoclonal primary antibody PrP (5B2) and Alexa Fluor 488-conjugated anti-mouse secondary (green fluorescence). The white arrows indicate the co-localization spots in the merged images. Scale bar = 10 µm.
Figure 10. Partial co-localization of PrP107-120 with endogenous cellular PrPC. Cells were treated for 1 h with the PrP107–120 sample described in the Figure 9 legend and indicated on top of the images (added to the cell medium). PrP107–120 was labelled with BODIPY TMR-X NHS Ester (red fluorescence) and cellular PrPC was labelled with mouse monoclonal primary antibody PrP (5B2) and Alexa Fluor 488-conjugated anti-mouse secondary (green fluorescence). The white arrows indicate the co-localization spots in the merged images. Scale bar = 10 µm.
Ijms 21 07273 g010

Share and Cite

MDPI and ACS Style

Rezvani Boroujeni, E.; Hosseini, S.M.; Fani, G.; Cecchi, C.; Chiti, F. Soluble Prion Peptide 107–120 Protects Neuroblastoma SH-SY5Y Cells against Oligomers Associated with Alzheimer’s Disease. Int. J. Mol. Sci. 2020, 21, 7273. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms21197273

AMA Style

Rezvani Boroujeni E, Hosseini SM, Fani G, Cecchi C, Chiti F. Soluble Prion Peptide 107–120 Protects Neuroblastoma SH-SY5Y Cells against Oligomers Associated with Alzheimer’s Disease. International Journal of Molecular Sciences. 2020; 21(19):7273. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms21197273

Chicago/Turabian Style

Rezvani Boroujeni, Elham, Seyed Masoud Hosseini, Giulia Fani, Cristina Cecchi, and Fabrizio Chiti. 2020. "Soluble Prion Peptide 107–120 Protects Neuroblastoma SH-SY5Y Cells against Oligomers Associated with Alzheimer’s Disease" International Journal of Molecular Sciences 21, no. 19: 7273. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms21197273

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop