1. Introduction
Alzheimer’s disease (AD) is the most prevalent form of dementia affecting millions of people worldwide. It is characterized by the presence of pathological markers such as senile plaques and neurofibrillary tangles in the brain of AD patients. Senile plaques are mainly composed of aggregated Amyloid-β peptide (Aβ; product of amyloid precursor protein (APP) cleavage) [
1]. Aβ peptides can range from 36 to 43 amino acids depending on the enzyme cleavage sites, but the most common pathological forms are Aβ
40 and Aβ
42 [
1,
2]. Aβ
42 is considered more toxic and highly prone to aggregation and accumulation as compared to Aβ
40. Another pathological feature, prominently found in AD brains, is cholinergic dysfunction that has been linked to cognitive decline and memory impairment in AD patients [
3]. Indeed, loss of basal nuclei cholinergic neurons is shown to be directly associated with cognitive impairment in AD [
4,
5]. However, the exact cause of cholinergic deficit and how it is related to other pathological changes in AD is still puzzling researchers around the world.
Cholinergic neurons, defined by the expression of the acetylcholine (ACh) synthesizing enzyme choline acetyltransferase (ChAT; EC 2.3.1.6), are projected from certain nuclei in the basal forebrain throughout the brain [
6] and regulate key cognitive functions such as attention, memory, and learning. It has been shown that the cholinergic signaling machinery, including high-affinity choline transporter (hChT) and ChAT, is highly susceptible to Aβ peptides and its aggregated forms [
4,
7,
8,
9].
The innate physiological function of Aβ peptides is still unknown and the focus of the research community has been on the toxicity of these peptides and their aggregates. However, studies have shown that Aβ can act as a ligand of RAGE (receptor for advanced glycation end products) and regulate key inflammatory signaling pathways [
10]. In addition, physiological concentrations of Aβ
42 peptide can activate the MAPK (mitogen-activated protein kinase) cascade in the hippocampus via the α7 nicotinic acetylcholine receptor [
11]. In mammals, MAPK cascade is involved in learning and memory formation [
12]. Most importantly, recent studies have suggested a more prominent role of Aβ peptides in synaptic and extrasynaptic ACh homeostasis. We have previously shown that Aβ peptides interact with acetyl- and butyryl-cholinesterase (AChE and BChE) in an apolipoprotein facilitated manner, and thereby can modulate cholinergic signaling by forming highly stable and ultra-active soluble ACh-degrading complexes called BAβACs [
13].
Other studies have shown that short- and long-term exposure to Aβ peptides can lead to a reduced ACh synthesis and release from neuronal cell cultures as well as neurodegeneration and cognitive impairment in animal models [
14,
15,
16,
17,
18,
19,
20,
21,
22,
23,
24,
25]. However, it is not clear whether these effects are caused by a direct or an indirect mode of action of Aβ peptides on the activity of the core cholinergic enzyme ChAT. In contrast, we have shown that Aβ peptides, in particular Aβ
42, directly interact and allosterically enhance rather than reduce the catalytic rate of ChAT at physiological concentrations representative of Aβ levels in the CSF of AD patients and healthy controls (
Figure 1) [
26].
Given that cholinergic deficit is one of the key features in major dementia disorders, the findings that Aβ peptides can allosterically increase biosynthesis of ACh opens a new avenue for the development of a new class of cholinergic enhancing therapeutic strategy based on improving ACh biosynthesis that we are calling ChAT-Potentiating-Ligands (CPLs), as opposed to the cholinesterase inhibitors approach based on inhibiting the degradation of ACh by cholinesterases. In the primary paper of our study, we have reported that Aβ
40 exhibited CPL activity with an EC
50 value of 800 pg/mL (or 185 pM) while Aβ
42 peptides showed a 10-fold higher CPL potency as deduced by an EC
50 value of 68 pg/mL (or 15 pM) [
26]. These EC
50 values are fully in the range of the expected concentration of Aβ peptides in the human brain, as deduced by the known concentration of Aβ peptides in human cerebrospinal fluid. Thereby, it is imperative to determine the molecular fingerprint of the interaction between ChAT and the Aβ peptides as the only known endogenous CPL.
Therefore, the current study had two main objectives. Firstly, we aimed to shed further light on the physiological role of Aβ peptides in ACh regulation, and secondly, to investigate and elucidate the molecular fingerprint of the binding mechanism of Aβ peptides on ChAT governing the CPL activity of the Aβ peptides, using advanced in silico molecular docking and molecular dynamics studies. We report here the results of both the docking and 100 ns molecular simulation analyses in terms of the molecular dynamic parameters, such as root means square deviation (RMSD), radius of gyration (Rg), root mean structure fluctuation (RMSF) and H-bond functionality values. We show that these analyses indicate that Aβ peptides can physically interact with ChAT protein, forming highly stable ChAT-Aβ complexes as deduced by a free-energy landscape analysis. In addition, we provide details on the stability and robustness of the ChAT-Aβ peptides simulated system.
2. Results
Amyloid-β and its exact native function are still questions of debate among researchers worldwide. In our recently published paper, we have shown that Aβ peptides can upregulate the activity of the core-cholinergic enzyme ChAT by direct interaction at physiological concentration ranges and this enhanced catalytic efficiency was persistent even at the physiological ratio of Aβ
40:Aβ
42 mixture (
Figure 1; adapted and reproduced from Kumar et al. [
26], under the terms of the CC BY license). Here, molecular docking studies were performed using a Cluspro protein-protein docking web server to elucidate the possible interaction mechanism of Aβ peptides with ChAT. The final protein-protein docking cluster analysis identified 29 clusters for ChAT-Aβ
40 and ChAT-Aβ
42 complexes. The most probable binding cluster for ChAT-Aβ
40 complex Cluster-0 with a maximum of 88 members and the lowest energy (−989.5 kcal/mol). The corresponding was Cluster-0 for ChAT-Aβ
42 complex had the maximum members of 134 and the lowest energy of −999.6 kcal/mol (
Table 1).
The predicted models were well justified based on the scores, as it implies the linear combinations of several energy terms, such as the attractive and repulsive van der Walls energy describing the complementary shape, and one or more terms denoting the electrostatic binding energy and cluster sizes. The 3D cartoon representations for the top three clusters of ChAT-Aβ
40 are given in (
Figure 2B–D). The putative entrances of the catalytic tunnel on ChAT are indicated by a
red circle (for choline/ACh entrance site) and
yellow circle (for ACoA/-CoA entrance site;
Figure 2). Intriguingly, in the cluster-1, Aβ
40 seemed to completely cover the mouth of the choline/ACh entrance site of ChAT’s catalytic tunnel (
red circle in
Figure 2C), whereas the most probable cluster-0 indicated that Aβ
40 nearly touched the ACoA/-CoA entrance site on ChAT (
yellow circle in
Figure 2B) as well as the cluster-2 indicating that the Aβ
40 covers the mouth of ACoA/-CoA entrance site on ChAT (
yellow circle in
Figure 2D). Thus, these binding cluster analyses intuitively suggest that the Aβ
40 binding at cluster-1, but not the cluster-0 or -2, could block the influx of the substrate, choline or ACh, and thereby the access of the enzyme to these substrates. Importantly, this could explain why in some of our in vitro experiments, the Aβ
40 either inhibited or induced minimal activation of ChAT (
Figure 1C) [
26].
For ChAT-Aβ
42 complex, the top three clusters are presented as a 3D cartoon in (
Figure 2F–H), illustrating the most probable binding sites of Aβ
42 on ChAT. Again, the position of the choline/ACh entrance site of the catalytic tunnel is highlighted by a
red circle (in
Figure 2E).
Surprisingly, like Aβ
40, Aβ
42 appeared to completely overlap the catalytic tunnel, but in this case it was in the most probable cluster-0 (
Figure 2F), and not in the cluster-1 like Aβ
40. Nonetheless, there are two additional binding sites for Aβ
42 that are in close vicinity of the mouth of the ACoA/-CoA entrance site on the catalytic tunnel, namely the clusters -1 and -2 (
yellow circle;
Figure 2G,H). These sites might best explain the improved catalytic efficiency of ChAT by Aβ
42 peptides (compare
Figure 1A,B). In addition, we have compiled the simultaneous possible bindings for Aβ
42 and Aβ
40 as depicted in (
Figure 2I). For instance, binding of Aβ
42 at cluster-1 (
Figure 2G) will most likely exclude binding of Aβ
40 at cluster-0 (
Figure 2B) or vice versa. In contrast, it is also possible that two Aβ bind simultaneously on a ChAT molecule, e.g., conditions C-H or F-H in (
Figure 2I).
Next, the predictions of the binding mode of Aβ peptides to ChAT by the docking analyses were challenged with advanced molecular dynamic (MD) simulation analysis in order to realize the conformational changes and stability during the course of interaction. The time required for the biological phenomenon to occur is considered as the optimum time for simulation. Based on the resources available, the calculations were tuned and, keeping our objective in mind to fulfil the criteria that the RMSD stabilizes over the period of time, were indicated by a stable line that concludes that the system has reached convergence and is comparatively stable. To explore the global minima in the energy landscape, the MD trajectories were run and analyzed under physiological environment for a 100 ns-long simulation time period in triplicates. The results of the molecular dynamic analysis of the first primary run are shown in (
Figure 3) as a measure of the stability and quality of the simulated system. The RMSD plot suggested that the ChAT protein remained stable from 0.1–0.28 nm, as shown in (
Figure 3A). The averages of the triplicates were plotted, as represented in
Supplementary Figure S1. The results are well within the limits with low uncertainties, which are denoted as the ‘Y’ bar error (standard deviation) obtained from three independent 100 ns simulations. The RMSD value of Aβ peptides for the first run (
Figure 3B) seems to be in the ranges of 0.3–1.73 nm, which indicates their stability in the complex system throughout the simulation time. Likewise, the averages obtained from the triplicate runs were plotted, as shown in
Supplementary Figure S2. The standard deviation is larger compared to the protein structure, which is due to the flexibility of the peptide as it interacts with the protein surface throughout the course of simulation. The radius of gyration (RoG) of the complex was also determined, as shown in (
Figure 3C), to further understand the compactness and overall dimensions of the protein structure. The RoG values for the first run were in line with the RMSD values as well as stable and relatively low ranging between 2.5 and 2.65 nm for the ChAT-Aβ complex. This indicated that the formed complex retained a compact folded conformation throughout the simulation time. The individual RoG plots for all three runs of the ChAT-Aβ complex clusters are shown in
Supplementary Figure S3. The results from the triplicates indicate that the RoG values remained stable with relatively low variation that were in an acceptable range.
Furthermore, to understand the dynamic behaviors of the amino acid residues and flexibility of the protein structure during the interaction period of the simulation, we recorded and analyzed the root mean square fluctuation (RMSF) as a measure of individual amino acid fluctuation throughout the course of simulation. The magnitude of individual amino acid residual flexibility is represented by the corresponding height of the peaks. RMSF plots of ChAT protein for the first run indicate no major fluctuation in the ChAT core structure (
Figure 3D), with a value of 0.05–0.4 nm for the amino acid residues. However, the C- and N-terminals of the ChAT protein exhibited higher structural flexibility with values ranging up to 0.7 nm. This is expected since they are naturally exposed to solvent surface. The average RMSF values of the ChAT protein, along with the standard deviation for the three runs are shown in
Supplementary Figure S4, which shows low fluctuations. Likewise, the RMSF value of the Aβ peptides (
Figure 3E) was observed between the acceptable limits of 0.1–1.2 nm indicating the relative stability of the Aβ peptides in the system. The C- and N-terminal of Aβ showed higher RMSF value, indicating the higher flexibility of the regions due to the unstructured nature, as reported based on NMR structural studies [
27,
28]. The average RMSF values of the Aβ peptides, along with the standard deviation for the three runs is shown in
Supplementary Figure S5. The larger variation most likely reflects the higher flexibility of the C- and N-terminal of Aβ peptides. Overall, the RMSF suggested that Aβ was tightly bound with the ChAT protein, despite these C- and N-terminal flexibilities and the overall ChAT-Aβ complexes remained stable for the whole simulated time.
Hydrogen bonding plays a central role during the interactions, formation and stabilization of protein complexes. The typical distance criterion for a hydrogen bond donor and acceptor atoms is ≤3.5 Å (0.35 nm) with an angle of 180° ± 30° between hydrogen bond donor and acceptor. Therefore, we also calculated the average number of hydrogen bonds formed during ChAT-Aβ peptide interactions throughout the 100 ns simulation with a cut-off value of 0.35 nm, as shown for the first run in
Figure 4A. The individual landscapes for the hydrogen bonds formed between ChAT-Aβ peptides during all the three runs is shown in
Supplementary Figure S6. The results were consistent with respect to each other, which indicates that a stable number of hydrogen bonds were maintained throughout the simulation period. The average hydrogen bond distance was also calculated, which was found to be about 0.3 nm, which is within the expected range (
Figure 4B). The results from all the triplicates were very similar (
Supplementary Figure S7), with an average hydrogen bond distance of about 0.3 nm.
Moreover, we performed a solvent accessible surface area (SASA) analysis, which measures the surface area of the biomolecule accessible by the solvent (biofluids), and helps in predicting the stability of the hydrophobic core of the protein. These hydrophobic contacts are crucial indicators of the compactness of the tertiary structures of a protein [
29]. We found that the SASA profile of each ChAT-Aβ cluster complex is consistent with its radius of gyration as there was a slight decrease in SASA. This in turn indicates a contracted conformation of the complex, as a higher SASA profile reflects a relative expansion of the protein surface area and would have resulted in a more fluctuating radius of gyration. The values ranged from 250–275 nm
2 for ChAT proteins (
Figure 4C) and 39–49 nm
2 for Aβ peptides (
Figure 4D) for the first run. The averages from the triplicates along with their standard deviation is shown in
Supplementary Figures S8 and S9. The results were consistent with reasonably low variations in the reproduction of the SASA profiles. Furthermore, SASA for the ChAT protein seemed to be reduced at the end of the simulation (
Figure 4C), indicating that the Aβ peptides are interacting with the ChAT protein and are causing changes in the structure and surface residues of the ChAT protein in such a way that it further shields its hydrophobic core accessible to the solvent. Whilst for the Aβ peptides, SASA showed an increase towards the end of the simulation with constant changes due to the interaction with ChAT protein.
To unravel the connection between the internal motion and the secondary structures, such as the coil, beta sheets, beta-bridge, bend, turn, and helices throughout the simulation time, a Dictionary of Secondary Structure of Protein (DSSP) analysis was instigated, which identifies protein sheets and helix assignments solely on the basis of backbone–backbone hydrogen bonds. We performed TESSE (time-evolution of the secondary structural elements) analyses for ChAT and Aβ peptides, which are visually summarized in
Supplementary Figures S10 and S11, respectively. The analyses in (
Figure S10) further reveal that the ChAT protein maintained its secondary structure throughout the simulation period without much conformational change, since all the amino acid residues maintained their structure during the simulation period.
Moreover, we observed that the ChAT protein retained a greater number of alpha helices, which is about 40% of the amino acids remained in this state, as compared to coils, beta sheet, turn, bend, 5-helices, and 3-helices. This observation is based on data presented in
Table 2 and
Table 3, which show the percentage of secondary structure distribution in the trajectory clusters for Aβ
40 and Aβ
42 on ChAT in each cluster system, respectively. The occurrence and maintenance of alpha helices is the most stable dominant structural element. This was a common feature of protein secondary structure in the ChAT protein throughout the simulation period, reflecting its compactness, energetically most favorable conformations, and slow steady change throughout the simulation. On the other hand, the Aβ peptides seemed to be constantly changing over the simulation time period, as is evident by containing a greater number of turns and random coils (
Figure S11), which could reflect an interplay with the ChAT catalytic mouth of Choline/ACh entrance site.
We did a Principal Component Analysis (PCA), also known as essential dynamics based on the principles of covariance matrix, to identify the most important conformational degrees of freedom of the simulation system. It helps in shortening the large dimensions of the data set to the major principal components that represents the considerable variations that can sufficiently explain the overall motions of the protein. The eigenvalues for the complex are shown in (
Figure 5), where it is clearly evident that the first four components explain the most variation in the data which is 53.58, 78.71, and 52.70 for cluster-0, -1, and -2 of Aβ
40 peptides, respectively, and in case of Aβ
42 peptides it is about 72.85, 49.25, and 52.43 for the cluster-0, -1, and -2, respectively. The plot shows that the eigenvalues after the fourth principal component starts to make a straight horizontal line. Thus, we have selected the first two principal components to plot our essential dynamics plot.
The plotted 2D graph for the first two principal components (
Figure 6), represents the various conformations taken by the protein in the course of simulated period. The changes in the ChAT-Aβ complex system indicates that Aβ peptides affected the flexibility of the ChAT protein. ChAT-Aβ
40 cluster-0 and -2 and ChAT-Aβ
42 cluster-1 and -2 complexes demonstrated concentrated conformational sampling, indicating a stable state conformation occupied by the overall system. While the ChAT-Aβ
40 cluster-1 and the ChAT-Aβ
42 cluster-0 display a widespread distribution of the conformational space, this can be due to larger conformational changes in the protein structure, thus indicating that a wide range of conformational states are occupied by the complex over the simulation time.
Free-energy landscape (FEL) is a representation of possible conformations taken by a protein in molecular dynamics simulation along with the Gibbs free energy. FEL represents two variables that reflect specific properties of the system and measure conformational variability. To visualize the energy minima landscape of bound ChAT-Aβ complex, we studied the free-energy landscape (FEL) against radius of gyration (RoG) and root-mean-square deviation (RMSD) as the two reaction coordinates. This revealed the changes in the Gibbs free energy (ΔG) values between 0 and 9.950 kJ/mol for Aβ
40 (
Figure 7) and from 0 to 10 kJ/mol for Aβ
42 (
Figure 8) complexes.
The shape and size of the minimal energy area (shown in blue) indicate the stability of the ChAT-Aβ complex system. Smaller and more centralized blue areas represent the complex within the cluster with the highest stability. The narrow funnel formed as seen in the 3D projections show the dynamic of the change in conformation with the time of simulation required to attain a native structure with least energy. Overall, the 3D plots illustrate that the overall complexes formed a single funnel, which together with the 2D contour plot clearly indicates the complexes having one local energy minima, and hence reflecting a stable folding process in the system.
Thereafter, we used the insights from the free-energy landscape analysis and constructed the energy minima structures (
Figure 9). These structures were in good agreement with those constructed through molecular docking analysis, reinforcing the insight about the potentially stable binding sites for Aβ peptides on ChAT protein.
The 3D cartoon representations for the top three clusters of ChAT-Aβ
40 are given in (
Figure 9K–M). The putative entrances of the catalytic tunnel on ChAT are indicated by a
red circle (for choline/ACh entrance site;
Figure 9J) and a
yellow circle (for ACoA/-CoA entrance site;
Figure 9). Intriguingly, in the cluster-1 (Aβ
40 from the energy minima conformations) it can be appreciated that the complex seems to stabilize in a way that the N-terminal of the Aβ
40 tends to hinder the opening of the choline/ACh entrance site of ChAT’s catalytic tunnel (
red circle in
Figure 9L), whereas the most probable cluster-0 indicated that Aβ
40 completely covered the ACoA/-CoA entrance site on ChAT (
yellow circle in
Figure 9K) as well as the cluster-2 indicating that the Aβ
40 covers the mouth of ACoA/-CoA entrance site on ChAT (
yellow circle in
Figure 9M). Thus, these binding clusters analyses intuitively suggest that the Aβ
40 binding at cluster-1 but not the cluster-0 or -2 could block the influx of the substrate, choline or ACh, and thereby the access of the enzyme to these substrates.
For ChAT-Aβ
42 complex, the top three clusters are presented as a 3D cartoon in
Figure 9O–Q, illustrating the most probable binding sites of Aβ
42 on ChAT. Again, the position of the choline/ACh entrance site of the catalytic tunnel is highlighted by a
red circle (in
Figure 9N). The most probable cluster-0 (
Figure 9O) of Aβ
42 appeared to completely overlap the catalytic tunnel. Nonetheless, there are two additional binding sites for Aβ
42, namely the clusters-1 and -2 (
yellow circle;
Figure 9P,Q) that completely block and are in close vicinity of the mouth of the ACoA/-CoA entrance site on the catalytic tunnel, respectively. In addition, we have compiled the simultaneous possible bindings for Aβ
42 and Aβ
40 as depicted in (
Figure 9R). For instance, binding of Aβ
42 at cluster-1 (
Figure 9P) will most likely exclude binding of Aβ
40 at cluster-0 (
Figure 9K), or vice versa. In contrast, it is also possible that two Aβ bind simultaneously on a ChAT molecule, e.g., conditions L-Q or O-Q in (
Figure 9R). The final conformations obtained at the end of the simulation of the ChAT-Aβ complexes are depicted in
Supplementary Figure S12. These were resembling the lowest energy conformations (shown in
Figure 9) indicating that towards the end of the simulation period, the ChAT-Aβ complexes stabilized into a low energy conformation.
Furthermore, we have generated the contacts map for the ChAT-Aβ complexes (
Figure 10 and
Figure 11). It shows the amino acid residues of the ChAT protein that make a close contact with the Aβ peptides. This can give hints about the necessary residues that are more likely to be interacting with each other during the complex formation.
3. Discussion
Our previous report clearly revealed that mainly Aβ
42 peptides potentiate the catalytic rate of ChAT by 20–30%. In contrast, Aβ
40 potentiates ChAT activity in some analysis, but inhibits it in the others (compare
Figure 1A and
Figure 1C). To shed light on such discrepancies, as well as to elucidate the molecular fingerprint of the interaction between Aβ peptides and ChAT governing the CPL activity of the Aβ peptides, we conducted the current investigation using two in silico approaches, namely a molecular docking and a 100 ns-long molecular dynamics (MD) simulation analysis.
Here, we show that the molecular docking analysis identified three high probability clusters for Aβ40, and three for Aβ42. Interestingly, molecular docking identified high-probability ChAT-Aβ complex clusters for both Aβ40 and Aβ42 that had overlapping binding sites with the mouth of the Choline/ACh entrance site of the ChAT catalytic tunnel, which is expected to inhibit rather than enhance ChAT catalytic efficiency. Nonetheless, we also identified binding clusters that blocked the ACoA/-CoA entrance site.
While the overlapping binding cluster of Aβ
40 with the choline/ACh entrance site could explain the inconsistency observed for the mode of action of Aβ
40 peptides on ChAT in the in vitro experiments, where in some experiments Aβ
40 inhibited, and in others it enhanced ChAT activity. Yet, such an entrance-overlapping cluster fails to explain why Aβ
42 did not exhibit this dual mode of action. Nonetheless, there were some differences that could shed light on this difference. (1) Molecular docking and MD analyses suggested that Aβ
42 had two additional high probability clusters that did not overlap with the mouth of the choline/ACh catalytic tunnel of ChAT (compare
Figure 2 and
Figure 9). (2) In addition, molecular docking analysis indicated that the Aβ
42 had two clusters with a binding site in close vicinity of the entrance of ACoA/-CoA into the catalytic tunnel (
Figure 2G,H) of which cluster-1 seems to cover the entrance of ACoA/-CoA in Lowest Energy Minima conformer obtained from MD analysis (
Figure 9P). In turn, such alternative bindings could mediate a positive effect on the substrate influx into the catalytic domain upon Aβ
42 interaction that is absent in case of Aβ
40, and thereby on the catalytic efficiency of ChAT, in a manner that we have previously reported for the interaction of Aβ with BuChE [
13].
Of note, there are two fundamental differences that exist between the catalytic modes of cholinesterases (AChE and BuChE) and ChAT. Firstly, ChAT can act reversibly, meaning that it can both produce and breakdown ACh, while ChEs only degrade ACh. Thus, ChAT always produces two products: ACh and -CoA in the forward reaction but choline and Acetyl-CoA in the reverse reaction. Secondly, the rate-limiting step in the cycle of ACh synthesis is the release of -CoA from its binding site (that is why -CoA is reported to act as an inhibitor of ChAT at concentrations higher than 50µM) [
30].
Overall, the following hypothetical explanations may account for the observed changes in the catalytic activity of ChAT in the presence of Aβ peptides as reported by us (
Figure 1) and the conflicting reports by others. (1) Some of the binding clusters may indicate that binding of Aβ, in particular Aβ
42, induces a conformational change in the enzyme structure, facilitating a faster release of -CoA from its binding sites and allowing the enzyme a faster re-entry into a new ACh cycle production. (2) Other in silico clusters may indicate that Aβ might reduce the reverse reaction of ChAT i.e., prevent the conversion of ACh and -CoA to choline and Acetyl-CoA. In other words, in these binding clusters, Aβ mainly directs the forward ACh production reaction. This could occur by favoring the entrance of Choline and/or Acetyl-CoA from their respective entrances or prevent/reduce the entrance of ACh and -CoA into the catalytic domain. (3) Given that both Aβ
42 and Aβ
40 are expected to be present together in vivo, the most likely scenario is that both alternative mechanisms are at work simultaneously, allowing Aβ
42 to mainly act as ChAT potentiating ligand (CPL) by implementing both alternatives, while allowing Aβ
40 to act as a weak CPL or a blocker of Aβ
42 and ChAT activity. The net effect would be a controlled increase in ACh production in some situations but not in others. However, based on the current data, we can neither exclude nor confirm any of these scenarios, although the overall in vitro data favors the third alternative, as indicated by the data shown in
Figure 1, where a biological 10:1 ratio of Aβ
40 to Aβ
42 resulted in ~15% CPL activity compared to ~25% for Aβ
42 alone.
The fourth hypothetical explanation concerns with the possibility of dual Aβ binding or exclusion of binding. For instance, the binding of Aβ
42 at the cluster-1 site seems to be excluding the binding of Aβ
42 to the cluster-2 site but not the cluster-0 site (
Figure 2F–H). This may increase the probability of simultaneous binding of two Aβ
42 peptides at the opposite cluster-0 and -1 site (or -0 and -2) on ChAT protein (
Figure 2F and 2G or 2F and 2H). In contrast, the binding sites for cluster-0 and cluster-1 for Aβ
40 are too close to each other, which most likely could exclude the simultaneous binding of two Aβ
40 peptides on the hAT protein (
Figure 2B,C). As Aβ
40 binds at cluster-1, but not at cluster -0 or -2, and completely blocking the entrance of choline/ACh into the catalytic tunnel, we can expect a 50% chance for either activation or inhibition, explaining the inconsistency in the activation of ChAT by Aβ
40 peptides.
For the docking and molecular dynamics analysis, we randomly chose one out of several conformations for Aβ
40 and Aβ
42 structures that were present in PDB deposition. Nonetheless, the Aβ
40 and Aβ
42 peptides have disordered/unstructured N- and C-terminals and can exist in several different conformations as shown by structural NMR studies [
27,
28]. In addition, despite the successful docking analyses by ClusPro docking server in the prediction of the binding mode of Aβ peptides to ChAT, it should be mentioned that the analyses represented a rigid-body docking, which does not allow conformational changes during protein–protein interactions. To overcome these limitations, we performed an advanced 100 ns molecular dynamics simulation on the representative clusters. These molecular dynamics analyses not only confirmed the unstructured nature of C- and N-terminals in Aβ peptides, but also provided evidence that the ChAT-Aβ complexes were stable, supporting a plausible mode of interaction between these two proteins. Overall, these in silico analyses strongly confirm a direct physical interaction between Aβ peptides and the ChAT protein, which further support our published in vitro enzyme kinetic analyses.