You are on page 1of 10

Biophysical Chemistry 297 (2023) 107011

Contents lists available at ScienceDirect

Biophysical Chemistry
journal homepage: www.elsevier.com/locate/biophyschem

Exploring the aggregation of amyloid-β 42 through Monte


Carlo simulations
Priya Dey , Parbati Biswas *
Department of Chemistry, University of Delhi, Delhi 110007, India

A R T I C L E I N F O A B S T R A C T

Keywords: Coarse-grained Monte Carlo simulations are performed for a disordered protein, amyloid-β 42 to identify the
Metropolis Monte Carlo simulations interactions and understand the mechanism of its aggregation. A statistical potential is developed from a selected
Coarse grained Cα chain backbone model dataset of intrinsically disordered proteins, which accounts for the respective contributions of the bonded and
Intrinsically disordered proteins
non-bonded potentials. While, the bonded potential comprises the bond, bend, and dihedral constraints, the
Aggregation propensity
Non-local contacts
nonbonded interactions include van der Waals interactions, hydrogen bonds, and the two-body potential. The
PMF profile two-body potential captures the features of both hydrophobic and electrostatic interactions that brings the chains
at a contact distance, while the repulsive van der Waals interactions prevent them from a collapse. Increased two-
body hydrophobic interactions facilitate the formation of amorphous aggregates rather than the fibrillar ones.
The formation of aggregates is validated from the interchain distances, and the total energy of the system. The
aggregate is structurally characterized by the root-mean-square deviation, root-mean-square fluctuation and the
radius of gyration. The aggregates are characterized by a decrease in SASA, an increase in the non-local in­
teractions and a distinct free energy minimum relative to that of the monomeric state of amyloid-β 42. The
hydrophobic residues help in nucleation, while the charged residues help in oligomerization and aggregation.

1. Introduction However, functional amyloids also exist in other forms of life like bac­
teria and fungi [7]. Specifically, Alzheimer's disease is caused by the
Proteins are susceptible to misfolding in a heterogeneous intracel­ aggregation of the intrinsically disordered protein amyloid-β, which is
lular environment, where they are locally trapped in some of the low- derived from the amyloid precursor protein.
energy near-native conformations rather than the thermodynamically Amyloid-β (Aβ) is a 40 or 42 residue protein that causes Alzheimer's
stable functional native state, which represents the free energy mini­ disease via the formation of soluble oligomers, which is reported to be
mum in the folding energy landscape [1,2]. These misfolded proteins the principal cytotoxic species as compared to the mature fibrils [8].
tend to associate to form dimers and oligomers, which subsequently self- However, in-vitro studies have demonstrated a significantly higher rate
assemble as fibrillar aggregates with characteristic cross beta-sheet of fibril formation in Aβ − 42 as compared to that of Aβ − 40 [9–11].
structures known as amyloids [3]. The formation of these fibrils in Oligomers may be either structured or unstructured. Since oligomers are
vitro are a generic attribute of many proteins, especially intrinsically difficult to characterize experimentally, computer simulations and ki­
disordered proteins, under specified experimental conditions. Intrinsi­ netic models are employed to provide significant insights regarding the
cally disordered proteins do not fold to a unique native structure but structure, stability and the potential mechanism of Aβ − 42 aggregation
rather exist as a dynamic ensemble of interconvertible conformations [12].
[4]. Such proteins are mostly amyloidogenic and readily form amyloids Protein aggregation involves a multitude of both length and time
[5]. Amyloids are highly toxic and disrupt the normal physiological scales that may be interpreted via a hierarchy of models ranging from
functioning of the cell [6]. Protein aggregation is often implicated in a simple coarse-grained models to the precise atomistic ones. [13]. Ag­
broad spectrum of debilitating human disorders, that comprise more gregation of protein monomers and oligomers were extensively studied
than 50 identified neurodegenerative diseases like Alzheimer's, Parkin­ using all-atom molecular dynamics simulations [14,15]. Such calcula­
son's, Huntington's diseases, type 2 diabetes and some forms of cancer. tions are however limited due to the high computational cost imposed

* Corresponding author.
E-mail address: pbiswas@chemistry.du.ac.in (P. Biswas).

https://doi.org/10.1016/j.bpc.2023.107011
Received 3 February 2023; Received in revised form 25 March 2023; Accepted 26 March 2023
Available online 5 April 2023
0301-4622/© 2023 Elsevier B.V. All rights reserved.
P. Dey and P. Biswas Biophysical Chemistry 297 (2023) 107011

by the size of the system and sampling of the protein conformational


space. Therefore, coarse-grained models of proteins at varying levels of
details constitute an essential component for exploring various generic
features of protein aggregation [16]. Such studies primarily investigated
the factors leading to fibril nucleation and formation. The precise
mechanism of aggregation is protein specific and varies from one protein
to the other. Atomistic simulations both with and without solvent are
used to probe the early events of the aggregation process for small
proteins [17,18]. Several simulation studies have been done on mono­
mers [19–21], formation of dimers [22] and oligomers [23,24], amyloid
fibril elongation [25,26] and amyloid fibril stability [27–29]. Recent
molecular dynamics (MD) simulation studies of Aβ − 42 and Aβ frag­
ments have shown that a β-hairpin structure promotes the formation of
intermolecular β-sheet structures [30,31]. A lattice model was devel­
oped to study protein aggregation in three-dimensional cubic lattices
using MC simulations [32] to investigate an interplay between amor­
phous aggregation and fibril formation. The oligomerization of Aβ
(16–22) peptides was explored using Metropolis MC simulations, which
demonstrated the spontaneous formation of low energy aggregates like
small oligomers and fibril-like structures [33]. Some simulation studies
have also confirmed that oligomers are more neurotoxic than amyloid
fibrils [34–36].
In this work, aggregation of the intrinsically disordered protein,
amyloid-β 42 is investigated via Metropolis [37] Monte Carlo simulation
algorithm based on a coarse-grained Cα chain backbone model [38]. A
coarse-grained statistical potential is developed that captures the salient
features of aggregation through hydrophobic effect, hydrogen bonding,
electrostatic and nonbonded interactions. The formation of protein ag­
gregates is analyzed by evaluating various structural properties like the
root-mean-square deviation (RMSD), root-mean-square fluctuation
(RMSF) and the radius of gyration (Rg), while the role of the non-bonded
interactions are quantified in terms of the intrachain and interchain
residue contacts. The structural transition from an extended monomeric
chain to a compact aggregate is reflected in the variation of the
conformational properties. The specific roles of the local and nonlocal
interactions in protein aggregation are investigated while the free en­
ergy profile is calculated from the potential of mean force. Fig. 1. Protein chain systems contained in the box with various size to keep the
system concentrations constant.
2. Methodology
DISPROT database is used to design a statistical potential that comprises
A coarse-grained Cα chain backbone model of the amyloid-β 42 both bonded and non-bonded contributions due to both interchain and
protein is considered where a residue is represented by a single Cα atom intrachain residue interactions. [41]. The bonded interactions include
while the side-chain conformations are neglected. The initial structure the energy function that accounts for the bond stretches, bond angles
of monomeric amyloid-β 42 (Aβ-42) protein is selected from the and dihedrals. These interactions are modeled as suitable constraints
Research Collaboratory for Structural Bioinformatics (RCSB) PDB with [42]. A data-set of 1135 intrinsically disordered proteins [43] is
the PDB ID:1Z0Q. Aβ-42 is derived by successive endoproteolytic pro­ compiled from the RCSB PDB and the distribution of bond lengths, Cα(i) -
cessing of the transmembrane amyloid precursor protein (APP) by β- and Cα(i + 1), bond angles Cα(i) - Cα(i + 1) - Cα(i + 2) and the dihedral angles,
γ-secretases [39]. The sequence of the amyloid-β 42 monomer is Cα(i) - Cα(i + 1) - Cα(i + 2) - Cα(i + 3) between all constituent residues are
‘DAEFRHDSGYEVHHQKLVFFAEDVGSNKGAIIGLMVGGVVIA'. Since calculated.
simulations of a single monomeric chain does not capture the interchain The range of the pseudo bond lengths and pseudo bend angles be­
interactions that lie at the core of the aggregation process, multi-chain tween the neighboring Cα backbone atoms is calculated from Fig. S1 and
systems are prepared using Packmol software [40]. The number den­ Fig. S2 of the Supplementary Material as 3.65–4.10 Å and 75◦ –160◦ ,
sity of the monomeric chains in the simulation box is kept constant, respectively. To restrict the dihedral angle, three most populated regions
which fixes the concentration of the system as shown in Fig. 1. The 10◦ –70◦ , 160◦ –180◦ and (− 180◦ ) – (− 100◦ ) are accessed and sampled
relevant information about the dimension of the boxes and the number from the distribution plots (refer to Fig. S3 of the Supplementary Ma­
density is given in Table S1 of the Supplementary Material. Systems with terial). The non-bonded interactions include van der Waals interactions,
two to eight chains are simulated and examined, while the findings for hydrogen bonding [44,45], and the distance dependent pairwise in­
the dimer, trimer and tetramer are analyzed and discussed. teractions between the non-bonded residues [46].

2.1. Potential 2.2. van der Waals interactions

The development of an appropriate force field is a key step in protein A 6–12 Lennard-Jones potential [47,48] is used to model the van der
aggregation. Since most force fields were developed for the globular Waals interactions between the intrachain and interchain residues,
proteins rather than the intrinsically disordered proteins (IDPs), sec­ which may be expressed as
ondary structures in the unfolded or misfolded states of IDPs is exces­
sively stabilized. Therefore, a selected data set of IDPs compiled from the

2
P. Dey and P. Biswas Biophysical Chemistry 297 (2023) 107011

[( ) ( )6 ]
12
rmin rmin
EVDW =ε − 2 (1)
rij rij

where EVDW denotes the van der Waals interaction potential between the
i-th and j-th residues separated by a spatial distance of rij and ε is the
depth of the energy well, which is assumed to be 1 for simplicity. The
distance between two particles at which the potential is minimum is
given by rmin.

2.3. Hydrogen bonding

A 10–12 Lennard-Jones potential is employed for modeling the


hydrogen bond potential as proposed by Scheraga and co-workers,
which may be given by [44].
[ ( )12 ( )10 ]
Rmin Rmin
EHB = ε 5 − 6 (2)
rij rij

where Rmin is the spatial distance between the residues for which the
hydrogen bond potential is minimum and ε is the depth of the potential
well, which is assumed to be unity. This coarse-grained model accounts
only for the Cα backbone-backbone hydrogen bond where the explicit
nature of the side chains is neglected. The formation of a hydrogen bond
between a pair of residues must satisfy a bond length and a bond angle
Fig. 2. Schematic representation of two-body intrachain and inter­
constraint. The distance between the backbone N atom of a residue and
chain potential.
the backbone O atom of another residue should be ≤3.5 Å and the CON
angle should be >90◦ . The value of Rmin is obtained by evaluating the Cα
− Cα backbone distances between the residues that form the hydrogen N(i, j, d)exp = N(d)χ i χ j (4)
bond, obtained from the selected data-set of 1135 intrinsically disor­
dered proteins. The minimum distance, Rmin, is defined to be the value where N(d) is the total number of interacting residue pairs and χ k is the
that corresponds to the peak at 5.25 Å in the plot of the distribution of mole fraction of k-th type of residue derived from the total mole frac­
residue-residue distances forming the hydrogen bond (refer to Fig. S4 of tions of the distinct residue types in the overall data-set. Two-body po­
Supplementary Material). Thus, hydrogen bond is formed between two tential is estimated for all residue pairs using the selected data-set of
residues if the two interacting resides are at a distance of 5.0–7.0 Å in the 1135 IDPs as shown in Fig. 3. Most of the residue pairs have a lower
Cα chain backbone model. probability of non-bonded contact in the first and second distant shells
resulting in lower potential values. However, two-body contact in­
2.4. Two-body potential teractions between any two residues commence at larger distance shells
4, 5, and 6. For larger distance shells, there is a higher probability of
Two-body interaction potential determines the non-bonded in­ contact between the residue pairs and most residue pairs exhibit an
teractions between any two residues along the sequence of the protein. attractive interaction. The hydrophobic residue pairs in contact pri­
The effective two-body potentials between residues are obtained from a marily experience attractive interactions, whereas the hydrophilic con­
distribution of the pairwise distances between them using the data set of tact residue pairs exhibit repulsive interactions. The characteristics of
1135 intrinsically disordered proteins [43]. the two-body potential are presented in Figs. 4 and 5 for different pairs
Two-body interaction potential for the Cα chain backbone model is of residues. Fig. 4 illustrates the striking contrast of the two-body po­
derived from the selected dataset of proteins, where the Cα atom is tential between the pairs of Leu-Ile and Leu-Asp residue pairs. A distinct
residue-specific indicating the Cα of one residue is different from that of potential well is observed for the pair Leu-Ile, which indicates the
the other. Hence, there are 20 different atom types forming 210 distinct attractive interaction present in the hydrophobic residue pair revealing
atom pairs for the two-body potential calculations. The separation be­ its hydrophobic characteristics, however no such pattern is observed for
tween a residue pair is calculated using distance shells which starts at the Leu-Asp residue pair.
1.5 Å from any residue in the pair, with each shell at a radial distance of The two-body potential of the charged residues is also plotted versus
1 Å in width as shown in Fig. 2. For instance, the third distance shell the number of distance shells in Fig. 5. The attractive interactions be­
covers the distance from 3.5 to 4.5 Å, while the thirtieth distance shell tween the oppositely charged residue pair Asp-Lys may be observed,
maps the distance from 30.5 Å to infinity. This two-body potential in­ while a repulsive interaction between the similarly charged residue pairs
corporates both electrostatic and hydrophobic interactions. The poten­ Asp-Asp and Lys-Lys is noted, which highlights the role of electrostatic
tial for any residue pair, i and j, within a distance shell d, is defined using interactions.
Boltzmann inversion method [49] as Thus, the total energy of the system is given by
[ ] E = EVDW + EHB + E2B (5)
N(i, j, d)obs
E2B (i, j, d) = − ln (3)
N(i, j, d)exp The Cα chain backbone model of the dimer, trimer and tetramer of
amyloid-β is selected and the chains are positioned randomly at a dis­
where N(i, j, d)obs and N(i, j, d)exp are the number of interacting pairs of tance of 35 Å. The initial energy is calculated using Eq. (5). At each MC
residues observed and the maximum number of such interacting residue step, a random residue is selected and moved to a new position. The
pairs expected in the distance shell d, respectively. N(i, j, d)obs is evalu­ energy of this newly generated conformation is calculated after applying
ated by computing the number of residue pairs i and j in a particular appropriate restrictions in the bond length, bend angle and dihedral
distance shell d, while N(i, j, d)exp is calculated from angle to obtain real protein conformations. The concentration is

3
P. Dey and P. Biswas Biophysical Chemistry 297 (2023) 107011

Fig. 3. The representation of residue-based pairwise two-body potential for the distance shells 1, 2, 3, 4, 5, 6, 8, 10, 15 and 30.

maintained using a PBC (periodic boundary condition) [50]. The newly


generated conformation is selected according to the Metropolis criterion
[51]. Any residue is moved using the Boltzmann probability, PT = exp
(-Eij/kBT), where Eij = Ej − Ei is the difference in energy between its new
(Ej) and old (Ei) conformations, T is the absolute temperature and kB is
Boltzmann constant. The Monte Carlo move is accepted if the initial
energy is more than that of the newly generated conformation, other­
wise the move is accepted conditionally. Then, a random number
generator subroutine ran1(idum) [52] is employed to produce a pseudo
random number which is compared with the the transition probability,
PT. The Monte Carlo move is accepted for PT > pseudo-random number,
otherwise the move is rejected. A unit Monte Carlo (MC) step is an
attempt to move a residue once.

3. Results and discussion


Fig. 4. The two-body interaction potentials between hydrophobic residue pair
Leu-Ile and between hydrophobic and hydrophilic residue pair Leu-Asp. The total energy of the dimer, trimer and tetramer is calculated from
Eq. (5) and plotted as a function of the number of MC steps as shown in
Fig. S6 of the Supplementary Material. The energy of the dimer, trimer
and tetramer decreases with an increase in the number of MC moves and
attains equilibrium due to the formation of a stable aggregate. The
components of the total energy is calculated using Eqs. (1), (2) and (3)
and are presented in Figs. 6 (a), Figs. 6 (b) and Figs. 6 (c) for Lennard-
Jones potential, Two-body potential and Hydrogen bond potential of
the dimer, trimer and tetramer of amyloid-β, respectively.
The two-body distance-dependent interaction potential captures the
features of both hydrophobic and electrostatic interactions. A higher
value of the two-body potential leads to increased hydrophobic in­
teractions due to the presence of more than 45% hydrophobic residues
in amyloid-β. The strength of hydrophobic interactions dominates that
of the hydrogen bond potential, which results in the formation of
amorphous rather than fibrillar aggregates as observed in earlier studies
[14,53] since hydrophobic interactions lead to the formation of amor­
Fig. 5. The two-body interaction potentials between oppositely charged res­ phous aggregates [54]. The generation of fibrils involves the trans­
idue pair Asp-Lys and between similarly charged residues pairs Lys-Lys and formation of small amorphous aggregates →β-sheets → ordered-nucleus
Asp-Asp. Each distance shell is 1 Å wide. which leads to a formation of a stable fibril nucleus [14]. Consequently,
the two-body potential is initially attractive for the inter and intrachain
interactions which helps in bringing the chains to a contact distance,

4
P. Dey and P. Biswas Biophysical Chemistry 297 (2023) 107011

Fig. 6. Variation of different components of total energy, i.e. Lennard-Jones potential, Two-body potential and Hydrogen bond potential for dimer, trimer and
tetramer of amyloid-β with MC moves.

Fig. 7. The intrachain and interchain contacts in the dimer, trimer and tetramer of amyloid-β 42.

5
P. Dey and P. Biswas Biophysical Chemistry 297 (2023) 107011

while the repulsive van der Waals interaction prevents the chains from amyloid-β reproduce the data reported earlier [57].
collapsing. The interchain distances in terms of the centre of mass distance, are
The intrachain and interchain contact map provides a statistical calculated. Fig. 8 shows a decrease in the distance between the two
description of the three-dimensional organization of dimer, trimer and chains with an increase in the MC simulation steps which indicates
tetramer structures. Two residues are considered to be in contact when increasing interchain interactions with the subsequent formation of the
the spatial distance of separation between them is ≤7.5 Å [55]. In Fig. S7 aggregate.
of the Supplementary Material, the probability contact map for the The root mean-square fluctuation (RMSF) of the Cα atoms is
trimer is shown. The intrachain contacts are observed in the initial computed in order to quantify the magnitude of deviation from their
conformations of the protein due to the contacts between the neigh­ respective average positions for the specific sites along the backbone of
boring residues, however both the intrachain and the interchain contacts the dimer, trimer, and tetramer of amyloid-β.
are found in the later conformations as a function of MC steps (for the ith It is defined as
residue, (i + 1)th and (i − 1)th residue are the neighboring residues). The √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
( )2̅

three different chains of the trimer are represented as ‘chain A', ‘chain B' √1 ∑ N
1 ∑ N

and ‘chain C'. It is observed that interchain contacts (i.e. between chain RMSF = √ rj,i − rk,i (8)
N j=1 N k=1
A and chain B or chain B and chain C) are also formed along with the
intrachain contacts. However, the interchain interactions lead to the Larger RMSF value indicates increased flexibility, whereas low RMSF
aggregate formation. values show restricted movements during simulation about its average
Fig. 7 reflects a marked increase in the number of numerous inter­ position.
molecular hydrophobic contacts and intramolecular contacts for some In the dimer, trimer and tetramer, individual residues of various
residues as a function of the MC steps. The formation of the aggregates as chains fluctuate in distinct ways. This finding demonstrates substantial
dimer, trimer or tetramer of amyloid-β is primarily driven by the energy fluctuations of the N- or C-terminal residues that indicate local unfold­
gain associated with an increase in the number of intermolecular hy­ ing. Analysis of the flexibility shows that the residue fluctuations in­
drophobic contacts. This is in agreement with a previous simulation crease with an increase in the number of chains as shown in Fig. 9. The
study [56]. RMSF is also plotted based on the amino acid type present in the dimer,
The change in the structure of a protein during simulations can be trimer and tetramer as shown in Fig. S9 of the Supplementary Material.
quantified through the radius of gyration, Rg, which is defined as: It is observed that the charged residues fluctuate more than the hydro­
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ phobic residues in the dimer and the cumulative effect of the negatively
√ N
√∑ (r(i) − rc )2 and positively charged residues are more pronounced than the hydro­
Rg = √ (6)
N phobic residues in the dimer, trimer and tetramer of the amyloid-β
n=1
protein. This shows that the charged residues play a significant role in
where N is the number of protein atoms and r(i) and rc are the co­ the formation of aggregates as shown in Fig. 10. This is in agreement
ordinates of an ith residue and the center of mass, respectively. The with an earlier work [58] suggesting that the behaviour of positively and
variation of the radius of gyration with simulation steps is shown in negatively charged residues has a substantial impact on the hydropho­
Fig. S8 of the Supplementary Material. The sharp decrease in Rg due to bicity of the protein that determines its aggregation propensity [59].
the formation of aggregates of protein chains provides an evidence that In Fig. 11, the local and nonlocal contacts are calculated for the
the extended structure gradually collapsed. dimer, trimer and tetramer for the Cα chain backbone coarse grained
The root-mean-square deviation (RMSD) for the dimer, trimer and model of amyloid-β. Local contact measures the contact between the
the tetramer of amyloid-β is monitored to measure of convergence of the nonbonded residues up to the (i + 4)th position along the sequence. The
MC simulation runs. RMSD is defined as: number of contacts between non-bonded residues within a threshold
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ distance of 7.5 Å separated by more than (i + 4) residues in the sequence
∑N represents nonlocal contact [38]. It is observed that the aggregation of
i=1 (ri (t) − ri (0) )
RMSD(t) = (7) the dimer, trimer, and tetramer of amyloid-β is initially regulated by the
N
local contacts. However, With the progress of simulation, the nonlocal
where N is the number of protein atoms and ri(t) and ri(0) denote the interactions play a more significant role resulting in increasing contacts
position coordinates of the ith atom of the initial structure and the between the residues of the dimer, trimer and tetramer of the amyloid-β.
structure after the tth MC step respectively. It may be due to this reason that the aggregation takes place when some
The evolution of RMSDs with the Monte Carlo moves of the dimer, of the most strongly interacting amino acids form local contacts which
trimer and tetramer of amyloid-β is shown in Fig. S8 of the Supple­ lead to the formation of a specific subset of the native structure [60].
mentary Material. The RMSD become stable after 5 × 107 MC steps for These structures provide a seed for the aggregation process and may be
the dimer, trimer and tetramer of amyloid-β indicating that the dimer or considered as partially folded intermediates that are involved in the
oligomers formed are stabilized. The RMSD values for the dimer of aggregation of a number of proteins [61,62]. Thus, formation of

Fig. 8. Variation in the distance between the chains of dimer, trimer and tetramer with MC moves.

6
P. Dey and P. Biswas Biophysical Chemistry 297 (2023) 107011

Fig. 9. Root-mean square fluctuation, RMSF of the dimer, trimer and tetramer of amyloid-β as a function of MC moves at different time scales.

Fig. 10. Normalised Root-mean square fluctuation, RMSF, for the hydrophobic and charged residues present in the dimer, trimer and tetramer of amyloid-β.

aggregates primarily occur due to the nonlocal interactions. force. The potential of mean force (PMF) at any coordinate can be
The solvent-accessible surface area (SASA) is calculated using the calculated by averaging over the forces for specific position coordinates
PCASA software [63] which is a rapid SASA estimator for coarse-grained (such as center to center interchain distances) for the dimer, trimer and
models and is shown in Fig. S10 of the Supplementary Material for the tetramer of amyloid-β. The potential of mean force may be expressed as
trimer of amyloid-β. A higher value of the SASA is correlated with the [64,65]:
extended chains in the trimer of amyloid-β. A decrese in the SASA is
U(r) = − kB Tln(g(r) ) (9)
observed for the aggregates relative to that of the monomeric form
which may be due to the formation of a compact structure of the trimer where g(r) denotes the radial distribution function from the centre of
of amyloid-β. mass of one chain to that of the other for the dimer, trimer and tetramer
The free energy profile was calculated from the potential of mean of amyloid-β. A dimer or oligomer of amyloid-beta is formed when the

7
P. Dey and P. Biswas Biophysical Chemistry 297 (2023) 107011

Fig. 11. Variation in local and nonlocal contacts of the (a)dimer, (b) trimer and (c) tetramer with the MC moves.

Fig. 12. The plot of the potential of mean force for the dimer, trimer and tetramer of amyloid-β.

CoM of the chains are separated by a distance of 4.5–5.5 Å as shown in 4. Conclusions


Fig. 12. The aggregate of amyloid-β appears as a narrow well located on
the the left-hand side of the PMF profile. The energy landscape is typi­ A coarse-grained Monte Carlo approach is used to investigate the
cally rugged, which represents an ensemble of closely spaced confor­ mechanism which leads to the protein aggregation. The development of
mations that may belong to the misfolded states [66,67]. These an accurate force field is one of the most important steps in under­
misfolded proteins through intermolecular interactions form amorphous standing protein aggregation. This potential is modeled in terms of the
aggregates, which are defined as a supramolecular assembly of poly­ bonded and non-bonded interactions using a selected data-set of
peptide chains without a defined shape [54,68]. Thus, the PMF profile intrinsically disordered proteins, since protein force fields developed for
exhibits a narrow basin corresponding to the formation of stable ag­ folded proteins produce highly compact structures for IDPs. The bonded
gregates. The plots of the energy, Rg, RMSD, RMSF and the free energy potential is modeled in terms of the bond, bend and dihedral restrictions,
profiles are shown for the pentamer, hexamer, heptamer and octamer of while the non-bonded interactions is comprise the van der Waals,
amyloid-β respectively in Figs. S11 - S15 of the Supplementary Material. hydrogen bond and two-body potential. The two-body potential be­
tween the atoms of the hydrophobic residues, Leu and Ile indicates
attractive interactions between them and like-charged hydrophilic

8
P. Dey and P. Biswas Biophysical Chemistry 297 (2023) 107011

residues show a repulsive two body interaction, while oppositely [9] D. Chakraborty, J.E. Straub, D. Thirumalai, Differences in the free energies
between the excited states of a β 40 and a β 42 monomers encode their aggregation
charged hydrophilic residues show an attractive two body interaction.
propensities, Proc. Natl. Acad. Sci. 117 (33) (2020) 19926–19937.
The two-body potential comprises an attractive interchain and intra­ [10] S. Vivekanandan, J.R. Brender, S.Y. Lee, A. Ramamoorthy, A partially folded
chain interaction and helps to bring the chains at the contact distance, structure of amyloid-beta (1–40) in an aqueous environment, Biochem. Biophys.
while the repulsive van der Waals interaction keeps the chains from Res. Commun. 411 (2) (2011) 312–316.
[11] J.C. Phillips, Why aβ42 is much more toxic than aβ40, ACS Chem. Neurosci. 10 (6)
collapsing. The total energy of the dimer, trimer and tetramer of amy­ (2019) 2843–2847.
loid-β decreases with an increase in the Monte Carlo moves confirming [12] G.A. Pantelopulos, J.E. Straub, D. Thirumalai, Y. Sugita, Structure of APP-C991–99
the formation of the amorphous oligomer. The structural properties of and implications for role of extra-membrane domains in function and
oligomerization, Biochim. Biophys. Acta Biomembr. 1860 (2018) 1698–1708.
the aggregate are characterized through the root-mean-square deviation [13] J.H. Meinke, U.H. Hansmann, Aggregation of β-amyloid fragments, J. Chem. Phys.
(RMSD), root-mean-square fluctuation (RMSF), the radius of gyration 126 (1) (2007), 01B603.
(Rg) and the PMF profile, while the role of interactions may be eluci­ [14] H.D. Nguyen, C.K. Hall, Spontaneous fibril formation by polyalanines;
discontinuous molecular dynamics simulations, J. Am. Chem. Soc. 128 (6) (2006)
dated by calculating the contacts formed by intra- and interchain resi­ 1890–1901.
dues respectively. The hydrophobic residues help in nucleation, while [15] W.M. Berhanu, U.H. Hansmann, Side-chain hydrophobicity and the stability of
the charged residues promote oligomerization and aggregation. The aβ16–22 aggregates, Protein Sci. 21 (12) (2012) 1837–1848.
[16] A.P. Heath, L.E. Kavraki, C. Clementi, From coarse-grain to all-atom: toward
transition from an extended structure to a compact form is exhibited by a multiscale analysis of protein landscapes, Proteins Struct. Funct. Bioinf. 68 (3)
decrease in the solvent-accessible surface area and the PMF profile ex­ (2007) 646–661.
hibits a narrow basin corresponding to the formation of stable aggre­ [17] A. Srivastava, P.V. Balaji, Molecular events during the early stages of aggregation
of gnnqqny: an all atom md simulation study of randomly dispersed peptides,
gates. This method may be modified to enhance our understanding
J. Struct. Biol. 192 (3) (2015) 376–391.
regarding the secondary structural changes that occurs during the ag­ [18] I.M. Ilie, A. Caflisch, Simulation studies of amyloidogenic polypeptides and their
gregation of intrinsically disordered proteins. aggregates, Chem. Rev. 119 (12) (2019) 6956–6993.
[19] Y. Tachi, Y. Okamoto, H. Okumura, Conformational change of amyloid-β 40 in
association with binding to gm1-glycan cluster, Sci. Rep. 9 (1) (2019) 1–11.
CRediT authorship contribution statement [20] S.G. Itoh, M. Yagi-Utsumi, K. Kato, H. Okumura, Effects of a hydrophilic/
hydrophobic interface on amyloid-β peptides studied by molecular dynamics
simulations and NMR experiments, J. Phys. Chem. B 123 (1) (2018) 160–169.
Priya Dey: Data curation, Formal analysis, Validation, Writing - [21] L.L.N. Ngoc, S.G. Itoh, P. Sompornpisut, H. Okumura, Replica-permutation
original draft. Parbati Biswas: Conceptualization, Methodology, Su­ molecular dynamics simulations of an amyloid-β (16–22) peptide and polyphenols,
pervision, Validation, Writing - original draft, Writing - review & Chem. Phys. Lett. 758 (2020), 137913.
[22] V.H. Man, P.H. Nguyen, P. Derreumaux, High-resolution structures of the amyloid-
editing. β 1–42 dimers from the comparison of four atomistic force fields, J. Phys. Chem. B
121 (24) (2017) 5977–5987.
[23] B. Barz, Q. Liao, B. Strodel, Pathways of amyloid-β aggregation depend on oligomer
Declaration of Competing Interest shape, J. Am. Chem. Soc. 140 (1) (2018) 319–327.
[24] H. Okumura, S.G. Itoh, Molecular dynamics simulations of amyloid-β (16–22)
peptide aggregation at air–water interfaces, J. Chem. Phys. 152 (9) (2020),
The authors declare no competiting financial interest. 095101.
[25] N. Schwierz, C.V. Frost, P.L. Geissler, M. Zacharias, Dynamics of seeded aβ40-fibril
growth from atomistic molecular dynamics simulations: kinetic trapping and
Acknowledgments reduced water mobility in the locking step, J. Am. Chem. Soc. 138 (2) (2016)
527–539.
The authors gratefully acknowledge DST-SERB, India (project no. [26] S. Sasmal, N. Schwierz, T. Head-Gordon, Mechanism of nucleation and growth of
aβ40 fibrils from all-atom and coarse-grained simulations, J. Phys. Chem. B 120
CRG/2022/000889) and the FRP Grant-IoE (Ref. No. IoE/2021/12/ (47) (2016) 12088–12097.
FRP) for financial support. Priya Dey acknowledges CSIR, India, for [27] D.S. Davidson, A.M. Brown, J.A. Lemkul, Insights into stabilizing forces in amyloid
providing financial support in the form of a Senior Research Fellowship fibrils of differing sizes from polarizable molecular dynamics simulations, J. Mol.
Biol. 430 (20) (2018) 3819–3834.
(09/045(1536)/2017-EMR-I). [28] P.N. Nirmalraj, J. List, S. Battacharya, G. Howe, L. Xu, D. Thompson, M. Mayer,
Complete aggregation pathway of amyloid β (1-40) and (1-42) resolved on an
atomically clean interface, Sci. Adv. 6 (15) (2020) eaaz6014.
Appendix A. Supplementary data
[29] I. Jahan, S.M. Nayeem, Destabilization of Alzheimer’s aβ 42 protofibrils with
acyclovir, carmustine, curcumin, and tetracycline: insights from molecular
Supplementary data to this article can be found online at https://doi. dynamics simulations, New J. Chem. 45 (45) (2021) 21031–21048.
org/10.1016/j.bpc.2023.107011. [30] Y. Tachi, S.G. Itoh, H. Okumura, Molecular dynamics simulations of amyloid-β
peptides in heterogeneous environments, Biophys. Physicobiol. 19 (2022),
e190010.
References [31] H. Okumura, S.G. Itoh, Molecular dynamics simulation studies on the aggregation
of amyloid-β peptides and their disaggregation by ultrasonic wave and infrared
laser irradiation, Molecules 27 (8) (2022) 2483.
[1] A.L. Goldberg, Protein degradation and protection against misfolded or damaged
[32] S. Abeln, M. Vendruscolo, C.M. Dobson, D. Frenkel, A simple lattice model that
proteins, Nature 426 (6968) (2003) 895–899.
captures protein folding, aggregation and amyloid formation, PLoS One 9 (1)
[2] A. Baruah, A. Bhattacherjee, P. Biswas, Role of conformational heterogeneity on
(2014), e85185.
protein misfolding, Soft Matter 8 (16) (2012) 4432–4440.
[33] S. Mohanty, Aggregation and coacervation with Monte Carlo simulations, Prog.
[3] L.C. Serpell, Alzheimer’s amyloid fibrils: structure and assembly, Biochim. Biophys.
Mol. Biol. Transl. Sci. 170 (2020) 505–520.
Acta Mol. basis Dis. 1502 (1) (2000) 16–30.
[34] C. Haass, D.J. Selkoe, Soluble protein oligomers in neurodegeneration: lessons
[4] R. Van Der Lee, M. Buljan, B. Lang, R.J. Weatheritt, G.W. Daughdrill, A.K. Dunker,
from the Alzheimer’s amyloid β-peptide, Nat. Rev. Mol. Cell Biol. 8 (2) (2007)
M. Fuxreiter, J. Gough, J. Gsponer, D.T. Jones, et al., Classification of intrinsically
101–112.
disordered regions and proteins, Chem. Rev. 114 (13) (2014) 6589–6631.
[35] G.M. Shankar, S. Li, T.H. Mehta, A. Garcia-Munoz, N.E. Shepardson, I. Smith, F.
[5] P.H. Nguyen, A. Ramamoorthy, B.R. Sahoo, J. Zheng, P. Faller, J.E. Straub,
M. Brett, M.A. Farrell, M.J. Rowan, C.A. Lemere, et al., Amyloid-β protein dimers
L. Dominguez, J.-E. Shea, N.V. Dokholyan, A. De Simone, et al., Amyloid
isolated directly from Alzheimer’s brains impair synaptic plasticity and memory,
oligomers: a joint experimental/computational perspective on Alzheimer’s disease,
Nat. Med. 14 (8) (2008) 837–842.
parkinson’s disease, type ii diabetes, and amyotrophic lateral sclerosis, Chem. Rev.
[36] E.E. Cawood, T.K. Karamanos, A.J. Wilson, S.E. Radford, Visualizing and trapping
121 (4) (2021) 2545–2647.
transient oligomers in amyloid assembly pathways, Biophys. Chem. 268 (2021),
[6] S. Baglioni, F. Casamenti, M. Bucciantini, L.M. Luheshi, N. Taddei, F. Chiti, C.
106505.
M. Dobson, M. Stefani, Prefibrillar amyloid aggregates could be generic toxins in
[37] N. Metropolis, A.W. Rosenbluth, M.N. Rosenbluth, A.H. Teller, E. Teller, Equation
higher organisms, J. Neurosci. 26 (31) (2006) 8160–8167.
of state calculations by fast computing machines, J. Chem. Phys. 21 (6) (1953)
[7] D.M. Fowler, A.V. Koulov, W.E. Balch, J.W. Kelly, Functional amyloid–from
1087–1092.
bacteria to humans, Trends Biochem. Sci. 32 (5) (2007) 217–224.
[38] A. Kumar, A. Baruah, P. Biswas, Role of local and nonlocal interactions in folding
[8] J.D. Harper, P.T. Lansbury Jr., Models of amyloid seeding in Alzheimer’s disease
and misfolding of globular proteins, J. Chem. Phys. 146 (6) (2017), 065102.
and scrapie: mechanistic truths and physiological consequences of the time-
[39] G.J. Morgan, Transient disorder along pathways to amyloid, Biophys. Chem. 281
dependent solubility of amyloid proteins, Annu. Rev. Biochem. 66 (1) (1997)
(2022), 106711.
385–407.

9
P. Dey and P. Biswas Biophysical Chemistry 297 (2023) 107011

[40] L. Martinez, R. Andrade, E.G. Birgin, J.M. Martnez, Packmol: a package for [56] B.R. Sahoo, S.J. Cox, A. Ramamoorthy, High-resolution probing of early events in
building initial configurations for molecular dynamics simulations, J. Comput. amyloid-β aggregation related to Alzheimer’s disease, Chem. Commun. 56 (34)
Chem. 30 (13) (2009) 2157–2164. (2020) 4627–4639.
[41] A. Irback, S. Mohanty, Profasi: a Monte Carlo simulation package for protein [57] B. Barz, B. Urbanc, Dimer formation enhances structural differences between
folding and aggregation, J. Comput. Chem. 27 (13) (2006) 1548–1555. amyloid β-protein (1–40) and (1–42): an explicit-solvent molecular dynamics
[42] A. Kumar, P. Biswas, Effect of site-directed point mutations on protein misfolding: study, PLoS One 7 (4) (2012), e34345.
a simulation study, Proteins Struct. Funct. Bioinf. 87 (9) (2019) 760–773. [58] S.-H. Chong, S. Ham, Interaction with the surrounding water plays a key role in
[43] C.J. Oldfield, A.K. Dunker, Intrinsically disordered proteins and intrinsically determining the aggregation propensity of proteins, Angew. Chem. Int. Ed. 53 (15)
disordered protein regions, Annu. Rev. Biochem. 83 (2014) 553–584. (2014) 3961–3964.
[44] R. McGuire, F. Momany, H. Scheraga, Energy parameters in polypeptides. v. [59] G. Reddy, J.E. Straub, D. Thirumalai, Influence of preformed Asp23-Lys28 salt
Empirical hydrogen bond potential function based on molecular orbital bridge on the conformational fluctuations of monomers and dimers of aβ peptides
calculations, J. Phys. Chem. 76 (3) (1972) 375–393. with implications for rates of fibril formation, J. Phys. Chem. B 113 (4) (2009)
[45] S.L. Mayo, B.D. Olafson, W.A. Goddard, Dreiding: a generic force field for 1162–1172.
molecular simulations, J. Phys. Chem. 94 (26) (1990) 8897–8909. [60] E. Shakhnovich, V. Abkevich, O. Ptitsyn, Conserved residues and the mechanism of
[46] H. Lu, L. Lu, J. Skolnick, Development of unified statistical potentials describing protein folding, Nature 379 (6560) (1996) 96–98.
protein-protein interactions, Biophys. J. 84 (3) (2003) 1895–1901. [61] M.A. Speed, T. Morshead, D.I. Wang, J. King, Conformation of p22 tailspike folding
[47] J.E. Jones, On the determination of molecular fields.-ii. From the equation of state and aggregation intermediates probed by monoclonal antibodies, Protein Sci. 6 (1)
of a gas, Proc. R. Soc. Lond. Ser. A 106 (738) (1924) 463–477. (1997) 99–108.
[48] X. Wang, S. Ramirez-Hinestrosa, J. Dobnikar, D. Frenkel, The lennard-jones [62] M.R. Hurle, L.R. Helms, L. Li, W. Chan, R. Wetzel, A role for destabilizing amino
potential: when (not) to use it, Phys. Chem. Chem. Phys. 22 (19) (2020) acid replacements in light-chain amyloidosis, Proc. Natl. Acad. Sci. 91 (12) (1994)
10624–10633. 5446–5450.
[49] D. Reith, M. Putz, F. Muller-Plathe, Deriving effective mesoscale potentials from [63] S. Wei, C.L. Brooks III, A.T. Frank, A rapid solvent accessible surface area estimator
atomistic simulations, J. Comput. Chem. 24 (13) (2003) 1624–1636. for coarse grained molecular simulations, J. Comput. Chem. 38 (15) (2017)
[50] M.P. Allen, D.J. Tildesley, Computer Simulation of Liquids, Oxford University 1270–1274.
Press, 2017. [64] N. Heilmann, M. Wolf, M. Kozlowska, E. Sedghamiz, J. Setzler, M. Brieg,
[51] R. Pandey, B. Farmer, Aggregation and network formation in self-assembly of W. Wenzel, Sampling of the conformational landscape of small proteins with Monte
protein (h3. 1) by a coarse-grained Monte Carlo simulation, J. Chem. Phys. 141 Carlo methods, Sci. Rep. 10 (1) (2020) 1–13.
(17) (2014), 175103. [65] V. Singh, P. Biswas, Conformational transitions of amyloid-β: a Langevin and
[52] W.H. Press, S.A. Teukolsky, B.P. Flannery, W.T. Vetterling, Numerical Recipes in generalized Langevin dynamics simulation study, ACS Omega 6 (5) (2021)
Fortran: The Art of Scientific Computing, Cambridge University Press, Cambridge, 13611–13619.
1992. [66] P. Biswas, Theoretical and computational advances in protein misfolding, in:
[53] B. Urbanc, L. Cruz, S. Yun, S.V. Buldyrev, G. Bitan, D.B. Teplow, H.E. Stanley, In Advances in Protein Chemistry and Structural Biology, Elsevier, 2019.
silico study of amyloid β-protein folding and oligomerization, Proc. Natl. Acad. Sci. [67] E.P. O’Brien, Y. Okamoto, J.E. Straub, B.R. Brooks, D. Thirumalai, Thermodynamic
101 (50) (2004) 17345–17350. perspective on the dock- lock growth mechanism of amyloid fibrils, J. Phys. Chem.
[54] J. Adamcik, R. Mezzenga, Amyloid polymorphism in the protein folding and B 113 (43) (2009) 14421–14430.
aggregation energy landscape, Angew. Chem. Int. Ed. 57 (28) (2018) 8370–8382. [68] J.R. Brender, A. Ghosh, S.A. Kotler, J. Krishnamoorthy, S. Bera, V. Morris, T.B. Sil,
[55] B. Adhikari, D. Bhattacharya, R. Cao, J. Cheng, Confold: residue-residue contact- K. Garai, B. Reif, A. Bhunia, et al., Probing transient non-native states in amyloid
guided ab initio protein folding, Proteins Struct. Funct. Bioinf. 83 (8) (2015) beta fiber elongation by NMR, Chem. Commun. 55 (31) (2019) 4483–4486.
1436–1449.

10

You might also like