ABSTRACT

WANG, YIMING. Understanding In Vitro Amyloid Formation and Inhibition Using Coarse- Grained Simulations. (Under the direction of Dr. Carol K. Hall).

The misfolding and aggregation of specific native proteins in mammalian brains, and the subsequent formation of amyloid plaques plays a central role in a number of fatal neurodegenerative diseases, including the Alzheimer’s, Parkinson’s, Huntington’s and Prion

Disease. Protein aggregation has thus become an active and multidisciplinary research subject.

The aim of this dissertation is to apply molecular modeling techniques in combination with biochemical and biophysical measurements to advance understanding of in vitro protein aggregation.

Firstly, we investigate the aggregation kinetics and amyloid core structures of three prion protein fragments corresponding to residues 120 to 144 of human (Hu), bank vole (BV) and Syrian hamster (SHa) prion protein using discontinuous molecular dynamics simulations (DMD) and the

PRIME20 force field. We find that HuPrP120-144 has a shorter aggregation lag phase than

BVPrP120-144, followed by SHaPrP120-144, consistent with experimental findings. We also investigate the molecular mechanism of seed-templating aggregation of three prion protein peptides. We find that the pre-formed protofilament (seed) accelerates the aggregation of the monomer peptides by eliminating the lag phase. The monomer aggregation kinetics are mainly determined by the structure of the seed. For cross-seeding aggregation, we show that Hu has a strong tendency to adopt the conformation of the BV seed and vice versa; the Hu and BV monomers have a weak tendency to adopt the conformation of the SHa seed.

Next, we develop a coarse-grained inhibitor model that is compatible with the PRIME20 protein model to examine the effect of phenolic molecules on the aggregation of Aβ(17-36) peptides. From simulating the dynamics of the interactions between four inhibitor molecules and Aβ(17-36) monomer, oligomer and protofilament, we find that the rank order of effectiveness in inhibiting Aβ(17-36) aggregation is : EGCG > resveratrol > curcumin > vanillin, consistent with experimental findings on their ability to inhibit full length Aβ fibrillation. We also learn that the inhibition effect of EGCG is specific to the peptide sequence while those of resveratrol and curcumin are non-specific in that they stem from strong interference with hydrophobic sidechain association, regardless of the residues’ location and peptide sequence.

We probe the aggregation properties of three peptides (P1, P2 and P3 corresponding to residues 326-337, 387-405 and 426-442, respectively) derived from the myocilin olfactomedin domain which is associated with inherited open angle glaucoma. DMD/PRIME20 simulations predict that only P1 and P3 are aggregation-prone; P1 consistently forms fibrillar aggregates with parallel in-register β-sheets whereas P3 forms heterogeneous β-sheet-rich aggregates. The natural abundance 13C solid-state NMR spectra supports our in silico observation.

We calculate an equilibrium concentration and temperature phase diagram for the amyloidogenic peptide, Aβ16-22. Our results reveal that the only thermodynamically stable phases are the solution phase and the macroscopic fibrillar phase, and that there also exists a hierarchy of metastable phases. The boundary line between the solution phase and fibril phase is found by calculating the temperature-dependent solubility of a macroscopic Aβ16-22 fibril consisting of an infinite number of β-sheet layers. Crucially, the in silico prediction of Aβ16-22 solubilities over the temperature range of 4-57°C agrees well with our fibrillation experiments.

We explore the structural mechanism of amyloid assembly by analyzing the co-aggregation of Aβ40 and Aβ16-22, two widely studied peptide fragments of Aβ42 implicated in Alzheimer’s disease. Using a combination of biochemical and biophysical analyses, we demonstrate that Aβ16-

22 increases the aggregation rate of Aβ40 through a surface catalysed secondary nucleation mechanism. The same system is also modelled using DMD/PRIME20, which provides molecular insights into the co-aggregation mechanism.

© Copyright 2018 by Yiming Wang

All Rights Reserved Understanding In Vitro Amyloid Formation and Inhibition Using Coarse-Grained Simulations.

by Yiming Wang

A dissertation submitted to the Graduate Faculty of North Carolina State University in partial fulfillment of the requirements for the degree of Doctor of Philosophy

Chemical Engineering

Raleigh, North Carolina

2018

APPROVED BY:

______Dr. Carol K. Hall Dr. Erik E. Santiso Committee Chair

______Dr. Jason M. Haugh Dr. Balaji Rao

DEDICATION

This dissertation is dedicated to my dear parents Deping Wang and Rong Yang for their love,

support, and encouragement, without which I would not be where I am today.

ii

BIOGRAPHY

Yiming Wang was born in Shanghai, China to Deping Wang and Rong Yang. He attended elementary, secondary and high school in Shanghai. He completed his Bachelor of Engineering

Degree in Chemical Engineering at East China University of Science and Technology in Shanghai,

China in June 2013. In Fall 2013, he moved to Raleigh, North Carolina and joined the Department of Chemical and Biomolecular Engineering in North Carolina State University to pursue his doctorate degree under the supervision of Dr. Carol K. Hall. He focused on simulation of protein aggregation. During his graduate study in 2016, he had a half-year visit to the Astbury Centre for

Structural Molecular Biology and the School of Chemistry at the University of Leeds, Leeds, UK under the support of a Cheney fellowship.

iii

ACKNOWLEDGMENTS

First, I express my deepest gratitude and appreciation to Dr. Carol K. Hall for guiding and training me throughout my graduate study. Specifically, I would like to thank her for inspiring and being supportive of new research ideas; for teaching me how to collaborate and communicate with scientists from different research fields; for improving my scientific writing and presentation skills; for guiding me to conduct great scientific research.

I would also like to thank my committee members, Dr. Jason M. Haugh, Dr. Balaji Rao, and Dr. Erik E. Santiso for their kind guidance during my graduate study at NC State. I would like to thank Dr. Haugh for teaching me reaction engineering and transport phenomena. I would also like to Dr. Keith E. Gubbins and Dr. Erik E. Santiso for teaching me quantum mechanics, statistical thermodynamics and modern molecular simulation techniques.

I would like to express my appreciation to Dr. Andrew J. Wilson and Dr. Sheena E. Radford of the University of Leeds for initiating collaborative work on the co-aggregation of full-length

Amyloid β and an Amyloid β fragment. I would like to thank Dr. Stefan Auer and Dr. Sergei

Krivov of the University of Leeds for initiating calculation of the peptide phase diagram and protein aggregation free energy landscape. I would also like to thank Dr. Raquel L. Lieberman and

Dr. Anant K. Paravastu for helpful discussions of our collaborative project on glaucoma-associated myocilin aggregation. Finally, I would like to thank Dr. Gregory A. Hudalla and Dr. Anant K.

Paravastu for helpful discussions on designing charge-complementary co-assembling peptides.

I am grateful to the Cheney fellowship for supporting my visit to the University of Leeds,

Leeds, UK. The knowledge and experience gained through this program has proven quite fruitful and valuable. I would also like to thank the National Science Foundation for providing financial

iv

support. Lastly, I would like to thank the National Institutes of Health for funding my work on protein aggregation.

I would like to acknowledge all of the members of the Hall research group. I want to thank

Dr. David Latshaw II for his discussions with me on the protein aggregation projects and PRIME20 force field. I would like to thank Dr. Xingqing Xiao, Dr. Qing Shao and Dr. Abhishek Singh for introducing me to atomistic simulation methods as well as for inspiring research ideas. I would also like to thank Dr. Emily Curtis, Dr. Binwu Zhao, Dr. David Rutkowski and Dr. Steven Benner for all the discussions on programming and for integrating me into the group, and Dr. Kye Won

Wang for sharing life philosophy during the coffee walks. All of these people helped me get through the challenging and unforgettable graduate life. To the newer members of our group: Ryan

Maloney, Amelia Chen and Sudeep Sarma: keep working hard and good luck with your graduate study!

Finally, I would like to thank my family and friends. Although my family members are far away from me in China and may not know much about my graduate research, they always showed strong care about my academic progress and that was all I needed. Mohammad Tuhin, Toufiq

Rahman, Chengxiang Liu and Yiliang Lin thank you for group study during our first-year core courses. Miaohua Liu and Han Jin, it was great to have you as roommates. Jianxun Cui, Chengjun

Li and Lan Luo, thank you for travelling, barbecuing, hiking, playing frisbee and organizing other recreation with me during one of the best five-year-periods of my life in beautiful North Carolina.

All of you have become great friends of mine, and I hope the friendship will go on. I would not be as happy as I am today without you.

v

TABLE OF CONTENTS

LIST OF TABLES ...... viii LIST OF FIGURES ...... ix Chapter 1 ...... 1 Motivation and Overview ...... 1 1.1 Motivation ...... 1 1.2 Overview ...... 4 1.3 Publications ...... 9 1.4 References ...... 11 Chapter 2 ...... 16 N-Terminal Prion Protein Peptides (PrP(120-144)) Form Parallel In -Register β-sheets via Multiple Nucleation-Dependent Pathways ...... 16 2.1 Introduction ...... 19 2.2 Results ...... 23 2.3 Discussion ...... 35 2.4 Methods...... 38 2.5 Acknowledgements ...... 43 2.6 References ...... 45 2.7 Supporting Information ...... 67 2.8 Erratum ...... 75 Chapter 3 ...... 78 Seeding and Cross-Seeding of N-Terminal Prion Protein Peptides PrP(120-144) ...... 78 3.1 Introduction ...... 81 3.2 Results ...... 84 3.3 Discussion and Conclusion ...... 90 3.4 Methods...... 92 3.5 Acknowledgements ...... 94 3.6 References ...... 95 3.7 Supporting Information ...... 109 Chapter 4 ...... 112 Aggregation of Aβ(17-36) in the Presence of Naturally Occurring Phenolic Inhibitors Using Coarse-Grained Simulations ...... 112 4.1 Introduction ...... 115 4.2 Results ...... 121 4.3 Discussion and Conclusion ...... 128 4.4 Methods...... 132 4.5 Acknowledgements ...... 137 4.6 References ...... 138 4.7 Supporting Information ...... 158 Chapter 5 ...... 160 Simulations and Experiments Delineate Amyloid Fibrillation by Peptides Derived from Glaucoma-Associated Myocilin ...... 160 5.1 Introduction ...... 163 5.2 Methods...... 164 5.3 Results and Discussion ...... 167

vi

5.4 Conclusion ...... 170 5.5 Acknowledgements ...... 171 5.6 References ...... 172 5.7 Supporting Information ...... 183 Chapter 6 ...... 188 Thermodynamic Phase Diagram of Amyloid-β (16-22) Peptide ...... 188 6.1 Significance Statement...... 191 6.2 Introduction ...... 191 6.3 Results and Discussions ...... 194 6.4 Conclusion ...... 203 6.5 Methods...... 203 6.5 Acknowledgements ...... 208 6.6 References ...... 210 6.7 Supplementary Information ...... 221 Chapter 7 ...... 228 Molecular Insights into the Surface Catalysed Secondary Nucleation of Aβ(1-40) by the Peptide Fragment Aβ(16-22) ...... 228 7.1 Introduction ...... 230 7.2 Results ...... 233 7.3 Discussion ...... 240 7.4 Materials and Methods ...... 242 7.5 Acknowledgements ...... 247 7.6 References ...... 248 7.7 Supplementary Materials ...... 259 Chapter 8 ...... 271 Future Works ...... 271 8.1 Parallelization of PRIME20 Code ...... 271 8.2 Construction of oligomerization free energy profile of Aβ16-22 peptide...... 272 8.3 References ...... 273

vii

LIST OF TABLES

Table 2.1 A comparison of total interaction energy (ETotal), hydrogen bonding energy (EHB), sidechain-sidechain interaction energy (ESidechain) of protofilaments formed by Hu, BV and SHaPrP120-144 ...... 63

Table 2.2 Lag phase time of Hu, BV, SHaPrP120-144 aggregation ...... 64

Table 2.3 Average percentage of parallel and anti-parallel β-sheet segments in Hu, BV and SHaPrP120-144 oligomers and protofilaments ...... 65

Table 2.4 Amino acid sequence of three PrP120-144 peptides ...... 66

Table S2.1 Square well interaction parameters between 19 side-chain centroids in reduced energy unit εHB which is set to be 12.47 kJ/mol ...... 72

Table 3.1 Number of HuPrP(120-144) peptides that adopt the peptide conformation in the seed or other conformations during the aggregation in the presence of the S-, Ω- and U-shaped Hu seeds ...... 106

Table 3.2 Number of Hu, BV and SHaPrP(120-144) peptides that adopt the peptide conformation in the seed during the seeding and cross-seeding aggregation ...... 107

Table 3.3 Amino acid sequence of three prion protein fragments ...... 108

Table 4.1 Coarse-grained group parameters for phenolic compounds ...... 157

Table 5.1 The secondary chemical shifts of the distinguishable 13C peaks in the CPMAS spectrum of P1 aggregates. For residues within β-strands, CO and Cα peak frequencies are expected to be at least 0.5ppm under corresponding random coil values (measured from random coil model peptides) 33-34 and Cβ peak frequencies are expected to be at least 0.5 ppm above corresponding random coil values ...... 182

Table S7.1 The expected and observed m/z values for monomeric and oligomeric Aβ40 in isolation and in the presence of a 1:1 ratio of Aβ16-22 ...... 266

Table S7.2 Assignments of each of the major peaks observed in Figure S6a ...... 269

viii

LIST OF FIGURES

Figure 2.1 Total interaction energy (kJ/mol) of HuPrP120-144 aggregation versus simulation time at reduced temperature T*=0.195 and peptide concentration of 20mM ...... 53

Figure 2.2 S-shaped (A), Ω-shaped (C) and U-shaped (E) protofilament formed by HuPrP120-144 from the 5th, 12th and 11th runs. B, D, F. Schematic representations of peptide conformations in the S-, Ω- and U-shaped protofilaments. The sizes of the sidechain beads don’t reflect the actual radii Hydrophobic residues (white), positively charged residues (red), negatively charged residues (blue), and polar residues (green) are shown ...... 54

Figure 2.3 The S-shaped (A) and Ω-shaped (C) protofilaments formed by BVPrP120-144 from the 5th and 4th runs. B, D. Schematic representations of peptide conformation in the S- and Ω-shaped protofilaments. The color scheme is same as that used in Fig.2.2 ...... 55

Figure 2.4 The S-shaped (A) and Ω-shaped (C) protofilaments formed by SHaPrP120-144 from the 7th and 1st runs. B, D. Schematic representations of peptide conformation in the S- and Ω-shaped protofilaments ...... 56

Figure 2.5 Secondary structure distribution of the S-shaped protofilament formed by HuPrP 120-144. The secondary structure content is calculated and averaged over the last 100 billion collisions using the STRIDE program. (27) β-sheet, random coil and turn structure are three major types. Other types of secondary structures are not shown here ...... 57

Figure 2.6 Simulation snapshots for the 3rd run of HuPrP120-144. Snapshots are taken at t = (A) 1, (B) 2900, (C) 4187, (D) 4559, (E) 6835 (ns). F shows the time evolution of system total interaction energy (kJ/mol) ...... 58

Figure 2.7 Simulation snapshots for the 5th run of HuPrP120-144. Snapshots are taken at t = (A) 1, (B) 1191, (C) 3065, (D) 6348, (E) 111393 (ns). F shows the time evolution of system total interaction energy (kJ/mol) ...... 59

Figure 2.8 Average sidechain interaction energy (kJ/mol) (A) and average number of sidechain contacts (B) experienced by residue 138, 139 and 143 during the nucleation stage ...... 60

Figure 2.9 Number of hydrogen bonds formed as a function of time during HuPrP120-144 and PrP120-137 aggregation ...... 61

Figure 2.10 Schematic representation of a parallel β-sheet segment (A) and two types of anti-parallel β-sheet segments (B, C) formed between two short peptide strands. Three residues are involved in each strand. The dash line means the N-H

ix

and C=O groups form a hydrogen bond ...... 62

Figure S2.1 Total interaction energy (kJ/mol) of BVPrP120-144 aggregation versus simulation time at reduced temperature T*=0.195 and peptide concentration of 20mM ...... 67

Figure S2.2 Total interaction energy (kJ/mol) of SHaPrP120-144 aggregation versus simulation time at reduced temperature T*=0.195 and peptide concentration of 20mM ...... 68

Figure S2.3 Final structures for 8 HuPrP120-144 peptides from 15 independent runs at T*=0.195. Structures are taken after at least 800 billion collisions (t ≈ 9μs) ...... 69

Figure S2.4 Final structures for 8 BVPrP120-144 peptides from 11 independent runs at T*=0.195. Structures are taken after at least 800 billion collisions (t ≈ 9μs) ...... 70

Figure S2.5 Final structures for 8 SHaPrP120-144 peptides from 11 independent runs at T*=0.195. Structures are taken after at least 800 billion collisions (t ≈ 9μs) ...... 71

Figure S2.6 Helical content of an alanine-rich polypeptide (Ac-AAAAKAAAAKAAAAKA -NH2) changes as a function of time. (A. DMD Simulation data and B. experimental data in literature. 1 C. Linear fitting correlation of simulation and real temperature where peptide adopt the same helical content percentage. R2 value stands for how close to the fitted regression line.) D. Plots of 5 linear correlation functions for 5 peptides. (1: Ac-AAQAAAAQAAAAQAAY- NH2; 2: Ac- AAAAKAAAAKAAAAKA-N NH2; 4: Ac-WDAAAKDAAAKDAAA KA-NH2; 5: Ac-EAEKAAKEAEKAAKEAEK-NH2) ...... 73

Figure E2.1 Plots of the MSD versus time calculated from atomistic simulation (A) and DMD simulation (B). The two points are chosen from each of two plots to calculate the slope of MSD line versus time ...... 76

Figure 3.1 S-shaped (A), Ω-shaped (C) and U-shaped (E) protofilament formed by HuPrP120-144 from the 5th, 12th and 11th runs. B, D, F. Schematic representations of peptide conformations in the S-, Ω- and U-shaped protofilaments. The sizes of the sidechain beads don’t reflect the actual radii. Hydrophobic residues (white), positively charged residues (red), negatively charged residues (blue), and polar residues (green) are shown. (This figure is reproduced from ref 32) ...... 100

Figure 3.2 (A) Plot of β-sheet percentage in HuPrP(120-144) aggregation simulations in the presence and absence of a preformed HuPrP(120-144) seed as a function of time. The percentage of β-sheet content is calculated by the STRIDE algorithm in the visual molecular dynamics (VMD) software. (B) and (C) are simulation snapshots of the non-seeding aggregation taken at t =1μs and 198μs, respectively. (D) and (E) are simulation snapshots of the seeding

x

aggregation taken at t = 1μs and 58μs, respectively ...... 101

Figure 3.3 (A) Potential energy versus time of HuPrP(120-144) peptides aggregating in the presence of U-, S-, Ω-shaped Hu seeds. (B) Average sidechain interaction energy per peptide of U-, S-, Ω-shaped HuPrP(120-144) protofilaments (seeds). (C) Potential energy versus time of Hu, BV and SHaPrP(120-144) peptides aggregating in the presence of S-shaped Hu, Ω-shaped BV and Ω-shaped SHa seeds. (D) Average sidechain interaction energy per peptide of S-shaped Hu, Ω-shaped BV and Ω-shaped SHa seeds ...... 102

Figure 3.4 (A) Plot of potential energy vs simulation time of HuPrP(120-144) peptides aggregation in the presence of S-shaped Hu, Ω-shaped BV and Ω-shaped SHa seed. (B-D) Plots of Plots of potential energy vs simulation time of Hu, BV, SHaPrP(120-144) peptides aggregating in the presence of S-shaped Hu, Ω-shaped BV and Ω- shaped SHa seeds ...... 103

Figure 3.5 Plots of the average side chain interaction energy (kJ/mol) experienced by residues 138, 139 and 143 of monomer Hu, BV and SHaPrP(120-144) peptides in the presence of (A) S-shaped Hu, (B) Ω-shaped BV and (C) Ω-shaped SHa seeds, respectively ...... 104

Figure 3.6 Five types of structural mismatches found in homogeneous and heterogeneous protofilaments formed by Hu, BV and SHaPrP(120-144) peptides. Simulation snapshots of (A) Hu with the Ω-shaped SHa protofilament from the 2nd run; (B) BV with the Ω-shaped SHa protofilament from the 3rd run; (C) BV with the Ω-shaped SHa protofilament from the 4th run; (D) SHa with the Ω-shaped BV protofilament from the 2nd run; (E) SHa with the Ω-shaped SHa protofilament from the 1st run. Hu, BV and SHaPrP(120-144) peptides in A-E are colored in red, blue and green, respectively, except that the SHa(120-144) monomer peptides in E are colored in purple ...... 105

Figure S3.1 (A) The simulation snapshot and (B) the schematic peptide representations of the Ω-shaped BVPrP(120-144) protofilaments are shown. (C) The simulation snapshot and (D) the schematic peptide representation of the Ω-shaped SHaPrP(120-144) protofilaments are shown. The sizes of the side chain beads do not reflect the actual radii. Hydrophobic residues (white), positively charged residues (red), negatively charged residues (blue), and polar residues (green) are shown. (this figure is reproduced from ref 1) ...... 109

Figure S3.2 Plots of potential energy vs simulation time of BV (blue) and SHaPrP(120-144) (red) peptides aggregating in the presence of Ω-shaped HuPrP(120-144) seed ...... 110

Figure 4.1 Simulation snapshots showing the pathway along which 8 Aβ(17-36) peptides form a parallel in-register β-sheet-rich U-shaped protofilament via a templated-growth mechanism. Snapshots are taken at t = (A) 0, (B) 6079,

xi

(C) 7383, (D) 12190, (E) 103775 (ns). F is a schematic side-view representation of the U-shaped protofilament in E. G shows the time evolution of the systems total interaction energy (kJ/mol) ...... 146

Figure 4.2 Root mean square displacement (RMSD) of 4 inhibitor molecules versus simulation time from both atomistic (A) amd coarse-grained (B) simulations, respectively. The reference structures for RMSD calculation are their initial structures. C. Number of contacts between inhibitors and individual amino acids on Aβ(17-36) peptide and; D. Plot of the radial distribution function g(r) of the center of mass distance between the F19 sidechain on Aβ(17-36) peptide and the four inhibitor molecules ...... 147

Figure 4.3 Simulation snapshots of 8 Aβ(17-36) peptides aggregating in the presence of 30 vanillin and EGCG molecules taken at different time points. The peptides are colored blue, the inhibitors are colored magenta, respectively ...... 148

Figure 4.4 Number of hydrogen bonds formed as a function of time during Aβ(17-36) peptides aggregation in the absence and presence of 30 vanillin, resveratrol, curcumin, EGCG molecules ...... 149

Figure 4.5 (A) Peptide sidechain-sidechain and (B) sidechain-inhibitor interaction energy distribution. (Hydrophobic residues are highlighted in blue) ...... 150

Figure 4.6 The interaction energy map of Aβ(17-36) peptide and (A) vanillin, (B) resveratrol, (C) curcumin and (D) EGCG. The scale bar applies for A-D and its unit is kJ/mol. The inhibitor CG sites are grouped by CG types ...... 151

Figure 4.7 The Aβ(17-36)-inhibitor interaction energy distribution for (A) vanillin, (B) resveratrol, (C) curcumin and (D) EGCG ...... 152

Figure 4.8 A-D are simulation snapshots of EGCG redirecting a U-shaped Aβ(17-36) protofilament into a disordered oligomer at t = 242 (A), 923 (B), 1421 (C) and 1990 (D) ns. E plots the number of peptide backbone hydrogen bonds versus simulation time ...... 153

Figure 4.9 A plot of the number of peptide backbone hydrogen bonds remained as a function of simulation time ...... 154

Figure 4.10 Coarse-grained representations of A) resveratrol, B) EGCG, C) vanillin, and D) curcumin to be used in DMD simulations. The side bar defines each coarse-grained group. The CG sites are labeled 1, 2, 3… for use in later discussion. (vanillin: 1-11, resveratrol: 1-17, curcumin: 1-25, EGCG: 1-32) ...... 155

Figure 4.11 Pseudo-bonds used for A) resveratrol, B) EGCG, C) vanillin, and D) curcumin. Red bonds constrain benzene and benzopyran rings, and green, purple and blue bonds constrain spheres once, twice, and three times

xii

removed from the rings ...... 156

Figure S4.1 Number of peptide-inhibitor (A) and peptide-peptide (B) contacts versus simulation time for four small molecules ...... 158

Figure S4.2 Comparison of most populated conformations of vanillin (A and F), resveratrol (B and G), curcumin (C, D, H and I) and EGCG (E and J) from atomistic and coarse-grained simulation trajectories. The carbon, hydrogen and oxygen beads are colored cyan, white and red ...... 159

Figure 5.1 mOLF peptide stretches P1 and P3 recapitulate disparate morphologies of amyloid aggregates derived from full-length mOLF seen by AFM. (A) mOLF fibrils grown at 37 °C with mechanical rocking exhibit straight morphology by AFM while (B) mOLF fibrils grown at 42 °C without rocking exhibit disparate circular morphology. (C) Fibrilization of predicted amyloidogenic peptide stretches, P1, P2 and P3, incubated in 10mM Na2HPO4/KH2PO4 pH 7.2 buffer containing 200mM NaCl at 36 °C, monitored by ThT fluorescence. (D) Amyloidogenic peptide stretches P1 (green) and P3 (purple) identified in (C) highlighted in the context of the native mOLF structure (PDB code 4WXS). (E) P1 fibrils appear straight while (F) P3 fibrils are circular when visualized by AFM. Scale bar is 300 nm for images A, B, E, and F. Images A, B, C, E, F reproduced from ref 1. Reprinted from Journal Molecular Biology, 426, S. E. Hill, R. K. Donegan and R. L. Lieberman, The Glaucoma-Associated Olfactomedin Domain of Myocillin Froms Polymorphic Fibrils that are Constrained by Partial Unfolding and Peptide Sequence, pages 921-935, Copyright (2014), with permission from Elsevier ...... 177

Figure 5.2 DMD/PRIME20 simulations of P1. (A) A plot of potential energy of P1 peptide aggregation versus simulation time. Also shown are simulation snapshots of eight random coil P1 peptides aggregating to form U-shaped (top panel corresponding to the red curve) and S-shaped (bottom panel corresponding to the black curve) P1 protofilaments. Each of the eight peptides has a distinct color. (B) and (C) are the final simulation snapshots of the S-shaped (B) and U-shaped (C) P1 protofilaments, respectively. (D) and (E) are the schematic representation of peptide conformation in the U-shaped and S-shaped P1 protofilaments. Hydrophobic and polar residues are shown in white and green, respectively ...... 178

Figure 5.3 Average number of inter-peptide backbone hydrogen bonds formed per residue of P1 (A) P2 (B) P3 (C). (D) β-sheet propensities calculated for P1, P2 and P3 peptides ...... 179

Figure 5.4 DMD/PRIME20 simulations of P3. (A) A plot of potential energy of P3 peptide aggregation versus simulation time. Also shown are simulation snapshots of eight random coil P3 peptides aggregating to form a disordered oligomer (top panel corresponding to the red curve) and a U-shaped P3

xiii

protofilament (bottom panel corresponding to the black curve). Each of the eight peptides has a distinct color. (B) and (C) are final simulation snapshots of the disordered P3 oligomer and U-shaped P3 protofilament, respectively. (D) The schematic representation of peptide conformation in the U-shaped P3 protofilaments. Hydrophobic and polar residues are shown in white and green, respectively ...... 180

Figure 5.5 1H-13C CPMAS 13C spectrum of P1 and P3 aggregates with no 13C enrichment (A) The spectrum of P1 fibril (red) and the predicted spectrum (generated by summing all Gaussian peaks at the chemical shifts of all 13C in the peptides) with FWHM = 1.2 ppm (black). (B) The region of aliphatic carbons in Panel A. (C) The predicted CPMAS spectrum of P3 fibril (red) and the predicted spectrum with FWHM = 14 ppm (black) ...... 181

Figure S5.1 Clustal Omega1 sequence alignment depicting bioinformatics analysis of amyloid forming region of mOLF by AmylPred 22 (amyl, pink), Zyggregator3 (Zygg, cyan), PASTA4 (Past, purple), Tango5 (yellow) and ZipperDB6 (Zipp, green) ...... 183

Figure S5.2 Simulation snapshots (A-J) of P1 aggregates from ten independent DMD simulations ...... 184

Figure S5.3 Simulation snapshots (A-J) of P2 aggregates from ten independent DMD simulations ...... 185

Figure S5.4 Simulation snapshots (A-J) of P3 aggregates from ten independent DMD simulations ...... 186

Figure 6.1 (A) Representation of the Aβ16-22 peptide in PRIME20 model. (B) and (C) are schematic representations of the square well interaction between two sidechains amd directional square well interaction (hydrogen bonding) between backbone NH and C=O beads. D-H are simulation snapshots of non-HB oligomer, HB oligomer, 2, 3 and 4 β-sheet fibril, respectively, generated by visual molecular dynamics software (VMD). The peptides in the five aggregates and in solution are shown in green and magenta, respectively ..... 216

Figure 6.2 The probability distribution for (A) the number of inter/intra-peptide hydrogen bonds (NHB) per peptide and (B) the average number of hydrogen bonds per residue for the five types of aggregates. The probability distribution for (C) the average number of sidechain contacts (NHP) per peptide and (D) the average number of sidechain contacts per residue for the five aggregates. The data for the non-HB oligomer (orange) and the HB oligomer (purple) are measured at T=193K and 250K, respectively, and the data for the 2β-sheet (red), 3β-sheet (green) and 4β-sheet (blue) are measured at T=330K. These five aggregates are characterized under different temperatures because their solubility measurements have relatively small standard deviations at low temperature, as

xiv

shown from Fig. 3A ...... 217

Figure 6.3 (A) a plot of the solubility 퐶푒 for non-HB oligomer, HB oligomer, 2β, 3β, 4β-sheet fibrils as a function of the temperature; (B) Solubility data for non-HB oligomer, HB oligomer, 2β, 3β, 4β-sheet fibrils at four to six given different temperatures; lines are fitted to Eq. (1); the color scheme is the same as plot A; (C) Comparison of the average binding energy per peptide within the aggregate and the corresponding latent heat of peptide aggregation into oligomer and fibril.; (D) Dependence of the fibril solubility on thickness (i=2, 3 and 4) for 2β, 3β, 4β-sheet fibrils, respectively; lines are fitted to Eq. (2) ...... 218

Figure 6.4 (A) Phase diagram for Aβ16-22 peptide. Solubility (Ce)-vs-temperature (T) data are presented for non-HB oligomer (orange dotted line), HB oligomer (purple dotted line), 2β-sheet fibril (red dotted line), 3β-sheet fibril (green dotted line), 4β-sheet fibril (blue dotted line) and the infinite layer (∞) β-sheet fibril (black line). The fibril and solution phases are colored yellow and cyan, respectively. (B) DMD/PRIME20 simulation snapshots at the concentration and temperature of phase point A and B are taken at different time points, respectively ...... 219

Figure 6.5 (A) Summary of both the simulation-predicted temperature-dependent solubility line 퐶푒(T) for Aβ16-22 peptide and the fibrillation experiments performed under given conditions. Red dots and red circles indicate conditions at which fibrils have been found via TEM to form, and not to form, respectively, (at temperature 277K, 296K, 310K and 330K). Blue dots indicate that fibrils have been reported in the literature to form under these conditions (see Fig. S6.6). (B) Six selected transmission electron micrographs (TEM) images (i-vi) showing that Aβ16-22 form fibrils under the conditions that correspond to the six red dots labeled i-vi in (A). Buffer: 100 mM ammonium bicarbonate, pH=7, with a final concentration of 1% DMSO (v/v). Scale bar: 200 nm ...... 220

2 Figure S6.1 Specific surface energy σh (mJ/m ) of Aβ16-22 fibril versus temperature calculated from fitting simulation data (푙푛퐶𝑖훽 vs 1/i) to Eq. (2) ...... 221

Figure S6.2 A plot of the estimated solubility of Aβ16-22 fibril versus fibril thickness (number of β-sheets) at various given temperatures Data for this plot are extracted the linear fit of 푙푛퐶𝑖훽 - 1/푖 (i=2,3,4) data to Eq. (2) from Fig. 3D ...... 222

Figure S6.3 Illustration of the Helmholtz free energy (A) versus peptide composition (x) for a binary (peptide-water) solution system at constant volume (V) and temperature (T). The Helmholtz free energy is considered here instead of the Gibbs free energy because the simulations are done at constant volume rather than constant pressure ...... 223

Figure S6.4 Snaphot from a simulation showing the structure of a two-layer β-sheet

xv

fibril of Aβ16-22 peptide (colored green and yellow) in an anti-parallel β-sheet configuration. The oxygen (O), nitrogen (NH) and carbon (Cα, CO) beads on the peptide backbone are shown in red, blue and cyan. Note that the sidechain beads are not represented here. The O beads (red) are incorporated into the C=O bead and do not exist independently in PRIME20 model, the positions of which are determined after simulation ...... 224

Figure S6.5 Plots of (A) total number of peptides and (B) the average peptide binding energy (kJ/mol) of a 4β-sheet fibril versus simulation time at T=330K ...... 225

Figure S6.6 Literature conditions (check marks) under which Aβ16-22 has been shown to form fibrillar structures (i.e. at neutral pH). The references used to form these figures can be found in the box next to the heat-map. N.D. means that there is no literature study performed under this conditions, however, fibrils would be presumed to form given the surrounding literature examples; ? means no experiments have been conducted under these conditions ...... 216

Figure 7.1 Co-aggregation of Aβ16-22 and Aβ40 results in accelerated aggregation

kinetics for Aβ40. (a) The primary sequence of Aβ16-22 and Aβ40, including the groups at each termini. (b) ThT fluorescence assays showing when mixing the two peptides (with the total peptide concentration held constant at 40

µM) the aggregation rate of Aβ40 increases as the ratio of Aβ16-22 to Aβ40 is

increased. (c) Simulation snapshots of the aggregation of six Aβ40 monomers into a β-sheet rich hexamer at CAβ40 = 5mM; at the start of the simulation (0 µs) all the peptides are in random coils but as the simulation progresses they aggregate into parallel, in-register β-sheets (104 µs). This oligomer then unfolds, losing some of its β-sheet structure (230 µs) prior to a rearrangement

in which the parallel β-sheets rearrange, forming a stable fibril with each Aβ40 peptide containing three β-strands (621 µs) ...... 253

Figure 7.2 Aggregation kinetics of Aβ16-22 are unaffected by the presence of Aβ40. (a) Schematic showing the principle behind the fluorescence quenching assay

used to determine the aggregation rate of Aβ16-22. (b) As self-assembly occurs, the TAMRA-labelled peptides (40 µM, total peptide, containing 5% (w/w)

TAMRA-Ahx-Aβ16-22 are sequestered into the fibril structure. This brings the fluorophores into close proximity, resulting in fluorescence quenching. (c)

Aggregation of Aβ16-22 (containing 5% (w/w) TAMRA-Ahx-Aβ16-22) and Aβ40 at a 1:1 mol/mol ratio (total peptide concentration 40 µM). (d, e) Sedimentation and separation of the pellet and supernatant of the 1:1 mixed

system after 1 h indicates that Aβ40 is present, as indicate by ESI-MS, in the (d) supernatant and only very small amounts within the (e) pellet. (f)

Under these conditions, fibrils of Aβ16-22 are present after 5 mins incubation. Scale bar: 500 nm ...... 254

xvi

Figure 7.3 Aβ16-22 can interact with both Aβ40 monomers and dimers. (a) Native

ESI-IMS-MS drift-scope images of Aβ40 indicate the presence of multiple

oligomeric species of Aβ40 (white numbers). (b) When incubated at a 1:1

mol/mol ratio with Aβ16-22 (yellow numbers) a number of heteromeric species are observed (green numbers). The oligomer size is given (1, 2, 3 etc) with the charge state in superscript. (c) DMD simulation showing the percent

β-sheet formed by Aβ40 during aggregation in the absence (red) or presence

(black) of Aβ16-22. (d) Energy contact map between one monomer of Aβ16-22

and one of Aβ40 scaled by energy (bar shown alongside) showing that residues 17-20 (LVFF) and 31-34 (IIGL) form the strongest interactions. (e) Co-aggregation can have differing effects on the 1° nucleation of each peptide, depending on whether the mixed oligomers formed can progress to form mixed fibrils or are off-pathway and take no further part in the aggregation reaction. Circles represent monomers and blocks represent fibrils,

with Aβ16-22/Aβ40 in red and blue respectively. Adapted from (35) ...... 255

Figure 7.4 Aβ16-22 fibrils increase the aggregation rate of Aβ40 to a greater extent than

Aβ16-22 monomers. (a) Increased concentrations (% w/w) of Aβ16-22 fibrils were

added to Aβ40 monomers (as shown in the key) and the aggregation rate measured by ThT fluorescence. (b) Direct comparison of the effect of

Aβ16-22 monomers (i.e. taken straight from a DMSO stock) and Aβ16-22 fibrils

on Aβ40 aggregation. (c) Effect of sonicating the Aβ16-22 fibrils on Aβ40 aggregation rate shows little effect compared with the data shown in (a) (see

text for details). (d) Plots of the percent β-sheet formed by Aβ40 in the absence (blue) or presence of preformed 2 (black), 3 (red) and 4 (green)

β-sheet Aβ16-22 fibrils showing that an increased number of Aβ16-22 fibrils

increases the rate of Aβ40 aggregation. (e) During co-aggregation experiments both elongation and surface catalysed mechanisms can occur, each has a different effect on the rate of assembly of each peptide (the same notation is used as in Fig. 3e, with circles representing monomers, blocks fibrils and

Aβ16-22/Aβ40 in red and blue, respectively). Adapted from (35) ...... 256

Figure 7.5 Aβ16-22 and Aβ40 do not co-assemble during co-aggregation. Negative stain

TEM analysis of Aβ40 incubated for 24 h in the (a) absence or (b) presence of

Aβ16-22. Scale bar = 200 nm. (c) PIC of mixtures of diazirine labelled Aβ16-22

(Aβ*16-22) and Aβ40 incubated for 24 h and then irradiated for 30 s. Only

homomolecular Aβ16-22 cross-links are observed, indicating that the fibrils are not co-polymerised at the end of the reaction (the inset depicts mechanism of PIC of the diazirine group (see also Fig. S12). (d) DMD simulation snapshots

of co-aggregation of Aβ40 (blue) and Aβ16-22 (red) indicate that separate homomolecular oligomers are formed at t = 202 µs ...... 257

xvii

Figure 7.6 Aβ16-22 fibrils catalyse Aβ40 assembly through secondary surface nucleation.

(a) Simulation snapshots of the process by which Aβ16-22 fibrils (red) increase

the aggregation rate of Aβ40 (blue) through a surface catalysed 2° nucleation.

(b) A schematic description of the mechanism is also included with Aβ40 in

blue and Aβ16-22 in red ...... 258

Figure S7.1 HRMS and analytical HPLC traces of Aβ16-22 and its variants ...... 262

Figure S7.2 SEC trace of Aβ40 indicates that there is a single peak and ESI-IMS-MS indicates that in the gas phase Aβ40 is largely monomeric ...... 263

Figure S7.3 Supplementary ThT data ...... 264

Figure S7.4 Supplementary negative stain TEM images ...... 265

Figure S7.5 Analysis of the CCS values for Aβ40 in the absence or presence of Aβ16-22 over different IMS experiments ...... 267

Figure S7.6 PIC analysis of 1:1 Aβ*16-22/Aβ40 at 5 min. and 24 h ...... 268

Figure S7.7 Plot of the average number of hydrogen bonding and sidechain-sidechain contacts between six Aβ40 monomers and a 3-β-sheet Aβ16-22 fibril during the simulation at CAβ40 = 5mM. As shown in Fig. 6b, Aβ40 monomers attach to both the lateral surface and the end of the Aβ16-22 fibril with its C-terminal part attaching more frequently to the lateral surface of the fibril than its two ends ...... 270

xviii

CHAPTER 1

Motivation and Overview

1.1 Motivation

The misfolding and aggregation of protein into amyloid plaques in the human brain is associated with over fifty types of fatal neurodegenerative diseases, including Alzheimer’s,

Parkinson’s, Huntington’s and the Prion diseases. Researchers hypothesize that amyloid formation is the primary cause of these fatal diseases; this is the so-called amyloid cascade hypothesis.1,2

Although amyloid research has been going on for over thirty years, the in vivo and in vitro mechanism underlying amyloid formation and the associated disease pathology still remain unclear.3 To date, there are no effective therapeutic treatments for any amyloid diseases, which pose a big threat to public health worldwide.4 The failures of clinical trials for many promising drug candidates over the last decade, including verubecestat5 and aducanumab 6, suggest that the development of effective treatments is still in its infancy. Fortunately, the good new is that a recent- developed antibody drug, BAN24017 aimed at binding to the Amyloid-β protofibril, has been found to slow down or alleviate patient dementia and other cognitive symptoms.

Over the past thirty years, researchers have developed a variety of biophysical and biochemical experimental approaches to characterize and investigate amyloid formation in vitro and in vivo. Atomic force microscopy (AFM) has been applied to examine fibril length and width distribution, 8 and to track the dynamics of fibril growth9 by attaching fibrils onto a surface. X-ray crystallography has been used to determine fibril structure 10 from diffraction patterns for fibril microcrystals formed by short amyloidogenic peptides. Circular dichroism spectroscopy (CD) and

Fourier-transform infrared spectroscopy (FTIR) have been applied to measure the secondary structure profile of proteins during the aggregation process 11,12. Florescent dyes such as Thioflavin

1

T13 and Congo red14 have been widely used to identify amyloid formation and to track fibril growth kinetics. Notably, the application of solid-state nuclear magnetic resonance spectroscopy (ssNMR) and cryo-electron microscopy (Cryo-EM) are magnificent achievements in elucidating the polymorphic molecular structure of amyloid fibrils formed by the amyloid β(1-40)15, amyloid β(1-

42)16, α-Synuclein protein17, β2 microglobulin18 and tau protein19. In addition, ion mobility spectrometry-mass spectrometry (IMS-MS) has been applied to detect the distribution of small oligomeric aggregates based on the difference in samples’ collision cross-sectional area and charge to mass ratio.20

Theoretical and computational approaches have also been developed to study protein aggregation. Firstly, atomistic molecular dynamics simulations are mainly applied to study the structural stability of a pre-formed fibril21 and the early-stage oligomerization22 process for small systems (

Over the last fifteen years, Hall and co-workers have been working on developing coarse- grained models for simulating protein aggregation. In 2004, Nguyen et al.30 applied discontinuous

2

molecular dynamics simulation (DMD) in combination with an intermediate resolution protein model called PRIME to simulate the dynamic aggregation process of 48 polyalanine peptides from random coils to fibrillar structures. In 2010, Cheon et al.31 extended the PRIME model to PRIME20 which could model sequence-specific peptides with twenty different amino acids. The PRIME20 force field takes into account the directional hydrogen bonding between backbone atoms. The sidechain-sidechain polar, electrostatic and hydrophobic interactions between each pair of interacting amino acid residues are unique. The energetic parameters were obtained by fitting to the structures of 711 native-state globular proteins in the Protein Data Bank (PDB) and 2 million decoyed structures. DMD/PRIME20 simulations are quite fast and can access time scales of the order of 50 μs32; the longest run to date is 572µs33. The ability to access long time scales allows us to simulate the spontaneous formation of fibrillar structures starting from a very large system of peptides in a random configuration. This is not possible in atomistic simulations except for very small systems. DMD/PRIME20 simulations have been extensively validated against experimental measurements on a number of systems including Aβ(16-22), Aβ(17-42) and tau protein fragment,

34-37 and found to give a relatively realistic snapshot of the structures that form and the events that occur as amyloid disease-related protein fragments aggregate in vitro.

The goal of the work in this thesis is to apply molecular simulation techniques, with the help of modern biophysical techniques (e.g. mass spectrometry, photo cross-linking techniques and solid-state NMR spectroscopy), to tackle key fundamental questions underlying in vitro amyloid formation. Specifically, we examine the non-seeding, seeding and cross-seeding aggregation of prion protein fragments, the inhibitory effect of polyphenols on Aβ(17-36) aggregation, the aggregation propensities and structural polymorphism of glaucoma-associated

3

myocilin protein fragments, the phase behavior of the Aβ(16-22) peptide and the co-aggregation of Aβ(1-40) with Aβ(16-22) peptides.

1.2 Overview

In this section a summary of the rest of the dissertation is provided. All chapters provide their own literature review and references.

1.2.1 N-Terminal Prion Protein Peptides (PrP(120-144)) Form Parallel In-Register β-sheets via Multiple Nucleation-Dependent Pathways

The prion diseases are a family of fatal neurodegenerative diseases associated with the misfolding and accumulation of normal prion protein (PrPC) into its pathogenic scrapie form

(PrPSc). Understanding the fundamentals of prion protein aggregation and the molecular architecture of PrPSc is key to unravelling the pathology of prion diseases. Our work investigates the early-stage aggregation of three prion protein peptides, corresponding to residues 120 to 144 of human (Hu), bank vole (BV) and Syrian hamster (SHa) prion protein, from disordered monomers to β-sheet-rich fibrillar structures. Using discontinuous molecular dynamics simulation combined with the PRIME20 force field, we find that the Hu, BV and Sha PrP120-144 aggregate via multiple nucleation-dependent pathways to form U-shaped, S-shaped and Ω-shaped protofilaments. The S-shaped HuPrP120-144 protofilament is similar to the amyloid core structure of HuPrP112-141 predicted by Zweckstetter. HuPrP120-144 has a shorter aggregation lag phase than BVPrP120-144, followed by SHaPrP120-144, consistent with experimental findings. Two amino acid substitutions I138M and I139M retard the formation of parallel in-register β-sheet dimers during the nucleation stage by increasing sidechain-sidechain association and reducing

4

sidechain interaction specificity. On average, HuPrP120-144 aggregates contain more parallel β- sheet content than those formed by BV and SHa PrP120-144. Deletion of C-terminal residues 138 to 144 prevents formation of fibrillar structures in agreement with experiment. This work sheds light on the amyloid core structures underlying prion strains and how I138M, I139M and S143N affect prion protein aggregation kinetics.

1.2.2 Seeding and Cross-Seeding of N-Terminal Prion Protein Peptides PrP(120-144)

Prion diseases are infectious neurodegenerative diseases that are capable of cross-species transmission, thus arousing public health concerns. Seed-templating propagation of prion protein is believed to underlie prion cross-species transmission pathology. Understanding the molecular fundamentals of prion propagation is key to unravelling the pathology of prion diseases. In this work, we use coarse-grained molecular dynamics to investigate the seeding and cross-seeding aggregation of three prion protein fragments PrP(120-144) originating from human (Hu), bank vole (BV) and Syrian hamster (SHa). We find that the seed accelerates the aggregation of the monomer peptides by eliminating the lag phase. The monomer aggregation kinetics are mainly determined by the structure of the seed. The stronger the hydrophobic residues on the seed associate with each other, the higher the probability that the seed recruits monomer peptides to its surface/interface. For cross-seeding aggregation, we show that Hu has a strong tendency to adopt the conformation of the BV seed and vice versa; the Hu and BV monomers have a weak tendency to adopt the conformation of the SHa seed. These two findings are consistent with the experimental findings of Apostol et al. on PrP(138-143) and partially consistent with the finding of Jones et al. on PrP(23-144). We also identify several conformational mismatches when SHa cross-seeds BV and Hu peptides, indicating the existence of a cross-seeding barrier between SHa and the other two

5

sequences. This work sheds light on the molecular mechanism of seed-templating aggregation of prion protein fragments underlying the sequence-dependent transmission barrier in prion diseases.

1.2.3 Aggregation of Aβ(17-36) in the Presence of Naturally Occurring Phenolic Inhibitors

Using Coarse-Grained Simulations

Although some naturally-occurring polyphenols have been found to inhibit amyloid β fibril formation and reduce neuron cell toxicity in vitro, their exact inhibitory mechanism is unknown.

In this work, discontinuous molecular dynamics (DMD) combined with the PRIME20 force field and a newly-built inhibitor model are performed to examine the effect of vanillin, resveratrol, curcumin and epigallocatechin-3-gallate (EGCG) on the aggregation of Aβ(17-36) peptides. Four sets of peptide/inhibitor simulations are performed in which inhibitors: (1) bind to Aβ(17-36) monomer; (2) interfere with Aβ(17-36) oligomerization; (3) disrupt a pre-formed Aβ(17-36) protofilament; (4) prevent the growth of Aβ(17-36) protofilament. The single-ring compound, vanillin, slightly slows down but cannot inhibit the formation of a U-shaped Aβ(17-36) protofilament. The multiple-ring compounds, EGCG, resveratrol and curcumin, redirect Aβ(17-

36) from a fibrillar aggregate to an unstructured oligomer. The three aromatic groups of the EGCG molecule are in a stereo (nonplanar) configuration, helping it contact the N-terminal, middle and

C-terminal regions of the peptide. Resveratrol and curcumin bind only to the hydrophobic residues near the peptide termini. The rank order of inhibitory effectiveness of Aβ(17-36) aggregation is :

EGCG > resveratrol > curcumin > vanillin, consistent with experimental findings on inhibiting full length Aβ fibrillation. Furthermore, we learn that the inhibition effect of EGCG is specific to the peptide sequence while those of resveratrol and curcumin are non-specific in that they stem from strong interference with hydrophobic sidechain association, regardless of the residues’ location

6

and peptide sequence. Our studies provide molecular-level insights into how polyphenols inhibit

Aβ fibril formation, knowledge that could be useful for designing amyloid inhibitors.

1.2.4 Simulations and Experiments Delineate Amyloid Fibrillation by Peptides Derived from

Glaucoma-Associated Myocilin

Mutant myocilin aggregation is associated with inherited open angle glaucoma, a prevalent optic neuropathy leading to blindness. Comprehension of mutant myocilin aggregation is of fundamental importance to glaucoma pathogenesis and ties glaucoma to amyloid diseases such as

Alzheimer’s. Here we probe the aggregation properties of peptides derived from the myocilin olfactomedin domain. Peptides P1 (residues 326-337) and P3 (residues 426-442) were found previously to form amyloid. Coarse-grained discontinuous molecular dynamics simulations using the PRIME20 force field (DMD/PRIME20) predict that P1 and P3 are aggregation-prone; P1 consistently forms fibrillar aggregates with parallel in-register β-sheets whereas P3 forms β-sheet- containing aggregates without distinct order. Natural abundance 13C solid-state NMR spectra validate that aggregated P1 exhibits amyloid signatures and is less heterogeneous than aggregated

P3. DMD/PRIME20 simulations provide a viable method to predict peptide aggregation propensities and the aggregate’s structure/order which cannot be accessed by bioinformatics or readily attained experimentally.

1.2.5 Thermodynamic Phase Diagram of Amyloid-β (16-22) Peptide

The aggregation of monomeric Aβ peptide into oligomers and amyloid fibrils in the mammalian brain is associated with Alzheimer’s disease. Insight into the thermodynamic stability of the Aβ peptide in different polymeric states is fundamental to defining and predicting the

7

aggregation process. Experimental determination of Aβ thermodynamic behavior is challenging due to the transient nature of Aβ oligomers and the low peptide solubility. Furthermore, quantitative calculation of a thermodynamic phase diagram for a specific peptide requires extremely long computational times. Here, using a coarse-grained protein model, molecular dynamics simulations are performed to determine an equilibrium concentration and temperature phase diagram for the amyloidogenic peptide fragment, Aβ16-22. Our results reveal that the only thermodynamically stable phases are the solution phase and the macroscopic fibrillar phase, and that there also exists a hierarchy of metastable phases. The boundary line between the solution phase and fibril phase is found by calculating the temperature-dependent solubility of a macroscopic Aβ16-22 fibril consisting of an infinite number of β-sheet layers. To our knowledge, this is the first in silico determination of an equilibrium (solubility) thermodynamic phase diagram for a real amyloid-forming peptide. Furthermore, the in silico prediction of Aβ16-22 solubilities over the temperature range of 4-57°C agrees well with fibrillation experiments and transmission electron microscopy measurements of the fibril morphologies formed. This in silico approach of predicting peptide solubility is also potentially useful for optimizing biopharmaceutical production and manufacturing nanofiber scaffolds for tissue engineering.

1.2.6 Molecular Insights into the Surface Catalyzed Secondary Nucleation of Aβ(1-40) by the

Peptide Fragment Aβ(16-22)

The formation of amyloid deposits from usually soluble peptides and proteins is implicated in a range of debilitating and prevalent diseases. Understanding, at the molecular level, the mechanism by which these complex and heterogeneous systems aggregate is a significant challenge facing the scientific community. Amyloid research in vitro often focuses on a single

8

highly pure peptide sequence, however, in vivo there can be a number of different amyloidogenic peptides that interact during the aggregation process. In the present work, we explore the aggregation of two widely studied peptides from the Aβ sequence, Aβ40 and Aβ16-22, when incubated as a mixture. Using a combination of biochemical and biophysical assays, such as ion mobility spectrometry mass spectrometry and photo-induced cross-linking, we demonstrate that

Aβ16-22 increases the aggregation rate of Aβ40 through a surface catalysed secondary nucleation pathway. The same system is also modelled using a powerful discontinuous molecular dynamics simulation and PRIME20 force field, which simulates the spontaneous aggregation of both peptides from random-coil monomer state to their final fibril structure. The simulation, consistent with experimental findings, provides molecular insights into the surface catalysed 2° nucleation, a key driver of Aβ aggregation.

1.2.7 Future Work

In Chapter 8, we provide an outline for future work on (1) parallelization of the PRIME20 code; (2) construction of the free energy landscape of peptide oligomerization.

1.3 Publications

Chapters 2-5 are based on the following publications: (* means equal contribution)

Chapter 2: Wang, Y., Shao, Q., Hall, C. K., (2016) N-terminal Prion Protein Peptides PrP(120-

144) Form Parallel In-register β-sheets via Multiple Nucleation-dependent Pathways. J. Biol.

Chem. 291(42), 22093-22105.

Chapter 3: Wang, Y., Hall, C. K., (2018) Seeding and Cross-seeding fibrillation of N-terminal

Prion Protein Peptide PrP(120-144). Protein Sci. 27(7), 1304-1313.

9

Chapter 4: Wang, Y., Latshaw, D. C., Hall, C. K., (2017) Aggregation of Amyloid Beta (17-36) in the Presence of Naturally Occurring Phenolic Inhibitors Using Coarse-Grained Simulations. J.

Mol. Biol. 429 (24), 3893-3908.

Chapter 5: Wang, Y., Gao, Y., Hill, S. E., Huard, D. J., Tomlin, M. O., Lieberman, R. L.,

Paravastu, A. K., Hall, C. K., (2018) Simulations and Experiments Delineate Amyloid

Fibrillization by Peptides Derived from Glaucoma-Associated Myocilin. J. Phys. Chem. B

122(12), 5845-5850

Chapter 6: Wang, Y., Bunce, S., Radford, S. E., Wilson A. J., Auer, S., Hall, C. K., (2018)

Thermodynamic Phase Diagram of Amyloid β(16-22) Peptide. Submitted to Proc. Natl. Acad. Sci.

U. S. A.

Chapter 7: Bunce S.*, Wang, Y.*, Stewart, K. L., Ashcroft, A. E., Radford, S. E., Hall C. K.,

Wilson, A. J., (2018) Molecular Insights into the Surface Catalyzed Secondary Nucleation of

Aβ(1-40) by the Peptide Fragment Aβ(16-22). Submitted to Sci. Adv.

10

1.4 References

1. Ferreira, S.T., Klein, W.L. (2011) The Aβ oligomer hypothesis for synapse failure and

memory loss in Alzheimer's disease. Neurobiol. Learn Mem. 96, 529-543

2. Knowles, T.P.J., Vendruscolo, M., Dobson, C.M. (2014) The amyloid state and its

association with protein misfolding diseases. Nat. Rev. Mol. Cell Biol. 15, 384-396.

3. Eisenberg, D., Jucker, M. (2012) The amyloid state of proteins in human diseases. Cell 148,

1188-1203.

4. Hardy, J., De Strooper, B. (2017) Alzheimer’s disease: where next for anti-amyloid

therapies? Brain 140, 835-855.

5. Kennedy, M.E., Stamford, A.W., Chen, X., Cox, K., Cumming, J.N., Dockendorf, M.F.,

Egan, M., Ereshefsky, L., Hodgson, R.A., Hyde, L.A. and Jhee, S., 2016. The BACE1

inhibitor verubecestat (MK-8931) reduces CNS β-amyloid in animal models and in

Alzheimer’s disease patients. Sci. Transl. Med. 8, 363ra150.

6. Sevigny, J., Chiao, P., Bussière, T., Weinreb, P.H., Williams, L., Maier, M., Dunstan, R.,

Salloway, S., Chen, T., Ling, Y. and O’Gorman, J. (2016) The antibody aducanumab

reduces Aβ plaques in Alzheimer’s disease. Nature, 537, 50-56.

7. Söllvander, S., Nikitidou, E., Gallasch, L., Zyśk, M., Söderberg, L., Sehlin, D., Lannfelt,

L. and Erlandsson, A. (2018) The Aβ protofibril selective antibody mAb158 prevents

accumulation of Aβ in astrocytes and rescues neurons from Aβ-induced cell death. J.

Neuroinflammation, 15, 98-98.

8. Adamcik, J., Jung, J.M., Flakowski, J., De Los Rios, P., Dietler, G. and Mezzenga, R. (2010)

Understanding amyloid aggregation by statistical analysis of atomic force microscopy

images. Nat. Nanotech. 5,423-428.

11

9. Watanabe-Nakayama, T., Ono, K., Itami, M., Takahashi, R., Teplow, D.B. and Yamada, M.

(2016) High-speed atomic force microscopy reveals structural dynamics of amyloid β1–42

aggregates. Proc. Natl. Acad. Sci. U. S. A. 201524807.

10. Sawaya, M.R., Sambashivan, S., Nelson, R., Ivanova, M.I., Sievers, S.A., Apostol, M.I.,

Thompson, M.J., Balbirnie, M., Wiltzius, J.J., McFarlane, H.T. and Madsen, A.Ø. 2007

Atomic structures of amyloid cross-β spines reveal varied steric zippers. Nature, 447, 453-

457.

11. Hsieh, M.C., Liang, C., Mehta, A.K., Lynn, D.G. and Grover, M.A. (2017) Multistep

conformation selection in amyloid assembly. J. Am. Chem. Soc. 139, 17007-17010.

12. Shim, S.H., Gupta, R., Ling, Y.L., Strasfeld, D.B., Raleigh, D.P. and Zanni, M.T. (2009)

Two-dimensional IR spectroscopy and isotope labeling defines the pathway of amyloid

formation with residue-specific resolution. Proc. Natl. Acad. Sci. U. S. A. 106, 6614-6619.

13. Naiki, H., Higuchi, K., Hosokawa, M. and Takeda, T. (1989) Fluorometric determination

of amyloid fibrils in vitro using the fluorescent dye, thioflavine T. Anal. Biochem. 177,

244-249.

14. Puchtler, H., Sweat, F. and Levine, M. (1962) On the binding of Congo red by amyloid. J.

Histochem. Cytochem. 10, 355-364.

15. Sachse, C., Fändrich, M. and Grigorieff, N. (2008) Paired β-sheet structure of an Aβ (1-40)

amyloid fibril revealed by electron microscopy. Proc. Natl. Acad. Sci. U. S. A. 105, 7462-

7466.

16. Wälti, M.A., Ravotti, F., Arai, H., Glabe, C.G., Wall, J.S., Böckmann, A., Güntert, P., Meier,

B.H. and Riek, R. (2016) Atomic-resolution structure of a disease-relevant Aβ (1-42)

amyloid fibril. Proc. Natl. Acad. Sci. U. S. A. 113, E4976-E4984.

12

17. Tuttle, M.D., Comellas, G., Nieuwkoop, A.J., Covell, D.J., Berthold, D.A., Kloepper, K.D.,

Courtney, J.M., Kim, J.K., Barclay, A.M., Kendall, A. and Wan, W. (2016) Solid-state

NMR structure of a pathogenic fibril of full-length human α-synuclein. Nat. Struct. Mol.

Biol. 23, 409-415.

18. White, H.E., Hodgkinson, J.L., Jahn, T.R., Cohen-Krausz, S., Gosal, W.S., Müller, S.,

Orlova, E.V., Radford, S.E. and Saibil, H.R. (2009) Globular tetramers of β2-microglobulin

assemble into elaborate amyloid fibrils. J. Mol. Biol, 389, 48-57.

19. Fitzpatrick, A.W., Falcon, B., He, S., Murzin, A.G., Murshudov, G., Garringer, H.J.,

Crowther, R.A., Ghetti, B., Goedert, M. and Scheres, S.H. (2017) Cryo-EM structures of

tau filaments from Alzheimer’s disease. Nature, 547,185-190.

20. Young, L.M., Saunders, J.C., Mahood, R.A., Revill, C.H., Foster, R.J., Tu, L.H., Raleigh,

D.P., Radford, S.E. and Ashcroft, A.E. (2015) Screening and classifying small-molecule

inhibitors of amyloid formation using ion mobility spectrometry-mass spectrometry. Nat.

Chem. 7, 73-81.

21. Xiao, Y., Ma, B., McElheny, D., Parthasarathy, S., Long, F., Hoshi, M., Nussinov, R. and

Ishii, Y. (2015) Aβ (1–42) fibril structure illuminates self-recognition and replication of

amyloid in Alzheimer's disease. Nat. Struct. Mol. Biol. 22, 499-505.

22. Brown, A.M., Bevan, D.R. (2016) Molecular dynamics simulations of amyloid β-peptide

(1-42): tetramer formation and membrane interactions. Biophys. J. 111, 937-949.

23. Monticelli, L., Kandasamy, S.K., Periole, X., Larson, R.G., Tieleman, D.P. and Marrink,

S.J., (2008) The MARTINI coarse-grained force field: extension to proteins. J. Chem.

Theory Comput. 4, 819-834.

24. Sterpone, F., Melchionna, S., Tuffery, P., Pasquali, S., Mousseau, N., Cragnolini, T.,

13

Chebaro, Y., St-Pierre, J.F., Kalimeri, M., Barducci, A. and Laurin, Y. (2014) The OPEP

protein model: from single molecules, amyloid formation, crowding and hydrodynamics

to DNA/RNA systems. Chem. Soc. Rev. 43, 4871-4893.

25. Davtyan, A., Schafer, N.P., Zheng, W., Clementi, C., Wolynes, P.G. and Papoian, G.A.

(2012) AWSEM-MD: protein structure prediction using coarse-grained physical potentials

and bioinformatically based local structure biasing. J. Phys. Chem. B. 116, 8494-8503.

26. Baftizadeh, F., Biarnes, X., Pietrucci, F., Affinito, F., Laio, A. (2012) Multidimensional

view of amyloid fibril nucleation in atomistic detail. J. Am. Chem. Soc. 134, 3886-3894.

27. Han, W. and Schulten, K. (2014) Fibril elongation by Aβ17-42: Kinetic network analysis

of hybrid-resolution molecular dynamics simulations J. Am. Chem. Soc. 136, 12450-12460.

28. Zheng, W., Tsai, M.Y., Chen, M. and Wolynes, P.G. (2016) Exploring the aggregation free

energy landscape of the amyloid-β protein (1–40). Proc. Natl. Acad. Sci. U. S. A. 113,

11835-11840.

29. Röder, K. and Wales, D.J. (2018) Energy Landscapes for the Aggregation of Aβ17-42. J.

Am. Chem. Soc. 140, 4018-4027.

30. Nguyen, H.D., Hall, C.K. (2004) Molecular dynamics simulations of spontaneous fibril

formation by random-coil peptides. Proc. Natl. Acad. Sci. U. S. A. 101, 16180-16185.

31. Cheon, M., Chang, I., Hall, C.K. (2010) Extending the PRIME model for protein

aggregation to all 20 amino acids. Proteins. Struct. Funct. Bioinf. 78, 2950-2960.

32. Wang, Y., Hall, C. K. (2018) Seeding and Cross seeding Fibrillation of Prion Protein

Peptides PrP(120-144). Protein Sci., 27, 1304-1313.

33. Wang, Y., Shao, Q., Hall, C. K. (2016) N-terminal Prion Protein Peptides (PrP(120–144))

Form Parallel In-register β-Sheets via Multiple Nucleation-dependent Pathways. J. Biol.

14

Chem. 291, 22093-22105.

34. Cheon, M., Chang, I., Hall, C. K. (2011) Spontaneous Formation of Twisted A beta(16-22)

Fibrils in Large-Scale Molecular-Dynamics Simulations. Biophys. J. 101, 2493-2501.

35. Wagoner, V. A., Cheon, M., Chang, I., Hall, C. K. (2012) Fibrillization Propensity for Short

Designed Hexapeptides Predicted by Computer Simulation. J. Mol. Biol. 416, 598-609.

36. Cheon, M., Chang, I., Hall, C. K. (2012) Influence of temperature on formation of perfect

tau fragment fibrils using PRIME20/DMD simulations. Protein Sci. 21, 1514-1527.

37. Cheon, M., Hall, C. K., Chang, I. (2015) Structural Conversion of A beta(17-42) Peptides

from Disordered Oligomers to U-Shape Protofilaments via Multiple Kinetic Pathways.

PLoS Comput. Biol. 11, e1004258.

15

CHAPTER 2

N-Terminal Prion Protein Peptides (PrP(120-144)) Form Parallel In-Register β-sheets via

Multiple Nucleation-Dependent Pathways

Chapter 2 is essentially a manuscript by Yiming Wang, Qing Shao, and Carol K Hall published on

Journal of Biological Chemistry.

16

N-Terminal Prion Protein Peptides (PrP(120-144)) Form Parallel In-register β-sheets via Multiple Nucleation-dependent Pathways

Yiming Wang, Qing Shao and Carol K. Hall*

Department of Chemical and Biomolecular Engineering, North Carolina State University, Raleigh, North Carolina 27695-7905, United States

Abstract

The prion diseases are a family of fatal neurodegenerative diseases associated with the misfolding and accumulation of normal prion protein (PrPC) into its pathogenic scrapie form

(PrPSc). Understanding the fundamentals of prion protein aggregation and the molecular architecture of PrPSc is key to unravelling the pathology of prion diseases. Our work investigates the early-stage aggregation of three prion protein peptides, corresponding to residues 120 to 144 of human (Hu), bank vole (BV) and Syrian hamster (SHa) prion protein, from disordered monomers to β-sheet-rich fibrillar structures. Using discontinuous molecular dynamics simulation combined with the PRIME20 force field, we find that the Hu, BV and Sha PrP120-144 aggregate via multiple nucleation-dependent pathways to form U-shaped, S-shaped and Ω-shaped protofilaments. The S-shaped HuPrP120-144 protofilament is similar to the amyloid core structure of HuPrP112-141 predicted by Zweckstetter. HuPrP120-144 has a shorter aggregation lag phase than BVPrP120-144, followed by SHaPrP120-144, consistent with experimental findings. Two amino acid substitutions I138M and I139M retard the formation of parallel in-register β-sheet dimers during the nucleation stage by increasing sidechain-sidechain association and reducing sidechain interaction specificity. On average, HuPrP120-144 aggregates contain more parallel β- sheet content than those formed by BV and SHa PrP120-144. Deletion of C-terminal residues 138

17

to 144 prevents formation of fibrillar structures in agreement with experiment. This work sheds light on the amyloid core structures underlying prion strains and how I138M, I139M and S143N affect prion protein aggregation kinetics.

18

2.1 Introduction

Prion Prion diseases are a family of infectious amyloid diseases that affect both humans and animals. They include Creutzfeldt-Jakob disease (CJD), Gerstmann-Sträussler-Scheinker

Syndrome (GSS), kuru and fatal familial insomnia (FFI) in humans, and scrapie, bovine spongiform encephalopathy (BSE) and chronic wasting disease (CWD) in other mammals. (1) The infectious agent is the scrapie form of the prion protein (PrPSc) which is a β-sheet-rich aggregated form of normal prion protein (PrPC) whose full structure is unknown. In contrast to viruses or other pathogens, which propagate by replicating themselves inside a host cell based on their nucleic acid genome, PrPSc propagates by templating the misfolding and accumulation of PrPC into misfolded

PrPSc. (2) In addition, within a single population, e.g. human, various strains of CJD are believed to be caused by the same prion protein but with diverse PrPSc conformations. (3) The molecular mechanism underlying PrPSc-assisted propagation still remains unclear. (4)

Various models for the full structure of PrPSc have been postulated based on experiment and simulation studies. PrPSc structures have been found to contain a 27-30 kDa protease-resistant core consisting of residues 90 to 231. (5) Govaerts et al. (6) postulated a trimeric model for PrPSc in which the N-terminal, residues 89-175, adopts a left-handed β-helix and residues 176-227 in the

C-terminal maintain an α-helical conformation. Using molecular dynamics simulation, DeMarco et al. (7) observed the conversion of native two-stranded β-sheets (residues 129-131 and 161-163) in the Syrian Hamster prion protein N terminus into three-stranded β-sheets (residues 116-119,

129-132, 160-164) and isolated β-strands (residues 135-140). Based on this, they proposed a spiral

β-helix filament model containing PrPSc-like trimers tightly packed along the fibril axis. Smirnovas et al. (8) applied hydrogen-deuterium exchange coupled with mass spectrometry to brain-derived

PrPSc and found that the proteinase K-resistant domain (from residues 90-231) of PrPSc consists

19

mainly of β-strands with relatively short connecting turns and loops and no α-helices, contradicting the notion that PrPSc retains native α-helices. Based on solid-state NMR data, Groveman et al. (9) proposed a parallel in-register architecture for the mouse PrP90-231 fibril. Their MD simulation results show that residues 124-227 adopt a more packed structure than residues 90-123.

The N-terminal prion protein fragments have been the subject of experiments aimed at learning the fundamental mechanisms underlying prion protein propagation. Examples of prion fragments that have been studied include the prion Y145Stop (PrP23-144) and Q160Stop mutants

(PrP23-159) which are associated with prion cerebral amyloid angiopathy (PrP-CAA), a GSS-like prion disease characterized by the presence of amyloid deposits in the cerebral nervous system.

(10) Vanik et al. (11) found that recombinant PrP23-144 from human, mouse and hamster form fibrils with distinct lag phases. They concluded that species-specific mutations (I138M and I139M) are critical in determining the aggregation kinetics as well as the cross-seeding specificity of

PrP23-144. Jones et al. (12) found that residues 130-140 are an essential region for amyloid formation because deletion of the conserved palindrome sequence 113AGAAAAGA120 results in an altered amyloid β-core but doesn’t affect amyloidogenicity or seeding specificity. Later on,

Chatterjee et al. (13) carried out seed-induced fibrillation experiments on recombinant full-length mouse PrP23-230 and found that residues 127-143 participate in the formation of the amyloid core.

They also found that synthetic MoPrP107-143 and MoPrP127-143 peptides can seed the fibrillation of recombinant full-length mouse prion protein while MoPrP107-126 can’t.

The fibrillation of short N-terminal prion peptides from the amyloid core region of N- terminal prion protein fragment has also been investigated in experiments. Tagliavini et al. (14) performed the fibrillation experiments of synthetic peptides homologous to prion protein residues

106-147 and found that PrP106-126 formed straight fibrils while PrP127-147 formed twisted

20

fibrils. Skora et al. (15) applied H/D exchange coupled with NMR spectroscopy to fibrils formed by recombinant human PrP108-143 and found that residue 129 is deeply buried in the amyloid core. They also found that mutations at position 129 greatly impact aggregation kinetics. Li et al.

(16) used vibrational spectroscopy to show that PrP118-135 starts to adopt a parallel β-sheet structure in a lipid bilayer at high peptide concentrations but prefers α-helical structure at low peptide concentrations. Apostol et al. (17) crystalized prion fragments corresponding to human

(Hu) PrP138-143, mouse (Mo) PrP137-142 and hamster (Ha) PrP138-143 and found that

HuPrP138-143 and MoPrP137-142 form parallel in-register steric-zippers while Ha138-143 forms antiparallel out-of-register steric-zippers. Chuang et al. (18) investigated how various amino acid substitutions in synthetic bovine PrP108-144 affect peptide amyloidogenesis and seeding efficiency. They found that the mutation L138M promotes amyloid formation and increases seeding efficiency, while the mutations I139M and N143S retard amyloid formation and decrease

PrP108-144 seeding efficiency.

There are many structural models proposed for short N-terminal prion peptides. Kuwata et al. (19) conducted MD simulations and confirmed the structural stability of tightly-packed parallel

β-sheets in the MoPrP106-126 octameric fibril. Helmus et al. (20) conducted solid state NMR on the human PrP23-144 amyloid structure and found that the fibril contains a parallel in-register β- sheet core comprising residues 112-141 near the peptide C-terminal. The remaining residues in human PrP23-144 amyloid fibrils were found to be flexible and largely disordered. Later on, based on solid-state NMR data, Lin et al. (21) presented a parallel in-register β-sheet model for the human

PrP127-147 fibril where each peptide has two straight β-strands connected by a kink at 137Pro.

Skora et al. (22) applied the ROSETTA method and predicted the conformation of human PrP112-141 within the amyloid core formed by the Y145stop prion mutant. Later on,

21

Zweckstetter (23) used a combination of NMR spectroscopy, hydroxyl radical probing combined with mass-spectrometry, electron microscopy and computational methods to propose dimeric and trimeric left-handed β-helix models for the PrP112-141 amyloid core.

In this work, we apply a combination of discontinuous molecular dynamics and the

PRIME20 force field to investigate the spontaneous aggregation of prion protein peptides PrP120-

144, an essential region for PrP23-144 amyloid formation. Although the fibrillation of prion protein and its N-terminal fragments has been widely studied in experiments, this is, to our knowledge, the first simulation work to investigate the aggregation mechanism for this particular peptide. We investigate three different PrP120-144s which originate from human (Hu), bank vole

(BV) and Syrian hamster (SHa). The spontaneous aggregation of a system of 8 PrP120-144 peptides is monitored, starting from a random configuration and eventually forming a mixture of oligomers and protofilaments. The structures of oligomers and protofilaments formed by Hu, BV and SHa PrP120-144 at the end of 37 independent simulations are analyzed. The aggregation kinetics of the three PrP120-144 peptides are compared. We also investigate how the amino acid substitutions I138M, I139M and S143N affect PrP120-144 aggregation. This work provides molecular level insight into the amyloid core structure and the aggregation mechanism for PrP120-

144 peptides.

Highlights of our simulation results are the following. PrP120-144 forms S-shaped, U- shaped and Ω-shaped protofilament structures via multiple nucleation-dependent pathways. The

PrP120-144 protofilaments mainly contain parallel in-register β-sheets which are consistent with previous experimental work. (20,23) The key event during the aggregation is the formation a nucleus consisting of a parallel in-register β-sheet dimer. In most cases, the nucleus first grows into a metastable oligomer, then undergoes backbone rearrangement and eventually grows into a

22

protofilament. In rare cases, the nucleus grows directly into a protofilament without intermediate steps. Hu, BV and SHaPrP120-144 all aggregate in a nucleation-dependent manner. In terms of aggregation kinetics, HuPrP120-144 has a shorter lag phase than BVPrP120-144, followed by

SHaPrP120-144, consistent with experimental findings. (11) The HuPrP120-144 aggregates also have more parallel in-register β-sheet content than BV and SHa PrP120-144. This is because the two species-specific amino acid substitutions I138M and I139M enhance sidechain-sidechain association and reduce sidechain-sidechain interaction specificity, thus decreasing the probability of forming parallel in-register β-sheets. S143N doesn’t have any effect on PrP120-144 aggregation kinetics because it does not engender significant change in the sidechain-sidechain interactions.

2.2 RESULTS

2.2.1 Hu, BV and SHaPrP120-144 aggregate into protofilament structures

We investigate the spontaneous fibrillation of prion protein fragments (PrP120-144) by performing DMD/PRIME20 simulations on a system containing 8 peptides. Eleven to fifteen independent simulations were conducted on each of the Hu, BV and SHaPrP120-144 peptides for a total of 37 simulations at T*=0.195 and peptide concentration of 20 mM. Fig. 1 show the total interaction energy versus reduced time from 15 runs of HuP120-144. Starting from a high energetic state, the system goes through various pathways and reaches its relatively low energy state by the end of each simulation. The protofilament structures formed for HuPrP120-144 in the 3rd, 5th,7th

,11th and 12th runs have the lowest interaction energies and hence the highest thermodynamic stabilities among all the final structures. The system total energy profiles for BV and SHa PrP120-

144 are shown in Fig. S1 and S2.

23

Fig. 2 shows the final structures for three different protofilaments formed by HuPrP120-

144 in our simulations. Fig. 2A shows an S-shaped parallel in-register protofilament formed by

HuPrP120-144 from the 5th run; Fig. 2B shows a schematic side view and the sidechain placement for this structure. Such S-shaped protofilaments are also formed in the 2nd, 3rd and 7th runs. From

Fig. 2B, we see that there are 4 turning points located at 124Gly, 126Gly, 137Pro and 142Gly along the backbone chain of each peptide in the protofilament. Two β-strands (residues 127-136 and 138-141) flank residues 136Arg and 137Pro, which is similar to the amyloid core structure suggested by Skora (22) and Zweckstetter (23) for HuPrP112-141. Having a proline at position

137 may explain why HuPrP120-144 adopts a “boomerang” structure, since proline facilitates turns and breaks in α-helices and β-sheets. (24) In contrast, in Zweckstetter’s PrP112-141 model

(23), the two β-strands lie on a nearly straight line. Jones et al. (12) probed the fibril structure of both wild type PrP23-144 and PrP23-144 without the nonessential palindromic sequence (residues

113-120) by solid-state NMR spectroscopy and found that deletion of residues 113-120 changes the secondary structure distribution of the fibril core region. Since our PrP120-144 simulation does not include residues 113-120, the structural differences between our PrP120-144 protofilament and the Zweckstetter’s PrP112-141 amyloid core structure (23) are not surprising.

Fig. 2C shows the Ω-shaped protofilament formed in the 12th run and Fig. 2D shows a side view. It is apparent from Fig. 2D that the Ω-shaped protofilament has 5 turning points at 124Gly,

126Gly, 131Gly, 137Pro and 142Gly. Note that the C-terminal of the yellow chain is not the part of the protofilament. The Ω-shaped protofilament has two triangular-shaped hydrophobic cores formed by the association of the hydrophobic sidechain beads in white. The first hydrophobic core contains 121Val, 125Leu and 130Leu. The second hydrophobic core contains 129Met, 134Met and 138Ile, which is similar to the triangular hydrophobic core of Het-s (218-289) fibrils. (25)

24

Fig. 2E shows the U-shaped protofilament formed by HuPrP120-144 in the 11th run and

Fig. 2F shows its side view. Compared to the Ω-shaped protofilament, the N-terminal residues of the U-shaped protofilament curve inwards to form a large hydrophobic sidechain core which is essentially a combination of the two separate hydrophobic cores formed in the S-shaped protofilament. Cheon et al. (26) found that Aβ17-42 form a similar U-shaped protofilament.

BVPrP120-144 can form both S-shaped and Ω-shaped protofilaments, as shown in Fig. 3.

Regardless of the amino acid difference between BV and Hu PrP120-144, the peptides in the

BVPrP120-144 S-shaped protofilament in Fig. 3A adopt nearly the same conformation as those in

HuPrP120-144 S-shaped protofilament in Fig. 2A. From Fig. 3C, the Ω-shaped protofilament formed by BVPrP120-144 has 121Val, 125Leu and 130Leu in the first hydrophobic core and

129Met, 132Ser, 134Met and 138Met in the second hydrophobic core.

SHaPrP120-144 can also form S-shaped and Ω-shaped protofilaments, as shown in Fig. 4.

The S-shaped protofilament formed by SHaPrP120-144 has a similar conformation to the

HuPrP120-144 S-shaped protofilament shown in Fig. 2A. The Ω-shaped protofilament formed by

ShaPrP120-144 has 121Val, 125Leu, 129Met in the first hydrophobic core and 130Leu, 132Ser,

135Ser and 138Met in the second hydrophobic core.

Our simulations reveal that the Hu, BV and SHa PrP120-144 peptides aggregate to form relatively heterogeneous aggregate structures. We believe that the main reason that PrP120-144 forms the S, Ω and U-shaped protofilaments is its conformational flexibility. Each of three different PrP120-144s contain six glycines and one proline which facilitate turn structures in a protein. For HuPrP120-144, the S, Ω and U-shaped protofilaments are formed in 4, 1 and 1 out of

15 runs. For BVPrP120-144, the S and Ω-shaped protofilaments are formed in 2 and 1 out of 11 runs. For SHaPrP120-144, the S and Ω-shaped protofilaments are formed in 2 and 1 out of 11 runs.

25

These S-shaped, Ω-shaped and U-shaped protofilaments are stable enough so that they maintain their conformation throughout the simulation. Aside from forming parallel β-sheets, in 9 of the 15 runs, 8 out of the 11 runs and 8 out of the 11 runs, Hu, BV, SHa PrP120-144 form metastable oligomers which are characterized by a mixture of parallel and anti-parallel β-strands, as shown in

Figs. S3-5, respectively.

Notably, all of the peptides considered, regardless of their amino acid differences, can form both S-shaped and Ω-shaped protofilaments. Comparing the structure of the three S-shaped protofilaments formed by three PrP120-144s in Figs. 2B, 3B and 4B, we find that peptides all have the same backbone alignment and the same sidechain placement. However, the peptides in the Ω- shaped protofilament adopt distinct sidechain placements as shown in Figs. 2D, 3D and 4D. The major structural difference between the three Ω-shaped protofilaments is that the two pairs of adjacent residues, 129Met and 130Leu and 134Met and 135Ser switch their position between the two hydrophobic pockets.

2.2.2 Structural analysis of Hu, BV, SHaPrP120-144 protofilament

The distribution of secondary structure in the HuPrP120-144 S-shaped protofilament was calculated using the STRIDE program (27) in the visual molecular dynamics (VMD) and is shown in Fig. 5. (28) The hydrophobic and polar residues, including 121Val to 135Ser and 138Ile to

141Phe, have at least 70% β-sheet structure except for residues near the two termini. The glycine and proline are most likely to be in random coil or turn structures. The secondary structure distributions for the non-S-shaped protofilaments formed by Hu, BV and SHa PrP120-144 are similar to those shown in Fig. 5 and are thus not displayed here.

26

Table 1 shows the equilibrium values for the energies of the different-shaped Hu, BV and SHa

PrP120-144 protofilaments in Figs. 2-4. The backbone hydrogen bonding energy, EHB, sidechain- sidechain interaction energy, ESidechain and their sum, total interaction energy, ETotal are listed. All of the energies are calculated by averaging over the last 100 billion collisions.

A good way to understand the conformational difference between S-shaped, Ω-shaped and

U-shaped protofilaments is to compare the equilibrium values for Esidechain within each species.

From Table 1, we find that for the protofilaments formed by HuPrP120-144, Esidechain(U-shaped) ≤

Esidechain(Ω-shaped) < Esidechain(S-shaped), indicating that the U-shaped protofilament has a stronger sidechain-sidechain association than the Ω-shaped protofilament, followed by the S-shaped protofilament. The U-shaped protofilament contains a large hydrophobic core having 7 sidechain spheres per peptide contacting each other (Fig. 2F) while the Ω-shaped protofilament contains two separate small hydrophobic core (Fig. 2D) and the S-shaped protofilament contains only one small hydrophobic core near the N-terminal and an open sidechain pocket near the C-terminal, (Fig. 2B).

In addition, the three different protofilaments formed by HuPrP120-144 have similar EHB because they all consist of parallel in-register β-sheet. Similar correlations between the side-chain energies and protofilament structure can be applied to those formed by BV and SHa PrP120-144.

2.2.3 Nucleation dependent aggregation of Hu, BV and SHaPrP120-144

Next we investigate the protein aggregation pathway and hence the aggregation mechanism for the three prion protein fragments. Due to the large number of degrees of freedom that a system of aggregating proteins can have, the free energy landscape is very rugged, making it difficult for researchers to elucidate information for the aggregation pathway. In this work, we trace out the important intermediate steps along the PrP120-144 aggregation pathway by applying hierarchical

27

clustering analysis. The data are analyzed by using the JMP® Pro 12.0.1 software (SAS Institute

Inc.). The three collective variables (CVs) are chosen to be the number of hydrogen bonds (NHB), the sidechain-sidechain interaction energy (Esidechain), and the end-to-end distance (Eendtoend). The clustering analysis is based on the Ward’s method (29) which proceeds in the following way. Each data point along the trajectory (which consists of 3 CVs) starts out as a single cluster. There are initially N clusters. The next step is to combine two of the clusters into one so that the total number of clusters is reduced to N-1. The two nearest clusters are selected from N(N-1)/2 possible pairs

ퟐ 풏푨풏푩 풏풌 ̅̅푨̅ ̅̅̅푩̅ of clusters. The distance of two clusters, 푫푨푩 = ∑풌=ퟏ(풙풌 − 풙풌 ) , where nA, nB are total 풏푨+풏푩

̅̅푨̅ ̅̅̅푩̅ number of elements in cluster A and B, nk = 3 is the total number of CVs in this case, and 풙풌, 풙풌 is the average value for the kth CV over all elements in cluster A and B. By iteratively merging two nearest clusters, the N clusters are eventually grouped into 3-4 major clusters that are termed the lag phase, nucleus growth, oligomer rearrangement and fibril formation. If the trajectory is grouped into 3 major clusters, we regard it as “one-step” fibrillation. If the trajectory is grouped into 4 major clusters, we regard it as “two-step” fibrillation. The results of the analysis for the 3rd

(“one-step” fibrillation) and 5th (“two-step” fibrillation) runs on HuPrP120-144 are discussed below.

We find from our simulation that the rate limiting step during the aggregation of Hu, BV and SHaPrP120-144 peptides is nucleus formation. We define nucleus formation to be the first time that a small oligomer containing a parallel in-register β-sheet dimer forms; this is the initial step in the development of a new aggregated phase. Figs. 6A-6E show simulation snapshots of

HuPrP120-144 aggregation from the 3rd run. Fig. 6F shows the total interaction energy versus simulation time for this run. Peptides are initially placed in a random configuration. After the temperature is cooled to T*=0.195, the peptides start to contact each other and aggregate. The

28

aggregation process for the 3rd run can be divided into three major stages: the nucleation stage, the nucleus growth stage and the equilibration stage as shown in Fig. 6F. During the nucleation stage, peptides keep attaching and detaching from each other through sidechain contacts or hydrogen bonding interaction. In Fig. 6B, peptides form a transient oligomer with peptides partially associated with each other to form hydrogen bonds. In Fig. 6C, peptides dissociate from and reversibly return to the disordered state due to the formation of transient antiparallel steric zippers.

The key event during HuPrP120-144 aggregation is the formation of a nucleus which contains a parallel in-register β-sheet dimer, as shown in Fig. 6D. We identify the time window where the first dimer is formed by keeping track of the number of parallel in-register β-sheet segments

(defined in Fig. 10) between any of two peptides in the system to see when the “18 parallel βsheets segments” criteria is satisfied. The nucleus rapidly grows into an S-shaped protofilament (Fig. 6E) by recruiting other peptides to its two ends. Fig. 6F shows that the system total interaction energy decreases sharply after the nucleus forms; the system here is stabilized by a large number of hydrogen bonds and maximal sidechain contacts. Baftizadeh et al. (30) had a similar observation in their biased-exchange meta-dynamics simulation where the rate limiting step during the aggregation of 18 Aβ35-40 peptides was associated with the formation of a specific type of steric zipper.

We find in our simulations that PrP120-144 goes through multiple aggregation pathways.

Figs. 7A-E show simulation snapshots of the 5th run for HuPrP120-144 and Fig. 7F shows the total interaction energy as a function of time. In this run, the eight HuPrP120-144s aggregate to form an S-shaped protofilament via a “two-step” fibrillation pathway (Fig. 7F). Compared with the

“one-step” fibrillation pathway described previously (Fig. 6F), the unstable nucleus on the “two- step” pathway requires a long time to overcome energy barriers, rearrange into a stable

29

conformation, and finally grow into a protofilament. This is reminiscent of the Lumry-Eyring nucleated polymerization (LENP) model for protein aggregation proposed by Andrews et al. (31)

In the LENP model, peptides first reversibly associate to form a pre-nucleus which eventually goes through a conformational rearrangement step to form an irreversible nucleus. In Fig. 7A, peptides are initially disordered. In Fig. 7B, a nucleus containing a β-sheet dimer (colored cyan and green) is formed. It quickly grows into a metastable oligomer (Fig. 7C) consisting of 5 peptides, the N terminal of which has a tilted β-roll conformation. Disordered peptides repeatedly attach and detach from the transient oligomer. The peptides within the oligomer rearrange into a stable parallel in-register β-sheet conformation (Fig. 7D) which eventually elongates and grows into an

S-shaped protofilament (Fig. 7E) by recruiting and templating the remaining disordered peptides at its two ends.

Although the “two-step” fibrillation pathway of HuPrP120-144 is similar to the LENP model, there are some differences. In the LENP model, the protein reversibly switches its conformation between its native state, an intermediate state and an unfolded state prior to nucleus formation. However, in our simulation, there are no native, intermediate or unfolded states because

PrP120-144 is too short to form any native domain, much less a domain that is characteristic of the whole prion protein. Also, residues 120-144 within the normal prion protein are intrinsically disordered, except that residues 128-131 form a short antiparallel β-sheet with residue 161-164 which is not included in our studies.

The PrP120-144 peptides in all 37 independent simulations aggregate in a nucleation- dependent manner, as evidenced by the fact that they all go through a lag phase before aggregating into β-sheet-rich structures. The “one-step” and “two-step” fibrillation pathways are applicable only for those simulation runs that result in the formation of parallel in-register β-sheet

30

protofilaments as shown in Figs. 2-4. For HuPrP120-144, aside from the 3rd run, the 2nd and 7th runs also follow the “one step” fibrillation pathway to form an S-shaped protofilament since the nucleus grows directly into a final fibrillar state without passing through an obvious metastable state. That is, there is no intermediate plateau between the lag phase and fibrillar phase along the total energy profile, as shown in Fig. 1. The formation of the U-shaped and Ω-shaped protofilaments in 11th and 12th runs also follows a “one-step” fibrillation pathway for the same reason. For BVPrP120-144, the 4th run follows a “one-step” and the 3rd, 5th runs follow a “two- step” fibrillation pathway. For SHaPrP120-144, the 10th run follows a “one-step” and the 1st and

7th follows a “two-step” fibrillation pathway. However, neither the one- or two-step fibrillation pathways are applicable for the trajectories where a metastable oligomer, e.g. β-barrel oligomer, is formed such as those in Figs. S3-5. In those cases we can’t predict whether the oligomers are off-pathway products or on-pathway toward forming protofilaments without using much more computing resources as well as applying biased-sampling techniques.

2.2.4 I138M and I139M retard nucleus formation

We then investigate how amino acid substitutions (I138M, I139M and S143N) between

Hu, BV and SHa PrP120-144 affect their aggregation kinetics with particular focus on the lag phase time. The lag phase time in our simulations is defined to be the time from the start of the simulation to the first time that two peptides form a stable parallel in-register β-sheet dimer. Our criterion for stability here is that the dimer should have at least 18 parallel β-sheet segments, as defined in Fig. 10. The rationale for using this definition is that the formation of a stable parallel

β-sheet dimer leads to an obvious increase in β-sheet content of the system and so is easy to detect.

If two peptides share less than 18 parallel β-sheet segments, they are metastable and may dissociate

31

into monomer state. The lag phase time for each species is averaged over 11 and 15 independent runs (15 runs for HuPrP120-144, 11 runs for both BV and SHa PrP120-144). Table 2 shows a comparison of the aggregation lag phase time for the three PrP120-144s. HuPrP120-144 has a shorter average lag phase time than BVPrP120-144, followed by SHaPrP120-144. The relatively large error in the lag phase time is a consequence of having a random initial configuration which may go down numerous nucleation pathways. From a kinetics point of view, the relative values of the lag time for the three peptides can be interpreted as evidence that HuPrP120-144 has a smoother free energy landscape for fibrillation than BVPrP120-144, followed by SHaPrP120-144.

Our findings that I138M and I139M increase the aggregation lag phase (HuPrP120-144 <

BVPrP120-144 < SHaPrP120-144) are consistent with Vanik et al. (11)’s fibrillation experiments on a longer prion protein fragment HuPrP23-144 and its I138M, I138M&I139M mutants that the

I138M and I138M&I139M mutants have much longer fibrillation lag phase than that of the native

HuPrP23-144. In addition, Chuang et al. (18) found in experiments that the I139M mutant of bovine PrP108-144 I139M has lag phase that are 4-fold times longer than those of bovine PrP108-

144.”

We find that the two amino acid substitutions I138M and I139M are responsible for retarding BV and SHa PrP120-144 dimer formation compared to HuPrP120-144 and thus retarding the aggregation kinetics. Fig. 8 describes the average sidechain interaction energy and average number of sidechain contacts experienced by residues 138, 139 and 143 in the Hu, BV and Sha

PrP120-144 peptides during the nucleation stage. From Fig. 8A and B, we see that the 138Mets on

BV and SHa PrP120-144 have the same number of side-chain contacts as the 138Ile on HuPrP120-

144, and the 139Met on SHaPrP120-144 has the same number of sidechain contacts as the 139Ile on HuPrP120-144, but the 138Mets on BV and SHa PrP120-144 have stronger sidechain

32

interaction energies than the 138Ile on HuPrP120-144, and also the 139Met on SHaPrP120-144 has a stronger sidechain energy than the 139Ile on HuPrP120-144. Thus the Mets on BV and SHa

PrP120-144 have a stronger sidechain association with the surrounding residues than the Iles on

HuPrP120-144 during the simulation. The stronger the attractive interaction that the C-terminal residues (138 and 139) experience with the surrounding residues, the more difficult it is for two peptides to form a perfect in-register β-sheet dimer. As shown in Table S1, which lists the energy parameter matrix between 19 amino acid sidechains in the PRIME20 force field, Ile has a square well attraction only with hydrophobic residues, e.g. Ala, Leu, Ile, Phe, while Met has a strong square well attraction not only with those hydrophobic residues but also with polar and charged residues, e.g. His, Arg, Ser and Asn. In other words, Table S1 indicates that Met has a broader set of attractive interaction partners than Ile. One reason is that the divalent sulfur atom on the Met sidechain can form hydrogen bonds with residues that have hydrogen donors in their sidechains.

(32) A calculation by Berka et al. (33) using the RI-DFT-D method with the COSMO model also showed that Met has a stronger sidechain interaction energy with Tyr, Arg, Ser and Asn than Ile does in a water-like environment. Our finding that I138M and I139M retard PrP120-144 aggregation is also consistent with predictions of the amyloid aggregation propensity of individual amino acids by Pawar et al. (34), showing that Met has a lower β-sheet-forming propensity than

Ile.

In contrast to the I138M or I139M, the amino acid substitution S143N does not retard the aggregation kinetics. From Fig. 8A, 143Ser (for HuPrP120-144) and 143Asn (for BV and SHa

PrP120-144) have nearly the same average sidechain interaction energy which means that they have similar degrees of sidechain association. In fact, Table S1 shows that Asn and Ser have the same sidechain energetic parameters in the PRIME20 force field. They have also been shown to

33

have similar β-sheet forming propensity. (34,35) From the above analysis of simulation data, we conclude that S143N doesn’t affect the aggregation kinetics of PrP120-144. Chuang et al. (18), however, had a different finding on how Asn-Ser mutation affects PrP108-144 aggregation kinetics; they found that the single amino acid substitution N143S can retard bovine PrP108-144 aggregation.

2.2.5 I138M and I139M retard parallel β-sheet formation

During the aggregation process, not only do I138M and I139M retard nucleus formation but they also retard the attachment of disordered peptides to the seed, thus decreasing the parallel

β-sheet content of the final aggregation configuration. Table 3 shows the average percentage of parallel β-sheet and anti-parallel β-sheet content in the oligomers and protofilaments obtained at the end of all simulations for Hu, BV and SHa PrP120-144. From Table 3, HuPrP120-144 on average has more parallel β-sheet content than BV and SHa PrP120-144, indicating HuPrP120-

144 has stronger propensity to form parallel β-sheets than BV and Sha PrP120-144. In addition, experiments by Apostol et al. (17) show that HuPrP138-143 and MoPrP138-143 (MoPrP has the same sequence as BVPrP for residues 138-143) both form parallel steric zippers while SHaPrP138-

143 preferentially forms anti-parallel out-of-register steric zippers. This effect is not obvious from our simulation.

2.3.6 Residues 138-144 are critical for PrP120-144 fibril formation

The C-terminal residues 138-144 are found to be necessary for PrP120-144 to form fibrils in our simulations. Deleting the C-terminal resides 138IIHFGSD144 results in a non-amyloidogenic peptide PrP120-137. The data is shown in Fig. 9 which describes the number of hydrogen bonds

34

formed during HuPrP120-144 (the 3rd run) and PrP120-137 aggregation simulations as a function of time. From Fig. 9, PrP120-144 forms β-sheet-rich aggregates while PrP120-137 fails to form a nucleus and thus maintains the disordered state throughout the simulations, indicating PrP120-144 is much more amyloidogenic than PrP120-137. One possible reason is that without the C-terminal residues 138 to 144, the β-sheet dimer of PrP120-137 becomes thermodynamically unstable, thus making it unable to form ordered aggregates. Our results are consistent with Kundu et al. (36) ’s experiments which show that HuPrP23-144 forms fibril but without the C-terminal residues 138-

144, HuPrP23-137 can’t form fibrils.

2.3 DISCUSSION

Using DMD simulation combined with PRIME20 force field, we investigated the early stage aggregation mechanism of amyloidogenic prion protein fragments PrP120-144. In this work, by simulating the aggregation of prion protein fragments PrP120-144 from three species (human, bank vole and Syrian hamster), we provide molecular-level insight into how the sidechain substitutions, I138M and I139M, affect the aggregation kinetics of PrP120-144 peptide. The parallel in-register protofilament formed by stacking of PrP120-144 peptides shed lights on the structure of the amyloid core of the PrP23-144 fibril which is associated with familial PrP-CAA.

(20,37).

Our simulation results on PrP120-144 are consistent with a number of experimental studies on PrP23-144 and it core region (residues 112 to 144). (No experiments have been conducted on

PrP120-144 so we cannot compare our results to experiments directly.) Firstly, we find that two species-specific amino acid substitutions (I138M and I139M) retard nucleus formation and the aggregation kinetics of N-terminal prion protein fragments. This is consistent with Vanik et al.’s

35

experimental findings that I138M and I139M greatly affect the aggregation kinetics of PrP23-144 and lead to cross-seeding barriers between PrP23-144 from different species. (11) In fact, the steric zipper structures formed by short PrP138-143 fragments have been shown to correlate with the cross-species transmission barrier of prion diseases by the Eisenberg group. (17,38) Secondly, we show that PrP120-144 is unable to form fibrillar structures without the C-terminal residues 138-

144. This is also consistent with Kundu et al.’s experiments that HuPrP23-144 form fibrils but

HuPrP23-137 cannot. (36) Our finding that residues 137-144 play a crucial role in the fibrillation of prion peptides is also consistent with the fact that the prion fragment PrP136-140 in experiment binds to PrPSc and may be responsible for conversion from PrPC to PrPSc. (39) Thirdly, we show that HuPrP120-144 aggregates to form β-sheet-rich protofilaments with a parallel in-register peptide alignment. This is also consistent with both Helmus et al’ s (20) and Zweckstetter’s (23) structural studies on the prion amyloid core PrP112-141 region based on NMR and other experimental data.

One of the main focus of this work is on learning the aggregation mechanisms that drive disordered peptides PrP120-144 to form fibrillar structures. We found that PrP120-144 aggregates in a nucleation-dependent manner but along multiple pathways. Several studies have found multiple aggregation pathways for amyloidogenic proteins. Natalello et al. (40) used a combination of Fourier transform infrared spectroscopy and atomic force microscopy to find that human PrP82-

146 undergoes multiple aggregation pathways to form fibrils with various types of secondary structure. Also, Cheon et al. (26) found in DMD/PRIME20 simulations that Aβ17-42 assembles from an oligomer state to a U-shaped protofilament via two distinct pathways. Unlike the nucleated polymerization that short amyloidogenic peptides like Aβ16-22 undergo, PrP120-144 seems to go

36

through heterogeneous oligomeric states before forming the nucleus because of numerous local energy minima resulting from the folding effects of peptide backbones.

The 8-peptide-system that we have simulated is large enough to allow investigation of the nucleation stage of prion protein fragments since generally only a few monomers are involved in the early stage of protein aggregation. (31) For example, Cho et al. (41) found in experiment that the first step along the aggregation of recombinant full-length prion protein is the coming together of 3 to 4 monomers to form a seed. Note that our definition of nucleation is different from the description in classical nucleation theory (42), where nucleus formation is characterized by overcoming a free energy barrier to form a new thermodynamic phase. Since our system size is much smaller than real systems, we are unable to quantitatively compare the aggregation kinetics in our simulation with those in experiments since the macroscopic fibrillation rate is an ensemble average value over all possible kinetic pathways. Instead, the kinetics-related quantities that we can calculate, e.g. the average lag phase time, are simple measurements that only allow for qualitative comparison with experiments.

The conditions chosen for our simulation, e.g. temperature and peptide concentration, have a significant effect on the protein aggregation kinetics. Our choice of high simulation temperature and high peptide concentration allows us to explore the aggregation mechanism of PrP120-144 within a reasonable timescale. The major advantage of simulating at a high temperature, just below the critical fibrillation temperature, is that the high entropic fluctuations in this state smooth the peptide aggregation free energy landscape and decrease the probability that the system will be trapped in metastable states. In addition, we use a relative high peptide concentration (20mM) compared to experimental conditions because the barrier to aggregation kinetics is greatly lowered

37

at supersaturated peptide concentration such as this, and is believed to be the driving force for protein aggregation. (43)

The 25-residue-length peptide simulated in this work is the longest prion protein fragment that we have simulated so far due to limitations in computational resources. The longer the peptide sequence is, the more rugged the aggregation free energy landscape is, thus requiring more computation resources and advanced sampling techniques to sample the whole ensemble of system. In future work we hope to simulate the aggregation of a longer prion peptide, PrP112-144, since it is thought to comprise the whole prion amyloid core. (44)

2.4 METHODS

2.4.1 Discontinuous molecular dynamics (DMD)

The main simulation method employed in this work is discontinuous molecular dynamics, a fast alternative to traditional molecular dynamics. (45) In DMD simulation, the potential energy between two particles is modeled as a discontinuous potential, such as the hard sphere, square well and square well shoulder potentials. In DMD, the velocities and positions of two beads (particles or sites on molecules) at each time step are recalculated only when a discontinuity in their interaction potential is encountered. The types of events considered include core collision, bond extension, well capture, dissociation and bounce. This is different from traditional molecular dynamics where the position and velocity of every particle in the system need to be recalculated at very small regularly-spaced time steps by solving Newton’s equation of motion.

38

2.4.2 PRIME20 model

PRIME20 is a coarse-grained model developed in the Hall group that was specifically designed for use in DMD simulations of protein aggregation. (46) Each amino acid is represented by three backbone spheres (N-H, Cα-H, C=O) and one sidechain sphere (R). PRIME, the predecessor of PRIME20, was used to simulate the spontaneous formation of fibrils containing polyalanine and polyglutamine. (47,48) PRIME was extended in 2010 to become PRIME20, a model that describes 20 different amino acids uniquely and considers sidechain-sidechain polar, charge-charge and hydrophobic interactions. (49) The geometric and energetic parameters for the

20 different sidechain beads were obtained by applying a perceptron learning algorithm and a modified stochastic learning algorithm to optimize the energy discrepancy between 711 native states protein and 2 million decoy conformations produced by gapless threading techniques. In

PRIME20, the known backbone phi-psi angles, Cα-Cα distances and L-isomerization constraints on each amino acid are modeled by imposing pseudo-bonds between the relevant beads. A non- biased distance-based criterion is used to judge when directional hydrogen bonds are formed and broken. (50) Glycine doesn’t have a sidechain bead and proline cannot form backbone hydrogen bonds with other residues. The total interaction energy consists of the sidechain-sidechain interaction energy, Esidechain, and the backbone-backbone hydrogen bonding energy, EHB. All the other non-bonded interactions (including backbone-backbone and backbone-sidechain interactions) are modeled as hard sphere interactions which represent excluded volume effects but do not contribute to the system energy. The interactions between two bonded (adjacent) beads include hard sphere repulsion when the distance between two beads reaches the minimum bond length and an infinite square well attraction when the distance reaches the maximum bond length.

As a result, the bond length fluctuates within 2% of its assigned value.

39

PRIME20 has been used in the Hall group to simulate the spontaneous formation of twisted

Aβ16-22 fibrils (51), tau fragment fibrils (50) and U-shaped Aβ17-42 protofilaments (26).

PRIME20’s ability to discriminate the fibrillation propensity of short peptides was tested on the

STVIIE designed sequence by López de la Paz and Serrano. (52,53) It has also been used to study the effect of macromolecular crowding on Aβ16-22 aggregation. (54,55) In this work, we use the same PRIME20 force field parameters as were used by Cheon et al. (51) in their simulations of the aggregation of Aβ16-22 peptides.

2.4.3 Definition of parallel and anti-parallel β-sheet segments

Two 3-residue peptide strands (labelled i, i+1, i+2 and j, j+1, j+2) form a parallel β-strand segment if both NHi+1-COj and NHj+2-COi+1 form hydrogen bonds, as shown in Fig. 10A. They form an anti-parallel β-sheet segment if 1) both NHi+1-COj+1 and NHj+1-COi+1 forms hydrogen bonds as shown in Fig. 10B; 2) NHj-COi+2, NHi+2-COj, NHi-COj+2 and NHj+2-COi all form hydrogen bonds, as shown in Fig. 10C. As an example, two 4-residue peptide strands (labelled A and B) could form two consecutive parallel β-sheet segments, with residues 1-3 of peptide A forming the first segment with the same residues on peptide B and residues 2-4 of peptide A forming the second segment with the same residues on peptide B. In our simulation two 25-residue

PrP120-144 peptides form a parallel β-sheet dimer if they have at least 18 parallel β-sheet segments.

2.4.4 Relating the PRIME 20 reduced temperature and time to real temperature and time

We relate our simulation reduced temperature T* to real temperature T by matching the

DMD/PRIME20 simulation predictions of the α-helical content versus temperature profiles of

40

alanine-rich polypeptides to those measured in experiments by Munoz and Serrano. (56)

DMD/PRIME20 simulations of the folding of the polypeptides, Ac-AAQAAAAQAAAAQAAY-

NH2, Ac-AAAAKAAAAKAAAAKA-NH2, Ac-AEAAAKEAAAKEAAAKA-NH2, Ac-

WDAAAKDAAAKDAAAKA-NH2, Ac-EAEKAAKEAEKAAKEAEK-NH2 at reduced temperatures ranging from 0.12 to 0.25 show that peptides adopt mostly α-helical structure at low temperature T*=0.12 undergo a folding transition at about T*=0.175 and become random coils at high temperatures T*=0.25 as shown in Fig. S6A. At each temperature point, we simulate a single peptide in an 87.25×87.25×87.25 Å3 box for at least 20 billion collisions. By comparing the simulation data for the percentage of α-helical hydrogen bonds with the experimental data (Fig.

S6B) from Munoz and Serrano (56), as shown in Fig. S6D we obtain a linear correlation between

T and T* for each of 5 peptides. Then after averaging over the slope (A) and interception (B) value of 5 T versus T* linear correlations functions, we obtain T/K = T* + = 2288.46T*-115.79.

Thus the simulations presented in this paper are conducted at real temperature of 330K.

Justification for conducting our simulations at a temperature that is higher than body temperature are given in the discussion section.

In our DMD/PRIME20 simulation, the system energy is scaled based on the hydrogen bonding energy unit εHB. In order to relate system energy value with realistic unit, we set εHB equal to 12.47 kJ/mol used by Ricchiuto et al. (57) in their DMD simulation of peptide aggregation, which is also close to the optimal backbone hydrogen bonding energy 11.72 kJ/mol used in the

STRIDE algorithm. (27) The εHB value is a simple estimation for the PRIME20 model since in nature the hydrogen bonding strength is considered to have temperature dependence. (58) Note

* ’ that the the reduced temperature in simulation is defined to be T =kBT /εHB, this would suggest that the real temperature from our simulations that corresponds to T* = 0.195 is 273K which is of

41

course unrealistically low. It is for this reason that we chose to use the method described above to relate the reduced temperature in our simulations to realistic temperature.

To convert reduced time to real time, we compare the self-diffusion coefficients (D) of the

Aβ16-22 peptide from DMD/PRIME20 simulations and atomistic simulation. Recall that the diffusion coefficient can be obtained in terms of the mean squared displacement (MSD) using the

Einstein diffusion equation(59), 6ΔtD = <|r(t+Δt)-r(t)|2>, where the angle brackets indicate that the center of mass MSD of the peptide is averaged over all peptides in the system during the simulation time period Δt. In DMD, we simulate 6 Aβ16-22 peptides in a 100×100×100 nm box

* 2 for 1 billion collisions at the reduced temperature T =0.19 and find that DDMD=0.072 nm /Δtreduced, where Δtreduced is the reduced time unit. We then perform a 1ns NPT followed by a 50 ns NVT explicit solvent atomistic simulation of a single Aβ16-22 peptide in a 6×6×6 nm3 water box at

T=298K using GROMACS 5.0.4 package (60)and Amber ffSB99 force field. (61) The N terminal of the peptide is capped with acetyl group and the C terminal is capped with an N-methyl group.

2 2 2 We find that DMD=1.0 nm /ns. By equating DDMD to DMD (1nm /1ns = 0.072nm /Δtreduced), we find that the reduced time unit Δtreduced is equivalent to 0.072 ns.

2.4.5 Description of DMD Simulation Techniques

We simulate three 25-residue-length prion peptides which correspond to residues 120 to

144 from three different species: human (Hu), bank vole (BV) and Syrian hamster (SHa). Note that bank vole PrP has the same sequence as mouse PrP between residues 120 and 144. The amino acid sequences of these three prion protein fragments are shown in Table 4. The primary sequence differences between three peptides are at positions 138, 139 and 143. HuPrP120-144 has Ile at

42

positions 138, 139, and Ser at position 143. Compared to HuPrP120-144, BVPrP120-144 has two mutations, I138M, S143N, and SHaPrP120-144 has three mutations, I138M, I139M, S143N.

The simulations are performed in the canonical ensemble (fixed number of particles, constant volume and temperature). The Andersen thermostat is implemented to maintain the system at a constant temperature. (62) The initial velocity for every bead in the system is generated based on a Maxwell-Boltzmann distribution centered at the desired temperature. The system is initialized by randomly placing eight peptides in a cubic box with box length equal to 87.25 Å corresponding to a relatively high peptide concentration of 20 mM. The peptide concentration is set to a high value to reduce the time required for monomer peptides separated far apart to move close together. The reduced temperature is set to be T*=0.195, which is right below the critical fibrillation temperature TC of the PrP120-144 peptide (no fibril forms when the temperature is above TC). Each independent simulation starts with a random configuration and a high reduced temperature (T*=0.50) to relax each peptide chain into a random coil conformation. The system is then gradually cooled to T*=0.195 and then proceeds for at least 1 trillion collisions (12 μs) which takes 1 thousand CPU hours (forty days) on a fast workstation.

2.5 ACKNOWLEDGMENTS

We appreciate Prof. Stefan Auer, Ilia V. Baskakov, Sheena E. Radford and Andrew J. Wilson for helpful discussion. This work was supported by the National Institutes of Health, USA, under grant

GM056766 and EB006006. It was also supported in part by the NSF’s Research Triangle MRSEC, DMR-

1121107.

Conflict of interest

The authors declare that they have no conflicts of interest with the contents of this article.

43

Author contributions

Y.W. and C.K.H designed the research and wrote the manuscript. Y.M. and Q.S. conducted the research and analyzed the data.

Abbreviations used

DMD, discontinuous molecular dynamics; PrP, prion protein; Hu, human; BV, bank vole; Mo, mouse; SHa, Syrian hamster; CJD, Creutzfeldt-Jakob disease, GSS, Gerstmann-Sträussler-

Scheinker Syndrome; FFI, fatal familial insomnia; BSE, bovine spongiform encephalopathy;

CWD, chronic wasting disease; PrPSc , scrapie form of prion protein; PrPC , cellular form of prion protein; NMR, nuclear magnetic resonance; CAA, cerebral amyloid angiophathy; VMD, visual molecular dynamics; STRIDE, a structural identification algorithm; LENP, Lumry-Eyring nucleated polymerization model; PRIME, protein intermediate resolution model; Aβ, Amyloid β;

I, Isoleucine; M, Methionine; S, Serine; N, Asparagine.

44

2.6 REFERENCES

1. Prusiner, S. B. (1991) Molecular-Biology Of Prion Diseases. Science 252, 1515-1522

2. Walker, L. C., and Jucker, M. (2015) Neurodegenerative Diseases: Expanding the Prion

Concept. Annu. Rev. Neurosci. 38, 87-103

3. Morales, R., Abid, K., and Soto, C. (2007) The prion strain phenomenon: Molecular basis

and unprecedented features. BBA-Mol. Basis Dis. 1772, 681-691

4. Soto, C., Estrada, L., and Castilla, J. (2006) Amyloids, prions and the inherent infectious

nature of misfolded protein aggregates. Trends Biochem. Sci. 31, 150-155

5. McKinley, M. P., Bolton, D. C., and Prusiner, S. B. (1983) A Protease-Resistant Protein Is

A Structural Component Of The Scrapie Prion. Cell 35, 57-62

6. Govaerts, C., Wille, H., Prusiner, S. B., and Cohen, F. E. (2004) Evidence for assembly of

prions with left-handed beta 3-helices into trimers. Proc. Natl. Acad. Sci. USA 101, 8342-

8347

7. DeMarco, M. L., and Daggett, V. (2004) From conversion to aggregation: Protofibril

formation of the prion protein. Proc. Natl. Acad. Sci. USA 101, 2293-2298

8. Smirnovas, V., Baron, G. S., Offerdahl, D. K., Raymond, G. J., Caughey, B., and Surewicz,

W. K. (2011) Structural organization of brain-derived mammalian prions examined by

hydrogen-deuterium exchange. Nat. Struct. Mol. Biol. 18, 504-506

9. Groveman, B. R., Dolan, M. A., Taubner, L. M., Kraus, A., Wickner, R. B., and Caughey,

B. (2014) Parallel In-register Intermolecular beta-Sheet Architectures for Prion-seeded

Prion Protein (PrP) Amyloids. J. Biol. Chem. 289, 24129-24142

10. Ghetti, B., Piccardo, P., Frangione, B., Bugiani, O., Giaccone, G., Young, K., Prelli, F.,

Farlow, M. R., Dlouhy, S. R., and Tagliavini, F. (1996) Prion protein amyloidosis. Brain

45

Pathol. 6, 127-145

11. Vanik, D. L., Surewicz, K. A., and Surewicz, W. K. (2004) Molecular basis of barriers for

interspecies transmissibility of mammalian prions. Mol. Cell 14, 139-145

12. Jones, E. M., Wu, B., Surewicz, K., Nadaud, P. S., Helmus, J. J., Chen, S. G., Jaroniec, C.

P., and Surewicz, W. K. (2011) Structural Polymorphism in Amyloids NEW INSIGHTS

FROM STUDIES WITH Y145Stop PRION PROTEIN FIBRILS. J. Biol. Chem. 286,

42777-42784

13. Chatterjee, B., Lee, C. Y., Lin, C., Chen, E. H. L., Huang, C. L., Yang, C. C., and Chen, R.

P. Y. (2013) Amyloid Core Formed of Full-Length Recombinant Mouse Prion Protein

Involves Sequence 127-143 but Not Sequence 107-126. Plos One 8, e67967

14. Tagliavini, F., Prelli, F., Verga, L., Giaccone, G., Sarma, R., Gorevic, P., Ghetti, B.,

Passerini, F., Ghibaudi, E., Forloni, G., Salmona, M., Bugiani, O., and Frangione, B. (1993)

Synthetic Peptides Homologous To Prion Protein Residues 106-147 Form Amyloid-Like

Fibrils In-Vitro. Proc. Natl. Acad. Sci. USA 90, 9678-9682

15. Skora, L., Fonseca-Ornelas, L., Hofele, R. V., Riedel, D., Giller, K., Watzlawik, J., Schulz-

Schaeffer, W. J., Urlaub, H., Becker, S., and Zweckstetter, M. (2013) Burial of the

Polymorphic Residue 129 in Amyloid Fibrils of Prion Stop Mutants. J. Biol. Chem. 288,

2994-3002

16. Li, H. C., Ye, S. J., Wei, F., Ma, S. L., and Luo, Y. (2012) In Situ Molecular-Level Insights

into the Interfacial Structure Changes of Membrane-Associated Prion Protein Fragment

118-135 Investigated by Sum Frequency Generation Vibrational Spectroscopy. Langmuir

28, 16979-16988

17. Apostol, M. I., Wiltzius, J. J. W., Sawaya, M. R., Cascio, D., and Eisenberg, D. (2011)

46

Atomic Structures Suggest Determinants of Transmission Barriers in Mammalian Prion

Disease. Biochemistry 50, 2456-2463

18. Chuang, C. C., Liao, T. Y., Chen, E. H. L., and Chen, R. P. Y. (2013) How do amino acid

substitutions affect the amyloidogenic properties and seeding efficiency of prion peptides.

Amino Acids 45, 785-796

19. Kuwata, K., Matumoto, T., Cheng, H., Nagayama, K., James, T. L., and Roder, H. (2003)

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

100, 14790-14795

20. Helmus, J. J., Surewicz, K., Apostol, M. I., Surewicz, W. K., and Jaroniec, C. P. (2011)

Intermolecular Alignment in Y145Stop Human Prion Protein Amyloid Fibrils Probed by

Solid-State NMR Spectroscopy. J. Am. Chem. Soc. 133, 13934-13937

21. Lin, N. S., Chao, J. C. H., Cheng, H. M., Chou, F. C., Chang, C. F., Chen, Y. R., Chang, Y.

J., Huang, S. J., and Chan, J. C. C. (2010) Molecular Structure of Amyloid Fibrils Formed

by Residues 127 to 147 of the Human Prion Protein. Chem. Euro. J. 16, 5492-5499

22. Skora, L., and Zweckstetter, M. (2012) Determination of amyloid core structure using

chemical shifts. Protein Sci. 21, 1948-1953

23. Zweckstetter, M. (2013) Conserved amyloid core structure of stop mutants of the human

prion protein. Prion 7, 193-197

24. Li, S. C., Goto, N. K., Williams, K. A., and Deber, C. M. (1996) Alpha-helical, but not

beta-sheet, propensity of proline is determined by peptide environment. Proc. Natl. Acad.

Sci. USA 93, 6676-6681

25. Wasmer, C., Lange, A., Van Melckebeke, H., Siemer, A. B., Riek, R., and Meier, B. H.

47

(2008) Amyloid fibrils of the HET-s(218-289) prion form a beta solenoid with a triangular

hydrophobic core. Science 319, 1523-1526

26. Cheon, M., Hall, C. K., and Chang, I. (2015) Structural Conversion of A beta(17-42)

Peptides from Disordered Oligomers to U-Shape Protofilaments via Multiple Kinetic

Pathways. Plos Comput. Biol. 11, e1004258

27. Frishman, D., and Argos, P. (1995) Knowledge-based protein secondary structure

assignment. Proteins: Struct., Funct., Genet. 23, 566-579

28. Humphrey, W., Dalke, A., and Schulten, K. (1996) VMD: Visual molecular dynamics. J.

Mol. Graph. Model. 14, 33-38

29. Ward, J. H. (1963) Hierarchical Grouping to Optimize an Objective Function. J. Am. Stat.

Assoc. 58, 236-244

30. Baftizadeh, F., Pietrucci, F., Biarnes, X., and Laio, A. (2013) Nucleation Process of a Fibril

Precursor in the C-Terminal Segment of Amyloid-beta. Phys. Rev. Lett. 110, 168103

31. Andrews, J. M., and Roberts, C. J. (2007) A Lumry-Eyring nucleated polymerization model

of protein aggregation kinetics: 1. Aggregation with pre-equilibrated unfolding. J. Phys.

Chem. B 111, 7897-7913

32. Biswal, H. S. (2015) Hydrogen Bonds Involving Sulfur: New Insights from ab Initio

Calculations and Gas Phase Laser Spectroscopy. in Noncovalent Forces (Scheiner, S. ed.),

Springer International Publishing. pp 15-45

33. Berka, K., Laskowski, R. A., Hobza, P., and Vondrasek, J. (2010) Energy Matrix of

Structurally Important Side-Chain/Side-Chain Interactions in Proteins. J. Chem. Theory

Comput. 6, 2191-2203

34. Pawar, A. P., DuBay, K. F., Zurdo, J., Chiti, F., Vendruscolo, M., and Dobson, C. M. (2005)

48

Prediction of "aggregation-prone" and "aggregation-susceptible" regions in proteins

associated with neurodegenerative diseases. J. Mol. Biol. 350, 379-392

35. Kim, C. W. A., and Berg, J. M. (1993) Thermodynamic Beta-Sheet Propensities Measured

Using A Zinc-Finger Host Peptide. Nature 362, 267-270

36. Kundu, B., Maiti, N. R., Jones, E. M., Surewicz, K. A., Vanik, D. L., and Surewicz, W. K.

(2003) Nucleation-dependent conformational conversion of the Y145Stop variant of

human prion protein: Structural clues for prion propagation. Proc. Natl. Acad. Sci. USA

100, 12069-12074

37. Helmus, J. J., Surewicz, K., Surewicz, W. K., and Jaroniec, C. P. (2010) Conformational

Flexibility of Y145Stop Human Prion Protein Amyloid Fibrils Probed by Solid-State

Nuclear Magnetic Resonance Spectroscopy. J. Am. Chem. Soc. 132, 2393-2403

38. Apostol, M. I., Sawaya, M. R., Cascio, D., and Eisenberg, D. (2010) Crystallographic

Studies of Prion Protein (PrP) Segments Suggest How Structural Changes Encoded by

Polymorphism at Residue 129 Modulate Susceptibility to Human Prion Disease. J. Biol.

Chem. 285, 29671-29675

39. Solforosi, L., Bellon, A., Schaller, M., Cruite, J. T., Abalos, G. C., and Williamson, R. A.

(2007) Toward molecular dissection of PrPC-PrPSc interactions. J. Biol. Chem. 282, 7465-

7471

40. Natalello, A., Prokorov, V. V., Tagliavini, F., Morbin, M., Forloni, G., Beeg, M., Manzoni,

C., Colombo, L., Gobbi, M., Salmona, M., and Doglia, S. M. (2008) Conformational

Plasticity of the Gerstmann–Sträussler–Scheinker Disease Peptide as Indicated by Its

Multiple Aggregation Pathways. J. Mol. Biol. 381, 1349-1361

41. Cho, K. R., Huang, Y., Yu, S. L., Yin, S. M., Plomp, M., Qiu, S. R., Lakshminarayanan, R.,

49

Moradian-Oldak, J., Sy, M. S., and De Yoreo, J. J. (2011) A Multistage Pathway for Human

Prion Protein Aggregation in Vitro: From Multimeric Seeds to beta-Oligomers and

Nonfibrillar Structures. J. Am. Chem. Soc. 133, 8586-8593

42. Debenedetti, P. G. (1996) Metastable liquids : concepts and principles, Princeton

University Press, Princeton, N.J.

43. Ciryam, P., Kundra, R., Morimoto, R. I., Dobson, C. M., and Vendruscolo, M.

Supersaturation is a major driving force for protein aggregation in neurodegenerative

diseases. Trends Pharmacol. Sci. 36, 72-77

44. Helmus, J. J., Surewicz, K., Nadaud, P. S., Surewicz, W. K., and Jaroniec, C. P. (2008)

Molecular conformation and dynamics of the Y145Stop variant of human prion protein.

Proc. Natl. Acad. Sci. USA 105, 6284-6289

45. Alder, B. J., and Wainwright, T. E. (1959) Studies In Molecular Dynamics .1. General

Method. J. Chem. Phys. 31, 459-466

46. Smith, A. V., and Hall, C. K. (2001) alpha-helix formation: Discontinuous molecular

dynamics on an intermediate-resolution protein model. Proteins: Struct., Funct., Genet. 44,

344-360

47. Marchut, A. J., and Hall, C. K. (2006) Spontaneous formation of annular structures

observed in molecular dynamics simulations of polyglutamine peptides. Comput. Biol.

Chem. 30, 215-218

48. Nguyen, H. D., and Hall, C. K. (2004) Molecular dynamics simulations of spontaneous

fibril formation by random-coil peptides. Proc. Natl. Acad. Sci. USA 101, 16180-16185

49. Cheon, M., Chang, I., and Hall, C. K. (2010) Extending the PRIME model for protein

aggregation to all 20 amino acids. Proteins: Struct., Funct., Bioinf. 78, 2950-2960

50

50. Cheon, M., Chang, I., and Hall, C. K. (2012) Influence of temperature on formation of

perfect tau fragment fibrils using PRIME20/DMD simulations. Protein Sci. 21, 1514-1527

51. Cheon, M., Chang, I., and Hall, C. K. (2011) Spontaneous Formation of Twisted A beta(16-

22) Fibrils in Large-Scale Molecular-Dynamics Simulations. Biophys. J. 101, 2493-2501

52. Wagoner, V. A., Cheon, M., Chang, I., and Hall, C. K. (2012) Fibrillization Propensity for

Short Designed Hexapeptides Predicted by Computer Simulation. J. Mol. Biol. 416, 598-

609

53. de la Paz, M. L., and Serrano, L. (2004) Sequence determinants of amyloid fibril formation.

Proc. Natl. Acad. Sci. USA 101, 87-92

54. Latshaw, D. C., Cheon, M., and Hall, C. K. (2014) Effects of Macromolecular Crowding

on Amyloid Beta (16-22) Aggregation Using Coarse-Grained Simulations. J. Phys. Chem.

B 118, 13513-13526

55. Latshaw, D. C., and Hall, C. K. (2015) Effects of Hydrophobic Macromolecular Crowders

on Amyloid beta (16-22) Aggregation. Biophys. J. 109, 124-134

56. Munoz, V., and Serrano, L. (1995) Elucidating The Folding Problem Of Helical Peptides

Using Empirical Parameters .3. Temperature And Ph-Dependence. J. Mol. Biol. 245, 297-

308

57. Ricchiuto, P., Brukhno, A. V., and Auer, S. (2012) Protein Aggregation: Kinetics versus

Thermodynamics. J. Phys. Chem. B 116, 5384-5390

58. Dougherty, R. C. (1998) Temperature and pressure dependence of hydrogen bond strength:

A perturbation molecular orbital approach. J. Chem. Phys. 109, 7372-7378

59. Einstein, A. (1905) On the movement of small particles suspended in stationary liquids

required by the molecular-kinetic theory of heat. Ann. Phys. 17, 549-560

51

60. Abraham, M. J., Murtola, T., Schulz, R., Páll, S., Smith, J. C., Hess, B., and Lindahl, E.

(2015) GROMACS: High performance molecular simulations through multi-level

parallelism from laptops to supercomputers. SoftwareX 1, 19-25

61. Hornak, V., Abel, R., Okur, A., Strockbine, B., Roitberg, A., and Simmerling, C. (2006)

Comparison of multiple Amber force fields and development of improved protein

backbone parameters. Proteins: Struct., Funct., Bioinf. 65, 712-725

62. Andersen, H. C. (1980) Molecular dynamics simulations at constant pressure and/or

temperature. J. Chem. Phys. 72, 2384-2393

52

A B

C

Figure 2.1. Total interaction energy (kJ/mol) of HuPrP120-144 aggregation versus simulation time at reduced temperature T*=0.195 and peptide concentration of 20mM.

53

A C E

B D F

Figure 2.2. S-shaped (A), Ω-shaped (C) and U-shaped (E) protofilament formed by HuPrP120- 144 from the 5th, 12th and 11th runs. B, D, F. Schematic representations of peptide conformations in the S-, Ω- and U-shaped protofilaments. The sizes of the sidechain beads don’t reflect the actual radii. Hydrophobic residues (white), positively charged residues (red), negatively charged residues (blue), and polar residues (green).

54

A C

B D

Figure 2.3. The S-shaped (A) and Ω-shaped (C) protofilaments formed by BVPrP120-144 from the 5th and 4th runs. B, D. Schematic representations of peptide conformation in S-and Ω-shaped protofilaments. The color scheme is same as that used in Figure 2.

55

A C

B D

Figure 2.4. The S-shaped (A) and Ω-shaped (C) protofilaments formed by SHaPrP120-144 from the 7th and 1st runs. B, D. Schematic representations of peptide conformation in S- and Ω-shaped protofilaments.

56

Figure 2.5. Secondary structure distribution of the S-shaped protofilament formed by HuPrP120- 144. The secondary structure content is calculated and averaged over the last 100 billion collisions using the STRIDE program. (27) β-sheet, random coil and turn structure are three major types. Other types of secondary structures are not shown here.

57

A B C

D E F

Figure 2.6. Simulation snapshots for the 3rd run of HuPrP120-144. Snapshots are taken at t = (A) 1, (B) 2900, (C) 4187, (D) 4559, (E) 6835 (ns). F shows the time evolution of system total interaction energy (kJ/mol).

58

A B C

D E F

Figure 2.7. Simulation snapshots for the 5th run of HuPrP120-144. Snapshots are taken at t = (A) 1, (B) 1191, (C) 3065, (D) 6348, (E) 11393 (ns). F shows the time evolution of system total interaction energy (kJ/mol).

59

A B

Figure 2.8. Average sidechain interaction energy (kJ/mol) (A) and average number of sidechain contacts (B) experienced by residue 138, 139 and 143 during the nucleation stage.

60

Figure 2.9. Number of hydrogen bonds formed as a function of time during HuPrP120-144 and PrP120-137 aggregation.

61

A

B

C

Figure 2.10. Schematic representation of a parallel β-sheet segment (A) and two types of anti- parallel β-sheet segments (B, C) formed between two short peptide strands. Three residues are involved in each strand. The dash line means the N-H and C=O groups form a hydrogen bond.

62

Table 2.1. A comparison of total interaction energy (ETotal), hydrogen bonding energy (EHB), sidechain-sidechain interaction energy (ESidechain) of protofilaments formed by Hu, BV and Sha PrP120-144.

Protofilament EHB (kJ/mol) ESidechain ETotal (kJ/mol) (kJ/mol) S-shaped HuPrP120-144 -1902.05±53.62 -396.42±15.46 -2298.47±50.63

U-shaped HuPrP120-144 -1884.22±64.35 -551.30±21.20 -2435.52±54.44 Ω-shaped HuPrP120-144 -1773.23±52.37 -494.19±20.45 -2275.40±61.98 S-shaped BVPrP120-144 -1907.66±58.36 -396.42±15.34 -2304.08±63.35 Ω-shaped BVPrP120-144 -1888.58±53.12 -525.74±17.83 -2414.32±55.49

S-shaped SHaPrP120-144 -1788.95±67.34 -380.58±22.45 -2169.53±72.82 Ω-shaped SHaPrP120-144 -1893.07±63.22 -441.81±15.34 -2334.88±67.94

63

Table 2.2. Lag phase time of Hu, BV, SHa PrP120-144 aggregation.

Lag phase time Species (ns) HuPrP120-144 5179±1161 BVPrP120-144 7298±1325 SHaPrP120-144 9037±1862

64

Table 2.3. Average percentage of parallel and anti-parallel β-sheet segments in Hu, BV and SHaPrP120-144 oligomers and protofilaments.

Species Parallel β-sheet Anti-parallel β-sheet Total β-sheet segments segments segments

HuPrP120-144 76.05±0.60 28.64±0.64 104.69±1.24

BVPrP120-144 67.70±0.71 31.05±0.98 98.75±1.69

SHaPrP120- 69.26±0.44 25.96±0.99 95.22±1.43 144

65

Table 2.4. Amino acid sequence of three PrP120-144 peptides.

12 130 140 144 0 Human A V V G G L G G Y M L G S A M S R P I I H F G S D Bank A V V G G L G G Y M L G S A M S R P M I H F G N D Vole Syrian A V V G G L G G Y M L G S A M S R P M M H F G N D Hamster

66

2.7 SUPPORTING INFORMATION

A B

Figure S2.1. Total interaction energy (kJ/mol) of BVPrP120-144 aggregation versus simulation time at reduced temperature T*=0.195 and peptide concentration of 20mM.

67

A B

Figure S2.2. Total interaction energy (kJ/mol) of SHaPrP120-144 aggregation versus simulation time at reduced temperature T*=0.195 and peptide concentration of 20mM.

68

1 2 3 4 5

6 7 8 9 10

11 12 13 14 15

Figure S2.3. Final structures for 8 HuPrP120-144 peptides from 15 independent runs at T*=0.195. Structures are taken after at least 800 billion collisions (t ≈ 9μs).

69

1 2 3 4 5

6 7 8 9 10

11

Figure S2.4. Final structures for 8 BVPrP120-144 peptides from 11 independent runs at T*=0.195. Structures are taken after at least 800 billion collisions (t ≈ 9μs).

70

1 2 3 4 5

6 7 8 9 10

11 Figure S2.5. Final structures for 8 SHaPrP120-144 peptides from 11 independent runs at T*=0.195. Structures are taken after at least 800 billion collisions (t ≈ 9μs).

71

Table S2.1. Square well interaction parameters between 19 side-chain centroids in reduced energy unit εHB which is set to be 12.47 kJ/mol.

72

A B

C D

Figure S2.6. Helical content of an alanine-rich polypeptide (Ac-AAAAKAAAAKAAAAKA-NH2) changes as a function of time. (A. DMD Simulation data and B. experimental data in literature. 1 C. Linear fitting correlation of simulation and real temperature where peptide adopt the same helical content percentage. R2 value stands for how close the data points are close to the fitted regression line.) D. Plot of 5 linear correlation functions for 5 peptides. (1: Ac- AAQAAAAQAAAAQAAY-NH2; 2: Ac-AAAAKAAAAKAAAAKA-NH2; 3: Ac- AEAAAKEAAAKEAAAKA-NH2; 4: Ac-WDAAAKDAAAKDAAAKA-NH2; 5: Ac- EAEKAAKEAEKAAKEAEK-NH2.)

73

Reference

1. Muñoz, V., and Serrano, L. (1995) Elucidating the Folding Problem of Helical Peptides

using Empirical Parameters. III> Temperature and pH Dependence. J. Mol. Biol. 245, 297-

308.

74

2.8 ERRATUM

In the Materials and Methods section, the conversion of the reduced time unit to a real time unit

* * (Δt = 0.072ns) was incorrect. The correct value should be Δt = 0.96ns. Thus, the longest time accessed in the single DMD simulation should have been 672μs which corresponds to 2000 billion collisions. The simulation times (x axes) in Figs. 1, 6, 7, 9 and S1-5 and Table 2 as well as in other places in the context of the article should all be multiplied by a factor of 29. The correct calculation procedures are described below:

The way that we relate the reduced time unit to the real time unit is to compare the self-diffusion coefficient of Aβ(16-22) peptide obtained from DMD simulation (퐷퐷푀퐷) and atomistic molecular dynamics (MD) simulation (퐷푀퐷 ) at T=298K. The reduced temperature of the self-diffusion coefficient measurement in DMD simulation is chosen to be T*=0.181, which corresponds to room temperature (T=298K) according to temperature scaling 1, T*= (T+115.79)/2288.46 =

(298+115.79)/2288.46 = 0.181.

The self-diffusion coefficients in both DMD and MD simulation are calculated using the

Einstein equation. We plotted the mean square displacement (MSD) of N Aβ(16-22) peptides versus time for both atomistic MD simulation (N=1) and DMD simulation (N=6), as shown in

Figs. 1A and B, by using the following equation.

2 ∑𝑖=푁|푟푡+∆푡 − 푟푡| 푀푆퐷(푡 + 훥푡) = 푀푆퐷(푡) + 푑̅̅2̅ = 푀푆퐷(푡) + 𝑖=1 𝑖 𝑖 푁 where 훥푡 is the time unit; N is the total number of peptides; d̄ 2 is the averaged square displacement

푡+∆푡 푡 푡ℎ of N peptides; and 푟𝑖 and 푟𝑖 are the coordinates of the 푖 peptide at 푡 + ∆푡 and 푡. For DMD simulation, 푡∗ is used in DMD instead of t to represent reduced time.

Using the MSD data in Figs. 1A and B, we obtain the self-diffusion coefficients 퐷푀퐷 and 퐷퐷푀퐷,

75

1 푀푆퐷(푡2)−푀푆퐷(푡1) 48.11−27.45 2 퐷푀퐷 = = = 0.48 푛푚 ⁄푛푠 6 푡2−푡1 6∗(17.80−10.66)

∗ ∗ 1 푀푆퐷(푡2)−푀푆퐷(푡1) 328.57−109.01 2 ∗ 퐷퐷푀퐷 = ∗ ∗ = = 0.46 푛푚 ⁄∆푡 6 푡2−푡1 6∗(120.01−40.01)

By equating the self-diffusion coefficients calculated from both atomistic MD simulation and

DMD simulation, we obtain the reduced time unit in DMD simulation in terms of the real time unit to be 1 (∆t∗) = 0.96 ns.

A B

Figure E2.1. Plots of the MSD versus time calculated from atomistic simulation (A) and DMD simulation (B). The two points are chosen from each of two plots to calculate the slope of MSD line versus time.

76

Reference

1. Wang Y, Shao Q, Hall CK. N-terminal Prion Protein Peptides (PrP120-144) Form Parallel

In-register β-sheets via Multiple Nucleation-dependent Pathways. Journal of Biological

Chemistry 2016;291:22093-105.

77

CHAPTER 3

Seeding and Cross-Seeding Fibrillation of N-terminal Prion Protein Peptides PrP(120-144)

Chapter 3 is essentially a manuscript by Yiming Wang and Carol K. Hall published on Protein

Science.

78

Seeding and Cross-Seeding Fibrillation of N-terminal Prion Protein Peptides PrP(120-144)

Yiming Wang and Carol K. Hall*

Department of Chemical and Biomolecular Engineering, North Carolina State University, Raleigh, North Carolina 27695-7905, United States

Abstract

Prion diseases are infectious neurodegenerative diseases that are capable of cross-species transmission, thus arousing public health concerns. Seed-templating propagation of prion protein is believed to underlie prion cross-species transmission pathology. Understanding the molecular fundamentals of prion propagation is key to unravelling the pathology of prion diseases. In this work, we use coarse-grained molecular dynamics to investigate the seeding and cross-seeding aggregation of three prion protein fragments PrP(120-144) originating from human (Hu), bank vole (BV) and Syrian hamster (SHa). We find that the seed accelerates the aggregation of the monomer peptides by eliminating the lag phase. The monomer aggregation kinetics are mainly determined by the structure of the seed. The stronger the hydrophobic residues on the seed associate with each other, the higher the probability that the seed recruits monomer peptides to its surface/interface. For cross-seeding aggregation, we show that Hu has a strong tendency to adopt the conformation of the BV seed and vice versa; the Hu and BV monomers have a weak tendency to adopt the conformation of the SHa seed. These two findings are consistent with Apostol et al.

’s experimental findings on PrP(138-143) and partially consistent with Jones et al ’s finding on

PrP(23-144). We also identify several conformational mismatches when SHa cross-seeds BV and

Hu peptides, indicating the existence of a cross-seeding barrier between SHa and the other two

79

sequences. This work sheds light on the molecular mechanism of seed-templating aggregation of prion protein fragments underlying the sequence-dependent transmission barrier in prion diseases.

80

3.1 Introduction

Prion diseases are a family of infectious neurodegenerative diseases that include kuru and

Creutzfeldt-Jakob disease (CJD) in humans, scrapie in goats, bovine spongiform encephalopathy

(BSE) in cattle and chronic wasting disease (CWD) in deer. 1-3 The infectious agent in prion disease is the prion protein scrapie (PrPSc) which has diverse β-sheet-rich conformations.4 PrPSc replicates without the help of nucleic acids by templating the misfolding and accumulation of normal prion protein into misfolded PrPSc, the mechanism of which is similar to the seeded polymerization of amyloids. 5-8

The significant threat of prion disease to human health is its potential cross-species transmission. For example, BSE (also known as Mad Cow Disease) is thought to originate from transmission of a scrapie prion from sheep to cattle. 9 Humans can acquire a CJD variant by consuming bovine products contaminated by BSE. 9, 10 Experimental mice can acquire variant CJD by intracerebral inoculation of brain homogenates from cattle affected by BSE. 11 Cross-species transmission of prion disease may be related to the direct interaction between the misfolded and cellular forms of prion proteins from different species, leading to cross-seeding of prion protein aggregation. 12, 13 Understanding the molecular mechanism underlying the cross-seeding aggregation of prion protein could pave the way to elucidating the pathology of cross-species transmission.

The cross-seeding efficiency and specificity of full-length prion protein and its N-terminal fragments have been investigated in vitro. Vanik et al. 14 found that Syrian hamster PrP(23-144) can seed the fibrillation of mouse PrP(23-144) while mouse PrP(23-144) cannot seed the fibrillation of Syrian hamster PrP(23-144), indicating that mouse and Syrian hamster PrP have an asymmetric cross-seeding barrier14. They further found14 that two species-specific mutations

81

(I138M and I139M) are critical in determining the cross-seeding specificity of PrP(23-144).

Experiments by Lee et al. 15 showed that residues 109, 112 and 139 determine the seeding and cross-seeding efficiency of mouse and hamster PrP(108-144). Chuang et al. 16 found that the mutation L138M promotes the amyloidogenesis of bovine PrP(108-144) peptide and increases its seeding efficiency while the mutations I139M and N143S do the opposite. Apostol et al. 17 found that human and mouse PrP(138-143) fibrils both form parallel steric zippers while hamster

PrP(138-143) fibril forms anti-parallel steric zippers. When cross-seeded with hamster PrP(138-

143) fibril, human and mouse PrP(138-143) fibrils never form anti-parallel steric zippers.

Molecular simulations have also been applied to investigate the seeding and cross-seeding mechanism underlying protein aggregation. A number of amyloid β fragments, e.g. Aβ(16-22) 18,

Aβ(35-40) 19 and Aβ(17-42)20, have been shown in silico to follow a two-step “dock-lock” mechanism when monomer peptides of the same sequence attach to the pre-formed fibril ends.

The cross-seeding interaction between two different peptides to form hetero-assemblies have also been investigated. Qi et al. found using molecular dynamics (MD) simulation that monomeric tau protein is stabilized by stretching its conformation and exposing its amyloidogenic motifs in the presence of the Aβ(1-42) protofilament. 21 Zhang et al. showed using MD simulation that the rat islet amyloid polypeptide (rIAPP) docks to the end of the human islet amyloid polypeptide

(hIAPP) protofilament to form stable rIAPP-hIAPP assemblies.22 Hu et al. also showed that the

U-shaped Aβ and hIAPP protofilaments can stack on top of each other to form double-layer assemblies. 23

PRIME20 is a knowledge-based four-bead-per-residue coarse-grained protein model developed in the Hall group that was specifically designed for discontinuous molecular dynamics

(DMD) simulation of protein aggregation. DMD/PRIME20 is able to simulate the dynamic process

82

of large-scale peptide aggregation from an initial random-coil configuration to disordered oligomers and then fibrillar structures. 24-26 In recent years, DMD/PRIME20 have been applied to investigate aggregation of prion protein fragments27, tau protein fragments28, Aβ(16-22)25, and

Aβ(17-42)29 peptides. In addition, it has been extended to investigate the effect of macromolecular crowding26, 30, and polyphenol inhibition 31 on peptide aggregation. In recent work, we performed coarse-grained simulations of the spontaneous aggregation of the three prion protein fragments,

Hu, BV and SHaPrP(120-144) and found that they form similar polymorphic β-sheet-rich fibrillar structures due to the similarity in their amino acid sequences. 32 HuPrP(120-144) aggregates to form the S-, Ω- and U-shaped protofilaments, while BV and SHaPrP(120-144) form the S- and Ω- shaped protofilaments, as shown in Figs. 1 and S1. In this work, we investigate the seeding and cross-seeding aggregation of these three peptides in the presence of the pre-formed S-, Ω- and U- shaped protofilaments.

Highlights of our results are as follows. In comparing seeding versus non-seeding aggregation, we find that the seed accelerates the aggregation kinetics by eliminating the lag phase for forming a stable nucleus. Analysis of the homogeneous seeding aggregation of HuPrP(120-

144) shows that the structure of the seed has a dominant effect on peptide aggregation kinetics.

The U-shaped Hu seed has a higher seeding efficiency than the S- and Ω-shaped Hu seeds because the peptides in the U-shaped seed have stronger sidechain-sidechain hydrophobic interactions, which help stabilize and template the incoming monomer peptides. From the cross-seeding simulation of Hu, BV and SHaPrP(120-144), we find that HuPrP(120-144) has a strong tendency to adopt the conformation of the BV seed and vice versa, indicating that Hu and BVPrP(120-144) may have a low cross-seeding barrier. Also, the BV seed has high efficiency in templating the Hu and SHa monomer while the SHa seed has low efficiency in templating the Hu, BV monomer.

83

This suggests that SHaPrP(120-144) may have high cross-seeding barrier with Hu and

BVPrP(120-144). The cross-seeding efficiency of the Hu, BV and SHa seeds in simulation is consistent with experimental findings on the species-dependent seeding specificity of PrP(138-

143) 17.

3.2 Results

3.2.1 Seeding vs non-seeding aggregation of HuPrP(120-144) peptides

Here we apply DMD simulation to investigate the effect of seeding on prion protein peptide aggregation. We first simulate the aggregation of eight initially disordered HuPrP(120-144) peptides in the presence of a pre-formed S-shaped HuPrP(120-144) protofilament. The pre-formed protofilament is regarded as the seed in our simulation. Fig. 1 describes the percentage of β-sheet versus time in seeding and non-seeding simulation of HuPrP(120-144) peptides. From Fig. 2A, in the absence of the seed, the aggregation of initially disordered HuPrP(120-144) peptides (Fig. 2B) has a lag phase before forming a parallel in-register S-shaped protofilament (Fig. 2C).32 In the presence of the seed, the aggregation lag phase is eliminated and the aggregation of HuPrP(120-

144) is significantly accelerated since monomer peptides can directly attach to the preformed seed without overcoming the nucleation barrier. From Figs. 2D and E, during the seeding aggregation, the pre-formed seed elongates by recruiting seven out of eight monomeric peptides and templating them into parallel in-register β-sheets, essentially adopting the same S-shaped conformation as the seed. Similarly, the aggregation of BV- and SHaPrP(120-144) peptides are also accelerated in the presence of the BV- and SHaPrP(120-144) seeds.

84

3.2.2 Homogeneous Seeding Aggregation of PrP(120-144)

We investigate the effect of polymorphic Hu seeds (U-, S- and Ω-shaped) on HuPrP(120-

144) aggregation kinetics and structures. The average potential energy versus time of HuPrP(120-

144) aggregation in the presence of the S, Ω and U-shaped seeds from five independent runs are shown in Fig. 3A. From Fig. 3A, the potential energy of the system using the U-shaped seed decreases faster than those of the S- and Ω-shaped seed, indicating that the U-shaped seed has the highest seeding efficiency among the three Hu seeds. In Table 1, we list the average number of

HuPrP(120-144) peptides that adopt the seed conformation, partially adopt the seed conformation or adopt other conformations by the end of the aggregation. The large error bars in the average number of peptides adopting the conformation of the seed indicates large deviations in the number of peptides adopting the conformation of the seed among the five independent runs for each type of seeding simulation. We find that the average numbers of HuPrP(120-144) peptides that adopt the conformation of the Ω- and U-shaped seeds are slightly higher than the number of peptides to adopt that conformation of the S-shaped seed. As shown from Fig. 3B, peptides in the original U and Ω-shaped seeds have stronger sidechain-sidechain hydrophobic interactions than in the S- shaped seed. The explanation is that the U-shaped seed has the most stable hydrophobic core among the three seeds, leading to high efficiency in stabilizing and templating the monomer peptide when attaching to it.

In addition, we compare the aggregation rate of Hu, BV and SHaPrP(120-144) in the presence of the S-shaped Hu, Ω-shaped BV and Ω-shaped SHaPrP(120-144) seeds, respectively.

As shown in Fig. 3C, the potential energy of the BV monomer with the BV seed decreases fastest among the three systems. From Table 2, we find that the number of BVPrP(120-144) peptides adopting the conformation of the Ω-shaped BV seed is higher than those of Hu and ShaPrP(120-

85

144) adopting the conformation of their corresponding seeds. This indicates that the Ω-shaped

BVPrP(120-144) seed has higher homogeneous seeding efficiency than the S-shaped Hu and Ω- shaped SHa seeds. The reason is that, from Fig. 3D, BVPrP(120-144) peptide in the original Ω- shaped BV seed has stronger sidechain interaction energy than those of the S-shaped Hu and Ω- shaped SHa seeds.

3.2.3 Cross-seeding aggregation of Hu, BV and SHaPrP(120-144) Peptides

We also performed simulations of six types of cross-seeding aggregation for the Hu, BV and SHaPrP(120-144) peptides. Similar to the homogeneous seeding simulation results, our cross- seeding simulation results also showed that the structure of the seed has a major effect on the monomer aggregation kinetics which is dominant over the monomer-seed sequence difference.

Fig. 4 plots the potential energy profiles of three homogeneous seeding and six cross-seeding aggregation simulations. From Fig. 4A, in the presence of the S-shaped Hu seed, Hu, BV and

SHaPrP(120-144) peptides all have similar aggregation kinetics regardless of the fact that

HuPrP(120-144) has higher aggregation propensity than BV and SHaPrP(120-144) 32, indicating that the sequence differences of the monomer play a minor role on the cross-seeding efficiency.

Note that mouse prion has the same PrP(120-144) sequence as bank vole. Lee et al. 15 had a similar experimental finding that using SHaPrP(108-144) to seed the fibrillation of its M139I mutant is as efficient as homologous seeding. In addition, Figs. 4B-D and S2 show that the aggregation rates of the Hu, BV and SHaPrP(120-144) peptides in the presence of the Ω-shaped BV seed are always higher than those in the presence of the S- and Ω-shaped Hu seeds and the Ω-shaped SHa seed.

This indicates that the Ω-shaped BV seed has the highest cross-seeding efficiency among the three seeds. We also notice from Fig. 3D that the peptides in the original Ω-shaped BV seed have the

86

strongest sidechain-sidechain interaction energy among the three seeds, indicating that the Ω- shaped BV seed has a strong hydrophobic core. Thus, the stronger the hydrophobic residues associate with each other within the seed, the higher the probability that the seed recruits monomer peptides on its surface/interface and, hence the faster it elongates.

For the cross-seeding aggregation, we compare the tendencies of the monomers to adopt the seed conformation. In Table 2, we list the average number of Hu, BV and SHaPrP(120-144) peptides that adopt the seed conformation from five independent runs. We found is that Hu and

BVPrP(120-144) have a relatively low cross-seeding barrier. From Table 2, when the HuPrP(120-

144) is cross-seeded with the Ω-shaped BV seed, most of the HuPrP(120-144) peptides adopt the peptide conformation of the BV seed; when the BVPrP(120-144) is cross-seeded with the S-shaped

Hu seed, most of the BVPrP(120-144) peptides adopt the peptide conformation of the Hu seed. In addition, we notice that BVPrP(120-144) peptides adopt the conformation (7.0±0.7 out of 8) of Hu

S-shaped seed slightly more than they adopt the conformation of the Ω-shaped BV seed (6.2±1.3 out of 8). Since these two numbers are within standard deviation of each other, we think that the tendency of BVPrP(120-144) peptide to adopt the conformation of the S-shaped Hu is similar to its tendency to adopt the conformation of the Ω-shaped BV seeds. This is consistent with experimental finding by Apostol et al.17 that the crystal structure of fibrils formed by Hu and

MoPrP(138-143) are similar, but are different from those formed by SHaPrP(138-143). Both the

HuPrP(138-143) monomer with the pre-formed MoPrP(138-143) fibril crystals and vice versa show seeding. This is also consistent with the experimental finding by Jones et al. 33 that the human and mouse PrP(23-144) fibrils both have twisted morphologies and that one can be used to seed the amyloid formation of the other. The major reason for the weak tendency of Hu and BV to adopt the conformation of the Ω-shaped SHa seed is that conformation of the Ω-shaped SHa seed is hard

87

for a monomer to adopt. From Table 2, not only Hu and BV but SHa have weak tendencies to adopt the conformation of the Ω-shaped SHa seed. Also, most of the SHa monomers adopt the BV seed conformation but most of the BV monomers do not adopt the SHa seed, which is inconsistent with the finding on PrP(23-144) by Vanik et al.14

To evaluate the impact of seeding on the three mutated residues of Hu, BV and

SHaPrP(120-144) peptides, we calculated the interaction energies experienced by each of the three mutated residues on the monomer peptides with all other residues in the system averaged over the entire trajectory in Fig. 5. Firstly, compared with the non-seeding simulations (Fig. 8A from ref

32), the interaction energies experienced by residue 138 on monomer Hu, BV and SHa in the seeding simulations are all increased, while those of residues 139 and 143 decreased. Secondly, residue 138 has the strongest interaction energy with the seeds regardless of whether it is on the

Hu, BV or SHa peptides. It also seems to be the most affected by the type of seed and has its strongest interaction energy in the presence of the BV seed, regardless of whether it is on Hu, BV and SHa peptides. Residue 139 is modestly affected by the seed and residue 143 seems hardly affected by the seeds.

We also noticed from Table 2 that only a small portion of the Hu and BV monomers adopt the conformation of the SHa seed, indicating a strong cross-seeding barrier between SHaPrP(120-

144) and the other two sequences. This is also consistent with experimental findings by Apostol et al.17 that when seeded with SHaPrP(138-143) fibril, Hu and MoPrP(138-143) cannot reproduce the morphology of the SHaPrP(138-143) fibril.

88

3.2.4 Structural Mismatch in Homogeneous and Heterogeneous PrP(120-144) fibril

We observed a number of different mismatches between the conformations of the monomers after they attach to the seed and the original seed conformation in the seeded and cross- seeded protofilaments. Fig. 6A shows that for Hu aggregating in the presence of the SHa seed, four Hu peptides and one SHa peptide form a parallel in-register β-sheet but with their C-terminal

(residues 137Pro to 144Asp) detached from the Ω-shaped SHa seed. This molecular-level structural mismatch is similar to the hybridized fibril model proposed by Makarava et al. 34 who found that when mouse full length prion protein is seeded with 30% hamster fibrils, a hybrid Mo-

SHa fibril is formed having two distinct Mo and SHa domains connected by a shared common local structural motif. We also found some other types of structural mismatch during the seeding and cross-seeding aggregation simulations. From Fig. 6B, in the case of BV aggregation with the

SHa seed, residues 120Ala to 131Gly of BV adopt a “triangle” conformation which is a mirror image of the conformation adopted by residues 120Ala to 131Gly of peptides in the seed. From

Fig. 6C, in the case of BV aggregating with the SHa seed, one BV monomer adopts a configuration that is antiparallel to the seed conformation. This conformation mismatch causes the other three

BV peptides to adopt the S-shaped conformation. From Fig. 6D, in the case of SHa aggregation with the BV seed, one SHa monomer adopts the seed conformation but only near the N-terminal region (residues 120Ala to 131Gly), leaving its tail (132Ser to 144Asp) to float in a disordered conformation. From Fig. 6E, in the case of SHa aggregation with the SHa seed, two SHa monomers adopt Ω-shaped conformations that are slightly different from the original Ω-shaped SHa seed.

Our finding from simulation indicates that during seeding and cross-seeding aggregation, the seed may not only template monomer peptides to form the same conformation as the seed but may also

89

induce monomer peptides to form new conformations which may eventually elongate into fibrils with new morphology, or new “strains”.

3.3 Discussion and Conclusion

Coarse-grained molecular dynamics simulations were used to investigate the early-stage aggregation of three prion protein fragments Hu, BV and SHaPrP(120-144). In a previous paper, we had investigated the spontaneous aggregation (nucleation) of PrP(120-144) peptides, similar to the work by Cheon et al. 29 for Aβ(17-42) peptides. In that paper, we showed that with the sequence differences in residues 138, 139 and 143 in Hu, BV and SHaPrP(120-144) causes these peptides to have different aggregation propensities (Hu > BV > SHa). It also causes these peptides to form various fibrillar structures including the S-, Ω- and U-shaped protofilaments. In this paper, we investigated the early-stage seeding aggregation of these three peptides. We show that the seeding aggregation kinetics depends primarily on the structure of the seed, or the hydrophobicity of the seed core. The hydrophobic sidechain-sidechain attraction, along with the backbone hydrogen bonding interaction, are the driving forces for seeding and cross-seeding aggregation. The stronger the hydrophobic residues associate with each other within the seed, the higher the probability that the seed recruits monomer peptides on its surface/interface, and hence the faster it elongates. We also surmise that in experiments, initially PrP(120-144) forms polymorphic fibrillar oligomers.

The ones with the lower energies are likely to have faster elongation rates.

In addition, conformational mismatches between the monomer and the seed are commonly observed during the fibril elongation, explaining the polymorphism in the final morphology of the macroscopic fibril. An implication of these conformational mismatches is that even when the monomer and the seed consist of the same prion protein, the seeded aggregation may still generate

90

new fibril structures, which could give rise to new variants of prion diseases. Within a single population, e.g. human, various strains of CJD are believed to be caused by the same prion protein but with diverse PrPSc conformations. 35

In the cross-seeding simulation of PrP(120-144), mutations on residues 138, 139 and 143 between species give rise to differences in the seeding efficiency, specificity and fibril morphology, suggesting that there is a species-dependent seeding barrier for prion proteins. On one hand, HuPrP(120-144) has a strong tendency to adopt the conformation of the BV seed and vice versa. On the other hand, the Hu and BV monomers have relatively weak tendencies to adopt the conformation of the SHa seed. The mismatch between the sequences of the SHa seed and the

Hu and BV monomer causes structural mismatches and decreases thermodynamic stability of the cross-seeded fibril, which leads to a decrease in the cross-seeding aggregation rate. These simulation results are consistent with the experimental findings for PrP(23-144) 14, 33 and PrP(138-

143) 17. In addition, bank vole is a unique species that has been shown in vivo to be a universal acceptor of various prion diseases. 36 This indicates that BVPrP has no inter-species transmission barrier and can likely subject to various infectious PrPSc from other species.

The small system size in the current simulations (eight peptides plus a pre-formed octameric seed) allow us to investigate the elongation of the PrP(120-144) protofilament. As revealed from our and other investigators’ simulations of fibrillation of short amyloidogenic peptides like Aβ(16-22) 25, 26, the complete aggregation pathway includes oligomerization, nucleation and fibril elongation stages. To investigate the complete aggregation pathway of Hu,

BV and SHaPrP(120-144) peptides, we may need to perform larger scale simulations of hundreds of peptides and fit the simulation data to the self-consistent solution of a master equation of the

91

fibril growth by Cohen et al. 37 to predict the primary nucleation rate, the secondary nucleation rate and the fibril elongation rate.

3.4 Methods

3.4.1 DMD and PRIME20 force field

Discontinuous molecular dynamics (DMD), a fast alternative to traditional molecular dynamics, is employed as our main simulation method. 38 PRIME20 is a four-bead-per-residue coarse-grained protein model developed in the Hall group that was specifically designed for DMD simulation of protein aggregation. 39 The PRIME20 force field models the 20 different amino acids

(each contains NH, Cα, CO and sidechain sphere R) with 20 different sets of geometric parameters.

Specifically, each sidechain bead of the 20 amino acids has a distinct hard sphere diameter

(effective van der Waals radius) and has distinct sidechain-to-backbone distances (R-Cα, R-NH,

R-CO). The potential function between two amino acid sidechain beads is modeled as a square well potential which contains two parameters, square well width and square well depth. We reduce the total number of the 210 independent square well depths between 20 different amino acids to

19 parameters while maintaining 210 independent square well widths to ensure physically meaningful pair interaction energies and reasonably accurate results in discriminating decoys from native structures in the PDB database. The backbone hydrogen bonding interactions are also modeled as a directional square well potential. All the other non-bonded interactions are modeled as hard sphere interactions. A detailed description of the derivation of the geometric and energetic parameters of the PRIME20 model is given in our earlier work. 24; 39, 40 The main difference between our force field and most of the atomistic (Amber, OPLS, CHARMM, etc.) or coarse- grained protein force fields (MARTINI41, OPEP42 and AWSEM43) is that PRIME20 models the

92

non-bonded interactions (backbone hydrogen bonding and sidechain-sidechain interactions) as discontinuous potentials while other models use continuous potentials, like the Lennard-Jones or

Columbic potentials. The square well potential for modeling sidechain-sidechain interaction in

PRIME20 is a sum of the van der Waals, charge-charge and hydrophobic interactions, and each of these contributions cannot be split separately. In addition, PRIME20 does not take into account the effect of water/counter-ions as it is an implicit solvent model. Instead the PRIME20 forced field is focused on predicting the structural changes of peptides in their oligomeric and fibrillar state.

3.4.2 Simulation Procedure

We simulate the seeding and cross-seeding aggregation of the Hu, BV and SHaPrP(120-

144) peptides in the presence of a pre-formed β-sheet-rich protofilament. The S-, Ω- and U-shaped protofilaments formed by spontaneous aggregation of eight Hu, BV and SHaPrP(120-144) peptides in the previous work 32 are used as seeds. The simulations are performed in the canonical ensemble (fixed number of particles, constant volume and temperature). The Andersen thermostat is implemented to maintain the system at a constant temperature. 44 The initial velocity for every bead in the system is generated based on a Maxwell-Boltzmann distribution centered at the desired reduced temperature. The system is initialized by randomly placing a pre-formed β-sheet-rich octamer surrounded by eight monomeric peptides in a cubic box with box length equal to 110.0 Å corresponding to a total peptide concentration of 20 mM. The reduced temperature is defined to

* be T = kBT/εHB, where the hydrogen bonding energy, εHB=12.47kJ/mol. The reduced temperature

T* is chosen to be 0.195, which corresponds to 330K in real temperature 32. We have five

93

independent runs for each of the fourteen seeding and cross-seeding simulation. Each simulation lasts for at least 75μs.

3.5 Acknowledgements

This work was supported by National Institutes of Health Grant R01 EB006006 and in part by National Science Foundation Research Triangle Materials Research Science and Engineering

Centers Grant DMR-1121107. The authors declare that they have no conflicts of interest.

94

3.6 References

1. Prusiner SB (1991) MOLECULAR-BIOLOGY OF PRION DISEASES. Science

252:1515-1522.

2. Miller G (2009) Could They All Be Prion Diseases? Science 326:1337-1339.

3. Andréoletti O, Lacroux C, Chabert A, Monnereau L, Tabouret G, Lantier F, Berthon P,

Eychenne F, Lafond-Benestad S, Elsen J-M, Schelcher F (2002) PrPSc accumulation in

placentas of ewes exposed to natural scrapie: influence of foetal PrP genotype and effect

on ewe-to-lamb transmission. J Gen Virol 83:2607-2616.

4. Soto C, Estrada L, Castilla J (2006) Amyloids, prions and the inherent infectious nature of

misfolded protein aggregates. Trends In Biochemical Sciences 31:150-155.

5. Walker LC, Jucker M. Neurodegenerative Diseases: Expanding the Prion Concept. In:

Hyman SE, Ed. (2015) Annu Rev Neurosci, Vol 38. pp. 87-103.

6. Pan KM, Baldwin M, Nguyen J, Gasset M, Serban A, Groth D, Mehlhorn I, Huang ZW,

Fletterick RJ, Cohen FE, Prusiner SB (1993) CONVERSION OF ALPHA-HELICES

INTO BETA-SHEETS FEATURES IN THE FORMATION OF THE SCRAPIE PRION

PROTEINS. Proc Natl Acad Sci U S A 90:10962-10966.

7. Alper T, Cramp WA, Haig DA, Clarke MC (1967) Does the Agent of Scrapie Replicate

without Nucleic Acid? Nature 214:764-766.

8. Jarrett JT, Lansbury Jr PT (1993) Seeding “one-dimensional crystallization” of amyloid: A

pathogenic mechanism in Alzheimer's disease and scrapie? Cell 73:1055-1058.

9. Prusiner SB (1997) Prion diseases and the BSE crisis. Science 278:245-251.

10. Parchi P, Giese A, Capellari S, Brown P, Schulz-Schaeffer W, Windl O, Zerr I, Budka H,

Kopp N, Piccardo P, Poser S, Rojiani A, Streichemberger N, Julien J, Vital C, Ghetti B,

95

Gambetti P, Kretzschmar H (1999) Classification of sporadic Creutzfeldt-Jakob disease

based on molecular and phenotypic analysis of 300 subjects. Ann Neurol 46:224-233.

11. Bruce ME, Will RG, Ironside JW, McConnell I, Drummond D, Suttie A, McCardle L,

Chree A, Hope J, Birkett C, Cousens S, Fraser H, Bostock CJ (1997) Transmissions to mice

indicate that /`new variant/' CJD is caused by the BSE agent. Nature 389:498-501.

12. Morales R, Moreno-Gonzalez I, Soto C (2013) Cross-Seeding of Misfolded Proteins:

Implications for Etiology and Pathogenesis of Protein Misfolding Diseases. PLoS Pathog

9:e1003537.

13. Sharma A, Bruce KL, Chen BX, Gyoneva S, Behrens SH, Bommarius AS, Chernoff YO

(2016) Contributions of the Prion Protein Sequence, Strain, and Environment to the Species

Barrier. J Biol Chem 291:1277-1288.

14. Vanik DL, Surewicz KA, Surewicz WK (2004) Molecular basis of barriers for interspecies

transmissibility of mammalian prions. Mol Cell 14:139-145.

15. Lee LYL, Chen RPY (2007) Quantifying the sequence-dependent species barrier between

hamster and mouse prions. J Am Chem Soc 129:1644-1652.

16. Chuang CC, Liao TY, Chen EHL, Chen RPY (2013) How do amino acid substitutions

affect the amyloidogenic properties and seeding efficiency of prion peptides. Amino Acids

45:785-796.

17. Apostol MI, Wiltzius JJW, Sawaya MR, Cascio D, Eisenberg D (2011) Atomic Structures

Suggest Determinants of Transmission Barriers in Mammalian Prion Disease.

Biochemistry 50:2456-2463.

18. Nguyen PH, Li MS, Stock G, Straub JE, Thirumalai D (2007) Monomer adds to preformed

structured oligomers of Aβ-peptides by a two-stage dock–lock mechanism. Proc Natl Acad

96

Sci U S A 104:111-116.

19. O’Brien EP, Okamoto Y, Straub JE, Brooks BR, Thirumalai D (2009) Thermodynamic

Perspective on the Dock-Lock Growth Mechanism of Amyloid Fibrils. J Phys Chem B

113:14421-14430.

20. Han W, Schulten K (2014) Fibril Elongation by Aβ17–42: Kinetic Network Analysis of

Hybrid-Resolution Molecular Dynamics Simulations. J Am Chem Soc 136:12450-12460.

21. Qi R, Luo Y, Wei G, Nussinov R, Ma B (2015) Aβ “Stretching-and-Packing” Cross-

Seeding Mechanism Can Trigger Tau Protein Aggregation. J Phys Chem Lett 6:3276-3282.

22. Zhang M, Hu R, Liang G, Chang Y, Sun Y, Peng Z, Zheng J (2014) Structural and Energetic

Insight into the Cross-Seeding Amyloid Assemblies of Human IAPP and Rat IAPP. J Phys

Chem B 118:7026-7036.

23. Hu R, Zhang M, Chen H, Jiang B, Zheng J (2015) Cross-Seeding Interaction between β-

Amyloid and Human Islet Amyloid Polypeptide. ACS Chem Neurosci 6:1759-1768.

24. Nguyen HD, Hall CK (2004) Molecular dynamics simulations of spontaneous fibril

formation by random-coil peptides. Proc Natl Acad Sci U S A 101:16180-16185.

25. Cheon M, Chang I, Hall CK (2011) Spontaneous Formation of Twisted A beta(16-22)

Fibrils in Large-Scale Molecular-Dynamics Simulations. Biophys J 101:2493-2501.

26. Latshaw DC, Cheon M, Hall CK (2014) Effects of Macromolecular Crowding on Amyloid

Beta (16-22) Aggregation Using Coarse-Grained Simulations. J Phys Chem B 118:13513-

13526.

27. Wagoner VA, Cheon M, Chang I, Hall CK (2011) Computer simulation study of amyloid

fibril formation by palindromic sequences in prion peptides. Proteins Struct Funct Bioinf

79:2132-2145.

97

28. Cheon M, Chang I, Hall CK (2012) Influence of temperature on formation of perfect tau

fragment fibrils using PRIME20/DMD simulations. Protein Sci 21:1514-1527.

29. Cheon M, Hall CK, Chang I (2015) Structural Conversion of A beta(17-42) Peptides from

Disordered Oligomers to U-Shape Protofilaments via Multiple Kinetic Pathways. PLoS

Comput Biol 11:e1004258.

30. Latshaw DC, Hall CK (2015) Effects of Hydrophobic Macromolecular Crowders on

Amyloid beta (16-22) Aggregation. Biophys J 109:124-134.

31. Wang Y, Latshaw DC, Hall CK (2017) Aggregation of Aβ(17–36) in the Presence of

Naturally Occurring Phenolic Inhibitors Using Coarse-Grained Simulations. J Mol Biol

429:3893-3908.

32. Wang Y, Shao Q, Hall CK (2016) N-terminal Prion Protein Peptides (PrP(120–144)) Form

Parallel In-register β-Sheets via Multiple Nucleation-dependent Pathways. J Biol Chem

291:22093-22105.

33. Jones EM, Surewicz WK (2005) Fibril conformation as the basis of species- and strain-

dependent seeding specificity of mammalian prion amyloids. Cell 121:63-72.

34. Makarava N, Ostapchenko VG, Savtchenko R, Baskakov IV (2009) Conformational

Switching within Individual Amyloid Fibrils. J Biol Chem 284:14386-14395.

35. Morales R, Abid K, Soto C (2007) The prion strain phenomenon: Molecular basis and

unprecedented features. BBA Mol Basis Dis 1772:681-691.

36. Watts JC, Giles K, Patel S, Oehler A, DeArmond SJ, Prusiner SB (2014) Evidence That

Bank Vole PrP Is a Universal Acceptor for Prions. PLoS Pathog 10:e1003990.

37. Cohen SIA, Vendruscolo M, Dobson CM, Knowles TPJ (2011) Nucleated polymerization

with secondary pathways. II. Determination of self-consistent solutions to growth

98

processes described by non-linear master equations. J Chem Phys 135:065106.

38. Alder BJ, Wainwright TE (1959) STUDIES IN MOLECULAR DYNAMICS .1.

GENERAL METHOD. J Chem Phys 31:459-466.

39. Smith AV, Hall CK (2001) alpha-helix formation: Discontinuous molecular dynamics on

an intermediate-resolution protein model. Proteins Struct Funct Genet 44:344-360.

40. Cheon M, Chang I, Hall CK (2010) Extending the PRIME model for protein aggregation

to all 20 amino acids. Proteins Struct Funct Bioinf 78:2950-2960.

41. Monticelli L, Kandasamy SK, Periole X, Larson RG, Tieleman DP, Marrink SJ (2008) The

MARTINI coarse-grained force field: Extension to proteins. J Chem Theory Comput

4:819-834.

42. Chebaro Y, Pasquali S, Derreumaux P (2012) The Coarse-Grained OPEP Force Field for

Non-Amyloid and Amyloid Proteins. J Phys Chem B 116:8741-8752.

43. Davtyan A, Schafer NP, Zheng WH, Clementi C, Wolynes PG, Papoian GA (2012)

AWSEM-MD: Protein Structure Prediction Using Coarse-Grained Physical Potentials and

Bioinformatically Based Local Structure Biasing. J Phys Chem B 116:8494-8503.

44. Andersen HC (1980) Molecular dynamics simulations at constant pressure and/or

temperature. J Chem Phys 72:2384-2393.

99

Figure 3.1. S-shaped (A), Ω-shaped (C), and U-shaped (E) protofilaments formed by HuPrP(120- 144) from the 5th, 12th, and 11th runs are shown. B, D, and F, schematic representations of peptide conformations in the S-, Ω- and U-shaped protofilaments. The sizes of the side chain beads do not reflect the actual radii. Hydrophobic residues (white), positively charged residues (red), negatively charged residues (blue), and polar residues (green) are shown. (This figure is reproduced from ref 32)

100

Figure 3.2. (A) Plot of β-sheet percentage in HuPrP(120-144) aggregation simulations in the presence and absence of a pre-formed HuPrP(120-144) seed as a function of time. The percentage of β-sheet content is calculated by the STRIDE algorithm in the visual molecular dynamics (VMD) software. (B) and (C) are simulation snapshots of the non-seeding aggregation taken at t = 1μs and 198μs, respectively. (D) and (E) are simulation snapshots of the seeding aggregation taken at t = 1μs and 58μs, respectively.

101

Figure 3.3. (A) Potential energy versus time of HuPrP(120-144) peptides aggregating in the presence of U-, S-, Ω-shaped Hu seeds. (B) Average sidechain interaction energy per peptide of U-, S-, Ω-shaped HuPrP(120-144) protofilaments (seeds). (C) Potential energy versus time of Hu, BV and ShaPrP(120-144) peptides aggregating in the presence of S-shaped Hu, Ω-shaped BV and Ω-shaped Sha seeds. (D) Average sidechain interaction energy per peptide of S-shaped Hu, Ω- shaped BV and Ω-shaped Sha seeds.

102

Figure 3.4. A Plot of potential energy vs simulation time of HuPrP(120-144) peptides aggregation in the presence of S-shaped Hu, Ω-shaped BV and Ω-shaped SHa seed. B-D. Plots of potential energy vs simulation time of Hu, BV, SHaPrP(120-144) peptides aggregation in the presence of S-shaped Hu, Ω-shaped BV and Ω-shaped SHa seed.

103

Figure 3.5. Plots of the average side chain interaction energy (kJ/mol) experienced by residues 138, 139, and 143 of monomer Hu, BV and SHaPrP(120-144) peptides in the presence of (A) S- shaped Hu, (B) Ω-shaped BV and (C) Ω-shaped SHa seeds, respectively.

104

Figure 3.6. Five types of structural mismatches found in homogeneous and heterogeneous protofilaments formed by Hu, BV and SHaPrP(120-144) peptides. Simulation snapshot of (A) Hu with the Ω-shaped SHa protofilament from the 2nd run; (B) BV with the Ω-shaped SHa protofilament from the 3rd run; (C) BV with the Ω-shaped SHa protofilament from the 4th run; (D) SHa with the Ω-shaped BV protofilament from the 2nd run; (E) SHa with the Ω-shaped SHa protofilament from the 1st run. Hu, BV and SHaPrP(120-144) peptides in A-E are colored in red, blue and green respectively, except that the SHa(120-144) monomer peptides in E are colored in purple.

105

Table 3.1. Number of HuPrP(120-144) peptides that adopt the peptide conformation in the seed or other conformations during the aggregation in the presence of the S-, Ω- and U-shaped Hu seeds.

Seed # Hu monomers that # Hu monomers that # Hu monomers that adopt seed partially adopt seed adopt other conformation conformation conformations Hu S-shaped 4.0±2.7 1.2±1.6 2.8±1.9 Hu Ω-shaped 5.6±1.8 1.4±1.1 1.2±1.8 Hu U-shaped 6.0±2.1 0.0±0.0 2.0±2.1

106

Table 3.2. Number of Hu, BV and SHaPrP(120-144) peptides that adopt the peptide conformation in the seed during the seeding and cross-seeding aggregation.

Seed/Monomer peptide # Hu monomers that # BV monomers that # SHa monomers adopt seed adopt seed that adopt seed conformation conformation conformation Hu S-shaped seed 4.0±2.7 7.0±0.7 4.8±2.4 BV Ω-shaped seed 6.0±0.8 6.2±1.3 6.4±1.3 SHa Ω-shaped seed 2.8±1.9 3.2±2.6 3.0±2.5

107

Table 3.3. Amino acid sequence of three prion protein fragments.

120 130 140 144 Human A V V G G L G G Y M L G S A M S R P I I H F G S D Bank Vole A V V G G L G G Y M L G S A M S R P M I H F G N D Syrian Hamster A V V G G L G G Y M L G S A M S R P M M H F G N D

108

3.7 Supporting Information

Figure S3.1. (A) The simulation snapshot and (B) the schematic peptide representations of the Ω- shaped BVPrP(120-144) protofilaments are shown. (C) The simulation snapshot and (D) the schematic peptide representation of the Ω-shaped SHaPrP(120-144) protofilaments are shown. The sizes of the side chain beads do not reflect the actual radii. Hydrophobic residues (white), positively charged residues (red), negatively charged residues (blue), and polar residues (green) are shown. (this figure is reproduced from ref 1)

109

-2500

-3000

-3500

-4000

-4500

-5000 Potential energy (KJ/mol) energy Potential -5500 0 10000 20000 30000 40000 Simulation time (ns)

Figure S3.2. Plots of potential energy vs simulation time of BV (blue) and SHaPrP(120-144) (red) peptides aggregating in the presence of Ω-shaped HuPrP(120-144) seed.

110

References

1. Wang Y, Shao Q, Hall CK (2016) N-terminal Prion Protein Peptides (PrP(120–144)) Form

Parallel In-register β-Sheets via Multiple Nucleation-dependent Pathways. J Biol Chem

291:22093-22105.

111

CHAPTER 4

Aggregation of Aβ(17-36) in the Presence of Naturally Occurring Phenolic Inhibitors Using

Coarse-Grained Simulations

Chapter 4 is essentially a manuscript by Yiming Wang, David C. Latshaw and Carol K. Hall published on Journal of Molecular Biology.

112

Aggregation of Aβ(17-36) in the Presence of Naturally Occurring Phenolic Inhibitors Using Coarse-Grained Simulations

Yiming Wang, David C. Latshaw and Carol K. Hall*

Department of Chemical and Biomolecular Engineering, North Carolina State University, Raleigh, North Carolina 27695-7905, United States

Abstract

Although some naturally-occurring polyphenols have been found to inhibit amyloid β fibril formation and reduce neuron cell toxicity in vitro, their exact inhibitory mechanism is unknown.

In this work, discontinuous molecular dynamics (DMD) combined with the PRIME20 force field and a newly-built inhibitor model are performed to examine the effect of vanillin, resveratrol, curcumin and epigallocatechin-3-gallate (EGCG) on the aggregation of Aβ(17-36) peptides. Four sets of peptide/inhibitor simulations are performed in which inhibitors: (1) bind to Aβ(17-36) monomer; (2) interfere with Aβ(17-36) oligomerization; (3) disrupt a pre-formed Aβ(17-36) protofilament; (4) prevent the growth of Aβ(17-36) protofilament. The single-ring compound, vanillin, slightly slows down but cannot inhibit the formation of a U-shaped Aβ(17-36) protofilament. The multiple-ring compounds, EGCG, resveratrol and curcumin, redirect Aβ(17-

36) from a fibrillar aggregate to an unstructured oligomer. The three aromatic groups of the EGCG molecule are in a stereo (nonplanar) configuration, helping it contact the N-terminal, middle and

C-terminal regions of the peptide. Resveratrol and curcumin bind only to the hydrophobic residues near peptide termini. The rank order of inhibitory effectiveness of Aβ(17-36) aggregation is :

EGCG > resveratrol > curcumin > vanillin, consistent with experimental findings on inhibiting full length Aβ fibrillation. Furthermore, we learn that the inhibition effect of EGCG is specific to the

113

peptide sequence while those of resveratrol and curcumin are non-specific in that they stem from strong interference with hydrophobic sidechain association, regardless of the residues’ location and peptide sequence. Our studies provide molecular-level insights into how polyphenols inhibit

Aβ fibril formation, knowledge that could be useful for designing amyloid inhibitors.

114

4.1 Introduction

Alzheimer’s Disease (AD) is a neurodegenerative disease that causes dementia, nervous system degradation, and death. Currently there are no therapeutic agents available for the treatment of AD despite great efforts from the medical research community. [1-3] The possible reasons for such inefficient drug development have been discussed by Doig et al. [4] The pathological hallmark of AD is the aggregation of the amyloid β (Aβ) peptide. Aβ monomers, which result from the enzyme-based cleavage of the amyloid precursor protein, APP, assemble extracellularly into progressively larger aggregate structures, monomers-to-oligomers-to-fibrils, and are eventually deposited in plaques in the brain. One possible avenue for AD treatment is to prevent or reverse aggregation of the Aβ peptide through interaction with small molecule aggregation inhibitors.

Many small molecules have been identified [5] or designed [6] for this purpose. Recent progress in experimental and computational investigations of Aβ/drug interaction are summarized by

Nguyen et al. [7] and Doig et al.[8]. A subset of these molecules, phenolic compounds, have been shown to be effective at preventing the formation of Aβ amyloidogenic structures in vitro. [9-11]

Many researchers have investigated the effectiveness of vanillin [5, 12], resveratrol [13-

20], curcumin [5, 12, 18, 19, 21-25, 26 , 27], and EGCG [18, 28-39] as AD therapeutics. Vanillin is the principal component found in vanilla bean extract. It is much smaller than the other phenolic compounds considered in this paper and it shares aromatic substitution patterns with curcumin.

The effectiveness of vanillin as an Aβ aggregation inhibitor is less clear than that of the other compounds considered in this paper, but since it only has one aromatic group, it is included to determine the impact of the number of aromatic inhibitor groups on inhibiting Aβ aggregation.

Necula et al. [5] identified vanillin as being capable of disrupting both Aβ(1-42) oligomer and

115

fibril formation, while Ladiwala et al. [38] and Reinke and Gestwicki [12] both indicate that vanillin has little-to-no effect on Aβ(1-42) aggregation.

Resveratrol is a naturally-occurring compound found in the skin of red grapes and is often consumed in the form of red wine. Although its concentration in red wine is about ten times lower than the amount shown to inhibit Aβ aggregation in vitro [17], Wang et al. showed that moderate consumption of Cabernet Sauvignon could promote the clearance of Aβ in transgenic mice. [20]

Ladiwala et al. found that resveratrol could remodel Aβ(1-42) monomers, soluble oligomers, non- toxic oligomers, fibrillar intermediates, and amyloid fibrils in vitro to form off-pathway, disordered, non-toxic, high molecular weight structures. [38] They suggested that resveratrol disassembles fibrillar structures by interacting with their β-sheets and stacked aromatic side chains.

Feng et al. tested the effects of resveratrol on Aβ(1-42) aggregation in vitro and found that it could inhibit Aβ fibril formation, disaggregate pre-formed fibrils, and inhibit oligomer cytotoxicity, but that it could not inhibit the formation of the oligomers themselves, which were instead stabilized.

[19]

Curcumin is one of the principal components of the spice turmeric, known for its potent antioxidant and anti-inflammatory properties. It has been shown to disrupt Aβ aggregation by direct interaction. Ono et al. found that curcumin inhibits Aβ(1-40) and Aβ(1-42) fibril formation and destabilizes pre-formed fibrils. [24] They were unsure of the inhibition mechanism but proposed that curcumin might bind to the ends of fibrils, destabilizing the structure and preventing further aggregation. Yang et al. studied curcumin’s effects on Aβ and determined that it could inhibit formation of Aβ(1-40) oligomers and fibrils, and prevent Aβ(1-42) oligomer formation.

[25] Reinke and Gestwicki examined the effect of the number of aromatic groups, aromatic group substitution patterns, and the length and flexibility of the chain linking each aromatic group on

116

Aβ(1-42) aggregation. [12] They found that at least two aromatic groups, each of which contains functional groups capable of hydrogen bonding, were required to inhibit Aβ aggregation. They also found that the linker between two aromatic groups should have a length of 6Å-19Å, and should contain one or two sp3 hybridized carbons. These results put curcumin squarely into their

“ideal” Aβ aggregation inhibitor range. Simulations have also been performed to examine curcumin binding to Aβ monomers and aggregate structures. Ngo and Li found from docking simulations that curcumin preferentially binds to the N-terminal amino acid D1 of Aβ(1-40). [26]

In an atomistic MD simulation of curcumin interacting with the Aβ(9-40) fibril (hexamer) with two different structures, they found that curcumin bound to S26 and K28 on an Aβ(9-40) fibril having two loop structures while curcumin bound to F20 on an Aβ(9-40) fibril having three loop structures. Zhao et al. used atomistic molecular dynamics simulations to examine the effects of curcumin on the stability of an Aβ(1-42) dimer. [27] They found that curcumin decreased the β- sheet content of Aβ oligomers but did not alter the inter-peptide contacts or break the dimer apart.

Epigallocatechin-3-gallate (EGCG) is an abundant molecule in green and white tea, with smaller amounts occurring in black tea. Like curcumin and resveratrol, EGCG is known for its potent antioxidant properties. EGCG is structurally different from the other phenolic compounds considered here in that it contains three aromatic groups and does not have a flexible carbon chain linker, making it more rigid. Bieschke et al. studied EGCG interactions with Aβ(1-42) and found that it can bind monomers and force them off pathway into large, disordered, non-toxic aggregates.

[29] Their experiments showed that EGCG has the capability to completely dissolve fibrils, and to remodel them without completely disassembling them. Similar results were obtained by

Ehrnhoefer et al. [32] and Bastianetto et al. [33] Wang et al. used isothermal titration calorimetry to understand how EGCG binds to Aβ(1-42). [35] Their major result was that at low EGCG/Aβ(1-

117

42) ratios, the inter-species interactions are primarily due to both hydrogen bonding and hydrophobic interactions but as the amount of EGCG increases the interactions shift toward hydrophobic interactions. Young et al. used ESI-IMS-MS to find that EGCG binds specifically to

Aβ(1-40) monomer leading to the formation of amorphous aggregates. [37] Atomistic simulations have also been used to probe the molecular level details of EGCG interaction with Aβ. Liu et al. found that non-polar interactions accounted for almost three quarters of the EGCG-Aβ(1-42) binding energy and that the remaining energy was due to hydrogen bonding. [34] The residues with the strongest EGCG interaction were F4, R5, F19, F20, E22, K28, G29, L34, G37, and I41.

Atomistic simulations of Aβ(1-42) dimer interacting with EGCG by Zhang et al. showed that

EGCG decreases β-sheet content and increases coil and α-helix secondary structure in the Aβ(1-

42) dimer. [36]

Studies have been conducted that directly compare the effectiveness of these small molecules. Chebaro et al. used docking simulations to show that resveratrol, curcumin and EGCG, among other inhibitors, have good binding affinity to Aβ(17-42) trimers, most specifically residues

17-21. [18] Resveratrol and curcumin both preferred to bind to the lowest energy Aβ oligomer tested, while EGCG preferred to bind with high-energy structures. This may mean that the aggregation inhibition from EGCG stems from binding early intermediate structures rather than later stable ones. Experiments by Reinke and Gestwicki showed that vanillin was significantly less effective than resveratrol and curcumin at inhibiting Aβ aggregation. [12] Feng et al. found curcumin to be effective, but not as effective as resveratrol at inhibiting fibril formation. [19]

Finally, Rajasekhar et al. [40] reported that EGCG has a much lower half maximal inhibition concentration (IC50≈2.4μm) than curcumin (IC50≈13.3μm) and resveratrol (IC50≈15.1μm) on

Aβ aggregation.

118

Researchers also found that different phenolic molecules may have distinct inhibition mechanisms. Necula et al. [5] found that inhibitors, e.g. curcumin, can inhibit Aβ oligomerization but don’t inhibit fibrillation while others can inhibit both Aβ oligomerization and fibrillation, or only fibrillation. Young et al. [37] used ion mobility spectrometry-mass spectrometry to classify small molecule inhibitors for Aβ(1-40) and hIAPP into: 1. specific inhibition (binding to peptide monomer/dimer with a fixed stoichiometry); 2. non-specific inhibition (binding to peptide monomer/dimer with a wide range of stoichiometries); 3. colloidal inhibition (inhibitor self- associates to form colloid).

In the present study, we take Aβ(17-36), a central fragment of Aβ(1-42), as our peptide model. This is different from the non-toxic Aβ fragments studied previously by our group and by others in investigating Aβ-inhibitor interactions, e.g. Aβ(12-28) [41], Aβ(25-35) [42, 43] and

Aβ(17-42) [18]. We choose this peptide because it is short enough to enable our simulation of spontaneous fibril formation but long enough to mimic full length Aβ(1-42) in many aspects. [44]

An even shorter fragment, Aβ(18-35), forms fibrils in vitro with a characteristic U-shape which is similar to those formed by Aβ(1-40) as shown by Chandrakesan et al. [45] using solid-state

NMR. Additionally, residues 17-36 consist of the central region of Aβ(1-42) peptide where small molecules frequently bind. Zhu et al. used a fragment-based mapping calculation from replica- exchange molecular dynamics (REMD) and found that curcumin and Congo red favorably bind to the central residues of the Aβ(1-42) monomer (e.g. Phe4, Tyr10, Leu17, Phe19, Ile31 and Met35).

[46]

In this work, we apply discontinuous molecular dynamics (DMD) simulations combined with a coarse-grained model for peptide and inhibitor to investigate the effect of the four naturally- occurring phenolic molecules (EGCG, resveratrol, curcumin and vanillin) on Aβ(17-36)

119

aggregation. We constructed an intermediate-resolution coarse-grained inhibitor model that is compatible with PRIME20, a coarse-grained peptide model developed by the Hall group. Five types of simulations were performed to give a full picture of how small molecules affect Aβ(17-

36) fibrillation, including: (1). 8 Aβ(17-36) peptides aggregating in the absence of small molecules; (2). a single Aβ(17-36) peptide interacting with 10 small molecules; (3). 8 Aβ(17-36) peptides aggregating in the presence of 30 small molecules; (4). a pre-formed Aβ(17-36) protofilament (octamer) interacting with 30 small molecules; (5). a pre-formed Aβ(17-36) protofilament elongating in the presence of 8 Aβ(17-36) peptides and 30 small molecules.

Highlights of our results are the following: In the peptide-only simulations, 8 initially disordered Aβ(17-36) peptides form a U-shaped fibrillar octamer via a templated-growth mechanism. In the simulation of small molecules binding to a single-peptide, resveratrol has the strongest binding energy with aromatic amino acids, e.g. F19 on Aβ(17-36), while EGCG has the highest number of contacts with the whole peptide. The four inhibitor molecules most frequently contact the hydrophobic residues F19, F20, L34, M35 on the N- and C- terminals of Aβ(17-36).

In the simulation of 8 peptides aggregating with 30 inhibitor molecules, EGCG, resveratrol and curcumin prevent Aβ(17-36) fibril formation and form disordered peptide-inhibitor oligomers while vanillin doesn’t inhibit the formation a U-shaped Aβ(17-36) protofilament. In the simulation of one protofilament interacting with 30 inhibitors, EGCG and resveratrol redirect the pre-formed protofilament into disordered oligomer while curcumin and vanillin bind but do not dissolve the protofilament structure. In the simulation of one protofilament elongating with 8 peptides and 30 inhibitors, vanillin and curcumin slow the elongation rate of the Aβ(17-36) protofilament; resveratrol stops its growth while EGCG dissolves the pre-formed protofilament. In summary,

EGCG is the most effective inhibitor for Aβ(17-36) fibril formation and the rank order of

120

effectiveness of the four candidates is: EGCG > resveratrol > curcumin > vanillin, which is consistent with experimental findings on full length Aβ fibrillation. [12, 40]

4.2 Results

4.2.1 Aβ(17-36) form a parallel in-register U-shaped protofilament

First, we performed 6 independent simulations of spontaneous aggregation of 8 Aβ(17-36) peptides at T* = 0.2 and peptide concentration of 5mM. The U-shaped Aβ(17-36) protofilament is formed in 5 out of 6 runs. Fig. 1 shows simulation snapshots from the 1st run, starting from a random configuration in Fig. 1A. The secondary structure content of the peptides at all time points are determined by using the STRIDE algorithm [47] within the VMD software [48]. In Fig. 1B, three peptides form a β-sheet-rich hairpin-shaped trimer. In Figs. 1B-E, the trimer grows into a U- shaped octamer by acting as a template and recruiting other random coil monomers to the two ends. The G25 in the middle of the peptide backbone serves as the hinge of the turn. [49] The inter- and intra-peptide hydrophobic sidechain associations between 19Phe and 32Ile and 34Leu lead to the formation of a tightly-packed hydrophobic core between the N- and C- terminals of the peptides, as shown in Fig. 1F. The salt bridge interaction between D23 and K28 helps stabilize the

U-shaped protofilament. Fig. 1G shows the decrease of potential energy over the course of the simulation. Formation of a parallel in-register U-shaped protofilament by Aβ(17-36) in our simulation is consistent with the experimental finding that Aβ(18-35) peptides form fibrils with a hairpin characteristic. [45]

121

4.2.2 Inhibitors bind to a single Aβ(17-36) peptide

We examined the binding affinity of the four inhibitor molecules to Aβ(17-36) peptide by simulating an Aβ(17-36) monomer interacting with 10 inhibitor molecules. The binding affinity of phenolic molecules to target peptide is a good indication of their potential inhibitory effect. The average number of inhibitor molecules in contact with Aβ(17-36) monomer over the last 25 billion collisions was calculated. The inhibitor molecule is defined as contacting the peptide if there is at least one hydrogen bond or one hydrophobic contact between the inhibitor and the peptide. The results are: NEGCG-Aβ17-36 = 4.29±0.83, Nresveratrol-Aβ17-36 = 3.70±0.72, Ncurcumin-Aβ17-36 = 3.22±0.93,

Nvanillin-Aβ17-36 = 0.91±0.86, indicating that the order of binding affinity of the four inhibitor molecules to Aβ(17-36) is: EGCG > resveratrol > curcumin > vanillin. In addition, we calculated the RMSD of the four inhibitor molecules in both atomistic and coarse-grained simulation as shown in Figs. 2A and B, respectively. We found that curcumin has the largest structural fluctuation among the four molecules in both the atomistic and coarse-grained simulations while the other three molecules are relatively rigid. This is because curcumin has two aromatic groups separated by a long and flexible carbon linker, resulting in large conformation fluctuations.

To find the specific binding sites on Aβ(17-36) for the four inhibitors, we calculated the number of contacts (hydrophobic contact or hydrogen bond) formed between the Aβ(17-36) backbone spheres and the inhibitors. Fig. 2C shows that EGCG has the highest number of contacts with Aβ(17-36) and that resveratrol and curcumin specifically contact the hydrophobic residues near the N- and C- terminals, including 17L, 19F, 20F, 21A, 31I, 32I, 34L and 35M. Vanillin has the least number of contacts with the peptide. Our results for the average number of contacts between small molecules and each residue of Aβ(17-36) (Fig. 2C) are comparable with those from a similar atomistic MD simulation study of Aβ(17-42)/inhibitor interactions.[18]

122

The π-π stacking interaction between polyphenols and aromatic sidechains on amyloidogenic peptides is considered to be key in inhibiting amyloid formation.[11] We thus calculated the radial distribution function, g(r), between the sidechain centroid of the F19 residue on Aβ(17-36) and the centers of mass of the four inhibitors as shown in Fig. 2D. Each g(r) has a peak at a peptide-inhibitor distance equal to around 6Å. The breadth of the peak in each curve indicates that the four inhibitors contact the F19 sidechain of Aβ(17-36) in many different modes.

For example, hydrogen bonds could form between the inhibitor CG groups (O-H, C=O, and O) and the peptide CG groups (C=O, N-H) or there could be hydrophobic interactions between the inhibitor CG sites (C, C-H, C-H2) and the sidechains of Aβ(17-36) peptide. From Fig. 2D, we also find that resveratrol binds to F19 with a higher probability than EGCG, although EGCG on average has higher total number of contacts with residues along the whole peptide than resveratrol.

4.2.3 Inhibitors interfere with Aβ(17-36) aggregation

We next perform DMD simulations of 8 Aβ(17-36) peptides aggregating in the presence of 30 inhibitors to evaluate the ability of the four molecules to inhibit peptide aggregation. In the monomer peptide binding simulations, EGCG has the highest Aβ(17-36)/inhibitor binding ratio

(1:4) among the four molecules. Thus the Aβ(17-36)/inhibitor molar ratio is chosen to be around

1:4 to ensure that there are enough inhibitor molecules in the simulation to interact with each peptide. Snapshots of the vanillin and EGCG cases are shown in Fig. 3. From Fig. 3 (top 3 panels), we find that vanillin doesn’t inhibit the formation of a U-shaped Aβ(17-36) protofilament. In the presence of vanillin, the peptides form an U-shaped β-sheet-rich trimer that binds with a few vanillin molecules, but later recruits other disordered peptides to form a U-shaped octamer.In other words , the aggregation pathway is similar to the case without inhibitors, shown in Fig. 1. In

123

comparison, as shown in Fig. 3(bottom 3 panels), EGCG quickly binds to the Aβ(17-36) monomers and oligomers and redirects them to form a disordered oligomeric peptide-inhibitor complex. Our simulation results indicate that EGCG has a much stronger inhibitory effect on Aβ(17-36) aggregation than vanillin. The inhibition mechanism of EGCG on Aβ(17-36) aggregation is consistent with experimental findings that EGCG inhibits the fibrillation of Aβ(1-42) [32] and

Aβ(1-40) [37] peptides by binding to the monomeric peptides and redirecting them to form amorphous aggregates. Resveratrol and curcumin have similar inhibitory effects on Aβ(17-36) aggregation as EGCG; the simulation snapshots are not shown.

We examine the relative effectiveness of the 4 small molecules in inhibiting Aβ(17-36) aggregation by comparing their ability to disrupt peptide-peptide hydrophobic interactions and hydrogen bonding, two key driving forces for peptide fibrillation. Firstly, we examine the ability of four inhibitor molecules to interfere with Aβ(17-36) backbone hydrogen bond formation. Fig.

4 plots the number of backbone hydrogen bonds formed, a measure of β-sheet content, over the course of the simulation. From Fig. 4, we find that the rank order of the ability to disrupt peptide- peptide hydrogen bonding is: EGCG > resveratrol > curcumin > vanillin. Vanillin slightly decreases the rate of backbone hydrogen bond formation while resveratrol, curcumin and EGCG greatly decrease it. There are some transient increases and decreases of β sheet content in the presence of curcumin while almost no β-sheet content forms in the presence of resveratrol and

EGCG.

We then examine the four inhibitor molecules’ ability to disrupt Aβ(17-36) sidechain- sidechain associations. Fig. 5 plots the (A) sidechain-sidechain and (B) sidechain-inhibitor interaction energy of the residues along the chain over the whole simulation trajectory. From Fig.

5A, we find that the order of effectiveness in interfering with peptide sidechain-sidechain

124

interaction is: EGCG > resveratrol > curcumin > vanillin. Resveratrol, curcumin and EGCG notably decrease the peptide sidechain-sidechain interaction. Vanillin slightly enhances the peptide sidechain-sidechain interaction between residues F20 and residue I31, indicating that in the presence of vanillin, the peptide-peptide interactions are mostly preserved within the protofilament. Fig. 5B shows that the four small molecules mainly interact with the N- and C- terminal residues of Aβ(17-36) over the whole simulation. Our results are consistent with Chebaro et al.’s docking simulation results which showed that resveratrol, curcumin and EGCG have good binding affinity to Aβ(17-42) trimers, most specifically to residues 17-21. [18]

To investigate the structure-function relationship of the four inhibitor molecules on Aβ(17-

36) aggregation, we also mapped the interaction energy between the inhibitor CG sites and Aβ(17-

36) residues in Fig. 6. Figs. 6A and 6B show that vanillin has a much weaker peptide-inhibitor interaction energy than resveratrol. This is because vanilllin has a single aromatic ring and thus has less hydropohbic sites than resveratrol for interfering with sidechain-sidechain hydrophobic association. This difference in peptide-inhibitor interaction strengths may explain why vanillin is much less effective than resveratrol at inhibiting Aβ(17-36) aggregation. Figs. 6B and 6C show that resveratrol and curcumin interact mainly with the hydrophobic residues near the N- and C- terminals because they are structurally similar in that they have two aromatic groups separated by a carbon linker, as shown in Fig. 10. However, resveratrol has a stronger interaction energy with peptides than curcumin, likely because the two aromatic groups on curcumin are further apart than those on resveratrol (curcumin has two ketone groups between two alkene groups). The structural flexibility associated with the long curcumin carbon linker decreases its binding affinity to the peptide, thus decreasing its inhibitory effectiveness, in agreement with Reinke and Gestwicki ’s finding. [12] Fig. 6D shows the OH groups (sites 3, 5 and 7) on the 1st aromatic group (sites 1 to

125

9) of EGCG interact with the middle region of the peptides (residues E22 to G29), and the OH groups (sites 16, 19, 29 and 31) on the 2nd and 3rd aromatic groups (sites 11 to 32) of EGCG mainly contact the N- and C- terminal of the monomer peptide. This is because the three aromatic groups of EGCG are in a stereo (non-planar) configuration, which is structurally different from the other three planar molecules. Thus while resveratrol and curcumin interact with the N- and C- terminii regions of Aβ(17-36), EGCG binds and stabilizes the whole monomeric peptide and prevents inter-peptide contacts.[43]

Next, we calculate the interaction energy between the individual CG groups on four inhibitor molecules and Aβ(17-36) peptide as shown in Fig. 7. Vanillin has a relatively evenly- distributed interaction energy over all of its CG sites except for the H- group. The OH groups on the aromatic rings of resveratrol (sites 4, 13 and 16 ) and curcumin (sites 6 and 20) have relatively strong interactions with the peptide while the interaction energies between the C and CH groups on the two aromatic groups and the peptides are evenly distributed. For EGCG, the two aromatic groups consisting of CG sites 12 to 32 have stronger interaction energy than the aromatic group consisting of CG sites 1 to 10, with the Aβ(17-36) peptides.

4.2.4 Inhibitors redirect pre-formed Aβ(17-36) protofilament into disordered oligomer

We examined the ability of the four inhibitor molecules to perturb a pre-formed fibrillar structure by simulating 30 inhibitors interacting with a pre-formed U-shaped octameric Aβ(17-36) protofilament. Figs. 8A-D are simulation snapshots of EGCG dissolving the pre-formed Aβ(17-

36) protofilament. EGCG initially binds to the hydrophobic residues near the peptide N- and C- terminals in the fibril, similar to the way that they bind to the monomeric peptides. Due to the

126

interference of EGCG with sidechain-sidechain association and backbone hydrogen bonds, the peptides near the two ends of the protofilament gradually become unstructured monomers.

From the peptide-peptide hydrogen bonds dissociation profile (Fig. 8), we find that EGCG and resveratrol are both able to disrupt most of the peptide hydrogen bonds, indicating that they redirect the preformed Aβ(17-36) fibril into unstructured oligomers. In comparison, in the presence of vanillin and curcumin, the total number of peptide hydrogen bonds stays constant throughout the simulation, indicating that the pre-formed fibrillar structure remains unchanged. Further analysis shows (in Fig. S1) that even though EGCG and resveratrol bind to peptides with a similar number of peptide-inhibitor hydrophobic contacts, EGCG causes an obvious decrease in the number of peptide-peptide hydrophobic contacts. This indicates that EGCG disrupts fibrillar structure more efficiently than resveratrol due to its strong ability to interfere with peptide-peptide contacts and stabilize unstructured oligomers.

4.2.5. Inhibitors retard the elongation of the Aβ(17-36) protofilament

Finally, we investigated the effect of the inhibitor molecules on the elongation of an Aβ(17-

36) protofilament. The simulation system includes 8 disordered Aβ(17-36) peptides, a pre-formed

U-shaped Aβ(17-36) protofilament and 30 inhibitors. Fig. 9 plots the total number of inter-peptide hydrogen bonds formed as a function of time. From Fig. 9, we find that vanillin doesn’t inhibit the seeded aggregation of Aβ(17-36). Instead it slightly increases the growth rate of the protofilament.

In the presence of curcumin, only three out of eight monomer peptides are recruited to join the protofilament. In the presence of resveratrol, the total number of hydrogen bonds remains constant, indicating that none of the monomer peptides are recruited into the preformed protofilament. In this case, resveratrol binds mainly to the monomer peptides. There are not enough resveratrol

127

molecules to disrupt the protofilament. In the presence of EGCG, the total number of hydrogen bonds decreases to a small value near zero, indicating that the 30 EGCG not only prevent the elongation of the preformed protofilament but they also redirect the protofilament into a disordered oligomer. These simulation results suggest that at the same inhibitor concentration, EGCG is more effective at inhibiting protofilament growth than resveratrol, curcumin and vanillin.

4.3 Discussion and conclusion

Phenolic compounds including vanillin, resveratrol, curcumin and EGCG are known to inhibit Aβ aggregation but the exact mechanisms by which this is accomplished are unknown. To examine the effect of naturally-occurring polyphenols on Aβ aggregation, we applied DMD simulation combined with the intermediate resolution protein model, PRIME20 and a coarse- grained inhibitor model. The simulation of peptide-only aggregation served as a benchmark. Four sets of peptide/inhibitor simulations were performed in which: small molecules (1) bind to Aβ(17-

36) monomer; (2) interfere with Aβ(17-36) oligomerization; (3) disrupt a pre-formed Aβ(17-36) protofilament; and (4) prevent the growth of Aβ(17-36) protofilament. Our results show that vanillin doesn’t inhibit the ability of Aβ(17-36) to form a protofilament, or dissolve a preformed protofilament structure or prevent the growth of the protofilament. Curcumin redirects the Aβ(17-

36) monomer to form β-sheet-rich oligomer but doesn’t dissolve the pre-formed protofilament.

Resveratrol and EGCG both convert Aβ(17-36) monomeric peptides and preformed protofilament into disordered oligomers.

Even though the simulation concentrations used here are significantly higher than in experiments and the model peptide is a fragment of the Aβ peptide instead of a full-length one, the hierarchy of the four candidates’ inhibitory effectiveness on Aβ aggregation (EGCG > resveratrol

128

> curcumin > vanillin) is consistent with experimental findings. Our simulation shows that the minimum effective Aβ(17-36)-inhibitor molar ratio for EGCG is below 1:2, for resveratrol is around 1:4, and for curcumin is above 1:4 . This indicates that to achieve the same inhibition effect in experiment, the required doses of resveratrol and curcumin need to be at least double that of

EGCG. Vanillin shows no inhibition effect on Aβ(17-36) aggregation. Experiments by Reinke and

Gestwicki [12] showed that vanillin was significantly less effective than resveratrol and curcumin.

Feng et al. [19] found resveratrol to be more effective in inhibiting amyloid β oligomer cytotoxicity than curcumin. Rajasekhar et al. [40] reported that EGCG has a much lower half maximal inhibition concentration than curcumin and resveratrol on Aβ aggregation. However, in general,

Aβ aggregation is strongly concentration dependent. The Aβ concentration in vivo (around μM) is much lower than those used in vitro or in silico (around mM).[4] Thus the effective stoichiometric ratio predicted in our simulation might also be specific to the high concentration conditions and may not apply to in vivo.

The self-association of these four small molecules plays a relatively minor role in their inhibitory effect. We find in our simulation that the average numbers of self-binding events per inhibitor molecule are (NEGCG-EGCG = 0.93±0.12, Nresveratrol-resveratrol = 1.28±0.14, Ncurcumin-curcumin =

0.91±0.14, Nvanillin-vanillin = 0.34±0.09), indicating that they are mostly likely to form dimers but not to further self-associate to form large clusters. In comparison, peptide-inhibitor binding events are more frequent than inhibitor-inhibitor binding events because the number of peptide-inhibitor binding events are three-times that of the self-binding events for all four inhibitors. In addition, as far as we know, these four small molecules do not have strong propensity to self-aggregate into large colloids.In our simulation, although EGCG, resveratrol and curcumin can all inhibit Aβ(17-

36) aggregation, their inhibitory mechanisms are different due to their unique chemical structures.

129

Resveratrol and curcumin seem to fit in the “non-specific” type of small molecule inhibitors, a concept introduced by Young et al. [37] since their inhibitory effect mainly comes from their strong interference with hydrophobic sidechain association, regardless of the residues’ location and peptide sequence. Kim et al. showed a similar Aβ inhibition mechanism where naproxen reduced the interpeptide contacts by binding primarily to hydrophobic interface between the

Aβ(10-40) dimer.[50] Furthermore, the effectiveness of the non-specific inhibitor may depend solely on the fraction of residues in the peptide sequence that are hydrophobic. In comparison,

EGCG seems to belong to the “specific” inhibitor type because its inhibitory effect comes from its ability to bind and thus stabilize the monomeric peptides and thus is sequence-dependent.

Furthermore, we learn from our simulation that the pharmacophore structure and the location of the functional groups of the small molecule have a crucial impact on its inhibitory effectiveness. Our simulation results are consistent with the rules introduced by Reinke and

Gestwicki [12] for designing effective potential small molecule inhibitors of amyloid β. Their first rule says that the inhibitor molecule needs at least two separate aromatic groups for effective inhibition. Thus, even though vanillin and resveratrol both have three hydrogen bonding sites, suggesting that they have similar capacity to interfere with backbone hydrogen bonding interaction, vanillin is a much less effective inhibitor. This is because vanillin only has a single aromatic group which doesn’t provide enough opportunities for strong hydrophobic association with sidechain beads on Aβ(17-36). Their second rule says that the molecule needs to have a rigid conformation and can’t be too flexible. Curcumin is structurally similar to resveratrol but is less effective than resveratrol because the long carbon linker allows considerable structural flexibility, thus decreasing its binding affinity to Aβ(17-36).

130

We add a third rule for designing effective inhibitor molecules for amyloid formation.

Inhibitor molecules should have a stereo (nonplanar) configuration which allows the functional sites on multiple aromatic groups to interact with multiple residues located in different regions of the peptide. EGCG is the most effective inhibitor among the four molecules because the OH groups on its three major aromatic rings contact the N-terminal (residues L17 to A21), middle (residues

E22 to G29) and C-terminal (residues A30 to V36) on Aβ(17-36), which helps stabilize the monomer peptide.

This is our first attempt to incorporate an inhibitor model into the PRIME20 force field.

Currently, the inhibitor and the peptide are represented with different coarse-grained levels. The inhibitor molecule is represented as a mostly united atom model while the peptide is represented with three backbone beads and one sidechain bead. In the future, we will improve the geometry of the peptide model. We plan to adopt multiple-bead sidechain representations to account for the directional π-π interaction between the phenolic molecules and aromatic sidechains on the amino acids, Phe, Trp and Tyr. We will also improve the potential function between inhibitor sites and peptide sidechain sites. Specifically, the energetic parameters between the aromatic carbon sites on inhibitor and peptide sidechain sites will be derived from atomistic simulation. In addition, experimental measurements could be used to further refine the energetic parameters of our force field, including measuring the peptide-molecule binding free energy, the inhibition constant and the IC50 of the small molecule.

131

4.4 Methods

4.4.1 Discontinuous Molecular Dynamics and PRIME20 Force Field

The simulation method we use in this work is discontinuous molecular dynamics (DMD), a fast alternative to traditional molecular dynamics and the PRIME20 force field. [51] In DMD simulations, the only time that the velocities and positions need to be recalculated is when a discontinuity in the potential is encountered. In contrast, with continuous potentials, the simulations are advanced at regularly-spaced time steps. Since DMD advances the simulation on a collision-to-collision basis rather than at small predetermined time steps, the technique allows us to simulate much longer time scales than traditional molecular dynamics. The PRIME20 force field use a four-sphere-per-residue model in which each amino acid residue is represented by three backbone spheres, one each for N-H, C-H, and C=O, and one side-chain sphere R. [52] While our coarse-grained model sacrifices some of the atomic-level detail and incorporates solvent effects implicitly, it captures the major biophysical features of the systems and allows us to simulate larger systems for longer time scales. A detailed description of DMD and PRIME20 can be found in other papers from our group. [53-55] Here we describe the modifications we have made to the PRIME20 force field in order to simulate small inhibitor molecules alongside of peptides.

4.4.2 Inhibitor Model - Coarse-Grained Sites, Geometric and Energetic Parameters

While PRIME20 is well suited for simulations of peptides and proteins, it does not contain geometric or energetic parameters for other types of molecules. We therefore had to come up with a way to incorporate an inhibitor model into PRIME20 that would be compatible with the

PRIME20 representation of peptides. The inhibitor model that we devised accounts for eight different coarse-grained groups, CH3, CH2, CH, C, OH, O, C=O and H; these are consistent with

132

the three original groups (CH, C=O, NH) used to represent the peptide backbone in PRIME20.

The hydrogen bonding scheme for the inhibitor model is the same as in PRIME20. [56] Fig. 10 shows the coarse-grained representations of A) resveratrol, B) EGCG, C) vanillin, and D) curcumin with a side bar that shows the 8 coarse-grained groups.

Next, we defined geometrical and energetic parameters that were consistent with the

PRIME20 force field. We based the geometric parameters for the new CG groups on the PRIME20 parameters for the CH3 group, the alanine sidechain group, in PRIME20. In our inhibitor model, the aromatic C, CH, CH2 groups all have bead diameters of 2.340Å, which is smaller than the diameter, 2.700Å, of an aliphatic carbon bead CH3. (The latter was chosen to be the same as the diameter of an alanine sidechain in PRIME20). [57] The diameter for the O group was scaled to that of the C group to obtain 2.150Å based on the ratio of atomistic carbon and oxygen van der

Waals radii, 1.700Å and 1.520Å, respectively. The O-H and O groups were chosen to be the same size. Similarly, the diameter for the H group was scaled to that of the C group to be 1.200 Å based on the van der Waals radii of carbon and hydrogen. The values for the mass, reduced mass, bead diameter σ, well diameter λ and square well depth ε of the eight different coarse-grained sites are summarized in Table 1.

4.4.3 Inhibitor Model – Bond and Pseudo-bond Parameters Obtained from Atomistic

Simulations

After defining the coarse-grained inhibitor geometric and energetic parameters, we determined the covalent bond length and fluctuation on the inhibitor molecules to model realistic molecular motion. In order to do this, the atomistic representation for each inhibitor was constructed using Accelrys Discovery Studio [58] and the topology files were generated using

133

SwissParam [59], an automated force field generator for small molecules. The GROMACS 4.5.5

[60] simulation package, the GROMOS53a6 [61] force field and SPC water are used. Each simulation was run with explicit water at 297K in a 5nm cubic box with periodic boundary conditions for a total of 10ns. Covalent bond lengths in our coarse-grained inhibitor molecule were assigned the values determined by SwissParam. The bond lengths between the C and O-H groups

(navy sphere and light blue spheres in Fig. 10), between the H and C=O (grey sphere and red spheres in Fig. 10) and between the C=O and C groups (red sphere and light blue spheres in Fig.

10) are taken to be the atomistic bond lengths between C and O, between H and C, and between C and C, respectively. The fluctuation ranges of all covalent bonds for each inhibitor molecule are determined over the whole simulation trajectory; they were within 3-6% of their assigned value.

We also determined the pseudo-bond length and fluctuation range. In PRIME20, pseudo- bonds are used to constrain coarse-grained sphere movement to match the backbone, the torsion angle, and the position of the sidechain groups to experimentally-measured values. For the inhibitor model, we also use pseudo-bonds to maintain the planar aromatic ring structures of the benzene and benzopyran rings as well as the alkene groups. Fig. 11 shows the location of the pseudo-bonds in A) resveratrol, B) EGCG, C) vanillin, and D) curcumin. Red bonds constrain benzene and benzopyran rings, and green, purple and blue bonds restrain spheres once, twice, and three times removed from the rings. The pseudo bond scheme chosen here is a first approximation to achieve our goals and will be refined in future work. The pseudo bond lengths for benzene and benzopyran rings fluctuate between 2.60Å and 2.92Å. Pseudo-bonds for spheres once removed, twice removed and three times removed from a ring fluctuate between 2.21Å and 2.50Å, between

2.24Å and 2.57Å, and between 2.73Å and 3.80Å.

134

We compare the most populated conformations for the four molecules in the coarse- grained simulations to those in the atomistic simulations in Fig. S2. In Fig. S2, we see that vanillin has a rigid conformation in both CG and atomistic simulations. The conformations of resveratrol, curcumin and EGCG observed in both CG and atomistic simulations are structurally similar except for the relative positions of the aromatic ring groups. Curcumin is more flexible than the other three molecules in that it adopts multiple conformations in both atomistic and CG simulations. The “extended” conformations of curcumin observed in both simulations are similar to each other. However, in atomistic simulation, the carbon linker can bend so strongly that the two aromatic groups stack together, but such “closed” conformations are not observed in CG simulations.

4.4.4 Inhibitor Model - Energetic Parameters

The well depth ε between the C, CH and CH2 groups has the same value as that of the alanine side chain group (CH3). The hydrogen bonding groups O-H, O and C=O on inhibitor molecules all have the same well width λ and well depth ε to maintain compatibility with our current hydrogen bonding routine. When the hydrogen bonding event involves the terminal groups, e.g. N-H2, C=O and O-H, it skips normal judging criteria and instead takes place with a probability of 0.2. The H group on vanillin is included as a separate group in the model to maintain the hydrogen bonding subroutine used in PRIME20 force field. It has a hard sphere interaction with all the other groups.

135

4.4.5 PRIME20 Force Field and Aβ(17-36) Peptide Model

In this work, we made a few adjustments to the original PRIME20 force field. [62] In previous DMD simulations of 8 Aβ(17-42) peptides by Cheon et al., the original PRIME20 force field was augmented to include: double-well potentials for all of the side-chain pair interactions, parallel preference constraints for backbone hydrogen bond angles [63], and enhanced salt-bridge interaction between D23 and K28 (εD23-K28=0.4εHB) [62] where εHB is the hydrogen bonding energy between NH and C=O. In this work, we keep the parallel preference constraints for backbone hydrogen bond angle and the enhanced salt-bridge D23-K28 interaction but we maintain single well potentials for all of the side-chain pair interactions. Additionally, we include an enhanced hydrophobic interaction between F19 and L34, and between F19 and I32 (εF19-L34=εF19-

I32=0.305εHB, 1.5 times its original value) since the hydrophobic sidechain interactions between

F19 and L34 are evidently important in stabilizing the U-shaped fibril. [45, 64] This treatment helps bias the system toward forming a U-shaped protofilament and reduces the complexity of aggreagtion pathway.

4.4.6. Simulation Procedure - Coarse-Grained Peptide/Inhibitor Simulation

The coarse-grained Aβ(17-36)/inhibitor simulations proceed in the following way. Peptides and inhibitor molecules (if inhibitors are also being simulated) are initially placed at random locations in a cubic simulation box with sides of lengths L = 138.5Å and periodic boundary

* conditions. In these simulations, the reduced temperature is defined as T = kBT/εHB where εHB is the hydrogen bonding well depth. The reduced temperature is related to real temperature by matching the folding temperature of helical polypeptides measured in experiments and our DMD simulation. The equation T/K = 2288.46T*- 115.79 was derived in our previous work. [65]The

136

reduced time is related to real time by using the conversion factor, 1 reduced time unit, Δtreduced =

0.46 ns which was used in our previous work. [65] Velocities for each peptide and inhibitor sphere are chosen randomly from a Maxwell-Boltzmann distribution that is centered at the desired temperature. The simulation temperature is set to T* = 0.20, corresponding to a real temperature of

342K, and is held constant using the Andersen thermostat. [66] Spheres in the simulation experience random “ghost collisions” with “ghost particles” so that the system temperature is maintained at the desired value. We first simulated a system of 8 Aβ(17-36) peptides as this could serve as the benchmark for peptide-inhibitor simulations. The six peptide-only simulations lasted between 35 and 49 μs. Then we ran four different types of peptide/inhibitor simulations for each of four molecules, vanillin, curcumin, resveratrol and EGCG: (1). one peptide interacting with ten inhibitors; (2). eight peptides aggregating with thirty inhibitors; (3).one protofilament (octamer) interacting with thirty inhibitors; (4). eight peptides aggregating in the presence of one protofilament (octamer) and thirty inhibitors.

4.5 Acknowledgement

This work was supported by National Institutes of Health Grants EB006006 and in part by

National Science Foundation Research Triangle Materials Research Science and Engineering

Centers (MRSEC) Grant DMR-1121107. The authors declare that they have no conflicts of interest.

Abbreviation: Aβ-Amyloid β protein; EGCG- epigallocatechin-3-gallate.

Keywords: discontinuous molecular dynamics; protein aggregation; polyphenol; Amyloid β; inhibitory mechanism.

137

4.6 References

1. L.K. Wolf. Getting to Alzheimer's Early. Chemical & Engineering News [Online]2014. p.

41.

2. G. Kolata. Doubt on Tactic in Alzheimer's Battle. The New York Times [Online]2010.

3. G. Kolata. Drug Trials Test Bold Plan to Slow Alzheimer's. The New York Times

[Online]2010.

4. A.J. Doig, M.P. del Castillo-Frias, O. Berthoumieu, B. Tarus, J. Nasica-Labouze, F.

Sterpone, et al., Why Is Research on Amyloid-β Failing to Give New Drugs for Alzheimer

Disease? ACS Chemical Neuroscience. 8 (2017) 1435-1437.

5. M. Necula, R. Kayed, S. Milton, C.G. Glabe, Small Molecule Inhibitors of Aggregation

Indicate That Amyloid β Oligomerization and Fibrillization Pathways Are Independent and

Distinct, Journal of Biological Chemistry. 282 (2007) 10311-10324.

6. Y. Nakagami, S. Nishimura, T. Murasugi, I. Kaneko, M. Meguro, S. Marumoto, et al., A

novel β-sheet breaker, RS-0406, reverses amyloid β-induced cytotoxicity and impairment

of long-term potentiation in vitro, British Journal of Pharmacology. 137 (2002) 676-682.

7. P. Nguyen, P. Derreumaux, Understanding Amyloid Fibril Nucleation and Aβ

Oligomer/Drug Interactions from Computer Simulations, Accounts of Chemical Research.

47 (2014) 603-611.

8. A.J. Doig, P. Derreumaux, Inhibition of protein aggregation and amyloid formation by

small molecules, Current Opinion in Structural Biology. 30 (2015) 50-56.

9. J. Kim, H.J. Lee, K.W. Lee, Naturally occurring phytochemicals for the prevention of

Alzheimer’s disease, Journal of Neurochemistry. 112 (2010) 1415-1430.

10. A.Y. Sun, Q. Wang, A. Simonyi, G.Y. Sun, Botanical Phenolics and Brain Health,

138

Neuromolecular Medicine. 10 (2008) 259-274.

11. Y. Porat, A. Abramowitz, E. Gazit, Inhibition of amyloid fibril formation by polyphenols:

Structural similarity and aromatic interactions as a common inhibition mechanism,

Chemical Biology & Drug Design. 67 (2006) 27-37.

12. A.A. Reinke, J.E. Gestwicki, Structure-activity relationships of amyloid beta-aggregation

inhibitors based on curcumin: Influence of linker length and flexibility, Chemical Biology

& Drug Design. 70 (2007) 206-215.

13. J.H. Jang, Y.J. Surh, Protective effect of resveratrol on beta-amyloid-induced oxidative

PC12 cell death, Free Radical Biology and Medicine. 34 (2003) 1100-1110.

14. E. Savaskan, G. Olivieri, F. Meier, E. Seifritz, A. Wirz-Justice, F. Muller-Spahn, Red wine

ingredient resveratrol protects from beta-amyloid neurotoxicity, Gerontology. 49 (2003)

380-383.

15. S.S. Karuppagounder, J.T. Pinto, H. Xu, H.L. Chen, M.F. Beal, G.E. Gibson, Dietary

supplementation with resveratrol reduces plaque pathology in a transgenic model of

Alzheimer's disease, Neurochemistry International. 54 (2009) 111-118.

16. Y.S. Han, W.H. Zheng, S. Bastianetto, J.G. Chabot, R. Quirion, Neuroprotective effects of

resveratrol against beta-amyloid-induced neurotoxicity in rat hippocampal neurons:

involvement of protein kinase C, British Journal of Pharmacology. 141 (2004) 997-1005.

17. P. Marambaud, H.T. Zhao, P. Davies, Resveratrol promotes clearance of Alzheimer's

disease amyloid-beta peptides, Journal of Biological Chemistry. 280 (2005) 37377-37382.

18. Y. Chebaro, P. Jiang, T. Zang, Y. Mu, P.H. Nguyen, N. Mousseau, et al., Structures of A

beta 17-42 Trimers in Isolation and with Five Small-Molecule Drugs Using a Hierarchical

Computational Procedure, Journal of Physical Chemistry B. 116 (2012) 8412-8422.

139

19. Y. Feng, X.P. Wang, S.G. Yang, Y.J. Wang, X. Zhang, X.T. Du, et al., Resveratrol inhibits

beta-amyloid oligomeric cytotoxicity but does not prevent oligomer formation,

Neurotoxicology. 30 (2009) 986-995.

20. J. Wang, L. Ho, Z. Zhao, I. Seror, N. Humala, D.L. Dickstein, et al., Moderate consumption

of Cabernet Sauvignon attenuates A beta neuropathology in a mouse model of Alzheimer's

disease, Faseb Journal. 20 (2006) 2313-2320.

21. S.A. Frautschy, W. Hu, P. Kim, S.A. Miller, T. Chu, M.E. Harris-White, et al., Phenolic

anti-inflammatory antioxidant reversal of A beta-induced cognitive deficits and

neuropathology, Neurobiology of Aging. 22 (2001) 993-1005.

22. G.P. Lim, T. Chu, F. Yang, W. Beech, S.A. Frautschy, G.M. Cole, The Curry Spice

Curcumin Reduces Oxidative Damage and Amyloid Pathology in an Alzheimer Transgenic

Mouse, The Journal of Neuroscience. 21 (2001) 8370-8377.

23. M. Garcia-Alloza, L.A. Borrelli, A. Rozkalne, B.T. Hyman, B.J. Bacskai, Curcumin labels

amyloid pathology in vivo, disrupts existing plaques, and partially restores distorted

neurites in an Alzheimer mouse model, Journal of Neurochemistry. 102 (2007) 1095-1104.

24. K. Ono, K. Hasegawa, H. Naiki, M. Yamada, Curcumin has potent anti-amyloidogenic

effects for Alzheimer's beta-amyloid fibrils in vitro, Journal of Neuroscience Research. 75

(2004) 742-750.

25. F.S. Yang, G.P. Lim, A.N. Begum, O.J. Ubeda, M.R. Simmons, S.S. Ambegaokar, et al.,

Curcumin inhibits formation of amyloid beta oligomers and fibrils, binds plaques, and

reduces amyloid in vivo, Journal of Biological Chemistry. 280 (2005) 5892-5901.

26. S.T. Ngo, M.S. Li, Curcumin Binds to A beta(1-40) Peptides and Fibrils Stronger Than

Ibuprofen and Naproxen, Journal of Physical Chemistry B. 116 (2012) 10165-10175.

140

27. L.N. Zhao, S.W. Chiu, J. Benoit, L.Y. Chew, Y.G. Mu, The Effect of Curcumin on the

Stability of A beta Dimers, Journal of Physical Chemistry B. 116 (2012) 7428-7435.

28. Y.T. Choi, C.H. Jung, S.R. Lee, J.H. Bae, W.K. Baek, M.H. Suh, et al., The green tea

polyphenol (-)-epigallocatechin gallate attenuates beta-amyloid-induced neurotoxicity in

cultured hippocampal neurons, Life Sciences. 70 (2001) 603-614.

29. J. Bieschke, J. Russ, R.P. Friedrich, D.E. Ehrnhoefer, H. Wobst, K. Neugebauer, et al.,

EGCG remodels mature alpha-synuclein and amyloid-beta fibrils and reduces cellular

toxicity, Proceedings of the National Academy of Sciences of the United States of America.

107 (2010) 7710-7715.

30. K. Rezai-Zadeh, G.W. Arendash, H.Y. Hou, F. Fernandez, M. Jensen, M. Runfeldt, et al.,

Green tea epigallocatechin-3-gallate (EGCG) reduces beta-amyloid mediated cognitive

impairment and modulates tau pathology in Alzheimer transgenic mice, Brain Research.

1214 (2008) 177-187.

31. Y. Levites, T. Amit, S. Mandel, M.B.H. Youdim, Neuroprotection and neurorescue against

A beta toxicity and PKC-dependent release of non-amyloidogenic soluble precursor protein

by green tea polyphenol (-)-epigallocatechin-3-gallate, Faseb Journal. 17 (2003) 952-954.

32. D.E. Ehrnhoefer, J. Bieschke, A. Boeddrich, M. Herbst, L. Masino, R. Lurz, et al., EGCG

redirects amyloidogenic polypeptides into unstructured, off-pathway oligomers, Nature

Structural & Molecular Biology. 15 (2008) 558-566.

33. S. Bastianetto, Z.X. Yao, V. Papadopoulos, R. Quirion, Neuroprotective effects of green

and black teas and their catechin gallate esters against beta-amyloid-induced toxicity,

European Journal of Neuroscience. 23 (2006) 55-64.

34. F.F. Liu, X.Y. Dong, L.Z. He, A.P.J. Middelberg, Y. Sun, Molecular Insight into

141

Conformational Transition of Amyloid beta-Peptide 42 Inhibited by (-)-Epigallocatechin-

3-gallate Probed by Molecular Simulations, Journal of Physical Chemistry B. 115 (2011)

11879-11887.

35. S.H. Wang, X.Y. Dong, Y. Sun, Thermodynamic Analysis of the Molecular Interactions

between Amyloid beta-Protein Fragments and (-)-Epigallocatechin-3-gallate, Journal of

Physical Chemistry B. 116 (2012) 5803-5809.

36. T. Zhang, J. Zhang, P. Derreumaux, Y.G. Mu, Molecular Mechanism of the Inhibition of

EGCG on the Alzheimer A beta(1-42) Dimer, Journal of Physical Chemistry B. 117 (2013)

3993-4002.

37. L.M. Young, J.C. Saunders, R.A. Mahood, C.H. Revill, R.J. Foster, L.-H. Tu, et al.,

Screening and classifying small molecule inhibitors of amyloid formation using ion

mobility spectrometry-mass spectrometry, Nature Chemistry. 7 (1) 73-81.

38. A.R.A. Ladiwala, J.C. Lin, S.S. Bale, A.M. Marcelino-Cruz, M. Bhattacharya, J.S. Dordick,

et al., Resveratrol Selectively Remodels Soluble Oligomers and Fibrils of Amyloid A beta

into Off-pathway Conformers, Journal of Biological Chemistry. 285 (2010) 24228-24237.

39. J.M. Lopez del Amo, U. Fink, M. Dasari, G. Grelle, E.E. Wanker, J. Bieschke, et al.,

Structural Properties of EGCG-Induced, Nontoxic Alzheimer's Disease A beta Oligomers,

Journal of Molecular Biology. 421 (2012) 517-524.

40. K. Rajasekhar, M. Chakrabarti, T. Govindaraju, Function and toxicity of amyloid beta and

recent therapeutic interventions targeting amyloid beta in Alzheimer's disease, Chemical

Communications. 51 (2015) 13434-13450.

41. M. Convertino, A. Vitalis, A. Caflisch, Disordered Binding of Small Molecules to A

beta(12-28), Journal of Biological Chemistry. 286 (2011) 41578-41588.

142

42. M. Naldi, J. Fiori, M. Pistolozzi, A.F. Drake, C. Bertucci, R. Wu, et al., Amyloid β-Peptide

25–35 Self-Assembly and Its Inhibition: A Model Undecapeptide System to Gain Atomistic

and Secondary Structure Details of the Alzheimer’s Disease Process and Treatment, ACS

Chemical Neuroscience. 3 (2012) 952-962.

43. C. Bleiholder, T.D. Do, C. Wu, N.J. Economou, S.S. Bernstein, S.K. Buratto, et al., Ion

Mobility Spectrometry Reveals the Mechanism of Amyloid Formation of Aβ(25–35) and

Its Modulation by Inhibitors at the Molecular Level: Epigallocatechin Gallate and Scyllo-

inositol, Journal of the American Chemical Society. 135 (2013) 16926-16937.

44. M. Ahmed, J. Davis, D. Aucoin, T. Sato, S. Ahuja, S. Aimoto, et al., Structural conversion

of neurotoxic amyloid-beta(1-42) oligomers to fibrils, Nature Structural & Molecular

Biology. 17 (2010) 561-567.

45. M. Chandrakesan, B. Sarkar, V.S. Mithu, R. Abhyankar, D. Bhowmik, S. Nag, et al., The

basic structural motif and major biophysical properties of Amyloid-beta are encoded in the

fragment 18-35, Chemical Physics. 422 (2013) 80-87.

46. M. Zhu, A. De Simone, D. Schenk, G. Toth, C.M. Dobson, M. Vendruscolo, Identification

of small-molecule binding pockets in the soluble monomeric form of the A beta 42 peptide,

Journal of Chemical Physics. 139 (2013).

47. D. Frishman, P. Argos, Knowledge-based protein secondary structure assignment, Proteins-

Structure Function And Genetics. 23 (1995) 566-579.

48. W. Humphrey, A. Dalke, K. Schulten, VMD: Visual molecular dynamics, Journal Of

Molecular Graphics & Modelling. 14 (1996) 33-38.

49. M. Chandrakesan, D. Bhowmik, B. Sarkar, R. Abhyankar, H. Singh, M. Kallianpur, et al.,

Steric Crowding of the Turn Region Alters the Tertiary Fold of Amyloid-beta(18-35) and

143

Makes It Soluble, Journal of Biological Chemistry. 290 (2015) 30099-30107.

50. S. Kim, Wenling E. Chang, R. Kumar, Dmitri K. Klimov, Naproxen Interferes with the

Assembly of A beta Oligomers Implicated in Alzheimer's Disease, Biophysical Journal.

100 (2011) 2024-2032.

51. B.J. Alder, T.E. Wainwright, STUDIES IN MOLECULAR DYNAMICS .1. GENERAL

METHOD, Journal Of Chemical Physics. 31 (1959) 459-466.

52. C. Wu, J.-E. Shea, Coarse-grained models for protein aggregation, Current opinion in

structural biology. 21 (2011) 209-220.

53. M. Cheon, I. Chang, C.K. Hall, Extending the PRIME model for protein aggregation to all

20 amino acids, Proteins-Structure Function And Bioinformatics. 78 (2010) 2950-2960.

54. H.D. Nguyen, C.K. Hall, Molecular dynamics simulations of spontaneous fibril formation

by random-coil peptides, Proceedings Of the National Academy Of Sciences Of the United

States Of America. 101 (2004) 16180-16185.

55. A.J. Marchut, C.K. Hall, Spontaneous formation of annular structures observed in

molecular dynamics simulations of polyglutamine peptides, Computational Biology and

Chemistry. 30 (2006) 215-218.

56. A.V. Smith, C.K. Hall, alpha-helix formation: Discontinuous molecular dynamics on an

intermediate-resolution protein model, Proteins-Structure Function And Genetics. 44

(2001) 344-360.

57. L.J. Yang, C.H. Tan, M.J. Hsieh, J.M. Wang, Y. Duan, P. Cieplak, et al., New-generation

amber united-atom force field, Journal Of Physical Chemistry B. 110 (2006) 13166-13176.

58. A.S. Inc., Discovery studio modeling environment, release 4.0. (2013).

59. V. Zoete, M.A. Cuendet, A. Grosdidier, O. Michielin, SwissParam: A Fast Force Field

144

Generation Tool for Small Organic Molecules, Journal Of Computational Chemistry. 32

(2011) 2359-2368.

60. S. Pronk, S. Pall, R. Schulz, P. Larsson, P. Bjelkmar, R. Apostolov, et al., GROMACS 4.5:

a high-throughput and highly parallel open source molecular simulation toolkit,

Bioinformatics. 29 (2013) 845-854.

61. C. Oostenbrink, A. Villa, A.E. Mark, W.F. Van Gunsteren, A biomolecular force field based

on the free enthalpy of hydration and solvation: The GROMOS force-field parameter sets

53A5 and 53A6, Journal Of Computational Chemistry. 25 (2004) 1656-1676.

62. M. Cheon, C.K. Hall, I. Chang, Structural Conversion of A beta(17-42) Peptides from

Disordered Oligomers to U-Shape Protofilaments via Multiple Kinetic Pathways, Plos

Computational Biology. 11 (2015) e1004258.

63. M. Cheon, I. Chang, C.K. Hall, Influence of temperature on formation of perfect tau

fragment fibrils using PRIME20/DMD simulations, Protein Science. 21 (2012) 1514-1527.

64. Y.L. Xiao, B.Y. Ma, D. McElheny, S. Parthasarathy, F. Long, M. Hoshi, et al., A beta(1-42)

fibril structure illuminates self-recognition and replication of amyloid in Alzheimer's

disease, Nature Structural & Molecular Biology. 22 (2015) 499-505.

65. Y. Wang, Q. Shao, C.K. Hall, N-terminal Prion Protein Peptides (PrP(120–144)) Form

Parallel In-register β-Sheets via Multiple Nucleation-dependent Pathways, Journal of

Biological Chemistry. 291 (2016) 22093-22105.

66. H.C. Andersen, Molecular dynamics simulations at constant pressure and/or temperature,

Journal of Chemical Physics. 72 (1980) 2384-2393.

145

Figure 4.1. Simulation snapshots showing the pathway along which 8 Aβ(17-36) peptides form a parallel in-register β-sheet-rich U-shaped protofilament via a templated-growth mechanism. Snapshots are taken at t = (A) 0, (B) 6079, (C) 7838, (D) 12190, (E) 103775 (ns). F is a schematic side-view representation of the U-shaped protofilament in E. G shows the time evolution of the systems total interaction energy (kJ/mol).

146

Figure 4.2. Root mean square displacement (RMSD) of 4 inhibitor molecules versus simulation time from both atomistic (A) amd coarse-grained (B) simulations, respectively. The reference structures for RMSD calculation are their initial structures. C. Number of contacts between inhibitors and individual amino acids on Aβ(17-36) peptide and; D. Plot of the radial distribution function g(r) of the center of mass distance between the F19 sidechain on Aβ(17-36) peptide and the four inhibitor molecules.

147

Figure 4.3. Simulation snapshots of 8 Aβ(17-36) peptides aggregating in the presence of 30 vanillin and EGCG molecules taken at different time points. The peptides are colored blue, the inhibitors are colored magenta, respectively.

148

Figure 4.4. Number of hydrogen bonds formed as a function of time during Aβ(17-36) peptides aggregation in the absence and presence of 30 vanillin, resveratrol, curcumin, EGCG molecules.

149

Figure 4.5. (A) Peptide sidechain-sidechain and (B) sidechain-inhibitor interaction energy distribution. (Hydrophobic residues are highlighted in blue).

150

Figure 4.6. The interaction energy map of Aβ(17-36) peptide and (A) vanillin, (B) resveratrol, (C) curcumin and (D) EGCG. The scale bar applies for A-D and its unit is kJ/mol. The inhibitor CG sites are grouped by CG types.

151

Figure 4.7. The Aβ(17-36)-inhibitor interaction energy distribution for (A) vanillin, (B) resveratrol, (C) curcumin and (D) EGCG.

152

Figure 4.8. A-D are simulation snapshots of EGCG redirecting a U-shaped Aβ(17-36) protofilament into a disordered oligomer at t = 242 (A), 923 (B), 1421 (C) and 1990 (D) ns. E plots the number of peptide backbone hydrogen bonds versus simulation time.

153

Figure 4.9. A plot of the number of peptide backbone hydrogen bonds remained as a function of simulation time.

154

Figure 4.10. Coarse-grained representations of A) resveratrol, B) EGCG, C) vanillin, and D) curcumin to be used in DMD simulations. The side bar defines each coarse-grained group. The CG sites are labeled 1, 2, 3… for use in later discussion. (vanillin: 1-11, resveratrol: 1-17, curcumin: 1-25, EGCG: 1-32)

155

Figure 4.11. Pseudo-bonds used for A) resveratrol, B) EGCG, C) vanillin, and D) curcumin. Red bonds constrain benzene and benzopyran rings, and green, purple and blue bonds constrain spheres once, twice, and three times removed from the rings.

156

Table 4.1. Coarse-grained group parameters for phenolic compounds.

CG Site Atoms Mass Reduced Mass σ(Å) λ(Å) ε(reduced unit) Interaction type 1 CH3 15.035 1.000 2.700 5.400 0.084 Hydrophobic 2 CH2 14.027 0.933 2.340 4.680 0.084 Hydrophobic 3 CH 13.019 0.866 2.340 4.680 0.084 Hydrophobic 4 C 12.011 0.799 2.340 4.680 0.084 Hydrophobic 5 OH 17.007 1.131 2.150 4.500 1.000 Hydrogen Bond 6 O 15.999 1.064 2.150 4.500 1.000 Hydrogen Bond 7 CO 28.010 1.863 4.000 4.500 1.000 Hydrogen Bond 8 H 1.008 0.067 1.200 0.000 0.000 Hard Sphere

157

4.7 Supporting Information

Figure S4.1. Number of peptide-inhibitor (A) and peptide-peptide (B) contacts versus simulation time for four small molecules.

158

Figure S4.2. Comparison of most populated conformations of vanillin (A and F), resveratrol (B and G), curcumin (C, D, H and I) and EGCG (E and J) from atomistic and coarse-grained simulation trajectories. The carbon, hydrogen and oxygen beads are colored cyan, white and red.

159

CHAPTER 5

Simulations and Experiments Delineate Amyloid Fibrilization by Peptides Derived from

Glaucoma-Associated Myocilin

Chapter 5 is essentially a manuscript by Yiming Wang, Yuan Gao, Shannon E. Hill, Dustin J. E.

Huard, Moya O. Tomlin, Raquel L. Lieberman, Anant K. Paravastu and Carol K. Hall published on Journal of Physical Chemistry B.

160

Simulations and Experiments Delineate Amyloid Fibrilization by Peptides Derived from Glaucoma-Associated Myocilin

Yiming Wanga, Yuan Gaob, Shannon E. Hillc, Dustin J. E. Huardc, Moya O. Tomlinc, Raquel L. Liebermanc, Anant K. Paravastub, Carol K. Halla*

aDepartment of Chemical and Biomolecular Engineering, North Carolina State University, Raleigh, North Carolina 27695-7905, United States bSchool of Chemical and Biomolecular Engineering, Georgia Institute of Technology, Atlanta, GA 30332-0100, USA. cSchool of Chemistry and Biochemistry, Georgia Institute of Technology, Atlanta, GA 30332- 0400, USA.

Abstract

Although Mutant myocilin aggregation is associated with inherited open angle glaucoma, a prevalent optic neuropathy leading to blindness. Comprehension of mutant myocilin aggregation is of fundamental importance to glaucoma pathogenesis and ties glaucoma to amyloid diseases such as Alzheimer’s. Here we probe the aggregation properties of peptides derived from the myocilin olfactomedin domain. Peptides P1 (residues 326-337) and P3 (residues 426-442) were identified previously to form amyloid. Coarse-grained discontinuous molecular dynamics simulations using the PRIME20 force field (DMD/PRIME20) predict that P1 and P3 are aggregation-prone; P1 consistently forms fibrillar aggregates with parallel in-register β-sheets whereas P3 forms β-sheet-containing aggregates without distinct order. Natural abundance 13C solid-state NMR spectra validate that aggregated P1 exhibits amyloid signatures and is less heterogeneous than aggregated P3. DMD/PRIME20 simulations provide a viable method to

161

predict peptide aggregation propensities and aggregate’s structure/order which cannot be accessed by bioinformatics or readily attained experimentally.

162

5.1 Introduction

A recent addition to the list of proteins associated with amyloid-based diseases is the myocilin olfactomedin domain (mOLF)1-2. Inherited non-synonymous mutations in mOLF are causative for early-onset open angle glaucoma, an ocular neurodegenerative disease leading to blindness3. Mutant myocilins aggregate instead of being secreted to the trabecular meshwork

(TM), a key anatomical structure responsible for maintaining eye pressure that is dysregulated in most cases of glaucoma4. Whereas cells typically deal with mutant proteins with an efficient degradation system mediated by molecular chaperones, this process is significantly compromised in TM cells expressing mutant myocilin. Mutant myocilins accumulate intracellularly5-7 where they exhibit aberrant interactions with molecular chaperones8-9, leading to TM cell death that hastens glaucoma onset. Cumulatively, such behavior suggests that mutant myocilins might be aggregating into amyloid, which are well-known for their high thermochemical stability and resistance to degradation. 10-13

Initial evidence for the amyloid behavior of mOLF includes thioflavin-T (ThT) positive aggregates generated in vitro, demonstrating a sigmoidal growth curve composed of a lag phase followed by exponential growth, as well as a positive ThT signal from full-length mutant myocilin that has accumulated within mammalian cells2. The mOLF amyloid aggregation parameters and associated biophysical features have been evaluated further1 (Fig. 1), leading to the identification of two different fibril morphologies, long straight fibrils (Fig. 1A) and more unusual lassoed oligomers (Fig. 1B). These morphologies are also observed in aggregates of glaucoma-associated variants, A427T and D380A, respectively (not shown). Subsequent bioinformatics analysis (Fig.

S1) to determine the amyloidogenic regions on full-length mOLF protein revealed three consensus

1 sequences : G326AVVYSGSLYFQ (P1), G387LWVIYSTDEAKGAIVLSK (P2) and

163

V426ANAFIICGTLYTVSSY (P3). Experimentally, two of these peptides, P3 and P1, formed ThT- positive aggregates (Fig. 1C) whereas P2 remained soluble. At low concentrations, P1 aggregation exhibits the expected lag phase, whereas conditions for observing a lag phase for P3 aggregation have not yet been obtained (data not shown). Upon solving the mOLF crystal structure14, we learned that P1 and P3 are in the interior of the β-propeller, confirming our earlier finding that fibril formation is initiated from a partially folded state1. The fibril morphologies of P1 and P3

(Figs. 1E, F) recapitulate morphologies seen for the full mOLF protein domain (Figs. 1A, B), suggesting that P1 and P3 are the core stretches responsible for the observed morphologies adopted by the full protein1.

To further comprehend amyloidogenesis by mOLF-derived peptides and its molecular relationship to better-studied amyloids like amyloid-β15-17, we performed discontinuous molecular dynamics simulations (DMD) combined with the PRIME20 force field 18-20 to predict fibrillar structure, as well as solid-state NMR to experimentally evaluate the extent of structural order of aggregated P1 and P3. In addition to providing new molecular insight into mOLF aggregation, our study demonstrates proof-of-concept for the applicability of DMD/PRIME20 to predict peptide amyloidogenicity and fibril structure.

5.2 Methods

5.2.1 DMD/PRIME20 simulations

PRIME20 is a 4-sphere-per-residue coarse-grained protein model (three backbone spheres

Cα, NH, CO and one sidechain sphere) developed in the Hall group that was specifically designed for DMD-based simulation of protein aggregation18-20, and has unique geometric and energetic parameters for each of the 20 different amino acids. Specifically, each sidechain sphere of the 20

164

amino acids has a distinct hard sphere diameter (effective van der Waals radius), and distinct sidechain-to-backbone distances (R-Cα, R-NH, R-CO). The potential energy function between two amino acid sidechain beads is modeled as a square well potential. The potential energy parameters between the 20 different amino acids (except for glycine) include 210 independent square well widths and 19 independent square well depths. The backbone hydrogen bonding interactions are modeled as a directional square well potential. The detailed description of the derivation of the

PRIME20 geometric and energetic parameters along with their values are given in our earlier work.

19, 21-22 * The reduced temperature is defined as T =kBT/εHB. It is related to a real temperature by T

[K] = 2288.46T*-115.79; this equation was obtained by matching the folding temperature of alanine-rich polypeptides in DMD simulation in our previous work 23 to the experimental values24.

The reduced time unit is Δtreduced = 0.96 ns; this was obtained by matching the self-diffusion coefficient of Aβ(16-22) obtained from DMD simulations to that calculated from atomistic MD simulation.23

The simulation system consists of eight peptides in a cubic box with peptide concentration

C=20mM at T=342K. The simulation temperature is chosen to optimize the likelihood that P1 and

P3 aggregate to form stable ordered β-sheet-rich protofilament. If the temperature is too high, the peptides do not aggregate and if it is too low they get trapped in local energy minimum states. The peptides are initially at random locations and in random coil conformations. We performed ten independent runs for the P1, P2 and P3 peptides. Over the course of the simulation, the peptides spontaneously aggregate, constantly rearranging into more-stable structures as they move towards the equilibrium state. The simulation times for the P1, P2 and P3 systems are 223μs, 353μs and

476μs, respectively.

165

5.2.2 Peptide synthesis and experimentation fibril formation

P1 and P3 peptides were synthesized by CPC Scientific (Sunnyvale, CA) to >95% purity for P1 and 88.8% for P3. Both peptides were prepared as a 5 mg mL-1 stock solution in DMSO and stored at room temperature. Fibrils were generated by dissolving 500 μM peptide into 10 mM

Na2HPO4/KH2PO4 buffer at pH 7.2 containing 200 mM NaCl plus 10 μM ThT and incubating at

36 °C for 24-48 hours, as previously published.1 For P1 the products of three 4-mL reactions were combined for NMR while for P3 a one 4-mL reaction was found to be sufficient for a robust signal.

Insoluble aggregates were isolated for solid-state NMR characterization by ultracentrifugation

(164,700 xg for 15 minutes at 25°C).

5.2.3 Solid-state NMR measurements.

The 1H-13C CPMAS experiments 25-26 were performed on a narrow-bore Bruker 11.75

Tesla (500-MHz 1H NMR frequency) solid-state NMR system with an Avance III console and a

3.2 mm magic angle spinning (MAS) HX NMR probe. The MAS spin rate was kept at 10 kHz for all measurements. For the 2 ms cross polarization, a 50 kHz pulse on 13C channel and a calibrated ramped pulse on 1H channel was used. After the cross polarization, decoupling of 1H was employed with a radio-frequency field power of 100 kHz during the detection period, with two-pulse-phase modulation 27 (TPPM). The predicted CPMAS spectrum was generated by a summation of all the

Gaussian peaks at the chemical shifts of all 13C in the peptides. The line width (full width at half maximum, FWHM) of each Gaussian peak was identical.

166

5.3 Results and Discussion

P1 peptides aggregated into ordered β-sheet-rich oligomers in all ten independent

DMD/PRIME20 simulation runs (Figs. 2 and S2), predicting that P1 has a strong amyloid-forming propensity. P1 peptides formed a U-shaped protofilament (Figs. 2A (red curve, top panel), B, D) in two out of ten runs and an S-shaped protofilament (Fig. 2A (black curve, bottom panel), C, E) in the remaining eight runs. P1 aggregates formed within the first 10μs (Fig. 2A), indicating fast aggregation kinetics during simulations. The kinetics are consistent with other simulation 18, 28 studies indicating that, in general, amyloidogenic peptides tend to form fibrils with fast aggregation kinetics if they contain sixteen or fewer amino acids. The U-shaped P1 protofilament adopted a backbone turn at 332Gly (Figs. 2B, D), forming a hydrophobic core containing residues 328Val,

330Tyr, 334Leu and 336Phe. The S-shaped P1 protofilament adopted two turns, one at 332Gly and the other near the C-terminal (Fig. 2C, E). Analysis of the average number of inter-peptide backbone hydrogen bonds formed within the last half of the ten trajectories by each of the residues reveals that the whole P1 sequence, except C-terminal 336Phe and 337Gln, has a high propensity to form β-sheet structure (Fig. 3A). The DMD/PRIME20 result is consistent with bioinformatics predictions (Fig. S1) and experiments (Fig. 1C, E) indicating that the P1 region

326GAVVYSGSLY335 is an amyloid forming peptide within mOLF, and adds key new molecular insight into possible arrangements of the amyloid.

In DMD/PRIME20 simulations, P2 formed disordered oligomers (Fig. S3) containing a small amount of β-hairpin or β-sheet structure in nine out of ten simulation runs. In just one of ten runs, P2 aggregated into a fibrillar structure. In contrast to bioinformatics analysis, but in line with experimental results1, P2 has a low aggregation propensity as it forms only a small number of inter-peptide hydrogen bonds along the whole sequence (Fig. 3B).

167

P3 peptides aggregated in DMD/PRIME20 simulations to form parallel in-register U- shaped protofilaments in two out of ten runs and polymorphic β-sheet-rich oligomers in the remaining eight runs (Fig. 4, S4). From Fig. 4A(black curve, bottom panel), P3 aggregated into a disordered β-sheet-rich oligomer at t =70.6μs, formed a U-shaped trimer at t =149.5μs and grew into a stable U-shaped octamer protofilament at t= 176.3μs (Figs. 4C, D). Note that the potential energy dropped precipitously as P3 grew from a trimer to an octamer. In contrast when P3 forms polymorphic structures (Fig. 4A (red curve, top panel), B) the potential energy is relatively constant. Note that all the disordered β-sheet-rich P3 oligomers (Fig. S4) are in local minimum states that are metastable with respect to ordered fibrillar state (global minimum). These P3 oligomers may eventually come out of the local minimum states and form a more ordered fibrillar structure, but, based on current computation resources, it’s hard to predict the occurrence probability of such events. P3 is found to have a much slower aggregation kinetics than P1. One reason is that P3 has six more amino acids than P1 and thus has more degrees of freedom in its backbone chain movement which adds complexity to the aggregation pathways and increases nucleation time23, 29-30. Within the U-shaped P3 protofilament, the sidechains of residues 430Phe,

432Ile, 436Leu, 439Val on P3 closely contact each other to form a hydrophobic fibril core (Fig.

4D). In addition, the stretch of P3 that forms a large number of inter-peptide hydrogen bonds is

428NAFIICGTLYTVS440 (Fig. 3C), consistent with bioinformatics analysis (Fig. S1). Simulations support experimental results in that ThT positive aggregates with variable-sized, lassoed oligomer morphologies were observed (Fig. 1C, F), but these aggregates grew quickly in experiment. The

ThT kinetic data further suggests that P3 forms less-ordered aggregate structures than P1, based on the intensity of the ThT fluorescence which was lower than P1 even though qualitatively the two peptides appeared to aggregate at sufficient quantities with similar amounts of insoluble

168

material. The simulations support this hypothesis, as the polymorphic β-sheet rich oligomers observed in the P3 simulations likely bind ThT, but possibly with a lower fluorescence emission intensity or affinity when compared to better ordered P1 aggregates. In addition, our simulations suggest that a high level of templated order in a P3 fibril may be achievable if experimental conditions can be driven towards the U-shaped conformation.

In order to quantitatively compare the amyloid forming propensities of P1, P2 and P3 peptides based on DMD/PRIME20 simulation results, we define β(Pj) as the propensity of a specific peptide sequence Pj (e.g. P1, P2 or P3) to aggregate and form a β-sheet amyloid:

푁 1 푛 (푖) β(Pj) = ∑ 퐻퐵 푁 푛푆𝑖푡푒(푖) 𝑖=1 where 푛퐻퐵(푖) is the total number of backbone hydrogen bonding sites (NH and C=O beads) that

푡ℎ form β-sheet hydrogen bonds formed by the 푖 peptide in the aggregate, and 푛푆𝑖푡푒(푖) is the total number of NH and C=O beads on the 푖푡ℎ peptide in the aggregate. N is the total number of peptides.

β(Pj) ranges from 1 for a perfect β-sheet structure with strong amyloid forming propensity, to 0 for a monomeric state or disordered oligomer with weak amyloid forming propensity. The values of β(Pj) for P1, P2 and P3 with standard deviations over the last one-third of the trajectories of ten runs are β(P1) = 0.64±0.06, β(P2) = 0.25±0.07 and β(P3) = 0.44±0.07, indicating that the order of amyloid forming propensity of the three mOLF peptides is P1>P3>P2 (Fig. 3D), which is consistent with experimental data1 (Fig. 1C, E, F) that P1 and P3 form fibrils but P2 does not.

To experimentally evaluate the extent of molecular order for aggregated P1 and P3, we aggregated peptides P1 and P3 under physiological conditions as before1 and then acquired 1H-13C

Cross-Polarization Magic Angle Spinning (CPMAS) solid-state NMR spectra. Consistent with

DMD/PRIME20 predictions, P1 and P3 showed dramatically different CPMAS spectra (Figs. 5A,

169

C), confirming dissimilar structures. The spectrum of P1 fibrils is consistent with structurally ordered amyloid fibrils12, 31-32, namely, narrow NMR line widths (~1.2 ppm, Fig. 5A). In addition, though many 13C peaks on the P1 CPMAS spectrum overlap with others, some peaks with special chemical shifts can be assigned to a residue type based on the known chemical shift ranges of different 13C sites 33-35 (Fig. 5B, Table 1). Indeed, by comparing the chemical shifts of the distinguishable peaks with those from random coil peptides, all the secondary chemical shifts imply a strong trend to β-strand structure33. The observed structural order in the P1 CPMAS spectrum suggests that this sample will be compatible with future solid-state NMR structural methodologies. By contrast, the spectrum of P3 fibrils provides experimental evidence for a heterogeneous aggregate mixture, namely, broad line widths (> 14 ppm, Fig. 5B). In contrast to

P1, the P3 sample is unlikely to yield readily interpretable structural constraints, unless experimental aggregation kinetics can be modified to yield a homogeneous sample.

5.4 Conclusion

In summary, this study demonstrates the capability of DMD/PRIME20 to predict peptide amyloidogenicity and fibril structure. DMD/PRIME20 simulations both recapitulated mOLF- derived peptide aggregation propensities seen experimentally1, and predicted the extent of homogeneity in assembly into particular structural arrangements, which we validated by solid- state NMR. Structure predictions obtained from DMD/PRIME20 can now serve as a basis for choices of 13C- and 15N-labeled sites required for P1 structural NMR experiments, and provide direction for optimizing P3 aggregation experiments with the goal of obtaining a homogeneous sample suitable for structure determination. Taken together, DMD/PRIME20 results contribute

170

new insights into cytotoxic myocilin aggregates, specifically, by predicting possible molecular arrangements of the amyloid core structure of the full-length mOLF for further study.

5.5 Acknowledgement

This work was supported by National Institutes of Health (NIH) Grant RO1EB006006 (to

C. K. H), by NIH R01EY021205 (to R.L.L.) and by National Institute on Aging of NIH

R01AG045703 (to A. K. P). Partial support was provided by the National Science Foundation

Research Triangle Materials Research Science and Engineering Centers Grant DMR-1121107.

The authors declare that they have no conflicts of interest.

171

5.6 References

1. Hill, S. E.; Donegan, R. K.; Lieberman, R. L. The Glaucoma-Associated Olfactomedin

Domain of Myocilin Forms Polymorphic Fibrils That Are Constrained by Partial Unfolding

and Peptide Sequence. J. Mol. Biol. 2014, 426, 921-935.

2. Orwig, S. D.; Perry, C. W.; Kim, L. Y.; Turnage, K. C.; Zhang, R.; Vollrath, D.; Schmidt-

Krey, I.; Lieberman, R. L. Amyloid Fibril Formation by the Glaucoma-Associated

Olfactomedin Domain of Myocilin. J. Mol. Biol. 2012, 421, 242-255.

3. Tamm, E. R. Myocilin and Glaucoma: Facts and Ideas. Prog. Retin. Eye Res. 2002, 21,

395-428.

4. Acott, T. S.; Kelley, M. J.; Keller, K. E.; Vranka, J. A.; Abu-Hassan, D. W.; Li, X. B.; Aga,

M.; Bradley, J. M. Intraocular Pressure Homeostasis: Maintaining Balance in a High-

Pressure Environment. J. Ocul. Pharmacol. Ther. 2014, 30, 94-101.

5. Yam, G. H. F.; Gaplovska-Kysela, K.; Zuber, C.; Roth, J. Aggregated Myocilin Induces

Russell Bodies and Causes Apoptosis - Implications for the Pathogenesis of Myocilin-

caused Primary Open-angle Glaucoma. Am. J.Pathol. 2007, 170, 100-109.

6. Liu, Y. H.; Vollrath, D. Reversal of Mutant Myocilin Non-secretion and Cell Killing:

Implications for Glaucoma. Hum. Mol. Genet. 2004, 13, 1193-1204.

7. Zhou, Z. H.; Vollrath, D. A Cellular Assay Distinguishes Normal and Mutant

TIGR/Myocilin Protein. Hum. Mol. Genet. 1999, 8, 2221-2228.

8. Stothert, A. R.; Suntharalingam, A.; Huard, D. J. E.; Fontaine, S. N.; Crowley, V. M.;

Mishra, S.; Blagg, B. S. J.; Lieberman, R. L.; Dickey, C. A. Exploiting the Interaction

Between Grp94 and Aggregated Myocilin to Treat Glaucoma. Hum. Mol. Genet. 2014, 23,

6470-6480.

172

9. Suntharalingam, A.; Abisambra, J. F.; O'Leary, J. C.; Koren, J.; Zhang, B.; Joe, M. K.; Blair,

L. J.; Hill, S. E.; Jinwal, U. K.; Cockman, M. et al. Glucose-regulated Protein 94 Triage of

Mutant Myocilin through Endoplasmic Reticulum-associated Degradation Subverts a

More Efficient Autophagic Clearance Mechanism. J. Biol. Chem. 2012, 287, 40661-40669.

10. Baldwin, A. J.; Knowles, T. P. J.; Tartaglia, G. G.; Fitzpatrick, A. W.; Devlin, G. L.;

Shammas, S. L.; Waudby, C. A.; Mossuto, M. F.; Meehan, S.; Gras, S. L. et al. Metastability

of Native Proteins and the Phenomenon of Amyloid Formation. J. Am. Chem. Soc. 2011,

133, 14160-14163.

11. Hartl, F. U.; Hayer-Hartl, M. Converging Concepts of Protein Folding in vitro and in vivo.

Nat. Struct. Mol. Biol. 2009, 16, 574-581.

12. Tycko, R. Molecular Structure of Amyloid Fibrils: Insights from Solid-state NMR. Q. Rev.

Biophys. 2006, 39, 1-55.

13. Knowles, T. P. J.; Vendruscolo, M.; Dobson, C. M. The Amyloid State and its Association

with Protein Misfolding Diseases. Nat. Rev. Mol. Cell Biol. 2014, 15, 384-396.

14. Donegan, R. K.; Hill, S. E.; Freeman, D. M.; Nguyen, E.; Orwig, S. D.; Turnage, K. C.;

Lieberman, R. L. Structural Basis for Misfolding in Myocilin-associated Glaucoma. Hum.

Mol. Genet. 2015, 24, 2111-2124.

15. Gremer, L.; Schölzel, D.; Schenk, C.; Reinartz, E.; Labahn, J.; Ravelli, R. B. G.; Tusche,

M.; Lopez-Iglesias, C.; Hoyer, W.; Heise, H.; Willbold, D.; Schröder, G. F. Fibril Structure

of Amyloid-β(1-42) by Cryo-electron Microscopy. Science 2017, 358, 116-119.

16. Wälti, M. A.; Ravotti, F.; Arai, H.; Glabe, C. G.; Wall, J. S.; Böckmann, A.; Güntert, P.;

Meier, B. H.; Riek, R. Atomic-resolution Structure of a Disease-relevant Aβ(1–42)

Amyloid Fibril. Proc. Natl. Acad. Sci. U. S. A. 2016, 113, E4976-E4984.

173

17. Colvin, M. T.; Silvers, R.; Ni, Q. Z.; Can, T. V.; Sergeyev, I.; Rosay, M.; Donovan, K. J.;

Michael, B.; Wall, J.; Linse, S.; Griffin, R. G. Atomic Resolution Structure of

Monomorphic Aβ42 Amyloid Fibrils. J. Am. Chem. Soc. 2016, 138, 9663-9674.

18. Nguyen, H. D.; Hall, C. K. Molecular Dynamics Simulations of Spontaneous Fibril

Formation by Random-coil Peptides. Proc. Natl. Acad. Sci. U. S. A. 2004, 101, 16180-

16185.

19. Cheon, M.; Chang, I.; Hall, C. K. Extending the PRIME Model for Protein Aggregation to

All 20 Amino Acids. Proteins: Struct., Funct., Bioinf. 2010, 78, 2950-2960.

20. Alder, B. J.; Wainwright, T. E. Stuides in Molecular Dynamics .1. General Method. J. Chem.

Phys. 1959, 31, 459-466.

21. Smith, A. V.; Hall, C. K. Alpha-helix Formation: Discontinuous Molecular Dynamics on

an Intermediate-resolution Protein Model. Proteins: Struct., Funct., Genet. 2001, 44, 344-

360.

22. Nguyen, H. D.; Hall, C. K. Spontaneous Fibril Formation by Polyalanines; Discontinuous

Molecular Dynamics Simulations. J. Am. Chem. Soc. 2006, 128, 1890-1901.

23. Wang, Y.; Shao, Q.; Hall, C. K. N-terminal Prion Protein Peptides (PrP(120–144)) Form

Parallel In-register β-Sheets via Multiple Nucleation-dependent Pathways. J. Biol. Chem.

2016, 291, 22093-22105.

24. Muñoz, V.; Serrano, L. Elucidating the Folding Problem of Helical Peptides using

Empirical Parameters. III.Temperature and pH Dependence. J. Mol. Biol. 1995, 245, 297-

308.

25. Schaefer, J.; Stejskal, E. O.; Buchdahl, R. High-resolution Carbon-13 Nuclear Magnetic

Resonance Study of Some Solid, Glassy Polymers. Macromolecules 1975, 8, 291-296.

174

26. Yannoni, C. S. High-resolution NMR in Solids: the CPMAS Experiment. Acc. Chem. Res.

1982, 15 (7), 201-208.

27. Bennett, A. E.; Rienstra, C. M.; Auger, M.; Lakshmi, K. V.; Griffin, R. G. Heteronuclear

Decoupling in Rotating Solids. J. Chem. Phys. 1995, 103, 6951-6958.

28. Cheon, M.; Chang, I.; Hall, C. K. Spontaneous Formation of Twisted A beta(16-22) Fibrils

in Large-Scale Molecular-Dynamics Simulations. Biophys. J. 2011, 101, 2493-2501.

29. Cheon, M.; Hall, C. K.; Chang, I. Structural Conversion of A beta(17-42) Peptides from

Disordered Oligomers to U-Shape Protofilaments via Multiple Kinetic Pathways. PLOS

Comput. Biol. 2015, 11, e1004258.

30. Wang, Y.; Latshaw, D. C.; Hall, C. K. Aggregation of Aβ(17–36) in the Presence of

Naturally Occurring Phenolic Inhibitors Using Coarse-Grained Simulations. J. Mol. Biol.

2017, 429, 3893-3908.

31. Paravastu, A. K.; Leapman, R. D.; Yau, W. M.; Tycko, R. Molecular Structural Basis for

Polymorphism in Alzheimer's Beta-amyloid Fibrils. Proc. Natl. Acad. Sci. U. S. A. 2008,

105, 18349-18354.

32. Petkova, A. T.; Ishii, Y.; Balbach, J. J.; Antzutkin, O. N.; Leapman, R. D.; Delaglio, F.;

Tycko, R. A Structural Model for Alzheimer's Beta-amyloid Fibrils Based on Experimental

constraints from solid state NMR. Proc. Natl. Acad. Sci. U. S. A. 2002, 99, 16742-16747.

33. Fritzsching, K. J.; Yang, Y.; Schmidt-Rohr, K.; Hong, M. Practical Use of Chemical Shift

Databases for Protein Solid-state NMR: 2D Chemical Shift Maps and Amino-acid

Assignment with Secondary-structure Information. J. Biomol. NMR 2013, 56, 155-167.

34. Wishart, D. S.; Bigam, C. G.; Holm, A.; Hodges, R. S.; Sykes, B. D. 1H, 13C and 15N

Random Coil NMR Chemical Shifts of the Common Amino Acids. I. Investigations of

175

Nearest-Neighbor Effects. J. Biomol. NMR 1995, 5, 67-81.

35. Ulrich, E. L.; Akutsu, H.; Doreleijers, J. F.; Harano, Y.; Ioannidis, Y. E.; Lin, J.; Livny, M.;

Mading, S.; Maziuk, D.; Miller, Z. et al. BioMagResBank. Nucleic Acids Res. 2008, 36,

D402-D408.

176

Figure 5.1. mOLF peptide stretches P1 and P3 recapitulate disparate morphologies of amyloid aggregates derived from full-length mOLF seen by AFM. (A) mOLF fibrils grown at 37 °C with mechanical rocking exhibit straight morphology by AFM while (B) mOLF fibrils grown at 42 °C without rocking exhibit disparate circular morphology. (C) Fibrilization of predicted amyloidogenic peptide stretches, P1, P2 and P3, incubated in 10mM Na2HPO4/KH2PO4 pH 7.2 buffer containing 200mM NaCl at 36 °C, monitored by ThT fluorescence. (D) Amyloidogenic peptide stretches P1 (green) and P3 (purple) identified in (C) highlighted in the context of the native mOLF structure (PDB code 4WXS). (E) P1 fibrils appear straight while (F) P3 fibrils are circular when visualized by AFM. Scale bar is 300 nm for images A, B, E, and F. Images A, B, C, E, F reproduced from ref 1. Reprinted from Journal of Molecular Biology, 426, S. E. Hill, R. K. Donegan and R. L. Lieberman, The Glaucoma-Associated Olfactomedin Domain of Myocillin Forms Polymorphic Fibrils that are Constrained by Partial Unfolding and Peptide Sequence, pages 921-935, Copyright (2014), with permission from Elsevier.

177

Figure 5.2. DMD/PRIME20 simulations of P1. (A) A plot of potential energy of P1 peptide aggregation versus simulation time. Also shown are simulation snapshots of eight random coil P1 peptides aggregating to form U-shaped (top panel corresponding to the red curve) and S-shaped (bottom panel corresponding to the black curve) P1 protofilaments. Each of the eight peptides has a distinct color. (B) and (C) are the final simulation snapshots of the S-shaped (B) and U-shaped (C) P1 protofilaments, respectively. (D) and (E) are the schematic representation of peptide conformation in the U-shaped and S-shaped P1 protofilaments. Hydrophobic and polar residues are shown in white and green, respectively.

178

Figure 5.3. Average number of inter-peptide backbone hydrogen bonds formed per residue of P1 (A) P2 (B) P3 (C). (D) β-sheet propensities calculated for P1, P2 and P3 peptides.

179

Figure 5.4. DMD/PRIME20 simulations of P3. (A) A plot of potential energy of P3 peptide aggregation versus simulation time. Also shown are simulation snapshots of eight random coil P3 peptides aggregating to form a disordered oligomer (top panel corresponding to the red curve) and a U-shaped P3 protofilament (bottom panel corresponding to the black curve). Each of the eight peptides has a distinct color. (B) and (C) are final simulation snapshots of the disordered P3 oligomer and U-shaped P3 protofilament, respectively. (D) The schematic representation of peptide conformation in the U-shaped P3 protofilaments. Hydrophobic and polar residues are shown in white and green, respectively.

180

Figure 5.5. 1H-13C CPMAS 13C spectrum of P1 and P3 aggregates with no 13C enrichment. (A) The spectrum of P1 fibril (red) and the predicted spectrum (generated by summing all Gaussian peaks at the chemical shifts of all 13C in the peptides) with FWHM = 1.2 ppm (black). (B) The region of aliphatic carbons in Panel A. (C) The predicted CPMAS spectrum of P3 fibril (red) and the predicted spectrum with FWHM = 14 ppm (black).

181

Table 5.1. The secondary chemical shifts of the distinguishable 13C peaks in the CPMAS spectrum of P1 aggregates. For residues within β-strands, CO and Cα peak frequencies are expected to be at least 0.5ppm under corresponding random coil values (measured from random coil model peptides) 33-34 and Cβ peak frequencies are expected to be at least 0.5 ppm above corresponding random coil values.

Random Coil Measured Chemical Secondary Chemical Carbon Shift (ppm) Shift (ppm) Shift (ppm)

S Cβ 62.1 64.3 +2.2

V Cα 60.5 58.9 -1.6

A Cα 50.8 49.8 -1.0

Q Cβ 27.7 30.6 +2.9

182

5.7 Supporting Information

Figure S5.1. Clustal Omega1 sequence alignment depicting bioinformatics analysis of amyloid forming region of mOLF by AmylPred 22 (amyl, pink), Zyggregator3 (Zygg, cyan), PASTA4 (Past, purple), Tango5 (yellow), and ZipperDB6 (Zipp, green).

183

Figure S5.2. Simulation snapshots (A-J) of P1 aggregates from ten independent DMD simulations.

184

Figure S5.3. Simulation snapshots (A-J) of P2 aggregates from ten independent DMD simulations.

185

Figure S5.4. Simulation snapshots (A-J) of P3 aggregates from ten independent DMD simulations.

186

References

1. Sievers, F.; Wilm, A.; Dineen, D.; Gibson, T. J.; Karplus, K.; Li, W.; Lopez, R.; McWilliam,

H.; Remmert, M.; Soding, J.; Thompson, J. D.; Higgins, D. G. Fast, Scalable Generation

of High-quality Protein Multiple Sequence Alignments Using Clustal Omega. Mol. Syst.

Biol. 2011, 7, 539.

2. Tsolis, A. C.; Papandreou, N. C.; Iconomidou, V. A.; Hamodrakas, S. J. A Consensus

Method for the Prediction of 'Aggregation-prone' Peptides in Globular Proteins. PLoS One

2013, 8, e54175.

3. Tartaglia, G. G.; Vendruscolo, M. The Zyggregator Method for Predicting Protein

Aggregation Propensities. Chem. Soc. Rev. 2008, 37, 1395-1401.

4. Walsh, I.; Seno, F.; Tosatto, S. C.; Trovato, A. PASTA 2.0: an Improved Server for Protein

Aggregation Prediction. Nucleic Acids Res. 2014, 42, W301-W307.

5. Fernandez-Escamilla, A. M.; Rousseau, F.; Schymkowitz, J.; Serrano, L. Prediction of

Sequence-dependent and Mutational Effects on the Aggregation of Peptides and Proteins.

Nat. Biotechnol. 2004, 22, 1302-1306.

6. Goldschmidt, L.; Teng, P. K.; Riek, R.; Eisenberg, D. Identifying the Amylome, Proteins

Capable of Forming Amyloid-like Fibrils. Proc. Natl. Acad. Sci. U.S.A. 2010, 107, 3487-

3492.

187

CHAPTER 6

Thermodynamic phase diagram of amyloid-β (16-22) peptide

Chapter 6 is essentially a manuscript by Yiming Wang, Samuel Bunce, Sheena E. Radford,

Andrew J. Wilson, Stefan Auer and Carol K. Hall submitted to Proceedings of the National

Academy of Sciences of the United States of the America.

188

Thermodynamic phase diagram of amyloid-β (16-22) peptide

Yiming Wang1, Samuel Bunce2,3, Sheena E. Radford2,4, Andrew J. Wilson2,3, Stefan Auer 3, Carol K. Hall1*

1. Department of Chemical and Biomolecular Engineering, North Carolina State University, Raleigh, North Carolina 27695-7905, United States

2. Astbury Centre for Structural Molecular Biology, University of Leeds, Woodhouse Lane, Leeds LS2 9JT, United Kingdom.

3. School of Chemistry, University of Leeds, Woodhouse Lane, Leeds LS2 9JT, United Kingdom.

4. School of Molecular and Cellular Biology, Faculty of Biological Sciences, University of Leeds, Woodhouse Lane, Leeds LS2 9JT, United Kingdom.

Abstract

The aggregation of monomeric Aβ peptide into oligomers and amyloid fibrils in the mammalian brain is associated with Alzheimer’s disease. Insight into the thermodynamic stability of the Aβ peptide in different polymeric states is fundamental to defining and predicting the aggregation process. Experimental determination of Aβ thermodynamic behavior is challenging due to the transient nature of Aβ oligomers and the low peptide solubility. Furthermore, quantitative calculation of a thermodynamic phase diagram for a specific peptide requires extremely long computational times. Here, using a coarse-grained protein model, molecular dynamics simulations are performed to determine an equilibrium concentration and temperature phase diagram for the amyloidogenic peptide fragment, Aβ16-22. Our results reveal that the only thermodynamically stable phases are the solution phase and the macroscopic fibrillar phase, and that there also exists a hierarchy of metastable phases. The boundary line between the solution

189

phase and fibril phase is found by calculating the temperature-dependent solubility of a macroscopic Aβ16-22 fibril consisting of an infinite number of β-sheet layers. To our knowledge, this is the first in silico determination of an equilibrium (solubility) thermodynamic phase diagram for a real amyloid-forming peptide. Furthermore, the in silico prediction of Aβ16-22 solubilities over the temperature range of 277-330K agrees well with fibrillation experiments and transmission electron microscopy measurements of the fibril morphologies formed. This in silico approach of predicting peptide solubility is also potentially useful for optimizing biopharmaceutical production and manufacturing nanofiber scaffolds for tissue engineering.

190

6.1 Significance Statement

Phase diagrams of atomic systems are calculated routinely by computer simulations, but such calculations are absent for even the simplest peptides. Previous simulations are mainly non- equilibrium, and focus on the assembly of peptides from the monomeric to the aggregated state.

To obtain accurate equilibrium solubilities, however, it is necessary to simulate many assembly and disassembly events of fibrillar aggregates, which is notoriously difficult, as it requires breaking many hydrogen bonds. We overcome these challenges and calculate the equilibrium phase diagram of Aβ16-22, the archetypal amyloid former, in the first quantitative calculation of a peptide phase diagram using a realistic protein model. Importantly, our prediction of Aβ16-22 solubility over temperatures from 296K to 330K agrees well with experimental measurements.

6.2 Introduction

Aβ16-22 is a 7-residue amyloidogenic peptide (Ac-K-L-V-F-F-A-E-CONH2) comprising the central, fibril-forming core of the full-length Aβ peptide, a major constituent of the extracellular plaques associated with Alzheimer’s disease. (1, 2) Aβ16-22 has been widely studied due to its relative ease of synthesis and its ability to form well-characterized antiparallel β-sheet fibril structures at concentrations above 55 μM.(3, 4) At neutral pH, monomeric Aβ16-22 is predicted to adopt a random coil configuration that then oligomerizes, passing through an intermediate out-of- register state prior to the final in-register anti-parallel alignment. (5-8) Atomistic simulations of a small number of peptides on the nanosecond time-scale are possible and have been used to examine the structural stability of a variety of Aβ16-22 oligomers, including β-barrels(9), β-sheet-rich dimers(10), trimers(11, 12) and hexamers.(13)-(14) Coarse-grained molecular dynamics (MD) simulations have succeeded in predicting spontaneous formation of octamer(15), tri-, tetra- (6) and

191

hepta-layer(16) twisted Aβ16-22 nanofibrils involving up to 192 peptides on the microsecond time- scale.(17) A prerequisite to understanding which structures Aβ16-22 peptides form, and under what conditions, is knowledge of the phase diagram, which characterizes the thermodynamic stability of the peptide phases as a function of relevant parameters, most importantly temperature and peptide concentration. However, despite being studied under a wide range of conditions, a quantitative thermodynamic phase diagram has never been generated for Aβ16-22, or any other peptide, using a realistic protein model.

In general, experiments (18-20) determining the coexistence lines (solubilities) between oligomeric or fibrillar aggregates and the protein solution are difficult to perform because of the metastable nature of the oligomers and the low solubilities of fibrils. Most experiments are performed by preparing a solution at a given concentration and temperature and left for up to weeks to determine whether oligomeric or fibrillar aggregates have formed. (21, 22) The so-observed concentrations and temperatures at which fibrils form can often be far from the true equilibrium concentrations (solubilities) as the observed structures can be kinetically trapped and thus not representative of the thermodynamically most-stable state (23). Experiments in which pre-formed fibrils are put in solution and the concentration at which they elongate or dissolve is measured can also provide estimates of solubility (24, 25).

The challenge in determining a peptide phase diagram for realistic protein models in silico is the long computational time needed to calculate the solubility. Molecular simulations aimed at constructing phase diagrams of peptides (26, 27) are all non-equilibrium and only simulate the self-assembly of peptides from the monomeric to the aggregate state. There is no guarantee, however, that structures so obtained will coincide with the thermodynamically stable phases.

Furthermore, previous studies do not consider the existence of meta-stable phases. It has only

192

recently become possible to perform Monte Carlo simulations that are capable of constructing a concentration and temperature phase diagram at thermodynamic equilibrium for simple generic peptide models. (28, 29) Given the complexity of biomolecules, the relevance of this finding to any real peptide system is not clear.

Here we construct a temperature-concentration phase diagram for the Aβ16-22 peptide by determining the solubility of oligomeric aggregates and fibrils using coarse-grained molecular dynamics and classical nucleation theory. The phase diagram reveals that the solution phase and fibril phase are the only two thermodynamic stable phases, and that there also exists a hierarchy of metastable aggregate phases. The thermodynamic boundary between the solution phase and the fibril phase is taken to be the solubility line of an infinite layer β-sheet fibril predicted using discontinuous molecular dynamics simulation and classical nucleation theory. The strength and soundness of this approach lie in the fact that the our force field, PRIME20, is knowledge-based, sequence specific and accurately predicts peptide orientation, strand-strand and sheet-sheet distances of Aβ16-22 fibril structures. (6) To validate the phase diagram predicted in silico, we conduct Aβ16-22 fibrillation experiments at peptide concentrations varying from 10 μM to 200μM at 277K-330K. Judging from TEM images, the experimental conditions under which the Aβ16-22 solution forms fibrils agrees well with the predicted solution-fibril phase boundary. To the authors’ knowledge, this is the first in silico determination of the thermodynamic phase diagram for a real amyloid-forming peptide and the first to be validated by experimental measurements.

193

6.3 Results and Discussions

6.3.1 Protein model, Structure of Aβ16-22 oligomer and fibril

The Aβ16-22 aggregates are modelled using a coarse-grained protein model, called

PRIME20, in which each amino acid residue is represented by three backbone spheres, one each for N-H, C-H, and C=O, and a single side-chain sphere R (Fig. 1A). The non-directional interaction between sidechain spheres and the directional hydrogen bond between N-H and C=O spheres are modelled as square-well potentials (Fig. 1B and 1C). While the strength of the directional backbone hydrogen bonding interaction is the same for all residues, the strength of the non- directional sidechain-sidechain interaction is sidechain specific. See Methods for more details.

Here we focus on five types of aggregates formed by Aβ16-22: non-hydrogen-bonded (non-

HB) oligomers (Fig1D), hydrogen-bonded (HB) oligomers (Fig 1E), and fibrils composed of 2, 3 and 4 β-sheet layers (Fig. 1F, G, H). The non-HB oligomer is defined to be an aggregate formed solely by inter-peptide hydrophobic sidechain-sidechain (HP) interactions with no inter-peptide hydrogen bonds; intra-peptide hydrogen bonds are allowed. The HB oligomer is defined to be an aggregate that has at most one inter-peptide hydrogen bond between any two neighboring peptides and any number of intra-peptide hydrogen bonds. We consider the non-HB and HB oligomers to be two distinct states. Although they may look like a continuum of states in Fig. 1, peptides in these two states have different numbers of hydrogen bonds and hydrophobic sidechain contacts with neighboring peptides (see Fig. 2). As shown in Fig. 2A (orange curve) and 2B, peptides in the non-HB oligomer on average have one intra-peptide helical HB formed mainly between Leu17 and Ala21 and sometimes between Val18 and Glu22. The HB oligomer, as shown in Figs. 2A-D, is stabilized by having both a large number of inter-peptide HP contacts (mainly by Leu17, Phe19 and Phe20) and a small number of both inter- and intra-peptide HBs. The β-sheet aggregates are

194

stabilized by having sidechain HP contacts between tightly-packed anti-parallel β-sheets and a maximum number of backbone HBs between fully-extended peptides within each β-sheet, consistent with X-ray diffraction and solid-state NMR measurements (2, 6). Our simulations show that the probability distributions for the number of HB and HP sidechain contacts for the five aggregates do not change significantly with temperature over the studied temperature range.

6.3.2 Measurement of solubility for Aβ16-22 oligomers and fibrils

In order to determine the solubility of aggregates (oligomers and fibrils) formed by the

Aβ16-22 peptides, we perform molecular dynamics simulations in the canonical ensemble. The solubility is defined to be the peptide concentration above which the peptide will join with others to form an aggregate and below which the peptide will remain in solution. In our simulations, the solubilities for the different Aβ16-22 aggregates (oligomers and fibrils) at a given temperature are measured as the equilibrium monomer concentration at which the aggregate neither grows nor shrinks. (See Methods for more details). This approach was used by Bai and Li (30) to measure solid-liquid phase equilibrium for the Lennard-Jones (LJ) fluid. Their calculation of the solid- liquid interfacial energy (surface tension) using classical nucleation theory was in qualitative agreement with results calculated using other theoretical approaches, e.g. interface fluctuation method (31) and reversible work integration (32). Later, Auer and co-workers used this method along with Monte Carlo simulation to calculate a phase diagram for a simple homo-polypeptide model where each amino acid is represented by one Cα bead (29). Their predicted values for the surface tension were in agreement with theoretical predictions for protein crystals using a simple lattice model (33).

195

We first measured the solubility for the non-HB and HB oligomers. Fig. 3A shows that the solubilities for the non-HB and HB oligomers increase with increasing temperature. This is expected as at higher temperatures the thermal energy makes it easier for peptides to detach from the oligomer so that higher concentrations are required to stabilize these aggregates. The latent heat of aggregation for monomeric peptides into oligomers can be obtained by fitting the van’t

Hoff equation,

퐿 퐶푒 = 퐶푟exp(− ) (1) 푘퐵푇 to the simulation data (Fig. 3B), where 퐶푒 is the equilibrium fibril solubility, 퐶푟 is a temperature- independent reference concentration, kB is the Boltzmann constant, 푇 is the simulation temperature, and 퐿 is the latent heat of monomer peptide aggregation into the oligomer or fibril.

From the simulation, we also calculated the average peptide binding energy (the sum of the HP and HB energies per peptide) within the non-HB and HB oligomers, including all the peptides on the surface and inside the aggregate, and found that the latent heat is about half of the average peptide binding energy EB, (see Fig. 3C) consistent with earlier work(29).

Although the solubility of the oligomers should depend on the oligomer size as described by the Ostwald formula , (28, 29, 34) (which predicts that solubility decreases with increasing oligomer size) , we do not consider the effect of oligomer size here . This is because large oligomers do not form for this Aβ16-22 system. The simulations show that the non-HB and HB Aβ16-

22 oligomers are relatively small (containing around ten peptides) and easily break apart into smaller aggregates when they contain more than fifteen peptides.

Next we determined the solubilities for 2, 3, 4 and ∞ β-sheet Aβ16-22 fibrils. The simulation results (Fig. 3A) show that at a given temperature, the solubilities for the 2, 3 and 4 β-sheets decrease with increasing thickness (number of β-sheet layers) of the fibril. As the number of β-

196

sheets in the fibril increases, the average number of HP contacts per peptide increases, as shown in Fig. 2C, leading to increased stability of the fibril. A quantitative description of the dependence of the solubility of a fibril with i layers (i=1, 2, 3, etc.) of a β-sheet, 퐶𝑖훽, on thickness at fixed temperature can be derived from classical nucleation theory and is given by (35)

2휓ℎ 1 퐶𝑖훽 = 퐶∞훽exp ( ∙ ) (2) 푘퐵푇 𝑖 where 퐶∞훽 is the solubility of an infinite thick fibril, 휓ℎ = 푎ℎ휎ℎ is the fibril surface energy parallel to the fibril thickening axis, 휎ℎ is the specific surface energy of the face perpendicular to the fibril axis, and 푎ℎ is the lateral surface area occupied by each peptide within one β-sheet as defined in Auer et al (35). A linear fit of our simulation data (푙푛퐶𝑖훽 vs 1/i) to Eq. (2) (Fig. 3D) yields an estimate for the solubility 퐶∞훽 as a function of temperature for the infinitely thick fibril.

As before, the latent heats of peptide aggregation into fibril phases from solution can be obtained by a fit of our data to Eq. (1), see Fig. 3D, and the values obtained are shown in Fig. 3C. From Fig.

3C, we find that for 2, 3 and 4 β-sheet fibrils, the values of the latent heat are comparable to the average peptide binding energy. The relation between latent heat and protein-protein interaction energy for a protein crystal was first derived by Haas and Drenth using a simple lattice model (33).

They found that the latent heat should be half of the average protein-protein interaction energy; this was derived by using the zero-temperature approximation, which does not consider the entropic effects. We speculate that it might be the neglect of the entropic effect that causes the relation between latent heat and peptide binding energy to hold for oligomers but not for 2, 3 and

4 β-sheet fibrils in this work. In addition, the fit of our simulation data to Eq. (2) also yields values for 휓ℎ = 푎ℎ휎ℎ, and knowledge of 푎ℎ enables us to estimate the average surface tension 휎ℎ of the

Aβ16-22 fibril (Fig. S1). The lateral surface area per peptide, 푎ℎ = 106.1Å, is the product of the length of a peptide within a β-sheet (22.1Å) and the inter-peptide distance in one β-sheet (4.8Å).

197

Hence, depending on temperature, the average surface tension 휎ℎ for the Aβ16-22 fibril ranges from

24 to 30 mJ/m2 (see Fig. S1), which is in the range of 0.1-30 mJ/m2 reported for protein crystals in aqueous solution (33, 36-38).

6.3.3 Aβ16-22 phase diagram

The major finding of this study is our quantitative calculation of the thermodynamic phase diagram for the Aβ16-22 peptide fragment (Fig. 4A). The peptide solution phase (cyan region) and the macroscopic fibrillar phase (yellow region) are the only two thermodynamically stable phases.

The thermodynamic boundary between these phases is taken to be the solubility line for a fibril with an infinite number of β-sheets. (Error bars are not shown in Figs 4 and 5(A) because the R2 value for the linear fit of 퐶∞훽 vs T data to obtain 퐶∞훽(T) is close to 1). Within the fibril phase region of the phase diagram, there exists a series of metastable aggregate phases separated by the metastable solubility lines for the non-HB oligomer, HB oligomer and 2, 3 and 4 β-sheet fibrils.

To justify the infinite-thickness approximation in calculating fibril solubility, we calculated

Aβ16-22 fibril solubility versus fibril thickness (Fig. S2) and found that fibril solubility remains unchanged after the thickness exceeds 30 layers. We think that an actual Aβ16 -22 fibril may contain more than 30 layers of β-sheets at neutral pH, because Aβ1-42 (39), Aβ10-35 (40) and Aβ18-28 (41) form fibrils in vitro with 2, 6 and 24 layers of β-sheets, respectively at neutral pH. This is consistent with a hypothesis by Lu et al. (42) that shorter amyloidogenic peptides tend to adopt more planar

β-sheets and maintain more layers in the fibril than longer peptides.

To better illustrate why the solubility line for the fibril is regarded as the coexistence line, and why the oligomer and various multi-sheet fibril phases are metastable with respect to the fibril phase, we put our calculations in the context of more familiar thermodynamic arguments in which

198

the compositions of two phases in equilibrium are found by minimizing a free energy. Figure S3 plots the Helmholtz free energies of the solution, oligomer and fibril phases as a function of the peptide mole fraction, 푥푝, (which is simply related to the peptide concentration) in a peptide-water mixture at fixed temperature and volume. A double tangent (also called convex-envelop) construction on the free energies of the solution and fibril phase determines the peptide mole

푠표푙푛(1) 푓𝑖푏 fractions, 푥푝 and 푥푝 at which the solution and fibril phases are in equilibrium. The construction locates the peptide mole fractions at which peptide chemical potentials (the 푥푝= 1

푠표푙푛(1) 푠표푙푛(1) 푓𝑖푏 푓𝑖푏 intercepts of lines tangent to the free energy curves) are equal, i.e. 휇푝 (푥푝 ) = 휇푝 (푥푝 )

푠표푙푛(1) . The composition, 푥푝 , is equivalent to the peptide concentration above which the monomer phase transitions to another phase, the solubility line determined in our simulations. Similarly a double tangent construction on the solution and oligomer phases yields peptide mole fraction

푠표푙푛(2) 표푙𝑖푔 푠표푙푛(2) 푠표푙푛(2) 표푙𝑖푔 표푙𝑖푔표 푥푝 and 푥푝 with equal peptide chemical potentials, 휇푝 (푥푝 ) = 휇푝 (푥푝 ), signifying equilibrium between solution and oligomer phase in this concentration range. This is,

푠표푙푛(1) however, a metastable equilibrium compared to the equilibrium between solution at 푥푝 and

푓𝑖푏 fibril at 푥푝 because the peptide chemical potential in solution is higher than that when the

푠표푙푛(2) 푠표푙푛(2) solution is in equilibrium with the fibril. Note also that having 휇푝 (푥푝 ) >

푠표푙푛(1) 푠표푙푛(1) 푠표푙푛(2) 푠표푙푛(1) 휇푝 (푥푝 ) means that 푥푝 >푥푝 , consistent with what is measured in simulations.

푠표푙푛(1) 푠표푙푛(2) Notice also that in Figure S3, there are no metastable phases when 푥푝 < 푥푝 < 푥푝 ; the

푠표푙푛(2) oligomer phase only becomes metastable with respect to the fibril phase when 푥푝 > 푥푝 . This is also reflected in the solubility versus temperature phase diagram, Fig. 4A, which shows a region of concentrations where the HB oligomer is soluble but unstable (above the solubility line), and a

199

region where the oligomer is metastable (above the dashed purple line). Similar arguments can be made for the other metastable phases, such as the 2, 3, and 4 β-sheet fibrils phases. We emphasize that we did not calculate free energies, as we were able to calculate the solubilities directly; the discussion here is mainly presented for pedagogical purposes.

Knowledge of the thermodynamically stable and metastable phases in the phase diagram allow us to determine the conditions under which the oligomer or fibril can form. A previous discontinuous molecular dynamics (DMD) study of Aβ16-22 kinetic aggregation by Cheon et al. (6) found that at 20mM and T*=0.2 corresponding to 342K (Fig. 4B top panel), a system of peptides in a random configuration first forms a fibril nucleus which then grows into fibrils. At the same concentration but at a lower temperature T*=0.17 corresponding to T=273K (Fig. 4B, bottom panel), the peptides first form HB oligomers that later merge and rearrange to form fibrils. Our thermodynamic phase diagram helps explain why Aβ16-22 aggregates via a “one-step” pathway at high temperature and a “two-step” pathway at low temperature. From Fig. 4, at point A (20mM,

T=342K), the peptide concentration is greater than the solubility of fibril, C>Ce,fibril(T=342K) but smaller than the solubility of the HB oligomer, CCe,fibril(273K) and C>Ce,HB oligomer

(T=273K) (34); this is conducive to two-step aggregation kinetics in which several oligomers from and later merge and rearrange to form fibrils. These arguments about the connection between the fibril formation kinetics and aggregate stability and metastability are meant here only to be suggestive. The best way to predict Aβ16-22 aggregation pathways at different temperatures and

200

concentrations would be to construct aggregation free energy landscapes along the appropriate reaction coordinates.

To validate the in silico prediction of Aβ16-22 solubility at biophysically relevant temperatures, transmission electron microscopy was used to determine whether fibrils had formed after a predetermined time (2 weeks incubation). Fig. 5 is a summary of our experimental results near the predicted solubility boundary line at temperatures ranging from 277K to 330K. At

T=330K, fibrils form at both 200μM and 300μM but no fibrils are observed when peptide concentration is at or below 100μM, which is consistent with the predicted solubility (177μM) at

330K. Furthermore, at 277K, 296K and 310K, fibrils form when the concentration is equal to or above 20μM, which also agrees with our predicted solubility at 277K, 396K and 310K of 0.2μM,

2.6μM and 16.5μM, respectively. It is important to note that even though fibrils were not observed in the TEM images at 10 μM at all temperatures tested, this does not rule out the presence of low populations of fibrils under these conditions which are below the detection limits of this method, i.e. low possible surface adsorption of the fibrils. Thus, we conclude that the boundaries predicted by the phase diagram quantitatively agree with the experimental findings, demonstrating the power of this approach to understand the thermodynamic stability of peptide assemblies.

The concentration and temperature conditions at which Aβ16-22 forms fibrils reported from a number of other in vitro studies also agree with our simulation prediction (Figs. 5 and S2). (2, 3,

8, 43, 44) It should be noted that the only studies included in this comparison use Aβ16-22 with capped N- and C-termini, since uncapping the termini can have a substantial twisting effect on the fibril supramolecular structure.(45) Most of these studies were performed at high concentrations

(>200μM), placing them in regions of the phase diagram that would predict fibril formation, even at the highest temperature studied (55 °C, 328K). (7) Only one study by Senguen et al., (3)

201

demonstrated that fibrils are formed at a concentration of 55 μM at 37 °C (310K), again in agreement with our phase diagram.(6)

Knowledge of the peptide solution-fibril phase boundary could allow us to modulate the quality of nanofiber structures formed. This can be accomplished by carefully controlling the driving force for fibril formation, e.g. highly-ordered fibrils can be obtained at solution conditions close to the phase boundary while less-ordered entangled fibrils can be obtained at conditions away from the phase boundary. This suggests that our computation-based approach could inform future fabrication of amyloid nanofiber substrates / scaffolds with desired microstructure properties for various tissue engineering and biomedical applications.(46, 47) In addition, controlling protein aggregation is a key problem for the purification and storage of therapeutic proteins.(48, 49) Thus, the knowledge of solubility of specific peptides and proteins over a wide range of temperatures could be used to effectively prevent their unwanted agglomeration.

Our approach can be applied to investigate the effect of mutations on peptide solubility.

For example, in preliminary simulations, we show that F19A mutant of Aβ16-22 remains soluble at

C=20mM and T=342K in contrast to the fibril-forming behavior of the wild type at the same conditions. In fact, our predicted solubility of the F19A mutant is about 600 mM at 307K. This result is consistent with experimental finding by Senguen et al. that the F19A mutant of Aβ16-22 is much more soluble (at least 350 μM) than its wild type at body temperature (3). One of our next steps will be to calculate the solubility of Aβ16-22 mutants at positions 19 and 20 and compare our results with experimental measurements.

202

6.4 Conclusion

In this work we demonstrated that by using a relatively realistic coarse-grained protein model it is possible to calculate via simulation an equilibrium concentration and temperature phase diagram for a short amyloidogenic peptide, Aβ16-22. The predicted phase diagram provides insight into the thermodynamic stability of the different types of aggregates formed by Aβ16-22 under biophysically relevant conditions. This helps us to understand the conditions under which they can form, and how to prevent them from forming. The solubilities predicted in silico are in the range expected from experiments and as such also provide a rigorous test of the validity of the model and methodology used. To our knowledge, this is the first in silico determination of an equilibrium

(solubility) thermodynamic phase diagram for a non-trivial peptide. Although peptide phase diagrams have been calculated via simulations (26, 27), they are usually based on the kinetics of phase formation as opposed to a direct solubility measurement. Furthermore, the in silico prediction of Aβ16-22 solubilities over the temperature range of 4-57°C agrees well with fibrillation experiments and transmission electron microscopy measurements of the fibril morphologies formed. This in silico approach of predicting peptide solubility is also potentially useful for engineering nanofiber scaffolds for biomedical applications as well as manufacturing biopharmaceuticals.

6.5 Methods

6.5.1 In silico peptide model

The molecular model that we use is a four-sphere-per-residue model in which each amino acid residue is represented by three backbone spheres, one each for N-H, C-H, and C=O, and one for the side-chain sphere, R. (50) For illustrative purposes, Fig. S4 shows a simulation snapshot of

203

the structure of a two-layer β-sheet fibril formed by Aβ16-22 peptides. The PRIME20 model has unique geometric and energetic parameters for each of the 20 amino acids. Specifically, each of the 20 different sidechain spheres (for the 20 amino acids) has a distinct hard sphere diameter

(effective van der Waals radius) and sidechain-to-backbone distances (R-Cα, R-NH, and R-CO).

The interactions between the atoms are described by discontinuous potentials including the hard sphere, square-well and square-shoulder potentials. The strength of the directional hydrogen bond is described by the well depth εHB and is the same for all residues, and the strength of the non- directional hydrophobic interaction between side chains i and j is described by a sidechain specific well depth (εHP-ij) . The potential energy parameters between the 20 different amino acids include

210 independent square well widths and 19 independent square well depths derived by using a machine-learning algorithm that optimizes the energy gap between 711 known native states from the PDB and decoy structures. (51) For example, the interaction strength, εHP-AA, between two alanine sidechains is 0.084εHB, where the hydrogen bonding strength εHB is chosen to be

12.47kJ/mol as was used in previous work(17). The details of the assignment of the values of the interaction parameters of the PRIME20 model are described in earlier work.(17, 50, 51) The

* reduced temperature is defined as T = kBT/εHB. It can be related to a real temperature by T [K] =

2288.46T*-115.79; this equation was obtained by matching the folding temperature of alanine-rich polypeptides in our previous discontinuous molecular dynamic (DMD) simulations (17) to the experimental values(52). The reduced time unit is Δtreduced = 0.96 ns, which has been obtained by matching the self-diffusion coefficient of Aβ16-22 obtained from DMD simulations to that calculated from atomistic MD simulation.(17)

DMD simulations with the PRIME20 model are ideally suited for simulating protein aggregation simulations in three respects. (i) They allow simulation of the complete aggregation

204

process from a system of random coil peptides to a fibrillar structure (up to 200 peptides) at time scales up to 100μs (16, 17). (ii) The PRIME20 model is realistic enough to ensure that the predicted fibril structures agree well with X-ray diffraction and solid-state NMR measurements (6). Its adequacy and efficiency has been demonstrated by its application to short peptide systems including Aβ16-22 (6), the designed hexapeptide sequences of Lopez de la Paz et al.(53, 54), the tau fragment(55), and longer peptide systems including Aβ17-36 (56) and Aβ17-42(57). (iii) The sidechain spheres of the twenty possible amino acids each have distinct geometric and energetic parameters, ensuring that the peptide is modelled in a sequence specific manner.

6.5.2 Simulation method

The solubilities for the different Aβ16-22 aggregates (oligomers and fibrils) were measured using an approach first proposed by Bai and Li (30) and later adopted by Auer and Kashchiev.(29)

The method is based on the definition of the solubility of an aggregate at a given temperature as the equilibrium monomer concentration at which the aggregate neither grows nor shrinks. The pre- formed aggregate is initially placed in the center of a cubic box surrounded by monomer peptides.

Periodic boundary conditions are imposed. At constant temperature, we constrain the fibril to avoid new β-sheet creation while allowing it to elongate or shrink; the concentration of monomeric peptides is monitored until it reaches a plateau and remains constant thereafter for a long time.

(Fig. S5) The solubility of the aggregate at any given temperature is defined to be the monomeric peptide concentration in that final equilibrium state. Although this method is simple in principle, it is very costly in terms of simulation time. In fact, to observe whether a pre-formed aggregate grows or shrinks requires simulation of hundreds of attachment and detachment events to and from the aggregate. Detachment events of peptides from fibrils are particularly rare as they require

205

breakage of several strong hydrogen bonds. Due to the infrequency of detachment events, the total number of peptides used in our simulations is relatively small (sixteen to eighty peptides).

Here we measured the solubilities of five types of Aβ16-22 aggregates including a non- hydrogen-bonded oligomer (non-HB oligomer), a hydrogen-bonded oligomer (HB oligomer), and fibrils containing 2, 3, and 4 β-sheets. The non-HB oligomer is defined to be an aggregate formed solely by inter-peptide hydrophobic sidechain-sidechain (HP) contacts; the formation of inter- peptide hydrogen bonds (HB) is forbidden but intra-peptide HBs are allowed. The HB oligomer is defined to be an aggregate formed by both HP contacts and inter-peptide HBs within which any two peptides are constrained to form at most one inter-peptide hydrogen bond. We choose to focus our attention on the non-HB and HB oligomers because they represent two limiting examples of the forces that can stabilize oligomers formed early in the fibril formation process. The solubility profiles of HB oligomers with more than one HB formed between any peptide pair are expected to be similar to the ones we study here and thus are omitted. The 2, 3 and 4 β-sheet fibrils are defined to be aggregates that contain two, three and four stacked β-sheets. The small non-HB and

HB oligomers contain 7 peptides and the number of monomer peptides in solution is 15. The 2, 3 and 4 β-sheet fibrils contain 24, 28 and 38 peptides, respectively, and the number of monomer peptides in solution is 40. The sizes of preformed fibrils are chosen to have around ten peptides in each β-sheet. The initial monomer peptide concentration is chosen to be relatively high so that the solution phase is supersaturated with respect to equilibrium state. In this way, we prevent the preformed fibrils from fully dissolving into monomers. The choice of a fibril with a different initial length should not affect the equilibrium monomer concentration (solubility) , as was verified in previous work by Auer et al(28). The total computation time for these simulations was 9 months using 25 central processing units (CPU), with each individual run lasting around 50μs to 100μs.

206

To measure the solubility of potentially metastable aggregates, constraints are required to avoid structural conversion of one type of metastable aggregate into another (i.e. the growth of a non-HB oligomer into a HB oligomer or the growth of a 2β-sheet fibril into a 3β-sheet fibril).

During measurement of the solubility of a non-HB oligomer, all peptides in the system can form intra-peptide hydrogen bonds, but not inter-peptide hydrogen bonds. The latter is accomplished by imposing constraints on the specific collision events that could lead to the formation of such bonds by turning off the attractive square well interactions between the C=O and NH spheres. Similarly, during measurement of the solubility of the HB oligomer, all peptide pairs in the system are constrained to form at most one inter-peptide hydrogen bond. During measurement of the solubility of fibrils with fixed thickness (2, 3, and 4 β-sheets), we start from a high initial peptide concentration to prevent fibrils from losing an existing β-sheet layer. The fibrils are also prevented from creating a new β-sheet layer during the solubility measurement by turning off the attractive square well interactions; in that case the only type of collision between the peptide sidechains spheres within the fibril and those of the peptides in the solution is a hard-sphere collision.

6.5.3 Aβ16-22 solid-phase peptide synthesis (SPPS)

All amino acids and resins were purchased from Novasyn (Merck), Fluorochem or Sigma-

Aldrich. All amino acids were N-Fmoc protected and the side-chains were protected with Boc

(Lys) or OtBu (Glu). Aβ16-22 was synthesized on an automated solid-phase peptide synthesiser

(CEM LibertyBlue). (58) DMF used in peptide synthesis was of ACS grade from Sigma-Aldrich.

Crude peptide identity was confirmed by LC-MS prior to HPLC purification. Aβ16-22 was purified by preparative scale HPLC using an X-bridge C18 preparative column (reversed phase) on an increasing gradient of MeCN to H2O (5 – 95% with 0.1% formic acid) over 15 minutes. The purity

207

of the peptide was confirmed to be above (>95%) by Analytical HPLC and the peptide identity was confirmed by LC-MS. The pure peptide was freeze dried and stored at - 20 °C prior to use.

6.5.4 Aggregation conditions and Transmission Electron Microscopy (TEM)

Aβ16-22 was diluted from a DMSO stock solution (1-30mM) to the required concentration (10-300μM) in 100 mM ammonium bicarbonate buffer (pH=7). Final DMSO concentration was kept at 1% (v/v) in all assays. After a week, aliquots were taken for TEM analysis. If no fibrils were observed, the samples were left to incubate for another week.

Conditions under which Aβ16-22 did not form visible fibrils using extensive TEM analysis after two weeks were counted as those under which fibrils do not form. TEM images were taken at the end of each experiment by removing 5 μL and placing the solution on carbon- formvar grids for 30s prior to staining with 2% (w/v) uranyl acetate solution for an additional 30s as described by Preston et al. (4). Images were taken on a JEM-1400 (JEOL

Ltd., Tokyo, Japan) transmission electron microscope within the Astbury Biostructure

Laboratory.

6.6 Acknowledgements

This work was supported by National Institutes of Health Grants R01 EB006006 and in part by National Science Foundation (NSF) Research Triangle Materials Research Science and

Engineering Centers (MRSEC) Grant DMR-1121107 and by NSF grant CBET 1743432, 1512059 and by EPSRC grants EP/N035267/1, EP/N013573/1 and EP/KO39292/1. SJB gratefully acknowledges BBSRC for a PhD studentship grant BB/J014443/1. The JEM-1400 (Jeol) was purchased with funding from the Wellcome Trust (094232/Z/10/Z) and fitted with a CCD camera

208

funded by (090932/Z/09/Z). We also gratefully acknowledge Dr. Matt Iadanza for help with processing the EM images. YW and CKH gratefully acknowledge the support of a Cheney Visiting

Scholar Fellowship from University of Leeds. The authors declare that they have no conflicts of interest.

209

6.7 References

1. Knowles TPJ, Vendruscolo M, & Dobson CM (2014) The amyloid state and its association

with protein misfolding diseases. Nat. Rev. Mol. Cell Biol. 15(6):384-396.

2. Balbach JJ, et al. (2000) Amyloid fibril formation by A beta(16-22), a seven-residue

fragment of the Alzheimer's beta-amyloid peptide, and structural characterization by solid

state NMR. Biochemistry 39(45):13748-13759.

3. Senguen FT, et al. (2011) Probing aromatic, hydrophobic, and steric effects on the self-

assembly of an amyloid-beta fragment peptide. Mol. Biosyst. 7(2):486-496.

4. Preston GW, Radford SE, Ashcroft AE, & Wilson AJ (2012) Covalent Cross-Linking

within Supramolecular Peptide Structures. Anal. Chem. 84(15):6790-6797.

5. Santini S, Wei GH, Mousseau N, & Derreumaux P (2004) Pathway complexity of

Alzheimer's beta-amyloid A beta(16-22) peptide assembly. Structure 12(7):1245-1255.

6. Cheon M, Chang I, & Hall CK (2011) Spontaneous Formation of Twisted A beta(16-22)

Fibrils in Large-Scale Molecular-Dynamics Simulations. Biophys. J. 101(10):2493-2501.

7. Petty SA & Decatur SM (2005) Experimental Evidence for the Reorganization of β-Strands

within Aggregates of the Aβ(16−22) Peptide. J. Am. Chem. Soc. 127(39):13488-13489.

8. Hsieh M-C, Liang C, Mehta AK, Lynn DG, & Grover MA (2017) Multistep Conformation

Selection in Amyloid Assembly. J. Am. Chem. Soc. 139(47):17007-17010.

9. Irbäck A & Mitternacht S (2008) Spontaneous β-barrel formation: An all-atom Monte Carlo

study of Aβ16–22 oligomerization. Proteins: Struct., Funct., Bioinf. 71(1):207-214.

10. Gnanakaran S, Nussinov R, & García AE (2006) Atomic-Level Description of Amyloid β-

Dimer Formation. J. Am. Chem. Soc. 128(7):2158-2159.

11. Klimov DK & Thirumalai D (2003) Dissecting the Assembly of Aβ16–22 Amyloid

210

Peptides into Antiparallel β Sheets. Structure 11(3):295-307.

12. Nguyen PH, Li MS, & Derreumaux P (2011) Effects of all-atom force fields on amyloid

oligomerization: replica exchange molecular dynamics simulations of the A[small beta]16-

22 dimer and trimer. Phys. Chem. Chem. Phys. 13(20):9778-9788.

13. Favrin G, Irbäck A, & Mohanty S (2004) Oligomerization of Amyloid Aβ16–22 Peptides

Using Hydrogen Bonds and Hydrophobicity Forces. Biophys. J. 87(6):3657-3664.

14. Rohrig UF, Laio A, Tantalo N, Parrinello M, & Petronzio R (2006) Stability and structure

of oligomers of the Alzheimer peptide A beta(16-22): From the dimer to the 32-mer.

Biophys. J. 91(9):3217-3229.

15. Lu Y, Derreumaux P, Guo Z, Mousseau N, & Wei G (2009) Thermodynamics and dynamics

of amyloid peptide oligomerization are sequence dependent. Proteins: Struct., Funct., and

Bioinf. 75(4):954-963.

16. Latshaw DC, Cheon M, & Hall CK (2014) Effects of Macromolecular Crowding on

Amyloid Beta (16-22) Aggregation Using Coarse-Grained Simulations. J. Phys. Chem. B

118(47):13513-13526.

17. Wang Y, Shao Q, & Hall CK (2016) N-terminal Prion Protein Peptides (PrP(120-144))

Form Parallel In-register β-Sheets via Multiple Nucleation-dependent Pathways. J. Biol.

Chem. 291(42):22093-22105; correction (2017) 292 (50): 20655.

18. Aggeli A, et al. (2001) Hierarchical self-assembly of chiral rod-like molecules as a model

for peptide beta-sheet tapes, ribbons, fibrils, and fibers. Proc. Natl. Acad. Sci. U. S. A.

98(21):11857-11862.

19. Hamley IW, et al. (2010) Influence of the Solvent on the Self-Assembly of a Modified

Amyloid Beta Peptide Fragment. II. NMR and Computer Simulation Investigation. J. Phys.

211

Chem. B 114(2):940-951.

20. Annunziata O, et al. (2005) Oligomerization and Phase Transitions in Aqueous Solutions

of Native and Truncated Human βB1-Crystallin. Biochemistry 44(4):1316-1328.

21. Hellstrand E, Boland B, Walsh DM, & Linse S (2010) Amyloid beta-Protein Aggregation

Produces Highly Reproducible Kinetic Data and Occurs by a Two-Phase Process. ACS

Chem. Neurosci. 1(1):13-18.

22. Mason TO, et al. (2017) Thermodynamics of Polypeptide Supramolecular Assembly in the

Short-Chain Limit. J. Am. Chem. Soc. 139(45):16134-16142.

23. Ricchiuto P, Brukhno AV, & Auer S (2012) Protein Aggregation: Kinetics versus

Thermodynamics. J. Phys. Chem. B 116(18):5384-5390.

24. Chen S & Wetzel R (2001) Solubilization and disaggregation of polyglutamine peptides.

Protein Sci. 10(4):887-891.

25. O'Nuallain B, Shivaprasad S, Kheterpal I, & Wetzel R (2005) Thermodynamics of Aβ(1-

40) Amyloid Fibril Elongation. Biochemistry 44(38):12709-12718.

26. Dima RI & Thirumalai D (2002) Exploring protein aggregation and self-propagation using

lattice models: Phase diagram and kinetics. Protein Sci. 11(5):1036-1049.

27. Nguyen HD & Hall CK (2004) Phase diagrams describing fibrillization by polyalanine

peptides. Biophys. J. 87(6):4122-4134.

28. Auer S & Kashchiev D (2010) Phase Diagram of alpha-Helical and beta-Sheet Forming

Peptides. Phys. Rev. Lett. 104(16):168105.

29. Auer S (2011) Phase diagram of polypeptide chains. J. Chem. Phys. 135(17):175103.

30. Bai XM & Li M (2006) Calculation of solid-liquid interfacial free energy: A classical

nucleation theory based approach. J. Chem. Phys. 124(12):124707.

212

31. Morris JR & Song XY (2003) The anisotropic free energy of the Lennard-Jones crystal-

melt interface. J. Chem. Phys. 119(7):3920-3925.

32. Broughton JQ & Gilmer GH (1986) Molecular dynamics investigation of the crystal–fluid

interface. VI. Excess surface free energies of crystal-liquid systems. J. Chem. Phys.

84(10):5759-5768.

33. Haas C & Drenth J (1995) The interaction energy between two protein molecules related

to physical properties of their solution and their crystals and implications for crystal growth.

J. Cryst. Growth 154(1):126-135.

34. Kashchiev D (2000) Nucleation: Basic Theory with Applications (Elsevier, Butterworth-

Heinemann,Oxford).

35. Kashchiev D & Auer S (2010) Nucleation of amyloid fibrils. J. Chem. Phys.

132(21):215101.

36. Garai K, Sahoo B, Sengupta P, & Maiti S (2008) Quasihomogeneous nucleation of amyloid

beta yields numerical bounds for the critical radius, the surface tension, and the free energy

barrier for nucleus formation. J. Chem. Phys. 128(4):045102.

37. Gosavi RA, Bhamidi V, Varanasi S, & Schall CA (2009) Beneficial Effect of Solubility

Enhancers on Protein Crystal Nucleation and Growth. Langmuir 25(8):4579-4587.

38. Bromberg L, Rashba-Step J, & Scott T (2005) Insulin Particle Formation in Supersaturated

Aqueous Solutions of Poly(Ethylene Glycol). Biophys. J. 89(5):3424-3433.

39. Gremer L, et al. (2017) Fibril structure of amyloid-beta(1-42) by cryo-electron microscopy.

Science 358(6359):116-119.

40. Burkoth TS, et al. (2000) Structure of the beta-amyloid((10-35)) fibril. J. Am. Chem. Soc.

122(33):7883-7889.

213

41. Kirschner DA, et al. (1987) Synthetic peptide homologous to beta protein from Alzheimer

disease forms amyloid-like fibrils in vitro. Proc. Natl. Acad. Sci. U. S. A. 84(19):6953-6957.

42. Lu K, Jacob J, Thiyagarajan P, Conticello VP, & Lynn DG (2003) Exploiting amyloid fibril

lamination for nanotube self-assembly. J. Am. Chem. Soc. 125(21):6391-6393.

43. Childers WS, Anthony NR, Mehta AK, Berland KM, & Lynn DG (2012) Phase Networks

of Cross-beta Peptide Assemblies. Langmuir 28(15):6386-6395.

44. Keith JA & Henry PM (2009) The Mechanism of the Wacker Reaction: A Tale of Two

Hydroxypalladations. Angew. Chem. Intl. Ed. 48(48):9038-9049.

45. Tao K, et al. (2011) Self-Assembly of Short A beta(16-22) Peptides: Effect of Terminal

Capping and the Role of Electrostatic Interaction. Langmuir 27(6):2723-2730.

46. Zhang SG (2003) Fabrication of novel biomaterials through molecular self-assembly. Nat.

Biotechnol. 21(10):1171-1178.

47. Krieg E, Bastings MMC, Besenius P, & Rybtchinski B (2016) Supramolecular Polymers

in Aqueous Media. Chem. Rev. 116(4):2414-2477.

48. Roberts CJ (2014) Therapeutic protein aggregation: mechanisms, design, and control.

Trends Biotechnol. 32(7):372-380.

49. Rosa M, Roberts CJ, & Rodrigues MA (2017) Connecting high-temperature and low-

temperature protein stability and aggregation. PLoS One 12(5):e0176748.

50. Nguyen HD & Hall CK (2004) Molecular dynamics simulations of spontaneous fibril

formation by random-coil peptides. Proc. Natl. Acad. Sci. U. S. A. 101(46):16180-16185.

51. Cheon M, Chang I, & Hall CK (2010) Extending the PRIME model for protein aggregation

to all 20 amino acids. Proteins: Struct. Funct. Bioinf. 78(14):2950-2960.

52. Muñoz V & Serrano L (1995) Elucidating the Folding Problem of Helical Peptides using

214

Empirical Parameters. III. Temperature and pH Dependence. J. Mol. Biol. 245(3):297-308.

53. de la Paz ML & Serrano L (2004) Sequence determinants of amyloid fibril formation. Proc.

Natl. Acad. Sci. U. S. A. 101(1):87-92.

54. Wagoner VA, Cheon M, Chang I, & Hall CK (2012) Fibrillization Propensity for Short

Designed Hexapeptides Predicted by Computer Simulation. J. Mol. Biol. 416(4):598-609.

55. Cheon M, Chang I, & Hall CK (2012) Influence of temperature on formation of perfect tau

fragment fibrils using PRIME20/DMD simulations. Protein Sci. 21(10):1514-1527.

56. Wang Y, Latshaw DC, & Hall CK (2017) Aggregation of Aβ(17-36) in the Presence of

Naturally Occurring Phenolic Inhibitors Using Coarse-Grained Simulations. J. Mol. Biol.

429(24):3893-3908.

57. Cheon M, Hall CK, & Chang I (2015) Structural Conversion of A beta(17-42) Peptides

from Disordered Oligomers to U-Shape Protofilaments via Multiple Kinetic Pathways.

PLoS Comput. Biol. 11(5):e1004258.

58. Collins JM, Porter KA, Singh SK, & Vanier GS (2014) High-Efficiency Solid Phase

Peptide Synthesis (HE-SPPS). Org. Lett. 16(3):940-943.

215

Figure 6.1. (A) Representation of the Aβ16-22 peptide in PRIME20 model. (B) and (C) are schematic representations of the square well interaction between two sidechains amd directional square well interaction (hydrogen bonding) between backbone NH and C=O beads. (The spheres are not drawn to scale for ease of viewing.) D-H are simulation snapshots of non-HB oligomer, HB oligomer, 2, 3 and 4 β-sheet fibril, respectively, generated by visual molecular dynamics software (VMD). The peptides in the five aggregates and in solution are shown in green and magenta, respectively.

216

Figure 6.2. The probability distribution for (A) the average number of hydrogen bonds (NHB) per peptide (counting both intra and inter-molecular HBs) and (B) the average number of hydrogen bonds per residue for the five types of aggregates. The probability distribution (C) for the number of HP sidechain contacts (NHP) per peptide and (D) the average number of sidechain contacts per residue for the five aggregates. The data for the non-HB oligomer (orange) and the HB oligomer (purple) are measured at T=193K and 250K, respectively, and the data for the 2β- sheet (red), 3β-sheet (green) and 4β-sheet (blue) are measured at T=330K. These five aggregates are characterized at different temperatures but similar solubilities, as shown in Fig. 3A.

217

Figure 6.3. (A) a plot of the solubility 퐶푒 for non-HB oligomer, HB oligomer, 2β, 3β, 4β-sheet fibrils as a function of the temperature; (B) Solubility data for non-HB oligomer, HB oligomer, 2β, 3β, 4β-sheet fibrils at four to six different temperatures; lines are fitted to Eq. (1); the color scheme is the same as plot A; (C) Comparison of the average binding energy per peptide within the aggregate and the corresponding latent heat of peptide aggregation into oligomer and fibril.; (D) Dependence of the fibril solubility on thickness (i=2, 3 and 4) for 2, 3 and 4 β-sheet fibrils, respectively; lines are fitted to Eq. (2).

218

Figure 6.4. (A) Phase diagram for Aβ16-22 peptide. Solubility (Ce)-vs-temperature (T) data are presented for non-HB oligomer (orange dashed line), HB oligomer (purple dashed line), 2β-sheet fibril (red dashed line), 3β-sheet fibril (green dashed line), 4β-sheet fibril (blue dashed line) and the infinite layer (∞) β-sheet fibril (black line). The fibril and solution phases are colored yellow and cyan, respectively. (B) DMD/PRIME20 simulation snapshots (reproduced from ref 6) at the concentration and temperature of phase point A and B are taken at different time points, respectively.

219

Figure 6.5. (A) Summary of both the simulation-predicted temperature-dependent solubility line 퐶푒(T) for Aβ16-22 peptide (black curve) and the fibrillation experiments performed under given conditions. Red dots and red circles indicate conditions at which fibrils have been found to form, and not to form, respectively, via TEM (at temperatures 277K, 296K, 310K and 330K). Blue dots indicate that fibrils have been reported in the literature to form under these conditions (see Figure S2). (B) Six selected transmission electron micrographs (TEM) images (i-vi) showing that Aβ16-22 form fibrils under the conditions that correspond to the six red dots labeled i-vi in (A). Buffer: 100 mM ammonium bicarbonate, pH=7, with a final concentration of 1% DMSO (v/v). Scale bar: 200 nm.

220

6.7 Supplementary Information

34

32

30 -2 28

/ mJ m mJ / 26

h σ 24

22

20 320 330 340 350 360 370 T / K 2 Figure S6.1. Specific surface energy σh (mJ/m ) of Aβ16-22 fibril versus temperature calculated from fitting simulation data (푙푛퐶𝑖훽 vs 1/i) to Eq. (2).

221

Figure S6.2. A plot of the estimated solubility of Aβ16-22 fibril versus fibril thickness (number of β-sheets) at various given temperatures. Data for this plot are extracted from the linear fit of 푙푛퐶𝑖훽 - 1/푖 (i=2,3,4) data to Eq. (2) in Fig. 3D.

222

Figure S6.3. Illustration of the Helmholtz free energy (A) versus peptide composition (x) for a binary (peptide-water) solution system at constant volume (V) and temperature (T). The Helmholtz free energy is considered here instead of the Gibbs free energy because the simulations are done at constant volume rather than constant pressure.

223

Figure S6.4. Snaphot from a simulation showing the structure of a two-layer β-sheet fibril of Aβ16-22 peptide (colored green and yellow) in an anti-parallel β-sheet configuration. The oxygen (O), nitrogen (NH) and carbon (Cα, C=O) groups on the peptide backbone are shown in red, blue and cyan. Note that the sidechain groups are not represented here. The O groups (red) are incorporated into the C=O sphere and do not exist independently in PRIME20 model, the positions of which are determined after simulation.

224

Figure S6.5. Plots of (A) total number of peptides and (B) the average peptide binding energy (kJ/mol) of a 4β-sheet fibril versus simulation time at T=330K.

225

Figure S6.6. Literature conditions (check marks) under which Aβ16-22 has been shown experimentally to form fibrillar structures (i.e. at neutral pH). The references used to make this table can be found in the reference list. N.D. means that fibrils are presumed to form but there is no literature study performed under this condition. ? means no experiments have been conducted under these conditions.

226

References

1. Hsieh MC, Liang C, Mehta AK, Lynn DG, Grover MA (2017) Multistep

Conformation Selection in Amyloid Assembly. J Am Chem Soc 139: 17007-17010.

2. Childers, WS, Anthony, NR, Mehta, AK, Berland, KM, Lynn, DG (2012) Phase

Networks of Cross-β Peptide Assemblies. Langmuir 28:6386-6395.

3. Petty SA, Decatur SM (2005) Experimental evidence for the reorganization of β-

strands within aggregates of the Aβ(16-22) peptide. J Am Chem Soc 127:13488-

13489.

4. Preston GW (2012) Analysis of biomolecular interactions using photo-induced

covalent cross-linking. PhD thesis, University of Leeds.

5. Balbach JJ, Ishii Y, Antzutkin ON, Leapman RD, Rizzo NW, Dyda F, Reed J, Tycko

R (2000) Amyloid fibril formation by A beta(16-22), a seven-residue fragment of the

Alzheimer's beta-amyloid peptide, and structural characterization by solid state NMR.

Biochemistry 39:13748-13759.

6. Senguen FT, Lee NR, Gu X, Ryan DM, Doran TM, Anderson EA, Nilsson BL (2011)

Probing aromatic, hydrophobic, and steric effects on the self-assembly of an amyloid-

β fragment peptide. Mol BioSyst 7:486-496.

227

CHAPTER 7

Molecular Insights into the surface catalyzed secondary nucleation of Amyloid-β40 (Aβ40)

by the peptide fragment Aβ16-22

Chapter 7 is essentially a manuscript by Samuel Bunce, Yiming Wang, Katie L. Stewart, Alison

E. Ashcroft, Sheena E. Radford, Carol K. Hall and Andrew J. Wilson submitted to Science

Advances.

228

Molecular Insights into the surface catalyzed secondary nucleation of Amyloid-β40 (Aβ40) by the peptide fragment Aβ16-22

Samuel J. Bunce, [1,2]† Yiming Wang, [3]† Katie L. Stewart, [2,4] Alison E. Ashcroft, [2,4] Sheena E. Radford,[2,4]* Carol K. Hall,[3]* and Andrew J. Wilson[1,2]*

1. School of Chemistry, University of Leeds, Leeds LS2 9JT, United Kingdom.

2. Astbury Centre for Structural Molecular Biology, University of Leeds, Woodhouse Lane, Leeds LS2 9JT, United Kingdom.

3. Department of Chemical and Biomolecular Engineering, North Carolina State University, Raleigh, NC 27695-7905, United States of America.

4. School of Molecular and Cellular Biology, Faculty of Biological Sciences, University of Leeds, Woodhouse Lane, Leeds LS2 9JT, United Kingdom.

Abstract

Understanding the structural mechanism by which proteins and peptides aggregate is crucial given the role of fibrillar aggregates in debilitating amyloid diseases and bioinspired materials, yet, this is a major challenge as assembly involves multiple heterogeneous and transient intermediates. Here, we analyze the co-aggregation of Aβ40 and Aβ16-22, two widely studied peptide fragments of Aβ42 implicated in Alzheimer’s disease. We demonstrate that Aβ16-22 increases the aggregation rate of Aβ40 through a surface catalyzed secondary nucleation mechanism.

Discontinuous molecular dynamics simulations allowed aggregation to be tracked from the initial random-coil monomer to the catalysis of monomer nucleation on the fibril surface. Together, the results provide insight into how dynamic interactions between Aβ40 monomers/oligomers on the surface of pre-formed Aβ16-22 fibrils nucleate Aβ40 amyloid assembly. This new understanding may facilitate development of surfaces designed to enhance or suppress secondary nucleation and hence to control the rates and products of fibril assembly.

229

7.1 Introduction

Understanding the molecular mechanisms of peptide self-assembly into amyloid fibrils is of key importance in understanding pathological disease states,(1) as well as in designing new functional materials.(2) Aberrant self-assembly of monomeric peptides or proteins into amyloid fibrils is associated with a number of degenerative conditions, notably Alzheimer’s and

Parkinson’s disease,(1, 3) in which considerable evidence now implicates soluble oligomers as the primary cause of cellular damage.(4, 5) Identifying and characterizing the structural changes that occur during peptide assembly into amyloid is essential in the quest to develop strategies to combat disease and manufacture bespoke materials (1, 6)

Peptide assembly into amyloid fibrils occurs via a complex nucleation-dependent mechanism in which subtle changes in lowly populated states can have dramatic effects on the rates and products of assembly.(7) Elegant work has resulted in kinetic models that are able to dissect the different contributing steps in assembly, including primary nucleation, elongation, fragmentation, and secondary nucleation.(8-10) However, elucidating structural insights into these different steps in assembly, including the nature of early oligomeric species is challenging, as CD,

IR and other spectroscopic techniques generally only observe population-average data for a whole system. Single molecule FRET and solid state NMR which have uncovered clues as to the structure of toxic versus non-toxic oligomeric species,(11, 12) provide information on the average properties of the different species at different times. Native ion mobility spectrometry-mass spectrometry

(IMS-MS) separates ions based on shape as well mass and charge,(13) and has been used to provide insights into the population, conformation and ligand-binding capability of individual peptide monomers and oligomers.(14) By using photo-induced cross-linking (PIC), fleeting inter/intra-peptide interactions may be trapped through covalent bond formation (to encode

230

supramolecular connectivity).(15) Molecular Dynamics (MD) simulations focusing on multi- peptide systems at short time scales (<1 ms)(16) can help fill the gaps between population- average data and individual structures. Such simulations can provide insights into self-assembly events in molecular detail, allowing the earliest stages of aggregation to be visualized and the course of aggregation to be tracked in all-atom detail.(17-19)

The amyloid-β peptide (Aβ) is a major component of the extracellular plaques observed in

Alzheimer’s disease.(20) Aggregation of Aβ40/42 (Fig. 1a) into amyloid fibrils has been widely studied both in vitro and in vivo,(21) although numerous questions remain about its structure and role in Alzheimer’s disease progression.(1, 21) Kinetic analysis of the sigmoid growth curves of

Aβ40/42 aggregation has enabled their assembly mechanisms to be deconvoluted into a number of microscopic steps.(7, 9) Assembly begins with a lag phase, during which time monomers and small amounts of oligomers persist.(7) Monomers then undergo a rearrangement step to form a nucleus

(primary nucleation) from which fibrils can grow. Further aggregate growth occurs through pathways that include elongation (whereby a monomeric peptide adds onto the end of a growing fibril), fragmentation (fibrils break into two smaller aggregates, exponentially increasing growth- competent fibril ends), and surface catalyzed secondary nucleation (whereby nucleation is catalyzed on the fibril surface).(22) Using MD simulations, the energy landscape of Aβ40 oligomer formation has also been modelled, demonstrating the different kinetic pathways that underlie the formation of pre-fibrillar and non-fibrillar oligomers.(16, 17) For WT Aβ40, primary nucleation has been shown to be a slower process than secondary pathways, such that surface-catalyzed secondary nucleation events dominate the growth rate of fibrils.(9) Under quiescent conditions, the contribution of fibril fragmentation to the growth of fibrils has been shown to be negligible.(8)

Co-aggregation processes (i.e. where two different peptide sequences interact during aggregation

231

but need not co-assemble) can result in more complex kinetics, due to the possibility of the sequences interacting with each other to modulate aggregation.(23, 24) Such a situation may occur in vivo wherein multiple sequences of different length of A are formed.(25)

Here, we combine fluorescence assays, IMS-MS measurements, and PIC experiments to determine the kinetic mechanism of co-assembly of the peptide fragment Aβ16-22, which contains the “core recognition motif” KLVFF(26) of Aβ40, with the parent Aβ40 sequence. Aβ16-22 has been shown to form fibrils with an in-register, antiparallel orientation at neutral pH,(27) and has been proposed to assemble via an intermediate with out-of-register β-sheet alignment prior to reaching the final in-register fibril structure.(28) The rate of aggregation is expected to be highly dependent on peptide concentration and ionic strength.(29-31) Discontinuous molecular dynamics (DMD), have also shown that the nucleation-dependent aggregation process of Aβ16-22 proceeds from a random coil configuration to form multi-layer β-sheet fibrils, with an in-register antiparallel β- sheet orientation, in accordance with the experimental data.(32) Here we show, using fluorescence quenching assays, that Aβ16-22 aggregates more rapidly than Aβ40, and that Aβ16-22 fibril formation then increases the aggregation rate of Aβ40 through a surface catalyzed, secondary nucleation mechanism. Using discontinuous molecular dynamics (DMD) simulations we also show that the preformed Aβ16-22 fibrils increase the early-stage aggregation rate of Aβ40, but that the monomeric

Aβ16-22 peptides do not, supporting secondary nucleation as the mechanism of enhanced Aβ40 aggregation by Aβ16-22. The results provide the first experimentally validated simulations portraying the structural mechanism of surface catalyzed nucleation. This new understanding may pave the way to the generation of surfaces able to enhance or suppress assembly, dependent on the desired outcome of the assembly reaction, inform effective design of ligands that modulate therapeutically important amyloid assembly. Finally, the study exemplifies an approach that may

232

be applied to the combined experimental and in silico analyses of hetreogenous peptide aggregation process with greater in vivo relevance.

7.2 Results

7.2.1 Aβ16-22 increases the aggregation rate of Aβ40

To determine whether the presence of Aβ16-22 affects the aggregation rate of Aβ40, the peptides were synthesized (see experimental methods, supplementary materials and Figs S1-S2), purified and mixed in different ratios at a constant total peptide concentration of 40 μM. The rate of aggregation was then measured using the fluorescence of Thioflavin-T (ThT) (Fig. 1b, experimental methods and Fig. S3). Initial experiments showed the expected sigmoid increase in

ThT fluorescence for Aβ40,(33) indicating the assembly of this peptide into amyloid fibrils (Fig.

1b, associated EM images can be seen in Fig. 5a). Interestingly, while Aβ16-22 formed fibrils under the conditions employed (Fig. 2f), as expected,(15) ThT fluorescence did not increase (Fig. 1b), indicating that the fibrils formed are either unable to bind ThT or do not enhance its fluorescence.

Other amyloid dyes were screened against Aβ16-22 fibrils; however, none produced a large enough signal with which to perform kinetic assays (data not shown). The increase in ThT signal in the peptide mixture thus reports on the aggregation rate of Aβ40 and whether/how this is affected by the presence of Aβ16-22. Interestingly, the experiments in Fig. 1b show that, as the molar ratio of

Aβ16-22 to Aβ40 is increased, the apparent aggregation rate of Aβ40 also increases, despite its concentration decreasing (see also Fig. S3), with the maximal apparent rate enhancement seen at a 1:1 ratio of the two peptides (Fig. 1b).

In order to characterize the extent to which Aβ40 aggregation is accelerated by the presence of Aβ16-22, the half-time (t1/2, the time at which the growth curve reaches 50% amplitude) was

233

calculated for each peptide mixture and normalized to the half time for the equivalent concentration of Aβ40 alone (Figs. S3). The results revealed a dramatic, and titratable, effect of the presence of

Aβ16-22 on the aggregation rate of Aβ40, demonstrating an interaction between the two peptides that accelerates the rate of assembly.

7.2.2 Aβ16-22 aggregates more rapidly than Aβ40 and is unaffected by the presence of Aβ40

As the assembly kinetics of Aβ16-22 could not be measured using any of the amyloid dyes surveyed at the concentrations employed here, a fluorescence quenching assay was developed to determine whether Aβ16-22 aggregates more or less rapidly that Aβ40 (Fig. 2a). Similar assays have been used previously to monitor the aggregation rates of Aβ40 and Aβ42,(34) with fluorescence quenching reporting on labelled monomers coming into close proximity as oligomers (or fibrils) form. For these assays Aβ16-22 N-terminally labelled with tetramethylrhodamine (TAMRA) was synthesized, including a 6-aminohexanoic acid linker (Ahx) to limit disruption to the native fibril structure that might arise due to the bulky fluorophore (TAMRA-Ahx-Aβ16-22) (supplementary materials and Fig. S1 and S4). When incubated in isolation, a 5% (w/w) TAMRA-Ahx-Aβ16-22:

95% Aβ16-22 mixture resulted in a rapid decrease in fluorescence intensity followed by a slower phase that plateaued after 1 h (Fig. 2b). In the presence of Aβ40 (1:1 (mol/mol) ratio, 40 μM total peptide concentration, and 2% (v/v) DMSO), no difference in the rate of fluorescence decrease was observed, indicating that the presence of Aβ40 has no effect on Aβ16-22 aggregation (Fig. 2c).

Analysis of these samples by negative stain TEM showed the presence of fibrils after only 5 mins

(Fig. 2f). Sedimentation of the mixed system by centrifugation after 1 h demonstrated that Aβ40 was present mainly in the supernatant (Figs. 2d, e). These results demonstrate that Aβ16-22 aggregates rapidly to form amyloid-like fibrils while Aβ40 remains soluble as

234

monomers/oligomers. Thus, although the rate of Aβ40 aggregation is increased by the presence of

Aβ16-22, limited or no co-assembly between the two peptides was observed. By contrast, Aβ16-22 aggregation is unaffected by the presence of Aβ40.

7.2.3 Monomeric Aβ16-22 can interact with monomeric and oligomeric Aβ40 through the self- recognition motif KLVFF

To determine whether Aβ16-22 and Aβ40 interact transiently in the early stages of assembly, native electrospray ionization (ESI) linked to ion mobility mass spectrometry (IMS-MS) was performed (see Experimental Methods). This soft ionization technique has been used to identify and structurally characterize amyloid oligomers formed from several different proteins and peptides.(13, 14) Under the conditions used here, ESI-IMS-MS revealed Aβ40 co-populates a number of oligomers, ranging from monomers to pentamers (Fig. 3a (white), see also Table S1), consistent with previous results. (33) When incubated with Aβ16-22, heteromolecular oligomers were observed (Fig. 3b (green)), along with homomolecular oligomers of Aβ40 (Fig. 3a, (white)).

Notably Aβ16-22 oligomers were not observed. The heteromolecular oligomers correspond to multiple Aβ16-22 monomers bound to either an Aβ40 monomer or dimer (Table S1). Collision cross- section (CCS) estimations from the ESI-IMS-MS analysis of the Aβ40 species in the presence or absence of Aβ16-22 indicate no significant difference in the gas phase cross-section of Aβ40, implying that a conformational change in monomer or oligomer structure is unlikely to be the provenance for the Aβ16-22 driven increase in Aβ40 aggregation rate (Fig. S5). Despite attempts to capture the interaction experimentally by PIC using a diazirine labelled Aβ16-22 (Aβ*16-22, see supplementary materials for synthesis, Scheme S1 and Fig. S1), the site of interaction could not be verified (Fig S6 and Table S2), likely due to the low percentage of any heterodimers present (as

235

assessed by total ion count, 1.0 ± 0.5%) and the lower solution concentration of Aβ16-22 arising as a consequence of its rapid aggregation.

To assess further the nature of the interactions between Aβ16-22 and Aβ40, discontinuous molecular dynamics (DMD) simulations were performed (see Experimental Methods) of Aβ40 alone (Fig. 1c) or a 1:1 mixture of the Aβ16-22 and Aβ40 peptide sequences at Cpeptide = 5mM (Fig.

3c). Simulations performed on six monomers of Aβ40 (Fig. 1c) showed that the initially unstructured peptides successfully assemble and adopt a metastable oligomer structure by 104 μs

(Fig. 1c); this structure comprises antiparallel intermolecular β-strands linked by disordered regions with β-strands stacked perpendicular to the long axis. During this oligomerization stage, the peptide conformation is similar to that observed by Zhang et al. (16) (35) As the simulation proceeds this oligomer loses some β-sheet content (t = 230 μs, Fig. 1c). By the end of the simulation (621 μs), peptides in oligomers undergo structural rearrangement from anti-parallel β- strand conformations to the parallel β-sheet conformation observed for Aβ40 fibrils (Fig. 1c).(35)

Interestingly, simulations of the peptide mixtures did not show an accelerating effect of Aβ16-22 on the aggregation rate of Aβ40 (see Fig. 3c and later). However, interactions between the two peptides were observed, consistent with the ESI-MS results in Fig. 3. From the DMD data, an energy contact map between the monomeric Aβ16-22 and Aβ40 peptides was calculated (Fig. 3d). The contact map indicated that Aβ16-22, residues 18-20 (VFF) interact strongly with residues 19-21 and 32-35 of

Aβ40 (FFA and IGLM, respectively), consistent with experimental data previously reported in the literature, which indicates KLVFF is a “self-recognition element”.(26) Such an interaction, however, does not appear to result in an acceleration of aggregation (Fig. 3c) suggesting that they generate transient, off-pathway species (Fig. 3e).

236

7.2.4 Aβ16-22 fibrils have a larger effect on the aggregation rate of Aβ40 than Aβ16-22 monomer

To determine whether rapidly formed Aβ16-22 fibrils are the causative agents of the enhanced rate of Aβ40 aggregation in the mixed samples (Fig. 1b), the effect of pre-formed Aβ16-

22 fibrils on Aβ40 aggregation was assessed. These experiments (Fig. 4a) showed that the presence of Aβ16-22 fibrils increases the rate of aggregation of Aβ40 in a fibril concentration-dependent manner (Fig. 4a) and this had a larger effect on aggregation rate when compared to the effect observed upon addition of monomeric (i.e. taken straight from a DMSO stock) Aβ16-22 (Fig. 4b).

This suggests that aggregation is enhanced either by cross-seeding (i.e. by the peptide adding directly to the ends of Aβ16-22 fibrils) or by secondary nucleation of Aβ40 on the Aβ16-22 fibril surface (Fig. 4e). Comparison of the effects of unsonicated fibrils (fewer ends) with the same fibrils fragmented by sonication (Figs. 4c) indicated that the former was not the case (Fig. 4c) since the average t1/2 for sonicated fibrils (6.2 ± 1.0 h) is similar to that of its unfragmented counterpart (7.2

± 0.7 h). Together, the results thus demonstrate that the presence of rapidly formed Aβ16-22 fibrils enhances aggregation of Aβ40 in peptide mixtures by secondary nucleation, despite the presence of small amounts of mixed oligomers (as demonstrated by the MS experiments).

DMD simulations of the aggregation of six Aβ40 peptides were also performed in the presence of pre-formed Aβ16-22 fibrils of different sizes (two, three and four β-sheets) at CAβ40

=1mM to model the dynamic process of the secondary nucleation event. The results showed that the largest Aβ16-22 fibril (i.e. four β-sheets) had the largest impact on the rate of β-sheet formation formed by Aβ40 (Fig. 4d). These simulations are thus qualitatively concordant with the experimental findings that the fibrillar structure of Aβ16-22 is the dominant influence on the aggregation rate of Aβ40.

237

7.2.5 Aβ40 and Aβ16-22 form distinct homomolecular fibrils

The peptide composition of the final fibril structure(s) represents a further means to discern the difference between surface catalyzed secondary nucleation and co-assembly exploiting fibril ends. A surface catalyzed mechanism would most likely produce homo-molecular fibrils, as once they have formed on the Aβ16-22 fibril surface, the Aβ40 nuclei would dissociate and form pure Aβ40 fibrils. In contrast, co-assembly involving fibril ends should result in mixed fibrils, in which Aβ16-

22 seeds are segmentally separated from fibril regions containing Aβ40 monomers.

Negative stain TEM images taken at the end of the aggregation reaction showed Aβ40 fibrils with similar gross morphology when incubated in isolation or co-aggregated with Aβ16-22 (Fig.

5a,b). Similarly, quantitation of ThT fluorescence at the end-point of aggregation in mixed samples and of the same concentration of Aβ40 incubated alone were indistinguishable (Fig. S3), supporting the hypothesis that homomolecular Aβ40 fibrils are formed at the end of the assembly reaction.

Finally, PIC was used to explore whether homo- or heteromolecular fibrils had formed

(Experimental Methods, (Fig. 5c). To perform PIC experiments, a diazirine label was placed on

F20 of Aβ16-22 (Aβ*16-22).(15) Control experiments demonstrated that Aβ*16-22 has a similar effect on the rate of Aβ40 aggregation as its unmodified counterpart (Fig. S3). PIC experiments performed

5 mins and 24 hours after initiating assembly failed to detect crosslinks between Aβ*16-22 and Aβ40

(Figs. 5c, S6, and Table S2). Instead, all identifiable cross-links were consistent with inter/intramolecular Aβ*16-22 or solvent adducts, as previously identified in Aβ*16-22-containing fibrils by Preston and co-workers(15), indicating that co-assembly into fibrils does not occur.

To provide a molecular image of co-assembly, DMD simulations were performed in which six Aβ40 and six Aβ16-22 monomers were mixed and their aggregation behavior was monitored versus time at Cpeptide = 5mM (Fig. 5d). The simulations showed that in the early stages of assembly

238

(t = 0.6 μs) a mixture of monomeric and oligomeric Aβ40 was present. As the simulation progressed

(t = 57 μs), all Aβ40 peptides coalesced into one β-sheet rich oligomer, with Aβ16-22 intercalated within the structure. Throughout the simulation, monomeric Aβ16-22 was observed to bind transiently to other monomeric Aβ16-22 peptides or the KLVFF motif of Aβ40, in accordance with the experimental data presented above. Finally, at the end of the simulation (t = 202 μs) the peptides tended to form distinct oligomeric domains with Aβ40 and Aβ16-22 forming separate sheets.

7.2.6 Aβ40 oligomer dynamics on the surface of Aβ16-22 fibrils

To obtain a molecular image of the process of secondary nucleation, DMD simulations were performed in which six Aβ40 monomers were mixed with preformed fibrils of a Aβ16-22 at

CAβ40 = 5mM (Fig. 6a). At the early stage of the simulation (t = 0.29 μs), three Aβ40 peptides were present in an oligomer, one other Aβ40 peptide was associated at the end of the fibril and the remaining two Aβ40 peptides were elongated across the fibril surface. At this stage (t = 0.29 μs), the Aβ40 peptides in the oligomer and on the surface were observed to adopt predominantly a random coil conformation with small amounts of β-strand structure (note that an elongated monomeric structure was also observed in simulations performed by Barz et al.(18) in exploring the secondary nucleation of Aβ42 on the surface of Aβ11-42) . The β-sheets were next observed to act as templates for peptides present in a random coil conformation (1.93 μs) and to pull them more fully to the fibril surface, thus, as the simulation progressed, the Aβ40 peptides remaining in solution were recruited by those on the fibril surface. Once the oligomer became fully associated with the fibril surface, the amount of β-sheet structure in the surface-associated oligomer increased

(t = 7.7 μs); antiparallel β-strands formed via inter and intramolecular hydrogen-bonding leading to sheet formation consistent with the early stages observed in the simulations performed for Aβ40

239

alone (Fig. 1c, t = 104 μs). Finally the surface-associated Aβ40 peptides were joined in an ordered oligomer (t = 28.99 and 77.0 μs). Related “bind and re-organize” processes for secondary nucleation were observed in simulations performed by Schwierz et al. using Aβ9-40 as a model.(36)

As noted above, Aβ40 peptides attached to both the lateral surface and to the end of the Aβ16-22 fibril during the simulation, with the Aβ40 C-terminal region attaching more frequently to the lateral surface of the fibril than to the fibril ends (Fig. S7). Consistent with these observations the simulation was repeated three times; two of the three independent runs gave results similar to those described above, whilst for the final run, a greater number of associations to the fibril end were observed. Also, in our simulation, the higher the Aβ40 concentration is , the higher the probability that the Aβ40 docks onto the Aβ16-22 fibril lateral surface. Collectively, these results provide molecular images of surface catalyzed nucleation in which a random coil peptide is catalytically converted into a β-sheet fibrillar structure on a fibril surface.

7.3 Discussion

In this work we used ESI-MS, PIC and DMD to study the co-assembly mechanism of Aβ16-

22 and Aβ40 into amyloid. We show that mixed Aβ16-22/ Aβ40 heteromeric oligomers form but that these are transient, lowly populated and do not significantly affect the rate of aggregation. In contrast Aβ16-22 has a high propensity to self-associate into homomolecular fibrils and these fibrils accelerate Aβ40 assembly by monomer/oligomer interactions through secondary nucleation at the fibril surface. Recent kinetic modelling of amyloid assembly kinetics has revealed the importance of primary nucleation, secondary nucleation and fibril elongation in fibril growth mechanisms.

Our DMD simulations illustrate that whilst all three processes occur simultaneously, secondary nucleation is the dominant process in Aβ40 fibril formation kinetics during co-assembly with Aβ16-

240

22 which is consistent with the findings for the assembly mechanism of Aβ40 observed previously.(8, 23, 37) Moreover, Aβ40 assembly intermediates on the surface of Aβ16-22 fibrils resemble those formed spontaneously in solution for Aβ40 alone, implying that the fibril surface catalyzes the assembly reaction without modifying the molecular mechanism, at least for the simulations performed here. Whether or not this holds for other sequences and co-assembly reactions will require further exploration. This demonstrates the power of using integrated approaches to study the kinetic and structural determinants of molecular assembly processes.

The current study thus serves to emphasize the dramatic differences in aggregation behavior that are observed during co-aggregation compared to self-assembly cases and underscores the need to employ multiple methods to understand aggregation mechanisms in molecular detail. Significant current interest centers on characterizing distinct molecular steps leading to amyloid fibril formation, with secondary nucleation considered as playing a key role in causing toxicity. Recently, kinetic analyses have been augmented by mapping the free-energy landscapes defining different microscopic phases in the aggregation pathway,(10) providing insight to facilitate development of strategies that modulate the thermodynamically distinct surface-monomer interactions characteristic of secondary nucleation. However, to design therapeutically useful modulators of amyloid aggregation requires that this understanding be complemented with an understanding of molecular recognition between fibrils and monomers set within the context of other interactions occurring during aggregation (e.g. monomer-nuclei interactions). We have shown that Aβ40 mononomers and oligomers dock onto the fibril surface which catalyzes assembly of antiparallel strand formation in close situ to the parent Aβ16-22 fibre.

Whether this is the end-point product or further re-organisation is required to generate the canonical amyloid structure requires further study (longer simulation time). Furthermore,

241

alternative structures have been observed by solid-state NMR studies, e.g. the Iowa mutant of Aβ40 in which antiparallel fibrils were observed as metastable intermediates in the assembly process.(38)

Together the results demonstrate that kinetic analyses and theory together with molecular dynamics provide a powerful arsenal and capability to visualize secondary nucleation in structural and kinetic detail. Such approaches may allow informed targeting of this process to either prevent or accelerate secondary nucleation for therapeutic purposes and peptide materials assembly. Co- aggregation adds an additional layer of complexity in understanding molecular assembly yet represents an opportunity to manipulate these supramolecular assembly processes.

7.4 Materials and Methods

7.4.1 Synthesis of N-Fmoc TFMD-Phe and amyloid-β peptides

N-Fmoc TFMD-Phe was synthesized using the method described by Smith et al. and further minor changes in protecting group (Scheme S1).(39) Aβ16-22, TAMRA-Ahx-Aβ16-22 and

Aβ*16-22 were synthesized via both automated and manual solid-phase peptide synthesis and dissolved into dimethylsulfoxide (DMSO) stock solutions prior to use (Fig. S1). Aβ40 was synthesized recombinantly using the method of Walsh and co-workers and modifications by

Stewart and co-workers.(40, 41) To ensure that Aβ40 was monomeric prior to use, the peptide was purified by size exclusion chromatography, lyophilized and stored at -4 °C (Fig. S2).

7.4.2 Thioflavin T fluorescence assays

Samples were prepared in a 96-well non-binding plate (Corning Costar 3881, Corning Life

Sciences, Amsterdam, The Netherlands) sealed with clear sealing film (BMG Labtech, Aylesbury,

242

Bucks, UK) and were incubated in a FLUOstar OPTIMA plate reader (BMG Labtech, Aylesbury,

Bucks, UK) for 20 hours at 37 °C without agitation. Samples had a volume of 95 μL containing

10 μM ThT in 100 mM ammonium bicarbonate, pH 7.4 and a final concentration of 1% (v/v)

DMSO. For seeding experiments, Aβ16-22 was incubated at 50 μM for at least 24 hours in the same buffer as described above with the presence of fibrils confirmed by Transmission Electron

Microscopy (TEM, described below). Prior to the assay, the fibrils were probe sonicated for 5 s at

22% amplitude to generate “seeds”. The ThT experiments used excitation and emission filters of

430 and 485 nm. Each ThT experiment shown was repeated in independent assays on three different occasions with the traces shown in this work being representative of all repeats.

7.4.3 Transmission Electron Microscopy

TEM images were taken at the end of each experiment by removing 5 μL from the necessary well and incubating this sample on carbon-formvar grids for 30 s prior to staining with

2% (w/v) uranyl acetate solution for an additional 30 s as described by Preston et al.27 Images were taken on a JEM-1400 (JEOL Ltd., Tokyo, Japan) or a Technai F12 transmission electron microscope. Images were taken on a JEM-1400 (JEOL Ltd., Toyko, Japan) or an Tecnai T12 (FEI,

Oregon, USA) transmission electron microscopes. Images were taken using either a ATM CCD camera or a Gatan Ultrascan 1000 XP (994) CCD camera (JEM-1400) or an Ultrascan 100XP

(994) CCD camera (Tecnai F12). Once taken, images were processed using ImageJ (NIH).

7.4.4 General Sedimentation Protocol

Samples were taken at the desired time point and centrifuged (20 mins, 14,000 g, 4 °C).

Each sample was then separated into pellet and supernatant fractions, lyophilised overnight and

243

disaggregated in hexafluoroisopropanol (HFIP) for at least 2 hours. The HFIP was removed under a stream of N2 and the peptides were taken up in DMSO prior to analysis by high-resolution mass spectrometry (Bruker HCT ion-trap MS).

7.4.5 Fluorescence Quenching Assays

Wild type Aβ16-22 was spiked with 5% w/w TAMRA-Ahx-Aβ16-22 and incubated either in isolation or at a 1:1 ratio with Aβ40 (total peptide concentration 40 μM) in 100 mM ammonium bicarbonate buffer, pH 7.4 with a final concentration of 2% (v/v) of DMSO. Samples were placed in quartz cuvettes and analysed using a temperature-controlled fluorimeter at 37 °C. Time points were taken every 30 s for the duration of the experiment and TEM images (as described above) were taken at the end of each experiment to ensure the presence of fibrils. The TAMRA fluorophore was excited at 520 nm and emission recorded at 600 nm to reduce the inner filter effect.

7.4.6 ESI-IMS-MS analysis

All samples were prepared as described above and left to incubate at 37 °C without agitation for 5 mins. A Synapt HDMS quadrupole time-of-flight mass spectrometer (Micromass

UK Ltd., Waters Corpn., Manchester, UK), equipped with a Triversa NanoMate (Advion

Biosciences, Ithaca, NY, USA) automated nano-ESI interface was used in this study. The instrument has a travelling-wave IMS device situated in-between the quadrupole and the time-of- flight analysers, as described in detail elsewhere. Samples were analysed by positive ionization nanoESI (nESI) with a capillary voltage of 1.4 kV and a nitrogen nebulizing gas pressure of 0.8 psi. The following instrumental parameters were set: cone voltage 60 V; source temperature 60

244

°C; backing pressure 4.7 mbar; ramped travelling speed 7-20 V; travelling wave speed 400 m s-1;

IMS nitrogen gas flow 20 mL min-1; IMS cell pressure 0.55 mbar. The m/z scale was calibrated using aq. CsI cluster ions. Collision cross-section (CCS) measurements were estimated by use of a calibration obtained by analysis of denatured proteins (cytochrome c, ubiquitin, alcohol dehydrogenase) and peptides (tryptic digests of alcohol dehydrogenase and cytochrome c) with known CCSs obtained elsewhere from drift tube ion mobility measurements.(14, 33) Data were processed by use of MassLynx v4.1 and Driftscope softwave supplied with the mass spectrometer.

7.4.7 Photo-induced covalent cross-linking (PIC)

A 1:1 ratio of Aβ16-22/ Aβ*16-22 or Aβ*16-22/ Aβ40 (40 μM total peptide concentration) in 100 mM ammonium bicarbonate buffer, pH 7.4 with a final concentration of 1% (v/v) DMSO was incubated in eppendorf tubes for either 5 mins or 24 h. Samples were then irradiated for 30 s using a home built LED lamp at 365 nm, then removed, lyophilized overnight, taken up in hexafluoroisopropanol (HFIP) for at least 2 hours and vortexed to ensure any aggregates were disrupted. The HFIP was then removed under a stream of N2 and the sample re-suspended in 50/50

MeCN/H2O (v/v) + 0.05% formic acid to a final concentration of ~40 μM. Any cross-links were then analysed using the method previously described and the ESI-IMS-MS system as described above.(15)

7.4.8 Discontinuous Molecular Dynamics and PRIME20 Force Field

The simulation approach applied in this work is discontinuous molecular dynamics

(DMD), a fast alternative to traditional molecular dynamics, in combination with the PRIME20 force field, a four-bead-per-residue coarse-grained protein model developed in the Hall group.(42)

245

In the PRIME20 model, each of the 20 different amino acids contains three backbone spheres NH,

CαH, CO and one sidechain sphere R with a distinct hard sphere diameter (effective van der Waals radius) and distinct sidechain-to-backbone distances (R-CαH, R-NH, R-CO). The backbone hydrogen bonding interaction is modelled as a directional square well potential. In the original

PRIME20 force field, the potential function between any two sidechain beads on the twenty different amino acids (except glycine) is modelled as a single well potential, containing 210 different square well widths and 19 different square well depths. In this work, we follow Cheon’s approach to apply a double square well potential instead of the single square well for sidechain- sidechain interaction.(43) All the other non-bonded interactions are modelled as hard sphere interactions. A detailed description of the derivation of the geometric and energetic parameters of the PRIME20 model is given in our earlier work.(44)

7.4.9 Simulation Procedure

DMD/PRIME20 simulations were performed on the following systems: 1. six Aβ40 monomeric peptides; 2. six monomeric Aβ40 peptides with six monomeric Aβ16-22 peptides; 3. six

Aβ40 monomeric peptides in the presence of pre-formed two, three and four β-sheet Aβ16-22 protofilaments, respectively. The 2, 3 and 4 β-sheet Aβ16-22 protofilaments contain 21, 42 and 71 peptides, respectively. Each simulation is performed at two different total peptide concentrations

(1 and 5 mM). Similar seeding simulations have been performed in previous works.(45) The simulations are performed in the canonical ensemble (fixed number of particles, volume and

* temperature). The reduced temperature is defined to be T = kBT/εHB, where the hydrogen bonding energy, εHB=12.47kJ/mol. The reduced temperature is related to real temperature by using the equation T/K = 2288.46 T*- 115.79. The reduced temperature T* is chosen to be 0.20, which

246

corresponds to a real temperature of 342K. The system is maintained at a constant temperature by applying the Andersen thermostat. We have performed three to ten independent runs for each system.

7.5 Acknowledgments

General: We thank Dr Matt Iadanza for help with taking and processing the EM images.

Funding: This work was supported by National Institutes of Health Grants R01 EB006006 and in part by National Science Foundation (NSF) Research Triangle Materials Research

Science and Engineering Centers (MRSEC) Grant DMR-1121107 and by NSF grant CBET

581606 and by EPSRC grants EP/N035267/1, EP/N013573/1 and EP/KO39292/1. SER acknowledges funding from the ERC (grant agreement number 322408) and the Wellcome

Trust (204963). SJB gratefully acknowledges BBSRC for a PhD studentship grant

BB/J014443/1. The EM was purchased with funding from Welcome Trust (108466/Z/15/Z) and (094232/Z/10/Z). YW and CKH gratefully acknowledge the support of a Cheney

Visiting Scholar Fellowship from University of Leeds.

Author contributions: SJB performed the experimental work; YW performed simulations; KLS performed expression and purification of Aβ40; SER, CKH and AJW designed the research; SJB,

YW, AEA SER, CKH and AJW wrote the paper.

247

7.6 References

1. T. P. J. Knowles, M. Vendruscolo, C. M. Dobson, The amyloid state and its

association with protein misfolding diseases. Nat. Rev. Mol. Cell Biol. 15, 384-396

(2014).

2. L. Adler-Abramovich, E. Gazit, The physical properties of supramolecular peptide

assemblies: from building block association to technological applications. Chem. Soc.

Rev. 43, 6881-6893 (2014).

3. M. G. Iadanza, M. P. Jackson, E. W. Hewitt, N. A. Ranson, S. Ranson, A New Era for

Understanding Amyloid Structures and Disease. Nature Reviews Molecular Cell

Biology, in press (2018).

4. R. Kayed et al., Common structure of soluble amyloid oligomers implies common

mechanism of pathogenesis. Science 300, 486-489 (2003).

5. C. Haass, D. J. Selkoe, Soluble protein oligomers in neurodegeneration: lessons from

the Alzheimer's amyloid beta-peptide. Nat. Rev. Mol. Cell Biol. 8, 101-112 (2007).

6. D. Eisenberg, M. Jucker, The amyloid state of proteins in human diseases. Cell 148,

1188-1203 (2012).

7. S. I. A. Cohen, M. Vendruscolo, C. M. Dobson, T. P. J. Knowles, From macroscopic

measurements to microscopic mechanisms of protein aggregation. J. Mol. Biol. 421,

160-171 (2012).

8. S. I. A. Cohen et al., Proliferation of amyloid-β42 aggregates occurs through a

secondary nucleation mechanism. Proc. Natl. Acad. Sci. U. S. A. 110, 9758-9763

(2013).

9. G. Meisl et al., Differences in nucleation behavior underlie the contrasting aggregation

248

kinetics of the Aβ40 and Aβ42 peptides. Proc. Natl. Acad. Sci. U. S. A. 111, 9384-9389

(2014).

10. S. I. A. Cohen et al., Distinct thermodynamic signatures of oligomer generation in the

aggregation of the amyloid-β peptide. Nat. Chem. 10, 523-531 (2018).

11. G. Fusco et al., Structural basis of membrane disruption and cellular toxicity by α-

synuclein oligomers. Science 358, 1440-1443 (2017).

12. J. A. Varela et al., Optical Structural Analysis of Individual α-Synuclein Oligomers.

Angew. Chem. Int. Ed. 57, 4886-4890 (2018).

13. S. L. Bernstein et al., Amyloid-β protein oligomerization and the importance of

tetramers and dodecamers in the aetiology of Alzheimer's disease. Nat. Chem. 1, 326-

331 (2009).

14. L. M. Young, L.-H. Tu, D. P. Raleigh, A. E. Ashcroft, S. E. Radford, Understanding

co-polymerization in amyloid formation by direct observation of mixed oligomers.

Chem. Sci. 8, 5030-5040 (2017).

15. G. W. Preston, S. E. Radford, A. E. Ashcroft, A. J. Wilson, Covalent Cross-Linking

within Supramolecular Peptide Structures. Anal. Chem. 84, 6790-6797 (2012).

16. W. Zheng, M.-Y. Tsai, M. Chen, P. G. Wolynes, Exploring the aggregation free energy

landscape of the amyloid-β protein (1-40). Proc. Natl. Acad. Sci. U. S. A. 113, 11835-

11840 (2016).

17. T. Gurry, C. M. Stultz, Mechanism of Amyloid ‑ β Fibril Elongation. Biochemistry 53,

6981-6991 (2014).

18. B. Barz, B. Strodel, Understanding Amyloid-β Oligomerization at the Molecular Level:

The Role of the Fibril Surface. Chem. Eur. J. 22, 8768-8772 (2016).

249

19. B. Barz, Q. Liao, B. Strodel, Pathways of Amyloid-β Aggregation Depend on

Oligomer Shape. J. Am. Chem. Soc. 140, 319-327 (2018).

20. W. Qiang, W.-M. Yau, J.-X. Lu, J. Collinge, R. Tycko, Structural variation in amyloid-

β fibrils from Alzheimer's disease clinical subtypes. Nature 541, 217-221 (2017).

21. B. De Strooper, E. Karran, The Cellular Phase of Alzheimer's Disease. Cell 164, 603-

615 (2016).

22. M. Törnquist et al., Secondary nucleation in amyloid formation. Chem. Commun. 54,

8667-8684 (2018).

23. R. Cukalevski et al., The Aβ40 and Aβ42 peptides self-assemble into separate

homomolecular fibrils in binary mixtures but cross-react during primary nucleation.

Chem. Sci. 6, 4215-4233 (2015).

24. C. J. Sarell, P. G. Stockley, S. E. Radford, Assessing the causes and consequences of

co-polymerization in amyloid formation. Prion 7, 359-368 (2013).

25. M. Jucker, L. C. Walker, Self-propagation of pathogenic protein aggregates in

neurodegenerative diseases. Nature 501, 45-51 (2013).

26. L. O. Tjernberg et al., Arrest of β-Amyloid by a Pentapeptide Ligand. J. Biol. Chem.

271, 8545-8549 (1996).

27. J. J. Balbach et al., Amyloid fibril formation by A 16-22, a seven-residue fragment of

the Alzheimer's beta-amyloid peptide, and structural characterization by solid state

NMR. Biochemistry 39, 13748-13759 (2000).

28. M. C. Hsieh, C. Liang, A. K. Mehta, D. G. Lynn, M. A. Grover, Multistep

Conformation Selection in Amyloid Assembly. J. Am. Chem. Soc. 139, 17007-17010

(2017).

250

29. S. A. Petty, S. M. Decatur, Experimental evidence for the reorganization of β-strands

within aggregates of the Aβ(16-22) peptide. J. Am. Chem. Soc. 127, 13488-13489

(2005).

30. F. T. Senguen et al., Probing aromatic, hydrophobic, and steric effects on the self-

assembly of an amyloid-β fragment peptide. Mol. BioSys. 7, 486-496 (2011).

31. M.-C. Hsieh, C. Liang, A. K. Mehta, D. G. Lynn, M. A. Grover, Multistep

Conformation Selection in Amyloid Assembly. J. Am. Chem. Soc. 139, 17007-17010

(2017).

32. M. Cheon, I. Chang, Carol K. Hall, Spontaneous Formation of Twisted A 16-22

Fibrils in Large-Scale Molecular-Dynamics Simulations. Biophys. J. 101, 2493-2501

(2011).

33. L. M. Young et al., Insights into the consequences of co-polymerisation in the early

stages of IAPP and Amyloid-beta assembly from mass spectrometry. Analyst 20, 6990-

6999 (2015).

34. K. Garai, C. Frieden, Quantitative analysis of the time course of Aβ oligomerization

and subsequent growth steps using tetramethylrhodamine-labeled Aβ. Proc. Natl. Acad.

Sci. U. S. A. 110, 3321-3326 (2013).

35. R. Aleksis, F. Oleskovs, K. Jaudzems, J. Pahnke, H. Biverstål, Structural studies of

amyloid-β peptides: Unlocking the mechanism of aggregation and the associated

toxicity. Biochimie 140, 176-192 (2017).

36. N. Schwierz, C. V. Frost, P. L. Geissler, M. Zacharias, From Aβ Filament to Fibril:

Molecular Mechanism of Surface-Activated Secondary Nucleation from All-Atom

MD Simulations. J. Phys. Chem. B 121, 671-682 (2017).

251

37. G. Meisl et al., Differences in nucleation behavior underlie the contrasting aggregation

kinetics of the Aβ40 and Aβ42 peptides. Proc. Natl. Acad. Sci. U. S. A. 111, 9384-9389

(2014).

38. W. Qiang, W.-M. Yau, Y. Luo, M. P. Mattson, R. Tycko, Antiparallel β-sheet

architecture in Iowa-mutant β-amyloid fibrils. Proc. Natl. Acad. Sci. U. S. A. 109,

4443-4448 (2012).

39. D. P. Smith et al., Trifluoromethyldiazirine: an effective photo-induced cross-linking

probe for exploring amyloid formation. Chem. Commun., 5728-5730 (2008).

40. D. M. Walsh et al., A facile method for expression and purification of the Alzheimer's

disease-associated amyloid β-peptide. FEBS J. 276, 1266-1281 (2009).

41. K. L. Stewart, E. Hughes, E. A. Yates, D. A. Middleton, S. E. Radford, Molecular

Origins of the Compatibility between Glycosaminoglycans and Aβ40 Amyloid Fibrils.

J. Mol. Biol. 429, 2449-2462 (2017).

42. M. Cheon, I. Chang, C. K. Hall, Extending the PRIME Model for Protein Aggregation

to All Twenty Amino Acids. Proteins 78, 2950-2960 (2010).

43. M. Cheon, C. K. Hall, I. Chang, Structural Conversion of Aβ17–42 Peptides from

Disordered Oligomers to U-Shape Protofilaments via Multiple Kinetic Pathways.

PLoS Comput Biol 11, e1004258 (2015).

44. M. Cheon, I. Chang, C. K. Hall, Influence of temperature on formation of perfect tau

fragment fibrils using PRIME20/DMD simulations. Protein Sci. 21, 1514-1527 (2012).

45. Y. Wang, D. C. Latshaw, C. K. Hall, Aggregation of Aβ(17–36) in the Presence of

Naturally Occurring Phenolic Inhibitors Using Coarse-Grained Simulations. J. Mol.

Biol. 429, 3893-3908 (2017).

252

Figure 7.1. Co-aggregation of Aβ16-22 and Aβ40 results in accelerated aggregation kinetics for

Aβ40 (a) The primary sequence of Aβ16-22 and Aβ40, including the groups at each termini. (b) ThT fluorescence assays showing when mixing the two peptides (with the total peptide concentration held constant at 40 µM) the aggregation rate of Aβ40 increases as the ratio of Aβ16-22 to Aβ40 is increased. (c) Simulation snapshots of the aggregation of six Aβ40 monomers into a β-sheet rich hexamer at CAβ40 = 5mM; at the start of the simulation (0 µs) all the peptides are in random coils but as the simulation progresses they aggregate into parallel, in-register β-sheets (104 µs). This oligomer then unfolds, losing some of its β-sheet structure (230 µs) prior to a rearrangement in which the parallel β-sheets rearrange, forming a stable fibril with each Aβ40 peptide containing three β-strands (621 µs).

253

Figure 7.2. Aggregation kinetics of Aβ16-22 are unaffected by the presence of Aβ40. (a) Schematic showing the principle behind the fluorescence quenching assay used to determine the aggregation rate of Aβ16-22. (b) As self-assembly occurs, the TAMRA-labelled peptides (40 µM, total peptide, containing 5% (w/w) TAMRA-Ahx-Aβ16-22 are sequestered into the fibril structure. This brings the fluorophores into close proximity, resulting in fluorescence quenching. (c)

Aggregation of Aβ16-22 (containing 5% (w/w) TAMRA-Ahx-Aβ16-22) and Aβ40 at a 1:1 mol/mol ratio (total peptide concentration 40 µM). (d, e) Sedimentation and separation of the pellet and supernatant of the 1:1 mixed system after 1 h indicates that Aβ40 is present, as indicate by ESI-MS, in the (d) supernatant and only very small amounts within the (e) pellet. (f) Under these conditions, fibrils of Aβ16-22 are present after 5 mins incubation. Scale bar: 500 nm.

254

Figure 7.3. Aβ16-22 can interact with both Aβ40 monomers and dimers. (a) Native ESI-IMS-MS drift-scope images of Aβ40 indicate the presence of multiple oligomeric species of Aβ40 (white numbers). (b) When incubated at a 1:1 mol/mol ratio with Aβ16-22 (yellow numbers) a number of heteromeric species are observed (green numbers). The oligomer size is given (1, 2, 3 etc) with the charge state in superscript. (c) DMD simulation showing the percent β-sheet formed by Aβ40 during aggregation in the absence (red) or presence (black) of Aβ16-22. (d) Energy contact map between one monomer of Aβ16-22 and one of Aβ40 scaled by energy (bar shown alongside) showing that residues 17-20 (LVFF) and 31-34 (IIGL) form the strongest interactions. (e) Co-aggregation can have differing effects on the 1° nucleation of each peptide, depending on whether the mixed oligomers formed can progress to form mixed fibrils or are off-pathway and take no further part in the aggregation reaction. Circles represent monomers and blocks represent fibrils, with Aβ16-

22/Aβ40 in red and blue respectively. Adapted from (35).

255

Figure 7.4. Aβ16-22 fibrils increase the aggregation rate of Aβ40 to a greater extent than Aβ16-

22 monomers. (a) Increased concentrations (% w/w) of Aβ16-22 fibrils were added to Aβ40 monomers (as shown in the key) and the aggregation rate measured by ThT fluorescence. (b) Direct comparison of the effect of Aβ16-22 monomers (i.e. taken straight from a DMSO stock) and Aβ16-22 fibrils on Aβ40 aggregation. (c) Effect of sonicating the Aβ16-22 fibrils on Aβ40 aggregation rate shows little effect compared with the data shown in (a) (see text for details). (d) Plots of the percent

β-sheet formed by Aβ40 in the absence (blue) or presence of preformed 2 (black), 3 (red) and 4

(green) β-sheet Aβ16-22 fibrils showing that an increased number of Aβ16-22 fibrils increases the rate of Aβ40 aggregation. (e) During co-aggregation experiments both elongation and surface catalysed mechanisms can occur, each has a different effect on the rate of assembly of each peptide (the same notation is used as in Fig. 3e, with circles representing monomers, blocks fibrils and Aβ16-22/Aβ40 in red and blue, respectively). Adapted from (35).

256

Figure 7.5. Aβ16-22 and Aβ40 do not co-assemble during co-aggregation. Negative stain TEM analysis of Aβ40 incubated for 24 h in the (a) absence or (b) presence of Aβ16-22. Scale bar = 200 nm. (c) PIC of mixtures of diazirine labelled Aβ16-22 (Aβ*16-22) and Aβ40 incubated for 24 h and then irradiated for 30 s. Only homomolecular Aβ16-22 cross-links are observed, indicating that the fibrils are not co-polymerised at the end of the reaction (the inset depicts mechanism of PIC of the diazirine group (see also Fig. S12). (d) DMD simulation snapshots of co-aggregation of Aβ40 (blue) and Aβ16-22 (red) indicate that separate homomolecular oligomers are formed at t = 202 µs.

257

Figure 7.6. Aβ16-22 fibrils catalyse Aβ40 assembly through secondary surface nucleation. (a)

Simulation snapshots of the process by which Aβ16-22 fibrils (red) increase the aggregation rate of

Aβ40 (blue) through a surface catalysed 2° nucleation. (b) A schematic description of the mechanism is also included with Aβ40 in blue and Aβ16-22 in red.

258

7.7 Supplementary Materials

7.7.1 General Materials and Methods for Organic Synthesis

Non-aqueous reactions were carried out in washed and oven-dried glassware. Solvents and reagents were used as received from major suppliers without prior purification unless stated.

Anhydrous tetrahydrofuran (THF), ethanol (EtOH), acetonitrile (MeCN), dichloromethane (DCM) and diethyl ether (Et2O) were obtained from the in-house solvent purification system from

Innovative Technology Inc. PureSolv®. Anhydrous dimethyl formamide (DMF), methanol

(MeOH) and chloroform (CHCl3) were obtained from major chemical suppliers equipped with a

SureSeal™ (or equivalent). For reactions under non-anhydrous conditions, the solvents used were of HPLC quality and provided by Fisher or Sigma-Aldrich. Water in aqueous solutions and used for quenching was deionized, and water used for buffers and HPLC was ultra-pure 18 MΩ from an ELGA Purelab system. 1H NMR spectra were recorded on Bruker DPX 300 (300 MHz) or

Avance 500 (500 MHz) spectrometers and referenced to either residual non-deuterated solvent

1 peaks or tetramethylsilane. H spectra are reported as follows: δH (spectrometer frequency, solvent): ppm to two decimal places (number of protons, multiplicity, J coupling constant in hertz, assignment). Chemical shifts are quoted in ppm with signal splitting recorded as singlet (s), doublet

(d), triplet (t), quartet (q), quintet (quin.) multiplet (m), broad (br) and apparent (app.). Coupling constants, J, are measured to the nearest 0.1 Hz.

7.7.2 Synthesis of N-Fmoc protected TFMD-Phe

The synthesis of TFMD-Phe was carried out according to the literature procedure originally developed by Fishwick and co-workers and improved upon by Smith and co-workers.5,6 A change in protecting group has been made from the original Smith and co-workers synthesis. All synthetic

259

products have been analysed by at least 1H NMR to confirm the identity and purity of each compound. Principally, 1H NMR, FTIR and LC-MS analysis were used. All data has been consistent with those previously reported within the group and in the literature.6,7

7.7.3 Synthesis of TFMD-Phe

7.7.4 General Materials and Methods for Aβ16-22 solid-phase peptide synthesis (SPPS)

All amino acids, coupling reagents and resins were purchased from Novasyn (Merck),

Fluorochem or Sigma-Aldrich. All amino acids were N-Fmoc protected and side chains were protected with Boc (Lys) or OtBu (Glu). Aβ16-22, TAMRA-Ahx-Aβ16-22 and Aβ*16-22 were

260

synthesized using standard SPPS conditions on an automated solid-phase peptide synthesiser

(CEM LibertyBlue) for Aβ16-22 and TAMRA-Ahx-Aβ16-22 or manually for Aβ*16-22. DMF, DCM and Et2O used in peptide synthesis was of ACS grade from Sigma-Aldrich. Two different N- termini capping strategies were used:

N-terminal acetyl capping (Aβ16-22 and Aβ*16-22): The deprotected resin was washed with

DMF (5 x 2 min x 2 mL), DCM (5 x 2 min x 2 mL) and Et2O (5 x 2 min x 2 mL). After this acetic anhydride (10 equivalents) and DIPEA (10 equivalents) were added to the resin, which was agitated overnight on a blood spinner at room temperature.

N-terminal TAMRA labelling (TAMRA-Ahx-Aβ16-22): Prior to coupling of the fluorescent label a 6-aminohexanoic acid (6-Ahx) linker would be coupled using standard SPPS conditions.

For fluorescent labelling, TAMRA (5 equivalents) was dissolved in a solution of DMF (2 mL) containing oxyma (5 equivalents), DIC (5 equivalents) and left to stir for 10 mins. The pre- activated TAMRA solution was added to resin, which was agitated overnight on a blood spinner at room temperature in the dark.

After cleavage from the resin using TFA:TIPS:H2O (98:1:1, 3 mL, 2 hr) the crude product was precipitated using ice cold Et2O (50 mL) and the supernatant removed. The crude peptides identity was confirmed by LC-MS prior to HPLC purification.

7.7.5 General Materials and Methods for HPLC purification

Peptides were purified by preparative scale HPLC using an X-bridge C18 preparative column (reversed phase) on an increasing gradient of MeCN to H2O. Crude peptides were dissolved in DMSO and injected into the column in 300 μL aquilots. The solvent gradient (5 –

95% MeCN in H2O + 0.1% formic acid) was increased linearly over a 15 min run time at a flow

261

rate of 10 mL/min-1. The fixed-wavelength detector was set to scan the eluent at either 220, 254 or

270 nm with peak-based collection for 30 s after the diode was triggered. Fractions were analyzed by LC-MS and fractions containing the desired peptide were pooled, concentrated under reduced pressure and lyophilized. The purity of the peptide samples was determined by the School of

Chemistry HPLC service. Identity was confirmed by mass using either the LC-MS or HRMS.

7.7.6 Analytical MS and HPLC data for synthetic peptides

a

b

c

262

Figure S7.1. HRMS (left) and analytical HPLC traces (right) of Aβ16-22 and its variants. The core molecular structure of Aβ16-22 (a) and the associated traces for Aβ16-22 (b), Aβ*16-22 (c) and TAMRA-Ahx-Aβ16-22 (d) are shown.

7.7.7 General materials and methods for recombinant peptide synthesis

Commercial E. coli strain BL21 (DE3) cells (Agilent) were transformed with a pETSAC plasmid containing the sequences for Aβ40 (a gift from Profs. Sara Linse and Dominic Walsh), including an N-terminal methionine residue that has no effect on the fibrillation of Aβ40 or the

1 morphology of fibrils formed . Cultures were grown in LB media and Aβ40 purified as described previously.2 Final protein concentrations were estimated from UV absorption in 7 M guanidinium hydrochloride at 280 nm using an extinction coefficient of 1490 M-1cm-1, and averaged 4 mg pure peptide/L culture. HRMS was used to confirm the identity of the final product (expected molecular weight 4459.21 Da). The size exclusion chromatography trace from the second purification step

(in 50 mM ammonium bicarbonate) is also shown. Fraction 11 and 12 were pooled and lyophilised prior to use.

Figure S7.2. SEC trace of Aβ40 indicates that there is a single peak and ESI-IMS-MS indicates that in the gas phase Aβ40 is largely monomeric. The observed m/z ratio is stated above each peak.

263

7.7.8 Additional Characterization and Analyses

Figure S7.3. Supplementary ThT data. The Aβ40 concentration series (a) used to normalise the half-times of Aβ40 aggregation in the presence of Aβ16-22 (b). Aβ*16-22 (blue) has the same effect on the aggregation rate of Aβ40 as WT Aβ16-22 (red, c). The end point ThT fluorescence data when increasing amounts of preformed Aβ16-22 fibrils are added to Aβ40 show no significant difference in fluorescence, indicating that majority homomolecular fibrils are formed at the end of the self- assembly reaction (d). Conditions: 100 mM ammonium bicarbonate, pH 7.4 with a final concentration of 1% (v/v) DMSO, 37 °C, quiescent.

264

Figure S7.4. Supplementary negative stain TEM images. Aβ16-22 forms fibrils that bundle together (a), after probe sonication the fibrils are shorter and have more fibril ends per aggregate mass (b). Buffer: total peptide concentration 50 μM, 100 mM ammonium bicarbonate, pH 7.4 with a final concentration of 1% (v/v) DMSO, sonication 5 s at 22% amplitude. When incubated at 5% of the total peptide concentration, mixtures of TAMRA-Ahx-Aβ16-22 and WT Aβ16-22 form fibrils that are morphologically similar to those formed by WT Aβ16-22 (c and d (zoomed). Buffer: total peptide concentration 40 μM, 100 mM ammonium bicarbonate, pH 7.4 with a final concentration of 2% (v/v) DMSO, quiescent, 37 °C.

265

Table S7.1. The expected and observed m/z values for monomeric and oligomeric Aβ40 in isolation and in the presence of a 1:1 ratio of Aβ16-22. The expected and observed m/z of the heteromeric oligomers of Aβ40/Aβ16-22 are also shown (bottom).

Monomers

Aβ40 Aβ40 1:1 1:1 alone alone Aβ /Aβ Aβ /Aβ Peak 40 16-22 40 16-22 m/z m/z m/z m/z expected observed expected observed 12+ 2230.60 2230.79 2230.60 2230.79 13+ 1487.40 1487.50 1487.40 1487.50 14+ 1115.80 1115.88 1115.80 1115.88 15+ 892.84 892.91 892.84 892.91 1+ Aβ16-22 1 N/A N/A 894.49 894.58 Dimers Aβ40 Aβ40 1:1 1:1 alone alone Aβ /Aβ Aβ /Aβ Peak 40 16-22 40 16-22 m/z m/z m/z m/z expected observed expected observed 24+ 2230.60 2230.79 2230.60 2230.79 25+ 1784.68 1784.98 1784.68 1784.98 26+ 1487.40 1487.42 N/A N/A Trimers, Tetramers and Pentamers Aβ40 Aβ40 1:1 1:1 alone alone Aβ /Aβ Aβ /Aβ Peak 40 16-22 40 16-22 m/z m/z m/z m/z expected observed expected observed 35+ 2676.53 2677.14 2676.53 2677.14 36+ 2230.60 2230.79 N/A N/A 46+ 2973.81 2974.03 N/A N/A 47+ 2549.12 2549.81 N/A N/A 58+ 2788.00 2788.9 N/A N/A Heteromeric species (Aβ40/ Aβ16-22) Peak m/z expected m/z observed 1:13+ 1785.33 1785.37 1:23+ 2083.17 2083.58 1:33+ 2381.00 2181.73 2:15+ 1963.50 1963.95 2:25+ 2142.20 2142.71 2:35+ 2320.90 2321.19

7.7.9 Collision Cross-section (CCS) analysis of Aβ40 in the presence and absence of Aβ16-22

Collision cross-section (CCS) measurements were estimated by use of a IMS-MS calibration obtained by analysis of denatured proteins (cytochrome c, ubiquitin, alcohol

266

dehydrogenase) and peptides (tryptic digests of alcohol dehydrogenase and cytochrome c) with known CCSs obtained elsewhere from drift tube ion mobility measurements.3,4 The proteins were denatured in 50/:50 H2O/MeCN + 0.1% formic acid prior to introduction to the mass spectrometer.

The CCS (Ω) of the peptide monomers/oligomers was then calculated according to the equation below:

2 퐵 1 1 Ω (Å ) = A × (푡퐷) × 푧 × √ + 푚𝑖표푛 푚푔푎푠 where A is the determined calibration constant, z is the charge state of the ion, B is the exponential

’ factor (determined experimentally), tD is the corrected absolute drift time, mion is the mass of the ion and mgas is the mass of the gas used in the ion-mobility cell (N2).

Figure S7.5. Analysis of the CCS values for Aβ40 in the absence (red) or presence (blue) of Aβ16-22 (blue, a – c) over different IMS experiments. The data are separated into monomer (a), dimer (b) and trimer (c) and broken down by charge state. The average CCS for the mixed oligomers is also shown (d).

267

a

b

c

Figure S7.6. PIC analysis of 1:1 Aβ*16-22/Aβ40 at 5 mins and 24 h. A full scan of the cross-linked region of m/z 800 – 2300 is shown at 5 mins (with the Aβ*16-22 cross-linked region shown as the inset, the assignments can be found in Table S2, a). Tandem MS/MS spectra of one of the dimer peaks (m/z 893.5) is shown at 5 mins (b) and 24 h (c), demonstrating that Aβ*16-22 forms in-register antiparallel fibrils.

268

Table S7.2. Assignments of each of the major peaks observed in Figure S6a (inset).

Observed at Peak Charge m/z Structure Comments 24 h?

intramolecular a 1+ 974.5 cross-link Yes (e = a + Na+)

intermolecular b 2+ 983.5 Aβ*16-22 - Aβ*16- Yes 22 cross-link

intermolecular c 2+ 988.5 Aβ*16-22 - Aβ*16- Yes 22 cross-link

H O adduct d 1+ 992.6 2 Yes (g = d + Na+)

Parent f 1+ 1002.5 No ion/diazoisomer

double H O h 1+ 1008.5 2 Yes adduct

269

Figure S7.7. Plot of the average number of hydrogen bonding and sidechain-sidechain contacts between six Aβ40 monomers and a 3-β-sheet Aβ16-22 fibril during the simulation at CAβ40 = 5mM. As shown in Fig. 6b, Aβ40 monomers attach to both the lateral surface and the end of the Aβ16-22 fibril with its C-terminal part attaching more frequently to the lateral surface of the fibril than its two ends. Error bars depict the standard deviation calculated from 3 independent runs of the seeding simulation.

PDB Files for individual snapshots are included as supporting data files number according to the figure and time at which they appear e.g. Fig 1c-0.pdb, Fig 1c-104.pdb

270

CHAPTER 8

Future Work

8.1 Parallelization of the DMD/PRIME20 code

Accelerating computational speed for molecular dynamics simulation is the key to investigating large biologically relevant systems at long time scales. In this work, one of the most important ways to improve the performance of DMD/PRIME20 simulation is to increase its computation speed. In the future, we will try to parallelize the current DMD/PRIME20 code so that each DMD/PRIME20 simulation job uses multiple central processing units (CPU) at one time.

Parallelized DMD/PRIME20 code is expected to have a two- to six-fold increase in speed depending on the number of CPUs assigned to each job and the system size. In fact, parallelization of traditional molecular dynamics code is straightforward and can be achieved by using the concept of “domain decomposition” 1. However, parallelizing DMD code is much more difficult because

DMD is an intrinsically serial algorithm in which the system proceeds by only one collision event at a time. The strategy we will apply for parallelizing DMD code is based on the concept developed by Khan et al.2 called “event-based decomposition” where multiple events can be processed at one time instead of processing one event at each time. Later, Shirvanyants et al.3 implemented Khan’s concept and successfully developed a parallelized atomistic DMD package for studying protein folding problems. They have showed that the parallelized DMD code can be speeded up to be six times faster than the serial one. In Shirvanyants’ algorithm, the collision event is processed serially in the master node but the computation burden is offloaded by assigning multiple collision events to multiple slave nodes after which the master node eliminates the invalid events and records the collision information for each of the valid events.

271

8.2 Construction of oligomerization free energy profile of Aβ16-22 peptide

According to the “toxic oligomer hypothesis” 4, the disease-causing toxic species that bring damage to the cerebral neuron cell are the transient and soluble oligomers formed during early-aggregation stage rather than the mature fibril formed at the end-point of the aggregation process. DMD/PRIME20 has proven in this thesis to be a robust technique to simulate early-stage protein oligomerization and nucleus formation. However, to accurately locate important oligomeric species along the aggregation pathway, we need good analysis tools to decipher long trajectories generated from DMD simulation. As is well-known to all, the free energy landscape of protein aggregation is extremely challenging to construct because the system contains a large number of degrees of freedom which lead to a complex high-dimension free energy surface.5 Dr.

Sergei Krivov of University of Leeds has developed an iterative algorithm6 to optimize a single reaction coordinate that delineates the free energy profile with preserved major free energy barrier.

This approach has been applied to construct the free energy landscape for the folding of a single protein.7 In future work, we will collaborate with Dr. Sergei Krivov to construct the oligomerization free energy landscape for Aβ16-22 peptide. To achieve this, we need to generate peptide aggregation trajectories containing millions of frames and we also need to calculate the number of inter-peptide hydrogen bonds and the root mean square displacement with respect to the reference structure as candidates for initial reaction coordinates. These informations, including the trajectory and system properties, will serve as initial input for constructing free energy profiles with an optimized reaction coordinate. Then we will need to select the most representative structures associated with each of the major peaks and valleys to determine the transition pathway of a multi-peptide system from the initially disordered state to the most stable sate.

272

8.4 References

1. Brown, D., Clarke, J. H. R. Okuda, M., Yamazaki, T. (1993) A domain decomposition

parallelization strategy for molecular dynamics simulations on distributed memory

machines. Comput. Phys. Comm. 74, 67-80

2. Khan, M. A., Herbordt, M. C. (2011) Parallel discrete molecular dynamics simulation with

speculation and in-order commitment. J. Comput. Phys. 230, 6535-6582

3. Shirvanyants, D., Ding, Feng., Tsao, D., Ramachandran, S., Dokholyan, N. V. (2012)

Discrete Molecular Dynamics: An Efficient And Versatile Simulation Method For Fine

Protein Characterization. J. Phys. Chem. B 116, 8375-8382

4. Ferreira, S. T., Klein, W. L. (2011) The Aβ oligomer hypothesis for synapse failure and

memory loss in Alzheimer's disease. Neurobiol. Learn Mem. 96, 529-543

5. Zheng, W., Tsai, M.-Y. , Chen, M., Wolynes, P. G. (2016) Exploring the aggregation free

energy landscape of the amyloid-β protein (1–40). Proc. Natl. Acad. Sci. U. S. A. 113,

11835-11840

5. Krivov, S., Karplus, M. (2006) One-dimensional free-energy profiles of complex systems:

Progress variables that preserve the barriers. J. Phys. Chem. B 110, 12689-12698

6. Krivov, S., Karplus, M. (2006) Protein folding free energy landscape along the committor-

the optimal folding coordinate. J. Chem. Theory Comput. 14, 1418-1427

273