You are on page 1of 87

MARMARA UNIVERSITY

INSTITUTE FOR GRADUATE STUDIES


IN PURE AND APPLIED SCIENCES

COMPUTATIONAL INVESTIGATION OF
PEPTIDE BINDING AFFINITY AND COMPLEX
STABILITY OF MAJOR
HISTOCOMPATIBILITY COMPLEX (MHC)

ASUMAN BUNSUZ

M.Sc. THESIS
Department of Bioengineering

ADVISOR
Assoc. Prof. Dr. PEMRA ÖZBEK

ISTANBUL, 2018
MARMARA UNIVERSITY
INSTITUTE FOR GRADUATE STUDIES
IN PURE AND APPLIED SCIENCES

COMPUTATIONAL INVESTIGATION OF
PEPTIDE BINDING AFFINITY AND COMPLEX
STABILITY OF MAJOR
HISTOCOMPATIBILITY COMPLEX (MHC)

ASUMAN BUNSUZ
524214012

M.Sc. THESIS
Department of Bioengineering

ADVISOR
Assoc. Prof. Dr. PEMRA ÖZBEK

ISTANBUL, 2018
i
ACKNOWLEDGEMENTS

I would like to present my greatfullness to my thesis supervisor Assoc. Prof. Dr. Pemra
ÖZBEK who accepted me as a MSc student for her project. Her vision, encouragement,
patience and support are valuable sources in developing my professional skills. I would like
to acknowledge The Scientific and Technological Council of Turkey TUBITAK in the frame
work of the Project No. 113M293 for the financial support during my thesis.

I would also thank to my colleague, Research Assistant Onur SERÇİNOĞLU who taught me
computational methods patiently, respondend to my questions so promptly and that
throughout these years had always encouraging words.

I wish to thank to my team member Elif Naz BİNGÖL for all the meals and laughs we had.
Thank you for love, friendship and encouragement.

Special thanks are going to Hüsniye GENÇ who encouraged me to apply for a master
program at Marmara University and for her love and friendship, Nilay YÖNET for her
positive energy and emotional support, Ekin UZUNHASANOĞLU for offering me advice,
and supporting me through the thesis process.

I owe my greatest thanks to my parents Selma and İhsan BUNSUZ, my brother Abdulkerim
BUNSUZ and my sister Reyhan BUNSUZ for love, constant support, encouragement, never
ending patience and inspiring me to follow my childhood dream. Thank you.

JUNE, 2018 ASUMAN BUNSUZ

i
TABLE OF CONTENT

PAGE

ACKNOWLEDGEMENTS ..................................................................... ii
TABLE OF CONTENT ........................................................................... ii
ABSTRACT ............................................................................................ iv
ÖZET .........................................................................................................v
LIST OF SYMBOLS .............................................................................. vi
LIST OF ABBREVATIONS ................................................................. vii
CHAPTER 1 AIM OF THE STUDY ..................................................13
CHAPTER 2 INTRODUCTION.........................................................15
2.1 Structure of Major Histocompatibility Complex (MHC) ............................... 15
2.2 HLA peptide binding affinity and stability ..................................................... 18
2.3 Molecular Dynamics (MD) Simulations......................................................... 21
2.3.1 Theoretical Background of Molecular Dynamics Simulations................... 22
2.3.2 Molecular Dynamics (MD) Simulations of peptide-HLA Class I
Complexes............................................................................................................... 24
2.4 Molecular Docking of HLA Class I Molecules .............................................. 26
2.4.1 DockTope Server ........................................................................................ 28
CHAPTER 3 MATERIALS AND METHODS ..................................30
3.1 HLA-A*02:01 Structures ................................................................................ 30
3.2 Molecular Docking of HLA-A*02:01 Structures ........................................... 33
3.3 Preparations of Structures for MD Simulations .............................................. 33
3.4 Analysis of the MD Simulation Trajectories .................................................. 34
3.4.1 Root Mean Square Deviation (RMSD) Calculations ................................. 34
3.4.2 Root Mean Square Fluctuation (RMSF) calculations ................................. 35
3.5 Calculation of pairwise amino acid non-bonded interaction energies via
gRINN tool.................................................................................................................. 35
CHAPTER 4 RESULTS AND DISCUSSIONS ...................................36
4.1 Analysis of RMSD profiles................................................................................... 36
4.2 Analysis of RMSF profiles ................................................................................... 39
4.3 Analysis of interaction energies using gRINN ..................................................... 42

ii
CHAPTER 5 CONCLUSION ............................................................67
CHAPTER 6 FUTURE WORK .........................................................69
REFERENCES ....................................................................................... 70
AUTOBIOGRAPHY .............................................................................. 83

iii
ABSTRACT

COMPUTATIONAL INVESTIGATION OF PEPTIDE BINDING


AFFINITY AND COMPLEX STABILITY OF MAJOR
HISTOCOMPATIBILITY COMPLEX (MHC)

Major histocompatibility complex (MHC) molecules are heterodimeric cell surface


glycoproteins playing a novel role in the regulation of the adaptive immune response by
binding antigenic peptides and presenting them at the cell surface for T cell recognition. In
humans, MHC molecules are called as Human Leukocyte Antigens (HLA). HLA-A2 is the
most common HLA-A allele in different ethnic populations. Among the HLA-A2 allelic
variants, HLA-A*02:01 is the most frequent one and it is commonly used to study HLA-A2-
restricted cytotoxic T lymphocyte (CTL) responses. Recently, studies have focused on
designing peptide based vaccines; hence the factors affecting the peptide immunogenicity has
become a popular subject for detailed studies. Several studies proposed that affinity and
stability are the two most significant parameters for peptide immunogenicity. Therefore, in
order to investigate the peptide binding affinity and the stability of peptide-MHC complexes,
we have employed various computational techniques. We performed molecular docking
studies for 12 peptide-HLA-A*02:01 complexes using DockTope web server. Then, we
carried out 100 ns parallel molecular dynamics simulations for the docked structures using
GROMACS software with the OPLS-AA/L all-atom force field at 310 K and 473 K. In the
final step, we calculated pairwise amino acid non-bonded interaction energies via gRINN tool
and revealed the critical residues that play a role in the observed stability differences for these
12 peptide- HLA-A*02:01 complexes. As a result, our computational results are found to be
inline with the experimental studies demonstrating that stability is a more significant
parameter for peptide immunogenicity and that computational methods can also be used for
the future studies in terms of determining stability of a peptide complex.

iv
ÖZET

MAJOR HISTOKOMPATİBİLİTE KOMPLEKSİNİN PEPTİT


BAĞLANMA AFİNİTESİNİN VE KOMPLEKS STABİLİTESİNİN
HESAPLAMALI ARAŞTIRILMASI

Majör Histokompatibilite Kompleks Molekülleri (MHC) antijenik peptitlere bağlanarak ve


onları T hücre tanınması için hücre yüzeyinde sunarak adaptif immune cevabın
düzenlenmesinde özgün bir rol oynayan heterodimerik hücre yüzey glikoproteinleridir.
İnsanlarda MHC molekülleri İnsan Lökosit Antijeni (HLA) olarak adlandırılır. HLA-A2
farklı etnik popülasyonlarda yaygın olarak bulunan HLA-A alelidir. HLA-A*02:01 HLA-A2
alel varyantları arasında en yaygın olan ve HLA-A2 sınırlı sitotoksik T lemfosit cevaplarını
çalışmak için çoğunlukla kullanılan alleldir. Son zamanlarda çalışmalar peptit temelli aşıların
tasarlanması üzerine odaklanmıştır bu nedenle peptit immunojenitesini etkileyen faktörler
detaylı araştırmalar için popüler konu olmuştur. Çeşitli çalışmalar o peptit immunojenitesi
için afinite ve stabilitenin iki en önemli parametre olduğunu ileri sürmektedir. Bu yüzden,
peptit bağlanma afinitesini ve peptit-MHC kompleks stabilitesini araştırmak için, çeşitli
hesaplamalı teknikler kullandık. 12 peptit-HLA-A*02:01 kompleksi için DockTope web
sunucusunu kullanarak moleküler doking çalışmaları gerçekleştirdik. Ardından, modellenmiş
yapıların 100 ns parallel moleküler dinamik simülasyonlarını OPLS-AA/L potansiyel enerji
fonksiyonunu kullanarak 310 K ve 473 K de gerçekleştirdik. Son aşamada, çiftli aminoasit
etkileşim enerjilerini gRINN aracılığıyla hesapladık ve bu 12 peptit-HLA-A*02:01
kompleksinin gözlenen stabilite farklılıklarında önemli rol oynayan kritik rezidüleri ortaya
çıkardık. Sonuç olarak, hesaplamalı sonuçlarımız peptit immunojenitesi için stabilitenin daha
anlamlı parametre olduğunu gösteren deneysel çalışmalarla aynı doğrultudadır.

v
LIST OF SYMBOLS

V Volume

N Number of Atoms

E Energy

μ Chemical Potential

T Temperature

P Pressure

H Enthalpy

F Force

m Mass

a Acceleration

r Atomic Position

v Velocity

p Potential Energy

X Coordinates of All Carbon Alpha Atoms

kB Boltzmann Constant

vi
LIST OF ABBREVATIONS

MHC Major Histocompatibility Complex

HLA Human Leukocyte Antigens

CTL Cytotoxic T Lymphocyte

SPR Surface Plasmon Resonance

MM/PBSA Molecular Mechanics Poisson-Boltzmann Surface Area

LIE Linear Interaction Energy

FEP Free Energy Perturbation

TI Thermodynamic Integration

MD Molecular Dynamics

BPTI Bovine Pancreatic Trypsin Inhibitor

PDB Protein Data Bank

NMR Nuclear Magnetic Resonance

AMBER Assisted Model Building with Energy Refinement

MOE Molecular Operating Environment

OPLS-AA/L All-Atom Optimized Potentials For Liquid Simulations

GROMACS GROningen MAchine for Chemical Simulations

RMSD Root Mean Square Deviation

RMSF Root Mean Square Fluctuation

PCA Principal Component Analysis

Cα Carbon Alpha

TOP Topology

gRINN get Residue Interaction eNergies and Networks

vii
ALA Alanine

ARG Arginine

ASN Asparagine

ASP Aspartic acid

CYS Cysteine

GLN Glutamine

GLY Glycine

GLU Glutamic acid

HIS Histidin

ILE Isoleucine

LEU Leucine

LYS Lysine

MET Methionine

PHE Phenylalanine

PRO Proline

SER Serine

THR Threonine

TRP Tryptophan

TYR Tyrosine

VAL Valine

viii
LIST OF FIGURES

Figure 2.1 Schematic representation of MHC Class I (A) and MHC Class II (B)[40]. .............. 15
Figure 2.2 Structures of peptide binding groove for MHC Class I (A) and Class II (B)[40]. .... 16
Figure 3.1 Peptide binding groove pockets of HLA-A*02:01 (PDB ID : 3HLA [51]) are shown
in PyMOL representation. Colored dots represent the pocket residues interacting with the peptide
residues (A: red, B: green, C: teal, D: pink, E: yellow, F: turquouse) ......................................... 31
Figure 3.2 Peptide binding groove pockets of HLA-A*02:01 (PDB ID : 1HHH [119]) and the
interacting peptide residues are shown in PyMOL representation. Macromolecule is shown in
gray and peptide is shown in lightgreen sticks. Colored dots represent the pocket residues
interacting with the peptide residues shown in blue stick. (A) P1, (B) P2, (C) P6, (D) P3, (E) P7,
(F) P9 ............................................................................................................................................ 32
Figure 4.1 Root mean square deviation (RMSD) profile of the modeled complexes during the
100 ns simulations at 310 K. ......................................................................................................... 36
Figure 4.2 Root mean square deviation (RMSD) profile of the modeled complexes during the
100 ns simulations at 473 K. ......................................................................................................... 36
Figure 4.3 Distribution of alpha carbon root mean square deviation (RMSD) values of the
modeled complexes during the 100 ns simulations at 310 K........................................................ 37
Figure 4.4 Distribution of alpha carbon root mean square deviation (RMSD) values of the
modeled complexes during the 100 ns simulations at 473 K........................................................ 38
Figure 4.5 Root mean square fluctuations (RMSF) values of the modeled complexes during the
100 ns simulations at 310 K. ......................................................................................................... 39
Figure 4.6 Root mean square fluctuations (RMSF) values of the modeled complexes during the
100 ns simulations at 473 K. ......................................................................................................... 40
Figure 4.7 Peptide RMSF profiles of the modeled complexes during the 100 ns simulations at
310 K. ............................................................................................................................................ 41
Figure 4.8 Peptide RMSF profiles of the modeled complexes during the 100 ns simulations at
473 K. ............................................................................................................................................ 41
Figure 4.9 Correlation between the experimental stability measurements and our computational

ix
energy values obtained from gRINN. ........................................................................................... 43
Figure 4.10 Motion of peptides and peptide binding grooves throughput the all trajectory.
Macromolecule is shown in gray, peptide binding groove is shown in military green, peptides are
shown in colored tubes. (ILDDNLYKV: red, GLFDFVNFV: orange, SLSAYIIRV: yellow,
YLPEVISTI: purple, RLYDYFTRV: magenta, LMYDIINSV: brown, VLYDEFVTI: pink,
RVYEALYYV: green, YLYFCSSDV: cyan, YLYQPCDLL: cream, HVDGKILFV: blue,
VLPFDIKYI: black) ..................................................................................................................... 44
Figure 4.11 Interaction energies between residues of HLA and β2-microglobulin. .................... 46
Figure 4.12 Key residues for the interaction between HLA and β2-microglobulin (ILDDNLYKV
peptide) are shown in PyMOL representation. Colored dots represent the important residues
interacting with β2-microglobulin (ARG48: magenta GLY119: red, ARG202: green, peptide:
orange, macromolecule: gray ) ..................................................................................................... 47
Figure 4.13 Interaction energies of residue GLY119. .................................................................. 47
Figure 4.14 Interaction energies of residue ARG48..................................................................... 48
Figure 4.15 Interaction energies of residue ARG202................................................................... 49
Figure 4.16 Interaction energies between residues of HLA and peptide residues. ...................... 50
Figure 4.17 Interaction energies between region 1 (residues 60-80) and peptide residues.. ...... 51
Figure 4.18 Interaction energies between region 2 (residues 140-170) and peptide residues. .. 52
Figure 4.19 Important residues for interaction between HLA and peptide (ILDDNLYKV
peptide) are shown in PyMOL representation. Colored dots represent the important residues
interacting with β2-microglobulin (GLU163: red, LYS66: blue, TRP146: cyan, ASP77: yellow,
peptide: orange, macromolecule: gray )........................................................................................ 53
Figure 4.20 Interaction energies of residue GLU63..................................................................... 54
Figure 4.21 Interaction energies of residue GLU66..................................................................... 54
Figure 4.22 Interaction energies of residue TRP146. .................................................................. 55
Figure 4.23 Interaction energy profiles of the peptide residues. .................................................. 56
Figure 4.24 Interaction energies of residue P1............................................................................. 57
Figure 4.25 Interaction energies of residue P2............................................................................. 58
Figure 4.26 Interaction energies of residue P3............................................................................. 59
Figure 4.27 Interaction energies of residue P4............................................................................. 60
Figure 4.28 Interaction energies of residue P5............................................................................. 60

x
Figure 4.29 Interaction energies of residue P6............................................................................. 61
Figure 4.30 Interaction energies of residue P7............................................................................. 61
Figure 4.31 Interaction energies of residue P8............................................................................. 62
Figure 4.32 Interaction energies of residue P9............................................................................. 63
Figure 4.33 Interacting residues of ILDDNLYKV, GLFDFVNFV, SLSAYIIRV, YLPEVISTI,
RLYDYFTRV and LMYDIINSV peptides. ................................................................................. 65
Figure 4.34 Interacting residues of VLYDEFVTI, RVYEALYYV, YLYFCSSDV,
YLYQPCDLL, HVDGKILFV and VLPFDIKYI peptides. ......................................................... 66

xi
LIST OF TABLES

Table 3.1 Dataset of HLA-A*02:01 alleles selected for the study. ............................................. 30
Table 3.2 Peptide binding groove pockets and the interacting peptide residue positions of HLA-
A*02:01 (PDB ID : 3HLA [51] and PDB ID : 1HHH [119]) ....................................................... 31
Table 4.1 Interaction energies of 12 peptide HLA-A*02:01 complexes obtained from gRINN
....................................................................................................................................................... 42

xii
CHAPTER 1 AIM OF THE STUDY

Major histocompatibility complex (MHC) molecules are heterodimeric cell surface glycoproteins

playing a novel role in the regulation of the adaptive immune response by binding antigenic

peptides and presenting them at the cell surface for T cell recognition [1]–[3].

In recent years, studies have focused on designing peptide based vaccines, hence the factors

affecting the peptide immunogenicity has become a popular subject for detailed studies [4]–[6].

Therefore, many experimental approaches have been developed to measure the peptide binding

affinity and pMHC complex stability [7]–[16].

Early experimental studies reported that peptide immunogenicity is highly related to peptide

binding affinity [17], [18]. However, several studies proposed that stability is a more accurate

parameter for peptide immunogenicity [19]–[22]. Moreover, in 2012, Harndahl et al [10]

demonstrated that although having similar binding affinities, immunogenic peptides form more

stable peptide-MHC complexes than non-immunogenic peptides. In addition, several

experimental studies have revealed that highly stable peptide-MHC complexes have longer half-

lives on the cell surface to activate CTLs, while complexes having low stabilities are released

from the cell surface in a shorter time duration [12], [23]. Therefore, stability of peptide-MHC

complexes seems to be a significant determinant for immunogenicity when compared with

peptide binding affinity [10], [12], [23].

Although there are several experimental methods in literature, molecular dynamics (MD)

simulation methods have been widely used to understand structural, kinetic and thermodynamics

properties of peptide-MHC complexes at the atomic level. MD simulations provide the most

detailed information related to intermolecular and intramolecular forces that are important for

13
protein stability [24], [25].

Furthermore, high temperature MD simulations have been extensively used in order to estimate

the stability of proteins comparatively. [26]–[31]. Through these simulations, differences in

folding processes and the protein stability comparisons can be conducted within shorter

simulation times. Therefore, in order to explore the stability of peptide-MHC complexes, high

temperature MD simulations can be utilized.

In humans, MHC molecules are called as Human Leukocyte Antigens (HLA). HLA-A2 is the

most common HLA-A allele in different ethnic populations. Among the HLA-A2 allelic variants,

HLA-A*02:01 is the most frequent one and it is commonly used to study HLA-A2-restricted

CTL responses [32].

In this study, we have investigated peptide binding affinity and complex stability on 12 peptide-

HLA-A*02:01 modeled complexes. Firstly, molecular docking studies were performed for these

peptides using Docktope web server [33]. Then, MD simulations were carried out on the docked

structures using GROMACS software [34] with the OPLS-AA/L all-atom force field [35] for 100

ns at 310 K and 473 K. RMSD, RMSF and PCA analysis were performed on the equilibrated

MD simulation trajectories. In the final step, pairwise amino acid non-bonded interaction

energies were calculated via gRINN tool [36]

We compared our computational results with previous experimental measurements [10], [13]

and investigated the correlations among them. We have also determined the critical residues

which play a role in the observed differences in the stability of these 12 peptide- HLA-A*02:01

complexes.

14
CHAPTER 2 INTRODUCTION

2.1 Structure of Major Histocompatibility Complex (MHC)

Major histocompatibility complex (MHC) molecules are heterodimeric cell surface glycoproteins

playing a novel role in regulation of the adaptive immune response by binding antigenic

peptides and presenting them at the cell surface for T cell recognition [2], [3], [37].

MHC molecules are classified in to two categories according to their structure and function;

MHC Class I and MHC Class II [38]. MHC Class I molecules are composed of a heavy chain (α

chain) consisting of three domains (α1, α2 and α3) and a light chain (β2-microglobulin), while

MHC Class II molecules are made up of an α (α1 and α2) and a β (β1 and β2) chain (Figure 1.1)

[39].

A B

Figure 2.1 Schematic representation of MHC Class I (A) and MHC Class II (B)[40].

In MHC Class I, α1 and α2 helices constitute the sides of the peptide binding groove while the

base of the binding groove is formed by the eight stranded β pleated sheet (Figure 1.2) [41] .

Binding groove is closed at terminal regions and an endogenous peptide between 8 to 11 amino

15
acids in length is accommodated. After peptide binding, peptide-MHC Class I complex is

recognized by CD8 T lymphocytes and activated cytotoxic T lymphocytes (CTLs) eliminates

target cells presenting MHC-bound antigenic peptides [39] .

A B

Figure 2.2 Structures of peptide binding groove for MHC Class I (A) and Class II (B)[40].

In MHC Class II, the sides of the peptide binding groove are constituted by α1 and β1 helices

and the base of the binding groove is constructed by a β sheet (Figure 1.2) [41]. Peptide binding

groove of MHC Class II is open at both ends, allowing peptides to extend out of the groove and

generally, an exogenous peptide of 13-25 amino acids in length is held in the binding groove via

hydrogen bonding interactions. After peptide binding, peptide-MHC Class II complex is

recognized by CD4 T lymphocytes [39].

In humans, MHC molecules are called as Human Leukocyte Antigens (HLA). Genes encoding

HLAs located on the short arm of chromosome 6 (6p21.3) are the most polymorphic genes in the

human genome. HLA Class I molecules are encoded by HLA-A, HLA-B and HLA-C genes,

while HLA Class II molecules are encoded by HLA-DP, HLA-DQ and HLA-DR genes [42].

16
Earlier experimental studies focused on identifying sequence of peptides presented by HLA

molecules. These studies revealed allele specific peptide binding motifs for various alleles of

HLA Class I consisting of HLA-A1 [43], A11 [44], A2.1 [45], A3 [46], A68 [47], B27 [48] and

B8 [49]. In 1987, three dimensional structure of HLA-A2 allele belonging to HLA Class I was

detailed by Bjorkman using X-ray crystallography [50]. Since the determination of the first

crystal structure of HLA Class I molecule, a variety of three dimensional structures of HLA

molecules have been elucidated. Experimental and structural analyses have revealed that

polymorphic residues in the peptide binding groove play essential roles in determining allele

specific repertoire of HLA molecules [51], [52].

To date, several computational algorithms depending on peptide sequence and structure of the

MHC molecules have been proposed for the prediction of possible T cell epitopes for specific

MHC alleles [53]. Sequence based prediction algorithms such as; sequence motif, motif

matrices, machine learning motifs (Artificial Neural Network and Support Vector Machine),

Hidden Markov models and Quantitative matrices make use of the knowledge related to peptides

known to bind HLA molecules [54]–[63]. On the other hand, structure based algorithms such as

molecular docking, threading and molecular dynamics simulations require the three dimensional

structures of MHC molecules [64]–[67].

Moreover, sequence based databases and web servers have been developed to facilitate

predictions of peptide binding affinity and stability of MHC-peptide complexes [68]–[73].

However, sequence based predictions are dependent on the knowledge related to peptides known

to bind HLA molecules [53]. On the other hand, structure based predictions can provide more

detailed information at the molecular level. Therefore, structure based predictions have become

increasingly important in recent years [74]–[76].

17
2.2 HLA peptide binding affinity and stability

In protein-ligand interactions, the term ‘binding affinity’ is used to describe the force of

attraction between the ligand and the receptor protein [77]. On the other hand, definition of

stability can be classified into two classes; thermodynamic stability and kinetic stability. The

thermodynamic protein stability is defined as the difference in free energies between the folded

and the unfolded states, whereas the kinetic stability of a protein is based on the energy barrier

between the folded and the unfolded states of the protein [78].

Recently, studies have focused on designing peptide based vaccines, hence the factors affecting

the peptide immunogenicity, defined as the ability of a peptide to stimulate CTL mediated

immune response, has become a popular subject for detailed studies [4]–[6]. Therefore, many

experimental approaches have been developed to measure the peptide binding affinity and

pMHC complex stability [7]–[16].

Labeling peptides by radioactive, biotinyl or fluorescent markers is one of the most common

approaches in the direct measurement of peptide binding to MHC molecules [12]–[16].

However, peptide binding features may change in response to alterations in the labeled peptides.

Therefore, need for alternative methodologies have emerged [7]–[9], [79]

On the other hand, indirect competitive binding assay approaches have been performed with the

use of synthetic peptides and monoclonal antibodies [80]. Moreover, Khilko et al. [7] used

surface plasmon resonance (SPR) technology for the investigation of peptide-MHC binding. In

1992, Parker et al. [11] have introduced a new indirect in vitro assay approach based on the

dissociation of radiolabelled β2m from MHC-peptide complexes having unlabeled heavy chain

and peptide. With this approach, interactions between heavy chain and peptide have been

18
measured without labeling. The first assay approach based on dissociation of labeled β2m led to

utilization of different methods for labeling in similar assays.

Harndahl et al. [9] have developed high throughput homogenous binding assay to determine the

peptide binding affinity for MHC molecules via application of Alpha Screen technology.

Furthermore, the same group have introduced high throughput homogenous β2m dissociation

assay for calculating the stability of peptide-MHC complexes [79]. These two binding assays are

also called scintillation proximity based immunoassays.

Besides these experimental approaches, computational methods for the calculation of binding

free energy are also extremely popular. These computational approaches can be divided into

two; endpoint and pathway approaches [81]. Endpoint approaches including linear interaction

energy (LIE) [82] and molecular mechanics Poisson-Boltzmann surface area (MM/PBSA)

methods [83] calculate the difference in binding free energy between the two extremes; unbound

and bound states of ligand-protein complexes. Pathway approaches including free-energy

perturbation (FEP) and thermodynamic integration (TI) methods [84] calculate binding free

energy along the transition pathway between the end point states [81], [85]. Although, pathway

approaches provide highly accurate results, the end-point approaches are more efficient and cost

effective [85].

Early experimental studies reported that peptide immunogenicity is highly related to peptide

binding affinity [17], [18]. However, several studies proposed that the stability is a more accurate

parameter for peptide immunogenicity [19]–[22]. Moreover, in 2012, Harndahl et al. [10]

demonstrated that immunogenic peptides form more stable peptide-MHC complexes than non-

immunogenic peptides having similar binding affinities. In addition, several experimental studies

have revealed that highly stable peptide-MHC complexes have longer half-lives on the cell

19
surface to activate CTLs, while complexes having low stabilities are released from the cell

surface in a shorter time duration [12], [23] . Moreover, high and intermediate affinity peptides

with high stability were classified as immunogenic peptides, while high affinity peptides with

low stability were classified as weak immunogenic or non-immunogenic peptides based on their

interaction with CTLs [23]. Therefore, stability of peptide-MHC complexes seems to be the

major determinant factor for immunogenicity when compared with the peptide binding affinity

[10], [12], [23].

Although there are several experimental and computational methods in literature, molecular

dynamics (MD) simulation methods have been widely used to understand structural, kinetic and

thermodynamics properties of peptide-MHC complexes at the atomic level. MD simulations

provide the most detailed information related to intermolecular and intramolecular forces that are

important for protein stability [24], [25] and can guide the future experiments in this field.

In order to estimate the stability of proteins comparatively, high temperature MD simulations

have been extensively used [26]–[31]. Through these simulations, differences in folding

processes and the protein stability comparisons can be conducted within shorter simulation

times. Therefore, in order to explore the stability of peptide-MHC complexes, high temperature

MD simulations can be utilized.

In 2004, Zacharias et al. [86] studied the changes in conformational dynamics of α1 and α2

domains of HLA*02:01 complexed with GVYDGREHTV peptide- by performing MD

simulations at two different temperatures 310 K and 355 K for 26 ns. Nevertheless, the short

simulations at 355 K did not achieve complete unfolding process and they observed partial

unfolding of α1 and α2 domains at high temperature. MD simulations of the peptide free HLA-A

at 355 K revealed that the binding of F-pocket region to C-terminus of the peptide is more

20
flexible than the binding to N terminus of the peptide. Since the peptide C and N terminus

provides the stabilization of the α1-α2 peptide binding domain, this study motivated us to carry

out high temperature MD simulations for the investigation of peptide-MHC complex stability.

2.3 Molecular Dynamics (MD) Simulations

In 1957, the first MD simulation was accomplished by Alder and Wainwright [87] in order to

investigate the equilibrium features of hard spheres and later in 1964, the first realistic MD

simulation was performed using Lennard-Jones potential for liquid argon by Rahman [88]. The

first application of MD simulations to proteins was carried out by McCommon [89] in 1977, in

order to study the folding dynamics of Bovine Pancreatic Trypsin Inhibitor (BPTI).

In MD simulations, the relation between macroscopic and microscopic properties of a system is

clarified using statistical mechanics. Thus, microscopic properties obtained from MD simulations

are converted to macroscopic properties. Thermodynamic systems have a number of

macroscopic states such as temperature, pressure and in MD simulations, the thermodynamic

states are called ensembles that are a collection of variety of microscopic states [90]. The five

most common ensembles used in MD simulations are canonical (NVT), micro canonical (NVE),

grand canonical (μVT), isobaric-isothermal and isobaric-isoenthalpic (NPH) [85].

In canonical ensemble (NVT), number of atoms, volume and temperature are held fixed. The

microcanonical ensemble (NVE) describes a system having constant number of atoms (N),

volume (V) and energy (E). Grand canonical ensemble (μVT) corresponds to fixed chemical

potential (μ), volume (V) and temperature (T). For isobaric-isothermal ensemble (NPT), number

of atoms (N), pressure (P) and temperature (T) are held constant. Isobaric-isoenthalpic ensemble

(NPH) has fixed number of atoms (N), pressure (P) and enthalpy (H) [85].

21
2.3.1 Theoretical Background of Molecular Dynamics Simulations

In MD simulations, Newton’s equations of motion [Eq. 2.1] is solved numerically for each atom

of the molecular system and trajectories. Each atoms position, velocity and energy is produced

asa function of time to give insight into time dependent development of the system.

2.1

Positions and velocities of all atoms are calculated via the [Eq. 2.2]

2.2

where F is force on an atom, m is the mass of the atom, r is the atoms position and V is the

potential energy. The force F on each atom is also computed using negative gradient of the

potential energy of the atom.

The starting positions of the molecules are taken from structures determined by X-ray or NMR

techniques. Starting velocities are assigned by Gaussian- Boltzman distribution [Eq. 2.3] where

pi denotes the momentum of the ith atom, vx represents the velocity of the atom, m is the mass, kB

is the Boltzmann constant and T indicates the temperature.

 mi   1 mi vi2, x 
=p ( vi , x )  exp  − 
 2π k B 
T  2 k B T 
2.3

The temperature is calculated based on the mean kinetic energy equation via the [Eq. 2.4]

mi ( vi2 )
N
2 1
T=
3 Nk B

i =1 2 2.4

22
where N is overall number of unconstrained degrees of freedom, v represents velocity of an atom

i at the time t, m displays mass and kB is the Boltzmann constant.

The integration of Newton’s equation of motion is performed numerically by variety algorithms

using finite difference methods. The Verlet algorithm depending on Taylor series expansion is the

most widely preferred algorithm. In Verlet algorithm [Eq. 2.5, Eq. 2.6, Eq. 2.7], r represents

position, v denotes velocity, a denotes acceleration and t is time

1
r ( t + ∆=
t ) r ( t ) + v ( t ) ∆t + a ( t ) ∆t 2 … 2.5
2

1
v ( t + ∆=
t ) v ( t ) + a ( t ) ∆t + b ( t ) ∆t 2 … 2.6
2

a ( t + ∆=
t ) a ( t ) + b ( t ) ∆t … 2.7

Velocities are computed non-explicitly via [Eq. 2.8, Eq. 2.9, Eq. 2.10]

1
r ( t + ∆=
t ) r ( t ) + v ( t ) ∆t + a ( t ) ∆t 2 2.8
2

1
r ( t − ∆=
t ) r ( t ) − v ( t ) ∆t + a ( t ) ∆t 2 2.9
2

r ( t +=
∆t ) 2r ( t ) − r ( t − ∆t ) + a ( t ) ∆t 2 2.10

Force field includes the parameters and the functions to calculate the potential energy of the

system. In this study, OPLS-AA/L force field is used. The OPLS-AA/L force field was first

introduced by Jorgensan [91] to study dynamics of organic liquids. Later, this force field has

been used for proteins, carbohydrates, nucleic acids and drug molecules. Moreover, OPLS-AA/L

force field has been developed for proteins by refitting the Fourier torsional coefficients [8]. The

total potential energy of the system is calculated by summing the non-bonded, bond strenging,

bending and torsional energies.

23
2.3.2 Molecular Dynamics (MD) Simulations of peptide-HLA Class I Complexes

There have been various MD studies on HLA molecules over the years. In 1992, Rognan et

al.[92] generated a peptide-HLA-A2 complex model for influenza virus matrix protein IMP 58-

66 peptide (GILGFVFTL) by manually docking the peptide into peptide binding cleft of HLA-

A2 crystal structure and carried out 100 ps long MD simulations of this modeled complex using

AMBER software [93]. This study revealed important residue positions for peptide binding,

including Tyr7, Tyr59, Glu63, Tyr84, Thr143, Tyr159 and Tyr171. Their results appeared to be

inline with the findings obtained from the crystal structure of the HLA-B*27 allele.

Moreover, Toh et al. [94] modeled HLA-B*02:17-ALPHAILRL complex by exchanging peptide

residues and the three residues 95, 97 and 99 within the binding groove of HLA-B*02:01-

ILKEPVHGV complex. Then, they performed 1000 ps long MD simulations using Discover 95.0

program. Although these three residues differ between HLA-B*02:17 and HLA-B*02:01, these

two alleles had distinct peptide motifs. For HLA-B*02:17 allele, prolin at P3 position seemed to

be significant for stable peptide binding.

In 2007, Joseph et al. [95] generated a variety of HLA*02:01-peptide complex models using

HLA-A*02:01-IISAVVGIL crystal structure by exchanging P6 and P7 positions and carried out

0.2 ns long MD simulations using Molecular Operating Environment (MOE) 2001.01 program

to study the effect of amino acid replacements on GP2 peptide. This study revealed that

GLU7PHE mutant affects the peptide-HLA complex stability the most and peptide residues P5-

P7 may have important roles for the stabilization of P9 side chain.

Pohlmann et al. [96] conducted 20 ns long MD simulations on the HLA-B*27:05 and HLA-

B*27:09 alleles bound to the m9 peptide using GROMACS [34] software in 2004. They

observed that m9 peptide-HLA-B*27:09 complex was more flexible compared to the m9

24
peptide-HLA-B*27:05 complex and revealed that the micropolymophism at residue 116 causes

the differential peptide dynamics.

In 2008, Fabian et al. [97] performed 40 ns long MD simulations of the HLA-B*27:05 and HLA-

B*27:09 alleles bound to pVIPR peptide using GROMACS software. For pVIPR-HLA-B*27:05

complex, the canonical and the non- canonical peptide conformations were simulated separately.

The results demonstrated that the heavy chain of the canonical form of HLA-B*27:05 was more

flexible compared to the heavy chain of HLA-B*27:09. However, the canonical form of HLA-

B*27:05 was more thermostable than HLA-B*27:09. But, similar flexibilities were observed

between the non- canonical forms of HLA-B*27:05 and HLA-B*27:09.

Narzi et al. [98] investigated the impact of micropolymorphism and the effect of peptide binding

on the dynamics of the binding groove of HLA-B*27:05 and HLA-B*27:09 alleles bound to a

viral (pLMP2) and three self-peptides (pVIPR, pGR, and TIS) by performing 400 ns long MD

simulations of these peptide-HLA complexes using GROMACS software. All peptides exhibited

higher conformational flexibilities at alpha 1 and peptide binding groove regions when bound to

HLA-B*27:05. This study demonstrated that the dynamics of the peptide binding groove is

related to the micropolymorphisms and the bound peptides.

In 2015, Abualrous et al. [99] performed 50 ns long MD simulations of HLA-B*27:05 and HLA-

B*27:09 alleles bound to IRAAPPPLF peptide. As a result of their study, they found that the

polymorphic residue HIS116 neutralizes the negative charge by binding surrounding residues to

provide stability, while ASP116 causes charge repulsion and higher flexibility in the F pocket.

Thus, it is concluded that HLA-B*27:05 is dependent upon tapasin for peptide loading.

25
Ozbek [100] studied the dynamics of HLA-B*44:02, HLA-B*44:03 and HLA-B*44:05 alleles

bound to three different peptides (EEFGRAFSF, EEYLKAWTF, EEYLQAFTY) by conducting

20 ns long MD simulations of these three peptide-HLA complexes in 2016. It is demonstrated

that peptide binding stability is more dependent on the peptide sequence rather than the allele.

In 2017, Sercinoglu et al. [101] performed 50 ns long MD simulations of HLA-B*27:05 and

HLA-B*27:09 alleles bound to four different peptides (m9, pLMP2, pCatA, and TIS) in order to

study the effect of residue interactions and the micropolymorphism at residue 116 on the global

dynamics of peptide-HLA complexes. As a result, they observed the strong interaction between

the different domains of the peptide-HLA complexes and revealed that in the absence of peptide,

dynamic interaction within the residues of HLA-B*27:09 allele is stronger than the HLA-B*27:05

allele.

In 2017, Ayres et al. [102] performed full atomistic 1 us MD simulations of structurally

determined peptide-HLA-A*02:01 complexes and generated a mathematical model for

calculating peptide motion in order to predict peptide immunogenicity. They revealed that

rigidity and hydrophobicity are the two most significant factors which have an effect on the

immune cellular response. They have also concluded that the stability of the peptide can be

varied by substituting peptide amino acids.

2.4 Molecular Docking of HLA Class I Molecules

Understanding the molecular basis of biological systems mediated through protein-protein

interaction mechanisms provide valuable information on the development of new drugs. Due to

the expensive and time consuming experimental methods, many protein structures have not yet

been defined experimentally, hence three dimensional structures of all these proteins are not

available in Protein Data Bank (PDB) [103], [104].


26
Recently, the advances in molecular modeling methods have enabled the modeling of protein

complexes with unknown three dimensional structures. One of the most important applications

of molecular modeling methods in drug development is protein-ligand docking [104], [105] .

Molecular docking methods generate a series of probable complex structures from its unbound

constituents by predicting the favored conformation of the ligand in the binding site of a protein.

Each probable complex structure is ranked depending on the global minimum free energy of

protein-ligand complex by scoring functions. Thus, the structure with the lowest score is ranked

as the best and is selected as a near native structure [105]–[107].

During the last two decades, various docking programs have been developed for docking of

ligand molecules into proteins such as Autodock [108] , Glide [109], GOLD [110] and

HADDOCK [111]. Due to the peptides having large number of rotatable bonds that are more

flexible than the small ligand molecules, most of the popular docking programs are not

applicable to peptide-protein docking problems. Therefore, in the recent years, much

improvement have been achieved in the field of peptide-protein docking programs such as DINC

[112], FlexPepDock [106] and DockTope [33].

Liu et al. [113] carried out docking studies using FlexPepDock for HLA Class I alleles including

HLA-A2, HLA-B27, HLA-B35, HLA-B44, and HLA-E to generate a special protocol. In their

study, RMSD (root mean square deviation) values less than 2 Å and 1 Å were determined for the

near native and subangstrom models, respectively. As a result, 66% of docking studies were

scored as subangstrom models. Rigo et al. [33] developed DockTope and a generated refinement

protocol for HLA-A*02:01 and HLA-B*27:05 alleles by using structurally determined peptide-

HLA complexes. Their calculated RMSD values are less than 1 Å and 2 Å for Cα and all atoms,

27
respectively. In 2018, Antunes et al. [114] achieved redocking studies using DINC for HLA Class

I alleles and found RMSD values less than 2 Å for all atoms.

In this study, Docktope server [33] is used for the modeling of the HLA-peptid complexes.

2.4.1 DockTope Server

DockTope is a web server [33], which aims to model the bound conformations of peptide-MHC

Class I complexes. Molecular docking is performed by an eight-step procedure involving

structural modeling of MHC Class I peptides, energy minimization of modeled peptides, first

molecular docking of peptides into MHC molecules, selection of the best pose, energy

minimization of the best pose generated in the first docking process, the final molecular docking

process, selection of the best pose created in final docking process and production of three

dimensional structures of peptide-MHC Class I complexes [33].

In the first step, target peptide sequence is provided in the form of a single aminoacid code by

the user and the target peptide is modeled using PyMol [115]. Then, energy minimization of the

modeled peptide is performed to eliminate unnecessary interactions. During the first molecular

docking process, Autodock tools and Autodock Vina [116] are employed. Previous to docking,

hydrogen and Gasteiger charges are added to both the peptide and MHC Class I molecule. For

the peptide, torsions are additionally specified. In a following step, search space is determined in

order to obtain the best pose in the peptide binding groove. Each run is compiled for 20 times.

Then, the best poses are selected from each run. After the first docking process, undesirable

interactions are removed by performing energy minimization using GROMACS [117]. After this

step, molecular docking is performed for a second time using the same procedure as above and

the best pose is selected based on the algorithm generated by Rigo et al [33]. In the final step,

28
three dimensional structure of peptide-MHC Class I complex is created and the structure file is

emailed to the user.

29
CHAPTER 3 MATERIALS AND METHODS

3.1 HLA-A*02:01 Structures

Among the HLA-A2 allelic variants, HLA-A*02:01 is the most frequent one and it is commonly

used to study HLA-A2-restricted CTL responses [32]. In this study, we have investigated the

peptide binding affinity and peptide-HLA-A*02:01 complex stabilities on 12 modeled

complexes. Experimental studies were conducted on these peptide-HLA complexes [10], [118]

previously. Table 3.1 demonstrates the details of these structures along with the peptide

sequences, their immunogenic characters, experimental affinity and stability measurements

generated by Assarsson [118] and Harndahl [10] and a predicted affinity measurement by

Harndahl [10].

Table 3.1 Dataset of HLA-A*02:01 alleles selected for the study.

Assarsson Harndahl Predicted


Observed
affinity, affinity, affinity, nM
HLA-A*02:01 Peptide Sequence Stability
nM nM
(T1/2)h
2 1 3
Immunogenic ILDDNLYKV 28

Immunogenic GLFDFVNFV 1 2 2 23
Immunogenic SLSAYIIRV 2 2 5 23
Immunogenic YLPEVISTI 2 1 5 19
11 4 3
Immunogenic LMYDIINSV 14

Immunogenic RLYDYFTRV 2 1 3 13
Immunogenic VLYDEFVTI 5 1 10 12
Immunogenic RVYEALYYV 2 1 2 11
Immunogenic HVDGKILFV 39 41 160 3

Nonimmunogenic YLYFCSSDV 2 1 7 10

Nonimmunogenic YLYQPCDLL 2 1 14 7
15 20 149
Nonimmunogenic VLPFDIKYI 1

30
For all HLA-A*02:01 alleles, peptide binding groove is divided into six binding pockets from A

to F [51]. Peptide binding motif is determined by the two primary anchors, P2 and P9. While P2

binds to B-pocket, P9 binds to F-pocket of the binding groove [119]. The pockets and the

interacting peptide residue positions are shown in Table 3.2, Figure 3.1 and Figure 3.2.

Table 3.2 Peptide binding groove pockets and the interacting peptide residue positions of HLA-

A*02:01 (PDB ID : 3HLA [51] and PDB ID : 1HHH [119])

Interacting Positions of
Pockets Residues
the Peptide
5, 7, 59, 63, 66, 159, 163, 167,
A 171 [51] P1 [119]

9, 24, 45, 63, 66, 67, 70, 99 [51]


B P2 [119]
9, 70, 73, 74, 97 [51]
C P6 [119]
99, 114, 156, 160 [51]
D P3 [119]
97, 114, 147, 152, 156 [51]
E P7 [119]
77, 80, 81, 84, 116, 123, 143,
F 146, 147 [51] P9 [119]

Figure 3.1 Peptide binding groove pockets of HLA-


A*02:01 (PDB ID : 3HLA [51]) are shown in PyMOL
representation. Colored dots represent the pocket
residues interacting with the peptide residues (A: red,
B: green, C: teal, D: pink, E: yellow, F: turquouse)

31
Figure 3.2 Peptide binding groove pockets of HLA-A*02:01 (PDB ID : 1HHH [119]) and the
interacting peptide residues are shown in PyMOL representation. Macromolecule is shown in
gray and peptide is shown in lightgreen sticks. Colored dots represent the pocket residues
interacting with the peptide residues shown in blue stick. (A) P1, (B) P2, (C) P6, (D) P3, (E) P7,
(F) P9

32
3.2 Molecular Docking of HLA-A*02:01 Structures

Molecular docking studies for the peptides given in Table 3.1 were performed by DockTope web

server [33] following the procedure given below:

1. An account is created to use the web server for modeling peptide-MHC Class I

complexes.

2. For each requested model, a special job name is entered to the text box provided.

3. Peptide sequence to be modeled is entered to the second text box.

4. Preferred MHC Class I allotype is selected among possible allotypes.

5. Job is submitted.

6. When the job is completed, the resulting peptide-MHC Class I structure pdb file is sent to

the registered email address.

3.3 Preparations of Structures for MD Simulations

MD simulations were performed for 100 ns at 310 K and 473 K by an eight step procedure

given below.

1. All hydrogen and water molecules were removed from the peptide-MHC Class I

structures obtained from the Docktope Server.

2. Structure pdb file was converted to GROMACS [34] topology file using the OPLS-AA/L

all-atom force field (2001 amino-acid dihedrals) [35] .

3. Protein was solvated in a cubic box placed at a distance of 2 nm using the Simple Point

Charge water system.

4. To neutralize the net charge of the system, an appropriate number of sodium and chloride

counter ions were added.

33
5. The potential energy of the solvated system was minimized using the steepest descent

energy minimization algorithm with an energy step size of 0.01 kcal/mol.

6. Minimization step was followed by an NVT equilibration, where the number of atoms,

volume and temperature were all fixed. This process was performed in 100 picoseconds.

7. The system was then equilibrated using a NPT ensemble, in which the pressure is

maintained isotropically at 1 bar using the Berendsen thermostat with a temperature

coupling constant of 0.1 ps.

8. In final step, MD simulation was performed for 50.000.000 steps with an integration time

0.002 picoseconds (ps).

MD simulations were carried out through TUBITAK TR GRID computing infrastucture and

workstations supplied in CompBio Laboratory of Bioengineering Department of Marmara

University.

3.4 Analysis of the MD Simulation Trajectories

3.4.1 Root Mean Square Deviation (RMSD) Calculations

Root mean square deviation (RMSD) is a quantative measurement between the two atomic
particles. Atom positions are identified with 3-dimensional vectors. Each vector is considered in
the RMSD calculation.

RMSD is calculated using the [Eq. 3.1]

∑(r − r0,i )
1 2
=
RMSDt t ,i
N i =1 3.1

where N is the number of the Cα atoms in the system, rt,i indicates ith atom atomic position at any
time t and r0,i indicates the Cα position in the minimized structure.

34
In this study, RMSD values were calculated only for the Cα atoms of residues after

superimposing the conformations produced in the trajectories on to the first conformation.

3.4.2 Root Mean Square Fluctuation (RMSF) calculations

Root mean square fluctuation (RMSF) is a measurement of the deviation between the
position of particle i and the reference position. RMSF is calculated using the [Eq. 3.2].

3.2

where T is the number of frames obtained by the MD simulation trajectory, is the position of

the atom i at time t and represents the average atomic position throughout the MD simulation

trajectory.

In this study, RMSF values were calculated from the average position of each Cα atom in the

equilibrated MD trajectory.

3.5 Calculation of pairwise amino acid non-bonded interaction energies via gRINN tool

Pairwise residue interaction energy was calculated using of the gRINN tool [36].

1. gRINN tool was downloaded from web page generated by Serçinoğlu et al.

2. For each calculation, special folder including topology (TOP), run input (TPR) and MD

trajectory (XTC) files were generated.

3. For each folder, gRINN interface was started and new calculation mode was selected for

pairwise residue interaction energy.

4. Stride value was set to be 150.

5. Number of processors was selected as 8.

6. Pairwise residue interaction energy was displayed using view results interface.

35
CHAPTER 4 RESULTS AND DISCUSSIONS

4.1 Analysis of RMSD profiles

All MD simulations were carried out for 100 ns and the structures are observed to reach

equilibrium after 40 ns. RMSD and RMSF analysis were performed on the last 60 ns of the MD

simulation trajectories. The RMSD figures are given in Figure 4.1-4.2 for 310 K and 473 K,

respectively.

Figure 4.1 Root mean square deviation (RMSD) profile of the modeled complexes during the
100 ns simulations at 310 K.

Figure 4.2 Root mean square deviation (RMSD) profile of the modeled complexes during the
100 ns simulations at 473 K.

36
When the simulations are conducted at a higher temperature, the RMSD values are observed to

be at a higher scale as well. Figure 4.3 demonstrates the distributions of the RMSD values at 310

K. In the presence of ILDDNLYKV peptide, HLA-A*02:01 allele displays the narrowest RMSD

distribution, while the widest RMSD distribution is observed in the presence of YLYYQPCDLL

and HVDGKILFV peptides. In addition, for ILDDNLYKV peptide, RMSD distribution profile is

observed at the leftmost compared to all other peptides. However, in presence of VLPFDIKYI

peptide, RMSD distribution profile is seen at the rightmost, meaning fluctuations at higher

RMSD values.

Figure 4.3 Distribution of alpha carbon root mean square deviation (RMSD) values of the
modeled complexes during the 100 ns simulations at 310 K.

37
Figure 4.4 Distribution of alpha carbon root mean square deviation (RMSD) values of the
modeled complexes during the 100 ns simulations at 473 K.

Figure 4.4 shows the distributions of the RMSD values at 473 K. We observe that the stability of

all peptide-HLA-A*02:01 complexes decrease with increasing temperature. For ILDDNLYKV

peptide-HLA*02:01 complex, RMSD distribution profile is seen at the leftmost with respect to

all other peptides, while for complexes with VLPFDIKYI and HVDGKILFV peptides, RMSD

distribution profile is observed to be at the rightmost.

According to the results of our RMSD analysis at 310 K and 473 K, ILDDNLYKV peptide forms

the most stable complex with the HLA-A*02:01 allele, while VLPFDIKYI and HVDGKILFV

peptides form the least stable complex with HLA-A*02:01.

38
4.2 Analysis of RMSF profiles

The RMSF figures are given in Figure 4.5 and Figure 4.6 for 310 K and 473 K, respectively.

Figure 4.5 Root mean square fluctuations (RMSF) values of the modeled complexes during the
100 ns simulations at 310 K.

For 310 K, the RMSF values are in the range of 0 to 4 Å as can be observed in Figure 4.5.

However at a higher temperature value (473 K), all the structures display higher flexibility

profiles, in the range of 1 to 14 Å as can be observed in Figure 4.6. As the temperature is

increased, the flexibility of the residues is increased as expected. The maximum flexibilities are

observed in the loop regions (residues 18, 30, 41, 56, 58, 83, 89, 106, 138, 177, 181, 196, 221,

226, 252, 256, 268, 275, 294, 318, 334, 351, 364, 374), while the minima is observed at 9, 26,

37, 48, 97, 99, 114, 158, 164, 187, 206, 242, 264, 286, 303, 315, 330, 342, 356, 371, 378.

39
Figure 4.6 Root mean square fluctuations (RMSF) values of the modeled complexes during the
100 ns simulations at 473 K.

In addition to the RMSF of the overall structure, we have also analysed the fluctuation profile of

the peptide residues as well, given in Figure 4.7. At 310 K, the residues of the peptides

demonstrate a very stable behavior as their alpha carbon atoms have rekatively small RMSF

values in the range of 1Å to 2Å. However, they display higher RMSF values (in the range of 2Å

to10Å) at 473 K as given in Figure 4.8. The residues of ILDDNLYKV, SLSAYIIRV and

GLFDFVNFV peptides have relatively smaller RMSF values with respect to other peptides at

both temperature values. These are also the top three most stable peptides that form complexes

with HLA-A*02:01 allele according to the experimental studies. The stability of these peptides is

40
reflected on their fluctuation profiles investigated in terms of RMSF profiles. The least stable

ones on the other hand, display a highly unstable profile at both temperatures.

Figure 4.7 Peptide RMSF profiles of the modeled complexes during the 100 ns simulations at
310 K.

Figure 4.8 Peptide RMSF profiles of the modeled complexes during the 100 ns simulations at
473 K.

41
4.3 Analysis of interaction energies using gRINN

In order to reveal the important interactions between the residues, pairwise residue interaction

energies at 310 K are computed via gRINN tool [36]. Total energy, posititive energy and negative

energy values obtained from gRINN are given in Table 4.1 and the correlation between

experimental stability measurements and our computational energy values are demonstrated in

Figure 4.9.

Table 4.2 Interaction energies of 12 peptide HLA-A*02:01 complexes obtained from gRINN.

Total Positive Negative Observed


Peptide
HLA-B*02:01 Energy Energy Energy Stability
Sequence
(kcal/mol) (kcal/mol) (kcal/mol) (T1/2)h
Immunogenic ILDDNLYKV -36355.856 2259.807 -38616 28

Immunogenic GLFDFVNFV -36068.013 2328.046 -38396.059 23


Immunogenic SLSAYIIRV -36113.743 2478.466 -38592.209 23
Immunogenic YLPEVISTI -36113.341 2457.059 -38570.4 19

Immunogenic LMYDIINSV -36112.99 2286.509 -38399.499 14

Immunogenic RLYDYFTRV -36152.396 2278.562 -38430.958 13


Immunogenic VLYDEFVTI -36121.993 2346.307 -38468.3 12
Immunogenic RVYEALYYV -36141.136 2355.525 -38496.661 11
Immunogenic HVDGKILFV -36028.416 2361.282 -38389.698 3
Nonimmunogenic YLYFCSSDV -36119.52 2327.923 -38447.443 10
Nonimmunogenic YLYQPCDLL -36031.463 2404.433 -38435.896 7

Nonimmunogenic VLPFDIKYI -35983.278 2233.361 -38216.639 1

42
Figure 4.9 Correlation between the experimental stability measurements and our computational
energy values obtained from gRINN.

Although an exact correlation can not be obtained among the experimental stability

measurements and the interaction energy results computed via gRINN, a general tendency can

still be captured. Especially the last three peptides (the least stable) and the first peptide (the

most stable) can be distinguished from the rest of the peptides in terms of having the highest and

the lowest total energies, respectively.

The most stable complex (ILDDNLYKV peptide-HLA-A*02:01 complex) has the highest total

energy (-36400 kcal mol), while the least stable complex (VLPFDIKYI peptide-HLA-A*02:01

complex) has the lowest total energy (-35950 kcal mol) (Figure 4.9).

43
Figure 4.10 Motion of peptides and peptide binding grooves throughput the all trajectory.
Macromolecule is shown in gray, peptide binding groove is shown in military green, peptides are
shown in colored tubes. (ILDDNLYKV: red, GLFDFVNFV: orange, SLSAYIIRV: yellow,
YLPEVISTI: purple, RLYDYFTRV: magenta, LMYDIINSV: brown, VLYDEFVTI: pink,
RVYEALYYV: green, YLYFCSSDV: cyan, YLYQPCDLL: cream, HVDGKILFV: blue,
VLPFDIKYI: black)

44
Moreover, motion of peptides and peptide binding grooves throughout the trajectories are

analysed. According to our analysis, differences in stability are also consistent with the motion

of the peptides and the peptide binding grooves (Figure 4.10). The last for peptides show a much

more flexible profile when compared with the rest. In all the others, peptides display a much

more stable profile, without much fluctuation.

Further residue based analyses are also conducted in order to gain insight about the nature of the

observed behavior. Through these extensive analyses, residues that play a key role in the stability

of the complex could be extracted. At first, the interaction energies between the residues of HLA

and β2-microglobulin are analysed and given in Figure 4.11. The important residues are also

shown on the structure in Figure 4.12.

In previous studies, the residues of HLA-A2 that interact with β2-microglobulin were stated as

PHE8, ARG48, GLN96, ALA117, ASP122, ARG202, GLU232, P235, ALA236 and GLY237 by

Hee [120].

As a general observation, according to our results, GLY119 has the highest energy value in all

peptide-HLA-A*02:01 complexes. This residue is also a close neighbooor of the interface

residues ALA117 and ASP122 that are mentioned in the previous study [120]. This residue can

be considered as the key residue of interaction between the HLA and the β2-microglobulin

sections in the complex structure. In all the peptides, without any exception, this residue displays

the highest interaction energy values as shown in Figure 4.11 and Figure 4.13.

45
Figure 4.11 Interaction energies between residues of HLA and β2-microglobulin.

46
Figure 4.12 Key residues for the interaction between
HLA and β2-microglobulin (ILDDNLYKV peptide) are
shown in PyMOL representation. Colored dots represent
the important residues interacting with β2-microglobulin
(ARG48: magenta GLY119: red, ARG202: green,
peptide: orange, macromolecule: gray )

We observe that residue ARG48 have lower interaction energy values in the least stable

complexes (YLYQPCDLL peptide-HLA-A*02:01, HVDGKILFV peptide-HLA-A*02:01 and

VLPFDIKYI peptide-HLA-A*02:01) when compared with the rest. This is further shown in

Figure 4.14, where the interaction energy of this residue is analysed alone for all the peptides.

Figure 4.13 Interaction energies of residue GLY119.

47
Additionally, residue ARG202, which is also a residue mentioned in the previous study [120] has

the lowest energy value in the least stable complex (VLPFDIKYI peptide-HLA-A*02:01

complex) as shown in Figure 4.11 and Figure 4.15.

For all the residues that interact with β2-microglobulin, there does not exist a specific difference

that can lead to the observed stability differences among the complexes of this study. It is already

recognized that the interaction between HLA and β2-microglobulin is crucial for the stability of

the peptide-HLA complexes [121][122]. Therefore, following our results, residue GLY119,

which displays an equally high interaction energy value in all the complexes, is an important

residue for HLA-A2 molecules in terms of HLA and β2-microglobulin connection. On the other

hand, residues ARG48 and ARG202 can be considered to have a role in the stability of these

complexes since they display lower interaction energy profiles for the least stable complexes.

Figure 4.14 Interaction energies of residue ARG48.

48
Figure 4.15 Interaction energies of residue ARG202.

Secondly, interaction energies between residues of HLA and peptide residues are calculated and

given in Figures 4.16 - 4.18. HLA residues that interact with peptide residues are widely known

from the previous studies [119], [123]. Accordingly, GLU63 (interacts with P1 and P2), LYS66

(interacts with P1, P2, P3 and P4), ASP77 (interacts with (P8 and P9), ARG97 (interacts with P3

and P6) and TRP146 (interacts with P9) are important sites for peptide binding of HLA-A*02:01

alleles.

Using our computational methods, we have also verified these residues and observed that

especially GLU63, LYS66, ASP77 and TRP146 are the four most important residues for stable

binding of HLA-A*02:01 alleles to their peptides (Figure 4.19). The peptide stabilities can be

dependant on the interaction energy differences observed at these sites. Figures 4.16 display the

interaction energy values of the binding groove residues with the peptide residues. Further

detailed representation of the important regions are given in Figures 4.17 and 4.18 to zoom into

the differences observed. The region between residues 60-80 and 140-170 seem to be the key

sites that needs attention.

49
Figure 4.16 Interaction energies between residues of HLA and peptide residues.

50
Figure 4.17 Interaction energies between region 1 (residues 60-80) and peptide residues.

51
Figure 4.18 Interaction energies between region 2 (residues 140-170) and peptide residues.

52
Figure 4.19 Important residues for
interaction between HLA and peptide
(ILDDNLYKV peptide) are shown in
PyMOL representation. Colored dots
represent the important residues interacting
with β2-microglobulin (GLU163: red,
LYS66: blue, TRP146: cyan, ASP77: yellow,
peptide: orange, macromolecule: gray )

Especially residues GLU63, LYS66 in the B pocket appear to have weak interactions for

YLYFCSSDV, YLYQPCDLL, HVDGKILFV and VLPFDIKYI peptides compared to the other

peptides. This is further detailed in the Figure 4.20 and Figure 4.21. The interactions of

HVDGKILFV and VLPFDIKYI peptides with residue TRP146 is relatively weaker than the

interactions of the other peptides (Figure 4.22). That might be the underlying reason of the low

stability value for VLPFDIKYI peptide-HLA-A*02:01 complex since residue TRP146 binds the

anchor residue P9, and the loss of interaction energy at this site might be the indicator of this

phenomena.

53
Figure 4.20 Interaction energies of residue GLU63.

Figure 4.21 Interaction energies of residue GLU66.

54
Figure 4.22 Interaction energies of residue TRP146.

In order to investigate the peptide residues’ point of view, the interaction energy profiles of the

peptide residues are plotted (Fig 4.23). Individually, each residue’s role on the varying peptides

can also be seen in Figures 4.24 - 4.32.

55
Figure 4.23 Interaction energy profiles of the peptide residues.

56
P1 and P9 display very high interaction energies for all the structures (Figure 4.24 and Figure

4.32). In most of the highly stable structures (top 7 out of 8), P4 has an emerging effect as well

(Figure 4.27). P9, which is an anchor residue, seem to be more effective than the other anchor

residue, P2, in terms of interaction energies due to having higher values (Figure 4.125 and figure

4.32).

Although in all the peptides, there is a high interaction energy profile observed for P1; the highly

stable peptides display a slightly higher interaction energy profile for this residue (Figure 4.24).

On the contrary, in the last four complexes in terms of stability (YLYFCSSDV peptide-HLA-

A*02:01, YLYQPCDLL peptide-HLA-A*02:01, HVDGKILFV peptide-HLA-A*02:01 and

VLPFDIKYI peptide-HLA-A*02:01), P1 has lower energy values than the rest.

Figure 4.24 Interaction energies of residue P1.

57
P2 has lower energies for the last five peptide complexes (RVYEALYYV peptide-HLA*02:01,

YLYFCSSDV peptide-HLA*02:01, YLYQPCDLL peptide-HLA-A*02:01, HVDGKILFV

peptide-HLA-A*02:01 and VLPFDIKYI peptide-HLA-A*02:01) with respect to the other

complexes (Figure 4.25). It is highly possible that these last five peptides form less stable

complexes with the HLA-A*02:01 allele due to the poor binding of anchor P2.

Figure 4.25 Interaction energies of residue P2.

P3 has higher energies in ILDDNLYKV peptide-HLA-A*02:01 and HVDGKILFV peptide-

HLA-A*02:01 complexes compared to other complexes (Figure 4.26). ILDDNLYKV and

HVDGKILFV peptides are the only ones having ASP in position P3. Hence, the observed

interaction energy might be caused due to the amino acid type.

58
Figure 4.26 Interaction energies of residue P3.

For P4, there is a higher interaction energy in almost all the highly stable peptides with an

exception of SLSAYIIRV peptide-HLA-A*02:01 complex (Figure 4.27). For the ones having

high interaction energies, there is either ASP (D) or GLU (E) at P4 location. Both of these are

negatively charged side groups. Whereas, for SLSAYIIRV, there is ALA at this position, which is

a nonpolar aminoacid. This might be the reason of exception for this peptide. But generally, it

can be concluded that having a negatively charged group (ASP or GLU) at this location has a

definite stabilizing factor, since all the peptides having these residues at this location have high

stability values.

59
For P5 to P8, there are no significant energy value differences observed (Figure 4.28 - 4.31).

Figure 4.27 Interaction energies of residue P4.

Figure 4.28 Interaction energies of residue P5.

60
Figure 4.29 Interaction energies of residue P6.

Figure 4.30 Interaction energies of residue P7.

61
Figure 4.31 Interaction energies of residue P8.

P9 has relatively lower energy values the least stable two complexes, VLPFDIKYI peptide-

HLA-A*02:01 and in HVDGKILFV peptide-HLA-A*02:01(Figure 4.32). An additonal

RVYEALYYV peptide-HLA-A*02:01 complex also has a low energy. This peptide is binding to

the HLA structure through the first peptide residue, P1, much more tightly when compared with

the rest (Figure 4.24). Hence, it the energy seems to switch to that location rather than this

section.

62
Figure 4.32 Interaction energies of residue P9.

As a final step, further detailed analysis is conducted and for each peptide, the residues involved

in interaction with each peptide residue throughout the MD simulation is calculated and shown in

Figures 4.33 and 4.34.

We have revealed all the interactions between residues through the MD simulations (Figures 4.33

and 4.34). According to our results, residue P1 can interact with 5, 7, 9, 26, 27,33, 45, 51, 52, 54,

55, 58-67, 70, 97, 99, 101, 114, 159, 160, 162-168, 170, 171, residue P2 can interact with 7, 9,

24, 25, 45, 58- 60, 62- 67, 69, 70, 97- 99, 114, 159, 163, 164, 167, 171, residue P3 can interact

with 6, 7, 9, 63, 64, 66, 68- 71, 97, 99, 113-115, 126, 133, 152, 153, 155- 160, 162- 164, residue

P4 can interact with 9, 22, 58, 61- 74, 77, 97, 99, 113-115, 126, 133, 150- 152, 154- 156, 158-

160, 165, 167, residue P5 interact with 62, 63, 65- 70, 72, 73, 77, 96, 97, 99, 112, 114, 116, 132,

133, 146, 147, 150- 156, 159, 160, residue P6 can interact with 9, 62, 63, 65-67, 69- 77, 97, 99,

114-117, 133, 147, 149-153, 155, 156, 159, residue P7 can interact with 9, 11, 22, 66- 70, 73,

63
74, 77-81, 84, 97, 99, 112, 114-116, 118, 123-125, 133, 139, 140, 142-157, residue P8 can

interact with 65, 66, 69, 70, 72- 81, 84, 97, 114- 116, 122-124, 133, 140, 142- 152, residue P9

can interact with 70, 72- 85, 96, 97, 115- 118, 122- 124, 133, 138-140- 152 .

From these heat maps, we can one more time acknowledge that in the most stable complexes,

residues P1-P4 have strong interactions with residues GLU63 and LYS66 compared to the least

stable four complexes (YLYFCSSDV peptide-HLA-A*02:01, YLYQPCDLL peptide-HLA-

A*02:01, HVDGKILFV peptide-HLA-A*02:01 and VLPFDIKYI peptide-HLA-A*02:01)

(Figure 4.33 and 4.34).

64
Figure 4.33 Interacting residues of ILDDNLYKV, GLFDFVNFV, SLSAYIIRV, YLPEVISTI,
RLYDYFTRV and LMYDIINSV peptides.

65
Figure 4.34 Interacting residues of VLYDEFVTI, RVYEALYYV, YLYFCSSDV,
YLYQPCDLL, HVDGKILFV and VLPFDIKYI peptides.

66
CHAPTER 5 CONCLUSION
In the present study, molecular docking studies were performed for 12 peptide-HLA-A*02:01

complexes by DockTope web server [33] and then 100 ns parallel MD simulations were carried

out on the docked structures using GROMACS software [34] with the OPLS-AA/L [91] all-atom

force field at 310 K and 473 K.

The RMSD and RMSF results of MD simulations demonstrate differences between the 12

different peptide-HLA-A*02:01 complexes. RMSD distributions of the complexes are inline

with the experimental measurements [10]. All modeled peptide-HLA-A*02:01 complexes show

higher fluctuations at 473 K compared to 310 K. But, three relatively most stable complexes still

remained stable throughout 100 ns at 473 K. However, two relatively less stable complexes lost

interactions with the HLA-A*02:01 allele. Therefore, high temperature MD verified the previous

experimental study revealing stability differences between these 12 peptide-HLA-A*02:01

complexes [10].

Additionally, pairwise amino acid non-bonded interaction energies at 310 K were calculated via

gRINN tool[36] to determine the critical residues which play a role in the observed difference in

the stability of these 12 peptide-HLA-A*02:01 complexes. We observed that our total energy

values correlate with the experimental stability measurements [10]. Therefore, we proved that the

stability is the most accurate parameter for peptide immunogenicity. Also, our results verified the

residues of HLA – β2-microglobulin interaction [120]. Residue GLY119, which displays an

equally high interaction energy value in all the complexes, is an important residue for HLA-A2

molecules. On the other hand, residue ARG48 and ARG202 can be considered to have a role in

the stability of these complexes since they display lower interaction energy profiles for the least

stable complexes.

67
Moreover, we analyzed the interaction energies between residues of HLA and peptide residues

and observe dramatic energy differences at residue positions GLU63 and LYS66 in the B pocket.

Our results prove that GLU63 and LYS66 interact with residues P2-P4 and are important for the

stable binding of HLA-A*02:01 alleles [123]. Moreover, these results are also inline with

previous experimental study by performed Madden [119].

Furthermore, we observed that P1 is the most critical residue for binding since it displays the

highest interaction energy values for all the peptides. Additionally, P4 is found to be important

for stable binding. In general the interaction between P4 and LYS66 seem to be crucial for the

stability. When P4 is ASP or GLU, there is a high stability observed for all the peptides because

of the nature of interaction among the amino acid types.

As a result, it can be concluded that the computational approaches used in this study provides

detailed information for the prediction of peptide-HLA-A*02:01 complex stabilities, for which

the three dimensional structures are unknown. Our methodology can be used to guide future

experiments in this field.

68
CHAPTER 6 FUTURE WORK

As future work, a large number of peptide-HLA-A*02:01 complexes can be modeled and

analyzed using our computational approaches. For peptide binding affinity prediction, new

methods can be developed. In addition, the effect of mutations on the stability of peptide HLA-

A*02:01 complexes can also be investigated.

69
REFERENCES

[1] M. Wieczorek et al., “Major Histocompatibility Complex (MHC) Class I and MHC Class
II Proteins: Conformational Plasticity in Antigen Presentation,” Front. Immunol., vol. 8, p.
292, Mar. 2017.

[2] J. B. Rothbard and M. L. Gefter, “Interactions between Immunogenic Peptides and MHC
Proteins,” Annu. Rev. Immunol., vol. 9, no. 1, pp. 527–565, Apr. 1991.

[3] P. A. van der Merwe and S. J. Davis, “M OLECULAR I NTERACTIONS M EDIATING


T C ELL A NTIGEN R ECOGNITION,” Annu. Rev. Immunol., vol. 21, no. 1, pp. 659–
684, Apr. 2003.

[4] M. Schumacher et al., “Restricted T Cell Epitopes − HLA-A2 Design through Chemically
Modified Altered Peptide Ligands Revisited: Vaccine Altered Peptide Ligands Revisited:
Vaccine Design through Chemically Modified HLA-A2–Restricted T Cell Epitopes,”
DCSupplemental.html J. Immunol. Med. Libr. Vrije Univ. Novemb. J. Immunol., vol. 193,
no. 9, pp. 4803–4813, 2014.

[5] S. Dedier, “Thermodynamic Stability of HLA-B*2705/Peptide Complexes: Effect of


Peptide and MHC Protein mutations,” J. Biol. Chem., vol. 275, no. 35, pp. 27055–27061,
2000.

[6] T. Blankenstein, P. G. Coulie, E. Gilboa, and E. M. Jaffee, “The determinants of tumour


immunogenicity,” Nat. Rev. Cancer, vol. 12, no. 4, pp. 307–313, 2012.

[7] S. N. Khilko, M. Corr, L. F. Boyd, A. Lees, J. K. Inman, and D. H. Margulies, “Direct


detection of major histocompatibility complex class I binding to antigenic peptides using
surface plasmon resonance. Peptide immobilization and characterization of binding
specificity.,” J. Biol. Chem., vol. 268, no. 21, pp. 15425–34, Jul. 1993.

[8] K. C. Parker, M. Dibrino, L. Hull, and J. E. Coligan, “The β2-microglobulin dissociation


rate is an accurate measure of the stability of MHC class I heterotrimers and depends on
which peptide is bound,” J. Immunol., vol. 149, no. 6, pp. 1896–1904, 1992.

[9] M. Harndahl, S. Justesen, K. Lamberth, G. Roder, M. Nielsen, and S. Buus, “Peptide

70
Binding to HLA Class I Molecules: Homogenous, High-Throughput Screening, and
Affinity Assays,” J. Biomol. Screen., vol. 14, no. 2, pp. 173–180, Feb. 2009.

[10] M. Harndahl et al., “Peptide-MHC class I stability is a better predictor than peptide
affinity of CTL immunogenicity,” Eur. J. Immunol., vol. 42, no. 6, pp. 1405–1416, Jun.
2012.

[11] K. C. Parker, M. DiBrino, L. Hull, and J. E. Coligan, “The beta 2-microglobulin


dissociation rate is an accurate measure of the stability of MHC class I heterotrimers and
depends on which peptide is bound.,” J. Immunol., vol. 149, no. 6, pp. 1896–904, Sep.
1992.

[12] D. H. Busch and E. G. Pamer, “MHC class I/peptide stability: implications for
immunodominance, in vitro proliferation, and diversity of responding CTL.,” J. Immunol.,
vol. 160, no. 9, pp. 4441–8, May 1998.

[13] E. Assarsson et al., “A quantitative analysis of the variables affecting the repertoire of T
cell specificities recognized after vaccinia virus infection.,” J. Immunol., vol. 178, no. 12,
pp. 7890–901, Jun. 2007.

[14] S. Dédier, S. Reinelt, S. Rion, G. Folkers, and D. Rognan, “Use of fluorescence


polarization to monitor MHC-peptide interactions in solution.,” J. Immunol. Methods, vol.
255, no. 1–2, pp. 57–66, Sep. 2001.

[15] A.-K. Binz, R. C. Rodriguez, W. E. Biddison, and B. M. Baker, “Thermodynamic and


Kinetic Analysis of a Peptide−Class I MHC Interaction Highlights the Noncovalent
Nature and Conformational Dynamics of the Class I Heterotrimer †,” Biochemistry, vol.
42, no. 17, pp. 4954–4961, May 2003.

[16] R. Buchli, R. S. VanGundy, H. D. Hickman-Miller, C. F. Giberson, W. Bardet, and W. H.


Hildebrand, “Real-Time Measurement of in Vitro Peptide Binding to Soluble HLA-
A*0201 by Fluorescence Polarization,” Biochemistry, vol. 43, no. 46, pp. 14852–14863,
Nov. 2004.

[17] a Sette et al., “The relationship between class I binding affinity and immunogenicity of
potential cytotoxic T cell epitopes.,” J. Immunol., vol. 153, no. 12, pp. 5586–92, 1994.

[18] M. E. Ressing et al., “Human CTL epitopes encoded by human papillomavirus type 16 E6
71
and E7 identified through in vivo and in vitro immunogenicity studies of HLA-A*0201-
binding peptides.,” J. Immunol., vol. 154, no. 11, pp. 5934–5943, 1995.

[19] R. R. Pogue, J. Eron, J. A. Frelinger, and M. Matsui, “Amino-terminal alteration of the


HLA-A*0201-restricted human immunodeficiency virus pol peptide increases complex
stability and in vitro immunogenicity.,” Proc. Natl. Acad. Sci. U. S. A., vol. 92, no. 18, pp.
8166–70, Aug. 1995.

[20] J. M. Brooks, R. A. Colbert, J. P. Mear, A. M. Leese, and A. B. Rickinson, “HLA-B27


subtype polymorphism and CTL epitope choice: studies with EBV peptides link
immunogenicity with stability of the B27:peptide complex.,” J. Immunol., vol. 161, no.
10, pp. 5252–9, Nov. 1998.

[21] U. M. Abdel-Motal, R. Friedline, B. Poligone, R. R. Pogue-Caley, J. A. Frelinger, and R.


Tisch, “Dendritic Cell Vaccination Induces Cross-Reactive Cytotoxic T Lymphocytes
Specific for Wild-Type and Natural Variant Human Immunodeficiency Virus Type 1
Epitopes in HLA-A*0201/Kb Transgenic Mice,” Clin. Immunol., vol. 101, no. 1, pp. 51–
58, Oct. 2001.

[22] S. Vertuani et al., “Improved immunogenicity of an immunodominant epitope of the


HER-2/neu protooncogene by alterations of MHC contact residues.,” J. Immunol., vol.
172, no. 6, pp. 3501–8, Mar. 2004.

[23] S. H. van der Burg, M. J. Visseren, R. M. Brandt, W. M. Kast, and C. J. Melief,


“Immunogenicity of peptides bound to MHC class I molecules depends on the MHC-
peptide complex stability.,” J. Immunol., vol. 156, no. 9, pp. 3308–14, May 1996.

[24] A. Stavrakoudis, “Conformational flexibility in designing peptides for immunology: the


molecular dynamics approach.,” Curr. Comput. Aided. Drug Des., vol. 6, no. 3, pp. 207–
22, 2010.

[25] C. J. Camacho, Y. Katsumata, and D. P. Ascherman, “Structural and thermodynamic


approach to peptide immunogenicity,” PLoS Comput. Biol., vol. 4, no. 11, 2008.

[26] A. Krammer et al., “Forced unfolding of the fibronectin type III module reveals a tensile
molecular recognition switch.,” Proc. Natl. Acad. Sci. U. S. A., vol. 96, no. 4, pp. 1351–6,
Feb. 1999.

72
[27] W. F. GUNSTEREN and A. E. MARK, “On the interpretation of biochemical data by
molecular dynamics computer simulation,” Eur. J. Biochem., vol. 204, no. 3, pp. 947–961,
Mar. 1992.

[28] V. Daggett and M. Levitt, “A model of the molten globule state from molecular dynamics
simulations.,” Proc. Natl. Acad. Sci. U. S. A., vol. 89, no. 11, pp. 5142–6, Jun. 1992.

[29] V. Daggett and M. Levitt, “Protein Unfolding Pathways Explored Through Molecular
Dynamics Simulations,” J. Mol. Biol., vol. 232, no. 2, pp. 600–619, Jul. 1993.

[30] M. Karplus and D. L. Weaver, “Protein folding dynamics: the diffusion-collision model
and experimental data.,” Protein Sci., vol. 3, no. 4, pp. 650–68, Apr. 1994.

[31] R. D. Sharma et al., “High temperature unfolding of a truncated hemoglobin by molecular


dynamics simulation,” J. Mol. Model., vol. 19, no. 9, pp. 3993–4002, Sep. 2013.

[32] K. Y. Chen, J. Liu, and E. C. Ren, “Structural and functional distinctiveness of HLA-A2
allelic variants,” Immunol. Res., vol. 53, no. 1–3, pp. 182–190, Sep. 2012.

[33] M. Menegatti Rigo et al., “DockTope: a Web-based tool for automated pMHC-I
modelling,” Sci. Rep., vol. 5, no. 1, p. 18413, Nov. 2016.

[34] M. J. Abraham et al., “Gromacs: High performance molecular simulations through multi-
level parallelism from laptops to supercomputers,” SoftwareX, vol. 1–2, pp. 19–25, 2015.

[35] G. A. K. and, R. A. Friesner*, J. T.-R. and, and W. L. Jorgensen, “Evaluation and


Reparametrization of the OPLS-AA Force Field for Proteins via Comparison with
Accurate Quantum Chemical Calculations on Peptides†,” 2001.

[36] O. Serçinoğlu and P. Ozbek, “gRINN: a tool for calculation of residue interaction energies
and protein energy network analysis of molecular dynamics simulations,” Nucleic Acids
Res., May 2018.

[37] J. M. Werner et al., “Major Histocompatibility Complex (MHC) Class i and MHC Class ii
Proteins: Conformational Plasticity in Antigen Presentation,” Front. Immunol, vol. 8, no.
8, pp. 2923389–292, 2017.

[38] K. Maenaka and E. Y. Jones, “MHC superfamily structure and the immune system.,”
Curr. Opin. Struct. Biol., vol. 9, no. 6, pp. 745–53, Dec. 1999.

73
[39] M. Wieczorek et al., “Major Histocompatibility Complex (MHC) Class I and MHC Class
II Proteins: Conformational Plasticity in Antigen Presentation,” Front. Immunol., vol. 8, p.
292, Mar. 2017.

[40] C. A. Janeway, P. Travers, M. Walport, and M. Shlomchik, “Immunobiology,” p. 884,


2001.

[41] D. R. Madden, “The Three-Dimensional Structure of Peptide-MHC Complexes,” Annu.


Rev. Immunol., vol. 13, no. 1, pp. 587–622, Apr. 1995.

[42] E. M. Lafuente and P. a Reche, “Prediction of MHC-peptide binding: a systematic and


comprehensive overview.,” Curr. Pharm. Des., vol. 15, no. 28, pp. 3209–20, 2009.

[43] R. T. Kubo et al., “Definition of specific peptide motifs for four major HLA-A alleles.,” J.
Immunol., vol. 152, no. 8, pp. 3913–24, Apr. 1994.

[44] Q. J. Zhang, R. Gavioli, G. Klein, and M. G. Masucci, “An HLA-A11-specific motif in


nonamer peptides derived from viral and cellular proteins.,” Proc. Natl. Acad. Sci. U. S.
A., vol. 90, no. 6, pp. 2217–21, Mar. 1993.

[45] K. Falk, O. Rötzschke, S. Stevanovié, G. Jung, and H.-G. Rammensee, “Allele-specific


motifs revealed by sequencing of self-peptides eluted from MHC molecules,” Nature, vol.
351, no. 6324, pp. 290–296, May 1991.

[46] M. Dibrino et al., “Endogenous peptides bound to HLA-A3 possess a specific


combination of anchor residues that permit identification of potential antigenic peptides
(major histocompatibility complex class I/I32-microglobulin/peptide binding motif),”
Immunology, vol. 90, pp. 1508–1512, 1993.

[47] H.-C. Guo, T. S. Jardetzky, T. P. J. Garrettt, W. S. Lane, J. L. Strominger, and D. C.


Wiley, “Different length peptides bind to HLA-Aw68 similarly at their ends but bulge out
in the middle,” Nature, vol. 360, no. 6402, pp. 364–366, Nov. 1992.

[48] T. S. Jardetzky, W. S. Lane, R. A. Robinson, D. R. Madden, and D. C. Wiley,


“Identification of self peptides bound to purified HLA-B27,” Nature, vol. 353, no. 6342,
pp. 326–329, Sep. 1991.

[49] J. Sutton et al., “A sequence pattern for peptides presented to cytotoxic T lymphocytes by

74
HLA B8 revealed by analysis of epitopes and eluted peptides,” Eur. J. Immunol., vol. 23,
no. 2, pp. 447–453, Feb. 1993.

[50] P. J. Bjorkman, M. A. Saper, B. Samraoui, W. S. Bennett, J. L. Strominger, and D. C.


Wiley, “Structure of the human class I histocompatibility antigen, HLA-A2,” Nature, vol.
329, no. 6139, pp. 506–512, Oct. 1987.

[51] M. A. Saper, P. J. Bjorkman, and D. C. Wiley, “Refined structure of the human


histocompatibility antigen HLA-A2 at 2.6 A resolution.,” J. Mol. Biol., vol. 219, no. 2, pp.
277–319, May 1991.

[52] D.R.Madden, “The structure of HLA-B27 reveals nonamer self-peptides bound in an


extended conformation,” Nature, vol. 354, pp. 56–58, 1991.

[53] A. Patronov and I. Doytchinova, “T-cell epitope vaccine design by immunoinformatics.,”


Open Biol., vol. 3, no. 1, p. 120139, 2013.

[54] H. Rammensee, J. Bachmann, N. P. Emmerich, O. A. Bachor, and S. Stevanović,


“SYFPEITHI: database for MHC ligands and peptide motifs.,” Immunogenetics, vol. 50,
no. 3–4, pp. 213–9, Nov. 1999.

[55] A. S. DE GROOT, B. M. JESDALE, E. SZU, J. R. SCHAFER, R. M. CHICZ, and G.


DEOCAMPO, “An Interactive Web Site Providing Major Histocompatibility Ligand
Predictions: Application to HIV Research,” AIDS Res. Hum. Retroviruses, vol. 13, no. 7,
pp. 529–531, May 1997.

[56] H. P. Adams and J. A. Koziol, “Prediction of binding to MHC class I molecules.,” J.


Immunol. Methods, vol. 185, no. 2, pp. 181–90, Sep. 1995.

[57] K. Gulukota, J. Sidney, A. Sette, and C. DeLisi, “Two complementary methods for
predicting peptides binding major histocompatibility complex molecules,” J. Mol. Biol.,
vol. 267, no. 5, pp. 1258–1267, Apr. 1997.

[58] P. Dönnes and A. Elofsson, “Prediction of MHC class I binding peptides, using
SVMHC.,” BMC Bioinformatics, vol. 3, p. 25, Sep. 2002.

[59] C.-W. Tung and S.-Y. Ho, “POPI: predicting immunogenicity of MHC class I binding
peptides by mining informative physicochemical properties,” Bioinformatics, vol. 23, no.

75
8, pp. 942–949, Apr. 2007.

[60] H. Mamitsuka, “Predicting peptides that bind to MHC molecules using supervised
learning of hidden Markov models.,” Proteins, vol. 33, no. 4, pp. 460–74, Dec. 1998.

[61] G. L. Zhang, A. M. Khan, K. N. Srinivasan, J. T. August, and V. Brusic, “MULTIPRED:


a computational system for prediction of promiscuous HLA binding peptides,” Nucleic
Acids Res., vol. 33, no. Web Server, pp. W172–W179, Jul. 2005.

[62] K. C. Parker, M. A. Bednarek, and J. E. Coligan, “Scheme for ranking potential HLA-A2
binding peptides based on independent binding of individual peptide side-chains.,” J.
Immunol., vol. 152, no. 1, pp. 163–75, Jan. 1994.

[63] B. Peters, W. Tong, J. Sidney, A. Sette, and Z. Weng, “Examining the independent
binding assumption for binding of peptide epitopes to MHC-I molecules,” Bioinformatics,
vol. 19, no. 14, pp. 1765–1772, Sep. 2003.

[64] R. Rosenfeld, Q. Zheng, S. Vajda, and C. DeLisi, “Flexible docking of peptides to class I
major-histocompatibility-complex receptors.,” Genet. Anal., vol. 12, no. 1, pp. 1–21, Mar.
1995.

[65] B. Zhao, V. S. Mathura, G. Rajaseger, S. Moochhala, M. K. Sakharkar, and P. Kangueane,


“A novel MHCp binding prediction model.,” Hum. Immunol., vol. 64, no. 12, pp. 1123–
43, Dec. 2003.

[66] O. Schueler-Furman, Y. Altuvia, A. Sette, and H. Margalit, “Structure-based prediction of


binding peptides to MHC class I molecules: Application to a broad range of MHC
alleles,” Protein Sci., vol. 9, no. 9, pp. 1838–1846, Sep. 2000.

[67] D. Rognan, “Molecular dynamics simulations: a tool for drug design,” Perspect. Drug
Discov. Des., vol. 9/11, no. 0, pp. 181–209, 1998.

[68] M. Nielsen et al., “NetMHCpan, a method for quantitative predictions of peptide binding
to any HLA-A and -B locus protein of known sequence.,” PLoS One, vol. 2, no. 8, p.
e796, Aug. 2007.

[69] I. Hoof et al., “NetMHCpan, a method for MHC class I binding prediction beyond
humans.,” Immunogenetics, vol. 61, no. 1, pp. 1–13, Jan. 2009.

76
[70] E. Karosiene, C. Lundegaard, O. Lund, and M. Nielsen, “NetMHCcons: a consensus
method for the major histocompatibility complex class I predictions,” Immunogenetics,
vol. 64, no. 3, pp. 177–186, Mar. 2012.

[71] C. Lundegaard, K. Lamberth, M. Harndahl, S. Buus, O. Lund, and M. Nielsen, “NetMHC-


3.0: accurate web accessible predictions of human, mouse and monkey MHC class I
affinities for peptides of length 8-11.,” Nucleic Acids Res., vol. 36, no. Web Server issue,
pp. W509-12, Jul. 2008.

[72] K. W. Jørgensen, M. Rasmussen, S. Buus, and M. Nielsen, “NetMHCstab - predicting


stability of peptide-MHC-I complexes; impacts for cytotoxic T lymphocyte epitope
discovery.,” Immunology, vol. 141, no. 1, pp. 18–26, Jan. 2014.

[73] M. Rasmussen et al., “Pan-Specific Prediction of Peptide-MHC Class I Complex Stability,


a Correlate of T Cell Immunogenicity.,” J. Immunol., vol. 197, no. 4, pp. 1517–24, 2016.

[74] W. W. P. Liao and J. W. Arthur, “Predicting peptide binding to Major Histocompatibility


Complex molecules,” Autoimmun. Rev., vol. 10, no. 8, pp. 469–473, Jun. 2011.

[75] N. Jojic, M. Reyes-Gomez, D. Heckerman, C. Kadie, and O. Schueler-Furman, “Learning


MHC I--peptide binding,” Bioinformatics, vol. 22, no. 14, pp. e227–e235, Jul. 2006.

[76] B. Knapp and B. Knapp, “Structural Bioinformatics and Immuoinformatics in the T Cell
Receptor / Peptide / Major Histocompatibility Complex Interface,” pp. 1–7, 2011.

[77] P. L. Kastritis and A. M. J. J. Bonvin, “On the binding affinity of macromolecular


interactions: daring to ask why proteins interact.,” J. R. Soc. Interface, vol. 10, no. 79, p.
20120835, Feb. 2013.

[78] J. M. Sanchez-Ruiz, “Protein kinetic stability,” Biophys. Chem., vol. 148, no. 1–3, pp. 1–
15, May 2010.

[79] M. Harndahl, M. Rasmussen, G. Roder, and S. Buus, “Real-time, high-throughput


measurements of peptide–MHC-I dissociation using a scintillation proximity assay,” J.
Immunol. Methods, vol. 374, no. 1–2, pp. 5–12, Nov. 2011.

[80] P. Parham, C. J. Barnstable, and W. F. Bodmer, “Use of a monoclonal antibody (W6/32)


in structural studies of HLA-A,B,C, antigens.,” J. Immunol., vol. 123, no. 1, pp. 342–9,

77
Jul. 1979.

[81] M. K. Gilson and H.-X. Zhou, “Calculation of Protein-Ligand Binding Affinities *,”
Annu. Rev. Biophys. Biomol. Struct, vol. 36, pp. 21–42, 2007.

[82] J. Aqvist, C. Medina, and J. E. Samuelsson, “A new method for predicting binding affinity
in computer-aided drug design.,” Protein Eng., vol. 7, no. 3, pp. 385–91, Mar. 1994.

[83] P. A. Kollman et al., “Calculating structures and free energies of complex molecules:
combining molecular mechanics and continuum models.,” Acc. Chem. Res., vol. 33, no.
12, pp. 889–97, Dec. 2000.

[84] D. L. Beveridge and F. M. DiCapua, “Free Energy Via Molecular Simulation:


Applications to Chemical and Biomolecular Systems,” Annu. Rev. Biophys. Biophys.
Chem., vol. 18, no. 1, pp. 431–492, Jun. 1989.

[85] S. a. Adcock and J. A. McCammon, “Molecular dynamics: Survey of methods for


simulating the activity of proteins,” Chem. Rev., vol. 106, no. 5, pp. 1589–1615, 2006.

[86] M. Zacharias and S. Springer, “Conformational flexibility of the MHC class I alpha1-
alpha2 domain in peptide bound and free states: a molecular dynamics simulation study.,”
Biophys. J., vol. 87, no. 4, pp. 2203–14, Oct. 2004.

[87] B. J. Alder and T. E. Wainwright, “Phase Transition for a Hard Sphere System,” J. Chem.
Phys., vol. 27, no. 5, pp. 1208–1209, Nov. 1957.

[88] A. Rahman, “Correlations in the Motion of Atoms in Liquid Argon,” Phys. Rev., vol. 136,
no. 2A, pp. A405–A411, Oct. 1964.

[89] J. A. McCammon, B. R. Gelin, and M. Karplus, “Dynamics of folded proteins,” Nature,


vol. 267, no. 5612, pp. 585–590, Jun. 1977.

[90] J. Meller, “Molecular Dynamics,” in Encyclopedia of Life Sciences, Chichester: John


Wiley & Sons, Ltd, 2001.

[91] W. L. Jorgensen and J. Tirado-Rives, “The OPLS [optimized potentials for liquid
simulations] potential functions for proteins, energy minimizations for crystals of cyclic
peptides and crambin,” J. Am. Chem. Soc., vol. 110, no. 6, pp. 1657–1666, Mar. 1988.

[92] D. Rognan, N. Zimmermann, G. Jung, and G. Folkers, “Molecular dynamics study of a

78
complex between the human histocompatibility antigen HLA-A2 and the IMP58-66
nonapeptide from influenza virus matrix protein.,” Eur. J. Biochem., vol. 208, no. 1, pp.
101–13, Aug. 1992.

[93] D. A. Pearlman et al., “AMBER, a package of computer programs for applying molecular
mechanics, normal mode analysis, molecular dynamics and free energy calculations to
simulate the structural and energetic properties of molecules,” Comput. Phys. Commun.,
vol. 91, no. 1–3, pp. 1–41, Sep. 1995.

[94] H. Toh, C. J. Savoie, N. Kamikawaji, S. Muta, T. Sasazuki, and S. Kuhara, “Changes at


the floor of the peptide-binding groove induce a strong preference for Proline at position 3
of the bound peptide: Molecular dynamics simulations of HLA-A*0217,” Biopolymers,
vol. 54, no. 5, pp. 318–327, Oct. 2000.

[95] M. A. Joseph et al., “Secondary anchor substitutions in an HLA-A*0201-restricted T-cell


epitope derived from Her-2/neu.,” Mol. Immunol., vol. 44, no. 4, pp. 322–31, Jan. 2007.

[96] T. Pöhlmann, R. A. Böckmann, H. Grubmüller, B. Uchanska-Ziegler, A. Ziegler, and U.


Alexiev, “Differential Peptide Dynamics Is Linked to Major Histocompatibility Complex
Polymorphism,” J. Biol. Chem., vol. 279, no. 27, pp. 28197–28201, Jul. 2004.

[97] H. Fabian et al., “HLA-B27 Subtypes Differentially Associated with Disease Exhibit
Conformational Differences in Solution,” J. Mol. Biol., vol. 376, no. 3, pp. 798–810, Feb.
2008.

[98] D. Narzi, C. M. Becker, M. T. Fiorillo, B. Uchanska-Ziegler, A. Ziegler, and R. A.


Böckmann, “Dynamical Characterization of Two Differentially Disease Associated MHC
Class I Proteins in Complex with Viral and Self-Peptides,” J. Mol. Biol., vol. 415, no. 2,
pp. 429–442, Jan. 2012.

[99] E. T. Abualrous et al., “F pocket flexibility influences the tapasin dependence of two
differentially disease-associated MHC Class I proteins,” Eur. J. Immunol., vol. 45, no. 4,
pp. 1248–1257, Apr. 2015.

[100] P. Ozbek, “Dynamic characterization of HLA-B∗44 Alleles: A comparative molecular


dynamics simulation study,” Comput. Biol. Chem., vol. 62, pp. 12–16, 2016.

[101] O. Serçinoğlu and P. Ozbek, “Computational characterization of residue couplings and


79
micropolymorphism-induced changes in the dynamics of two differentially disease-
associated human MHC class-I alleles,” J. Biomol. Struct. Dyn., vol. 36, no. 3, pp. 724–
740, Feb. 2018.

[102] C. M. Ayres, T. P. Riley, S. A. Corcelli, and B. M. Baker, “Modeling Sequence-


Dependent Peptide Fluctuations in Immunologic Recognition,” J. Chem. Inf. Model., vol.
57, no. 8, pp. 1990–1998, Aug. 2017.

[103] H. M. Berman et al., “The Protein Data Bank.,” Nucleic Acids Res., vol. 28, no. 1, pp.
235–42, Jan. 2000.

[104] S.-Y. Huang, “Search strategies and evaluation in protein–protein docking: principles,
advances and challenges,” Drug Discov. Today, vol. 19, no. 8, pp. 1081–1096, Aug. 2014.

[105] S. F. Sousa et al., “Protein-ligand docking in the new millennium--a retrospective of 10


years in the field.,” Curr. Med. Chem., vol. 20, no. 18, pp. 2296–314, 2013.

[106] N. London, B. Raveh, E. Cohen, G. Fathi, and O. Schueler-Furman, “Rosetta


FlexPepDock web server--high resolution modeling of peptide-protein interactions.,”
Nucleic Acids Res., vol. 39, no. Web Server issue, pp. W249-53, Jul. 2011.

[107] D. L. Mobley and K. A. Dill, “Binding of small-molecule ligands to proteins: ‘what you
see’ is not always ‘what you get’.,” Structure, vol. 17, no. 4, pp. 489–98, Apr. 2009.

[108] F. Österberg, G. M. Morris, M. F. Sanner, A. J. Olson, and D. S. Goodsell, “Automated


docking to multiple target structures: Incorporation of protein mobility and structural
water heterogeneity in AutoDock,” Proteins Struct. Funct. Bioinforma., vol. 46, no. 1, pp.
34–40, Jan. 2002.

[109] M. P. Repasky, M. Shelley, and R. A. Friesner, “Flexible Ligand Docking with Glide,” in
Current Protocols in Bioinformatics, vol. Chapter 8, Hoboken, NJ, USA: John Wiley &
Sons, Inc., 2007, p. Unit 8.12.

[110] G. Jones, P. Willett, R. C. Glen, A. R. Leach, and R. Taylor, “Development and validation
of a genetic algorithm for flexible docking 1 1Edited by F. E. Cohen,” J. Mol. Biol., vol.
267, no. 3, pp. 727–748, Apr. 1997.

[111] C. Dominguez, R. Boelens, and A. M. J. J. Bonvin, “HADDOCK: a protein-protein

80
docking approach based on biochemical or biophysical information.,” J. Am. Chem. Soc.,
vol. 125, no. 7, pp. 1731–7, Feb. 2003.

[112] A. Dhanik, J. S. McMurray, and L. E. Kavraki, “DINC: a new AutoDock-based protocol


for docking large ligands.,” BMC Struct. Biol., vol. 13 Suppl 1, no. Suppl 1, p. S11, 2013.

[113] T. Liu et al., “Subangstrom accuracy in pHLA-I modeling by Rosetta FlexPepDock


refinement protocol.,” J. Chem. Inf. Model., vol. 54, no. 8, pp. 2233–42, Aug. 2014.

[114] D. A. Antunes, D. Devaurs, M. Moll, G. Lizée, and L. E. Kavraki, “General Prediction of


Peptide-MHC Binding Modes Using Incremental Docking: A Proof of Concept,” Sci.
Rep., vol. 8, no. 1, p. 4327, Mar. 2018.

[115] “The PyMol Molecular Graphics System, Version #, Schrodinger, LLC.” .

[116] O. Trott and A. J. Olson, “AutoDock Vina: improving the speed and accuracy of docking
with a new scoring function, efficient optimization, and multithreading.,” J. Comput.
Chem., vol. 31, no. 2, pp. 455–61, Jan. 2010.

[117] M. J. Abraham et al., “GROMACS: High performance molecular simulations through


multi-level parallelism from laptops to supercomputers,” SoftwareX, vol. 1, pp. 19–25,
2015.

[118] E. Assarsson et al., “A quantitative analysis of the variables affecting the repertoire of T
cell specificities recognized after vaccinia virus infection.,” J. Immunol., vol. 178, no. 12,
pp. 7890–901, Jun. 2007.

[119] D. R. Madden, D. N. Garboczi, and D. C. Wiley, “The antigenic identity of peptide-MHC


complexes: a comparison of the conformations of five viral peptides presented by HLA-
A2.,” Cell, vol. 75, no. 4, pp. 693–708, Nov. 1993.

[120] C.-S. Hee et al., “Dynamics of free versus complexed β2-microglobulin and the evolution
of interfaces in MHC class I molecules,” Immunogenetics, vol. 65, no. 3, pp. 157–172,
Mar. 2013.

[121] L. O. PEDERSEN, A. S. HANSEN, A. C. OLSEN, J. GERWIEN, M. H. NISSEN, and S.


BUUS, “The Interaction between Beta 2-Microglobulin (ssm) and Purified Class-I Major
Histocompatibility (MHC) Antigen,” Scand. J. Immunol., vol. 39, no. 1, pp. 64–72, Jan.

81
1994.

[122] M. Shiroishi et al., “Structural basis for recognition of the nonclassical MHC molecule
HLA-G by the leukocyte Ig-like receptor B2 (LILRB2/LIR2/ILT4/CD85d).,” Proc. Natl.
Acad. Sci. U. S. A., vol. 103, no. 44, pp. 16412–7, Oct. 2006.

[123] A. Simon, Z. s Dosztányi, E. Rajnavölgyi, and I. Simon, “Function-related regulation of


the stability of MHC proteins.,” Biophys. J., vol. 79, no. 5, pp. 2305–13, Nov. 2000.

82
AUTOBIOGRAPHY

Name-Surname : Asuman BUNSUZ

Birthplace and Date : İstanbul, 07/11/1988

Foreign Language : English

E-Mail : asumanbunsuz@gmail.com

Education Statue

University /
Degree Section / Program Graduation Year
Highschool

Kadir Has Anatolian


High School Science 2007
High School

Bachelor of Molecular Biology and Gebze Institute of


2014
Science Genetics Technology

Scientific Works

Conference Publications:

1. Bunsuz, A., Sercinoglu, O., Ozbek, P, Effect of temperature on the molecular dynamics
simulations of HLA-A*02 alleles, International Symposium on Chemistry via
Computation Applications on Molecular Nanoscience, Istanbul,Turkey, 30 October 2017.
(Poster Presentation)
2. Bunsuz, A., Sercinoglu, O., Ozbek, P, Immunogenic Peptide - HLA-A*02:01 Complexes
Studied by Comparative Molecular Dynamics Simulation, 5th International BAU-Drug
Design Congress, İstanbul, Turkey, 19-21 October 2017. (Poster Presentation)
3. Bunsuz, A., Sercinoglu, O., Ozbek, P, Investigation of peptide binding affinity and
thermal stability of Human Leukocyte Antigens (HLAs), 19th IUPAB and 11th EBSA
83
Congress, Edinburgh, UK, 16-20 July 2017. (Poster Presentation)
4. Bunsuz, A., Sercinoglu, O., Ozbek, P, Investigation of peptide binding affinity and
complex stability of Human Leukocyte Antigens (HLAs), 4th International BAU-Drug
Design Congress, İstanbul, Turkey, 12-15 October 2016. (Poster Presentation)
5. Bunsuz, A., Sercinoglu, O., Ozbek, P, HLA-B*44 Alellerinin Peptid Bağlanma Davranış
Mekanizmalarının Hesaplamalı Olarak araştırılması, Biruni Üniversitesi Bilgisayar
Destekli İlaç Tasarımı, İstanbul, Türkiye, 16-17 Mayıs 2016 (Poster Award)
6. Bunsuz A., Sercinoglu, O., Ozbek, P., Computational Study on the Binding Behaviour of
HLA-B44 Alleles, 3rd International BAU-Drug Design Congress, Istanbul, Turkey,
October 1-3, 2015. (Poster Presentation)

Project Assignment:

1. Researcher , " Dynamic Characterization of HLA Proteins" , TUBITAK-113M293

84

You might also like