130
Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a Model Thesis submitted for the degree "Doctor of Philosophy" by Assaf Ganoth Submitted to the Senate of Tel-Aviv University July 2006

Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

Computer Simulations of

Protein-Peptide Complexes Using the

Myosin Light Chain Protein as a Model

Thesis submitted for the degree

"Doctor of Philosophy"

by

Assaf Ganoth

Submitted to the Senate of Tel-Aviv University

July 2006

Page 2: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

II

This work was carried out under the supervision of

Prof. Menachem Gutman

Page 3: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

III

ACKNOWLEDGMENTS

It has been a long journey. During the journey, I was fortunate to come in touch

with a variety of people who helped me to pursuit my dreams.

It is my pleasure to thank them.

Prof. Hemi Gutman,

For guidance, unconditional support, brainstorm sessions and, most important,

believing in me. His special personality and profound knowledge accompanied me

throughout the research. I doubt I will ever be able to convey my appreciation fully,

but I owe him my scientific career.

Dr. Eti Nachliel,

For continuous help, professional advices, inspirational ideas, and endless assistance.

Ran Friedman,

For teaching me Molecular Dynamics, expert knowledge, being for me whenever I

encountered scientific problems or difficulties, and for bicycle trips.

Elad Project,

For assistance with Molecular Dynamics simulations, programming skills, and

unlimited help with solving computer problems.

Dr. Dani Canaani, Dr. Yossi Tsfadia and Prof. Rimona Margalit,

For professional and moral support during the last years. Each provided unique

insights and challenged my thinking, substantially contributing to a successful

completion of this thesis.

Finally, I would like to express my deepest gratitude to my friends in the lab, past and

present, for close collaboration: Anna Seltzer, Assaf Amitay, Cintia Sbarsky, Dana

Baron, Eran Bosis, Tom Mark, Aviv Mezer, Limor Radozkowicz, and Yael Rabin.

Page 4: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

IV

"Two roads diverged in a wood, and I-

- I took the one less traveled by,

And that has made all the difference"

Robert Lee Frost (1874-1963)

Page 5: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

V

TABLE OF CONTENTS

ABSTRACT IX

LIST OF ABBREVIATIONS XII

I. INTRODUCTION

1. Biological and Biochemical Background................................................................ 1

1.1 The Myosin Family......................................................................................... 1

1.2 The Myosin V Protein..................................................................................... 2

1.3 The Mlc1p Protein.......................................................................................... 5

1.4 IQ Motif Peptides............................................................................................ 6

1.5 The Calmodulin Protein.................................................................................. 7

1.6 Calmodulin-Peptide Complexes..................................................................... 8

1.7 Mlc1p-IQ Complexes...................................................................................... 9

2. Computer Simulations............................................................................................. 9

2.1 Molecular Dynamics....................................................................................... 10

2.1.1 General Overview.................................................................................. 10

2.1.2 Methodology.......................................................................................... 11

2.1.3 Molecular Dynamics of Calmodulin...................................................... 13

2.1.4 Molecular Dynamics of Calmodulin's Complexes................................. 14

2.2 Free Energy..................................................................................................... 14

2.2.1 Free Energy of Interaction..................................................................... 15

2.2.2 The MM-PBSA Approach..................................................................... 15

3. Significance of Study.............................................................................................. 16

Page 6: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

VI

II. METHODS

1. MD Simulations...................................................................................................... 18

1.1 Simulations of Protein-Peptide Complexes.................................................... 18

1.1.1 Mlc1p-IQ2 Complex.............................................................................. 18

1.1.2 Mlc1p-IQ4 Complex.............................................................................. 19

1.2 Simulations of Peptides................................................................................... 19

1.2.1 IQ2 Peptide............................................................................................ 19

1.2.2 IQ4 Peptide............................................................................................ 20

2. Visual Presentations................................................................................................ 20

3. Inter-Helical Angles................................................................................................ 21

4. Dihedral Angle Calculations................................................................................... 21

5. The Electrostatic Potential Around the Peptides..................................................... 21

6. The Protein-Peptide Interaction Free Energies....................................................... 22

6.1 Molecular Mechanics Calculations................................................................. 23

6.2 Polar Solvation Calculations........................................................................... 23

6.3 Nonpolar Solvation Calculation...................................................................... 24

6.4 Entropy Calculations....................................................................................... 25

III. RESULTS & DISCUSSION

1. Simulations of Protein-Peptide Complexes............................................................ 26

1.1 Mlc1p-IQ2 Complex....................................................................................... 26

1.1.1 The Crystallographic and Simulated Structures..................................... 26

1.1.2 The Dynamics of the Protein-Peptide Complexes................................. 27

1.1.3 Summary................................................................................................ 29

Page 7: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

VII

1.2 Mlc1p-IQ4 Complex....................................................................................... 30

1.2.1 Overall Conformational Changes During the Simulation...................... 30

1.2.2 Relative Rotation of the Helices During the Simulation........................ 34

1.2.3 The Structural Characteristics of the Compaction Event....................... 37

1.2.4 The Forces that Stabilize the Refolded Model Structure of the

Complex.................................................................................................

39

1.2.5 The Interactions Between Residues During the Structural Change of

the Complex...........................................................................................

44

1.2.6 MD Simulation of the Mlc1p-IQ4 Complex at 400 K........................... 47

1.2.7 Summary................................................................................................ 49

2. Comparison Between the Mlc1p-IQ2 and the Mlc1p-IQ4 Complexes................... 52

2.1 Crystallographic and Simulated Structures Comparison................................ 52

2.2 Structural Evolution of the Simulated Mlc1p Protein at the Protein-Peptide

Complexes.......................................................................................................

54

2.3 The Root Mean Square Fluctuation (RMSF) of the Mlc1p Protein at the

Protein-Peptide Complexes.............................................................................

57

2.4 The Electrostatic Field Around the IQ Peptides............................................. 60

2.5 The Protein-Peptide Interaction Free Energies............................................... 62

2.6 The Contacts Between the Protein and the Peptides....................................... 67

2.7 Summary......................................................................................................... 70

3. Simulations of Free IQ Peptides............................................................................. 72

3.1 Synopsis of the Presented Simulations........................................................... 72

3.2 Overall Conformational Changes During the Simulations............................. 73

3.3 Structural Characteristics of the Refolding Process of the IQ Peptides.......... 78

3.4 Secondary Structures of the IQ Peptides......................................................... 81

Page 8: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

VIII

3.5 Salt Bridges Analysis of the IQ4 Peptide Simulation..................................... 84

3.6 Dynamics of the IQ4 Peptide at Different Salt Concentrations...................... 86

3.7 Summary......................................................................................................... 88

4. The Structure of the Light Chain-Binding Domain (LCBD) of Myosin V............ 89

4.1 The Current Structural Model......................................................................... 89

4.2 Structure and Dynamics of an Extended IQ4 Peptide..................................... 92

4.3 The Suggested Solution Model....................................................................... 94

4.4 Summary......................................................................................................... 96

IV. OVERALL GENERAL DISCUSSION................................................ 98

V. SUPPLEMENT............................................................................................... 102

1.1 Articles............................................................................................................ 102

(I) Ganoth, A., E. Nachliel, R. Friedman, and M. Gutman. 2006.

Molecular dynamics study of a calmodulin-like protein with an IQ peptide:

Spontaneous refolding of the protein around the peptide. Proteins 64: 133-146

(II) Ganoth, A., R. Friedman, E. Nachliel, and M. Gutman. 2006. A

molecular dynamics study and free energy analysis of complexes between the

Mlc1p protein and two IQ motif peptides. Biophys J 91:2436-2450

REFERENCES..................................................................................................... 103

Page 9: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

IX

ABSTRACT

The Mlc1p protein, which is a member of the Calmodulin family, is an

essential component of the mechano-chemical myosin system. At the budding yeast

Saccharomyces cerevisiae, six Mlc1p proteins bind the neck domain of myosin V,

composed of six IQ motif peptides. The structures of a few Mlc1p complexes with IQ

peptides had been resolved by X-ray crystallography. The present thesis expands our

knowledge, by investigating, through molecular dynamics simulations and subsequent

calculations, of the structures, dynamics and energetics of Mlc1p-IQ complexes.

Besides grossly extending the structural data embedded in the crystalline structures,

our study provides atomistic understanding about the movement mode of myosin V

over the actin filament. We found out how bending of a specific site, located on the

neck of the myosin, can serve as the flexible joint along the myosin, securing a proper

function of its stroke lever arm.

The thesis includes four main chapters: (I) Molecular dynamics simulations of

the IQ2 and the IQ4 peptides in a complex with the Mlc1p protein; (II) A detailed

comparison between the simulations of both Mlc1p-IQ complexes; (III) Molecular

dynamics simulations of free IQ peptides; (IV) Reevaluation of the structure of the

light chain-binding domain of myosin V.

In the first chapter of the thesis, we present molecular dynamics simulations of

the Mlc1p-IQ2 and the Mlc1p-IQ4 complexes, following their relaxation in a

physiological salt solution. The Mlc1p-IQ2 complex relaxed without loosing its main

packing features, exhibiting a limited conformational change throughout the

simulation. The other complex, with the IQ4 peptide, experienced a major refolding

process, where the protein transformed its conformation from an extended to a

compact one, and the peptide was snapped into two sections. In the second chapter of

Page 10: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

X

the thesis, we offer a detailed comparative study between the molecular dynamics

simulations of both complexes. We performed a comprehensive comparison between

the structure and the dynamics of the Mlc1p-IQ complexes and an extensive

interaction free energy analysis. The latter includes an analysis of the various forces

operating on the protein-peptide complexes by indicating their specific contributions.

In the third chapter of the thesis, molecular dynamics simulations of free IQ peptides

at different conditions for various durations are given. Variations between the

conformations of the bound and the free peptides, and differences between the

configurations that the two IQ peptides make take in solution, are discussed. The

results of the molecular dynamics simulations of the protein-peptide complexes and

the free peptides enable to assess the structural model of the light chain-binding

domain of myosin V, being composed of six complexes between light chain proteins

(such as the Mlc1p) and IQ peptides, presented in the fourth chapter of the thesis. At

the core of our suggested model stands the notion that the light chain-binding domain

is a dynamic cellular entity, and hence the proposed model incorporates the ability of

the Mlc1p protein and the IQ peptides to flex and curve in a mutual manner.

The results of the thesis were summed up in three articles. The first article,

which describes the molecular dynamics simulation of the Mlc1p-IQ4 complex, was

recently published (Ganoth, A., E. Nachliel, R. Friedman, and M. Gutman. 2006.

Molecular dynamics study of a calmodulin-like protein with an IQ peptide:

Spontaneous refolding of the protein around the peptide. Proteins 64:133-146). The

second paper presents a comprehensive comparison between the simulations of the

Mlc1p-IQ4 and the Mlc1p-IQ2 complexes, including an energetic analysis and a

reevaluation of the structure of the light chain-binding domain of myosin V (Ganoth,

A., R. Friedman, E. Nachliel, and M. Gutman. 2006. A molecular dynamics study and

Page 11: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

XI

free energy analysis of complexes between the Mlc1p protein and two IQ motif

peptides. Biophys J 91:2436-2450 (Cover article)); A third paper, dealing with

simulations of the free IQ peptides, is in preparation.

The current research is a novel MD study of IQ peptides in a complex with a

CaM-like protein and at its absence. In view of the major role taken by the IQ

peptides in mechano-chemical processes, the thesis provides a comprehensive

description at atomic resolution for their interaction with CaM and CaM-like proteins.

The study, which portrays a computerized journey, starting from molecular dynamics

simulations and ending at physiological insights concerning myosin V, exemplifies

that cooperation between crystallographers and biophysicists may contribute to a

better understanding of structure-function relationship.

Page 12: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

XII

LIST OF ABBREVIATIONS

APBS Adaptive Poisson-Boltzmann Solver

Apo-CaM Ca+2

-deprived CaM

ATP Adenosine Triphosphate

CaM Calmodulin

CL C-lobe

FIONA Fluorescence Imaging with One-Nanometer Accuracy

GTP Guanosine Triphosphate

Holo-CaM Ca+2

-bound CaM

ID Inter-domain

LCBD Light Chain-Binding Domain

LJ Lennard-Jones

MD Molecular Dynamics

MM-PBSA Molecular Mechanics - Poisson-Boltzmann Surface Area

mRNA messenger RNA

NL N-lobe

NMR Nuclear Magnetic Resonance

PB Poisson-Boltzmann

PDB Protein Data Bank

PME Particle Mesh Ewald

RMSD Root Mean Square Deviation

RMSF Root Mean Square Fluctuation

SASA Solvent Accessible Surface Area

SMFP Single Molecule Fluorescence Polarization

SPC Single Point Charge

Page 13: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

XIII

TFP Trifluoperazine

VdW Van der Waals

VMD Visual Molecular Dynamics

Page 14: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

1

I. INTRODUCTION

1. Biological and Biochemical Background

1.1 The Myosin Family

Within all eukaryotic cells, the generation of mechanical force is provided by

specific motor proteins necessary for performing vast array of sub-cellular tasks required

to sustain life. The motors that move along actin filaments make up a diverse protein

family, collectively referred to as myosins (1). The myosin family of molecular motors

consists of at least 20 structurally and functionally distinct classes. Myosins have been

implicated in a variety of intra-cellular functions, including cell migration and adhesion,

intra-cellular transport and localization of organelles and macromolecules, signal

transduction, and tumor suppression. The myosins constitute a large family of actin-

dependent motors found in many organisms from yeast to humans. Upon interaction with

actin filaments, they convert energy from ATP hydrolysis into mechanical force

(reviewed in (2)).

The first identified motor protein was skeletal muscle myosin, which is

responsible for generating the force for muscle contraction (3). This protein, called

myosin II (for two-headed), was also found in non-muscle cells, including protozoan

cells. Myosin II is composed of two heavy chains; each bears a globular motor domain

that includes a binding site for ATP and a domain that interacts with actin. Two light

chains wrap around the elongated neck region of each heavy chain. In the late 1970's (4),

one-headed myosin (called myosin I), was found in the fresh water amoeba

Acanthamoeba castellanii. Myosin I is considered an unconventional myosin since it

functions as a monomer and consists of very short neck domain. Since then, many other

Page 15: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

2

myosin types were discovered. The new types of myosins include a number of one-

headed and two-headed varieties that are related to myosin I and myosin II, and the

nomenclature now reflects their approximate order of discovery (myosin III through at

least myosin XX).

The myosin tails (and the tails of motor proteins generally) have apparently

diversified during evolution to permit the proteins to dimerize with other subunits and to

interact with different cargoes. Some myosins (such as VIII and XI) have been found

only in plants, and some are exclusive to vertebrates (IX). However, most of them are

common to all eukaryotes, suggesting that myosins arose early in eukaryotic evolution

(reviewed in (5, 6)).

1.2 The Myosin V Protein

The family of myosin V is a class that differs structurally from other myosins by

having an extended neck domain and a tail domain that allows dimerization, but not the

formation of filaments due to the presence of a globular carboxy-terminus. Members of

the myosin V family have been identified in humans, mice, chickens, flies, fungi, worms,

yeast and plants. The myosin V proteins are molecular motors involved in a range of

organelle-transporting functions, including the transport of melanosomes and synaptic

vesicles in mammals and vacuoles and mRNA in yeast (reviewed in (7, 8)). Myosin V is

essential for transport of vesicles in actin-rich cortical regions of neurons, i.e. dendritic

spines and axon termini. Moreover, vesicle-associated myosins, like myosin V, may do

more than serve as mechanical feet, transporting vesicles from one location to another

along actin tracks. They may interact with membrane-associated F-actin to "stitch"

Page 16: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

3

membranes together during membrane fusion (9), and to sort proteins from the fluid

phase into lipid rafts during the assembly of membrane protein complexes (10). Based on

phylogenetic analysis of the myosin family, myosin V evolved after myosins II and I, the

two most ancient members of the myosin super-family (1, 11).

Myosin V is composed of two identical heavy chains, each one is capable of

binding six light chains (reviewed in (12, 13)). The two heavy chains dimerize to produce

a two-headed protein consisting of three distinct domains (Fig. 1).

The first domain is the head domain (green), the second domain constitutes the

neck (red), and third domain is the tail domain (colored yellow for its proximal and

medial modules, and blue for its distal module). When bound to the actin filaments, it has

the ability to convert the energy released by ATP hydrolysis into mechanical work, i.e.

Figure 1: A cartoon diagram of myosin V. The three domains are color coded: the head

domain is green, the neck domain is red, and the tail domain is yellow/blue. The light

chains, which engulf the neck domain, are colored in gray.

Page 17: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

4

movement (14). Successive cycles of ATP hydrolysis allow the two heads to "step" on

the actin filament towards its barbed or plus end (15).

The head domain of myosin V contains ATP and actin binding sites (16). Since

the force required for transport of cargoes is generated in the head domain, it is often

referred also as motor domain. The neck domain contains six repeating amino acid

sequences called IQ motifs (designated IQ1-IQ6), each of which binds a light chain. The

light chains have been shown to be CaM or CaM-like proteins, and their primary function

is to regulate the ATPase activity of the globular heads (17, 18). According to the lever

arm hypothesis (19, 20), the step size of myosin V motors is proportional to the length of

the neck domain, which functions as a lever. The long neck domain of myosin V gives

rise to a step size of ~ 36 nm, the largest step size thus far measured for a myosin motor.

Not surprisingly, the step size of myosin V shortens when one or more of the IQ motifs

from the neck domain is truncated (21).

It has been clearly established that myosin V motors from vertebrates are

processive, i.e. have a long duty cycle, enabling it to undergo multiple steps before

dissociating from the actin filament. This allows a single motor to transport cargo, in ~ 36

nm steps, for several micrometers along the actin filament (20). Yet, not all class V

myosins are processive. A recent kinetic analysis of myosin V from drosophila concluded

that it is not a processive motor (22). The two classes of myosin V from the yeast

Saccharomyces cerevisiae, Myo2p and Myo4p (the former is the myosin V discussed in

this research), have been reported to be low duty non-processive motors, based upon

motility and landing rate assays (23).

The tail domain can be divided into two modules: proximal/medial and distal. The

Page 18: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

5

proximal/medial module is the site of dimerization of the two heavy chains. In this

module of the tail, alpha-helical regions of the heavy chains are predicted to interact to

form a coiled coil segment that is interrupted by small globular regions (16). A PEST site,

which has been shown to be a cleavage site for the calcium-dependent protease calpain

(14), is located in the N-terminus of the proximal/medial tail module. The proximal/distal

module of the tail domain, in addition to its primary role in dimerization, also serves as

the link between the neck domain and the cargo-binding module of the tail. The distal

module of the tail domain is the cargo-binding domain (24). When bound to vesicles by

the distal tail module, myosin V becomes part of a large proteinous complex that includes

the membrane receptor.

1.3 The Mlc1p Protein

The Mlc1p protein is a CaM-like protein from the budding yeast Saccharomyces

cerevisiae, associated with the yeast's mechano-chemical myosin system. It was

originally identified as a light chain protein of the Myo2p protein, which is a myosin type

V (25). The Mlc1p protein is essential for viability and secretory vesicle delivery at the

mother bud neck domain during cytokinesis due to its ability to bind to the IQ motifs of

the Myo2p protein (26). While calcium binding to CaM promotes its binding/release

to/from the IQ motifs, the Mlc1p protein is unable to bind calcium since its EF-hand

motifs are abortive (27, 28). The signal that stimulates its interaction with target IQ

motifs is still unknown, but recently it was found that its interaction with the Myo2p

protein is mediated by GTP (29). Though also the exact function of the Mlc1p protein has

not been clarified yet, it is assumed that its binding to the neck domain of the Myo2p

Page 19: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

6

protein has a stabilizing effect on the latter (25). The three-dimensional structure of the

Mlc1p was determined when it was co-crystallized with the IQ2, IQ2/3 and IQ4 peptides

of the Myo2p protein (28). The Mlc1p protein shares a structural similarity with CaM;

both possess a dumbbell-shaped conformation, consisting N- and C- globular lobes

(referred, at the thesis, as NL for the N-lobe and CL for the C-lobe) that are connected by

a flexible inter-domain (referred, at the thesis, as ID).

1.4 IQ Motif Peptides

The IQ motifs constitute common target peptides for CaM and CaM-like proteins.

These motifs are widely distributed among different kinds of proteins, contributing >3700

Pfam entries in the database (30). The motifs are ~ 25 amino acids long, alpha-helical,

and carry a net positive charge. They conform to the general consensus sequence

IQXXXRGXXXXR, but in many cases, the sequence is rather loosely adhered to this

consensus. For example, the isoleucine in the first position is frequently replaced by other

branched-chain amino acids (leucine or valine) or, rarely, by a methionine. The arginines

in both the sixth and the terminal positions are sometimes replaced by lysine or histidine,

and the seventh position glycine is poorly conserved. Despite the lack of strict

conservation, there is no doubt that this sequence is a recognizable protein motif that

binds CaM and CaM-related proteins, such as Troponin C and myosin light chains.

IQ motifs are present in a number of neuronal growth proteins, sodium and

calcium voltage-dependent channels, EF-hand-containing protein phosphatases, spindle

pole and centrosomal proteins, transient receptor potential proteins, plant cyclic

nucleotide-regulated channels, certain signaling molecules, and at the neck region of the

Page 20: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

7

myosins. In most cases, the IQ motifs appear in tandem repeats, and therefore enable to

bind multiple copies of CaM and CaM-like proteins. IQ motifs were first identified as

Apo-CaM (Ca+2

-deprived CaM) binding sites; however, it is now clear that IQ motifs can

show different levels of Ca+2

dependency (31-33).

1.5 The Calmodulin Protein

The Calmodulin (CaM) protein is recognized as a prototypical calcium sensor and

regulation orchestrator of cellular events through its interaction with a diverse group of

proteins. The CaM protein is a small (16.7 kDa), highly acidic, ubiquitous, evolutionary

conserved protein. It is considered as the major calcium sensor and is expressed in all

eukaryotic cells, where it participates in signaling pathways that regulate crucial

processes such as growth, proliferation, and movement. The concentration and location of

CaM appear to play an important role in regulating its biological activity. CaM

constitutes at least 0.1% of the total protein present in cells (10-6

M – 10-5

M) and is

expressed at even higher levels in rapidly growing cells, especially those undergoing cell

division and differentiation. The local intra-cellular availability of CaM is likely to be

biologically significant because various CaM-dependent proteins are regulated over a

wide range of CaM concentrations. The dynamics of CaM and its interactions with target

proteins had been extensively studied (reviewed in (34, 35)).

CaM and CaM-like proteins belong to a family of soluble proteins that share a

common structure. These proteins are built of three structural domains: the NL, the CL,

and an elongated, mostly helical, ID that connects the two lobes to form a dumbbell-like

shape. Each lobe possesses two Ca+2

-binding domains, each of which is made of two

Page 21: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

8

helix-loop-helix EF-hand motifs. These proteins are present at two configurations: a

dumbbell extended state, where the ID is stretched; and a folded configuration, where the

ID bends, bringing the lobes into close contact. The CaM and CaM-like proteins can form

protein-protein complexes with specific target peptides, a reaction that is a key event in

many intra-cellular regulatory processes. On the binding of Ca+2

and in the presence of

target peptides, the CaM undergoes a conformational change that enables it to bind its

target. This basic mechanism activates well over 100 distinct target proteins. However, in

some cases (as at the Mlc1p protein), the reaction with target peptides can be mediated by

the Apo-CaM. The promiscuity of CaM on one hand, and binding specificity on the other

hand, made it one of the most studied proteins (reviewed in (36-38)).

1.6 Calmodulin-Peptide Complexes

Investigations of the CaM protein in a complex with different bound peptides

have been carried out by genetic methods, such as site-directed mutagenesis of the

protein (39-41), or the target peptide (42-44), and their thermodynamical properties were

explored (45, 46). In addition to these studies, the structure of the protein-peptide

complexes was investigated by NMR (47), and crystallographic studies (48-54). These

studies revealed that, upon binding of the peptides, the ID of Holo-CaM (Ca+2

-bound

CaM) adopts a bent conformation accompanied by its partial unwinding. The bending of

the ID brings together the two lobes of the protein. Thus, the flexibility of the ID region is

critical for the ability of CaM to interact with target peptides. The various aspects of these

interactions, such as recognition and activation, are reviewed by Vetter et al. (55).

Page 22: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

9

1.7 Mlc1p-IQ Complexes

The structure of the Mlc1p protein was solved by X-ray crystallography in

complexes with the IQ2, IQ2/3, and IQ4 peptides (28, 56). These Mlc1p-IQ protein-

peptide complexes were formed and crystallized in the absence of Ca+2

ions. According

to these structures, when bound to the IQ2 peptide, the Mlc1p protein adopts a compact

conformation in which both the NL and CL interact with the IQ motif. However, in the

complex with the IQ4 peptide, the NL of the protein does not interact with the IQ motif,

resulting in an extended conformation of the protein. Based on these crystal structures,

combined with comparative sequence analysis, a model for the entire poly-IQ (IQ1-IQ6)

sequence of the neck domain of myosin V, together with six light chain proteins, was

built (57). The model of the Light Chain-Binding Domain (LCBD) of myosin V has

important implications for the understanding of the structure-function relationship of the

lever arm of myosin V. In the present study, the crystal structures of the Mlc1p-IQ4,

Mlc1p-IQ2 and the model of the LCBD are discussed, analyzed, and reevaluated.

2. Computer Simulations

Dynamic simulation methods are widely used to obtain information on the time

evolution of conformations of proteins and other biological macromolecules (58-61), as

well as kinetic and thermodynamic information. Simulations can provide fine details

concerning the motions of individual particles as a function of time. They can be utilized

to quantify the properties of a system at a precision and on a time scale that is otherwise

inaccessible, providing a valuable tool in extending our understanding of model systems.

Theoretical computational consideration of a system additionally allows one to

Page 23: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

10

investigate specific contributions of property through “computational alchemy” (62), that

is, modifying the system in a way that is nonphysical but nonetheless allows a model’s

characteristics to be probed.

2.1 Molecular Dynamics

MD simulations have become valuable tools for investigating the basis of protein

structure and function. Given the structure of a biomolecular system, that is, the relative

coordinates of the constituent atoms, MD is used to investigate and study the dynamics of

that system. A brief outline of its history and methodology is offered below.

2.1.1 General Overview

The conformational dynamics of proteins, which is encoded in their structures, is a

critical element of their function. A fundamental appreciation for how proteins work

requires an understanding of the connection between three-dimensional

structures,

obtained by X-ray crystallography and NMR, and dynamics, which is much more difficult

to probe experimentally. Molecular Dynamics (MD) simulations provide links

between

structure and dynamics by enabling the exploration of

the conformational energy

landscape accessible to protein molecules (59, 63, 64). The first MD simulation of a

protein was reported in the year 1977 and consisted of a 9.2-ps trajectory for

a small

protein in vacuum (65). Eleven years later, a 210-ps simulation of the same protein in

water was reported (66), and the phenomenal increase in computing power since then

makes it routine to run simulations of much larger proteins for 1000-10000 times longer

than the original simulation (tenths of nanoseconds), in which the protein is surrounded

Page 24: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

11

by water and salt. Significant improvements in the potential

functions have also been

achieved, making the simulations much more stable and accurate (67).

Biomolecular dynamics simulations find two major areas of application today.

Firstly, MD simulations are used to bring biomolecular structures alive, giving insights

into the natural dynamics on different timescales of biomolecules in solution. Secondly,

MD simulations can explore which conformations of a molecule or a complex are

energetically accessible, by searching the conformational space. Driven by improvements

in simulation methodology, increasing accuracy of biomolecular force fields and the

ever-increasing power of computers, MD simulations is a rapidly progressing scientific

discipline. Nowadays, it is well established that MD is not a mere historical phenomenon,

but an important fundamental developing field of science (68-71).

2.1.2 Methodology

We briefly sum up the methodology of MD, but for a comprehensive inclusive

literature the reader is referred to (72, 73). MD simulations calculate the "real" dynamics

of the system, from which time averages of properties can be calculated. Sets of atomic

positions are derived in sequence by applying Newton's equations of motion. MD is a

deterministic method, i.e. the state of the system at any future time can be predicted from

its current state. The first MD simulations were performed using very simple potentials

such as the hard-sphere potential. The behavior of the particles in this potential is similar

to that of billiard balls: the particles move in straight lines at a constant velocity between

collisions. The collisions are perfectly elastic and occur when the separation between a

pair of spheres equals the sum of their radii. After a collision, the new velocities of the

Page 25: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

12

colliding spheres are calculated using the principle of conservation of linear momentum.

The hard-sphere model has provided many useful results but is obviously not ideal for

simulating atomic or molecular systems. In potentials such as the Lennard-Jones (LJ)

potential, the force between two atoms or molecules changes continuously with their

separation. The continues nature of more realistic potentials requires the equations of

motion to be integrated by breaking the calculation into a series of very short time steps

(in our simulations the time step was usually 2-fs). At each step, the forces operating on

the atoms are computed and combined with the current positions and velocities to

generate new positions and velocities. The force acting on each atom is assumed to be

constant during the time interval. The atoms are then moved to the new positions; an

updated set of forces is computed, and so on. In this way, an MD simulation generates a

trajectory that specifies how the positions and velocities of the particles in a system

change with time.

In biomolecular simulations, such as presented in this study, successive

configurations of the system are produced by iterative numerical calculations of

instantaneous forces present in a molecular mechanical system and the consequential

movements in that system. The molecular mechanical system consists of a set of particles

that move in response to their interactions according to the equations of motion defined in

classical Newtonian laws of motion. Newton's laws of motion can be stated as follows:

(I) A body continues to move in a straight line at a constant velocity unless a force acts

upon it.

(II) Force equals the rate of change of momentum.

(III) To every action there is an equal and opposite reaction.

Page 26: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

13

The trajectory is obtained by solving the differential equation embodied in

Newton's second law (F = m·a):

(1)

Equation 1 describes the motion of a particle of mass (mi) along one coordinate

(xi) with (Fxi) being the force on the particle in that direction.

2.1.3 Molecular Dynamics of Calmodulin

The unique structure of CaM and its fundamental role in Ca+2

-signaling processes

has also drawn the attention of computational scientists besides the experimental

biologists and biochemists. Early MD simulations of Holo-CaM confirmed the flexibility

of its ID (74-76). However, due to limited computational power available at that time,

these simulations were carried out over sub nanosecond time scales and included

relatively few, if any, water molecules. In the recent decade, MD simulations of CaM

were extensively performed under more realistic conditions, addressing issues concerning

the relative positions of its NL and CL and the nature of its flexible ID. In the year 1996,

Van Der Spoel et al. (77) performed an MD simulation of the ID of Holo-CaM, and

observed its bending. Since this pioneering MD simulation of CaM in an aqueous

solution, simulations of Apo- (78, 79) and Holo- CaM (79-87) have been performed for

durations ranging from hundreds of picoseconds to 20-ns. These MD studies

demonstrated the unwinding and curvature of the ID of the protein. Moreover, it was

found out that, although bound Ca+2

ions harden the structure of the protein, it collapses

i

Xi

m

F

dt

xdi=

2

2

Page 27: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

14

to a form resembling a dumbbell with a hinged bar connecting its lobes.

2.1.4 Molecular Dynamics of Calmodulin's Complexes

Only a few MD simulations of CaM together with target molecules or peptides

had been performed. An MD simulation of CaM-TFP complex was carried out,

comparing the binding affinities of the NL and CL of Holo-CaM to TFP (88). The

binding dynamics of melatonin to Holo-CaM were examined by means of MD after

molecular docking (89). The dynamics and entropy of a Holo-CaM complexed with a

peptide were studied by NMR and MD. A good agreement was found between amide

order parameters measured by NMR, and those obtained from the simulation (90).

Finally, an MD simulation of a complex between Holo-CaM and a peptide examined the

structure, dynamics and the interaction mode between the protein and the peptide. In a 4-

ns MD simulation, the CaM-peptide complex was quite rigid and did not exhibit any

large amplitude domain motions (91, 92).

2.2 Free Energy

The calculation of free energy from molecular simulations is an area of intense

research activity (reviewed in (93)). This is because free energy is at once one of the most

central and one of the most difficult thermodynamic quantities to compute from atomic

level simulations. It is of paramount importance in efforts to relate microscopic details

stemming from atomic interactions to measurable macroscopic quantities, and to

understand the physical and structural basis of biological phenomena. In particular, free

energy of interaction is a measure of the stability of a complex, a measure that is

Page 28: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

15

fundamental to all studies of biomolecular binding processes.

2.2.1 Free Energy of Interaction

The interactions between proteins and other molecules are critical to all biological

systems and processes. Signal transduction, metabolic regulation, enzyme cooperativity,

physiological response, and other processes are all dependent upon noncovalent binding.

These processes may be investigated through modeling and computer simulations. At this

study, we estimate interaction free energy for the protein-peptide complexes through MD

simulations and complementary computation techniques. Approaches available for

estimating interaction free energies cover a broad range of accuracies and computational

requirements. Given that the purpose of this research is not to carry a comparative study

of interaction free energy calculations' techniques, but rather use the interaction free

energy to characterize the protein-peptide complexes, we chose to calculate the

interaction free energy in a nonrigorous technique, called the MM-PBSA approach.

2.2.2 The MM-PBSA Approach

Molecular Mechanics Poisson-Boltzmann Surface Area (MM-PBSA) is basically

a post-processing method to evaluate the standard free energies of molecules or the

interaction free energies of molecular complexes in a relatively computationally efficient

manner. The MM-PBSA approach was developed by Srinivasan and co-workers in 1998

(94). Since then, it has been widely used for estimation of free energies of different RNA

(94), DNA (95, 96) and protein conformations (97), binding affinities of protein

complexes and mutational analysis on them (98-102), binding affinities of small

Page 29: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

16

compound-protein complexes (103, 104), interaction free energies of RNA-protein (105),

RNA and metal ions (106), and RNA-ligand (107) complexes. The strategy of

calculations is based on applying of a continuum model to solute configurations derived

from an MD simulation in explicit solvent. For each selected solute configuration,

molecular mechanics energy is determined. Free energies of solvation are estimated by

using Poisson-Boltzmann (PB) calculations for the electrostatic contribution and a

surface-area-dependent term for the nonelectrostatic contribution to solvation. Solute

entropic contributions are estimated from a normal mode analysis. To get a statistically

meaningful value of the interaction free energy of a complex, calculations are commonly

carried out on several snapshots extracted from an MD trajectory with explicit solvent.

3. Significance of Study

The general significance of this research lies in the fact that it demonstrates that

careful application of MD can be used for evaluation of proteinous structures, offering

close analysis as well as deeper insights of crystalline configurations. We have employed

a multi-stage approach in order to investigate the Mlc1p-IQ complexes. For this purpose

we performed MD simulations of the Mlc1p-IQ2 complex, the Mlc1p-IQ4 complex, and

the free IQ2 and IQ4 peptides at various conditions. Our MD simulations enable to

propose solution conformations of protein complexes, which were not offered by X-ray

crystallography studies. Besides grossly expanding the structural data embedded in the

crystalline configurations of these protein complexes, our study provides detailed analysis

of the protein-peptide interaction energy, and atomistic understanding about the

movement mode of myosin V over the actin filament. Thus, this research, which starts

Page 30: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

17

from MD simulations and ends at an enlightment of the dynamic aspects of a

physiological phenomenon, illustrates the biophyiological relevance of the MD

methodology.

Page 31: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

18

II. METHODS

1. MD Simulations

1.1 Simulations of Protein-Peptide Complexes

1.1.1 Mlc1p-IQ2 Complex

The MD simulations were performed using the GROMACS 3.2.1 package of

programs (108-110), with the GROMOS96 43a1 force field (111). The crystal structure

of the Mlc1p protein bound to an IQ2 peptide of the Myo2p protein (PDB file 1M45),

determined by X-ray crystallography at 1.65 Å (56), was downloaded from the Protein

Data Bank (112). Four missing residues (D-50, S-51, R-54, D-55) were added to the

structure using the PROFIX program, which is incorporated in the JACKAL molecular

modeling package (113). The protein-peptide complex was embedded in a box containing

the SPC water model (114), that extended to at least 12 Å between the protein-peptide

structure and the edge of the box. Assuming normal charge states of ionizable groups

corresponding to pH 7, the net charge of the Mlc1p-IQ2 structure is -7e. Hence, 35

sodium and 28 chloride ions were added to the simulation box at random positions, to

neutralize the system at a physiological salt concentration of ~ 100 mM. Prior to the

dynamics simulation, internal constraints were relaxed by energy minimization.

Following the minimization, an MD equilibration run was performed under position

restraints for 40-ps. Then, an unrestrained MD run was initiated. The first 100-ps of the

run were treated as a further equilibration simulation, and the remainder 12-ns were saved

and used for the analysis. During the MD run, the LINCS algorithm (115) was used in

order to constrain the lengths of all bonds; the waters were restrained using the SETTLE

algorithm (116). The time step for the simulation was 2-fs. The simulation was run under

Page 32: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

19

NPT conditions, using Berendsen's coupling algorithm for keeping the temperature and

the pressure constant (117) (P = 1 bar; τP = 0.5 ps; τT = 0.1 ps; T = 300 K). Van der

Waals (VdW) forces were treated using a cutoff of 12 Å. Long-range electrostatic forces

were treated using the PME method (118). The coordinates were saved every 1-ps.

1.1.2 Mlc1p-IQ4 Complex

The simulations' conditions for the Mlc1p-IQ4 protein-peptide complex were the

same as those for the Mlc1p-IQ2 protein-peptide complex. The calculations were carried

out using the crystal structure of the Mlc1p protein bound to an IQ4 peptide of the Myo2p

protein (PDB code 1M46) determined by X-ray crystallography at 2.1 Å (56), that was

downloaded from the Protein Data Bank (112). The net charge of the complex is -3e, and

hence 43 sodium and 40 chloride ions were added in random positions to neutralize the

system at a physiological salt concentration of ~ 100 mM. The simulations were

performed at two temperatures, 300 K and 400 K. The simulations' conditions were

similar at both temperatures, except for the time step and τP that, at the 400 K, were 1.5-fs

and 2-ps, respectively.

1.2 Simulations of Peptides

1.2.1 IQ2 Peptide

The coordinates for the IQ2 peptide were derived from the crystal structure of the

Mlc1p-IQ2 protein-peptide complex. The net charge of the IQ2 peptide is +2e, and hence

13 sodium and 15 chloride ions were added in random positions to neutralize the system

at a physiological salt concentration of ~ 100 mM. The conditions of the simulation were

Page 33: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

20

similar to those described for the Mlc1p-IQ2 complex.

1.2.2 IQ4 Peptide

The coordinates for the IQ4 peptide were derived from the crystal structure of the

Mlc1p-IQ4 protein-peptide complex. MD simulations were performed at various

conditions as elaborated on Table 1. The net charge of the IQ4 peptide is +6e, whereas

the number of ions that were added at the simulations varied according to the specific

conditions of each simulation. The conditions of the simulations were similar to those

described for the Mlc1p-IQ4 complex.

Table 1

Summary of the MD Simulations of the IQ Peptides

* High temperature (400 K). †

Low salt concentration. ‡

High salt concentration. ¶ Very high salt concentration.

§ An IQ4 peptide containing 10 additional amino acids flanking each of its terminals was

built using Swiss PDB Viewer (119). These added 20 residues belong to the IQ3 and IQ5

peptides. Thus, the 45 amino acids elongated IQ4 peptide constitutes a portion of the

poly-IQ sequence as present at the neck of the myosin.

2. Visual Presentations

All protein and peptides figures were created using the VMD computer program

(120).

Duration (ns)

Concentration

of Salt

Number of

added Na+

ions

Number of

added Cl-

ions

IQ4 peptide 100 ~ 100 mM 12 18

IQ4 peptide * 30 ~ 100 mM 12 18

IQ4 peptide † 20 ~ 30 mM 4 10

IQ4 peptide ‡ 20 ~ 300 mM 36 42

IQ4 peptide ¶ 20 ~ 2.4 M 288 294

Extended IQ4 peptide § 20 ~ 100 mM 10 17

Page 34: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

21

3. Inter-Helical Angles

For the protein-peptide complexes, inter-helical angles were calculated according

to the following procedure: The secondary structure of the protein and the peptide were

determined using the STRIDE algorithm (121). Three residues were selected at the N-

and C- terminals of each alpha helix. The C-alpha atoms' center of mass of each triplet of

residues was determined. A vector-defining-helix was drawn between these centers of

mass along each helix. The inter-helical angle was calculated between each two

successive vectors. For the IQ peptides, intra-helical angle was calculated according to

the same procedure.

4. Dihedral Angle Calculations

The position of the protein’s lobes towards each other can be expressed by

measuring the dihedral angle between the planes defined by the two lobes and the ID.

Each plane was defined by the straight section of the ID and a selected representative

residue located at each lobe. The C-alpha atoms of residues N-47, L-58, and V-69 defined

one plane; whereas the C-alpha atoms of residues L-58, V-69, and E-129 defined the

other. The calculation of the dihedral angles was performed for the last snapshot of the

simulations at t = 12-ns using a standard GROMACS utility.

5. The Electrostatic Potential Around the Peptides

The electrostatic potential around the IQ peptides was calculated for the model

structures of the peptides as derived from the Mlc1p-IQ protein-peptide simulations. The

coordinates of 21 snapshot structures, extracted every 100-ps from t = 10 until t = 12-ns,

Page 35: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

22

were used for the electrostatic potentials calculations. The electrostatic potential surface

around each peptide was calculated by solving the nonlinear Poisson-Boltzmann (PB)

equation through the use of the APBS (Adaptive Poisson-Boltzmann Solver) software

package (122) with a grid spacing of 0.5 Å. The volumes of the averaged positive and

negative electrostatic fields of both peptides (for the time frame t = 10 until t = 12-ns) and

the Coulomb cages at the last snapshot (t = 12-ns) of the simulations are presented.

6. The Protein-Peptide Interaction Free Energies

The general strategy used for calculating the protein-peptide interaction free

energy is based on the MM-PBSA method. This method was successfully employed by

numerous studies (94-107), and involves calculating energies for snapshot configurations

taken from the MD trajectories of the Mlc1p-IQ complexes. The configurations of the

protein-peptide complexes, the protein and the peptides were obtained from the MD

simulations of the Mlc1p-IQ2 and the Mlc1p-IQ4 structures. The coordinates of 21

snapshot model structures, extracted at 100-ps intervals during the last 2-ns of the

simulations, where both complexes appeared to gain a stable configuration, were used for

the calculations. The calculations, which were performed for each of these snapshots and

their average values are presented, were intended for estimation of the protein-peptide

free energy interaction.

The changes in the Gibbs free energy of interaction were calculated from the

atomic model structures of the protein and the peptide undergoing the binding to form the

protein-peptide complex. Thus, the free energy of interaction was defined as presented in

equation 2:

Page 36: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

23

(2) ∆Ginteraction = (Gcomplex) − (Gprotein) − (Gpeptide)

The calculations of the free energy of each molecule were carried out according to

equation 3:

(3) TSGGE)(G solvation nonpolar,solvation polar,MMmolecule −⟩⟨+⟩⟨+⟩⟨=

where the free energy was decomposed into molecular mechanics (⟨EMM⟩), polar

solvation (⟨Gpolar,solvation⟩), nonpolar solvation (⟨Gnonpolar,solvation⟩) and entropy (TS)

contributions. ⟨ ⟩ denotes an average over a set a snapshots along an MD trajectory. Each

term on the right side of the equation was calculated as detailed below.

6.1 Molecular Mechanics Calculations

The molecular mechanics contribution to the free energy of interaction energy

was calculated according to equation 4:

(4) ⟩⟨+⟩⟨+⟩⟨=⟩⟨ VdWticelectrostaintMM EEEE

(⟨Eint⟩) includes bond, angle, and torsional angle energies, while (⟨Eelectrostatic⟩) and

(⟨EVdW⟩) denote the intra-molecular electrostatic and VdW energies. (⟨Eelectrostatic⟩) was

calculated using the APBS software package (122). (⟨EVdW⟩) was calculated using a

standard GROMACS utility.

6.2 Polar Solvation Calculations

The electrostatic contribution to the solvation energy, (⟨Gpolar,solvation⟩), was

determined by using a continuum electrostatic with the Poisson-Boltzmann (PB)

approach (123). We used the APBS software package (122), with a grid spacing of 0.5 Å

and solution of 100 mM NaCl, for the numerical solution of the nonlinear PB equation.

Page 37: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

24

The term (⟨Gpolar,solvation⟩) refers to the energy associated with the transfer of the solute

from a continuum medium with a low dielectric constant (ε = 4) to a continuum medium

with the dielectric constant of water (ε = 78.4).

Crucial to the application of PB models, and a source of many scientific disputes,

is the so-called macromolecule dielectric constant, ε. It is generally accepted that, in a

continuum electrostatics approach, the dielectric constant of the solute is a scaling factor

that represents all the contributions that are not treated explicitly, rather than a true

dielectric constant (124, 125). Different protein-associated dielectric constants are

frequently used in the literature, and we chose to perform the calculation with a dielectric

constant of 4 as commonly employed (123, 126-128). To make sure that the conclusions

derived from our calculations are not dependent upon the choice of the value used, we

repeated the calculation for representative snapshots with a lower (ε = 2) and a higher (ε

= 8) dielectric constants. Comparison between the results indicated that, although the

value of the calculated (⟨Gpolar,solvation⟩) varies with ε, its trend is independent from the

dielectric constant used.

6.3 Nonpolar Solvation Calculations

The nonpolar contribution to the solvation free energy, (⟨Gnonpolar,solvation⟩), was

determined by using the Solvent Accessible Surface Area (SASA). The SASA was

calculated by a standard GROMACS utility, which implements the double cube lattice

method (129) with a probe radius of 1.4 Å. The nonpolar solvation energy was described

as Gnonpolar,solvation= γ·(SASA) + β. The constants γ and β are 2.2 kJ mol-1

nm-2

and 3.84 kJ

mol-1

, respectively. These values of γ and β are in accord with the MM-PBSA approach

Page 38: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

25

(101, 103-105, 107).

6.4 Entropy Calculations

Entropy changes upon binding of the peptide were calculated by use

of normal mode analysis. The conformations of the protein, the peptide and

the protein-peptide complex were extracted from the last frame of each of the trajectories

as performed in (101). The model structures were subjected to rigorous energy

minimization, until the maximal force operating on an atom was less than 10-6

kJ mol-1

nm-1

. Normal mode analysis (130-132) was performed by calculating and diagonalizing

the mass-weighted Hessian matrix. The frequency of the normal mode was then used to

calculate the vibration entropy (133) as given by equation 5:

(5)

where Svib is the vibrational entropy, h is Planck's constant, ν0 is the frequency of the

normal mode, k is the Boltzmann constant, T is the absolute temperature and NA is

Avogadro's number. All calculations were performed with the GROMACS program,

compiled with double precision.

)T(

hvN)(R

vibee

eS /kThv

/kThv

A/kThv

0

0

0

11ln

0

−+−−=

Page 39: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

26

III. RESULTS & DISCUSSION

1. Simulations of Protein-Peptide Complexes

In this chapter of the thesis, we describe an investigation, through MD

simulations, of the dynamics of two Mlc1p-IQ complexes: the Mlc1p-IQ2 and the Mlc1p-

IQ4 complexes.

1.1 Mlc1p-IQ2 Complex

The Mlc1p-IQ2 complex, which had been resolved by crystallography to 1.65 Å,

confers to a Ca+2

-independent stable structure. Its 12-ns MD simulation is presented.

1.1.1 Crystallographic and Simulated Structures

The crystallographic structures and those obtained by the simulations of the

Mlc1p protein with the IQ2 peptides are presented in Fig. 1.1. The NL, the ID, the CL

and the IQ peptides are colored in blue, red, green and yellow, respectively.

At the crystalline structure of the Mlc1p-IQ2 complex (Fig. 1.1A), the Mlc1p

protein is found at a compact state as evident by its curved ID. In this configuration, the

CL of the protein engulfs the IQ2 peptide, which interacts also with the NL and the ID of

the protein. Overall, the configurations of the Mlc1p protein and the IQ2 peptide only

slightly change throughout the 12-ns long simulation, and hence their final simulated

state structures (Fig. 1.1B) resemble the crystalline ones. Thus, the simulated structure of

the Mlc1p-IQ2 complex exhibits just a few minor conformational deformations compared

to its crystalline structure. These deformations consist of appearance of a new kink

located at helix D of the protein's ID (Fig. 1.1B, see arrow), a consequent rotation of the

Page 40: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

27

NL, and minor changes of the inter-helical angles observed mainly at its CL, as

elaborated below.

1.1.2 The Dynamics of the Protein-Peptide Complexes

A quantitative expression of the conformational change is given by the RMSD

(Root Mean Square Deviation) of the backbone atoms of the protein-peptide complexes.

Fig. 1.2A depicts the RMSD values as calculated for the whole Mlc1p-IQ2 complex

(black), and its components: the Mlc1p protein (red), and the IQ2 peptide (green). The

RMSD of the protein-peptide complex exhibits some structural fluctuations that can be

fully attributed to the Mlc1p protein. It increases until ~ 0.35 nm, stays around this value

for ~ 4-ns, decreases for a short while and then stabilizes at ~ 0.28 nm. The RMSD track

Figure 1.1: Cartoon diagrams of the crystal and the simulated structures of the Mlc1p

protein when it binds the IQ2 peptide (PDB 1M45). The NL (residues 1−59), the ID

(residues 60−92), the CL (residues (93−148), and the IQ peptide are shown in blue, red,

green and yellow, respectively. Both crystal structures, and both simulated solution

structures, are presented with the same orientation, where the NLs are structurally

aligned. (A) The crystal structure of the Mlc1p-IQ2 complex; (B) The simulated

structure of the Mlc1p-IQ2 complex after 12-ns simulation. The arrow points towards a

kink discussed at the text.

B A

C-lobe

N-lobe Inter-domain

Page 41: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

28

of the IQ2 peptide, which contributes little to the RMSD of the complex, exhibits a

different pattern. It appears to explore the configurational space until ~ 3.2-ns, and then

increases to a value of ~ 0.23 nm. From this time point until the end of the simulation, the

RMSD of the IQ2 peptide is relatively stable.

The fluctuations of the Mlc1p protein can be resolved into a contribution of its

structural domains. Accordingly, the RMSD of its NL (black), ID (red), and CL (green)

are presented in Fig. 1.2B.

The RMSD of the NL exhibits a sharp increase at ~ 5-ns, and then stabilizes at a

value of ~ 0.24 nm. The RMSD of the ID, which is the flexible domain of the Mlc1p

protein, hardly changes. The stability of the ID throughout the simulation time is not

surprising since its structure is already bent and curved at the Mlc1p-IQ2 protein-peptide

A B

Figure 1.2: (A) The RMSD of the backbone atoms of the Mlc1p-IQ2 complex (black), the

Mlc1p protein (red), and the IQ2 peptide (green) as a function of the simulation time; (B)

The RMSD of the backbone atoms of the different domains of the Mlc1p protein as a

function of the simulation time. The domains of the protein are defined as follows: residues

1−59 for the NL (black), 60−92 for the ID (red), and 93−148 for the CL (green).

Page 42: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

29

crystalline structure. Apparently, the CL is much more flexible than the other structural

domains, exhibiting the largest variations in its RMSD value. Inspection of the Mlc1p-

IQ2 model structure reveals that most of its eight alpha helices retain their structures and

the angles between them, except the angles found at the CL. The angle between its G and

H helices increases from ~ 100° at the beginning of the simulation to ~ 115° at the end of

it. The angle between its E and F helices changes from ~ 110° to ~ 123° during the time

frame of 2.7 until 5.5-ns, but settles back at its original value. The fluctuations of the

RMSD of the CL are due to these inter-helical motions.

1.1.3 Summary

Overall, the configurations of the Mlc1p protein and the IQ2 peptide barely

change throughout the 12-ns long simulation. This relative stability of the complex

inspired us to perform MD simulation for another complex, the Mlc1p-IQ4 complex,

which is composed of the same protein and a slightly different peptide.

Page 43: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

30

1.2 Mlc1p-IQ4 Complex

The Mlc1p-IQ4 complex, which had been resolved by crystallography to 2.1 Å,

confers to a Ca+2

-independent stable structure. During its MD simulations, the complex

undergoes a complicated modulation process, which involves bending of the angles

between the alpha-helices of the protein, breaking of the alpha-helical structure of the

IQ4 peptide into two sections, and formation of new contact points between the protein

and the peptide. The dynamics of the process consist of fast sub picosecond events and

much slower ones that take a few nanoseconds to completion. Our study expands the

information embedded in the crystal structure of the Mlc1p-IQ4 complex by describing

its dynamic behavior as it evolves from the crystal structure to a form stable in solution.

1.2.1 Overall Conformational Changes During the Simulation

Fig. 1.3 depicts the structure of the Mlc1p-IQ4 complex as present in the crystal

structure (Fig. 1.3A), and after 12-ns of simulation at 300 K (Fig. 1.3B). The structures

were aligned over their NLs, to present them from the same orientation. In its crystalline

configuration, the Mlc1p-IQ4 complex, (PDB file 1M46), has an extended structure,

where the IQ4 peptide is mainly bound to the CL of the Mlc1p protein (for details see

Fig. 1.3 below). The ID (residues 60−92, colored red) that connects the NL (1−59,

colored blue) with the CL (residues 93−148, colored green) is a straight alpha helix. The

IQ4 peptide (colored in yellow) also keeps a straight alpha helix configuration,

perpendicular to the long axis of the ID. It forms few contacts with the CL and the ID of

the Mlc1p protein. At the end of the 12- ns simulation, the complex reached a new

conformation. The most stable part of the Mlc1p-IQ4 complex is the NL, which even

Page 44: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

31

after 12-ns retained most of its original conformation and its contact points with the N-

terminal section of the ID. The ID had refolded, so that its two helices (helices D and E),

became perpendicular each to the other and are in contact with the NL and CL. The CL

itself rotated and refolded into a shape that engulfs the IQ4 peptide, snapping its straight

helix into two, almost perpendicular sections. In the final configuration, the number of

contact points between the peptide and protein increased and the three sections of the

protein are in contact with the peptide.

The modulation of the protein and peptide with time can be followed by

examination of the backbone RMSD relative to the crystal structure. Fig. 1.4A depicts the

RMSD values as calculated for the whole complex (black), and its components: the

A

C-lobe

N-lobe

Inter-domain

B

Figure 1.3: Cartoon diagram of the crystal and the simulated structures of the Mlc1p

protein bound to the IQ4 peptide (PDB 1M46). The N-lobe, the inter-domain, the C-lobe,

and the IQ4 peptide are shown in blue, red, green, and yellow, respectively. The domains

of the protein are defined as residues 1–59, NL; 60–92, ID; and 93–148, CL. (A): The

crystal structure; (B): The simulated structure at 300 K after the 12-ns simulation in the

presence of water and ions. Both structures are presented with the same orientation.

Page 45: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

32

Mlc1p protein (red) and the IQ4 peptide (green).

Some 3.2-ns after the initiation of the simulation, the RMSD trace reveals a major

structural modulation. The RMSD value of the entire Mlc1p-IQ4 complex and that of the

Mlc1p protein almost double (from ~ 0.4 to ~ 0.7 nm). The rise continues in a moderate

fashion till the end of the simulation. In contrast to the complex and the protein, the

RMSD calculated for the IQ4 peptide (green line) exhibits different dynamics. The trace

A B

C D

Figure 1.4: (A) The RMSD of the backbone atoms of the Mlc1p-IQ4 complex (black), the

Mlc1p protein (red), and the IQ4 peptide (green) as a function of the simulation time; (B)

The RMSD of the backbone atoms of the different domains of the Mlc1p protein as a

function of the simulation time. The domains of the protein are defined as residues 1−59

for the NL (black), 60−92 for the ID (red), and 93−148 for the CL (green); (C) Expansion

of frame B between 2- and 3.5- ns; (D) The RMSD of the backbone atoms of the last two

ns of the simulation (10−12 ns) relative to the position of the same set of atoms at t = 10-

ns.

Page 46: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

33

reveals a reversible increment at ~ 1.75-ns, followed by a minor response to the structural

change of the Mlc1p protein at 3.2-ns. A short time later, the peptide undergoes some

structural modification, which increases its RMSD to ~ 0.4 nm. From this time point on,

the RMSD of the peptide is practically constant. The fact that the RMSD values of the

peptide hardly reflect the transition detected in the RMSD of the Mlc1p protein, implies

that the structural changes are driven by the Mlc1p protein itself, as evident in Fig. 1.4B.

Fig. 1.4B presents the RMSD values of the structural elements of the protein: the

NL, the CL and the ID as shown (marked in black, green and red, respectively). The NL

and the CL exhibit only moderate deviations from their initial structures, reaching, at the

end of the simulation, RMSD values of ~ 0.27 nm and ~ 0.36 nm, respectively. The ID

section of the protein is much more flexible than the lobes, as its RMSD exhibits large

variations before and after the conformational change of the complex. These events are

depicted with higher temporal resolution in Fig. 1.4C. During the 2.1−2.8-ns time

interval, the ID flexes for a very short time. Its RMSD increases to > 0.3 nm, but

eventually the ID relaxes to its original structure. Within this time frame, the CL slowly

regains a new conformation, roughly doubling its RMSD from 0.1 to 0.2 nm. Once the

new conformation of the CL was attained, the next structural modulation of the ID

appears to lead to the main conformational change at t ~ 3.2-ns. The RMSD of the CL

further increases and, after a short delay, the NL undergoes its conformation change.

Thus, the structural transition process experiencing the protein-peptide complex is

composed of a set of sequential steps, not a concerted motion.

Once the complex had reached its new stable conformation, the protein and the

peptide do not exhibit further major conformational changes. The calculation of the

Page 47: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

34

backbone RMSD value for the last 2-ns of the simulation, relative to the backbone heavy

atoms’ position at t = 10-ns, reveals only moderate deviations around 0.25 nm (Fig.

1.4D). This is an indication that, in the refolded state, the complex does not sample many

other conformations but rather settles down to a stable configuration with no further

evolution.

1.2.2 Relative Rotation of the Helices During the Simulation

The Mlc1p protein, like CaM, consists of 8 well-defined helices (A−H).

Inspection of the Mlc1p-IQ4 complex's structure (Fig. 1.3) reveals that most of the alpha

helices retained their structure. The variations of the RMSD reflect mostly the rotation of

the helices one with respect to the other. This motion can be expressed in changes of the

angle between the helices. Thus, in order to examine the relative rotation of the helices

during the simulation, we have calculated the inter-helical angles between each

successive pair of them, and present the results in Fig. 1.5A−C. A snapshot from the time

point t = 12-ns is shown on the right of each panel in order to highlight the discussed

helices. It should be noted that the cartoon diagram of the Mlc1p-IQ4 complex is

presented solely clarification purposes, and we can not deduce the angle itself from it.

The first three helices (A−C) are located at the NL, which hardly experienced

conformational changes (see Fig. 1.3). The angles between the helices of this lobe barely

changed throughout the simulation course (data not shown). The angle between the NL

and the first helix of the ID was also stable with time, and not affected by the structural

transition of the complex. The rest of the protein was more flexible, and the dynamics of

the relative motions of its helices is presented in Fig. 1.5. The first global motion of the

Page 48: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

35

Mlc1p protein is a slow (~ 1-ns), ~ 35° modulation in the F-G angle (Fig. 1.5A),

representing a deformation of the CL (~ 1−2 ns). Immediately later, this angle exhibits a

reversible bending, a process that may be considered as a preparation for the major

structural deformation event. This event is characterized by rotations of two pairs of

helices (D vs. E and G vs. H).

A

F G

B

D

E

C

H

G

D

N-section

C-section

Figure 1.5: The dynamics of the relative rotation of the alpha helices of the Mlc1p

protein and the IQ4 peptide during the 12-ns simulation time. In order to clarify which

angle is presented on each panel, a cartoon diagram of the complex after 12-ns of

simulation is shown on the right of each panel in black, whereas the discussed helices

are shown in the domain's characteristic color. (A) The inter-helical angle between

helix F (residues 102−111) and helix G (residues 119−126) of the CL; (B) The inter-

helical angle between helix D (residues 60−70) and helix E (residues 81−92) of the

ID; (C) The inter-helical angle between helix G (residues 119−126) and helix H

(residues 138−147) of the CL; (D) The inter-helical angle between the N- (residues

3−8) and C- (residues 18−23) sections of the IQ peptide.

Page 49: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

36

The D and E helices are the components of the ID, with a hinge located between

them at residues 71−80. This hinge exhibits a considerable flexibility, demonstrated by

NMR studies of Apo-CaM (134, 135), crystal structures of Holo-CaM by itself (136) and

in a complex with peptides (49, 50), and MD simulations of Apo- and Holo-CaM (78,

81). Hence, it is well established that the ID is a flexible bar rather than a stiff stick. Also

in the present simulation the hinge flexes, initiating the collapse of the extended structure.

The angle between the helices of the ID increases by almost 80° within 400-ps starting at

~ 3.2-ns (Fig. 1.5B). This bending leads to the refolding of the dumbbell-like shape of the

Mlc1p protein to its final model solution structure.

Another pair of rotating helices (G and H) is located at the CL (Fig. 1.5C). The

angle between them decreases gradually by ~ 50° during a time frame that precedes the

major structural change event (~ 2−3 ns). This change is followed by a sharp increase in

the angle, corresponding with the snapping of the straight helix of the peptide into two,

almost perpendicular, sections.

In order to demonstrate the modulation of the IQ4 peptide during the simulation,

two alpha helices, in its N- and C- sections, were defined. The angle between them is

presented in Fig. 1.5D. The IQ4 peptide, which at the initial state consists of a single

alpha helix, shows a significant flexibility. Its straight alpha helical structure is snapped

in its middle during the complex's structural transition, in a process that is slightly

delayed with respect to the major conformational change, when the two shorter helices

attain an almost perpendicular relation. The process, in which the angle between the

helices increases from ~ 20° to ~ 70°, is sharp and quick.

The major conformational change of the Mlc1p protein is preceded by reversible

Page 50: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

37

changes at the RMSD values of the ID (Fig. 1.4C), and the inter-helical angle between

helices of the CL (Figs. 1.5A and 1.5C). Analysis of these events shows that the protein

undergoes several minor conformational changes that precede the major conformational

modulation. These changes may be needed to allow the modulation of the protein

structure, and to enable the structural change (i.e. the accumulation of many minor

changes drives the major refolding of the complex). It is of interest to point out that the

rotation of the CL protein’s helices (Figs. 1.4A and 1.4C) precedes the sharp rise at the

RMSD values (Figs. 1.3A−C), and then proceeds, although after that the RMSD function

seemed to reach constant values. This may indicate that the structural modulation of the

structure is not an instantaneous event, but rather a complex “Plate Tectonics”, whereas

the forces are operating well before and after the deformation event.

1.2.3 The Structural Characteristics of the Compaction Event

The variations of three major structural characteristics of the Mlc1p protein with

simulation time (namely, the length of its ID, the distance between its N- and C- lobes'

centers of mass, and its gyration radius) are presented in Fig. 1.6A.

A B

Figure 1.6: (A) The modulation of the structure of the Mlc1p protein as a function of the

simulation time. ID's length (green); Distance between the NL center of mass and the CL

center of mass (red); Radius of gyration (black); (B) Solvent Accessible Surface Area

(SASA) of the structure of the Mlc1p-IQ4 complex as a function of the simulation time.

Page 51: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

38

As the ID of the Mlc1p protein experiences the most dramatic changes of

structure, we quantified its contraction during the simulation. The distance between the

first and the last alpha carbons of the ID of the protein was followed throughout the

course of the 12-ns simulation (Fig. 1.6A, green). The ID gradually shortens from 4.77 to

2.78 nm, demonstrating a significant conformational flexibility. The flexible ID

extensively curves, bringing the two lobes of the Mlc1p protein closer. The approach of

the lobes towards each other is also expressed by a decline of the distance between the

mass centers of the NL and the CL of the protein (Fig. 1.6A, red), from 4.11 to 2.99 nm.

This modulation reflects the transition from the extended complex structure at the crystal

to a more compact structure at the end of the simulation. The decrease of the distance

between the lobes occurs due to a twist of the ID. This can be manifested by calculation

of the gyration radius of the Mlc1p protein (Fig. 1.6A, black), which drops from 2.22 to

1.78 nm. The decrease in the radius of gyration demonstrates the decline in the overall

dimension of the Mlc1p protein. Overall, the conformational change at t ~ 3.2 ns leads to

a compaction of the protein. This can be observed by variations of the ID's length, the

distance between the N- and C- lobes' centers of mass, and the gyration radius.

The effect of the compaction of the protein on the Solvent Accessible Surface

Area (SASA) of the complex was calculated for both the hydrophobic and the hydrophilic

residues, and the combined SASA is presented in Fig. 1.6B. Both parameters exhibited a

significant reduction. The total SASA decreased with time, exhibiting a sharp drop (at t ~

3.2-ns) of the surface area within less than 100-ps from ~ 112 nm2 to ~ 102 nm

2. The time

point of this drop is associated with the major structural change of the complex. This

suggests that when one inspects the SASA criteria, taken as a whole, the structural

Page 52: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

39

change is quick rather than gradual.

The dynamics of the various parameters reflecting the contraction, as presented in

Fig. 1.6A, demonstrate that the structural changes are not an instantaneous event, but

rather a prolonged event stretching over ~ 1500-ps (~ 2.5-4 ns). Thus, the transition that

appears to be rapid when the RMSD (as displayed in Figs. 1.4A−C), the inter-helical

angles of the protein and the peptide (Figs. 1.5B and 1.5D), or the SASA (as displayed in

Fig. 1.6B) are examined, is actually a sum of individual steps as presented in Fig. 1.6A.

Some aspects of the overall compaction are abrupt, while other aspects are gradual. What

is more, after the major contraction takes place, the protein-peptide complex continues to

reshape for a long period. This shows that the post compaction relaxation is also a

complex, multi-component process.

1.2.4 The Forces that Stabilize the Refolded Model Structure of the Complex

To gain a comprehensive evaluation of the energetic aspects of the structural

modulation process, one needs to examine not only the protein-peptide complex but also

its interaction with the solvent, and the solvent-solvent interactions. However,

determination of parameters such as the solvent's cavitation energy, solvent's entropy,

solvation energy and the Long-Range electrostatic energies in a periodic system pose

both practical and conceptual difficulties when dealing with MD simulations. Therefore,

we shall limit our discussion only to certain specific aspects of the protein's energy, being

aware that only part of the overall system is analyzed. Accordingly, the variations of the

contraction of the protein in terms of VdW interactions (the Lennard-Jones component of

the potential energy), electrostatic interactions (Short-Range Coulomb potential), and

Page 53: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

40

entropy changes are presented below.

The Lennard-Jones (LJ) component of the potential energy interaction among the

residues of the Mlc1p protein, and the LJ component of the potential energy of the

protein-peptide interaction, are presented in Figs. 1.7A and 1.7B, respectively. The LJ

contribution to the potential energy, calculated for the interaction between the residues of

the Mlc1p protein, is changing rapidly, covering the time frame associated with the major

contraction event (Fig. 1.7A). The gain in stability, due to the better interaction between

hydrophobic residues of the Mlc1p protein, is ~ 200 kJ/mol. The contraction also

increases the contact between the protein and the IQ4 peptide, a process that contributes a

similar amount of stabilization but follows a different time course (Fig. 1.7B). Instead of

a rapid formation of LJ stabilization energy, the process is stretched over ~ 4-ns, slowly

gaining its final level. Thus, the slow evolution of the LJ stabilization, in a time frame

where there is hardly a change at the RMSD value, is attributed to minor rearrangements

of the side chains; these small motions make a large contribution to the stabilizing

energy. The LJ potential of the peptide with itself was constant in time (data not shown).

Figure 1.7: The contribution of the Lennard-Jones interactions to the stabilization of the

refolded model state of the Mlc1p-IQ4 complex. (A) The LJ component of potential

energy of the interactions among the residues of the Mlc1p protein; (B) The LJ

component of potential energy of the interactions between the Mlc1p protein and the

IQ4 peptide.

B A

Page 54: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

41

The contribution of the Short-Range (SR) electrostatic interactions to the stability

of the contracted complex is rather small (Fig. 1.8). However, it should be mentioned that

the mode of calculation is more qualitative than quantitative as described below.

The Coulomb SR intra-complex electrostatic interactions between the charges on

the protein and the peptide exhibit a transient nature. As the structural elements of the

system shift their relative position (2.5–4 ns time frame), there is a gradual decrease of

the overall electrostatic potential. Yet, when the side chains readjust (4–6 ns time frame),

the Coulomb SR potential increases and becomes less favorable. This leads to initial and

final levels of the overall electrostatic interactions being about the same.

Though the calculation of the Coulomb SR interactions was performed using a

cut-off of 12 Å, inspection of the variation of the Coulomb SR interactions, calculated for

longer cut-off distances (14, 16, 18, and 20 Å), revealed pattern similar in shape with that

presented in Fig. 1.8. As the amplitude of the electrostatic potential varies with the cut-off

distance, the values presented in Fig. 1.8 should not be taken as a quantitative

Figure 1.8: The electrostatic potential of the Mlc1p-IQ4 complex as a function of the

simulation time. The data is a summation of all Short-Range Coulomb potentials

between charges on the complex.

Page 55: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

42

representation of the electrostatic interactions. Apparently, the qualitative description of

the process is independent of the cut-off used. Thus, we conclude that the overall

electrostatic potential responses transiently to the structural modification process,

relaxing after few nano-seconds to its original value. According to our simulation results,

the stabilization of the Mlc1p-IQ4 complex is the sum of VdW interactions and a

transient variation of the electrostatic potential of the system. However, the VdW

interactions seem to be the dominant stabilizing force operating on the complex, while

the electrostatic interactions exhibit reversible changes. Apparently, the stability gained

by the rearrangements of the side chains, located both on the Mlc1p protein and on the

IQ4 peptide, follows the major changes of the main structural elements of the protein.

While the compaction event is brief, the LJ forces leading to the refolding operating for a

relatively long time.

In addition to the potential energy and electrostatic terms, one should also

consider the entropic changes associated with the compaction of the complex's structure.

The refolding of the protein and the tighter packing of its side chains reduced the freedom

of motion of the interacting residues, thus affecting the entropy of the system. To

estimate the contribution of this term, we calculated the entropy of the protein-peptide

complex (prior to and following the compaction event). These calculations indicated that

the entropy of the complex decreased, upon compaction, from an average value of 26.02

± 0.16 kJ mol-1

K-1

(calculated in time frames of 1000-ps between 1000 and 3000 ps) to

25.31 ± 0.29 kJ mol-1

K-1

(calculated in time frames of 1000-ps between 4000 and 12000

ps). This is equivalent to an entropic destabilization of the refolded conformation by ~

210 kJ/mol (300 K). The decrease of the complex’s entropy is due to formation of a

Page 56: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

43

tighter, more compact conformation of the refolded protein. However, the entropy of the

solvent, which cannot be directly obtained from the MD simulation, is likely to increase

upon the compaction event due to release of water molecules from the protein's surface to

the bulk.

To account for the release of water molecules from the solvation layer of the

Mlc1p protein, its first solvation layer was defined as a 2.8-Å-thick shell around it. The

number of water molecules found within the solvation layer, as it varies with time, is

presented in Fig. 1.9.

It can be seen that the compaction process is accompanied with a sharp decrease

in the number of water molecules that are in the VdW contact with the protein. Some 20

water molecules are released from the interface of the protein to the bulk at a time point

associated with the major structural modification event. Besides the abrupt fast reduction,

a slow trend of a decrease in the number of water molecules in the solvation layer

throughout the time is also observed. The detachment of the water molecules from the

Water Molecules at the First Hydration Shell

Figure 1.9: The number of water molecules at the first solvation layer (2.8Å) of the

Mlc1p protein as a function of the simulation time.

Page 57: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

44

hydration shell of the solute and their rejoining to the solvent, produce an entropic

advantage for the new, refolded model structure of the protein-peptide complex.

Apparently, the release of water molecules from the solvation layer compensates for the

reduction at the protein's entropy, thus assisting in the overall free energy balance.

1.2.5 The Interactions Between Residues During the Structural Change of the

Complex

The detailed interactions between individual residues of the Mlc1p protein and the

IQ4 peptide were followed at three stages: at the crystal structure, during the simulation

but before the compaction event (from the beginning of the simulation till 2-ns), and at

the last half of the simulation when the new compact configuration was already formed

(6−12-ns). For each state we listed the residues on the peptide which are in contact (less

than 4 Å) with the protein. In the case of the crystal structure, we searched for present

contacts between residues at the static presentation. For the simulated solution structures,

either before or after the compaction, we chose to present only contacts between residues

that were present in at least 80 % of the snapshots. This restriction eliminated transient

interactions that make a negligible contribution to the stabilization of the compacted

complex. The results are summarized in Fig. 1.10.

At the crystal structure, 31 contacts between the IQ4 peptide and the Mlc1p

protein are found (Fig. 1.10A). 24 of these contacts involve the CL of the protein (green),

and the rest are with residues of ID (red). During the MD simulation, the complex

underwent an initial relaxation and assumed the initial solution model structure, in which

the number of contacts between the IQ4 peptide and the CL was reduced to 18 (Fig.

Page 58: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

45

1.10B), 3 of them not being present at the crystal structure. In this intermediate state, the

number of contacts with the ID was reduced to 5.

Figure 1.10: Contacts between residues of the IQ4 peptide with side chains of the

Mlc1p protein. The residues of the protein are marked by their location for identification

of their presence. The color code: NL (blue), ID (red), CL (green). (A) Contacts (less

than 4 Å) between residues of the Mlc1p protein and the bound IQ4 peptide obtained

from the crystal structure PDB 1M46; (B) Contacts (less than 4 Å) between residues of

the Mlc1p protein and the bound IQ4 peptide obtained from the simulation at t <= 2 ns.

Only contacts persist at least 80 % of the examined period are presented; (C) Contacts

(less than 4 Å) between residues of the Mlc1p protein and the bound IQ4 peptide

obtained from the simulation at t >= 6 ns. Only contacts persist at least 80 % of the

examined period are presented. In Figs. 1.10B and 1.10C, contacts that are common for

the solution states and the crystal structure, are marked by bold letters.

Page 59: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

46

After the structural transition of the complex's structure, the number of contacts

between the peptide and the protein increased to 42 (Fig. 1.10C). 25 of these contacts are

with the CL, 4 new contacts are with the NL (colored in blue), and 13 with the ID (8 of

which were not present at the crystal structure). It is of interest to point out that the

residues 21−23 of the IQ4 peptide, which at the crystallized complex made no contacts

with the Mlc1p protein, interacted with its CL during the simulation.

Each of the interactions summarized in Fig. 1.10 had its own time evolution. Most

of the contacts appear in conjunction with the major structural transition event, yet in

some cases the interaction appears to be composed from more than two states. Hence, in

order to examine more closely the interaction between the amino acids formed the

contacts presented in Fig. 1.10, the minimal distances between all of the pairs were

followed as a function of time. Two representative examples are shown in Fig. 1.11.

As presented in Fig. 1.11A, the interaction of R4 (of the IQ4 peptide) with N36

(located at the NL of the Mlc1p protein) has at least two states. Before the compaction

event, the distance between the residues was larger than 1 nm; after that, it fluctuated

wildly since there was no interaction between the residues. At t ~ 3.2-ns, the structural

deformation of the complex' model structure imposed a stable separation of ~ 0.7 nm

between the residues. Yet, at t ~ 5-ns, a new structural organization took place and atoms

of the two residues almost reached a contact of their VdW radii (d = 2.5 Å). A different

scenario was observed concerning the interaction of Q10 of the IQ4 peptide (a conserved

residue of the IQ family), with residue F92 (located at the ID of the Mlc1p protein),

presented in Fig. 1.11B. In this case, the approach of the two residues towards each other

is a slow, gradual process, which resembles the slow evolution of the LJ stabilization of

Page 60: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

47

the complex (Fig. 1.7B). Apparently, the initiation of the approach precedes the structural

deformation event, yet proceeds for several ns.

1.2.6 MD Simulation of the Mlc1p-IQ4 Complex at 400 K

CaM and CaM-like proteins are rather stable and, in some cases, their purification

employs a brief boiling of the proteins’ mixture (137-139). Most of the proteins, unlike

the CaM and CaM-like proteins, cannot stand this treatment and precipitate. To check

whether the elevated temperature may cause further structural changes of the protein, and

detect changes not observed at 300 K, we repeated the simulation at 400 K.

According to our simulation results, the elevated temperature does not affect the

refolding mechanism of the protein-peptide complex, but rather its rate. Fig. 1.12A

depicts the structure of the Mlc1p-IQ4 complex after 12-ns of simulation at 400 K., while

superposition between the structures of the complex at both temperatures is presented on

Fig. 1.12B. Inspection of the model structure of the complex at the end of the simulation

A B

Figure 1.11: Representative examples of contacts between the IQ4 peptide and the

Mlc1p protein as a function of the simulation time. (A) Minimal distance between

residue R4 of the peptide and residue N36 of the protein; (B) Minimal distance between

residue Q10 of the peptide and residue F92 of the protein.

Page 61: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

48

Figure 1.12: (A) Cartoon diagram of the Mlc1p protein bound to the IQ4 peptide (PDB

file 1M46). Simulated model structure, after 12-ns of MD simulation at 400 K in the

presence of water and ions. The NL, the ID, the CL and the IQ4 peptide are shown in

blue, red, green and yellow, respectively. The domains of the protein are defined as

residues 1−59, NL; 60−92, ID; and 93−148, CL. Structure is presented with the same

orientation as in Fig. 1; (B) Superposition of the cartoon diagrams of the simulated model

structures of the Mlc1p-IQ4 complexes at 300 K and 400K. The Mlc1p protein and the

IQ4 peptide at 300 K are shown in black and gray, respectively; whereas the Mlc1p

protein and the IQ4 peptide at 400 K are shown in red and orange, respectively. (C) The

backbone atom RMSD of the different domains of the Mlc1p protein as a function of the

simulation time, for the MD simulation at 400 K. The NL, ID and CL of the Mlc1p

protein are shown in black, red and green, respectively.

A

C B

at 400 K reveals similarity with the 300 K simulation: 1. The Mlc1p protein lost its

extended configuration, exhibiting a compact configuration; 2. The NL retained two of its

helices and a good contact with the N-terminal section of the peptide; 3. The ID folded in

a mode similar to that gained at 300 K; 4. The CL retained its three helices while

engulfing the IQ4 peptide; 4. At the elevated temperature the IQ4 peptide lost a fair

fraction of its alpha helix structure. Yet, as can be seen from the superimposed structures,

it is located between the lobes of the protein at both temperatures.

Page 62: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

49

The mechanism of the conformational transition is basically comparable at the

two temperatures with a clear acceleration at the higher temperature. The RMSD values,

calculated for the structural elements of the protein: the NL, the CL and the ID (marked

in black, green and red, respectively), are presented in Fig. 1.12C. The quick rise of the

RMSDs of the structural elements of the protein within the first nano-second of the

simulation, especially that of the ID, illustrates the fast alternation of the protein's initial

structure when the temperature is elevated. Evidently, the RMSDs of the MD simulation

at 400 K increased at the beginning of the simulation to levels achieved only after ~ 3.2-

ns by the MD simulation at 300 K. A careful analysis of the trajectories at both

simulations reveals high similarity. Moreover, since also the final level of RMSDs is akin

at both simulations, we suggest that the modulation's speed of the complex' structure is

temperature-dependent.

1.2.7 Summary

In the present study we investigated, through MD simulations, the dynamics of a

complex between a CaM-like protein and its target peptide, the IQ4 peptide. This

complex, which is a Ca+2

-independent stable structure, underwent a complicated

structural transition process. The process appeared to be initiated by fluctuations of the

model structure in the multi-dimensional parameter space, with no clear triggering event.

Once the proper situation was attained, the reaction became unidirectional and the system

gained stability in few steps. Only 15 out of 31 contact points detected at the crystal

structure were retained in the solution model structure, while 27 new interactions were

established. The structural modification of the complex involved bending of the angles

Page 63: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

50

between the alpha helices of the Mlc1p protein, and breaking of the alpha helical

structure of the IQ4 peptide into two, almost perpendicular sections. Some of the

structural changes coincide with the main conformational transition event. This event was

preceded by preparatory events that proceed in time, and followed by some “after-

shocks” that helped to seal and stabilize the refolded model structure.

It is likely that the progression of the simulation, even before the major structural

transition event, included stepwise preparatory steps. Since the MD simulation at 400 K

yielded a final conformation similar to the one at 300 K, we suggest that the pattern of the

structural transition is embedded as an inherent instability of the crystal structure of the

Mlc1p-IQ4 complex. Once the restraints imposed by the crystallization conditions were

released, the system progressed into a stable model solution structure at a temperature-

dependant rate.

The refolding of the protein-peptide complex is composed of both fast, sub pico-

seconds events, and also slow events that take up a few ns. The stabilization of the

compact state is mostly hydrophobic in nature, although electrostatic interactions also

take part in this process. The main deformation is probably directed by introductory

events, followed by the rapid process of compaction. The continuing stabilization of the

protein-peptide complex can be considered as optimization of the packing of the refolded

model structure, minimizing the LJ potential energy by local shuffling of atoms at the

Mlc1p-IQ4 complex contact surface. These proceeding events help to “close the deal”,

thus completing the process till a new configuration is formed.

The number of residues involved in the interaction between the IQ4 peptide and

the Mlc1p protein is much larger than the number of conserved residues that define the

Page 64: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

51

IQ motif. What is more, the 4 conserved motif residues L9, Q10, R14 and R20 contribute

only 15 out of the 42 contact points that stabilize the refolded complex. Apparently, the

plethora of residues that participate in the interaction between the Mlc1p protein and the

IQ4 peptide allows some tolerance to a single point mutation of a given site.

We wish to note that the proposed simulated model solution structure of the

Mlc1p-IQ4 complex presented in this study differs from its crystalline structure. Our

suggested solution model structure resembles structures of Holo-CaM and Apo-CaM in a

complex with peptides, obtained by NMR (47) and crystallography studies (48-54).

Moreover, the crystal structure of the Mlc1p protein in a complex with the IQ2 peptide

(28, 56) is similar to the simulated model structure of the Mlc1p-IQ4 complex. However,

experimental methods, like Dynamic Light Scattering or NMR, may be carried out to

verify the proposed simulated model solution structure of the Mlc1p-IQ4 complex.

Nevertheless, this kind of experiments is beyond the scope of the present study.

Page 65: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

52

2. Comparison Between the Mlc1p-IQ2 and the Mlc1p-IQ4 Complexes

In this chapter of the thesis, the MD simulations of the Mlc1p-IQ2 and the Mlc1p-

IQ4 are compared. The crystal and the simulated structures are judged against each other,

the dynamic properties of both complexes are described, and a comprehensive energetic

detailed analysis using the MM-PBSA approach is presented.

2.1 Crystallographic and Simulated Structures Comparison

The crystallographic structures and those obtained by the simulations of the

Mlc1p protein with the IQ2 and IQ4 peptides are presented at the previous chapter of the

thesis. At the crystal state of the Mlc1p-IQ2 complex, the protein assumes a compact

state; while at the crystal state of the Mlc1p-IQ4 complex, the Mlc1p confers to an

extended configuration. These distinctive states can be attributed either to complex-

specific interactions, to the different crystallization conditions of each complex (28, 56),

or to a combination of both. A close examination of the crystal and simulated structures

of both protein-peptide complexes (Fig. 2.1) reveals that, following the simulations, the

protein’s model structures are more similar than in their crystalline states.

The calculated backbone atoms RMSD between the Mlc1p proteins found at the

Mlc1p-IQ2 and the Mlc1p-IQ4 crystal structures (Figs. 2.1A and 2.1C), is as large as

1.296 nm. However, this value drops to 0.959 nm when calculated between the solution

model of the protein after 12-ns simulations (Figs. 2.1B and 2.1D). A value of 0.959 nm

may still seem to be quite high, but it should be noted that, due to the shape of the

protein, any attempt to align two of its structures is expected to result in a relatively high

RMSD. Therefore, the decline of the RMSD from 1.296 nm to 0.959 nm is structurally

Page 66: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

53

notable, as seen in Fig. 2.1. On the top of the RMSD calculation, other structural

indicators (such as length of the ID, distance between the lobes center-of-mass and

gyration radius) were calculated for the crystal and the simulated structures of the protein

(data not shown). All these point out that the solution configurations of the Mlc1p protein

are more similar than its crystalline ones. Additionally, it should be mentioned that the

simulated model structures of the protein acquire special conformational features not

present in either of the crystals. Besides exhibiting a common compact form of the

simulated protein, during the simulations a new kink appears at both of its model

structures (Figs. 2.1B and 2.1D, see arrow). This kink is missing from the crystal

structures of neither of the Mlc1p-IQ complexes.

A CB D

Figure 2.1: Cartoon diagrams of the crystal and the simulated structures of the Mlc1p

protein when it binds the IQ2 peptide (PDB 1M45), and the IQ4 peptide (PDB 1M46).

The NL (residues 1−59), the ID (residues 60−92), the CL (residues (93−148), and the IQ

peptides are shown in blue, red, green and yellow, respectively. Both crystal structures,

and both simulated solution structures, are presented with the same orientation, where

the N-lobes are structurally aligned. (A) The crystal structure of the Mlc1p-IQ2

complex; (B) The simulated model structure of the Mlc1p-IQ2 complex after 12-ns

simulation; (C) The crystal structure of the Mlc1p-IQ4 complex; (D) The simulated

model structure of the Mlc1p-IQ4 complex after 12-ns simulation. The black arrows in

frames B and D mark the kink of helix D.

Page 67: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

54

Though both simulated model structures are characterized by an overall similarity,

there is still a difference between them, as the CL of the protein points towards opposite

directions. The position of the CL of the Mlc1p protein in respect to its NL may be

expressed by calculation of the dihedral angle between the planes defined by the two

lobes and the ID. We found that, at the end of the simulations, the dihedral angle at the

Mlc1p-IQ2 simulated model structure is 128.54°, while the dihedral angle at the Mlc1p-

IQ4 simulated model structure is significantly smaller, 72.1°. The difference represents

two binding modes and orientations of the IQ peptides in respect to the protein, and two

different conformations of the latter. Apparently, various Mlc1p-binding partners may

affect and dictate its structure. The Mlc1p protein is very flexible, enabling it to wrap

around the IQ peptides in different ways. Its tolerance and adaptivity towards different IQ

peptides make it an appropriate candidate to bind a variety of target helical peptides.

Evidently, this unique property enables it to bind six IQ peptides of the LCBD of myosin

V, each distinguished by a unique sequence.

2.2 Structural Evolution of the Simulated Mlc1p Protein at the Protein-Peptide

Complexes

A dynamic comparison of the trajectories of the Mlc1p protein in both simulations

is shown in Fig. 2.2. The RMSD of the C-alpha atoms of the Mlc1p protein obtained

from the Mlc1p-IQ4 simulation's trajectory is presented in relation to the C-alpha atoms

of the Mlc1p protein obtained from the Mlc1p-IQ2 simulation's trajectory. The two-

dimensional matrix representation exemplifies the time evolution of the protein's

configurations as a function of the simulations time. A color code is used for visualizing

Page 68: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

55

Figure 2.2: Matrix representation of mutual C-alpha atoms' RMSD of the Mlc1p protein

obtained upon comparison of the two simulations. The RMSD values of the Mlc1p protein

for the Mlc1p-IQ4 simulation's trajectory were calculated in relation to those of the Mlc1p

protein obtained from the Mlc1p-IQ2 simulation's trajectory and vice versa. The values

are given by color codes; where blue and red represent high and low similarity,

respectively.

how the two model structures of the Mlc1p protein, in the Mlc1p-IQ protein-peptide

complexes, approach a common compact shape during the simulations. The pattern of the

colors exhibits progressive yet reversible changes, indicating that both model structures

are fluctuating, and the similarity between them varies. The mutual evolution of the

conformational changes, as shown in Fig. 2.2, suggests that the convergence and settling

of the Mlc1p protein's model structures towards relatively similar compact configurations

can be divided into three phases: a relaxation phase (from the beginning of the simulation

till ~ 3.2-ns), a progression phase (from ~ 3.2 till ~ 10-ns), and a quiescent phase (from ~

10-ns till the end of the simulations).

Page 69: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

56

The relaxation phase is represented by the yellow-reddish color laying at the

bottom of the figure, stretching over its full width and extending up to ~ 3.2-ns mark of

the ordinate. At this phase, the Mlc1p protein at the Mlc1p-IQ4 complex responds to the

absence of the packing forces present at the crystal lattice. Hence, the model structures

sampled by the protein from both simulations are still remarkably different, each

remaining close to its crystal form. At the progression phase, the Mlc1p protein at the

Mlc1p-IQ4 simulation undergoes a major conformational change (140), rendering it more

similar to that of the Mlc1p-IQ2 simulation. This newly gained configuration is stabilized

by hydrophobic interactions, where minor rearrangements of the side chains contribute to

a relatively slow progression stabilizing process. The refolded conformation of the Mlc1p

protein obtained from the Mlc1p-IQ4 simulation becomes more similar to the model

structure of the Mlc1p protein obtained from the Mlc1p-IQ2 simulation. This tendency

increases along the ordinate as seen by the shift from the yellow-greenish to the green-

bluish colors. Finally, at the quiescent phase, the new model structure of the Mlc1p

protein obtained from the Mlc1p-IQ4 simulation is already stabilized. The higher degree

of similarity between the protein’s model structures, as reflected by the smaller RMSD

values one with respect to the other, is observed at this phase (represented by the bluish

hue). Evidently, as the simulations progress, the model structures of the proteins become

more similar in a time-dependent manner. Thus, the final MD-derived solution model

structures of the protein are more similar than its crystal structures states. Furthermore,

these final MD-derived solution model structures of the protein resemble also compact

configurations of the CaM protein presented at crystal structures of Holo-CaM with target

peptides (48-54, 141).

Page 70: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

57

It is of interest to point out that the model structures of the protein at both

simulations do not co-evolve in parallel. The model derived from the Mlc1p-IQ2

simulation experiences limited changes, whereas that derived from the Mlc1p-IQ4

simulation assumes a significant modification. Thus, the Mlc1p' model structures evolve

at different rates towards a more similar configuration.

Carrying out the same analysis only for the ID of the Mlc1p protein (data not

shown) reveals a pattern resembling that of the whole protein. The structure of the ID of

the Mlc1p protein, obtained from the Mlc1p-IQ2 simulation, is almost invariable, while

that obtained from the Mlc1p-IQ4 simulation evolves with time. A high degree of

similarity is obtained after ~ 10-ns, as observed for the whole protein.

2.3 The Root Mean Square Fluctuation (RMSF) of the Mlc1p Protein at the Protein-

Peptide Complexes

In order to further analyze the trajectories of the Mlc1p protein at both

simulations, we computed the standard deviation from the RMSD for each of its residues,

i.e. their Root Mean Square Fluctuations (RMSF). Fig. 2.3A presents the RMSF of the

Mlc1p protein at the Mlc1p-IQ2 structure simulation (solid line), and at the Mlc1p-IQ4

structure simulation (dashed line). Residues of the Mlc1p protein, that comprise alpha

helices at the crystalline configurations of the Mlc1p-IQ protein-peptide complexes, are

shown as bold horizontal bars parallel to the abscissa. The RMSF curves reveal the

different behavior of the protein at both simulations, characterizing the mobility of each

of its residues during the MD runs.

In general, the structural sections of the protein, namely its eight alpha helices, are

Page 71: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

58

more rigid and confined and tend to be less flexible than its other sections (e.g. residues

39−50).

Correspondingly, the protein's unstructured sections show an increased motility.

Thus, at both simulations, the RMSF data indicate large fluctuations of segments

belonging to loops that connect secondary structure elements (e.g. residues 14−20 and

128−137), as well as of residues located at the edges of the alpha helical sections (e.g.

residues 90−92, and 123−125). While these features are common to both complexes, the

Mlc1p protein exhibits a different mobility when it binds the IQ2 or the IQ4 peptides as

the absolute values of the RMSF differ. The RMSF curve of the Mlc1p protein at the

Mlc1p-IQ4 simulation consistently reveals a higher extent of motion than that of the

protein at the Mlc1p-IQ2 simulation. Despite these different degrees of motion, the

A B

Figure 2.3: The Root Mean Square Fluctuation (RMSF) as a function of the residue

number of the Mlc1p protein. The RMSF was calculated for the backbone atoms of the

Mlc1p protein for each residue at both simulations. The solid line represents the RMSF of

the Mlc1p protein at the simulation of the Mlc1p-IQ2 structure, while the dashed line

represents the RMSF of the Mlc1p protein at the simulation of the Mlc1p-IQ4 structure.

The bold horizontal bars, drawn in parallel to the abscissa, represent the alpha helices that

the Mlc1p protein comprises. The RMSF of both MD trajectories is presented for the

entire simulations time (A), and for the time frame t = 10-ns till t = 12-ns (B).

Page 72: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

59

RMSF of some sections of the protein is correlated (for example, the RMSF of the NL of

the protein, between both simulations, is correlated, with R2 = 0.69). However, other

sections of the protein do not exhibit such correlation. The variation in the correlations

between the structural domains of the protein implies that the main differences regarding

the dynamics of the protein at both simulations are located at its ID and CL.

It can be argued that the relatively high RMSF of the protein at the Mlc1p-IQ4

simulation is due to its structural modification process, by which it refolds, and that

process is still going-on. However, detailed structural and energy analysis presented in

our previous publication (140) suggests that its major conformational change had been

completed in the course of the simulation. In order to observe the behavior of the Mlc1p

protein at the stable period at both simulations, we repeated the RMSF analysis during the

time frame 10-ns till 12-ns (Fig. 2.3B). The RMSF curves reveal a pattern similar to that

calculated for the entire simulations time, although the extent of the fluctuations is

smaller. Nevertheless, the RMSF curves resemble each other more than those obtained

for the entire simulations' time. This indicates as well that the Mlc1p protein at the

Mlc1p-IQ4 simulation had already experienced its structural modification and may not

significantly evolve at the discussed time frame. We wish to note that, even at this time

frame, we found that the RMSF of the protein derived from the Mlc1p-IQ4 simulation is

still higher than that obtained by the protein from the Mlc1p-IQ2 simulation. This is in

accord with the observation that more contacts are involved in the Mlc1p-IQ2 interaction

than in the Mlc1p-IQ4 interaction (See Fig. 2.5, below). Thus, the protein at the latter

simulation is not strongly retained to its position and may be more mobile.

Page 73: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

60

2.4 The Electrostatic Field Around the IQ Peptides

The Mlc1p-IQ protein-peptide complexes are composed of the same protein, and

similar, although not identical, peptides. Yet, the protein-peptide complexes differ one

from the other by their crystalline structures (Figs. 2.1A and 2.1C), and to some extent by

the simulated model structures obtained by means of MD simulations (Figs. 2.2B and

2.2D). Evidently, the differences between the structures of the protein-peptide complexes

reflect the variation sequence of the bound IQ peptides. To account for the difference

between the peptides, we calculated the electrostatic field surrounding the IQ2 and IQ4

peptides. The volumes of the averaged (for the time frame t = 10 until t = 12-ns) positive

and negative electrostatic fields of both peptides are presented in Table 2, while in Fig.

2.4 we show the Coulomb cages of the IQ peptides at the last snapshot (t = 12-ns) of the

simulations.

Table 2

The Electrostatic Field Around the IQ Peptides

* Calculated by summation of the volumes of the positive and the negative Coulomb

cages.

The positive (transparent blue) and negative (transparent red) domains are drawn

where the electrostatic potential equals 1 kBT/e. The C-alpha traces of both peptides are

IQ2 Peptide IQ4 Peptide

Charge +2 +6

Positively-charged residues K-7, K-14, R-19, R-21 R-4, K-11, K-12, R-14,

K-15, K-18, R-20, K-23

Negatively-charged residues D-24, E-25 E-16, E-25

Volume of the positive

Coulomb cage (Å3)

4257.12 ± 264.48 10180.42 ± 426.12

Volume of the negative

Coulomb cage (Å3)

2080.56 ± 147.25 796.02 ± 144.94

Total volume of the

Coulomb cage* (Å

3)

6337.69 ± 395.11 10976.76 ± 498.02

Page 74: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

61

Figure 2.4: The electrostatic potential surface around the IQ2 peptide (A) and the IQ4

peptide (B) at the last snapshot (t = 12-ns) of both simulations. Both peptides are

presented in yellow with the same orientation, while their positive and negative residues

are drawn in blue and red, respectively. The Coulomb cages for the positive (transparent

blue) and negative (transparent red) domains are drawn at the distance where the

electrostatic potential equals 1 kBT/e.

A B

colored in yellow, while their positive and negative residues are shown in blue and red,

respectively. The potential field of the IQ peptides consists of two main lobes, one

positive and the other negative. However, although the IQ2 and IQ4 peptides are both

basic, alpha-helical, 25 amino acids long, they produce different electrostatic fields

around them. The volume of the positive Coulomb cage around the IQ2 peptide is

4257.12 ± 264.48 Å3, while the volume of its negative Coulomb cage is 2080.56 ± 147.25

Å3. The volume of the positive Coulomb cage around the IQ4 peptide is 10180.42 ±

426.12 Å3, while the volume of its negative Coulomb cage is 796.02 ± 144.94 Å

3.

Page 75: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

62

The differences between the electrostatic fields surrounding the peptides are

caused by variations in their local charge and charge distribution. The total charge of the

IQ2 peptide is Z = +2, while that of the IQ4 peptide is Z = +6. The peptides differ not

only in their total net charge, but in the distribution of the charges along them as well.

Four positive residues (K-7, K-14, R-19, R-21) are scattered along the IQ2 peptide, while

its negative residues (D-24, E-25) are concentrated at its C-terminal edge. The IQ4

peptide consists a series of four positive residues (K-11, K-12, R-14, K-15) located in its

middle section, whereas its other positive residues (R-4, K-18, R-20, K-23) are

distributed along it. The mid-cluster of positive charge contributes to the high volume

positive Coulomb cage bulb at the center of the peptide. The differences between the

electrostatic fields of the two peptides may suggest that electrostatic forces play a major

key role in the protein-peptide interactions.

2.5 The Protein-Peptide Interaction Free Energies

In order to analyze the energetics of peptide binding to the Mlc1p protein, the

various components of the interaction free energy of the two protein-peptide complexes

were evaluated during the last 2-ns for each simulation. This time frame corresponds with

the stable MD-derived model solution structures of the protein-peptide complexes at both

simulations, from which we can calculate the energy associated with protein-peptide

interaction. The detailed results of the energetic analysis are presented in Table 3.

The analysis was based on the MM-PBSA approach (94), in which the interaction

free energy, (∆Ginteraction), for each of the complexes is composed of three energetic terms:

The molecular mechanics energy term (⟨∆EMM⟩), the solvation energy term (⟨∆Gsolvation⟩),

Page 76: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

63

and the solute entropic contribution (T∆S). The first term includes internal (⟨∆Eint⟩), VdW

(⟨∆EVdW⟩) and electrostatic (⟨∆Eelectrostatic⟩) components. The second term consists of

electrostatic (⟨∆Gpolar,solvation⟩) and nonpolar (⟨∆Gnonpolar,solvation⟩) contributions. Solute

entropies were determined at the last snapshots of the MD trajectories. Note that the

internal component of the molecular mechanics energy, (⟨∆Eint⟩), is set per definition as

zero and thus cancels out, making no contribution at all (101, 103, 104).

Table 3

Components of the Mlc1p-IQ Interaction Free Energy

Energies are presented in kJ/mol. The calculations present the average values

obtained from t = 10 till t = 12-ns for both Mlc1p-IQ simulations. ⟨ ⟩ denotes an average

over a set a snapshots along an MD trajectory. Atomic charge and radii values were taken

from the PARSE parameter set (142).

Definitions of the energetic components are as follows: (⟨∆Eelectrostatic⟩),

electrostatic molecular mechanics energy; (⟨∆EVdW⟩), VdW molecular mechanics energy;

(⟨∆EMM⟩), total molecular mechanics energy defined as ⟨∆EMM⟩ = (⟨∆Eint⟩ + ⟨Eelectrostatic⟩ +

⟨∆EVdW⟩); ⟨∆Gpolar,solvation⟩, electrostatic contribution to the solvation energy calculated by

the PB approach; ⟨∆Gnonpolar,solvation⟩, nonpolar contribution to the solvation energy;

⟨∆Gsolvation⟩, total solvation energy defined as (⟨∆Gpolar,solvation⟩ + ⟨∆Gnonpolar,solvation⟩); T∆S,

solute entropic contribution; and (∆Ginteraction), total free energy of interaction defined as

(⟨∆EMM⟩ + ⟨∆Gsolvation⟩ – (T∆S)).

On combining the (⟨∆EMM⟩) with the (⟨∆Gsolvation⟩) and the (T∆S) terms, we end

up with interaction free energy, (∆Ginteraction), for the complexes' formation. The estimated

interaction free energy of the Mlc1p-IQ2 and the Mlc1p-IQ4 complexes is ~ -560 kJ/mol

Mlc1p-IQ2 Complex Mlc1p-IQ4 Complex

⟨∆Eelectrostatic⟩ -590.77 ± 49.96 -1534.02 ± 84.28

⟨∆EVdW⟩ -664.18 ± 25.2 -547.7 ± 25.88

⟨∆EMM⟩ -1254.95 -2081.72

⟨∆Gpolar,solvation⟩ 678.25 ± 41.85 1594.88 ± 78.84

⟨∆Gnonpolar,solvation⟩ -58.69 ± 5.59 -53.98 ± 5.48

⟨∆Gsolvation⟩ 619.56 1540.9

-T∆S 75 372

∆Ginteraction -560 -169

Page 77: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

64

and ~ -169 kJ/mol respectively, consistent with their observed stability. These data

represent a balance between enthalpy and entropy in which, according to our calculations,

the complexes' formation is an enthalpically driven process and is entropically

unfavorable. The favorable formation of both Mlc1p-IQ complexes is driven by the

electrostatic (⟨∆Eelectrostatic⟩) and the VdW (⟨∆EVdW⟩) terms of the molecular mechanics

energy and the nonpolar component of the solvation energy (⟨∆Gnonpolar,solvation⟩).

Of particular interest is the total solvation energy, (⟨∆Gsolvation⟩), composed of

polar (⟨∆Gpolar,solvation⟩) and nonpolar (⟨∆Gnonpolar,solvation⟩) terms. The total solvation energy

is unfavorable at both complexes (619.56 kJ/mol for the Mlc1p-IQ2 complex, and 1540.9

kJ/mol for the Mlc1p-IQ4 complex). Thus, considering the solvation energy, it appears

that the protein-peptide complexes would rather not be formed at all. Yet, the molecular

mechanics energy component of the interaction energy strongly favors the complexes

over the unbound molecules.

Electrostatic interactions were assumed to play a dominant major role in the

interaction between CaM and target peptides (49, 143-145). This view was emerged from

X-ray structures of complexes showing close proximity between the negatively charged

CaM and positively charged peptides (47-54, 141). Accordingly, electrostatic interactions

have been anticipated to be significant in the Mlc1p-IQ systems, as the Mlc1p-IQ

complexes also present a close distance between a negatively charged protein and highly

positively charged peptides. It is of high importance to consider the electrostatic

component of the molecular mechanics energy, (⟨∆Eelectrostatic⟩), together with the

electrostatic contribution to solvation, (⟨∆Gpolar,solvation⟩), when examining the role of

electrostatics in the protein-peptide complexes formation. At both protein-peptide

Page 78: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

65

complexes, the nature of the electrostatic interactions is similar: The molecular

mechanics electrostatic term per se favors the bound state of the complexes (⟨∆Eelectrostatic⟩

< 0), while the electrostatic PB solvation energy favors the unbound state of the protein-

peptide complexes (⟨∆Gpolar, solvation⟩ > 0). As the latter is dominant (|⟨∆Gpolar,solvation⟩| >

|⟨∆Eelectrostatic⟩|), their sum, representing the total electrostatic energy, opposes the

formation of the protein-peptide complexes. Thus, the positive solvation energy

electrostatic term penalty paid by the electrostatics of solvation is not completely covered

by favorable electrostatic interactions within the resulting protein-peptide complexes.

Evidently, the same phenomenon was also demonstrated by numerous studies (101, 103,

104, 128, 146-148), in which the total electrostatics between two interacting molecules

unfavors their bound state over the unbound due to intense solvation forces. Interestingly,

the electrostatic energy terms (⟨∆Eelectrostatic⟩) and (⟨∆Gpolar,solvation⟩) are more prominent in

the Mlc1p-IQ4 complex compared to the Mlc1p-IQ2 complex by approximately a factor

of three, which is proportional to the charge of the peptides (The charges of the IQ2 and

IQ4 peptides are +2 and +6, respectively). Hence, the differences in charges between the

protein and the peptide are more substantial in the Mlc1p-IQ4 complex, and consequently

its electrostatic terms are more profound.

It is not surprising that, at both complexes, the entropic term does not support the

interaction between the protein and the peptide. The Mlc1p protein, as well as the IQ

peptides, are characterized by higher entropy in their unbound states. In its free state, the

lobes of the Mlc1p protein can tumble more or less independently of one another,

constrained only by the ID. However, in the bound state, its lobes interact with the

peptide and hence are relatively at fixed positions. Similarly, the free IQ peptides may

Page 79: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

66

also acquire more structural freedom when they are not constrained by the protein.

The energy calculations agree fairly well with isothermal titration calorimetry

(45, 149) and NMR relaxation (150) experiments, in which it was found that the binding

of a peptide to Holo-CaM is favored by enthalpy and opposed by entropy. Our results are

also in accord with an MD study of Holo-CaM complexed with a target peptide

suggesting that the protein-peptide free energy is enthalpy-dependent and not entropy-

dependent (92). As pointed out by these authors, identification of changes in entropy (45,

150) or enthalpy (92) upon complex formation is fraught with difficulty. Therefore, the

qualitative agreement between our calculations regarding the Mlc1p-IQ complexes

obtained by computer simulations, and experimental and theoretical calculations

regarding Holo-CaM peptide complexes, is encouraging and promising.

Finally, it must be stressed that the values given in Table 3 are model-dependent

and reflect all of the approximations implemented in the MM-PBSA formalism. Thus, the

numerical values should be taken as indicative, representing qualitative trends rather than

actual quantitative ones. The consistency of the results with the observed stability of the

complexes supports the acceptance of this mode of calculation as a proper representation

of the operating forces.

Owing to the opposite charges of the protein and the peptide, the electrostatic

interactions may serve as the initial driving force for long-range molecular recognition

between the Mlc1p protein and the IQ peptides. On the other hand, the highly charged

protein and peptides strongly interact with the solvent, leading to intensive solvation

forces. Upon formation of the protein-peptide complexes, the Mlc1p protein approaches

into a close vicinity to the IQ peptides. This desolvation process, which is unfavorable, is

Page 80: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

67

accompanied by a release of water molecules from their interfaces, replacing solute-

solvent interactions by intra-complex interactions. The unfavorable change in the

electrostatics of solvation is mostly, but not fully, compensated by the favorable

electrostatic charge-charge interactions within the resulting Mlc1p-IQ complexes. The

close interaction of the Mlc1p protein with the IQ peptides is grossly mediated by the LJ

interactions, whereas their opposite highly charged surfaces contribute to their initial

attraction. The major role played by the molecular mechanics VdW interactions

demonstrates how inter-residues contacts, where the tight fitting of the surface atoms

occurs, contribute to the LJ stabilization energy term and consequently to the stability of

the complexes. In conclusion, electrostatic interactions seem to operate mostly during the

long-range attraction between the protein and the peptides before the complexes are

formed. Once protein-peptide contact occurs, VdW and non-specific hydrophobic

interactions stabilize the Mlc1p-IQ structures, whereas the contribution of salt bridges is

relatively negligible.

2.6 The Contacts Between the Protein and the Peptides

The detailed interactions between individual residues of the Mlc1p protein and the

IQ peptides were followed during the last half of the simulations (6−12-ns). Residues of

the peptides, which were in a contact (less than 4 Å) with the Mlc1p protein, were listed.

The rapid motion of the residues during the simulations led to numerous encounters, but

most of them were temporary and made a marginal contribution to the protein-peptide

interaction. To account for that, we selected only those contacts, between residues of the

IQ2 or the IQ4 peptides with the Mlc1p protein, which were present in at least 80 % of

Page 81: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

68

the snapshots. These lasting interactions are presented in Fig. 2.5.

Altogether, 51 contacts between the peptide and the protein were found at the

Mlc1p-IQ2 simulation. 15 of these contacts involve the NL of the protein (blue), 14 the

ID (red), and 22 the CL (green). Similarly, at the Mlc1p-IQ4 simulation, 42 contacts were

found between the peptide and the protein (140). Only 4 of these contacts involve the NL

of the protein (blue), 13 the ID (red) and 25 the CL (green). It is of interest to point out

that 20 residues of the Mlc1p protein interact with both IQ peptides (underlined in Fig.

2.5), 12 of its residues interact only the IQ2 peptide, whereas 7 of its residues exclusively

interact the IQ4 peptide. Residues 24−25 of both IQ peptides made no contact with the

Mlc1p protein in neither of the simulations.

Figure 2.5: Contacts (less than 4 Å) between residues of the Mlc1p protein and the

bound IQ peptides obtained from the MD simulations at t >= 6-ns. Data regarding the

Mlc1p-IQ4 model structure simulation (140) are presented at the upper half of the

illustration, whereas data regarding the Mlc1p-IQ2 model structure simulation are

presented at its lower half. The IQ peptides, the NL of the protein, the ID of the protein,

and the CL of the protein are colored in black, blue, red, and green, respectively.

Residues of the Mlc1p protein that interact with both peptides are underlined. Only

contacts persist at least 80 % of the examined period are presented.

The IQ2 peptide has 4 positive residues (K-7, K-14, R-19, R-21), and 2 negative

Page 82: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

69

residues (D-24, E-25) located in its C-terminal edge. These negative residues repel the

negatively charged Mlc1p protein, and hence only one residue of the protein (I-9)

interacts with the last 5 residues of the IQ2 peptide. In comparison, the IQ4 peptide is

more positive. It has a substantial cluster of positive residues (K-11, K-12, R-14, K-15,

K-18, R-20, K-23) located at its mid- and C-terminal parts. These positive residues are

bound to the CL of the Mlc1p protein mainly through hydrophobic interactions (i.e. R-14

− L-116, R-20 − V-128), but two salt bridges are also present (K-11 − E-114, R-14 − E-

120).

The different binding modes of the two peptides are reflected by their different

regions of contact with the Mlc1p protein at each simulation. At the Mlc1p-IQ2

simulation, the mid- and C-terminal regions of the peptide (residues 13−25) interact with

8 residues of the NL and 12 residues of the CL of the protein. On the other hand, at the

Mlc1p-IQ4 simulation, the mid- and C-terminal regions of the peptide (residues 13−25)

do not interact with the NL of the protein as they are bound to 17 residues of its CL. The

fact that the mid- and C-terminal sections of the IQ2 peptide interact with the NL of the

protein, while those regions of the IQ4 peptide do not interact with it, is manifested by

the various orientations of the peptides when bound to the protein (Figs. 2.2B and 2.2D).

Although the protein and the peptides are highly charged in opposite charges, out

of the 51 contacts between the IQ2 peptide and the Mlc1p protein, only 2 involve

electrostatic interactions (K-7 − E-114, K-14 − D-28). Similarly, out of the 42 contacts

between the IQ4 peptide and the Mlc1p protein, only 2 involve electrostatic interactions

(K-11 − E-114, R-14 − E-120). Besides these few electrostatic interactions, all the other

protein-peptide interactions are hydrophobic in nature. Matter of fact, even the positive

Page 83: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

70

residues of the peptides interact with the positive residues of protein (K-14 − R-31, R-19

− R-147 for the Mlc1p-IQ2 protein-peptide complex; R-4 − R-31, R-14 − K-115, R-19 −

R-147, K-23 − R-147 for the Mlc1p-IQ4 protein-peptide complex). Considering the

repulsive force between positive charges, the interactions between the basic residues are

clearly hydrophobic. These data are consistent with the findings regarding the dominant

role played by the LJ component of the molecular mechanics energy and the nonpolar

component of the solvation energy in the stabilization of the protein-peptide complexes

(Table 3).

2.7 Summary

The present study provides a fundamental understanding of the Mlc1p protein’s

solution behavior in a complex with IQ peptides by sampling the conformational space of

two Mlc1p-IQ complexes. Our findings suggest that, although the IQ2 and the IQ4

peptides share similar sequence and structure, the fine details of each individual IQ

sequence determine its binding mode to the Mlc1p protein. The ability of the Mlc1p

protein to assume different conformations, which is driven by the specific IQ peptides, is

crucial. The flexibility of the protein and the dominance of its nonspecific hydrophobic

interactions with the IQ peptides are probably correlated with its ability to bind a wide

range of targets. Besides describing the structure and dynamics of the protein in the

presence of the peptides, we analyze the interaction free energy that governs the protein-

peptide formation. Using a combination of energies derived from MD simulations in an

explicit solvent, a continuum solvent model, and solute entropies contributions derived

from normal mode analysis, we have obtained approximate values for the protein-peptide

Page 84: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

71

interaction free energy of both complexes. We found that favorable molecular mechanics

energy contribution profoundly supports this protein-peptide interaction, while the polar

solvation energy and the entropy oppose it.

Page 85: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

72

3. Simulations of Free IQ Peptides

In the first two chapters of the thesis we found out that the solution model

structures of the Mlc1p-IQ complexes are both compact, while the IQ peptide is located

between the lobes of the protein. These observations may imply that the Mlc1p protein

contributes to the relative stability of the IQ peptides, while the peptides affect the

specific folding form of the protein. Since the only difference between the Mlc1p-IQ2

and the Mlc1p-IQ4 simulations is the IQ peptide, and in order to check how the IQ

peptides refold at the absence of the Mlc1p protein, we have performed MD simulations

of the peptides in solution. Accordingly, this chapter deals with MD simulations of free

IQ peptides. These simulations complement our research of the Mlc1p-IQ complexes, and

aim for elucidation the dynamic behavior of isolated IQ peptides without the presence of

the Mlc1p protein. Since none of the IQ peptides were crystallized at the absence of the

protein, their structures, which are basically alpha-helical, were derived from the

published crystal structures of the Mlc1p-IQ protein-peptide complexes (28, 56).

Considering the long duration of our simulations and the repetitive pattern of the models

obtained at different simulations' conditions, we argue that the predicted model structures

of both IQ peptides represent their solution conformations. Comparisons between the

dynamics of the free and the bound IQ peptides, between the free IQ2 and IQ4 peptides,

and between the simulations of the free IQ4 peptide at different conditions, were

performed.

3.1 Synopsis of the Presented Simulations

A series of MD simulations, presented in Table 4, was performed for various

Page 86: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

73

durations and conditions. These simulations include one MD run of the IQ2 peptide, and

five MD runs of the IQ4 peptide. Additionally, an extended IQ4 was constructed and

subjected to an MD run as further elaborated.

Table 4

Summary of the MD Simulations of the IQ Peptides

* 400 K. †

~ 30 mM NaCl. ‡

~ 300 mM NaCl. ¶ ~ 2.4 M NaCl (mimicking the salt concentration in which the Mlc1p-IQ4 complex was

crystallized (56)). § An IQ4 peptide containing 10 additional amino acids flanking each of its terminals.

These added 20 residues belong to the IQ3 and IQ5 peptides. Thus, the 45 amino acids

elongated IQ4 peptide constitutes a portion of the poly-IQ sequence as present at the neck

of the myosin. This simulation is described in chapter 4 of the thesis.

3.2 Overall Conformational Changes During the Simulations

The crystallographic and simulated structures of the free IQ2 peptide are

presented in Fig. 3.1 (Snapshots from the simulation are shown every 20-ns). At its initial

configuration (Fig. 3.1A), the peptide is almost linear alpha-helix. After 20-ns, the

peptide stretches (Fig. 3.1B), keeping its linearity. At the next snapshots (Figs. 3.1C-F),

its N- and C- edges loose, while its mid-section retains its straight conformation. From

40-ns till 100-ns, the model structure of the peptide hardly changes.

In contrast to the IQ2 peptide, snapshots of the IQ4 peptide reveal a different

Duration

(ns) Conditions of simulation

IQ2 peptide 100 As written in section 1.2 of the Methods

IQ4 peptide 100 As written in section 1.2 of the Methods

IQ4 peptide 30 High temperature *

IQ4 peptide 20 Low salt concentration †

IQ4 peptide 20 High salt concentration ‡

IQ4 peptide 20 Very high salt concentration ¶

Extended IQ4 peptide 20 An elongated IQ4 peptide §

Page 87: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

74

scenario (Fig. 3.2). The initial configuration of the IQ4 peptide is a linear alpha-helix

(Fig. 3.2A). After 20-ns of simulation it bends (Fig. 3.2B), whereas most of its secondary

structure is no longer alpha-helix. The conformation of the peptide continues to explore

the configurational space reaching a new configuration, observed at the next snapshots

(Figs. 3.2C-F). At this newly gained conformation, the IQ4 peptide is refolded in a

manner that its N- and C- edges are relatively close to each other. At the final

configuration, the IQ4 peptide is composed of two helices, separated by a coiled hinge.

The refolding process of the peptide, in which it loses most of its alpha-helical structure,

occurs in less than 20-ns. It can be seen that its conformation after only 20-ns of

simulation (Fig. 3.2B) grossly deviates from its initial one (Fig 3.2A). The secondary

structure analysis, presented in section 3.4 of this research, emphasizes the abrupt

transition of the peptide from an elongated alpha-helix into a conformation of two helices

separated by a coiled section.

A B C D E F

Figure 3.1: Snapshots of the crystal and the simulated structures of the free IQ2 peptide.

The structure of the IQ2 peptide as derived from the crystal structure of the Mlc1p-IQ2

complex (A). The simulated model structure of the IQ2 peptide after 20-ns (B); 40-ns (C);

60-ns (D); 80-ns (E) and 100-ns (F) of MD run.

Page 88: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

75

From the comparison of the snapshots between the two free IQ peptides, it is

evident that the IQ2 peptide is more rigid and stiff than the IQ4 peptide. Whereas the IQ2

peptide keeps most of its structural features, the IQ4 peptide modifies into a helix-loop-

helix conformation through an intermediate state. This argument will be further

strengthened and elaborated

We repeated the simulation of the IQ4 peptide at a high temperature (400 K). Our

aim was to check whether the elevated temperature may cause further structural changes

Figure 3.2: Snapshots of the crystal and the simulated structures of the free IQ4 peptide.

The structure of the IQ4 peptide as derived from the crystal structure of the Mlc1p-IQ4

complex (A). The simulated model structure of the IQ4 peptide after 20-ns (B); 40-ns (C);

60-ns (D); 80-ns (E) and 100-ns (F) of MD run. The simulated model structure of the IQ4

peptide after 30-ns (G) of MD run at 400 K.

A B C D

GE F

Page 89: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

76

to the peptide and detect changes not observed at 300 K. Fig. 3.2G depicts the model

structure of the IQ4 peptide after 30-ns of simulation at 400 K. Inspection of the peptide's

model at the end of the simulation at 400 K reveals a high similarity to the model

structure of the peptide after 100-ns simulation at 300 K. According to our simulations,

the elevated temperature does not affect the refolding mechanism of the peptide but,

rather, its rate (as also observed for the Mlc1p-IQ4 simulation at 400 K presented in

chapter 1 of the thesis). Thus, the progression of the conformational transition is basically

comparable at the two temperatures with a clear acceleration at the higher temperature.

Hence, we suggest that the modulation's speed of the peptide's model structure is

temperature-dependent.

The modulation of the IQ peptides' model structures with time can be followed by

examination of their backbone RMSD relative to their crystal structures in the protein-

peptide, as given in Fig. 3.3 (black for the IQ2 peptide, red for the IQ4 peptide). The

RMSD of the free IQ2 peptide (Fig. 3.3A, black) increases in a moderate fashion for the

first 13-ns of the simulation. Then, it rises abruptly to a value of ~ 0.76 nm, drops and

stabilizes at ~ 0.48 nm till the end of the simulation. The RMSD of the free IQ4 peptide

(Fig. 3.3A, red) increases at the very first nanoseconds of the simulation till a value of ~

0.5 nm, then gradually rises till it stabilizes on t ~ 50-ns at ~ 0.9 nm. Once the peptides

had reached their new stable conformations, they do not exhibit further configurational

changes. The calculation of the backbone RMSD value for the last 20-ns of the

simulations, relative to the backbone heavy atoms’ position at t = 80-ns (Fig. 3.3B),

reveals that both free IQ peptides hold steady solution model structures. These model

structures do not further evolve and thus represent settled down configurations.

Page 90: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

77

Figure 3.3: (A) The RMSD of the

backbone atoms of the free IQ2 (black)

and IQ4 (red) peptides as a function of

the simulation time; (B) The RMSD of

the backbone atoms of the last 20-ns of

the simulations (80−100 ns) relative to

the position of the same set of atoms at t

= 80-ns; (C) The RMSD of the backbone

atoms of the bound IQ2 (black) and IQ4

(red) peptides as a function of the

simulation time. Note the differences in

the abscissa between the simulations of

the free and the bound peptides.

B

C

A

The evaluation of the structural stability of the free peptides calls for comparison

with their stability in a complex with the Mlc1p protein. The RMSDs of the bound IQ

peptides were reported in Figs. 1.2A and 1.4A, and presented also in Fig. 3.3C. That for

the bound IQ2 peptide increases after ~ 3-ns to a value of ~ 0.23 nm and stabilizes,

whereas that for the bound IQ4 peptide doubles after ~ 3.3-ns from ~ 0.2 nm to ~ 0.4 nm,

and stays at this value till the end of the simulation. When one compares the simulations

of the free and the bound peptides, it can be clearly seen that free peptides tend to loose

their structure in a more profound manner than the bound ones.

Page 91: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

78

The Mlc1p protein is likely to shield or protect the bound peptides from a gross

deviation from their initial structure observed for the free peptides. It causes a steric

hindrance that prevents the peptide to assume conformations observed when simulated at

its absence. It should be mentioned that at both the free and at the bound peptide's

simulations, the IQ4 peptide is more flexible and tends more to deviate from its original

structure than the IQ2 peptide.

3.3 Structural Characteristics of the Refolding Process of the IQ Peptides

To demonstrate the modulation of the IQ peptides during the simulations, two

alpha-helices in their N- and C- sections, as present in their crystalline structures, were

defined. The angle between them is presented in Fig. 3.4.

Figure 3.4: The dynamics of the relative rotation of the free and the bound IQ peptides

during the simulations time. (A) The intra-helical angle between the N- and C- sections

of the free IQ2 (black). The arrow points towards a conformation of the free IQ2 peptide

at 15-ns; (B) The intra-helical angle between the N- and C- sections of the bound IQ2

(black) and IQ4 (red) peptides.

A B

Page 92: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

79

For the free IQ2 peptide, the angle is relatively constant at a value of ~27°,

besides a transient increase to a value of ~125° at the time range ~13-ns till ~20-ns. This

temporary rise, corresponds to the short-lived configuration of the peptide marked with

an arrow, coincides with the increment of the RMSD observed at Fig. 3.3. Thus, even

though the IQ2 peptide explores the parametrical space and tries to refold, it relatively

returns to its initial conformation and does not grossly deviate from it. For the free IQ4

peptide, the intra-helical calculation is not given since the peptide does not hold two

defined sections throughout the entire simulation. However, its structural modulation is

clearly detected on Fig. 3.2.

A different scenario is observed when dealing with the bound IQ peptides (Fig.

3.4B). The angle of the bound IQ2 peptide remains constant throughout the entire

simulation time. The bound IQ4 peptide, which at the initial state consists of a single

alpha-helix, shows a significant flexibility. Its straight alpha-helical structure is snapped

in its middle during the complex's structural transition, when the two shorter helices

attain an almost perpendicular relation. The process, in which the angle between the

helices increases from ~ 20° to ~ 70°, is sharp and quick.

As demonstrated, the IQ4 peptide is characterized by a greater flexibility than the

IQ2 peptide, both in its free and its bound configurations. While the former experienced a

refolding process, the latter keeps its linearity. Upon comparison of the extent of the

curvature of the free and the bound IQ4 peptides in solution, it is evident that the free

peptide bends much more than the bound one. The considerable modification of the free

IQ4 peptide enables it to form intra-peptide bonds (between its N- and C- terminals),

whereas the bound IQ4 peptide exhibits two defined, almost vertical, sections.

Page 93: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

80

Figure 3.5: The end-to-end distance (from C-alpha of the first residue till C-alpha of the

last residue) of the free IQ2 (black) and IQ4 (red) peptides.

The refolding of the IQ peptides can be also expressed when calculating the

distance between their first and last alpha-carbons (Fig. 3.5). For the free IQ2 peptide

(black), the end-to-end distance hardly changes during the simulation. After decreasing

from ~ 3.7 nm to ~ 2.5 nm, it keeps a relatively constant distance. However, this end-to-

end distance decreases for a short while (~ 13-ns till ~ 20-ns) to a distance that is even

shorter than 1 nm. The short end-to-end distance reflects its transient curved

conformation, which is not stable and alternates back to the initial straight configuration.

The refolding of the free IQ4 peptide (red) is clearly seen as its length shortens from ~

3.7 nm to ~ 2 nm. Afterwards, at t ~ 22-ns it drops to ~ 0.75 nm, rises to ~ 1.5 nm and

finally stabilizes at ~ 0.7 nm. The free IQ4 peptide shows a significant conformational

flexibility, shown when its bending brings its ends into a closer contact than seen at their

initial state. Overall, the structural characteristics of the extensive conformational change

of the peptide are presented by variation of its length as well as by the increase of the

angle between its N- and C- sections.

Page 94: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

81

3.4 Secondary Structures of the IQ Peptides

We followed the secondary structure of the free IQ peptides as a function of the

simulation time and the analyses of the free IQ2 (Fig. 3.6) and the free IQ4 (Fig. 3.7) are

presented.

The initial alpha helical conformation of the free IQ2 (Fig. 3.6) peptide remains

stable for more than 20-ns, and then transforms into a 5-helix conformation while

keeping its linearity. A 5-helix conformation, also known as a Π-helix, is a helical

structure in which the backbone C=O group of the i residue forms a hydrogen bond with

Figure 3.6: Secondary structure analysis of the free IQ2 peptide as a function of the

simulation time. The residue number runs along the ordinate and time along the abscissa.

Color codes are used to represent secondary structure elements.

Page 95: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

82

Figure 3.7: Secondary structure analysis of the free IQ4 peptide as a function of the

simulation time. The residue number runs along the ordinate and time along the abscissa.

Color codes are used to represent secondary structure elements.

the backbone N-H group of the residue five amino acids ahead (i → i + 5 hydrogen

bonding). The 5-helix conformation becomes the dominant secondary structure element

and holds till the end of the simulation of the free IQ2 peptide.

The secondary structure analysis of the free IQ4 peptide (Fig. 3.7) reveals a

different pattern than that for the IQ2 peptide.

After a short relaxation period, the peptide is no longer composed of only alpha

helical elements. While its first ~ 15 residues still keep their alpha helical configuration,

the remaining 10 residues become composed of bends and turns. At ~ 16-ns, a 5-helix

Page 96: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

83

appears and gradually controls over the model structure till ~ 38-ns (besides a middle

coil/turn section). Apparently, the 5-helix conformation is not stable enough to persist for

the entire helix, and the first ~ 15 residues reverse back to the alpha-helical conformation.

From this time point on, the model structure of the peptide exhibits three secondary

structure elements, reflecting an energetically stable configuration of the peptide: the N-

section, composed of an alpha-helix; the C-section, composed of a 5-helix; and a hinge,

composed of turn and coil, separating between these N- and C- sections.

The secondary structure analyses of both peptides add an additional value to the

data presented by the snapshots and the RMSD of the peptides. Subtle changes in the

secondary structure composition of the peptides, that are not necessarily observed in Figs.

3.1−3.3, are clearly seen in this type of analysis. For the IQ2 peptide, although one can

allegedly think that the peptide barely changes during its simulation; this analysis

demonstrates that it loses most of its alpha-helical structure at the expense of gaining a 5-

helix structure. For the IQ4 peptide, it can be assumed that its alpha-helix had practically

lost. However, we demonstrated that it attains about 50 % from its initial alpha-helix

content, while the newly appearing 5-helix and coil/turn configurations form the other

half. Interestingly, the alpha- to 5- helix transition (full-length and partial for the IQ2 and

the IQ4 peptides, respectively), which was detected in both simulations, was observed, as

well, in previous MD simulations of peptides. 5-helix propagation at the expense of an

alpha-helix was noticed in simulations of the transmembrane domain of ErbB-2 (151,

152), the central domain of caldesmon (153) and synthetic peptides (154). Besides

theoretical MD simulations reporting a formation of the 5-helix structure, there are also

experimental evidences to the existence of this secondary structure element in proteins

Page 97: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

84

and peptides. For example, 5-helix was identified by x-ray crystallography and NMR in

fumarase C (155), glycogen phosphorylase b (156), lipoxygenase (157), catechol O-

methyltransferase (158), serine carboxypeptidase (159), and the protein NMA1147 (160).

5-helix was also identified at the vasoactive intestinal protein, using an empirical

approach involving peptide design, construction, and distribution frequency techniques

(161).

Both peptides explore the conformational space till they find a stable

configuration. At more (for the IQ2 peptide) or less (for the IQ4 peptide) 20-ns, both

loose they initial structures, and then, following more exploration, new stable model

structures emerge.

3.5 Salt Bridges Analysis of the IQ4 Peptide Simulation

An analysis for the presence of salt bridges was performed for the 100-ns long

simulation of the IQ4 peptide.

The IQ4 peptide is a highly-charged peptide, containing 8 positive residues and 2

negative ones. The oppositively-charged residues can form salt bridges between them,

and thus we followed the minimal distance between each pair of oppositively-charged

residues, and present the main results in Fig. 3.8. A ribbon diagram of the IQ4 peptide,

presenting its positive and negative residues (blue and red, respectively) is shown on the

right.

Only one salt bridge (R20-E16) persisted over the whole simulation time (Fig.

3.8A), keeping the minimal distance between the residues as ~ 0.25 nm. Two salt bridges

(K12-E16, K11-E16) were broken during the simulation at a time point corresponding

Page 98: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

85

Figure 3.8: Salt bridges analysis of the free IQ4 peptide simulation. (A) Minimal

distance between R20 and E16; (B) Minimal distance between K12 and E16; (C)

Minimal distance between K11 and E16; (D) Minimal distance between R20 and E25;

(E) Minimal distance between K11 and E25; (F) Minimal distance between R4 and

E25. The crystalline conformation of the IQ4 peptide, as derived from the crystal

structure of the Mlc1p-IQ4 complex, is presented on the right side. The positive and

negative residues of the IQ4 peptide are shown in blue and red, respectively.

A B C

D E F

R20

E25

R4

E16

K11

K12

K23

K18

R14

K15

with the major conformational change of the peptide (Figs. 3.8B-C). Two salt bridges

(R20-E25, K11-E25) were detached at ~ 62-ns (Figs. 3.8D-E). One salt bridge (R4-E25)

was formed (Fig. 3.8F) at the end of the simulation when the two edges of the peptide got

close enough to be linked through a salt bridge. This latter is of particular interest since it

represents the structural modification of the peptide and reflects its curvature. The

residues R4 and E25 are located at the edges of the IQ4 peptide. At the beginning of the

simulation, they are at a distance that exceeds 3 nm. As the simulation progresses, these

residues become closer to each other, keeping a distance of ~ 1.1 nm. Finally, the

structural deformation of the peptide imposed a stable contact of the residues, in which

their VdW radii almost reached (d = 0.25 nm).

Page 99: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

86

From the salt bridge analysis it becomes clear that only two salt bridges maintain

the refolded model structure of the peptide, while four pre-existing salt bridges detached.

However, these bonding salt bridges contribute to the enhanced stability of the refolded

model structure of the IQ4 peptide.

3.6 Dynamics of the IQ4 Peptide at Different Salt Concentrations

We repeated the MD simulations of the free IQ4 peptide, for a duration of 20-ns

each, in three different salt concentrations: ~ 30 mM (low), ~ 300 mM (high) and ~ 2.4 M

(very high). Snapshots of the structures after 20-ns simulations (Fig. 3.9) are presented.

At the low salt concentration, the IQ4 peptide refolded into two distinctive

sections separated by a hinge (Fig. 3.9A). However, at the high salt concentration, it did

not refold, although its N- and C- terminals were loosen (Fig. 3.9B). The high salt

concentration prevented the peptide to assume a different conformation from its initial

Figure 3.9: Snapshots of the crystal and the simulated structures of the IQ4 peptide at

different salt concentrations. The simulated model structures of the IQ4 peptide after 20-

ns simulation at low (A), high (B), and very high salt (C) concentrations.

A B C

Page 100: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

87

one, keeping it as an elongated helix.

Given that the Mlc1p-IQ4 complex was crystallized at a very high salt

concentration (56) (~ 2.4 M), we suggest that its crystal structure was imposed by its

crystallization conditions. According to this argument, which is compatible with the

results we present in chapter 1 of the thesis, the crystal structure of the Mlc1p-IQ4

complex does not represent its physiological solution structure. Thus, we repeated the

simulation of the free IQ4 peptide at salt conditions mimicking its crystallographic ones,

and a snapshot of the simulated model structure after 20-ns simulation (Fig. 3.9C) is

shown. The snapshot resembles the one obtained by the simulation at the high salt

because that already at ~ 300 mM of NaCl the screening effect caused by the ions is

extensive. The 8-fold increase at the NaCl concentration did not significantly change the

screening effect and thus the results of the high and the very high salt concentrations'

simulations are close.

Since the peptide hardly bends at the high salt concentrations (as opposed to the

bending at ~ 30 mM and ~ 100 mM salt concentrations), it is obvious that its refolding

pattern is influenced by the concentration of ions in its surrounding. High salt

concentrations probably prevent the peptide to assume a conformation that reflects its

solution structure, and thus yield a configuration that differs from a physiological one.

We can generalize the argument and suggest that the conditions of the crystallization,

such as the salt concentration of the crystallization buffer, may influence the

configuration of the protein inside the crystal lattice. A clear influence of the crystal

environment on protein structures was also found by others (162, 163). Therefore, the

simulations' results at the high salt concentrations strengthen our conclusion that the

Page 101: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

88

crystallization conditions of the Mlc1p-IQ4 affected its crystal packing forces, while the

latter determined the crystalline structure. Hence, the simulations of the peptide at the

high and the very high salt concentrations exemplify how crystallographic conditions

may determine the outcome structure of the crystal.

3.7 Summary

MD simulations of IQ peptides indicate that the free (and also the bound) IQ2

peptides almost do not refold in a solution and maintain a stable helical structure. In

comparison, both the free and the bound IQ4 peptides are less stiff and tend to curve and

flex in a profound manner. This is attributed to the higher net charge (Z = +6) of the IQ4

peptide than this of the IQ2 peptide (Z = +2), and reflected by the electrostatic attraction

between positive and negative residues located on its N- and C- terminals (for example,

R4-E25). The IQ4 peptide presents an intrinsic instability and an increased tendency to

flexibility at a more or less defined location (further discussed in the next chapter).

Though the IQ4 peptide exhibits a structural modification when simulated at the presence

of the protein, its extent of refolding is considerably higher when it is simulated at its

absence. Secondary structure analysis of both free IQ peptides followed their secondary

structure change throughout the simulations, revealing their stable solution

conformations. The curved shape of the free IQ4 peptide was observed when it was

simulated under sub-physiological and physiological salt concentrations (~ 30 mM and ~

100 mM), and was not detected when the simulations were performed under over-

physiological salt concentrations (~ 300 mM and ~ 2.4 M).

Page 102: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

89

4. The Structure of the Light Chain-Binding Domain (LCBD) of Myosin

V

Given our results presented in chapters 1−3 of the thesis, and an MD simulation

of extended IQ4 peptide presented below, we propose a dynamic solution model of the

LCBD of myosin V (section 4.3). The model, which involves mutual modulations of the

structures of the light chain proteins in respect to the IQ peptides of the myosin's neck

towards each other, may have important implications regarding the structure-function

relationship of the lever arm of myosin V.

4.1 The Current Structural Model

Myosin V is a versatile motor involved in the short-range transport of vesicles in

the actin-rich cortex of the cell. Its long neck domain, serving as a lever arm (19, 20),

gives rise to a step size of ~ 36 nm, the largest step size thus far measured for a myosin

motor. The LCBD neck of myosin V consists of six tandem IQ motifs, to which light

chain proteins, such as the CaM and the Mlc1p, are bound. The primary function of the

light chains is to regulate the ATPase activity of the globular head of myosin V (13, 17,

164). The crystal structures of the Mlc1p-IQ2, Mlc1p-IQ4 and Mlc1p-IQ2/3 complexes

had been determined by Terrak and co-workers (28, 56). On the basis of these structures,

and sequence similarity among the six IQ motifs, Terrak and co-workers suggested a

model for the LCBD (Fig. 4.1) (57). According to this model, the six IQ motifs, that

constitute the neck domain of the myosin V, adopt a straight long alpha helical

configuration (green). Moreover, two of the light chain proteins retain an extended

configuration (see the arrows pointing towards the light chains that bind the IQ4 and the

Page 103: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

90

IQ6 peptides), in which their NL does not interact with the IQ motif, as determined by the

crystalline form of the Mlc1p-IQ4 complex.

Figure 4.1: Model of the LCBD of myosin V. The light chains, which can be either CaM

or CaM-related molecules such as the Mlc1p, are colored cyan (NLs) and magenta (CLs),

and the six-IQ fragment of the heavy chain is colored green. The black arrows point

towards the light chains that bind the IQ4 and the IQ6 peptides. An enlargement

illustrates the interaction between adjacent light chains. Adapted from (57).

This proposed model of the LCBD does not take into account the conformations

that the proteins may reveal in solution; rather, it is constrained by the packing forces of

the Mlc1p-IQ crystal structures. These packing forces may vary between the Mlc1p-IQ2

and the Mlc1p-IQ4 structures because each complex was crystallized solely under unique

crystallographic conditions. Attempts to grow crystals of both complexes under identical

or similar conditions failed (Terrak-M, personal communication). The elucidation, by

MD simulations, of the solution structures of two Mlc1p-IQ complexes (presented in

chapters 1-2), and with a combination of MD simulation of free IQ2 and IQ4 peptides

(presented in chapters 3-4), calls for reevaluation of the Mlc1p-LCBD model structure.

Our reevaluation of the current structural model and suggestion for a solution

model, which are discussed in section 4.3, are based on a wide range of experimental data

Page 104: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

91

as well. Fluorescence imaging with one-nanometer accuracy (FIONA) (165-167) and

time-resolved single-molecule fluorescence polarization studies (SMFP) (168) suggest a

fundamental role to the elasticity of the LCBD during the movement of the myosin V.

Myosin V "walks" following an asymmetric hand-over-hand mechanism, where its heads

alternate leading and trailing positions along the actin filament, analogous to the hands of

a rope climber. In the course of its stride, a conformational change was demonstrated

during the transition of the lever-arm from a pre-stroke to post-stroke state. This change

is manifested by a tilting of the LCBD between two distinct conformations, a straight one

and a bent one. The curvature of the LCBD deduced from these experiments is in a very

good accord with the one predicted in our MD studies as further discussed. In addition,

when actin-bound myosin V was imaged by electron microscopy (Fig. 4.2) (169, 170), a

bent lever-arm was observed.

Figure 4.2: Two electron microscopy images showing filaments of actin with two-

headed myosin V "walking". The red arrows point towards the myosin's lead arm, seen in

a bent conformation. These electron microscopy images, by presenting myosins with

curved lever arms, strengthen our suggested model. The scale bar is 50 nm (note that the

myosin was found with fringes and thus spans ~ 50 nm). Images adapted from (169).

The leading head was curved backwards, whereas the rear head was straighter,

Page 105: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

92

resembling a skier in a telemark stance. Thus, the electron microscopy data confirms the

curved shape of the LCBD of the lead head pre-stroke state. Furthermore, when the

crystal structure of scallop’s myosin S1 was determined (171), a ~ 90° curvature of the

myosin's IQ4 motif neck domain was observed. It was found that the lever arm of the

scallop's myosin S1 does not move as a rigid body, but rather flexes when the myosin is

in motion.

4.2 Structure and Dynamics of an Extended IQ4 Peptide

In order to verify that the kink presented by the bound (Fig. 3.4B and chapter 1)

and the free IQ4 peptides (Figs. 3.2 and 3.4A) has a physiological relevance, we build an

extended IQ4 peptide and performed an MD simulation for it. The sequence of this

peptide consists of the IQ4 peptide, as present at the crystal structure of the Mlc1p-IQ4

complex, and an additional twenty residues. Ten of these residues flank its N-terminus

and are part of the IQ3 peptide sequence, while the other ten of these residues flank its C-

terminus and belong to the IQ5 peptide sequence. Hence, this 45 long amino acids

sequence represents a section from the poly-IQ (IQ1-IQ6) of the neck of myosin V.

Comparison between the model solution structures of the free IQ4 peptide after

100-ns of simulation, the bound IQ4 peptide after 12-ns with the Mlc1p protein, and the

extended IQ4 peptide after 20-ns of simulation are, presented in Fig. 4.3 (A-C,

respectively). All model structures exhibit a major kink (see the black arrows), that is

roughly located at the same location along the peptides. The common position of the

angular shift of the peptide (from a linear peptide with a ~ 180° angle to a curved one

with a hinge) can be referred as a "hot spot", which tends to curve rather than assume a

Page 106: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

93

Figure 4.3: Snapshots of the simulated model structures of the IQ4 peptides. (A) After

100-ns simulation of free IQ4 peptide; (B) After 12-ns simulation of bound IQ4 peptide;

(C) After 20-ns simulation of extended IQ4 peptide (the IQ4 peptide is drawn in blue,

while its extended residues are colored in red). The black arrows indicate the "hot spot"

position.

C B A

straight conformation. This "hot spot" is also seen at the simulation of the free IQ4

peptide at the low salt concentration (Fig. 3.9A).

We use the term "hot spot" not only as a phrase to describe the location of the

peptide's twist but literally as an idiom to characterize a sequence encompassing a high

B-factor. B-factor or temperature factor is an indicator of thermal motion of an atom, and

is commonly used as a measure of how much an atom oscillates or vibrates around a

specified position. Thus, the high B-factor that is observed at the location of the kink in

the simulations of the IQ4 peptide reflects its local inherent instability. Moreover, when

the RMSF was calculated for the IQ4 peptide's simulations, it turned out that residues that

present high RMSF values correspond to those that curve. Therefore, intrinsic flexibility

of the IQ4 peptide was already embedded in its X-ray structure, and, as such, our finding

is not surprising.

The RMSD analysis of the extended IQ4 peptide is presented in Fig. 4.4. It

increases from ~ 0.1 nm to ~ 0.5 nm, and stabilizes on this value. Since the RMSD

Page 107: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

94

Figure 4.4: The RMSD of the backbone atoms of the MD run of the extended IQ4

peptide as a function of the simulation time.

reached to a relatively constant value for the last 10-ns of the simulation, further major

conformational changes are not expected. Thus, the curved extended IQ4 peptide seems

to be structurally steady and hence may represent a physiological entity. This finding

strengthens and supports our reevaluation regarding the structural model of the LCBD

presented in section 4.3.

4.3 The Suggested Solution Model

The following suggested solution model of the LCBD is a consequence of our

overall findings presented throughout this study. We propose that the light chains of the

myosin, namely the CaM and the Mlc1p proteins, may maintain a compact conformation

and not an extended one as their ID is bent (as in Figs. 1.1B and 1.3B). Their NL is

probably not free to engage in protein-protein interactions as claimed by Terrak and co-

workers, but rather interacts with the IQ peptides, coming into a close contact with their

CL. In addition, and not less important, we propose that the poly-IQ sequence of the

Page 108: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

95

myosin's neck may curve when bound to the light chain proteins. This structural

flexibility of the poly-IQ is associated with the myosin's walking over the actin filament,

generating the bended knee conformation of the lead arm of the LCBD, which is a crucial

element in its mechano-chemical mechanism. Therefore, the banana-shaped arm of the

myosin, which was not included in the structural model of Terrak and co-workers, is

revealed by our simulations. Besides supplying a dynamic model of the poly-IQ

sequence, we point out the exact location of its knee. According to our simulations, it

appears that the curvature of the neck may be located within the IQ4 peptide sequence.

The hinge presented by the IQ4 peptide, observed in all our simulations under

physiological salt conditions (and also for the extended IQ4 peptide), may correspond to

a physiologically relevant bent solution structure of the lever arm of myosin. This unique

observation regarding the location of the curve, which was not specified by electron

fluorescence or electron microscopy studies, exemplifies the ability of MD simulations to

provide additional information to the one obtained by experimental biologists.

The tight refolding of the protein around the peptides and the bending of the IQ4

peptide point out the difficulty of predicting the solution structure of a large protein-

peptide complex based on crystal structures of some of its isolated components. The

schematic description of the IQ motifs as constitutes of a rigid straight alpha helix (as

shown on Fig. 4.1), which is not consistent with the dynamics of the peptides and

complexes, represents an unlikely physiological conformation.

Our view about myosin V is corroborated by an illustration, in which the

movement of myosin V over an actin filament is shown (Fig. 4.5). Though this is only an

illustration, we feel that it faithfully delivers our proposed solution model. The lead arm

Page 109: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

96

of the neck is curved at a position that corresponds to the IQ4 peptide (see the white

arrow), exemplifying that the flexibility of the neck is embedded in the myosin's motion.

Thus, the flexibility of the neck, as opposed to the model proposed by Terrak and co-

workers, is necessary for the proper function of the mechano-chemical function of the

myosin molecules.

Figure 4.5: Myosin V (green), a biomolecular motor that moves in nanometer size steps

on actin (red), transports cargo within cells. By placing a fluorophore near one foot

(rainbow colored oval), and following the motion a single myosin V, it was determined

that myosin V "walks", placing one foot over the other, and does not "crawl". The

illustration demonstrates that the neck domain of myosin V, as predicted by our

simulations, curves when it strides over the actin filament. The white arrow points

towards the IQ4 peptide of the lead arm of the myosin. The light chains were omitted

from the illustration for simplicity. Adapted from Yildiz et al. (166).

4.4 Summary

On this chapter we addressed the structure of the LCBD of myosin V from the

yeast Saccharomyces cerevisiae. We presented the current structural model and proposed

a modified model, based upon the various MD simulations discussed on the thesis. The

Page 110: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

97

crystal structure of the Mlc1p-IQ4 protein-peptide complex was obtained under high salt

conditions. As such, the IQ4 peptide exhibits a straight alpha-helical configuration and

the protein, restricted by the linear conformation of the peptide, warps itself around it.

According to our suggestion, which was elaborated throughout the thesis, the crystal

structure of the Mlc1p-IQ4 protein-peptide complex may not represent its physiological

one, and thus the model proposed by Terrak and co-workers may be misleading. We wish

to stress out that although a wide plethora of evidences point out to this conclusion, it

would be interesting to perform a simulation of that complex under high salt conditions.

Such an MD simulation will provide a further verification for the influence of the

crystallization conditions, i.e. high salt and ionic strength of the crystallization buffer, on

the configuration of the protein inside the crystal lattice.

Apparently, the solution structure of the LCBD of myosin V is more complicated

than the one based on crystal structures of a few individual Mlc1p-IQ complexes, as

predicted by Terrak and co-workers. When our simulations are taken into account

together with the experimental data presented above, it is reasonable to conclude that the

LCBD is not a passive structural device but dynamic proteinous machinery. We argue

that mutual structural flexibility of the light chain proteins and the IQ peptides represents

a more realistic model of the neck region of myosin V. Our proposed solution model of

the flexed neck of myosin V IQ4 peptide is in accord with all our MD simulations and

also in agreement with other research groups reporting the flexible nature of the neck.

However, a word of caution should be stated. We cannot preclude the possibility that the

neck may curve and break its linearity along other IQ peptides, besides the primary kink

at the IQ4 sequence.

Page 111: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

98

IV. OVERALL GENERAL DISCUSSION

The current research is the first MD study of IQ peptides in a complex with a

CaM-like protein and at its absence. Since IQ motifs are widely spread in nature, serving

as binding sites for CaM and CaM-like proteins, and as the structure of CaM resembles

that of Mlc1p, it was of great importance to conduct this research in order to elucidate the

structural mechanism underlying the interaction of CaM and CaM-like proteins with IQ

peptides. Analysis of the structure, dynamics and energetic aspects of the Mlc1p-IQ

complexes was of high significance also due to their important role in the regulation of

the mechano-chemical myosin system of the yeast Saccharomyces cerevisiae. Given that

the Mlc1p protein functions and binds myosin V at the absence of Ca+2

, fulfilling roles

taken by Holo-CaM in other classes of myosins, our conclusions presented in this study

may, as well, hold for the entire family of myosins.

In this study, we described a usage of the MD methodology as a tool to analyze

the crystal structures of Mlc1p-IQ complexes. The reported simulations addressed the

manifold structures of the Mlc1p-IQ complexes as found in crystal structures and in

solution. We have clearly demonstrated that the crystal structure of the Mlc1p-IQ4

complex, obtained under high salt conditions, does not represent its solution structure. At

this high ionic strength, the IQ4 peptide assumes a straight alpha helix and does not bend.

It supplies a scaffold on which the protein binds but the latter keeps an extended

conformation. At a physiological ion strength, the IQ4 peptide curves and consequently

the protein may hold a compact structure. We speculate that at low-medium ionic

strength the Mlc1p-IQ4 could not well diffract, representing an inhomogeneous crystals

population.

Page 112: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

99

We can generalize the above mentioned argument and claim that crystal structures

of proteins do not necessarily represent their solution structures. Though this claim may

not seem to be too adventurous or innovative, our research clarifies that a crystal structure

can be influenced by the crystallographic conditions, while the latter determine the

packing forces operating at the crystal lattice. Furthermore, crystalline forms of proteins

can be imposed by specific crystallographic conditions, consequently leading to a

fundamental deviation of the obtained X-ray structure from the native configuration.

We have noticed that the Mlc1p protein had a structural versatility, allowing it to

bind the IQ peptides in a mode reflecting their precise sequence, hardly using the "IQ

motif residues" as key anchoring sites. Apparently, the Mlc1p protein can be flexed in

many configurations, each suitable for a certain IQ structure. Thus, the sequence variation

among the IQ motif peptides is sufficiently large to induce non-identical interactions with

the same protein. Taking a broader view, we suggest that when two similar (but not

identical) peptides, which belong to the same family, bind the same protein, each protein-

peptide structure may assume a unique conformation.

Following a comprehensive energy analysis, we found that the protein-peptide

complexes formation is mediated by enthalpy, while entropy opposes it. The close

interaction between the Mlc1p protein and the IQ peptides is grossly mediated by LJ

interactions, whereas their opposite highly charged surfaces contribute to their initial

encounter. Electrostatic interactions may serve as the driving force for long-range

molecular recognition, but once protein-peptide contact occurs, highly specific VdW

interactions and non-specific hydrophobic interactions stabilize the Mlc1p-IQ complexes.

This finding may also be valid for pairs of highly opposite charged proteins that form

Page 113: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

100

protein-protein or protein-ligand complexes.

The simulations of the protein-peptide complexes, combined with the simulations

of the free IQ peptides, allow us to offer a comprehensive solution model of the LCBD.

According to our simulations, and in contrast to a structural model for the LCBD

suggested by Terrak and co-workers, the solution structures of the Mlc1p-IQ protein-

peptide complexes may impose refolding of the light chain proteins around the IQ

peptides. Thus, the light chain proteins may reveal a compact configuration and not an

extended one. Considering a poly-sequence of IQ motifs, as present in the LCBD, it

appears that its final structure, saturated by the CaM-like proteins, is under intensive

internal stress. The internal stress may be associated with the mechano-chemical function

of the myosin molecules. When the myosin "walks" over the actin filament, its lead arm

of the neck domain must flex in every stride. Our simulations, by adding a dynamic

nature to the structural model, demonstrate the curvature of the LCBD. As suggested by

our simulations, the LCBD assumes a banana-like conformation, while the location of its

knee corresponds to the IQ4 peptide. The knee-shape structure that the LCBD may hold

is essential to its functionality. We wish to note that our model emphasizes that the

LCBD is a dynamic proteinous machinery, and not a passive structural device as

previously suggested.

In a broader perspective, our research shows that careful application of MD

simulations can be used for extending the structural information presented by crystal

structures, thereby revealing the dynamic configurations of proteins in their physiological

environment. It is to be hoped that experimental structural biologists will make increasing

use of MD simulations for obtaining a deeper understanding of particular biological

Page 114: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

101

systems. When MD simulations for long durations will become routine procedures, a

constructive interplay between the simulations and experiments will be useful and

informative regarding both sides. Given the availability of several MD programs, large

amounts of computer time, and examples for which MD has really played a role in

furthering our understanding of protein structure and functions (as presented in this

research), we look forward to a wide field of biological and biophysical applications for

MD in the future.

Page 115: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

102

V. SUPPLEMENT

1.1 Articles

The outcome of the thesis is three articles:

(I) Ganoth, A., E. Nachliel, R. Friedman, and M. Gutman. 2006. Molecular dynamics

study of a calmodulin-like protein with an IQ peptide: Spontaneous refolding of the

protein around the peptide. Proteins 64:133-146. (The article is attached at the following

pages).

(II) Ganoth, A., R. Friedman, E. Nachliel, and M. Gutman. 2006. A molecular

dynamics study and free energy analysis of complexes between the Mlc1p protein and

two IQ motif peptides. Biophys J 91:2436-2450. (Cover article. Both article and cover are

attached at the following pages).

(III) Ganoth, A., R. Friedman, E. Nachliel, and M. Gutman. The structural basis for

myosin V movement: A molecular dynamics study (temporary title). In Preparation.

Page 116: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

103

REFERENCES

1. Berg, J. S., B. C. Powell, and R. E. Cheney. 2001. A millennial myosin census.

Mol Biol Cell 12:780-794.

2. Sellers, J. R. 2000. Myosins: a diverse superfamily. Biochim Biophys Acta

1496:3-22.

3. Weeds, A. G., and R. S. Taylor. 1975. Separation of subfragment-1 isoenzymes

from rabbit skeletal muscle myosin. Nature 257:54-56.

4. Pollard, T. D., W. F. Stafford, and M. E. Porter. 1978. Characterization of a

second myosin from Acanthamoeba castellanii. J Biol Chem 253:4798-4808.

5. Krendel, M., and M. S. Mooseker. 2005. Myosins: tails (and heads) of functional

diversity. Physiology (Bethesda) 20:239-251.

6. Baker, J. P., and M. A. Titus. 1998. Myosins: matching functions with motors.

Curr Opin Cell Biol 10:80-86.

7. Reck-Peterson, S. L., D. W. Provance, Jr., M. S. Mooseker, and J. A. Mercer.

2000. Class V myosins. Biochim Biophys Acta 1496:36-51.

8. Kalhammer, G., and M. Bahler. 2000. Unconventional myosins. Essays Biochem

35:33-42.

9. Eitzen, G., L. Wang, N. Thorngren, and W. Wickner. 2002. Remodeling of

organelle-bound actin is required for yeast vacuole fusion. J Cell Biol 158:669-

679.

10. Luna, E. J., K. N. Prestonjamasp, T. Nebl, Y. Chen, C. Gatto, S. W. Oh, and R. K.

Pope. 2001. Motile Membrane Skeletons: What Neutrophils and Muscle Have in

Common. Cell Mol Biol Lett 6:219.

11. Korn, E. D. 2000. Coevolution of head, neck, and tail domains of myosin heavy

chains. Proc Natl Acad Sci U S A 97:12559-12564.

12. Sellers, J. R., and C. Veigel. 2006. Walking with myosin V. Curr Opin Cell Biol

18:68-73.

13. Langford, G. M. 2002. Myosin-V, a versatile motor for short-range vesicle

transport. Traffic 3:859-865.

14. Nascimento, A. A., R. E. Cheney, S. B. Tauhata, R. E. Larson, and M. S.

Mooseker. 1996. Enzymatic characterization and functional domain mapping of

brain myosin-V. J Biol Chem 271:17561-17569.

15. Langford, G. M., S. A. Kuznetsov, D. Johnson, D. L. Cohen, and D. G. Weiss.

1994. Movement of axoplasmic organelles on actin filaments assembled on

acrosomal processes: evidence for a barbed-end-directed organelle motor. J Cell

Sci 107 ( Pt 8):2291-2298.

16. Cheney, R. E., M. K. O'Shea, J. E. Heuser, M. V. Coelho, J. S. Wolenski, E. M.

Espreafico, P. Forscher, R. E. Larson, and M. S. Mooseker. 1993. Brain myosin-V

is a two-headed unconventional myosin with motor activity. Cell 75:13-23.

17. Espreafico, E. M., R. E. Cheney, M. Matteoli, A. A. Nascimento, P. V. De

Camilli, R. E. Larson, and M. S. Mooseker. 1992. Primary structure and cellular

localization of chicken brain myosin-V (p190), an unconventional myosin with

calmodulin light chains. J Cell Biol 119:1541-1557.

18. Espindola, F. S., D. M. Suter, L. B. Partata, T. Cao, J. S. Wolenski, R. E. Cheney,

S. M. King, and M. S. Mooseker. 2000. The light chain composition of chicken

Page 117: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

104

brain myosin-Va: calmodulin, myosin-II essential light chains, and 8-kDa dynein

light chain/PIN. Cell Motil Cytoskeleton 47:269-281.

19. Rayment, I., H. M. Holden, M. Whittaker, C. B. Yohn, M. Lorenz, K. C. Holmes,

and R. A. Milligan. 1993. Structure of the actin-myosin complex and its

implications for muscle contraction. Science 261:58-65.

20. Mehta, A. D., R. S. Rock, M. Rief, J. A. Spudich, M. S. Mooseker, and R. E.

Cheney. 1999. Myosin-V is a processive actin-based motor. Nature 400:590-593.

21. Sakamoto, T., F. Wang, S. Schmitz, Y. Xu, Q. Xu, J. E. Molloy, C. Veigel, and J.

R. Sellers. 2003. Neck length and processivity of myosin V. J Biol Chem

278:29201-29207.

22. Toth, J., M. Kovacs, F. Wang, L. Nyitray, and J. R. Sellers. 2005. Myosin V from

Drosophila reveals diversity of motor mechanisms within the myosin V family. J

Biol Chem 280:30594-30603.

23. Reck-Peterson, S. L., M. J. Tyska, P. J. Novick, and M. S. Mooseker. 2001. The

yeast class V myosins, Myo2p and Myo4p, are nonprocessive actin-based motors.

J Cell Biol 153:1121-1126.

24. Schott, D., J. Ho, D. Pruyne, and A. Bretscher. 1999. The COOH-terminal domain

of Myo2p, a yeast myosin V, has a direct role in secretory vesicle targeting. J Cell

Biol 147:791-808.

25. Stevens, R. C., and T. N. Davis. 1998. Mlc1p is a light chain for the

unconventional myosin Myo2p in Saccharomyces cerevisiae. J Cell Biol 142:711-

722.

26. Wagner, W., P. Bielli, S. Wacha, and A. Ragnini-Wilson. 2002. Mlc1p promotes

septum closure during cytokinesis via the IQ motifs of the vesicle motor Myo2p.

Embo J 21:6397-6408.

27. May, K. M., F. Z. Watts, N. Jones, and J. S. Hyams. 1997. Type II myosin

involved in cytokinesis in the fission yeast, Schizosaccharomyces pombe. Cell

Motil Cytoskeleton 38:385-396.

28. Terrak, M., G. Wu, W. F. Stafford, R. C. Lu, and R. Dominguez. 2003. Two

distinct myosin light chain structures are induced by specific variations within the

bound IQ motifs-functional implications. Embo J 22:362-371.

29. Bielli, P., E. C. Casavola, A. Biroccio, A. Urbani, and A. Ragnini-Wilson. 2006.

GTP drives myosin light chain 1 interaction with the class V myosin Myo2 IQ

motifs via a Sec2 RabGEF-mediated pathway. Mol Microbiol 59:1576-1590.

30. Bateman, A., L. Coin, R. Durbin, R. D. Finn, V. Hollich, S. Griffiths-Jones, A.

Khanna, M. Marshall, S. Moxon, E. L. Sonnhammer, D. J. Studholme, C. Yeats,

and S. R. Eddy. 2004. The Pfam protein families database. Nucleic Acids Res

32:D138-141.

31. Martin, S. R., and P. M. Bayley. 2004. Calmodulin bridging of IQ motifs in

myosin-V. FEBS Lett 567:166-170.

32. Rhoads, A. R., and F. Friedberg. 1997. Sequence motifs for calmodulin

recognition. Faseb J 11:331-340.

33. Bahler, M., and A. Rhoads. 2002. Calmodulin signaling via the IQ motif. FEBS

Lett 513:107-113.

34. Chin, D., and A. R. Means. 2000. Calmodulin: a prototypical calcium sensor.

Trends Cell Biol 10:322-328.

Page 118: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

105

35. Yamniuk, A. P., and H. J. Vogel. 2004. Calmodulin's flexibility allows for

promiscuity in its interactions with target proteins and peptides. Mol Biotechnol

27:33-57.

36. Vogel, H. J., R. D. Brokx, and H. Ouyang. 2002. Calcium-binding proteins.

Methods Mol Biol 172:3-20.

37. Soderling, T. R. 1999. The Ca-calmodulin-dependent protein kinase cascade.

Trends Biochem Sci 24:232-236.

38. Jurado, L. A., P. S. Chockalingam, and H. W. Jarrett. 1999. Apocalmodulin.

Physiol Rev 79:661-682.

39. Persechini, A., P. M. Stemmer, and I. Ohashi. 1996. Localization of unique

functional determinants in the calmodulin lobes to individual EF hands. J Biol

Chem 271:32217-32225.

40. VanBerkum, M. F., S. E. George, and A. R. Means. 1990. Calmodulin activation

of target enzymes. Consequences of deletions in the central helix. J Biol Chem

265:3750-3756.

41. Craig, T. A., D. M. Watterson, F. G. Prendergast, J. Haiech, and D. M. Roberts.

1987. Site-specific mutagenesis of the alpha-helices of calmodulin. Effects of

altering a charge cluster in the helix that links the two halves of calmodulin. J Biol

Chem 262:3278-3284.

42. Lukas, T. J., W. H. Burgess, F. G. Prendergast, W. Lau, and D. M. Watterson.

1986. Calmodulin binding domains: characterization of a phosphorylation and

calmodulin binding site from myosin light chain kinase. Biochemistry 25:1458-

1464.

43. Findlay, W. A., S. R. Martin, K. Beckingham, and P. M. Bayley. 1995. Recovery

of native structure by calcium binding site mutants of calmodulin upon binding of

sk-MLCK target peptides. Biochemistry 34:2087-2094.

44. Findlay, W. A., M. J. Gradwell, and P. M. Bayley. 1995. Role of the N-terminal

region of the skeletal muscle myosin light chain kinase target sequence in its

interaction with calmodulin. Protein Sci 4:2375-2382.

45. Wintrode, P. L., and P. L. Privalov. 1997. Energetics of target peptide recognition

by calmodulin: a calorimetric study. J Mol Biol 266:1050-1062.

46. Brokx, R. D., M. M. Lopez, H. J. Vogel, and G. I. Makhatadze. 2001. Energetics

of target peptide binding by calmodulin reveals different modes of binding. J Biol

Chem 276:14083-14091.

47. Osawa, M., H. Tokumitsu, M. B. Swindells, H. Kurihara, M. Orita, T. Shibanuma,

T. Furuya, and M. Ikura. 1999. A novel target recognition revealed by calmodulin

in complex with Ca2+-calmodulin-dependent kinase kinase. Nat Struct Biol

6:819-824.

48. Meador, W. E., A. R. Means, and F. A. Quiocho. 1993. Modulation of calmodulin

plasticity in molecular recognition on the basis of x-ray structures. Science

262:1718-1721.

49. Meador, W. E., A. R. Means, and F. A. Quiocho. 1992. Target enzyme

recognition by calmodulin: 2.4 A structure of a calmodulin-peptide complex.

Science 257:1251-1255.

50. Kurokawa, H., M. Osawa, H. Kurihara, N. Katayama, H. Tokumitsu, M. B.

Swindells, M. Kainosho, and M. Ikura. 2001. Target-induced conformational

Page 119: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

106

adaptation of calmodulin revealed by the crystal structure of a complex with

nematode Ca(2+)/calmodulin-dependent kinase kinase peptide. J Mol Biol

312:59-68.

51. Clapperton, J. A., S. R. Martin, S. J. Smerdon, S. J. Gamblin, and P. M. Bayley.

2002. Structure of the complex of calmodulin with the target sequence of

calmodulin-dependent protein kinase I: studies of the kinase activation

mechanism. Biochemistry 41:14669-14679.

52. Aoyagi, M., A. S. Arvai, J. A. Tainer, and E. D. Getzoff. 2003. Structural basis

for endothelial nitric oxide synthase binding to calmodulin. Embo J 22:766-775.

53. Yamauchi, E., T. Nakatsu, M. Matsubara, H. Kato, and H. Taniguchi. 2003.

Crystal structure of a MARCKS peptide containing the calmodulin-binding

domain in complex with Ca2+-calmodulin. Nat Struct Biol 10:226-231.

54. Matsubara, M., T. Nakatsu, H. Kato, and H. Taniguchi. 2004. Crystal structure of

a myristoylated CAP-23/NAP-22 N-terminal domain complexed with

Ca(2+)/calmodulin. Embo J 23:712-718.

55. Vetter, S. W., and E. Leclerc. 2003. Novel aspects of calmodulin target

recognition and activation. Eur J Biochem 270:404-414.

56. Terrak, M., L. R. Otterbein, G. Wu, L. A. Palecanda, R. C. Lu, and R.

Dominguez. 2002. Crystallization, X-ray characterization and selenomethionine

phasing of Mlc1p bound to IQ motifs from myosin V. Acta Crystallogr D Biol

Crystallogr 58:1882-1885.

57. Terrak, M., G. Rebowski, R. C. Lu, Z. Grabarek, and R. Dominguez. 2005.

Structure of the light chain-binding domain of myosin V. Proc Natl Acad Sci U S

A 102:12718-12723.

58. Cheatham, T. E., and P. A. Kollman. 2000. Molecular dynamics simulation of

nucleic acids. Annual Review of Physical Chemistry 51:435-471.

59. Karplus, M., and J. A. McCammon. 2002. Molecular dynamics simulations of

biomolecules. Nat Struct Biol 9:646-652.

60. Norberg, J., and L. Nilsson. 2002. Molecular dynamics applied to nucleic acids.

Accounts Chem Res 35:465-472.

61. Moraitakis, G., A. G. Purkiss, and J. M. Goodfellow. 2003. Simulated dynamics

and biological macromolecules. Rep Prog Phys 66:383-406.

62. Straatsma, T. P., and J. A. Mccammon. 1992. Computational Alchemy. Annual

Review of Physical Chemistry 43:407-435.

63. Wang, W., O. Donini, C. M. Reyes, and P. A. Kollman. 2001. Biomolecular

simulations: recent developments in force fields, simulations of enzyme catalysis,

protein-ligand, protein-protein, and protein-nucleic acid noncovalent interactions.

Annu Rev Biophys Biomol Struct 30:211-243.

64. Hansson, T., C. Oostenbrink, and W. van Gunsteren. 2002. Molecular dynamics

simulations. Curr Opin Struct Biol 12:190-196.

65. McCammon, J. A., B. R. Gelin, and M. Karplus. 1977. Dynamics of folded

proteins. Nature 267:585-590.

66. Levitt, M., and R. Sharon. 1988. Accurate simulation of protein dynamics in

solution. Proc Natl Acad Sci U S A 85:7557-7561.

67. Mackerell, A. D., Jr. 2004. Empirical force fields for biological macromolecules:

overview and issues. J Comp Chem 25:1584-1604.

Page 120: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

107

68. Karplus, M. 2002. Molecular dynamics simulations of biomolecules. Acc Chem

Res 35:321-323.

69. Karplus, M., A. T. Brunger, R. Elber, and J. Kuriyan. 1987. Molecular dynamics:

applications to proteins. Cold Spring Harb Symp Quant Biol 52:381-390.

70. Karplus, M., and J. Kuriyan. 2005. Molecular dynamics and protein function.

Proc Natl Acad Sci U S A 102:6679-6685.

71. Lu, D., A. Aksimentiev, A. Y. Shih, E. Cruz-Chu, P. L. Freddolino, A. Arkhipov,

and K. Schulten. 2006. The role of molecular modeling in bionanotechnology.

Phys Biol 3:S40-53.

72. van Gunsteren, W. F., and H. J. C. Berenden. 1990. Computer simulation of

molecular dynamics: methodology, applications and perspectives in chemistry.

Angew Chem Int Ed Engl 29:992-1023.

73. Leach, A. R. 2001. Molecular modeling: principles and applications. Prentice

Hall.

74. Vorherr, T., O. Kessler, A. Mark, and E. Carafoli. 1992. Construction and

molecular dynamics simulation of calmodulin in the extended and in a bent

conformation. Eur J Biochem 204:931-937.

75. Mehler, E. L., J. L. Pascual-Ahuir, and H. Weinstein. 1991. Structural dynamics

of calmodulin and troponin C. Protein Eng 4:625-637.

76. Weinstein, H., and E. L. Mehler. 1994. Ca(2+)-binding and structural dynamics in

the functions of calmodulin. Annu Rev Physiol 56:213-236.

77. van der Spoel, D., B. L. de Groot, S. Hayward, H. J. Berendsen, and H. J. Vogel.

1996. Bending of the calmodulin central helix: a theoretical study. Protein Sci

5:2044-2053.

78. Yang, C., and K. Kuczera. 2002. Molecular dynamics simulations of calcium-free

calmodulin in solution. J Biomol Struct Dyn 19:801-819.

79. Komeiji, Y., Y. Ueno, and M. Uebayasi. 2002. Molecular dynamics simulations

revealed Ca(2+)-dependent conformational change of Calmodulin. FEBS Lett

521:133-139.

80. Vigil, D., S. C. Gallagher, J. Trewhella, and A. E. Garcia. 2001. Functional

dynamics of the hydrophobic cleft in the N-domain of calmodulin. Biophys J

80:2082-2092.

81. Shepherd, C. M., and H. J. Vogel. 2004. A molecular dynamics study of Ca2+-

calmodulin: evidence of interdomain coupling and structural collapse on the

nanosecond timescale. Biophys J 87:780-791.

82. Yang, C., G. S. Jas, and K. Kuczera. 2001. Structure and dynamics of calcium-

activated calmodulin in solution. J Biomol Struct Dyn 19:247-271.

83. Fiorin, G., R. R. Biekofsky, A. Pastore, and P. Carloni. 2005. Unwinding the

helical linker of calcium-loaded calmodulin: a molecular dynamics study. Proteins

61:829-839.

84. Likic, V. A., P. R. Gooley, T. P. Speed, and E. E. Strehler. 2005. A statistical

approach to the interpretation of molecular dynamics simulations of calmodulin

equilibrium dynamics. Protein Sci 14:2955-2963.

85. Wriggers, W., E. Mehler, F. Pitici, H. Weinstein, and K. Schulten. 1998. Structure

and dynamics of calmodulin in solution. Biophys J 74:1622-1639.

Page 121: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

108

86. Goto, K., A. Toyama, H. Takeuchi, K. Takayama, T. Saito, M. Iwamoto, J. Z.

Yeh, and T. Narahashi. 2004. Ca2+ binding sites in calmodulin and troponin C

alter interhelical angle movements. FEBS Lett 561:51-57.

87. Project, E., R. Friedman, E. Nachliel, and M. Gutman. 2006. A molecular

dynamics study of the effect of Ca2+ removal on calmodulin structure. Biophys J

90:3842-3850.

88. Yamaotsu, N., M. Suga, and S. Hirono. 2001. Molecular dynamics simulation of

the calmodulin-trifluoperazine complex in aqueous solution. Biopolymers 58:410-

421.

89. Turjanski, A. G., D. A. Estrin, R. E. Rosenstein, J. E. McCormick, S. R. Martin,

A. Pastore, R. R. Biekofsky, and V. Martorana. 2004. NMR and molecular

dynamics studies of the interaction of melatonin with calmodulin. Protein Sci

13:2925-2938.

90. Prabhu, N. V., A. L. Lee, A. J. Wand, and K. A. Sharp. 2003. Dynamics and

entropy of a calmodulin-peptide complex studied by NMR and molecular

dynamics. Biochemistry 42:562-570.

91. Yang, C., and K. Kuczera. 2002. Molecular dynamics simulations of a

calmodulin-peptide complex in solution. J Biomol Struct Dyn 20:179-197.

92. Yang, C., G. S. Jas, and K. Kuczera. 2004. Structure, dynamics and interaction

with kinase targets: computer simulations of calmodulin. Biochim Biophys Acta

1697:289-300.

93. Rodinger, T., and R. Pomes. 2005. Enhancing the accuracy, the efficiency and the

scope of free energy simulations. Curr Opin Struct Biol 15:164-170.

94. Srinivasan, J., T. E. Cheatham, P. Cieplak, P. A. Kollman, and D. A. Case. 1998.

Continuum solvent studies of the stability of DNA, RNA, and phosphoramidate -

DNA helices. J Am Chem Soc 120:9401-9409.

95. Hazel, P., G. N. Parkinson, and S. Neidle. 2006. Predictive modelling of topology

and loop variations in dimeric DNA quadruplex structures. Nucleic Acids Res

34:2117-2127.

96. Cheatham, T. E., 3rd, J. Srinivasan, D. A. Case, and P. A. Kollman. 1998.

Molecular dynamics and continuum solvent studies of the stability of polyG-

polyC and polyA-polyT DNA duplexes in solution. J Biomol Struct Dyn 16:265-

280.

97. Lee, M. R., Y. Duan, and P. A. Kollman. 2000. Use of MM-PB/SA in estimating

the free energies of proteins: application to native, intermediates, and unfolded

villin headpiece. Proteins 39:309-316.

98. Villacanas, O., and J. Rubio-Martinez. 2006. Reducing CDK4/6-p16(INK4a)

interface. Computational alanine scanning of a peptide bound to CDK6 protein.

Proteins 63:797-810.

99. Gohlke, H., C. Kiel, and D. A. Case. 2003. Insights into protein-protein binding

by binding free energy calculation and free energy decomposition for the Ras-Raf

and Ras-RalGDS complexes. J Mol Biol 330:891-913.

100. Massova, I., and P. A. Kollman. 1999. Computational alanine scanning to probe

protein-protein interactions: A novel approach to evaluate binding free energies.

Journal of the American Chemical Society 121:8133-8143.

Page 122: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

109

101. Chong, L. T., Y. Duan, L. Wang, I. Massova, and P. A. Kollman. 1999. Molecular

dynamics and free-energy calculations applied to affinity maturation in antibody

48G7. Proc Natl Acad Sci U S A 96:14330-14335.

102. Huo, S., I. Massova, and P. A. Kollman. 2002. Computational alanine scanning of

the 1:1 human growth hormone-receptor complex. J Comput Chem 23:15-27.

103. Huo, S., J. Wang, P. Cieplak, P. A. Kollman, and I. D. Kuntz. 2002. Molecular

dynamics and free energy analyses of cathepsin D-inhibitor interactions: insight

into structure-based ligand design. J Med Chem 45:1412-1419.

104. Kuhn, B., and P. A. Kollman. 2000. Binding of a diverse set of ligands to avidin

and streptavidin: an accurate quantitative prediction of their relative affinities by a

combination of molecular mechanics and continuum solvent models. J Med Chem

43:3786-3791.

105. Reyes, C. M., and P. A. Kollman. 2000. Structure and thermodynamics of RNA-

protein binding: using molecular dynamics and free energy analyses to calculate

the free energies of binding and conformational change. J Mol Biol 297:1145-

1158.

106. Tsui, V., and D. A. Case. 2001. Calculations of the absolute free energies of

binding between RNA and metal ions using molecular dynamics simulations and

continuum electrostatics. J Phys Chem B 105:11314-11325.

107. Gouda, H., I. D. Kuntz, D. A. Case, and P. A. Kollman. 2003. Free energy

calculations for theophylline binding to an RNA aptamer: MM-PBSA and

comparison of thermodynamic integration methods. Biopolymers 68:16-34.

108. Berendsen, H. J. C., D. van der Spoel, and R. van Drunen. 1995. GROMACS: A

message-passing parallel molecular dynamics implementation. Comp Phys Comm

91:43-56.

109. Lindahl, E., B. Hess, and D. van der Spoel. 2001. Gromacs 3.0: A package for

molecular simulation and trajectory analysis. J Mol Mod 7:306-317.

110. van Der Spoel, D., E. Lindahl, B. Hess, A. R. van Buuren, E. Apol, P. J.

Meulenhoff, D. P. Tieleman, A. L. T. M. Sijbers, A. K. Feenstra, R. van Drunen,

and H. J. C. Berendsen. 2004. GROningen Machine for Molecular Simulations.

Groningen.

111. van Gunsteren, W. F., S. R. Billeter, A. A. Eising, P. H. Hunenberger, P. Kruger,

A. E. Mark, W. R. P. Scott, and I. G. Tironi. 1996. Biomolecular Simulation: The

GROMOS96 Manual and User Guide. Vdf Hochschulverlag AG an der ETH

Zurich, Zurich:1-1024.

112. Berman, H. M., T. Battistuz, T. N. Bhat, W. F. Bluhm, P. E. Bourne, K.

Burkhardt, Z. Feng, G. L. Gilliland, L. Iype, S. Jain, P. Fagan, J. Marvin, D.

Padilla, V. Ravichandran, B. Schneider, N. Thanki, H. Weissig, J. D. Westbrook,

and C. Zardecki. 2002. The Protein Data Bank. Acta Crystallogr D Biol

Crystallogr 58:899-907.

113. Xiang, Z. 2002. JACKAL: A Protein Structure Modeling Package. Columbia

University, New York, NY.

114. Berendsen, H. J. C., J. P. M. Postma, W. F. van Gunsteren, and J. Hermans. 1969.

Interaction Models for Water in Relation to Protein Hydration. Nature 224:175-

177.

Page 123: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

110

115. Hess, B., H. Bekker, H. J. C. Berendsen, and J. G. E. M. Fraaije. 1997. LINCS: A

linear constraint solver for molecular simulations. J Comp Chem 18:1463-1472.

116. Miyamoto, S., and P. A. Kollman. 1992. SETTLE: An Analytical Version of the

SHAKE and RATTLE Algorithms for Rigid water models. J Comp Chem 13:952-

962.

117. Berendsen, H. J. C., J. P. M. Postma, A. DiNola, and J. R. Haak. 1984. Molecular

dynamics with coupling to an external bath. J Chem Phys 81:3684-3690.

118. Essman, U., L. Perela, M. L. Berkowitz, T. Darden, H. Lee, and L. G. Pederse.

1995. A smooth particle mesh Ewald method. J Chem Phys 103:8577-8592.

119. Guex, N., and M. C. Peitsch. 1997. SWISS-MODEL and the Swiss-PdbViewer:

an environment for comparative protein modeling. Electrophoresis 18:2714-2723.

120. Humphrey, W., A. Dalke, and K. Schulten. 1996. VMD: visual molecular

dynamics. J Mol Graph 14:33-38, 27-38.

121. Frishman, D., and P. Argos. 1995. Knowledge-based protein secondary structure

assignment. Proteins 23:566-579.

122. Baker, N. A., D. Sept, S. Joseph, M. J. Holst, and J. A. McCammon. 2001.

Electrostatics of nanosystems: application to microtubules and the ribosome. Proc

Natl Acad Sci U S A 98:10037-10041.

123. Sharp, K. A., and B. Honig. 1990. Electrostatic interactions in macromolecules:

theory and applications. Annu Rev Biophys Biophys Chem 19:301-332.

124. King, G., F. S. Lee, and A. Warshel. 1991. Microscopic Simulations of

Macroscopic Dielectric-Constants of Solvated Proteins. Journal of Chemical

Physics 95:4366-4377.

125. Warshel, A., and J. Aqvist. 1991. Electrostatic energy and macromolecular

function. Annu Rev Biophys Biophys Chem 20:267-298.

126. Kasper, P., P. Christen, and H. Gehring. 2000. Empirical calculation of the

relative free energies of peptide binding to the molecular chaperone DnaK.

Proteins 40:185-192.

127. Sharp, K. A. 1998. Calculation of HyHel10-lysozyme binding free energy

changes: effect of ten point mutations. Proteins 33:39-48.

128. Novotny, J., R. E. Bruccoleri, M. Davis, and K. A. Sharp. 1997. Empirical free

energy calculations: a blind test and further improvements to the method. J Mol

Biol 268:401-411.

129. Eisenhaber, F., P. Lijnzaad, P. Argos, C. Sander, and M. Scharf. 1995. The

Double Cube Lattice Method: Efficient Approaches to Numerical Integration of

Surface Area and Volume and to Dot Surface Contouring of Molecular

Assemblies. J Comp Chem 16:273-284.

130. Levitt, M., C. Sander, and P. S. Stern. 1983. The normal modes of a protein:

native bovine pancreatic trypsin inhibitor. Int J Quantum Chem Quantum Biol

Symp 10:181-199.

131. Go, N., T. Noguti, and T. Nishikawa. 1983. Dynamics of a small globular protein

in terms of low-frequency vibrational modes. Proc Natl Acad Sci U S A 80:3696-

3700.

132. Brooks, B., and M. Karplus. 1983. Harmonic dynamics of proteins: normal modes

and fluctuations in bovine pancreatic trypsin inhibitor. Proc Natl Acad Sci U S A

80:6571-6575.

Page 124: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

111

133. McQuarrie, D. A. 1973. Statistical thermodynamics. Harper & Row, New York.

134. Kuboniwa, H., N. Tjandra, S. Grzesiek, H. Ren, C. B. Klee, and A. Bax. 1995.

Solution structure of calcium-free calmodulin. Nat Struct Biol 2:768-776.

135. Zhang, M., T. Tanaka, and M. Ikura. 1995. Calcium-induced conformational

transition revealed by the solution structure of apo calmodulin. Nat Struct Biol

2:758-767.

136. Fallon, J. L., and F. A. Quiocho. 2003. A closed compact structure of native

Ca(2+)-calmodulin. Structure (Camb) 11:1303-1307.

137. Hayashi, N., M. Matsubara, A. Takasaki, K. Titani, and H. Taniguchi. 1998. An

expression system of rat calmodulin using T7 phage promoter in Escherichia coli.

Protein Expr Purif 12:25-28.

138. Gopalakrishna, R., and W. B. Anderson. 1982. Ca2+-induced hydrophobic site on

calmodulin: application for purification of calmodulin by phenyl-Sepharose

affinity chromatography. Biochem Biophys Res Commun 104:830-836.

139. Garbers, D. L., J. R. Hansbrough, E. W. Radany, R. V. Hyne, and G. S. Kopf.

1980. Purification and characterization of calmodulin from sea urchin

spermatozoa. J Reprod Fertil 59:377-381.

140. Ganoth, A., E. Nachliel, R. Friedman, and M. Gutman. 2006. Molecular dynamics

study of a calmodulin-like protein with an IQ peptide: Spontaneous refolding of

the protein around the peptide. Proteins 64:133-146.

141. Petegem, F. V., F. C. Chatelain, and D. L. Minor. 2005. Insights into voltage-

gated calcium channel regulation from the structure of the Ca(V)1.2 IQ domain-

Ca(2+)/calmodulin complex. Nat Struct Mol Biol.

142. Sitkoff, D., K. Sharp, and B. Honig. 1994. Accurate calculation of hydration free

energies using macroscopic solvent models. J Phys Chem 98:1978-1988.

143. Cox, J. A., M. Comte, J. E. Fitton, and W. F. DeGrado. 1985. The interaction of

calmodulin with amphiphilic peptides. J Biol Chem 260:2527-2534.

144. Crivici, A., and M. Ikura. 1995. Molecular and structural basis of target

recognition by calmodulin. Annu Rev Biophys Biomol Struct 24:85-116.

145. Ikura, M., G. M. Clore, A. M. Gronenborn, G. Zhu, C. B. Klee, and A. Bax. 1992.

Solution structure of a calmodulin-target peptide complex by multidimensional

NMR. Science 256:632-638.

146. Misra, V. K., J. L. Hecht, K. A. Sharp, R. A. Friedman, and B. Honig. 1994. Salt

effects on protein-DNA interactions. The lambda cI repressor and EcoRI

endonuclease. J Mol Biol 238:264-280.

147. Misra, V. K., K. A. Sharp, R. A. Friedman, and B. Honig. 1994. Salt effects on

ligand-DNA binding. Minor groove binding antibiotics. J Mol Biol 238:245-263.

148. Sharp, K. A. 1996. Electrostatic interactions in hirudin-thrombin binding. Biophys

Chem 61:37-49.

149. Andre, I., T. Kesvatera, B. Jonsson, and S. Linse. 2006. Salt enhances

calmodulin-target interaction. Biophys J 90:2903-2910.

150. Lee, A. L., S. A. Kinnear, and A. J. Wand. 2000. Redistribution and loss of side

chain entropy upon formation of a calmodulin-peptide complex. Nat Struct Biol

7:72-77.

Page 125: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

112

151. Duneau, J. P., N. Garnier, and M. Genest. 1997. Insight into signal transduction:

structural alterations in transmembrane helices probed by multi-1 ns molecular

dynamics simulations. J Biomol Struct Dyn 15:555-572.

152. Duneau, J. P., S. Crouzy, Y. Chapron, and M. Genest. 1999. Dynamics of the

transmembrane domain of the ErbB-2 receptor. Theor Chem Acc 101:87-91.

153. Shepherd, C. M., D. van der Spoel, and H. J. Vogel. 2004. Molecular dynamics

simulations of peptides from the central domain of smooth muscle caldesmon. J

Biomol Struct Dyn 21:555-566.

154. Lee, K. H., D. R. Benson, and K. Kuczera. 2000. Transitions from alpha to pi

helix observed in molecular dynamics simulations of synthetic peptides.

Biochemistry 39:13737-13747.

155. Weaver, T., and L. Banaszak. 1996. Crystallographic studies of the catalytic and a

second site in fumarase C from Escherichia coli. Biochemistry 35:13955-13965.

156. Sprang, S. R., S. G. Withers, E. J. Goldsmith, R. J. Fletterick, and N. B. Madsen.

1991. Structural basis for the activation of glycogen phosphorylase b by

adenosine monophosphate. Science 254:1367-1371.

157. Boyington, J. C., B. J. Gaffney, and L. M. Amzel. 1993. The three-dimensional

structure of an arachidonic acid 15-lipoxygenase. Science 260:1482-1486.

158. Vidgren, J., L. A. Svensson, and A. Liljas. 1994. Crystal structure of catechol O-

methyltransferase. Nature 368:354-358.

159. Bullock, T. L., K. Breddam, and S. J. Remington. 1996. Peptide aldehyde

complexes with wheat serine carboxypeptidase II: implications for the catalytic

mechanism and substrate specificity. J Mol Biol 255:714-725.

160. Liu, G., D. K. Sukumaran, D. Xu, Y. Chiang, T. Acton, S. Goldsmith-Fischman,

B. Honig, G. T. Montelione, and T. Szyperski. 2004. NMR structure of the

hypothetical protein NMA1147 from Neisseria meningitidis reveals a distinct 5-

helix bundle. Proteins 55:756-758.

161. Musso, G. F., S. Patthi, T. C. Ryskamp, S. Provow, E. T. Kaiser, and G.

Velicelebi. 1988. Development of helix-based vasoactive intestinal peptide

analogues: identification of residues required for receptor interaction.

Biochemistry 27:8174-8181.

162. Eyal, E., S. Gerzon, V. Potapov, M. Edelman, and V. Sobolev. 2005. The limit of

accuracy of protein modeling: influence of crystal packing on protein structure. J

Mol Biol 351:431-442.

163. DePristo, M. A., P. I. de Bakker, and T. L. Blundell. 2004. Heterogeneity and

inaccuracy in protein structures solved by X-ray crystallography. Structure

12:831-838.

164. Espindola, F. S., E. M. Espreafico, M. V. Coelho, A. R. Martins, F. R. Costa, M.

S. Mooseker, and R. E. Larson. 1992. Biochemical and immunological

characterization of p190-calmodulin complex from vertebrate brain: a novel

calmodulin-binding myosin. J Cell Biol 118:359-368.

165. Yildiz, A., H. Park, D. Safer, Z. Yang, L. Q. Chen, P. R. Selvin, and H. L.

Sweeney. 2004. Myosin VI steps via a hand-over-hand mechanism with its lever

arm undergoing fluctuations when attached to actin. J Biol Chem 279:37223-

37226.

Page 126: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

113

166. Yildiz, A., J. N. Forkey, S. A. McKinney, T. Ha, Y. E. Goldman, and P. R.

Selvin. 2003. Myosin V walks hand-over-hand: single fluorophore imaging with

1.5-nm localization. Science 300:2061-2065.

167. Yildiz, A., and P. R. Selvin. 2005. Fluorescence imaging with one nanometer

accuracy: application to molecular motors. Acc Chem Res 38:574-582.

168. Forkey, J. N., M. E. Quinlan, M. A. Shaw, J. E. Corrie, and Y. E. Goldman. 2003.

Three-dimensional structural dynamics of myosin V by single-molecule

fluorescence polarization. Nature 422:399-404.

169. Walker, M. L., S. A. Burgess, J. R. Sellers, F. Wang, J. A. Hammer, 3rd, J.

Trinick, and P. J. Knight. 2000. Two-headed binding of a processive myosin to F-

actin. Nature 405:804-807.

170. Burgess, S., M. Walker, F. Wang, J. R. Sellers, H. D. White, P. J. Knight, and J.

Trinick. 2002. The prepower stroke conformation of myosin V. J Cell Biol

159:983-991.

171. Houdusse, A., A. G. Szent-Gyorgyi, and C. Cohen. 2000. Three conformational

states of scallop myosin S1. Proc Natl Acad Sci U S A 97:11238-11243.

Page 127: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

סימולציות מחשב של קומפלקסים

בין חלבון לפפטיד תוך שימוש

מודלמערכתבשרשרת הקלה של מיוזין כ

לשם קבלת התוארחיבור

"דוקטור לפילוסופיה"

מאת

אסף גנות

אביב-הוגש לסנאט אוניברסיטת תל

2006 יולי

Page 128: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

עבודה זו נעשתה בהדרכת

מנחם גוטמן' פרופ

Page 129: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

תקציר

. מית של מיוזיןיכ-הוא רכיב חיוני במערכת המכנו, חבר במשפחת הקלמודולין, Mlc1pהחלבון

ואר של ו קושרים את מתחם הצMlc1pשישה חלבוני , Saccharomyces cerevisiaeבשמר ההנצה

לבין Mlc1p מבנים של מספר קומפלקסים בין החלבון .IQ פפטידים מסוג המורכב משישה, Vמיוזין

, המחקר הנוכחי מרחיב את ההבנה שלנו.X של קרני קריסטלוגרפיהבאמצעות נפתרו IQפפטידי

דינמיקה ואנרגטיקה של , על מבנים,י דינמיקה מולקולרית וחישובים נוספים"באמצעות מחקר ע

המחקר , הרחבה ניכרת של המידע המבני האצור במבנים הגבישייםל בנוסף. Mlc1p-IQהקומפלקסים

כיצד קיפול באתר מסוים ו מצאנ. על סיב האקטיןVמספק הבנה אטומית אודות דגם התנועה של מיוזין

מנוף תקינה שלה תופעולאת ובכך להבטיח , כואר המיוזין יכול לשמש כמפרק גמיש לאורוהממוקם על צו

.זרוע המיוזין

4IQ ו 2IQסימולציות של דינמיקה מולקולרית של הפפטידים ) I: (כוללת ארבעה פרקיםהתיזה

-Mlc1p השוואה מפורטת בין הסימולציות של שני הקומפלקסים ) Mlc1p;) IIבקומפלקס עם החלבון

IQ;) III ( סימולציות של דינמיקה מולקולרית של פפטידיIQחופשיים ;) IV (של הערכה מחודשת

.Vתחם הקושר שרשראות קלות של מיוזין המבנה של המ

פלקסיםמבפרק הראשון של התיזה אנחנו מציגים סימולציות של דינמיקה מולקולרית של הקו

Mlc1p-IQ4 ו Mlc1p-IQ2 ,הקומפלקס.לאחר רלקסציה שלהם בתמיסת מלח פיזיולוגית Mlc1p-

IQ2 בות הסימולציה אופיין בשינוי עיקריים שלו ובעקהעבר רלקסציה ללא איבוד של כוחות האריזה

עבר תהליך של קיפול מחדש במהלכו החלבון שינה את , 4IQעם הפפטיד , פלקס הנוסףמ הקו.מבני מוגבל

אנו מציגים השוואה , בפרק השני של התיזה. מבנהו ממתוח לקומפקטי והפפטיד התקפל לשני חלקים

, בצענו השוואה מקיפה בין המבנים. קסיםפלממפורטת בין סימולציות הדינמיקה המולקולרית של שני הקו

אנליזת האנרגיה כוללת ניתוח של . פפטיד- חלבוןנמיקה והאנרגיה החופשית של האינטראקציהיהד

. פטיד וציון תרומתם של רכיבים אנרגטיים שונים והפחלבון בין הפלקסיםמהכוחות הפועלים על הקו

ך פרקי זמן משתנים ובתנאים מש חופשיים לIQהפרק השלישי של התיזה עוסק בסימולציות של פפטידי

שלושינויים בין מבני התמיסה, שינויים בין הקונפורמציות של פפטידים בנוכחות ובהעדר חלבון. שונים

פלקסים מקוה התוצאות של סימולציות הדינמיקה המולקולרית של . נידונים, החופשייםIQשני פפטידי ה

ושל פפטידים חופשיים מאפשרות לבצע הערכה של המבנה של המתחם הקושר שרשראות פפטיד -חלבון

. IQלבין פפטידי ) Mlc1pכמו (לקסים בין שרשראות קלות פה קומשהמורכב משי, Vקלות של מיוזין

במרכז המודל המוצע עומדת התפיסה שהמתחם הקושר . הערכה זו מוצגת בפרק הרביעי של התיזה

ולכן המודל לוקח בחשבון את יכולתם של החלבון ,נמיתי הוא ישות תאית דVוזין שרשראות קלות של מי

Mlc1p ושל פפטידי IQלהתגמש ולהתכופף באופן הדדי .

מתאר סימולציות של דינמיקה המאמר הראשון. תוצאות התיזה מסוכמות בשלושה מאמרים

4IQ-p1Mlc) . and M, Friedman.R, Nachliel. E, .A, Ganothמולקולרית של הקומפלקס

Gutman. 2006. Molecular dynamics study of a calmodulin-like protein with an IQ

Page 130: Computer Simulations of Protein-Peptide Complexes Using the Myosin Light Chain Protein as a

peptide: Spontaneous refolding of the protein around the peptide. Proteins 64:133-

ללת הכו, Mlc1p-IQ2 ו Mlc1p-IQ4 מציג השוואה נרחבת בין הקומפלקסים המאמר השני;)146

V ). R, .A, anothGאנליזה אנרגטית והערכה מחודשת של המתחם הקושר שרשראות קלות של מיוזין

Friedman, E. Nachliel, and M. Gutman. 2006. A molecular dynamics study and free

energy analysis of complexes between the Mlc1p protein and two IQ motif peptides.

Biophys J 91:2436-2450(; העוסק בסימולציות של פפטידי , המאמר השלישיIQבהכנה, חופשיים.

IQהמחקר הנוכחי מהווה מחקר חדשני של סימולציות דינמיקה מולקולרית של פפטידי

- בתהליכים מכנוIQפפטידי תפקידם העיקרי של לאור . קלמודולין ובהעדרו בקומפלקס עם חלבון דמוי

ה מספקת תיאור ברזולוציה אטומית של האינטראקציות שלהם עם קלמודולין וחלבונים התיז, כימיים

ולציות דינמיקה מולקולרית ונגמר מחיל בסיתמשרטט טיול ממוחשב המ המחקר. דמויי קלמודולין

מדענים ששיתוף פעולה בין ומדגים ,Vבתובנות אודות המתחם הקושר שרשראות קלות של מיוזין

.מבנה-נה טובה יותר של יחסי תפקודהבפיזיקאים יכול לתרום ל-ביובין להגרפילוטקריסהעוסקים ב