c 2006 by prasanth sankar. all rights...

118
c 2006 by Prasanth Sankar. All rights reserved.

Upload: others

Post on 06-Jul-2020

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

c© 2006 by Prasanth Sankar. All rights reserved.

Page 2: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

PHENOMENOLOGICAL MODELS OF MOTOR PROTEINS

BY

PRASANTH SANKAR

M. S., University of Illinois at Urbana-Champaign, 2000

DISSERTATION

Submitted in partial fulfillment of the requirementsfor the degree of Doctor of Philosophy in Physics

in the Graduate College of theUniversity of Illinois at Urbana-Champaign, 2006

Urbana, Illinois

Page 3: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Abstract

Motor proteins produce directional movement and mechanical output in cells by converting

chemical energy into mechanical work. Innovative applications of physics experimental tech-

niques to the study of motor proteins have generated a wealth of data. By careful analysis

and modeling of this data, we try to gain a quantitative understanding of this biological

system. We have proposed a way to incorporate the biochemical data on motor proteins

into the modeling and thus minimize the number of fitting parameters and increase the rel-

evance of the models. Mechanical understanding of motor proteins are obtained by using

an experimental technique of observing the motion of a large cargo elastically attached to

it. On modeling the effect of the elastic motor-cargo link on the motion of the cargo, we

see that the conclusions on motor mechanism derived from such techniques are dependent

on the size of the cargo as well as the stiffness of the link. By proposing a new modeling

scheme, we devise a unified description of such experiments for various motor proteins. This

description is based on the picture that the motor thermal fluctuations are rectified in a pre-

ferred direction by means of barriers whose heights are modulated by the chemical reactions

taking place at the motor.

iii

Page 4: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Acknowledgments

I would like to thank my advisor Yoshitsugu Oono for his constant encouragement, guidance,

and support. I would also like to thank Satwik Rajaram for carefully reading the manuscript

and suggesting improvements, and Bojan Tunguz for being a nice office mate all these years.

This work was supported in part by teaching assistantship from University of Illinois, and

summer research assistance from University of Illinois Campus Research Board.

iv

Page 5: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Table of Contents

Chapter 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.2 Motor Proteins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.2.1 ATP: The Energy Source . . . . . . . . . . . . . . . . . . . . . . . . . 31.2.2 Types of Motor Proteins . . . . . . . . . . . . . . . . . . . . . . . . . 5

1.3 Experimental Studies of Motor Proteins . . . . . . . . . . . . . . . . . . . . 61.3.1 Biochemical Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . 61.3.2 Structural Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71.3.3 Single Molecule Studies . . . . . . . . . . . . . . . . . . . . . . . . . . 9

1.4 Theoretical approaches . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121.5 Our Point of View . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141.6 Road Map . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

Chapter 2 Phenomenology of Motor Protein Kinesin . . . . . . . . . . . . 172.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

2.1.1 Theoretical Approaches . . . . . . . . . . . . . . . . . . . . . . . . . . 172.1.2 General Considerations on Phenomenological Modeling . . . . . . . . 18

2.2 Empirical Facts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202.3 Proposed Kinetic Scheme . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

2.3.1 Kinetic Steps and States . . . . . . . . . . . . . . . . . . . . . . . . . 232.3.2 Modeling Details . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

2.4 Explanation of the Mechanochemical Data . . . . . . . . . . . . . . . . . . . 332.4.1 Force-velocity relation . . . . . . . . . . . . . . . . . . . . . . . . . . 352.4.2 Velocity-ATP Concentration Relation . . . . . . . . . . . . . . . . . . 362.4.3 Force-Randomness Relation . . . . . . . . . . . . . . . . . . . . . . . 372.4.4 Randomness-ATP Concentration Relation . . . . . . . . . . . . . . . 392.4.5 Force-Run Length Relation . . . . . . . . . . . . . . . . . . . . . . . 402.4.6 Run Length-ATP Concentration Relation . . . . . . . . . . . . . . . . 412.4.7 Velocity and Run Length Variation with ADP . . . . . . . . . . . . . 42

2.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 432.5.1 Uniqueness of Kinetic Scheme . . . . . . . . . . . . . . . . . . . . . . 442.5.2 Motor Energetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 452.5.3 Is There a Power Stroke in Kinesin Force Production? . . . . . . . . . 452.5.4 How Unique is the model? . . . . . . . . . . . . . . . . . . . . . . . . 47

v

Page 6: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Chapter 3 Effects of the Elastic Motor-Cargo Link on Motor Transport 493.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 493.2 Motor-Cargo System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 513.3 Variation of Motor Velocity with Link Stiffness . . . . . . . . . . . . . . . . . 533.4 Stokes Efficiency of Motor-Cargo System . . . . . . . . . . . . . . . . . . . . 593.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

Chapter 4 Interpretation of Single Molecule Experiments of Motor Proteins 634.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 634.2 Motor-Cargo System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

4.2.1 An Effective Cargo Only Representation of a Coupled Motor-CargoSystem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

4.2.2 Mathematical Formulation of Effective Cargo Only Representation . . 684.3 Explanation of Mechanochemical Data . . . . . . . . . . . . . . . . . . . . . 71

4.3.1 F1-ATPase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 724.3.2 Kinesin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 734.3.3 Myosin V . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

4.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

Chapter 5 Affinity Switch Model for Rotary Motor F1-ATPase . . . . . . 795.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 795.2 Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 815.3 Algorithmic Description of the Model . . . . . . . . . . . . . . . . . . . . . . 845.4 Comparison With Analytically Solvable Model . . . . . . . . . . . . . . . . . 865.5 Comparison of Model Predictions with Empirical Results . . . . . . . . . . . 875.6 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

Chapter 6 Summary and Open Questions . . . . . . . . . . . . . . . . . . 93

Appendix A Solution of the Kinetic Model . . . . . . . . . . . . . . . . . . 96

Appendix B Stochastic Energetics . . . . . . . . . . . . . . . . . . . . . . . 102

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

Author’s Biography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

vi

Page 7: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Chapter 1

Introduction

1.1 Motivation

Ever since the application of X-ray crystallography to the determination of the structure of

DNA, incorporation of physical techniques into biological research have produced a steady

stream of advances in the field of cellular and molecular biology. This laid the foundation for

one of the most exciting insights into the machinery of life, the structure-function relation

which postulated that the function of nucleic acids and proteins are coded in their three

dimensional structure. Attempts to determine the static structure of proteins (the protein

folding problem) and understanding the structural contribution to function dominated the

attention of physicists initially contributing to biological understanding. At the same time,

physical techniques and theoretical tools which were developed to solve complex problems in

a quantitative way were also incorporated into the available tools of biological research. This

produced a new realm of quantitative understanding which was a break from the qualitative

descriptions which characterized biological research.

Development of these new experimental techniques and their application to the study

of function of proteins have generated a wealth of quantitative data. With this, a new

consensus has emerged that to have a complete understanding of a biological phenomenon,

theoretical and experimental collaboration between biologists and physicists is needed. Ac-

cumulated experimental data allow proposals of qualitative hypotheses. These proposals are

substantiated by using quantitative models based on the understanding of underlying physics

and quantitative agreement of such models with available data. Such models are useful in

1

Page 8: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

extracting further information from experimental data, making predictions regarding the

unknowns and thus directing further experimental studies.

While proposing such models, we must pay due attention to certain peculiarities of

biological systems, nature of required understanding, and the way physical approaches should

be modified while approaching a biological problem. Understanding of a biological system is

complete only if its function is explained in terms of underlying physics and the interaction

of the system with its environment. Even the simplest constituents of biological systems

such as individual protein molecules are made up of many atoms. This makes application of

direct reductionist methods of physics which seeks to explain the a phenomena in terms of

known interactions of underlying atomic constituents less successful. It is difficult to bridge

the timescale of atomic motions occurring in picoseconds or less to the biologically relevant

dynamic processes of the system as a whole which occur in microseconds or slower. Another

traditional physics approach of explaining relevant phenomena in terms of coarse-grained

statistical descriptions also needs modification when applied to biological systems. Since

the function is coded in relevant substructures of the system, coarse-graining should retain

these structural domains which may be unique to individual systems. In addition, biological

systems are in nonequilibrium. Though the statistical descriptions of equilibrium systems

are well developed, the nonequilibrium extensions are not.

Motor proteins such as myosins and kinesins which are used for directional transport

in cells constitute one such biological system in which development and application of new

experimental techniques have generated a wealth of information at various time and length

scales [1]. At a fundamental physical level, these are energy transducers which convert

available chemical energy in cells to mechanical energy which is used for directional motion.

The mesoscopic size of the system forces them to do this energy conversion while being

influenced by interactions with the thermal environment. The physically relevant quantities

are the speed at which these motors move, the mechanical force generated, the role of

structural elements in the energy conversion, the efficiency of chemical to mechanical energy

2

Page 9: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

conversion, etc. An important question is how the system uses or minimizes the effects of

thermal fluctuations in various processes of molecular machines. It is also interesting to

know whether the function of different motors can be explained by a unified mechanism at

a coarse-grained level. In this thesis, by incorporating relevant experimental information on

the chemical processes, biochemical and structural information and the dynamics of relevant

functional subdomains, we propose models which attempt to explain the working of motor

proteins in a quantitative way. While proposing these models, attempts are made to minimize

the number of fitting parameters by utilizing available biochemical and structural data on

motors.

1.2 Motor Proteins

Proteins are a class of macromolecules. Individually or in association with other macro-

molecules they perform a variety of biological functions needed to sustain life. Many proteins

act as enzymes catalyzing chemical reactions. Others have structural and mechanical roles

in the cellular machinery. Proteins are made of amino acids. Depending on the sequence of

amino acids constituting them they can have unique three dimensional structures in cellular

environment.

Of the many required cellular functions, some involve directional movement and transport

of macromolecules and supramolecular complexes. Without specialized motors, efficient

transport of large complexes over eucaryotic cells is impossible. Motor proteins are such

molecular motors which produce directional movement in cells and mechanical output by

converting chemical energy into mechanical work [1].

1.2.1 ATP: The Energy Source

Motor proteins produce work by converting the free energy liberated by the hydrolysis of

adenosine triphosphate (ATP) to mechanical energy. ATP is an organic compound com-

3

Page 10: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

posed of adenine, the sugar ribose, and three phosphate groups. The breaking of the bond

connecting the last phosphate ion to the rest of the molecule by the following hydrolysis

reaction creates an adenosine diphosphate (ADP) and inorganic phosphate Pi.

ATP + H2O ↔ ADP + Pi, (1.1)

The breaking of the phosphate bond has a high negative free energy of reaction. The

total free energy released depends on the standard free energy, ∆G0, and the concentrations

of ATP, ADP and Pi. It is given as

∆G = ∆G0 + kBT log[ADP][Pi]

[ATP], (1.2)

where ∆G0 = −54 × 10−21J/mol is the standard free energy change at pH 7, and [ATP],

[ADP] and [Pi] are the concentrations of ATP, ADP and Pi, respectively, in molarity, kB is

the Boltzmann constant and T is the absolute temperature.

Since the typical length scales involved in motor proteins are nanometers (nm), and the

forces generated are of the order of piconewtons (pN), it is convenient to use these units

instead of SI units. In these units, the standard free energy change ∆G0 = −54 pNnm,

kBT = 4.1 pNnm at room temperature. Under physiological conditions, [ATP]∼ 10−3 M,

[ADP]∼ 10−4 M, and [Pi]∼ 10−3 M, ∆G∼ −90 pNnm. ∆G is negative under all practical

conditions, implying that the reaction always proceed in the direction of hydrolysis, and

that spontaneous net synthesis of ATP never happens in solution. However in the absence

of any enzymes the reaction rate to the right for the above reaction is very low and it is this

stability that makes the phosphate bond such an ideal high-energy source.

4

Page 11: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

1.2.2 Types of Motor Proteins

Diversity of cellular functions are realized through the evolution of many different classes of

motor proteins. Latest estimate suggests there are 100 different motor proteins in eucaryotic

cells performing various tasks. Different motor proteins can be classified according to their

specific functions [2] .

Kinesins are a family of motor proteins involved in cargo transport along microtubule

tracks [3]. Microtubules are formed by the polymerization of αβ tubulin heterodimers. They

have a hollow cylinder shape with diameter of 24 nm with tubulin dimers oriented parallel

to the cylinder axis. In addition, they have a polarity which allows fixing the direction of

motor motion with respect to the track as moving towards the plus end or the minus end of

microtubule.

Of the many members of kinesin family, some of them such as kinesin-1 form dimers and

move processively (motor takes many steps before detaching from the track) toward the plus

end of microtubule. Ncd is another kinesin which dimerizes and moves toward the minus end

of microtubule nonprocessively. There are monomers such as Unc104, KIF1A which are also

processive motors. In addition to transporting cargo such as vesicles, they are also involved

in cell division and microtubule assembly. Defective functioning of kinesins are implicated

in neurological disorders, cancer and various other diseases.

Myosins are a family of motor proteins involved in muscle contraction as well as cargo

transport along actin filaments [4]. An actin filament is a left-handed helix of actin monomers

with a periodicity of 36 nm. Similar to microtubule they also have a polarity. The thick

filaments in a muscle are composed of myosin II, and by sliding with the thin filaments of

actin, they produce the muscle contraction force. Myosin V dimerizes and moves processively

toward the plus end of actin whereas myosin VI dimers move towards minus end. Both are

involved in cellular transport.

Dyneins are another class of motor proteins [5]. One type of dynein powers cilia and

5

Page 12: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

flagella (structures with a core made of microtubules which oscillate at high frequency by

virtue of force generated by dynein molecules arrayed along microtubule). Another form of

dynein, cytoplasmic dynein, is involved in the transport of organelles along microtubules.

Polymerases such as DNA and RNA polymerases, move along the strands of DNA in

order to replicate them and to transcribe them into RNA, respectively [6].

In addition to the linear motors above there is a class of rotary motors known as ATPases

which are involved in transport of ions across membranes. A prominent member of rotary

motors is F1-ATPase [7] which is involved in the synthesis of ATP from ADP and Pi using the

proton flow across the membrane as the energy source. Similarly, flagellar motors involved in

the directed motion of bacteria also use ion gradient across membrane as the energy source.

Another class of motors are involved in packaging of viral DNA into protein capsids [8].

In this thesis we will be considering kinesin, myosin V, and F1-ATPase.

1.3 Experimental Studies of Motor Proteins

Experimental characterization of motor proteins can be categorized into the general cate-

gories of biochemical, structural, mechanical and optical studies.

1.3.1 Biochemical Studies

Biochemical studies assume that, while converting chemical energy of ATP to mechanical

work, motors cycle through a series of long lived states with different positions along the

track. After completion of a cycle, the motor advances by one step. These states depend on

the state of the nucleotide (ATP or its hydrolysis intermediates) bound to it and whether the

motor is attached to the track. The state transitions are represented by chemical transitions

of the form,

X0

k1⇀↽

k−1

X1 ⇀↽ · · ·Xi−1

ki⇀↽

k−i

Xi ⇀↽ · · ·Xn, (1.3)

6

Page 13: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

where Xi is an intermediate state of the motor and ki are the rates at which one state

transforms to another. In addition to identifying these long lived states, biochemical studies

also seek to determine the rate constants of state transition. A summary of application of

biochemical methods to the study of motor proteins is given in [9].

1.3.2 Structural Studies

Successful crystallization of protein molecules allows the determination of three dimensional

structures using X-ray crystallography. Structural characterization of nucleotide and track

binding regions and the regions connecting these two have contributed greatly to the under-

standing of motor function. The goal of such studies is to identify the possible structural

transformations that the motor undergoes while hydrolyzing ATP, and generating force and

movement. For this, it is sometimes possible to crystallize the intermediates with different

nucleotide bound states and determine their structures [10]. Difficulties with crystallization

mean there is not enough structural data to give a complete picture of motor conformational

changes. Sometimes low resolution cryo-EM images [11] are used to fill the gaps in structural

knowledge.

Kinesin: Figure 1.1 shows the crystal structure of kinesin dimer with both heads bound

to ADP [12]. The structure comprises two motor heads connected through a coiled-coil

domain. The head is attached to this coiled coil through a structurally important region

called the neck linker (β9 and β10 in the figure). Structures of kinesin bound to microtubule

are not yet determined and the docking of known structure to low resolution cryo-EM images

are used to infer the structure of microtubule bound states and possible structural changes

associated with microtubule binding.

Myosin: Structural studies of myosin are more advanced than kinesin since more hydrol-

ysis intermediate structures available. Crystal structures of myosin II with no nucleotide or

bound to ADP or ADP-vanadate have been reported. Structures of myosin bound to actin

have not yet been reported. Here also docking of known structures to low resolution cryo-

7

Page 14: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Figure 1.1: Crystal structure of kinesin dimer [12].

EM images are used to infer the possible changes associated with actin binding. Fig. 1.2

shows the known crystal structures of myosin II and the structural changes associated with

nucleotide changes [13]. One of the most important insights gained from crystal structures

of myosin is that there is a lever like region (grey helix in Fig. 1.2) which rotates in response

to the changes in nucleotide states.

F1-ATPase: Another motor whose crystal structure determination has provided further

insights into its force generation is the F1-ATPase motor. As shown in Fig. 1.3, the motor

structure consists of α3β3γδǫ subunits [14]. The central γ subunit can rotate inside a cylinder

made of three α and three β subunits arranged alternately. ATP binding sites are located

primarily on a β subunit. Crystal structures show that, depending on the state of ATP

present at the pocket, β subunit can be in either a closed state or open state.

In addition to knowing the structures of individual motors, comparative studies of struc-

8

Page 15: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Figure 1.2: Known crystal structures of myosin II and a proposed schematic representationof possible structural changes during the ATP hydrolysis [13].

tures of different motors are also useful. It is seen that the core of myosin and kinesin

structures have similar structure. This structural similarity suggests nucleotide binding, hy-

drolysis, and release of hydrolysis products may trigger similar motions in motor proteins.

Thus it is possible that both motors have similar mechanisms. Such structural similarities

are encouraging. It is possible that understanding the function of one motor will help in

identifying the functional mechanism of other structurally similar motors.

1.3.3 Single Molecule Studies

The advent of in vitro motility assays [15] allowed the study of motion produced under

biochemically defined conditions using one or a few proteins. Such single molecule studies of

motor proteins can be classified into two general categories. Mechanical studies monitors the

motion of the motor against an external force applied to it. Optical studies with fluorescence

9

Page 16: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Figure 1.3: Schematic representation of F0F1-ATPase structure. Crystal structure is knownonly for F1 part (the top part of the figure) [14].

tags use fluorescent probes attached to the motor to determine distances involved in the

motion of the motor as a whole or structural changes within the motor.

Mechanical studies: Motor proteins typically produce piconewton forces. To observe

the motion of motors against a load, such piconewton forces can be applied to it using an

optical trap (a focused laser beam) or by attaching the motor to a very fine glass needle. In

a typical optical trap experiment shown in Fig. 1.4, the motor is attached to a large bead

which is trapped in the laser beam. The motion of the bead is tracked at high resolution

(to have accurate detection of bead position, beads which are many times motor size are

used in experiments) [16]. To move the bead away from the trap center, motor has to exert

a force on it to overcome the force exerted on the bead by the trap. Optical trap studies

allow the determination of the variation of motor velocity with external load acting on it,

the maximum work that the motor can produce, and the step size of motor displacements.

10

Page 17: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Another variation of this set up uses a large cargo attached to the motor and the motion of

this cargo is observed. In this case, the motor work is measured as the work done against

the viscous drag produced by the cargo.

Application of optical trapping experiments on kinesins have shown that they move on

microtubule by taking 8 nm steps, and generates a maximum force of 8 pN [16]. For myosin

V, the step size is determined as 36 nm and the maximum force produced is 2 pN [17].

Figure 1.4: Optical trap setup used to study the load dependence of kinesin motion [16]

Optical studies with fluorescence tags: Fluorescence resonance energy transfer (FRET)

studies [18] are useful in characterizing the structural changes involved in motor operation.

Here a donor and an absorber fluorophore are attached at two different locations on the pro-

teins. If there is a structural change which modifies the distance between these fluorophores,

the changes in emission spectra of the donor can be used to determine the distance change

involved. Recent advances have allowed tracking of a single fluorophore attached to the

11

Page 18: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

motor with very high resolution. For example, fluorescence imaging at nanometer accuracy

(FIONA) [19] is able to track the location of a fluorophore to nanometer resolution. Appli-

cation of FRET studies to kinesin shows that the neck linker undergoes a conformational

change. Application of FIONA to kinesin and myosin V has shown that the dimers move in

a hand-over-hand fashion in which the two dimers alternate their relative position.

1.4 Theoretical approaches

The goal of any theoretical approach is to understand the mechanism by which motors

convert chemical energy from ATP binding, hydrolysis and product release into mechanical

work. The pioneering work was done by A. F. Huxley [20] to explain muscle contraction. A

cross-bridge model involving relative sliding of actin and myosin filaments was proposed as

causing muscle contraction. In the model, myosin molecules in the filament moves back and

forth about an equilibrium position as a result of thermal fluctuations. On reaching near a

binding site on actin, the strained myosin binds to it. The relaxation of the strain causes

the sliding of the actin filament. On completion of actin sliding, myosin detaches from it

and the cycle repeats.

The determination of crystal structures and the identification of possible structural

changes occurring during ATP hydrolysis has given rise to newer approaches to modeling

motor operation. At the microscopic level, molecular dynamics simulations seek to explain

motor operation in terms of the atomic structures and forces involved. Atomic resolution

structures are used for detailed calculations of molecular energetics and dynamics. The

ultimate aim is to understand how the conformational changes that engender the motion

are produced by ATP and its hydrolysis. The presently available computing power is not

yet sufficient to bridge the time scale of atomic motions to functionally relevant timescales.

However, there is a simulation method which seeks to accelerate microscopic dynamics by

the application of external forces [21]. This approach is ideally suited to the study of chem-

12

Page 19: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

ical reaction paths. It has clarified the details of the energetics of ATP hydrolysis at the

nucleotide pockets of known motor crystal structures [22]. Still, it is difficult to study the

actual conformational changes during the motion cycle of motors.

The time scale limitations of full molecular dynamics simulations have given rise to

intermediate mesoscopic models which use some sort of abstraction of structural details as

the starting point. In a class of models known as power stroke models, the protein structure

is abstracted as a machine made up of a spring-like or elastic element to produce force, a

lever to amplify the force, and a latch to regulate nucleotide binding or release. Fig. 1.5

shows such a representation of the operation of kinesin motors [23].

Figure 1.5: Kinesin motor as a molecular machine [23].

These power stroke models postulate the storage of energy of ATP hydrolysis in the

spring. The relaxation of this stored energy through the motion of the lever is postulated as

the mechanism driving the motor forward. β closing models in F1-ATPase [24], neck linker

docking models of kinesin [25], and lever arm models of myosin [10] are examples of this

modeling approach.

Another class of models identify a few long-lived states in the mechanochemical cycle

13

Page 20: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

of the motor and assume stochastic transitions with mean kinetic rates between them [26].

These models assume the whole motor as a particle that can go through a certain number

of states with different positions along the track. After completion of the cycle the motor

advances by one step. All the stochastic transitions between states are assumed to be

reversible. In presence of an external load acting on the motor, the kinetic rates are modified

as dependent exponentially on the load. The load dependence and the rate constants are

determined by curve fitting to the experimentally determined mechanochemical data on the

motor such as its load dependent velocity.

1.5 Our Point of View

As summarized above, the presently available models capable of explaining the mechanochem-

ical data, the power stroke models, imagine motors as miniature versions of macroscopic

engines working by means of levers, springs, etc. In this picture, ATP binding, hydrolysis,

or product release induces conformational changes in the protein that under load create

strain. The strain drives movement of any load attached to the motor, and this movement

which is referred as working stroke of the motor, relieves strain. For such a model, the

free energy difference between the pre- and post-work stroke states must be comparable to

motor work output. Though this is possible, there are no clear experimental confirmations

of this assertion. For the putative work stroke of kinesin, the conformational change of neck

linker, experiments show that the associated free energy change of nearly 8 pNnm [27] is

much less than the motor output which is greater than 60 pNnm. For myosin, the work

stroke is associated with lever arm rotation, but as of now there are no clear experimental

observations which show that there is a significant free energy change associated with this.

Similar is the case for the assigned work stroke of F1-ATPase, the closing of the β structure.

On the other hand, it is well recognized that physics of small particles in solution are

strongly affected by viscous drag and thermal noise which dominate the inertial forces which

14

Page 21: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

determine system behavior of macroscopic systems. In ratchet systems (for a review, see

[28]), the thermal fluctuations are rectified in a preferred direction by maintaining the system

at nonequilibrium provided there is an asymmetry built in the system. For track binding

motors, it is assumed that the interaction between motor and track produces an asym-

metric potential. Application of such simple ratchets to the quantitative explanation of

mechanochemical data of motor proteins have been less successful till now. Velocity and en-

ergy conversion efficiency obtainable from simple ratchet models are much lower than that

of motor proteins. Nor are they able to explain the crucial role of the structural changes in

motor force generation.

In this thesis we will show that models without any explicit motor driving can explain

the mechanochemical data on motor proteins quantitatively with a suitable chemically tight-

bound ratchet model. The ingredients needed in the models are, (i) incorporation of bio-

chemical experimental data on motors, (ii) role of an elastic link that connects the motor to

load, (iii) reinterpretation of the role of conformational changes as modulating the attach-

ment and detachment of motor to the track in a preferred direction.

1.6 Road Map

In Chapter 2, a phenomenological scheme to incorporate the experimental data on motors

into modeling is proposed. The resulting model is used to explain the mechanochemical data

on kinesin motor. It is found that mechanochemical data alone cannot be used to determine

the unique motor force production mechanism. Indirect experimental observation of using a

large cargo elastically attached to the motor is identified as the possible explanation for the

inadequacy of mechanochemical data to identify the motor mechanism. In Chapter 3 the role

of the elasticity of the link connecting motor to cargo is studied using simple motor models.

It is shown that motor velocity and energetics are dependent on the link flexibility. Chapter

4 proposes a unified scheme incorporating motor biochemical data and motor-cargo link.

15

Page 22: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

The resulting scheme is applied to explaining the mechanochemical data on motor proteins

kinesin, myosin V, and F1-ATPase. In Chapter 5 incorporating the effects of motor-cargo

link and its flexibility, a simple model (a null model) is proposed for the F1-ATPase motor

which can explain the available mechanochemical data. Chapter 6 summarizes the results

of this thesis.

16

Page 23: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Chapter 2

Phenomenology of Motor ProteinKinesin

2.1 Introduction

As introduced in the previous chapter, active transport of molecules and molecular com-

plexes are needed inside cells because of their sheer size. Motor proteins serve as the engines

for this intracellular transport. Kinesins are one such family of motor proteins that move

uni-directionally along the microtubule while hydrolyzing ATP. An interdisciplinary effort

involving different experimental fields and theoretical approaches are needed to have an un-

derstanding of the most interesting aspect of molecular motors: its mechanism of converting

chemical energy gained from the fuel molecules such as ATP into mechanical energy.

Experimental characterization of these motors can be divided into two general categories.

One set of experiments mainly seek qualitative characterization of the motors. Structural

studies involving crystallization, identifying structural changes during motor motion, bio-

chemical characterization of intermediate motor states and their correlation with ATP hy-

drolysis processes and qualitative descriptions of the way motor moves during its operation

fall in this category. The other set of experiments (mechanochemical experiments) measures

the motor output, its velocity, force production, and the efficiency of energy conversion.

2.1.1 Theoretical Approaches

A detailed understanding will involve describing the motor operation at a microscopic level in

terms of atomic motions. Molecular dynamics (MD) simulations involving motor structure,

17

Page 24: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

interaction between motor and ATP and its hydrolysis intermediate, chemical details of ATP

hydrolysis process, and motor track interaction will give such a detailed description. Such an

approach has not yet been very successful since the time scale of molecular motor motions

is still much longer than the time scale routinely available by the current MD.

At the other end of description are the abstract models. Here, structural and biochemical

details of motor operation are neglected and motors are modeled as point particles. In

ratchet models [28], rectification of thermal fluctuations of motors by asymmetric potentials

produced by motor-track interaction is identified as responsible for motor motion and force

generation. In another class of models [26, 29, 30, 31, 32], a set of abstract chemical states

localized along different track positions are introduced. The motor motion is identified

with stochastic transitions between these states. An example is the modeling by Fisher

and Kolomeisky [26], which tries to explain the mechanochemical data on kinesin. This

presents an abstract kinetic scheme with four intermediate states with exponentially load

dependent rate constants. Although the scheme is able to fit the data with the aid of more

than fifteen fitting parameters, it is difficult to translate such a scheme into biochemically

relevant statements concerning the internal mechanism of the motor operation. Besides, it

is hard to incorporate structural or other biological information into the scheme.

2.1.2 General Considerations on Phenomenological Modeling

As described above, theoretical contributions to the understanding of motor proteins, other

than detailed molecular dynamics simulations, try to propose abstract models. Ability to

fit the output from the quantitative experiments are used as a measure of success of these

models. There are two chief obstacles in making a phenomenological description. One is that

the mechanochemical data on motors alone do not adequately constrain the space of possible

models as explicitly noted by Duke and Leibler [33]. The other difficulty is the number of

parameters: if there were many fitting parameters (say, more than 10), even if we could fit

the kinetic scheme to empirical results, it would be difficult to check whether the obtained

18

Page 25: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

set of parameters is the best choice for a given kinetic scheme (let alone the verification of

the scheme itself). Therefore, we wish to minimize the number of adjustable parameters as

much as possible before trying to explain the empirical mechanochemical data quantitatively.

To this end, based on the intermediate states and substeps of motor stepping inferred from

structural and other experimental data, a possible kinetic scheme can be constructed first

(without referring to the availability of rate constants). Then the biochemical experimental

results can be used to determine the phenomenological rate constants appearing in the

scheme. The descriptive power of such a scheme can be tested by trying to explain the

mechanochemical data quantitatively. In essence, if successful, this will produce a kinetic

scheme which may be regarded as a phenomenological summary (or a minimal model, so to

speak) of biochemical and mechanochemical empirical results.

Therefore, it is desirable to construct good mesoscopic models of molecular motors that

can facilitate understanding of single molecule experiments (e.g., [34, 35, 36]) at the time

scale of microseconds or longer. We feel the first step in this direction is the incorporation of

qualitative information on motors into modeling. The general approach can be summarized

as follows. Identify the relevant qualitative information on motors. Propose a model which

incorporates the identified information, and the additional assumptions regarding the motor

force production, and check the agreement of model results with quantitative measurements

on motors. The methodology of phenomenological modeling of kinesin is developed in this

chapter. While formulating the phenomenological model it is assumed that all the empir-

ical data are really reliable. As we will see later, some structure-related empirical results

are not very reliable. This makes phenomenological models less effective in giving a clear

understanding of motor mechanism.

19

Page 26: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

2.2 Empirical Facts

As already discussed in the introduction, our strategy is to make a kinetic description max-

imally utilizing currently available biochemical and structural information, and then to try

to reproduce the mechanochemical experimental results quantitatively with minimal fitting

parameters. Therefore, we first outline the experimental results relevant to the kinesin dimer

procession along the microtubule (Mt).

(i) Observation of kinesin motion using optical traps while it is transporting a cargo shows

that a kinesin dimer takes rapid 8 nm forward steps [35] spaced with (often long) waiting

periods. The periodicity of microtubule track is also 8 nm. There can be occasional backward

steps whose probability increases with increasing external load on the cargo [36, 37]. For

low loads each 8 nm step takes place with consumption of a single ATP molecule [38, 39].

The kinesin dimer processivity (i.e., the capability to take many consecutive steps on the

track before detaching from it) is due to the coordination of two monomer heads [40, 41].

Direct observation of stepping of each head by fluorescent tags shows that kinesin moves by

a hand-over-hand mechanism in which the two heads switches their relative position on the

track [42]. There are strong evidences indicating that the two consecutive steps are left-right

asymmetric (or without left-right symmetry) [43].

(i) is the basic observation and our kinetic scheme is based on: the coordination of two

heads.

(ii) The 8 nm step can be resolved into fast and slow substeps, each corresponding to a

(cargo) displacement of 4 nm. This was first demonstrated by Higuchi et al. [44] and has

been fully confirmed by Nishiyama et al. [36]:

a. The duration of the faster substeps is about 50 µs and is insensitive to the force (for 3-8

pN).

b. The second substep duration is variable (120 ∼ 300 µs for higher loads).

It is known that each head has four major states depending on the state of nucleotide

20

Page 27: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

ATP or its hydrolysis intermediates that are attached to the kinesin head: E: without any

nucleotide, D: with an ADP, T: with an ATP, and DP: with a hydrolyzed ATP (before

releasing the phosphate ion Pi). We can approximately describe the state of a double-

headed kinesin as (A, B), where state A is the state of the trailing head (closer to the minus

end of Mt) and state B that of the advanced head. Also let us denote by the underline the

binding to Mt. Thus, e.g., (DP, D) implies that the trailing head is with a hydrolyzed ATP

and is attached to Mt, and the advanced head is with an ADP but not attached to Mt. We

use X for a not specified state (this does not mean that any state is allowed).

(iii) The reactivity of kinesin head with nucleotides may be summarized as follows:

a. if a kinesin head is free from Mt, releasing ATP or accepting ATP is not easy. This

exchange becomes much easier when the head is attached to Mt [40].

b. ATP → ADP + Pi (without releasing the phosphate ion) becomes difficult when the head

is attached to Mt; only when kinesin is a dimer is the release of phosphate ion possible [40].

c. Exchange of ADP is not strongly affected by the presence or absence of Mt. Under the

presence of ADP, (E, E) is not observed [45].

d. The T and E states have nearly the same binding strength to Mt [46, 45]. D is significantly

weaker than E or T [45]. DP is intermediate between T and D [46].

(iv) Transition from (X, E) to (X, T) triggers the force production.

a. ATP binds to the motor with a second order rate constant 2±0.8µM−1s−1 [47, 48] and

this is followed by the 4 nm (cargo) displacement, which is completed within 50 µs [36].

b. It is strongly suggested [41] that X here is actually DP. (As yet there is no direct

experimental demonstration, but many researchers assume the existence of such a state as

(DP, T) [25].

(v) Detachment of one of the heads occurs in (DP, X).

a. Hancock and Howard [41] have inferred that for a kinesin monomer the detachment occurs

in the DP state.

b. Hackney [40] suggests that phosphate ion dissociation precedes ADP dissociation. The

21

Page 28: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

measured phosphate ion release rate from the monomeric kinesin is very small compared

with the case of dimers [41, 48]. Therefore, the phosphate release process must be assisted

by the other head.

c. Study of the kinetics of a mutant defective in ATP hydrolysis [49] suggests that the

hydrolysis of ATP in the rear head is required for the strong attachment of the front head

to Mt.

d. The step (forward or backward) is tightly coupled to the consumption of one ATP and

the waiting state for both seems the same. The stalling state is interpreted to be where the

probabilities of forward and backward steps are equal [36].

e. The measurement by Nishiyama et al. [36] and Carter and Cross [37] of the ratio of

forward to rear steps as a function of load shows that the number of backward steps increase

as a function of load.

(vi) a. The ATP binding leads to two sequential isomerizations, the second of which reorients

the neck linker relative to the Mt surface [50].

b. The neck linker docks, pointing in the forward direction, to the kinesin catalytic core in the

T or DP state, but is not docked in the E or D state; this transition has an enthalpy change

of ∆H ∼ −200 pNnm [51]. It is also known that the ATP hydrolysis is not a prerequisite

of neck linker docking [51]. This fact combined with a suggests that neck linker docking

precedes ATP hydrolysis [50].

c. For a related protein KIF1A the neck linker docks in the T state but not in the D state

[52].

d. Under the condition that both heads are attached to Mt, if the neck linker of the trailing

head is docked, it is sterically impossible for the front head to have a docked neck linker

[25, 53].

e. Tomishige and Vale [54, 55] have shown that when both heads are bound to Mt, the two

neck linkers (each with length 4 nm) must be in opposite-directing conformations, pointing

backward in the front head and forward in the rear. When the length of the neck linker is

22

Page 29: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

decreased by two residues, processivity is lost. Inserting a 4 nm polyproline segment between

the neck linker and the coiled coil allows the motor to move processively with 16 nm steps

[54].

All these imply that (DP, T) assumed above must have transient substates (denoted with

1 and 2) with respect to the neck linker conformation.

(vii) Transition from (X, D) to (X, E) is the transition from the singly to doubly bound

conformation; there is a consensus about this process [41, 47, 56].

a. The Kawaguchi-Ishiwata experiments [57] have shown that with a head in the T state

and the other in the D state the motor is singly attached to Mt.

2.3 Proposed Kinetic Scheme

2.3.1 Kinetic Steps and States

Based on the summary in the preceding section, we propose the following kinetic scheme for

the kinesin dimer processivity (Fig. 2.1):

(DP, E)k1[ATP]

⇀↽k−1

(DP, T)k2⇀↽

k−2[P](D, T)

k3⇀↽

k−3(T/DP, D)1k4⇀↽

k−4(T/DP, D)2

k5⇀↽

k−5[ADP](DP, E).

Here, [ATP] denotes the ATP concentration, [ADP] the ADP concentration, and [P] the

phosphate ion concentration. Detailed justification will follow the outline of our scenario.

Our scenario may be outlined as follows. Supporting empirical facts summarized in the

previous section are specifically mentioned with key steps.

(1) After ATP goes into the front empty head attached to Mt, state (DP, T) is formed (cf.

(iva)). However, the neck linker does not dock to the motor core of the front head with ATP

immediately (cf. (vid)).

23

Page 30: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Figure 2.1: Kinetic scheme for kinesin dimer processivity. The nucleotide state of each headis denoted by, E: nucleotide free; T: with ATP; DP: hydrolyzed ATP, D: with ADP. Dockedand undocked portions of neck linker are denoted by thick and thin lines respectively.

(2) Next, release of a phosphate ion from the rear head (cf. (va, b)) and its detachment from

microtubule occur in a concerted way, and (D, T) is reached. The cooperativity needed for

accelerated phosphate release in the dimer (vb) comes from the attempts of the neck linker

of the front head in T state to dock.

(3) This allows the completion of docking of the neck linker of the front head to the core (cf.

(vid)) and the rear head moves a distance of 8 nm forward and is poised to move forward

further (perhaps slightly ahead of the previous front head that is still attached to Mt; the

cryo-electron microscopy studies of Hoenger et al. [58] suggests such a state). This is our

interpretation of the power stroke proposed by Vale et al. (cf. (vi)), and is correspondent to

the fast substep (iia). The still Mt-bound head (the previous front head) is in the T or DP

state with the docked neck linker (cf. (vib)) (likely to be in the DP state according to (iv)).

24

Page 31: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

This is the state (T/DP, D)1.

(4) From this state the unattached head in the D state (the previous rear head) diffuses

forward a distance of another 8 nm to state (T/DP, D)2, which is just before reattaching

of the new front head to the new position on Mt. Due to the steric constraint, the neck (or

load) diffusion is required for the diffusing head to reach the new position on Mt. The still

attached now-rear head is in the T or DP state (with the docked neck linker; very likely to

be in DP [41, 49]). This corresponds to the second substep (iib). Strictly speaking, this is

our interpretation, but is the simplest interpretation of the facts summarized in (ii).

(5) The release of ADP and reattachment to Mt occurs to produce (DP, E), which corre-

sponds to the same state as the starting state but having advanced 8 nm. This is the waiting

state for ATP.

(6) The backward steps at higher loads is taken into account by assuming that once ATP goes

in (step (1)) the attempts of the front head linker to dock produce an internal tension and

either of the heads can detach from microtubule (vd). There is a higher probability for the

rear head to detach (ve). A possible kinetic scheme for backward steps is given in Discussion.

Once the rate constants are known, with the aid of the standard procedure summarized

in [59] the velocity, randomness, and their [ATP] and [ADP] as well as force dependences

can be determined.

The detailed explanation and justification of the states and processes follow.

(DP, E)

Supporting facts for this state to be the waiting state for a step have already been summarized

in (iv) and (va) in the preceding section.

The transition from this state to the next state (DP, T) involves the binding of ATP to

the front head with a second order rate constant k1.

(DP, T)

In this state the neck linker of the rear head is being docked but that of the front head in the

25

Page 32: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

T state has not yet been docked (cf. (vid)). The binding of ATP is assumed to be contingent

to a displacement of some part(s) of the front head. This displacement may correspond to

the one observed in KIF1A due to Kikkawa et al. [52]. Another observation that may have

relevance to this displacement is that the kinesin head has an intrinsic bias towards forward

displacement independent of the neck linker [25]. If the monomer velocity v measured by

Inoue et al. [60] is interpreted as due to a biasing displacement of δ associated with ATP

binding followed by diffusion, then we have v = kδ, where k is the ATP turnover rate. Their

data are approximately compatible with δ =1 nm.

Observations by Higuchi et al. [44] have demonstrated that there is a waiting step after

ATP binds the front head and before the motor starts to produce force. An interpretation

is that the initial binding of ATP is a kind of collision complex, and it requires some dis-

placements of the parts of the head to come into the (DP, T) state which is the state that

allows the subsequent phosphate ion release from the rear head and its detachment from

microtubule (the change to (D, T) ).

This state is likely to be identified with the starting point of the neck linker docking

(see Discussion). From this state (D, T) is formed with a rate constant k2, and a reverse

transition to the initial state (DP, E) that releases the bound ATP [41, 47, 56] with a rate

constant k−1. Another process that can happen especially under high load is the detachment

of the front head in T state and kinesin taking a backward step (vd, e).

(D, T)

This state is obtained by the release of a phosphate ion from the rear DP state (cf. (vb)),

and detachment of the rear head from microtubule into the D state assisted by the front

head (cf. (va, b)).

The driving force for this change is interpreted to come from the initial stage of neck

linker docking to the front head core.

Further neck linker docking happening at the front head causes a forward transition to

(T/DP, D)1 state with a rate constant k3. Reattachment to microtubule and phosphate ion

26

Page 33: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

rebinding produces (DP, T) state.

(T/DP, D)1

This state is obtained through the neck linker docking to the front head. This process

corresponds to the first substep mentioned in (iia) [25].

The hydrolysis of ATP may not have taken place as summarized in (v). The neck linker

of the T/DP state is docked and that of the D state is undocked (cf. (vib)). Gilbert et al.

and Ma and Taylor as well as Rice et al. [47, 56, 51] have shown that ATP hydrolysis is not

a prerequisite for moving one single step.

It is observed that the free energy change associated with neck linker docking is small

(∼ 1-2 kBT) [35].

Note that the enthalpy change for docking is very negative (△H ∼ −200 pNnm (vb)).

There is a forward transition from this state to (T/DP, D)2 by the forward movement of

the detached head (in the D state, the former rear head) in the D state to the next binding

site. Also there is a reverse transition to (D, T) (by the undocking of the neck linker).

(T/DP, D)2

The new front head which is now in the D state is near the next binding site but not yet

attached to it tightly. The rear head is in the T/DP state (likely to be in DP) with the neck

linker docked (cf. (vib)). It is possible that the front head in this state is attached weakly

to the next binding site.

The experimental evidence for such a state comes from all the chemical studies [47, 56]).

Ma and Taylor [56] propose that the D state on Mt can exist in two states, one weakly bound

and the other strongly bound. The observations by Sosa et al. [61] show that the D state is

weakly attached to Mt and is very flexible. Other observations [46, 57] suggest that the D

state is weakly attached to Mt.

There is a forward transition from this state to (DP, E) by the detachment of ADP from

the front head. Also there is a reverse transition to (T/DP, D)1 by the detachment of the

weakly bound front head.

27

Page 34: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Returning to (DP, E)

This state is recovered by the attachment (or weak to strong attachment) of the front D head

of (T/DP, D)2 to Mt and subsequent (or almost concerted) release of ADP. This release is

fast (100-300 s−1, which is sensitive to kinesin species) and completes the transition from

a weakly bound to strongly bound head [46, 57]. In this state the neck linker of the rear

head is docked and points forward, while the neck linker of the front head that is not docked

points rearward (cf. (vie)).

This is the waiting state for the next 8 nm movement, but can get back to state (T/DP,

D)2 (very likely to be (DP, D)2, that is, the nucleotide is hydrolyzed) by the attachment of

ADP to the nucleotide-free head E. Ma and Taylor [56] propose that this happens with a

second order rate constant ∼ 1.5-4 µM−1s−1. The observations of Yajima et al. [62] showing

that the velocity and processivity of the motor decreases with increased concentration of

ADP also points to the presence of this reversibility.

2.3.2 Modeling Details

The choice of the rate constants (see Table 2.1 for a summary) and the general mathematical

expression of the proposed kinetic scheme are summarized here with supporting arguments.

(DP, E) to (DP, T)

In the presence of an external force F (> 0 implies the force against procession), the second

order rate for ATP binding is assumed to be k1(F ) = k1[ATP] exp(−Fδ/kBT ), where kB

is the Boltzmann constant, T the absolute temperature (chosen to be a room temperature

300K), and [ATP] the ATP concentration. Here, δ describes the initial displacement associ-

ated with T binding.

For the reverse transition from state (DP, T) to state (DP, E) the rate constant is k−1.

A possible unified picture of the ATP binding and the nature of this substep are given in

Discussion.

There is a general consensus on the value of the second order rate constant for ATP bind-

28

Page 35: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

ing, k1 [47, 56] in the absence of external force: k1 = 2±0.8 µM−1s−1 according to Gilbert

et al. [47]. The reverse transition rate in the absence of external force is: k−1 = 71 ± 9 s−1

according to Gilbert et al. [47]. We have chosen k1 = 2 µM−1s−1, k−1 = 71 s−1 and δ = 1.2

nm. As already discussed in the entry of (DP,T), δ ≃ 1 nm is consistent with the data by

Inoue et al. [60].

(DP, T) to (D, T)

For low loads this transition corresponds to the phosphate ion release and detachment of the

rear head from microtubule. We assume both these processes happen in a concerted way

with a net rate constant k2. Ma and Taylor [56] show that the maximum ATPase rate is

100s−1, which is k2k5/(k2 + k5) if k5 is the ADP release rate. The measured value of k5 is in

the range 100-150 s−1. We adopt k5 = 150 s−1, giving k2 = 300 s−1.

At higher loads as mentioned before there is a finite probability for the front head to

detach and a backward step to take place. The ratio of the probability of forward (Pf ) to

backward (Pr) steps may be written as

Pf

Pr

= e(GT−GDP−Fδd)/kBT (2.1)

to be consistent with detailed balance condition with exponential load dependence, where

GT is the free energy change due to the detachment of the head in T state, GDP that in DP

state and δd accounts for the force dependence of head detachment rate from the track. We

adopt GT − GDP ∼ 8kBT that is obtained by the thermodynamic consideration on phos-

phate ion release (details are given in Discussion). The assumption behind this is that the

rear head detachment which happens together with phosphate ion release is assisted at least

in part by the free energy change due to phosphate ion release. The displacement distance

δd ∼4 nm is chosen, because Pf/Pb = 1 when F is the stalling force (∼ 8 pN). Such a value

is compatible with the experimental observation of the force dependence of unbinding in the

29

Page 36: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

T and E state by Uemura et al. [45] who get δd in the range of 3-4 nm.

The reverse transition rate which is proportional to the phosphate ion concentration [P]

is written as k−2[P]. There are no measurements of k−2 available. As explained below in

4.G, a study of phosphate ion dependence of the velocity can be used to estimate k−2 to be

∼ 0.1 µM−1s−1.

(D, T) to (T/DP, D)1

This transition corresponds to the fast substep of Nishiyama et al. [36]. It is shown that this

takes place within 50 µs (as observed by the displacement of a 0.2 µm diameter bead attached

to the neck coiled-coil) and is independent of the external force (for their observation range

3-8 pN). We interpret that the actual length of this step is 3 nm with the experimental

limitations of step resolution taking the 1 nm of ATP binding as part of a 4 nm step. Our

molecular picture of state (T/DP, D)1 is that once the rear head is detached, the neck-linker

docking proceeds immediately and this docking moves the neck coiled coil forward by 3 nm

and this places the previously trailing head slightly ahead of the previously advanced head

(see Fig. 1).

A possible explanation for the insensitivity of this step to the external force is that the

docking of the neck linker produces forces much larger than even the stalling force of the

motor which is in the range of 6-8 pN. The docking process is probably due to zipping of

the neck linker to the motor core. The very high △H ∼ −200 pNnm associated with this

conformational change could be a source of a very large force. Note that in this picture

the time taken will be inversely proportional to the friction constant, implying its direct

proportionality to radius of the bead. (The radius of the head ∼ 3 nm is much less than

the radius of the beads used in mechanochemical experiments which is of the order of 0.1

µm, so that the displacement of the head can be neglected as a rate determining process in

comparison to that of the bead.)

The rate constant k3 is taken as the inverse of the above mean reaction time. For the

30

Page 37: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

explanation of the mechanochemical data of Visscher et al. [16] who use beads of diameter

0.5 µm, the rate constant k3 = (1/50 µs) × 0.2/0.5 = 8, 000s−1 (i.e., scaled according to the

diameter). As mentioned earlier the free energy decrease associated with neck linker docking

is −△G ∼1-2 kBT . Using this and the fact that ratio of the forward and the backward rate

constants satisfies k3/k−3 = exp[−(△G−F×d)/kBT ], with d = 3 nm being the displacement

against the load F , we can estimate the reverse rate constant k−3 for getting back to the

(D, T) state. We choose −△G ∼2 kBT (The results are not very sensitive to this choice as

long as it is in the range 1-2 kBT ).

(T/DP, D)1 to (T/DP, D)2

It is assumed that this step corresponds to the free diffusion of the previous rear head over

a distance of 8 nm, if no external load exists. Reflecting and absorbing boundary conditions

are chosen at the two ends to model the diffusion process. In presence of an external load

applied to the bead attached to the coiled-coil as in mechanochemical experiments, the rate

constant is taken as the inverse of the mean first passage time for this bead to move against

the load by a distance of 4 nm. Here, we assume that the link connecting the motor and

cargo is rigid. For the front head to bind to the forward binding site on Mt, the bead has

to diffuse forward by 4 nm. Once the bead reaches this position, the forward head would be

in a favorable position to attach to the binding site on Mt. With the displacement d =4 nm

we get the mean first passage time as [32],

d2

Db

[

eFd/kBT − (1 + Fd/kBT )]

(Fd/kBT )2, (2.2)

where Db is the diffusion constant of the bead to which results are sensitive at high loads, but

not very at smaller loads. For the explanation of the mechanochemical results of Visscher

et al. [16] we use Db = 8.79 × 10−13m2/s, estimated from the bead diameter of 0.5 µm, and

the viscosity of solution ∼0.01poise. k4(F ) is taken to be the reciprocal of the above mean

passage time.

31

Page 38: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

The front head, which is attached to microtubule after the completion of the diffusion

process discussed above, can detach from Mt, making the reverse transition to (T/DP, D)1.

This process is assumed to have a rate constant, k−4 which is close to the detachment rate

from microtubule. From the results of Hancock and Howard [41], the detachment rate for

the head with ADP from microtubule is 1.01 s−1 ± 0.28 and we take k−4 to be 1.01 s−1.

(T/DP, D)2 to (DP, E)

For this transition, Gilbert et al. [47], who use Drosophila kinesin, give the rate k5 = 300

±100 s−1. Ma and Taylor, who use human kinesin, give the rate 100-150 s−1. Gilbert et al.

attribute this difference to the different types of kinesin used. To reproduce approximately

the measured velocity vs. force data of Visscher et al., who used squid kinesin, we have to

choose k5 ≃ 150 s−1 (See also the section on (DP, T) to (D, T) transition above). The speed

of kinesin procession depends strongly on this parameter, because ADP release is the factor

determining the maximum speed. This process is sensitive to the molecular structure as can

be exemplified by the fast fungal kinesin required for hypha elongation of Neurospora crassa

with a large nucleotide binding pocket [63]. Thus, k5 is a species-dependent parameter.

Therefore, we must treat this as an adjustable parameter. Notice, however, that the selected

value is within the known range and is a very natural one. For the reverse transition the

rate constant k−5 = 1.5-4.5 µM−1 s−1 according to Ma and Taylor [56]. We choose k−5 = 2

µM−1s−1. The model results are not sensitive to this choice of k−5.

Detachment from Mt

Hancock and Howard [41] measured the detachment rates of two-headed kinesin from mi-

crotubule at zero load in different nucleotide states. The measured rates are, 1.01±0.28 s−1,

1.67±0.50 s−1, 0.0009±0.0002 s−1 and 0.0010±0.0004 s−1 for states D, DP (from state D with

excess phosphate ions), E and T (with ATP and AMP-PNP), respectively. Experiments by

Coppin et al. [34] have shown that the detachment rate of kinesin from microtubule increases

with external load. In our scheme, the intermediate states involve states analogous to E (for

(DP, E)), T (for (DP, T) and (D,T)) and DP (for (T/DP, D)1 and (T/DP, D)2) states. We

32

Page 39: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

choose the detachment rates for each of the above states at zero load as, 0.001 s−1, 0.001 s−1,

and 1.7 s−1, respectively. The run length is sensitive to the last kinetic parameter, so just

as in the case of k5, strictly speaking, we treat this as an adjustable parameter. However,

notice that the chosen value 1.7 s−1 is very close to the average value cited above.

To account for the force dependence of the detachment, each of these rates are multi-

plied by an exponential term e0.25F (F in pN) corresponding to an interaction range of 1 nm

between kinesin head and microtubule. The above rate constants and the exponential term

produces good agreement with the experimental data on processivity.

Note on the low-load data of Yajima et al.

Yajima et al. [62] uses kinesins fused to gelosin to determine the low load variation of

kinesin velocity and runlength with ATP, ADP and phosphate ion concentration. For the

explanation of this data the previous rate constants and mechanisms are modified slightly,

owing to the fact that they are using a different kinesin (rat kinesin) and there are no beads

attached to kinesin. The necessary modifications are: changing the ATP attachment rate

k1 to 4.5 µM−1 s−1, and detachment rate from Mt in the DP state to 3.2 s−1. The diffusion

step is taken to be the escape time for the head to travel 8 nm and the fast substep rate is

taken to be > 20,000 s−1. Note that at low loads neither the velocity nor the run length is

sensitive to these rates. The ADP reattachment rate is taken as 4.5 µM−1 s−1. The other

rate constants have the same value used in the previous case.

2.4 Explanation of the Mechanochemical Data

We will show below that the kinetic scheme constructed in the above through distilling the

known biochemical and structural information explains the mechanochemical experimental

results of kinesin dimer processivity quantitatively.

It should be emphasized that except for

(i) the choice of the species-dependent ADP release rate k5 = 150 s−1, and the DP state

33

Page 40: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Process Rate constant Reference

ATP binding (k1) 2 µM−1s−1 [47]ATP unbinding (k−1) 71 s−1 [47]Phosphate ion release (k2) 300 s−1∗ [56]“Neck linker docking” (k3) 20,000 s−1# [36]ADP unbinding (k5) 150 s−1 [56]ADP binding (k−5) 2 µM−1s−1 [56]Detachment from microtubulein state D (k−4) 1.01 s−1 [41]Detachment from Mt in state DP 1.7 s−1 [41]Detachment from Mt in state T 0.001 s−1 [41]Detachment from Mt in state E 0.001 s−1 [41]

* This is not a directly measured rate but one inferred from the measured maximum ATP turnover rate and

the measured ADP detachment rate. See the (DP, T) to (D, T) section below for details.

# This is when a bead of 0.2 µm diameter is attached to the neck coiled-coil.

Table 2.1: The adopted rate constants. Except for k4 and the detachment rate from Mtin the DP state, they are chosen from the literature (representative or average values arechosen). The rate constants k2, k5 and the detachment rate from Mt in the DP state areadjusted but are still within the error bars of the reported data; we say we have adjustedthese values simply because they are not the mean (center) values given in the literature.

detaching rate from Mt (1.7 s−1),

and

(ii) the following four qualitative information:

(a) The ATP binding rate is force-dependent (e−0.3F consistent with the displacement δ = 1.2

nm as discussed before, and F in pN),

(b) The neck linker docking is insensitive to load,

(c) The ratio of forward to reverse steps is of the form Pf/Pr = e8−F (F in pN) that has

also been discussed already,

(d) The detachment rate of kinesin from Mt increases with external load (assumed to be

e0.25F , F in pN),

all the parameters and the structure of the kinetic scheme are determined without fitting

to any of the available mechanochemical experimental results. Even the chosen values to

the adjustable parameters in (i) are within the empirically obtained ranges. The agreement

34

Page 41: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

of our prediction and the available mechanochemical results is comparable to that of the

descriptions as Fisher and Kolomeisky [26] with sheer fitting parameters.

2.4.1 Force-velocity relation

0

200

400

600

800

1000

0 2 4 6 8

Ve

loci

ty (

nm

/s)

Force (pN)

[ATP]=2,000µM

[ATP]=5µM

Figure 2.2: The force-velocity relationship for ATP concentrations of 2,000 µM and 5 µM.The results due to our kinetic scheme and experimental data of Visscher et al., (Visscher etal., 1999) are given. If we define the stalling force as the force that reduces velocity below 5nm/s, it is of the order of 6 pN for [ATP] = 5 µM and 8 pN for [ATP] = 2,000 µM. Positiveforce increases the velocity only for intermediate and low levels of ATP concentrations. Theindependence of velocity for assisting force under high [ATP] is a prediction of our kineticscheme.

The variation of velocity with force at two different [ATP] is shown in Fig. 2.2. For

comparison, the experimental data of Visscher et al. [16] are also given. See the appendix for

an expression of velocity in terms of force dependent rate constants and ATP concentration.

Clearly, the velocity obeys the Michaelis-Menten type formula as a function of [ATP].

35

Page 42: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

The velocity predicted by our kinetic scheme agrees with the experimental data of Viss-

cher et al. [16]; the velocity saturates for low loads and drops to low values at higher loads.

There is another published velocity-force relation by the Vale group [34]. Qualitatively, their

results exhibit this saturation effect and agree with our functional shape for higher [ATP]. At

high ATP and with intermediate to large forces, the diffusion step is the only rate limiting

and force dependent step, and this range of the curve is supposed to be explained by this

diffusion process.

At saturating ATP levels, the effect of forward (i.e., assisting) forces is negligible, whereas

at lower [ATP] the velocity increases with force. This is a feature not yet verified experi-

mentally, although there are certain indications in this regard in the experimental data of

Coppin et al. [34].

From the plots it is obvious that the stalling force (defined as the force for which velocity

falls below a certain threshold) of the motor increases with increasing [ATP]. If we choose

the threshold as 5 nm/s, the stalling force varies from ∼6 pN for [ATP] = 5 µM to ∼8 pN

for 2,000 µM. This is in agreement with the data presented by Visscher et al.

2.4.2 Velocity-ATP Concentration Relation

Figure 2.3 exhibits the variation of velocity with [ATP] for four different load levels. For

loads of 1.05 pN, 3.59 pN and 5.63 pN the predictions due to our kinetic scheme agree with

the experimental data of Visscher et al. [16]. For zero load case the experimental data of

Yajima et al. [62] is used for comparison and the agreement is good. The maximum velocity

is found to decrease as the load increases. The velocity at very low [ATP] is also found

to decrease with increasing force. This is indicative of the fact that the Michaelis-Menten

constant increases with force in agreement with the results of Visscher et al. [16].

A very low ATP concentration data can be found in Hua et al. [38]: for 5 nM of ATP

the rate is 0.09 ± .01 nm/s, and for 400 nM of ATP the rate is 5.3 ± 0.9 nm/s. Our model

gives 0.07 nm/s and 5.1 nm/s for respective concentrations.

36

Page 43: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

1

10

100

1000

1 10 100 1000 10000

Ve

loci

ty (

nm

/s)

[ATP] (µM)

F=5.63 pNF=3.59 pNF=1.05 pN

F=0 pN

Figure 2.3: Velocity-[ATP] relation. Experimental data of Visscher et al., [16] are given forloads of 1.05 pN, 3.59 pN and 5.63 pN. Data for 0 pN are taken from Yajima et al. [62]. Theplot is of the Michaelis-Menten form with the Michaelis-Menten constant increasing withload.

2.4.3 Force-Randomness Relation

Randomness is defined by [16]

r = limt→∞

〈x2(t)〉 − 〈x(t)〉2d〈x(t)〉 =

2D

vd, (2.3)

where x(t) is the load position at time t, 〈 〉 is the sample average, D is the effective diffusion

constant of the motor, v its velocity, and d the step length. It is suggested that randomness

is a measure of the number of rate limiting steps affecting the motion of the motor. Fig.

2.4 shows the variation of randomness with force at a saturating (2,000 µM) [ATP]. It was

found that the there is a significant disagreement with the data of Visscher et al. at large

forces unless backward steps are taken into account.

37

Page 44: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

0

0.5

1

1.5

0 2 4 6 8

Ra

nd

om

ne

ss

Force (pN)

[ATP]=2,000µM

Figure 2.4: Randomness vs force for [ATP] = 2,000 µM. Experimental data of Visscheret al. [16] for 2,000 µM is also given. For 2,000 µM and for low forces the value of r isapproximately 0.5 and the value approaches unity for high forces.

Our kinetic scheme has (i.e., the available biochemical information suggests) two rate

determining step candidates for low loads and saturating [ATP], the ADP release step and

phosphate ion release step. At saturating ATP levels, the ATP binding rate cannot be rate

limiting (4,000 s−1 for [ATP] = 2,000 µM and with low forces) and is much greater than

the ADP release rate (100-150 s−1) and phosphate ion release rate (>300 s−1). The fact

that these two rates are nearly of the same magnitude indicates constant randomness value

near 0.5 at low loads. At larger forces the diffusion step also becomes rate limiting and the

randomness approaches 1. The presence of backward steps at higher loads accounts for the

randomness values greater than 1.

38

Page 45: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

0

0.5

1

1.5

2

1 10 100 1000 10000

Ra

nd

om

ne

ss

[ATP] (µM)

F=5.69 pNF=3.59 pNF=1.05 pN

Figure 2.5: Randomness with [ATP] for loads of 1.05 pN, 3.6 pN and 5.69 pN and theexperimental data taken from Visscher et al. [16]. See text for explanation.

2.4.4 Randomness-ATP Concentration Relation

Figure 2.5 exhibits the variation of randomness with ATP concentration. For intermediate

and high ATP concentrations the data agree with the experimental results of Visscher et al.

[16]. For low forces, at low ATP concentrations, irrespective of the load, the ATP binding

process is the sole rate limiting process and gives a randomness value should be close to

1. As the ATP concentration is increased, the ATP binding, and ADP and phosphate ion

detachment rates becomes rate limiting and the randomness approaches 0.33, and at higher

ATP concentration, ADP and phosphate ion detachment rates becomes rate limiting and

randomness approaches 0.5. The higher value of randomness at larger force is due to the

presence of substantial number of back steps.

39

Page 46: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

2.4.5 Force-Run Length Relation

0

500

1000

1500

2000

0 2 4 6 8

Ru

n le

ng

th (

nm

)

Force (pN)

[ATP]=2,000µM[ATP]=5µM

Figure 2.6: Variation of run length with force for [ATP] = 2,000 µM and 5 µM and experi-mental data taken from Schnitzer et al. [35]. The result of our kinetic scheme for these two[ATP] values are almost indistinguishable.

Figure 2.6 exhibits the variation of run length with load for two different [ATP]. For

comparison, the experimental data of Schnitzer et al. [35] are also given. The predictions

due to our kinetic scheme agree with the experimental data in the intermediate and large

force ranges. There is some discrepancy at very low loads, but is likely due to the large

uncertainty in the experimental results (See the next paragraph).

The predicted run length does not strongly depend on the ATP concentration. At low

[ATP] the (DP, E) state dominates the mechanochemical cycle whose detachment rate is

very small. This smallness compensates for the low velocity associated with low [ATP]. The

above conclusion is in agreement with the recent experiments of Yajima et al. [62] which

show that for low forces the run length is independent of [ATP], and disagrees with [35].

40

Page 47: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

2.4.6 Run Length-ATP Concentration Relation

0

200

400

600

800

1000

1200

1400

1 10 100 1000 10000

Ru

n le

ng

th (

nm

)

[ATP] (µM)

F=5.6 pNF=3.6 pNF=1.1 pN

F=0 pN

Figure 2.7: Run length vs [ATP] for loads of 0 pN, 1.1 pN, 3.6 pN and 5.6 pN and experi-mental data taken from Yajima et al. [62] for 0 pN and Schnitzer et al. [35] for loads of 1.1pN, 3.6 pN and 5.6 pN. The run length is found to decrease not significantly for low loadseven for very small levels of ATP concentration.

Figure 2.7 exhibits the variation of run length with [ATP] for four different loads. For

comparison, experimental data of Schnitzer et al. [35] for loads of 1.1 pN, 3.6 pN and 5.6

pN and Yajima et al. [62] for 0 load are also given(the fact that 1.1 pN data of Schnitzer et

al. has greater randomness value than the 0 pN data of Yajima et al. might be due to the

different kinesin used). It is found that the run length saturates with higher levels of [ATP]

and the saturated value decreases with increasing force. Although the general shape of this

curve is in agreement with Schnitzer et al. [35], there is one notable difference. Our kinetic

scheme shows that at low loads even at very low [ATP] the processivity does not decrease

significantly. This is in agreement with more recent results of Yajima et al. [62]. In our

41

Page 48: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

kinetic scheme this happens because at low [ATP] the (DP, E) state is dominant from which

detachment rate is small. Since the velocity also is low in this case, the net processivity

(velocity/detachment rate) is not significantly different from the higher ATP level cases.

2.4.7 Velocity and Run Length Variation with ADP

0

200

400

600

800

1000

0 500 1000 1500 20000

250

500

Ve

loci

ty (

nm

/s)

Ru

n le

ng

th (

nm

)[ADP](µM)

Run length

Velocity

Figure 2.8: Velocity and run length vs [ADP] in the presence of 2,000 µM ATP with 0 pNforce, and experimental data taken from Yajima et al. [62]. Both the velocity and the runlength are found to decrease with added ADP.

Figure 2.8 exhibits the variation of velocity and run length of the motor with increasing

[ADP] in presence of saturating ATP levels and low force. The prediction of our kinetic

scheme is in agreement with Yajima et al. [62] as seen in the figure. Both the velocity and

the processivity are found to decrease with the increasing ADP level. This result can be

interpreted in the following way. The binding of ADP to kinesin in state (DP, E) slows down

the net ADP detachment rate from state (T/DP, D)2. As noted earlier the velocity is very

42

Page 49: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

sensitive to this rate. For processivity a similar mechanism is also in operation. The binding

of ADP increases steady state probability of state (T/DP, D)2 from which detachment rate

is high compared to the low rate detaching from state (DP, E), resulting in the decrease in

processivity.

Yajima et al. also studied the variation of velocity and run length for two different

phosphate ion concentrations. If the phosphate ion concentration is increased, the velocity

is found to decrease slightly with no change in run length. Fitting the velocity data by

keeping all the rate constants fixed except k−2 gives the phosphate ion rebinding rate k−2 ∼

0.1 µM−1s−1. This rebinding slows down the net forward transition from (DP, T) state to

(D, T) state and explains the velocity decrease with increasing phosphate ion concentration.

The fact that the run length is not affected is explained by the fact that the rebinding of

phosphate ion does not alter the microtubule binding strength. (It takes kinesin from one

strongly bound state (D, T) to another strongly bound state (DP, T)).

2.5 Discussion

It has turned out that we can use the biochemical data for the ATP hydrolysis process and the

overall biochemical scheme is consistent with the one given by biochemical experiments [47].

Our scheme also incorporates structural and other physical information of motor operation

and is capable of explaining mechanochemical data, while at the same time being compatible

with biochemical data.

We have tried to distill a kinetic scheme from the available biochemical and structural

information, and have given a detailed account of the scenario we believe most likely. Our

basic kinetic scheme follows the general scheme for motors summarized by Bustamante et

al. [64]. We have assumed (as the most natural scheme compatible with almost all the

available biochemical and structural information) strict coordination of the two heads and

tight coupling of motor motion to ATP. The obtained scheme explains the mechanochemical

43

Page 50: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

data of kinesin quantitatively. That is, the available biochemical information augmented

with certain details on neck linker docking has turned out to be sufficient to explain the

mechanochemical data quantitatively. The scheme also demonstrates the compatibility of

the experimental information about the parts of the motor, say, the neck linker described in

Rice et al. [51], with overall functioning of the motor.

2.5.1 Uniqueness of Kinetic Scheme

As mentioned before, the waiting step for forward and backward steps are the same, that

is the ATP binding. On ATP binding and the associated displacement of the front head, if

the rear head is detached, a forward step takes place. On the other hand if the front head

in ATP state is detached, a backward step is produced. An approximate scheme for such a

step is

( , DP, E) ⇀↽ ( , DP, T) ⇀↽ ( , DP, T) ⇀↽ (DP, DP/D, )

⇀↽ (DP, D/E, ) ⇀↽ (DP, E, ),

where the empty sites have been explicitly denoted.

Since there are 8 states T, T, DP, DP, D, D, E, E for each head, a possibility is that

we must consider the transition table among 64 states at least for a single kinesin dimer as

mechanically (or blindly) done in [65]. However, the sequence T → DP → D → E never

changes (reversing this reaction sequence is generally very hard), and with or without Mt

only a couple of transitions occur without difficulty. They are T → DP, DP → D, E → D

without Mt (however, note that T is hard to realize and so is DP). With Mt T → DP (vc),

D ↔ E (iiic), and E → T (iva) can occur. Furthermore, ease of detachment (iiid) tells us

that essentially only D → D is easy. Therefore, no parallel kinetic path other than the one

we proposed can contribute significantly.

44

Page 51: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

2.5.2 Motor Energetics

The energetics of the kinesin dimer may be inferred as follows:

If we assume cellular conditions with [ATP] = 4,000 µM, [ADP] = 20 µM and the phosphate

ion concentration 2,000 µM, the free energy supplied by the ATP hydrolysis is about 25 kBT

(= 100 pNnm). Using the chemical kinetic constants for ATP hydrolysis process [47] and

the microtubule binding equilibrium constants [46], free energy changes involved in each of

the kinetic steps can be estimated. The energy transduction can be explained as follows.

(1) Major sources of free energy available are from binding of ATP (∼ 5 kBT ), phosphate

ion release (∼16 kBT ) and strong binding to microtubule (∼8 kBT ). (2) The ATP binding

energy is transfered to the load by means of partial neck linker docking. (3) Nearly half of

phosphate ion release energy is used to detach the rear head from microtubule and the other

half is dissipated. (4) Diffusing front head is captured by microtubule and the interaction is

stabilized by its strong binding. In this energetic picture the maximum available energy for

work is nearly half of ATP hydrolysis energy and this can explain the efficiency measured

in stall experiments which is near 50%. Since the detachment of the rear head is coupled

to phosphate release with an available free energy of ∼ 8 kBT , this justifies the assumption

that GT − GDP ∼ 8kBT .

2.5.3 Is There a Power Stroke in Kinesin Force Production?

For a motor displacement to be characterized as a power stroke, two conditions have to be

satisfied. There should be a motor conformational change associated with this displacement,

and the free energy change associated with this conformational change should be at least as

large as the maximum work produced by this displacement. Such a large free energy change

implies that the load dependence of the time taken to complete this step will be independent

or weakly dependent on load for low loads. Vale et al. [25] has proposed the neck linker

docking as producing the power stroke. Rice et al. [51] has shown that the free energy

45

Page 52: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

change associated with neck linker docking is small (∼ −10 pNnm). But Nishiyama et al.

[36] has shown that the time taken to complete the first 4 nm step is independent of load

for small loads. These seemingly contradictory information can be reconciled as follows.

We have suggested that the initial ATP binding to state (DP, E) is a kind of collision

complex. The power stroke is, as suggested by Vale and coworkers, driven by the neck

linker docking. We know the enthalpy change for this docking is very negative (∆H ∼ −200

pNnm), but the Gibbs free energy change is small (∼ −10 pNnm). This implies that

the entropy change is also very negative, almost compensating the potential energy gain

due to docking. The most natural interpretation of this entropy decrease is by a precise

conformation required for the neck linker docking; relatively unconstrained conformation of

the linker must assume a ‘dockable’ arrangement.

The microscopic picture we have behind our scheme is as follows. The state (DP, T) is

actually this prerequisite state for docking (or for the zipping of the neck linker) or the initial

stage of zipping (nucleated stage). For zipping, some sort of nucleation is needed. This is

probably the initial zipping of a small portion of the neck linker close to the motor core.

This initial zipping produces sufficient conformation change to induce the phosphate ion

release and subsequent detachment of the rear head. Thus, the state (D, T) is formed. The

neck linker continues to dock to the motor and this docking moves the motor forward. This

is the subsequent fast substep to (T/DP, D)1. Thus, the force dependent barrier between

(DP, E) and (DP, T) is interpreted as the largely entropic barrier for the preparation of the

dockable state. The relaxation from the top of this barrier in the forward direction into the

deep energy valley (large negative ∆H = deep energy valley; notice that the valley looks

deep from the top of the barrier, but its absolute depth is small as given by ∆G) can explain

the force independence of step in the forward direction, despite the total free energy change

associated with this process being small.

Note that the above interpretation precludes the existence of a power stroke in kinesin

operation. For the fast forward displacement of 4 nm associated with neck linker docking,

46

Page 53: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

the reverse step is also fast unlike in the power stroke picture. On completion of 4 nm

substep if motor takes a fast diffusive step forward, it is prevented from going backward due

to the head attachment to track. As shown below, such a picture can explain the energetics

of motor force production. The role of neck linker docking is interpreted as producing a bias

in the forward direction, not as a power stroke.

2.5.4 How Unique is the model?

At the beginning of the paper we quoted Duke and Leibler [33] asserting that the mechanochem-

ical data alone do not adequately select models. Thus, we have extensively used available

biochemical data. The scheme proposed by biochemical observations is unique. Modeling

of mechanical motion as due to 4 + 4 nm steps with the first one as a load independent

step produced by neck linker docking, the scheme that we obtained is unique and can ex-

plain all the available mechanochemical data. This uniqueness is contradictory to Duke and

Leibler [33] assertion. This is due to the fact that Nishiyama et al. [36] could probe the

mechanical stepping with higher resolution and give more information than that obtained

by mechanochemical measurements of motor velocity. Still, the fact that mechanochemical

experiments observe the motion of a large cargo attached to the motor elastically, not the

motor motion itself, makes such substep measurements unreliable. For example, there is a

recent observation [37] claiming that there are no substeps in the 8 nm step.

With slight modifications of kinetic constants it is found that the following schemes could

also explain mechanochemical data.

(1) We can assume that both the substeps in Nishiyama et al. [36] correspond to the rectified

diffusion of bead. In the first step, the bead moves against the load by a distance of 4 nm

and then the neck linker docks, preventing the backward motion of the bead. In the second

substep the bead moves over another 4 nm and the rebinding of the front head to microtubule

prevents its going back (as the model adopted in this paper).

(2) We may also assume that there is no substep associated with neck linker docking. The

47

Page 54: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

bead moves against the load by 8 nm and is prevented from going back by the binding of

the front head to microtubule.

The inability of mechanochemical measurements to distinguish between the different pos-

sible schemes described above can be attributed to the nature of experimental observation.

In the single molecule mechanochemical experiments such as optical trap experiments, a large

cargo elastically attached to the motor is observed to infer the motor mechanism. This may

mean the unique mechanism of motor stepping and force production is not resolved, thus

allowing more than one scheme to fit the data. A first step in trying to get more information

out of single molecule experiments will be to understand the effects of the size differences

of the motor and cargo and the elasticity of the motor-cargo link on the mechanochemical

observations.

In the next chapter we study the effects of motor-cargo link on the motor velocity and

efficiency. This leads to the conclusion that since mechanochemical experiments cannot

observe any detailed motion of motor itself, a simple mechanism involving motor diffusion

combined with modulation of motor-track interaction depending on the nucleotide state of

the motor and possibly one or more structural changes of the motor is enough to explain the

mechanochemical data. If one wishes to understand the motor mechanism, one must devise

experiments to reject this null model.

48

Page 55: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Chapter 3

Effects of the Elastic Motor-CargoLink on Motor Transport

3.1 Introduction

As discussed at the end of Chapter 2, the attempts to develop motor phenomenology require

a careful study of the effects of motor-cargo link. Attempts to explain the mechanochemical

data on kinesin in the previous chapter showed that more than one motor schemes are

compatible with these data, virtually without fitting parameters, if we remove marginally

informative experimental results. We have seen that this marginality is closely connected to

the nature of the present mechanochemical experiments: they can observe only large probes

(cargoes) attached to the motor. If we take this limitation into account, it is hard to select

the mechanism uniquely. This may render attempts to use these data to understand the

motor mechanism less useful unless we have a clearer understanding of the limitations of

these experiments. In this chapter using simple motor models we will try to understand the

effects of the elasticity of the link on mechanochemical measurements such as motor velocity

and energetics.

In addition, such a study has relevance is understanding the cellular function of motors.

The primary function of motor proteins such as kinesins and myosin V is the intracellular

transport carrying cargo that are often much larger than themselves along particular tracks

[1]. The motor and cargo are connected by a flexible link (for example, the measured stiffness

of the link for kinesin is ∼0.1 pN/nm [66]).

The study of the effect of the motor-cargo link has direct relevance to interpreting

mechanochemical experimental data often used to study the motility and force generation

49

Page 56: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

of motor proteins [16]. In optical trap experiments, as shown in Fig. 1.4 in Chapter 1, the

motion of an optically trapped bead elastically attached to the motor is analyzed to obtain

information about the motor motion. The motor can be made to move against a force ap-

plied to the bead and the kinetic features of the bead motion are interpreted in terms of

motor properties (e.g., [26]). In such studies, the effect of the link stiffness and the bead

diffusion constants are absorbed into effective kinetic rates for motor transition between

different states and are obtained by fitting to the experimental data. The spatiotemporal

resolution limits of the experimental setup mean to resolve the motion of the motor accu-

rately, a large bead has to be used in optical trap experiments. Another example is the

experimental studies of F1-ATPase motor [67] in which motion of a long fluorescent actin

filament elastically connected to the motor is observed to infer motor mechanism. In these

experiments, it is hoped that the observation of cargo motion can be used to infer the details

of motor mechanism such as the nature of force production, and efficiency. An example is

the dissipative experiments such as that of F1-ATPase, where the motor output to the cargo

is dissipated. It is found that to have a high motor efficiency (the Stokes efficiency of [24]

defined as the ratio of the cargo dissipation to motor input energy) the force experienced

by the cargo filament should be a constant. Thus, it is suggested that the motor itself pro-

duces a constant force output. As we show below, the velocity of the motor and thus the

Stokes efficiency are dependent on the link stiffness, so such a conclusion about the motor

mechanism itself is likely to be misleading (at best).

A previous study in this direction [68, 69] considered the variation of motor velocity with

the stiffness of the link when the motor is modeled as a ratchet. Detailed analytic results

for motor velocity were given for the asymptotic regions of large and small link stiffness and

cargo diffusion constant. Here we will consider the effect of the link on the efficiency of

the motor-cargo system and the limitations caused by such a setup in inferring the motor

mechanism.

50

Page 57: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

3.2 Motor-Cargo System

Since we are interested in having a physical understanding of the effect of motor-cargo link

on motor operation, we will consider simpler physical representations of motor mechanism

rather than detailed descriptions as done in the previous chapter. There can be two types of

simple motor representations depending on the coupling between energy consumption and

mechanical motion. In tight coupling models, consumption of one unit of energy (hydrolysis

of one ATP molecule in motor protein case) produces one mechanical step. The other simple

representation of motor mechanisms is in terms of loosely coupled thermal ratchets. Using

such a loosely coupled ratchet models is at variance with the reality of actual motors, but as

an illustrative system capturing physics of the directional motion of a particle in a thermal

environment, the model is useful to the understanding of real motors. There are many

possible physical realizations of the ratchet [28]. We adopt the simplest scheme proposed

in [70, 71]. The motor is modeled as a strongly damped particle moving in a (spatially

asymmetric) ratchet potential that fluctuates between two different states 1 and 2 with

potential Φ1 and Φ2, respectively. Fig. 3.1 shows the motor-cargo system moving in such a

potential. The motor-cargo link is modeled by a harmonic spring.

C MKs

γ1

a

u

γ2

Φ1

Φ2

Figure 3.1: The motor-cargo system modeled as a fluctuating ratchet with potentials Φ1 andΦ2. u is the potential maximum height, a < ℓ/2 denotes the potential asymmetry. γ1 andγ2 are the transition rates of the potential between the two states. Motor (M) and Cargo(C) are connected by a harmonic link of stiffness Ks.

51

Page 58: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

The system can be described by a set of Fokker-Planck equations of the form:

∂P1

∂s= −∂J1,m

∂X− ∂J1,c

∂Y− Γ1(X)P1 + Γ2(X)P2, (3.1)

∂P2

∂s= −∂J2,m

∂X− ∂J2,c

∂Y− Γ2(X)P2 + Γ1(X)P1, (3.2)

where X is the motor coordinate, Y the cargo coordinate, Pi(X,Y, s) the probability density

for state i ∈ {1, 2} at time s, Γi(X) the transition rate or fluctuation rate of the potentials

when the motor is at X. Here, the motor and the cargo fluxes in state i, Ji,m(X,Y, s) and

Ji,c(X,Y, s), respectively, are given by

Ji,m = −Dm

(

∂Pi

∂X+

1

kBT

∂Vi

∂XPi

)

, (3.3)

Ji,c = −D′

c

(

∂Pi

∂Y+

1

kBT

∂Vi

∂YPi

)

. (3.4)

In these formulas, kB is the Boltzmann constant, T is the temperature, Dm is the motor

diffusion constant, D′

c is the cargo diffusion constant, and the potential Vi is given by

Vi(X,Y ) = Φi(X) +1

2Ks(X − Y )2 − FY, (3.5)

where Ks is the stiffness of the link, and F is the external force acting on the cargo.

At steady state,

−∂J1,m

∂X− ∂J1,c

∂Y− Γ1(X)P1 + Γ2(X)P2 = 0, (3.6)

−∂J2,m

∂X− ∂J2,c

∂Y− Γ2(X)P2 + Γ1(X)P1 = 0. (3.7)

In a steady state, the boundary conditions to be satisfied are continuous periodic boundary

conditions in the X-direction for probability densities and fluxes: Pi(X+ℓ, Y +ℓ) = Pi(X,Y ),

and Ji,r(X+ℓ, Y +ℓ) = Ji,r(X,Y ) for r = m or c. Here, the spatial periodicity of the potential

52

Page 59: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

is ℓ. In the Y -direction, the probability densities are assumed to vanish at infinity. To have

a numerical solution, a simple finite difference scheme was found to be sufficiently accurate

to obtain the steady state.

From now on, we will work with the following dimensionless quantities: x = X/ℓ, y =

Y/ℓ, v = V/kBT , φi = Φi/kBT , Dc = D′

c/Dm, γi = ΓiDm/ℓ2, t = sℓ2/Dm, ks = Ksℓ2/kBT

and f = Fℓ/kBT .

3.3 Variation of Motor Velocity with Link Stiffness

Before studying the efficiency of the motor we will study the general behavior of the motor-

cargo system. Fig. 3.2 exhibits the variation of the motor velocity with the link stiffness

obtained by numerically solving the Fokker-Planck equations (3.6) and (3.7). We see that

the variation is non-monotonic and for moderate values of the potential transition rates

(here, γ1 = γ2 = γ as a representative case) there is an optimal stiffness of the link which

maximizes the motor velocity.

The numerical results can be understood intuitively with the aid of the case with u ≫ 1

that can be analytically studied. Assume that the motor is in state 2 at t = 0. The

probability distribution in state 2 is localized at the potential minima because u ≫ 1. Once

the motor makes a transition from state 2 to 1, the motor starts diffusing freely. Let p(x, y, t)

be the joint motor-cargo probability in state 1 at time t with the initial motor position at

x = 0. The probability of the motor being captured around x = 1 when the potential

switches back to 2 is given by

Pf =∫

adx

−∞

dy∫

0dt p(x, y, t)R(t), (3.8)

53

Page 60: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

0

0.075

0.15

0.225

0.75 1.25 1.75 2.25

Vel

ocity

log10(ks)

γ=1.6γ=160

Figure 3.2: Variation of motor velocity of the motor with the stiffness of the link for threedifferent transition rates. The potential parameters are, u = 20, a = 0.3, Dc = 0.1, andexternal load f = 0.

and the probability of being captured around x = −1 is

Pb =∫

−(1−a)

−∞

dx∫

−∞

dy∫

0dt p(x, y, t)R(t), (3.9)

where R(t) is the waiting time distribution for the state to return from 1 to 2 and may be

modeled by a Poisson process: R(t) = γe−γt.

The time evolution of p(x, y, t) is given by the coupled diffusion equation,

∂p

∂t=

∂x

(

∂p

∂x+

∂v0

∂xp

)

+ Dc∂

∂y

(

∂p

∂y+

∂v0

∂yp

)

, (3.10)

with v0(x, y) = (1/2)ks(x−y)2. To estimate the velocity, we need a time dependent solution

to the coupled diffusion equation above. The coordinate transformation z = (Dcx+y)/(Dc+

54

Page 61: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

1) and r = x − y allows the separation of these equations as,

∂Pz(z, t)

∂t= De

∂2Pz(z, t)

∂x2, (3.11)

and

∂Pr(r, t)

∂t= Dt

∂r

(

∂Pr(r, t)

∂r+

∂v0(r)

∂rPr(r, t)

)

. (3.12)

where De = Dc/(1 + Dc) and Dt = 1 + Dc. Using the time dependent solutions for the dif-

fusion equation, and the Ornstein-Uhlenbeck process, we get the solution for the probability

distribution as

p(x, y, t) =∫

−∞

dx0

−∞

dy0 p(x, y, t|x0, y0, t0) p(x0, y0, t0), (3.13)

where

p(x, y, t|x0, y0, t0) = C1exp

[

−(

(Dct(x − x0) + Dmt(y − y0))2

α+

((y − x) − δ(y0 − x0))2

β

)]

,

(3.14)

with S(t, t0) = 1 − exp (−4(t − t0)/τ), τ = 2/ksDt, α = 4De(t − t0), β = 2S(t, t0)/ks,

δ = exp(−2(t − t0)/τ), Dct = Dc/Dt, Dmt = 1/Dt and C1 is a normalization constant. The

initial distribution p(x0, y0, t0) for sufficiently high barrier in state 1 can be approximated as

p(x0, y0, t0) = C2δ(x0 − xi) exp(

−ks(y0 − x0)2/2

)

, (3.15)

where C2 is a normalization constant and δ(x0 − xi) is the Dirac delta function.

For sufficiently large γ, R(t) can be approximated as a Dirac delta function, ∼ δ(t−1/γ).

Thus, the velocity of the motor can be approximated as v ∼ (γ/2)(Pf − Pb). Using the

analytically obtained p(x, y, t), we obtain

v ∼ γ

4[Erfc(

√γ1a) − Erfc(

√γ1(l − a))] , (3.16)

55

Page 62: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

where Erfc is the complementary error function and γ1 = 1/[α + (1/Dt)2(β + (δ − 1)2ǫ)]

with α = 4De/γ, β = 2(1 − exp (−4/γτ)), τ = 2/ksDt, δ = exp(−2/γτ), ǫ = 2/ks, De =

Dc/(1 + Dc) and Dt = 1 + Dc. Figure 3.3 illustrates the velocity dependence on the link

stiffness given by (3.16). It is qualitatively similar to the behavior in Fig. 3.2; due to a

stiffer potential the main features are exaggerated in Fig. 3.3.

0.12

0.14

0.16

0.18

0.2

-1 0 1 2 3 4 0

0.2

0.4

0.6

0.8

1

Vel

ocity

(γ=

2)

Vel

ocity

(γ=

10)

log10(ks)

γ=2γ=10

Figure 3.3: Variation of the velocity of the motor determined in the analytically solvablelimit with the link stiffness ks for transition rates γ = 2 and 10. The cargo diffusion constantDc = 0.1. The potential asymmetry a = 0.3.

The various regimes in Fig. 3.3 can be understood as follows.

∗ Small transition rate and small stiffness: Here, the waiting time for motor in state 1 is

long (∼ 1/γ). The motor initially localized at x = 0 has enough time to spread. Because

wide spreading is facilitated by the small stiffness of the link, when the state changes from

1 to 2 there is nearly the same probability of motor being captured at the forward (x = 1)

and backward sites (x = −1). Thus, the motor velocity is small.

∗ Small transition rate and intermediate stiffness: As the link stiffness is increased, the width

of the motor probability distribution in state 1 decreases. This and the potential asymmetry

of state 2 cause the motor to be captured with much larger probability by the forward site

56

Page 63: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

than the backward site when transition to state 2 is made. Consequently the motor velocity

increases.

∗ Small transition rate and large stiffness: The motor distribution in state 1 remains sharply

peaked even after the long diffusion time. This decreases both the probabilities of the motor

being captured by the forward and backward sites when transition to state 2 occurs. Thus,

the motor velocity decreases.

∗ Large transition rate and small stiffness: Although the waiting time for motor in state 1

is not long, still the motor in state 1 can spread enough thanks to the weak link to feel the

asymmetry of the potential in state 2 when transition to state 2 occurs. Thus the motor can

efficiently proceed.

∗ Large transition rate and intermediate and large stiffness: the short diffusion time and

larger stiffness combine to produce a sharply peaked motor probability in state 1 and de-

creases the forward capture probability when state changes to 2. Motor remains at the same

position for most of the state changes. This causes the velocity to decrease.

For the ratchet model, the cargo effect on motor motion can be interpreted as modifying

the effective forward or backward transition probabilities of the motor. For small stiffness,

the relative probabilities of forward and backward motor steps are not much affected by the

link to the cargo, so the motor velocity is not significantly affected by the cargo irrespective

of its size. Increasing the link stiffness reduces the motor fluctuations and localizes the

motor position especially if the cargo is large. This can be beneficial in ratchet models for

which energy input is dependent on the motor position, because efficient supply of energy

to a localized motor may realize. This can, however, be disadvantageous if the motor needs

spatial exploration by thermal fluctuation to obtain energy supply.

As seen in Chapter 2 this happens for tightly coupled models of motors for which energy

supply is localized along certain track positions. Efficient diffusion of the motor allows it to

reach the supply point quickly, but the stiff link with a large cargo hinders this motion. We

consider three such scenarios shown in Fig. 3.4: in Case (1) energy is supplied continuously

57

Page 64: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

along the motor coordinate. Cases (2) and (3) require diffusive search by the motor for the

completion of a step. Fig. 3.5 shows the variation of velocity with the link stiffness obtained

by solving the associated Fokker-Planck equations numerically.

In Case (1), since there is continuous dragging of the motor to the forward direction,

motor velocity is independent of the link stiffness and is given by v = △µ/(1+Dc). In Cases

(2) and (3), however, the increasing stiffness of the link decreases the motor velocity. This

decrease is understandable if the localizing of the motor due to the increasing stiffness of the

link is taken into account. For low stiffness, the motor starting at the left is able to diffuse

easily and reach the other end. Increased link stiffness makes this process difficult and slows

down the motor. A similar scenario must be operational in the cellular conditions in which

motor carries cargo which is usually much larger than themselves. The motor-cargo link

must be sufficiently flexible for the transport to be sufficiently fast. Similarly, the velocity

measured in optical trap experiments will also be dependent on the size of the bead used

and the trap stiffness.

C

12

3∆µ

¼l½l l

M

Figure 3.4: △µ is the free energy which is released to the motor during one step of its motion.In (1) the energy is gradually released along the motor coordinate. (2) and (3) involve freediffusing regions within the step.

58

Page 65: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

0.2

0.45

0.7

0.95

0.75 1.25 1.75 2.25 2.75

Ve

loci

ty

log10(ks)

a=1a=0.5

a=0.25

Figure 3.5: Variation of the velocity of the motor with the stiffness of the link for threedifferent ways of free energy release within a step. The potential parameters are, △µ = 9,Dc = 0.1, and external load f = 0.0.

3.4 Stokes Efficiency of Motor-Cargo System

Molecular motors often drag cargoes that are much larger than themselves. In the absence

of a measurable external force applied to the motor it is difficult to give a thermodynamic

definition of motor work output and thus to define its efficiency. This is also the case

in the common experimental setup used in the study of motor in which a large element

(cargo) is attached elastically to the motor and its observed motion is used to infer the

motor mechanism. An example is the study of the F1-ATPase motors in which a large

actin filament attached to the motor is observed [67]. In these experiments the measurable

quantities are the rate at which energy is supplied to the motor and the velocity of the

attached element. Using these two observables, Oster and coworkers [24] proposed a new

definition of motor efficiency called the Stokes efficiency defined as the ratio of the rate of

mechanical dissipation of the energy at the cargo 〈Qc〉 and the motor energy input rate

〈Rm〉 (Here, all the quantities are dimensionless. Energy rates are in units of kBTD′

c/ℓ2 and

59

Page 66: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

velocity is in units of D′

c/ℓ).

ηStokes =〈Qc〉〈Rm〉

, (3.17)

where

〈Qc〉 = ζc〈v〉2 (3.18)

with ζc = 1/Dc being the cargo friction constant and 〈v〉 its average velocity.

Oster et al. has shown that the Stokes efficiency can give information about the nature of

motor energy output. For example, a high Stokes efficiency implies a constant motor energy

output without any spatial irregularities in the potential. However, the conclusion is true

only when one neglects the link between the motor and the cargo.

To calculate the Stokes efficiency we need the energy supply rate. Each change of po-

tential from φ1 to φ2 supplies energy φ2 − φ1 ≡ v2 − v1. Therefore, the mean rate of energy

input to the motor is

〈Rm〉 =∫ 1

0dx∫

−∞

dy(v2(x, y) − v1(x, y)) · (γ1P1(x, y) − γ2P2(x, y)). (3.19)

Figure 3.6 shows that the Stokes efficiency for the thermal ratchet described in Section

3.2 is dependent on the stiffness of the link and there is an optimal stiffness which maximizes

the Stokes efficiency for γ = 1.6. The explanation for this peak and variation is the same

as that for the velocity in the preceding section. The effective motor transition rates to the

forward or backward site is dependent on the potential transition rate and the link stiffness.

A much more serious effect of the link stiffness can be seen in the case of the tightly

coupled model discussed at the end of Section 3.3. In this case the Stokes efficiency is defined

as v2/Dcr△µ, where r is the rate of motor energy input and △µ the free energy input per

step. For unit step length, r = v and hence the Stokes efficiency is given as v/Dc△µ. Figure

3.7 shows the variation of the Stokes efficiency with the link stiffness corresponding to the

three cases in Fig. 3.4. We see that the softer link increases the Stokes efficiency for a given

60

Page 67: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

0

0.002

0.004

0.006

0.008

0.01

0.75 1.25 1.75 2.25

Sto

kes

Eff

icie

ncy

log10(ks)

γ=1.6γ=160

Figure 3.6: Variation of the Stokes efficiency of the motor for the ratchet model with thestiffness of the link for two different transition rates. The potential parameters are, u = 20,a = 0.3, Dc = 0.1, and external load f = 0.

energy release pattern within a step. This dependence on the stiffness of the link imposes a

serious limitation on the quantitative theoretical argument based on the Stokes efficiency on

the motor mechanism. For example, suppose the measured Stokes efficiency is around 0.45.

The theoretical argument in [24] disregarding the link effect would conclude that a = 1/2

(under the assumption that the potential has only one sloped portion). Actually, however,

any value of a between 1/2 and 1/4 is possible dependent on the stiffness. If we use realistic

(i.e., rather soft) link, a ≃ 1/4 is likely.

If we use stochastic energetics formalism [72], a more detailed energetics study can be

performed. For completeness sake, some results are summarized in Appendix B.

3.5 Discussion

We have demonstrated that motor velocity, and efficiency depend on the motor-cargo link

and cargo diffusion constant significantly. As shown explicitly, quantities such as the Stokes

61

Page 68: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

0

0.2

0.4

0.6

0.8

1

0.75 1.25 1.75 2.25 2.75

Sto

kes

Eff

icie

ncy

log10(ks)

a=1a=0.5

a=0.25

Figure 3.7: Variation of the Stokes efficiency of the tight coupled motor with the stiffness ofthe link for three different ways of free energy release within a step. The potential parametersare, △µ=9, Dc =.1, and external load f = 0.0

efficiency depend on the motor-cargo link, so the inference one draws from Stokes efficiency

alone about the nature of motor force production may be wrong if the effect of cargo size and

link stiffness are not modeled explicitly. This indicates that discussing the motor mechanism

based on the mechanochemical experiments requires caution. Incidentally, we found that for

a simple loose coupling ratchet model of the motor there is an optimal stiffness of the link

which maximizes the velocity and the efficiency of the motor. For a tightly coupled model of

the motor, both the velocity and efficiency increase as the stiffness of the link is decreased.

As suggested by the above observations, the nature of information one gets out of optical

trap experiments on motor proteins are dependent on the experimental setup. This combined

with the observation of Chapter 2 that these experiments are not good at resolving the

mechanism of motor force production suggests that there might be a unified interpretation

of these experiments which can be obtained without modeling motor mechanism. In the

next chapter, we attempt to develop such a unified interpretation of these single molecule

experiments.

62

Page 69: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Chapter 4

Interpretation of Single MoleculeExperiments of Motor Proteins

4.1 Introduction

Single molecule experiments are currently one of the most powerful methods for the physical

understanding of motor proteins [1, 16, 73]. In one class of experiments motion of a cargo

(an optically trapped bead much larger than the motor) elastically attached to a motor is

analyzed to obtain information about the kinetic and energetic aspects of the motor motion

[16, 74]. A variation of this setup observes an attached cargo (a fluorescent element) in a

viscous medium [75, 76]. Because the observable is the mechanical motion of the cargo, as

shown in the previous chapter, modeling of the motor-cargo system is required to infer the

motor mechanism and energetics from the observed results. Such models may be formulated

by assuming particular motor mechanisms. If the model results are in agreement with

experimentally measured quantities such as motor velocity, efficiency, etc. [26, 29, 31, 24],

these assumed mechanisms are taken as representing the actual motor mechanism.

As shown in Chapter 3, motor properties such as velocity and efficiency are dependent

on the size of the cargo and the stiffness of the motor-cargo link. Thus, the results of models

which neglect the effect of motor-cargo link (for e. g. [32, 26]) should be treated with caution

even if they show agreement with experimental measurements such as motor velocity. In

this chapter we will show that if we take into account the time scale distinction between the

motor and the cargo fluctuations, then:

(1) we can devise a unified description of single molecule experiments of various molecular

motors;

63

Page 70: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

MM M

Cargo Cargo Cargo

Figure 4.1: A large cargo is connected to motor M through a long elastic link. The dashedvertical line shows the starting position of the cargo.

(2) no information about the details of the driving force generated by the motor is needed

for quantitative understanding of single molecule experiments so far obtained.

4.2 Motor-Cargo System

In the system used in the single molecule studies as shown in Fig. 4.1, a large cargo is

connected elastically to the motor which moves along a track with evenly spaced binding

sites. The motor moves along a linear track with evenly spaced binding sites. A typical

motor step involves a chemical process (e.g., ATP binding) at one binding site, unbinding

of the motor from the site, its diffusive motion biased in the forward direction by a driving

potential or by some other mechanism, and rebinding at the next binding site which is

stabilized by another chemical process (e.g., the release of products of ATP hydrolysis).

Forward motion of the motor stores energy in the elastic link connecting it to the cargo and

it is the tensile force in the link that drives the cargo forward.

A single molecule mechanochemical experiment typically monitors the cargo stepping

with nanometer resolution when a load is applied to it using an optical trap or by the

viscosity of the surrounding medium. In Chapter 3, we modeled the elastic link as a har-

64

Page 71: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

monic spring studied the cargo response when the motor potential is modeled as a fluctu-

ating ratchet. Such a description cannot be used for a quantitative understanding of single

molecule experiments, because it assumes that chemical reaction time scales are much faster

than the mechanical motion time scales.

In the typical single molecule experimental setup the dimensions of the cargo is ∼1 µm

and that of the motor ∼1 nm. Thus, the Einstein relation tells us that the ratio of the

diffusion time scales (which are proportional to the dimension) of the motor and the cargo

is about 103. This implies that the fluctuation time scales which are orders of magnitude

different.

While motor jumps between one track binding position to next, it is working against the

strain in the link connecting it to the cargo and storing energy in the link by stretching it.

The time taken for this motor diffusive motion against the strain in the link can be estimated

as [1] τ = τ0 exp(E)/Eα, where E is the required energy storage (in units of kBT ), τ0 is the

diffusive time scale of 0.1 µs, and α = 1.5 (if energy is stored in a harmonic spring) or 2

(if stored in a linear potential). The energy stored in the link can be estimated using the

maximum force generated by the motor (also called the stalling force). For kinesin, using

the known stalling force which is in the range of 7 pN [16], we estimate τ < 1 ms, which is

well within the order of slowest chemical step completion time and is compatible with the

observed cargo velocities (this implies motor diffusion time scales does not affect the cargo

velocity). For the F1-ATPase, the known stalling force gives E ≃ 80 pNnm [73]. Here, using

the experimental information [78] that single step of 120 degree consists of substeps of 80

and 40 degrees, the diffusive time scale again is less than 1 ms.

65

Page 72: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

4.2.1 An Effective Cargo Only Representation of a Coupled

Motor-Cargo System

As shown in Chapter 2, to ensure tight coupling (consumption of one ATP molecule to pro-

duce one step) a particular reaction in ATP hydrolysis cycle occurs only when the motor

position is within a small range around a particular locations on the track (modulo peri-

odicity) [1]. When the motor reaches a reaction position, the cargo is still lingering in the

position corresponding to the previous reaction position of the motor, because due to its

much smaller size the motor moves much faster than the cargo. This stretches the elastic

link connecting the motor and cargo. While the motor is waiting for the next chemical

reaction to occur, the strain in the link pulls the cargo forward. Before the motor moves to

the next reaction position, cargo reaches its equilibrium position corresponding to the motor

position for the just completed reaction. Therefore, the occurrence of various chemical reac-

tions associated with nucleotide states/motor-track binding states and the cargo positions

are roughly one-to-one correspondent. We may use these cargo locations as the ‘checking

positions’ for reactions. (These checking positions depend on external forces, etc., but we

need only relative distances between successive checking positions which are expected to

be independent of external forces, etc.). This suggests that the kinetic constants obtained

from the biochemical observations of the motor alone can be used in the explanation of the

experiments. Between such ‘checking positions’ the cargo is pulled by the motor. The time

scales of the motor and the cargo are very different and the slowest time scale is introduced

by the chemical reaction time scale (essentially motor waiting time along fixed track posi-

tions). This allows a simplified representation of the coupled motor-cargo motion as that

of the cargo alone in an effective potential exerted on it by the strain in the motor-cargo

link. For reasonably elastic links this potential can be approximated as a linear potential

U = fy made by the cargo-motor link (y > 0 is the usual moving direction), where f (< 0)

is the average tensile force in the link, irrespective of the details of the potential acting on

66

Page 73: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

the motor [68].

As is seen below the cargo only model with a constant force may be justified with a

singular perturbative reduction of the motor-cargo full model augmented by a further sim-

plifying approximation that can be checked numerically. Let the cargo position be y and the

motor position x. Since the time scale of x is much faster, the coupled stochastic differential

equations has the following form with a small parameter ǫ > 0:

dx = [−A(x − y) + F (x)]dt +√

2AdB,

dy = −ǫA(y − x)dt +√

2ǫAdB′, (4.1)

where A > 0 corresponds to the stiffness of the link connecting motor and cargo, F (x) is the

potential experienced by the motor during its stepping due to structural changes introduced

by ATP hydrolysis processes or due to motor track interactions, B and B′ are independent

Wiener processes and units are chosen such that kBT = 1. If as in our case F allows a stable

fixed point corresponding to the next motor position, the equation for y reads, to the lowest

nontrivial order,

dy = −ǫA(y − 〈x〉y)dt +√

2ǫAdB′, (4.2)

where 〈x〉y is the average position of x when the cargo position is y. Since F exerted on the

trapped motor is strong, 〈x〉y is almost y-independent and is the next reaction position. Up

to this point the approximation is a systematic expansion with respect to ǫ (we could employ

the standard averaging method). In our approach, to simplify further the term −ǫA(y−〈x〉)

is replaced with an appropriate average −f (where f (< 0) is the average tensile force in

the link). Needless to say, with this approximation we lose stepwise sample motions of the

cargo, but still overall average results can be recovered. The magnitude of required f is later

shown to be realistic. One benefit of this constant force assumption is that the model can

67

Page 74: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

be solved analytically. In Chapter 5, this constant force assumption is further verified by

the simulation of combined motor-cargo system.

Thus, the observed motor-cargo motion can be modeled as the cargo diffusion under

constant force (due to the link tensile force + external force on it, which need not be a

potential force) between appropriate ‘checking positions’ that correspond to the chemical

reactions. The above consideration allows us to disregard the details of the potential (in

other words, the mechanochemical experimental results are expected not to reflect motor

mechanisms) experienced by the motor during its motion between binding sites on the track.

The time scale separation between the slower chemical reaction times (usually of the order

of tens of milliseconds or longer) and motor diffusion times (even for nondriven diffusive

motion) makes this consideration reasonable.

4.2.2 Mathematical Formulation of Effective Cargo Only

Representation

A unified framework to explain the observable dynamics is illustrated for the following

mechanochemical scheme with two substeps which can describe the mechanochemistry of

F1-ATPase:

(a)k′1[ATP ]

⇀↽k−1

(b) Substep1⇀↽

(c)k2⇀↽

k′

−2[P ](d)

k3⇀↽

k′

−3[ADP ](e) Substep2

⇀↽(f).

Here the forward and backward kinetic rate constants are k′

1, · · · , k′

−3 and [ATP], [P] and

[ADP] are the concentrations of ATP, phosphate ion and ADP, respectively. Once ATP

binds to the motor in state (a) located at x1 = 0, state (b) is formed and motor detaches

from the track. Motor makes a substep of length x2 and forms state (c) with hydrolyzed ATP

attached to it. When the motor displacement is complete and the cargo reaches the next

‘checking position,’ phosphate ion P is released and state (d) is formed. With the release of

68

Page 75: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

ADP state (e) is formed and motor is again detached from the track. On completion of the

next substep the starting state of the motor with one step (of length xl) advanced state (f)

is formed.

To treat the diffusion and chemical reaction processes in a unified fashion, state proba-

bilities are introduced. Let Px (for x = a, · · ·, e) be the probability of the cargo with the

motor being in state x which we assume is localized within a length scale ∆l along the track

with uniform probability density Px/∆l. In state (b) or (e) the cargo can move: ρb(y) and

ρe(y) are the probability densities of the cargo at angle y for states (b) and (e), respectively.

Time evolution of probabilities can be described by the following equations:

dPa

dt= −k1Pa + k−1Pb + J ′

a,

dPb

dt= −k−1Pb + k1Pa − Jb,

∂ρb(y)

∂t= −∂Jb(y)

∂y,

dPc

dt= −k2Pc + k−2Pd + J ′

c,

dPd

dt= − (k3 + k−2) Pd + k2Pc + k−3Pe,

dPe

dt= −k−3Pe + k3Pd − Je,

∂ρe(y)

∂t= −∂Je(y)

∂y, (4.3)

where k1 = k′

1[ATP], k−2 = k′

−2[P] and k−3 = k′

−3[ADP]. Jx and J ′

x are, respectively, the

outgoing and incoming fluxes at state x. The cargo diffusive motion is described by the

Fokker-Planck equations with the flux Jx(y) in the above equation given as

Jx(y) = −Dc

(

∂ρx(y)

∂y+

1

kBTfρx(y)

)

, (4.4)

for x = b or e. Here, f is the effective driving force experienced by the cargo. If there is a

load fL acting on the cargo, f in the above expression must be replaced by f − fL. In this

69

Page 76: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

model the boundary conditions to solve partial differential equations are Dirichlet conditions

for concentration matching. However, in the steady state, we are interested in the system

with a periodic boundary condition, so continuity of probabilities and fluxes are required.

In a steady state, all the time derivatives vanish. Integrating dJb/dy = 0 from 0 to

y gives Jb(y) = J , where J is the steady probability flux. With the boundary conditions

ρb(0) = Pb/∆l, ρb(x2) = Pc/∆l, and the flux continuity requirement Jb(x) = J = Jb = J ′

c

we obtain

Jb(y) = −Dc e−βfy d

dy

(

ρb(y) eβfy)

= k1Pa − k−1Pb. (4.5)

Integrating this from y = 0 to x2 gives

(Pc/∆l) eβfx2 − (Pb/∆l) = − 1

Dc

∫ x2

0

eβfydy × (k1Pa − k−1Pb). (4.6)

Analogous calculation for Je(y) yields

(Pa/∆l) eβfxl − (Pe/∆l) eβfx2 = − 1

Dc

∫ xl

x2

eβfydy × (k3Pd − k−3Pe). (4.7)

With the normalization condition

Pa + Pb +∫ x2

0ρb(y)dy + Pc + Pd + Pe +

∫ xl

x2

ρedy = 1 (4.8)

we have a closed set of equations for steady state probabilities and the steady flux J .

Though cumbersome, analytical solution of the above equations can be obtained very

easily with the aid of, e.g., Mathematica. The exact expression for the motor velocity v

(= xlJ) is given as

v = C

(

1 − eβfxlk−1k−2k−3

k1k2k3

)

, (4.9)

where the coefficient C is a function of Dc, effective motor driving force f (< 0) and the

reaction rates. It has an explicit analytic formula which is very long and hence is not shown

here. Note that if there is an external load fL acting on the cargo, the effective driving force

f is replaced by f − fL and this way external load enters into velocity expression.

70

Page 77: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Since we know that Pi release rate is much faster than ADP release rate and both

processes occur at the same motor location, we combine these two processes into a single

process with forward rate k∗

2 and reverse rate k∗

−2 and get the following expression for motor

velocity v as

v = C

(

1 − eβfxlk−1k

−2

k1k∗

2

)

, (4.10)

where the coefficient C has the following explicit form,

C−1 =1

k1k2β2Dceβfx2∆l[β2 Dc f 2 (eβ f x2 (k1 + k∗

2) + e2 β f x2 (k1 + k−1) + eβ f xl (k∗

2 + k∗

−2)

+ eβ f (x2+xl) (k−1 + k∗

−2)) ∆l + (e2 β f x2 (k1 − k−1) (k∗

2 − k∗

−2) + eβ f xl (k1 − k−1)

× (k∗

2 − k∗

−2) + eβ f x2 (−2 k1 k∗

2 + k∗

2 k−1 + k1 k∗

−2) + eβ f (x2+xl) (k∗

2 k−1 + k1 k∗

−2

− 2 k−1 k∗

−2)) ∆l + β f (Dc (−(eβ f x2 (k1 + k∗

2)) + e2 β f x2 (k1 − k−1) + eβ f xl (k∗

2 − k∗

−2)

+ eβ f (x2 + xl) (k−1 + k∗

−2)) + ∆l (e2 β f x2 (k1 + k−1) (k∗

2 − k∗

−2) ∆l + eβ f xl (k1 − k−1)

× (k∗

2 + k∗

−2) ∆l − eβ f x2 (2 k1 k∗

2 ∆l + k∗

2 k−1 ∆l + k1 k∗

−2 ∆l + k1 k∗

2 xl) + eβ f (x2 + xl)

× (k∗

2 k−1 ∆l + k1 k∗

−2 ∆l + 2 k−1 k∗

−2 ∆l + k−1 k∗

−2 xl)))]. (4.11)

4.3 Explanation of Mechanochemical Data

The general scheme illustrated above can explain quantitatively the load dependence of

velocity of F1-ATPase, kinesin, and myosin V obtained by optical trap as well as viscous

load experiments with the aid of the published reaction rates and a single fitting parameter:

f , the effective driving force exerted by the motor on the cargo which is a constant for a

given motor.

71

Page 78: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Table 4.1: The adopted rate constants for F1-ATPase (from Pake et al. [79])

Process Rate constant

ATP binding 2.08 µM−1s−1

ATP unbinding 270 s−1

Phosphate ion release 2030×105 s−1

Phosphate ion binding 0.81 µM−1s−1

ADP unbinding 490 s−1

ADP binding 8.9µM−1s−1

4.3.1 F1-ATPase

F1-ATPase is a part of F0F1-ATP synthase. It consists of a ring-like structure α3β3 and

a central shaft γ that can be rotated using the free energy of ATP hydrolysis and thus

produce work with high efficiency [76]. The motor rotation (which is identified with the

γ rotation) is tightly coupled to the consumption of ATP. The energetics of this motor is

studied by attaching a large fluorescent actin filament of varying length to the rotating shaft

and observing its rotation [73]. It is also known that each motor step of 120 degrees consists

of substeps of 80 and 40 degrees [78]. The experimentally measured chemical rate constants

are given in Table 4.1. The load on the motor is that of viscous drag acting on the filament.

The load variation of motor velocity can be explained by above scheme with substeps with

available kinetic constants from biochemical experiments. Fig. 4.2 shows the variation of

velocity of the motor with the increasing length of the attached actin filament. Even though

[73] presents data for different ATP concentrations, most of the data were obtained by using

ATP regenerating system which did not monitor the concentrations of ADP and Pi. There

was only one data set which monitored the concentrations of ADP and Pi and this data

set was quantitatively explained by our model. The other data could also be explained by

assuming particular realistic ADP and P concentrations. The diffusion constant Dc was

estimated as [76] Dc = kBT/γ with γ = (4π/3)ηL3/log(L/2r)) − 0.447, where L is the

length of the actin filament, η the viscosity whose value is taken as that of viscosity of water

72

Page 79: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

0.01

0.1

1

10

100

0 1 2 3 4

Ro

tatio

n r

ate

(r

p s

)

Actin length (µm)

Figure 4.2: Variation of actin rotation rate with increasing length. [ATP] = 20 µM , [ADP]= 1 µM and [P] = 100 µM . Solid line is the results from the analytic model with an effectivedriving torque of −40 pNnm for both substeps and ∆l = 5 degrees.

(10−3Ns/m2), and r = 5 nm the radius of the actin filament. It is found that results are

not very sensitive to the exact value of ∆l as long as it is less than 10 degrees.

4.3.2 Kinesin

Kinesin hydrolyzes one ATP while moving along the microtubule track by one step (xl = 8

nm) [16]. There are two approaches to study the load dependence of the velocity of kinesin.

[16] uses an optical trap setup which can apply a constant load to a bead of diameter 0.5

µm attached to the motor. Dependence of motor velocity on the external load for different

concentrations of ATP was studied. It is also observed that at very high loads the motor

can make reverse steps [37]. The probability of these reverse steps are very low and for

loads away from stalling they can be neglected. In this case, the model with the following

mechanochemical scheme is able to explain the load variation with external load.

(a)k′1[ATP ]

⇀↽k−1

(b) Mechanical motion⇀↽

(c)k2⇀↽

k′

−2[P ](d)

k3⇀↽

k′

−3[ADP ](e)

73

Page 80: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Table 4.2: The adopted rate constants for kinesin (from Cross [80])

Process Rate constant

ATP binding 2 µM−1s−1

ATP unbinding 80s−1

Phosphate ion release 300s−1

Phosphate ion binding 0.25 µM−1s−1

ADP unbinding 300s−1

ADP binding 4.5 µM−1s−1

The experimentally measured rate constants given in Table 4.2 were used to explain the

velocity data.

Figure 4.3 shows the variation of velocity of kinesin with load for two different ATP

concentrations. The same model also explains the variation of the velocity of the motor

with changing ATP concentration [16] as well as the case of motor response to a viscous

load as in [75]. Also note that as shown in [81] the velocity of the motor does not change

significantly when a forward load is applied to the motor. In this case, even if the forward

load accelerates the mechanical motion, the net motor stepping rate will be limited by the

chemical kinetic rates which are independent of load.

The data in [16] were obtained by using ATP regenerating system which did not monitor

the concentrations of ADP and P concentrations. This makes comparisons with experimental

data difficult especially for low ATP concentration. For [ATP] = 5 µM , to fit the data it had

to be assumed that the concentration of ADP present also increases with external load. For

[ATP] = 2 mM, we chose [P] = 100 µM and [ADP] = 10 µM to get the stall force in agreement

with experimental data. For the [ATP] = 5 µM, with the above kinetic constants and ADP

and P concentrations, the load dependent velocity decrease was much less pronounced than

that of experimental data. Velocity stays nearly constant for a range of loads (up to 4 pN)

and then suddenly decreases. To get the curve above which fits the data we assumed that

74

Page 81: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

0

200

400

600

800

1000

-6 -4 -2 0 2 4 6 8

Velo

city

(nm

/s)

Load (pN)

[ATP]=2mM[ATP]=5µM

Figure 4.3: Variation of velocity of kinesin with load for [ATP] = 2mM and [ATP] = 5 µM.Experimental data of [16] is also shown. Driving force f = −6 pN, and ∆l = 1 nm.

the unmonitored ADP concentration also increases with load fL as, [ADP] × exp(2fLβ).

Here the diffusion constant Dc of the cargo is obtained as Dc = kBT/6πηa, where η is taken

as the viscosity of water (10−3Ns/m2) and a = 0.25 µm, the radius of the bead used in [16].

4.3.3 Myosin V

Myosin V walks processively along actin with step lengths of 36 nm [42]. [82] measured the

variation of velocity of this motor with an external load using an optical trap setup with bead

diameter of 0.356 µm. Using the experimentally determined rate constants given in Table

4.3, the chemical scheme without substeps used to explain kinesin data can quantitatively

explain the velocity variation with load for myosin V as well. Figure 4.4 shows the variation

of velocity of the motor with load.

75

Page 82: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Table 4.3: The adopted rate constants for myosin V

Process Rate constant Reference

ATP binding 0.9 µM−1s−1 [83]ATP unbinding 1.2s−1 [84]Phosphate ion release 220s−1 [85]Phosphate ion binding 0.01 µM−1s−1 [84]ADP unbinding 12s−1 [83]ADP binding 12.6 µM−1s−1 [86]

0

100

200

300

400

500

-5 -4 -3 -2 -1 0 1 2 3

Ve

loci

ty (

nm

/s)

Load (pN)

Figure 4.4: Variation of velocity with load for myosin V at [ATP] = 2mM. Experimentaldata from [82] is also shown. It was assumed that ADP and P were also present with aconcentration of 5 µM. ∆l = 1 nm and the driving force f=−1.75 pN.

76

Page 83: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

4.4 Discussion

A simple scheme to interpret the single molecule experimental data on motor proteins derived

from optical trap and viscous load experiments is proposed. A significant feature of the

scheme is that it treats diffusion and chemical reactions in a unified fashion. With a fitting

parameter (the constant driving force exerted by motor on cargo, which is a given constant

for a given motor and linkage independent of the experimental condition) the available motor

experiments for F1ATPase, kinesin, and myosin V are reproduced quantitatively. Also, it

is straightforward to modify the scheme to include the possibility of backward steps (the

fact that the probability of backward steps is very low at loads below stalling and in viscous

experiments implies that incorporation of backward steps in our scheme is not necessary to

explain the available experimental data).

One requirement for the application of the scheme is the availability of experimentally

determined biochemical rate constants. In the absence of known kinetic constants, it is

still possible to assume reasonable rate constants (or treat them as fitting parameters). An

example is the application of the scheme to the recently published results on viral DNA

packaging motor φ29 [87]. With reasonable rate constants the observed velocity variation

with load can be explained.

In the modeling we have assumed that the effective driving force exerted by the motor on

the cargo is constant. This is an approximation that is strictly valid at very low stiffness of the

link connecting them. As shown in Chapter 5, numerical simulation of the complete motor-

cargo system with a realistic link demonstrates that, the linear potential approximation is

quantitatively reliable over a wide range of link stiffness.

The most important conclusion is that models which do not assume anything about the

details of motor force production mechanisms can explain the single molecule mechanochem-

ical experimental data. The orders of magnitude differences in motor and cargo fluctuation

time scales and the presence of a link connecting motor and cargo imply that the single

77

Page 84: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

molecule experiments cannot probe the details of motor force production. The fact that

models which assume a particular motor mechanism (such as the existence of a power stroke)

can explain single molecule mechanochemical experimental results does not guarantee the

existence of power stroke in the actual motor operation. Additional supporting information

from experiments which probe the motor motion, or conformational changes directly are

necessary.

In the next chapter, we will test the conclusions of this chapter by modeling the combined

motor-cargo system (i.e., without reducing it to a cargo only model) of the rotary motor

F1-ATPase.

78

Page 85: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Chapter 5

Affinity Switch Model for RotaryMotor F1-ATPase

5.1 Introduction

As described in Chapter 1, F1-ATPase is an ATP factory in cells [7]. To make this Chapter

self contained we will restate some of the experimental observations on this motor and general

modeling approaches to motor proteins. The motor consists of the threefold symmetric outer

ring α3β3 which surrounds the shaft γ that lacks three fold symmetry (as shown in Fig. 1.3

in Chapter 1) [88]. If the γ shaft is rotated, the resultant mechanical work drives the

ATP synthesis by combining ADP and inorganic phosphate Pi. Conversely, utilizing ATP

molecules, F1-ATPase can rotate as a molecular motor, using the free energy available from

ATP hydrolysis [88]. Its successful crystallization [89] and advancements in single molecule

experimental techniques [73, 78] greatly facilitated physical characterization of this motor.

Observation of a large cargo attached to the γ subunit has shown that coordinated ATP

binding and hydrolysis processes at three binding pockets at the α-β interface rotate the γ

rotor. One revolution is made up of three discrete steps of 120 degrees, each of which is

tightly coupled to the consumption of one ATP molecule and is made up of substeps of 80

and 40 degrees [78]. It is also known that the β structural subunit changes its conformation

[88] depending on the state of nucleotide (ATP or its hydrolysis intermediates ADPPi, the

hydrolysed ATP, ADP and inorganic Pi) at the pocket.

Development of ratchet models [20, 28, 90, 91] have provided physical insights into the

operation of mesoscopic motors working in a thermal environment. For a system maintained

at nonequilibrium, thermal fluctuations can be rectified to produce directional motion by

79

Page 86: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

using asymmetric potentials. Application of abstract ratchet models to specific biological

motors has been less successful in producing quantitative agreement with mechanochemical

data. Models which attempt to produce quantitative agreement have relied on identifying

a conformational change within the motor which converts the free energy associated with

ATP hydrolysis into mechanical work needed for cargo motion. Such power stroke models for

F1 motor [24, 92, 93, 94] identify the open to closed conformational change of β, following

ATP binding and hydrolysis, as pushing the γ forward. The underlying assumptions are

that there are no strong binding interactions between β and γ and that the asymmetric γ

structure produces steric interactions which convert the closing motion of β into forward γ

motion.

Though such models are appealing for their simplicity and correspondence with macro-

scopic motors with levers and crank-shafts, the underlying assumptions are indirect interpre-

tations of known experimental data. There are localized specific amino acid residues along

both β and γ [95, 96], replacement of which by any non-specific residues causes the the mo-

tor to lose its activity completely or partially. These specific residues have opposite charges,

allowing for strong binding of γ onto β. Even in the absence of γ and ATP, β can close

spontaneously [21, 97]. There are no clear experimental data indicating large free energy

change associated with closing of β. The above two observations can be used to question the

assumption that the conformational change of β alone produces the necessary free energy

change required to drive γ forward. In addition, as shown in Chapters 3 and 4, the fact

that a model can explain the mechanochemical data does not mean the force production

assumption of the model is necessarily correct.

The system is of mesoscopic size operating in a thermal environment. Therefore, even

without explicit driving, if the motions in wrong directions are appropriately checked by

conformational changes in β, there is a possibility that the desired motion of γ in the allowed

time scales is accomplished by thermal fluctuations alone. The conformational changes of β

provide affinity switches between γ and β as well as barriers checking wrong motions and

80

Page 87: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

thus acting as the rectifiers of thermal motion of γ. In this chapter, we demonstrate that this

scenario can quantitatively explain single molecule experimental results without any fitting

parameter.

5.2 Model

The structural data on F1 can be abstracted as the motor having three nucleotide pockets

interacting with γ that lacks three-fold symmetry. A γ rotation step consists of two substeps

of 80 degrees and of 40 degrees [78]. Between these substeps, the γ is in an immobile waiting

state and the steps are triggered by changes in the occupancy of the ATP binding pocket.

Biochemical data shows that the first substep follows ATP binding and the second follows

Pi release and possibly ADP release [98]. Below, T, E, DP, D imply, respectively, as in the

previous chapters, the occupancy of the pocket with ATP, nothing (empty), hydrolyzed ATP

before releasing Pi, and ADP. Structural data shows that the conformation of β depends on

the state of nucleotide in the pocket. It is in a closed conformation for the T state, half open

conformation in the DP state and open in the E state [89].

We use the following assumptions:

(1) the immobile waiting state of γ is produced by the strong β-γ binding interaction con-

trolled by the state of nucleotide in the pocket; (2) the orientation of γ and the chemical

reactions associated with ATP hydrolysis are tightly coupled (a particular reaction in the

nucleotide pocket is realizable only when it makes a right relative angle with γ); (3) the

conformational changes of β associated with nucleotide changes acts as an affinity switch

between β and γ or as barriers which blocks the ‘wrong’ motions of γ. According to the

experimental observations [102, 99] γ in a waiting state cannot be pulled in the backward

direction even with applying large forces on it. A natural interpretation of these facts is

that there are barriers preventing wrong directional motions. In addition, such barriers are

logically required to ensure tight coupled chemical reactions associated with ATP hydrolysis

81

Page 88: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

to occur in a correct sequence; (4) in between such binding sites or barriers γ undergoes

Brownian rotation due to thermal noise.

Based on these assumptions we propose the following scheme (see Fig. 5.1) to explain the

mechanism of this motor. The waiting state is (a). On binding of ATP, γ is released from

T DP

E

DP

T

DP D

T

DPT

T

DP DP

T

E

T

DP

ATP binding

80 deg rotation

Phosphate release ADP release

40 deg rotation

(a) (b) (c) (d) (e) (f )

(a)

(b)

(c)

(d)

E

Figure 5.1: Model for the rotation of F1-ATPase. The three nucleotide binding pockets arelabeled according to the state of occupying nucleotide. As the motor cycles through states(a) · · · (f), γ shaft (the triangle at the center) rotates by 120 degrees in substeps of 80 and40 degrees. The lower part of the figure shows the potential experienced by the γ elementin the different motor states.

the ring (b) and undergoes thermal fluctuations. There is a barrier preventing the backward

motion of γ beyond the starting point. When thermal fluctuation of γ takes it to 80 degrees

(c), γ is trapped in another potential well produced by the strong γ-β interaction. The

previously bound ATP is hydrolyzed at this location forming DP bound state. Pi release

at the DP bound β gives the D bound β (d). On ADP release from this β (e), γ is again

released from the well and starts undergoing Brownian motion. There is a potential barrier

82

Page 89: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

that prevents the Brownian motion to go back from the 80 degree position. If a 40 degree

rotation of γ is completed (f), it is in the potential well of the waiting state and 1/3 rotation

is complete.

The lower part of Fig. 5.1 shows the energetic picture of the model. γ starts in a

state strongly bound to β (the leftmost well of the potential curve (a)). On binding of

ATP, conformational changes at the β-γ interface modify the β-γ interaction potential curve

(b). γ now diffuses forward because of the barrier preventing the backward motion. Upon

moving forward by 80 degrees, it binds strongly to β (c). On Pi and ADP release, further

conformational changes at β-γ modifies the potential curve to (d), allowing the free diffusion

of γ forward by another 40 degrees. γ cannot go back due to the barrier on the left. After

completion of the 40 degree diffusion, γ again falls into the right well by forming strong

binding interactions with β. In the model the entire free energy of ATP hydrolysis is used

to alter the β-γ interaction.

The above model uses a tri-site reaction mechanism in which all three nucleotide binding

pockets are occupied while the motor takes a 120 degree step. Such a reaction scheme is not

completely inferable from the presently available experimental data. There are proposed bi-

site schemes which are compatible with the same experimental information [88]. Note that

our results are not affected by the exact nature of the scheme as long as any alternate scheme

also assumes that ATP binding produces 80 degree substep and product release produces

the remaining 40 degree substep.

In cargo observation experiments [73, 78], it is seen that the motor can drive cargoes

that are many times motor size. In this case, the cargo diffusion times without any driving

from motor is known to be much longer than the time scales at which the motor operates.

Using the fact that γ to which cargo is attached is elastic and is capable of storing energy

[100, 101], our model can be made compatible with this observation of motor driving cargoes

many times its size.

83

Page 90: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

5.3 Algorithmic Description of the Model

The model can be summarized for one step as follows:

(a)k1[ATP]

⇀↽k−1

(b) Substep1⇀↽

(c)k2⇀↽

k−2[P](d)

k3⇀↽

k−3[ADP](e) Substep2

⇀↽(f).

Here, k1, · · · , k3 are the kinetic constants associated with ATP binding and release of hy-

drolysis products, [ATP], [P], and [ADP] are the concentrations of ATP, Pi, and ADP,

respectively. In the following, β = 1/kBT , where kB is the Boltzmann constant and T the

temperature, ∆l ∼ 5 degrees, ks the stiffness of the link connecting motor and cargo (in-

cluding the stiffness of γ) which is modeled as a harmonic spring, Dm the motor diffusion

constant, Dc the cargo diffusion constant. x is the γ angle coordinate, y the cargo angle

coordinate, and t time. Here, the state of the motor is denoted by (x±), where x denotes

the nucleotide states (a-e in Fig. 5.1), and ± denotes the state of γ: + implies that γ is

movable diffusively and − implies that γ is fixed to the ring (not all the combinations make

sense; e.g., (a+) or (b−) does not exist). The elementary evolution of time step dt is taken

as 10−7s. In the following algorithm, the cargo position y is updated at every time step as

y(t + dt) = y(t) − βDcks(y − x)dt +√

2DcdtNy, (5.1)

where Ny is a Gaussian random number with zero mean and unit variation.

(A1) Start from state (a−): x = 0, y = 0 at time t = 0.

(A2) Let n be a uniform random number in [0, 1]. When the motor position is x < ∆l:

if n < k1dt, go to state (b+); if n > k1dt but n < k1dt + k−3dt, go to e− (binding of ADP

instead of ATP at x = 0) ; otherwise, the state remains unchanged.

(A3) For (b+) update the motor positions according to

x(t + dt) = max{0, x(t) − βDmks(x − y)dt +√

2DmdtNx(mod 120 deg)}, (5.2)

84

Page 91: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

which includes the barrier constraints. Nx is a Gaussian random number with zero mean

and unit variation.

(A4) When the motor coordinate is x > x2 −∆l (x2 = 80 degrees), the motor completes the

first substep (its position is updated to x2) and the state changes to (c−).

(A5) In state (c−) if n < k2dt, go to state (d−) (Pi unbinds from motor). Otherwise, the

state is not changed. x is not changed (x = x2).

(A6) In state (d−) if n < k−2[Pi]dt, go to (c−) (rebinding of Pi). x is not changed; if

n > k−2[Pi]dt but n < k3dt + k−2[Pi]dt (ADP release) go to state (e+).

(A7) For (e+) update the motor positions according to

x(t + dt) = max{x2, x(t) − βDmks(x − y)dt +√

2DmdtNx(mod 120 deg)}, (5.3)

which includes the barrier constraints. (A8) In state (e+) if the motor position is x < x2+∆l,

and if n < k−3[ADP]dt, go to state (d−) (ADP rebinding). x = x2; if n ≥ k−3[ADP]dt stay

in (e+) and update x according to (A7).

(A9) In state (e+) if x > xl − ∆l (xl = 120 degrees), then x = xl and the state becomes

(f−) = (a−); the 120 degree step is completed; if x < xl − ∆l, stay in (e+) and update x

according to (A7).

It is possible to have external load fL acting on the cargo. In this situation, in y position

update expressions of the above algorithm, the force experienced by the cargo ks(y − x) is

replaced with ks(y − x) − fL.

Table 5.1 gives the values of kinetic rate constants used in the simulation. The remaining

constants are chosen as follows.

∗ The stiffness of the link ks is 40 pNnm. The choice is consistent with the equilibrium

fluctuations given in [102].

∗ The γ rotational diffusion constant Dm is chosen to be ∼4000 rad2/s. The model results

85

Page 92: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Table 5.1: The adopted rate constants taken from Panke and Rumberg [103]

Process Rate constant

ATP binding 2.08 µM−1s−1

ATP unbinding 270 s−1

Phosphate ion release 2030×105 s−1

Phosphate ion binding 0.81 µM−1s−1

ADP unbinding 490 s−1

ADP binding 8.9 µM−1s−1

are not sensitive to the actual value chosen. Any value greater than 1000 rad2/s gives the

saturating motor hydrolysis rate in the absence of any cargo (In terms of the viscosity η of

water, the radius r = 0.9 nm of γ and its length L = 6 nm, the rotational diffusion constant

Dm of γ is given as kBT/4πηr2L ≃ 108 rad2/s, so the adopted value of 4000 rad2/s is nearly

105 times as large as this number. Effectively, we are assuming a viscosity of 105 times that

of water to account for the possible weak interactions between γ and the α-β ring. This

is an outrageously big friction and the actual friction may be smaller. However, as noted

above, a faster diffusion constant does not alter the result quantitatively. Therefore, the

choice here is rather technical: to have a reasonably large time step for simulation smaller

diffusion constants are convenient).

∗ For the experiment involving the observation of an actin filament attached to γ as a cargo,

its diffusion constant is estimated with the aid of Dc = kBT/[(4π/3)ηL3/log(L/2r)− 0.447],

where L is the length of the actin filament, η is the viscosity of water, and r = 5 nm is the

radius of the actin filament.

5.4 Comparison With Analytically Solvable Model

In Chapter 4, we developed an analytically solvable limit of the elastically coupled motor-

cargo system, when the size of the cargo is many times that of the motor so that there is

a clear time scale separation between motor and cargo fluctuations. It was shown that the

86

Page 93: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

coupled description of motor and cargo can be reduced to that of the cargo alone as being

driven by an effective motor driving torque. Using the simulation results in this chapter, it

is shown that the simplifying approximation of constant driving force in the analytic model

is quantitatively reliable, since the analytical results agree with simulation results as well as

with experimental results.

5.5 Comparison of Model Predictions with Empirical

Results

Figure 5.2 exhibits typical displacement records of cargo for varying ATP concentrations

obtained by the simulation of the model of ATP hydrolysis (Section 3). Most of the time

the motor is waiting for ATP binding or product release. Once ATP binds to the motor,

it makes quick diffusive forward steps during which there is very little cargo displacement.

Once motor reaches the next binding site and is waiting for ATP binding, the cargo is

dragged forward.

Figure 5.3 exhibits the variation of the cargo rotational rate with ATP concentration.

In this case, in order to explain the rotational rate at low ATP concentrations a different

ATP binding rate had to be used; the values used were k1 =27 µ M−1s−1 for [ATP] = 0.02

mM and 22 µ M−1s−1 for [ATP] < 10 µM. These values were taken from the analysis of

waiting time data given in [73]. With k1 = 2.08µ M−1s−1 at low ATP concentrations waiting

times occasionally exceeded 10 s. It is possible that in the actual single molecule experiment

such data were neglected, because they were indistinguishable from such long waiting times

caused by the formation of contacts between the actin rod and the surface or other such

irregularities. This probably is the reason for an effective higher binding constant.

Figure 5.4 exhibits the actin rotation rate as a function of the length of the actin rod. In

this case concentrations of ATP, ADP and Pi were monitored [73]. The simulation data and

the results from the analytical model (using the effective driving torque f of −40 pNnm for

87

Page 94: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

-200

100

400

700

1000

1300

0 5 10 15 20 25 30

Rota

tional A

ngle

(D

egre

es)

Time (s)

[ATP]=0.02 µM[ATP]=0.06µM[ATP]=0.2 µM[ATP]=0.6 µM

Figure 5.2: Displacement records for different ATP concentrations. [ADP] = 0 µM and [Pi]= 100 µM. The length of actin filament is 1 µm.

both 80 and 40 degree substeps. This value of f is compatible with the stiffness of the link

as well as the experimental observations) agree with the experimental data. Since our model

reproduces the rate-load relation, the efficiency of the motor is very high for our model.

Figure 5.5 exhibits the variation of actin rotation rate with increasing actin length at an

ATP concentration of 20 µM.

Figure 5.6 shows the variation of the actin rotation rate with increasing externally im-

posed torque for two different ATP concentrations. The analytical model predictions for a

driving torque of −35 pNnm is also shown. As of now there are no experimental data for

external torque. We expect the recent advances in designing magnetic traps makes such an

experiment feasible in the near future.

88

Page 95: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

0.01

0.1

1

10

0.01 0.1 1 10 100

Rota

tional R

ate

(r.

p.s

)

[ATP] (µM

exptl. data

Figure 5.3: Variation of the cargo rotational rate with ATP concentration. The length ofthe actin filament is 1 µm. [ADP] = 0 µM and [Pi] = 100 µM. Experimental data from [73]is also shown.

5.6 Discussion

We have introduced a simple model to explain the F1 motor with the following ingredients:

(1) the modulation of biding affinity between γ and β elements by the state of nucleotide

during the ATP hydrolysis process, (2) the mechanical motion of γ caused by thermal fluc-

tuation without any driving, and (3) experimentally inferred barriers checking motions in

wrong directions. It is shown that by incorporating the known biochemical data on the

motor, the model can quantitatively explain the available mechanochemical data without

adjustable parameters.

In most single molecule experiments that allow imposing external forces on the motor,

a large cargo (probe) is attached to the motor shaft, so the actual motor motion is never

directly observed. As seen in Chapters 3 and 4, in such situations the velocity of the motor

is dependent on the cargo size and stiffness of the link connecting them, irrespective of

the nature of the potential exerted by the motor. The indirectness of motor observation

89

Page 96: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

0.25

1

4

1 2 3

Rota

tional r

ate

(r.

p.s

)

Actin Length (µm)

[ATP]=2mM[ATP]=2mM (expt. data)

Figure 5.4: Variation of actin rotation rate with increasing actin length. [ATP] = 2 mM,[ADP] = 10 µM and [Pi] = 100 µM. Solid line shows the result from the analytic model withan effective driving torque of −40 pNnm for both substeps and experimental data from [73]is also shown.

may be the reason why many different models [24, 92, 93, 94] are compatible with available

quantitative experimental results.

It is possible to put a fluorescent probe directly on the shaft, but for the currently

available probes the space-time resolution of the obtained results is not good enough for

inferring any mechanism of the motor.

In the model, we assumed that the motion of γ is entirely due to thermal fluctuations.

It is possible to modify the minimal model to incorporate the driven motion of γ due to

some potential slope mimicking, e.g., the so-called power stroke. We found that no such

modification is needed to explain the mechanochemical data. So it is difficult to support the

current models with special driving mechanisms if the only criterion of their goodness is an

agreement with single molecule experiments.

If ATPase uses essentially simple diffusion to rotate, then substeps produces faster ro-

tation. Thus we can expect that ATPase without substeps should be considerably slower

90

Page 97: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

0.125

0.25

0.5

1

2

4

1 2 3

Rota

tional r

ate

(r.

p.s

)

Actin Length (µm)

[ATP]=0.02mM

Figure 5.5: Variation of actin rotation rate with increasing length. [ATP] = 20 µM, [ADP]= 1 µM and [Pi] = 100 µM. Solid lines are the results from the analytic model with aneffective driving torque of −40 pNnm for both substeps.

than F1-ATPase. V1 motor [104] behaves as such a motor which, similar to F1, uses ATP

hydrolysis to drive 120 degree rotation of a shaft. It is known that these motor steps do not

involve substeps [104]. On applying our model without any substeps to this motor, we find

that the motor velocity can be explained with reasonable rate constants of the order of that

of F1 (there are no known rate constant measurements for this motor).

Possible ways of proceeding from the current status that cannot reject the simple model

are:

(1) Experimentally demonstrate that there are no strong interactions between γ and β.

(2) Experimentally demonstrate that barrier heights cannot be modulated by nucleotide

changes alone.

(3) Attach small probes directly to γ, find its fluctuations in the absence of any cargo.

(4) Identify some mutations which make a motor that can rotate as fast as or faster than F1

but with no substeps.

(5) Using a highly sensitive magnetic trap, determine the load dependence of substep com-

91

Page 98: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

1e-04

0.001

0.01

0.1

1

10

0 10 20 30 40

Rota

tional r

ate

(r.

p.s

)

Load (pN-nm)

[ATP]=2mM[ATP]=0.02mM

Figure 5.6: The variation of motor rotational rate with increasing external torque. [ADP]= 1 µM and [Pi] = 100 µM and length of actin filament L = 1 µm. The solid lines are theanalytical model predictions for a driving torque of −35 pNnm for both substeps.

pletion times. This can say something about possible barriers.

(6) Measure the position dependence of the stalling forces.

Since (1) and (2) check the main ingredients of the minimal model, needless to say, they

directly test the simple model without any driving. However, as discussed already, structural

and other experimental data suggest that these demonstrations are not possible. (3) may

be an ideal experiment, but sufficient space-time resolution of such small probes is currently

impossible. (4) could strongly support active driving of γ by some potential formed by the

ring. (5) and (6) crucially depend on the progress of magnetic trap technology; at present,

we feel these experiments are not feasible, but they are the most direct single molecule

experiment conceivable.

92

Page 99: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Chapter 6

Summary and Open Questions

In this thesis, through careful analysis and modeling of available experimental data we have

tried to develop a phenomenological understanding of motor proteins. We have demonstrated

the following:

(1) In Chapter 2 we have developed a phenomenology of motor protein kinesin. We have

distilled a kinetic scheme from the available biochemical and structural information. This

has allowed us to develop a phenomenology compatible with the available data, ignoring the

flexibility of motor-cargo link. For the kinesin dimer, preferential detachment ( attachment)

of one head from (to) the track dependent on the conformation of the linker (neck linker

precisely) connecting the two heads, and the state of nucleotide at both heads is identified

as the biasing mechanism which introduces directionality. At the same time, we found that

biochemical data and mechanochemical data still cannot single out a unique motor force

generating scheme.

(2) In Chapter 3 by modeling the motor as simple tight and loosely coupled ratchets, we

have showed that the measured velocity of a large cargo attached to it elastically depends

on the stiffness of the link and the cargo diffusion constant. We have also showed that

the efficiency measures which are proposed as tools to identify the details of motor force

production also depends on the link stiffness. This suggests that inference of the details of

motor force production from cargo observations alone should be treated with caution. It is

possible to have qualitatively wrong inferences.

(3) In Chapter 4 we have proposed a scheme to interpret the single molecule mechanochem-

ical experimental data on motor proteins. The scheme is capable of treating motor diffu-

93

Page 100: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

sion and chemical reactions in a unified fashion. The model can quantitatively explain the

available mechanochemical data for F1ATPase, kinesin, and myosin V with just one fitting

parameter. A significant finding is that models without any detail of motor force production

mechanisms suffice to reproduce mechanochemical experimental results, if the cargo used in

these experiments are much larger than motor and the motor-cargo linkage is sufficiently

flexible.

(4) In Chapter 5 a simple model has been given to explain the motor F1-ATPase with

the following ingredients: (1) the modulation of binding between the rotary element γ and

the stator element β by the state of nucleotide during the ATP hydrolysis process, (2) the

mechanical motion of γ caused by thermal fluctuation without any driving, and (3) ex-

perimentally inferred barriers checking motions in wrong directions. The resultant model

incorporating the biochemical data has turned out to be able to describe the mechanochem-

ical data quantitatively without any adjustable parameters.

We have considered the mesoscopic models with the least ad hoc features compatible

with available experimental data. The model may be regarded as the null model, because

any experiment that can give a nontrivial statement about the motor mechanism must be

able to produce a result that cannot be explained by the model. We hope in the near future

experiments with better spatio-temporal resolutions will be able to reject the null model.

Then, minimal revisions of the null model to incorporate the new experimental results will

stimulate the next experimental progress. In addition to the explanation of mechanisms

of individual motors, such studies may tell us something about universal features of motor

proteins in general.

Another problem which can be approached at the mesoscopic level is understanding the

biasing mechanisms in motors which introduces directionality. In our studies, careful analysis

of available experimental data suggested that for kinesin dimer a possible biasing mechanism

is preferential detachment from track or attachment to track of a motor head depending on

the nucleotide state and the conformation of the neck linker. For F1-ATPase, it is preferential

94

Page 101: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

accessibility to binding sites on the stator (β) by the rotor (γ), depending on the state of

nucleotide on the stator. A common chemical description for both of these mechanisms is

the modulation of barrier height by the changes in nucleotide state of the motor or some

other conformational change. Physical descriptions of such biasing mechanisms in terms of

coarse-grained modeling of motor structures will be useful in the future.

Another unresolved question in the motors that we have considered is whether the opera-

tion is due to power stroke or due to rectification of thermal fluctuations. As we have showed,

currently available mechanochemical data is not sufficient to resolve the motor mechanism.

A clear resolution of this is possible, if we can estimate the free energy changes associated

with the conformational changes implicated in motor force production. Advances in com-

puting power and molecular dynamics simulation methods will be useful in resolving this

issue.

95

Page 102: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Appendix A

Solution of the Kinetic Model

In this Appendix we present the expressions for the velocity v, the randomness r, and the

run length L obtained by the general method summarized by Elston [59]. d denotes the step

length (= 8 nm). The 5-state model of Fig. 2.1 can be represented by the following chemical

kinetic scheme, where Mi represents a particular motor state.

M1

k1⇀↽

k−1

M2

k2⇀↽

k−2

M3

k3⇀↽

k−3

M4

k4⇀↽

k−4

M5

k5⇀↽

k−5

M1 (A.1)

The time evolution equations of such a scheme is given by

dFj

dt= LFj + L+Fj−1 + L−Fj+1, (A.2)

where index j denotes the spatial step number and Fk for k = j, j ± 1 is defined as

Fk =

(M1)k

(M2)k

(M3)k

(M4)k

(M5)k

,

(A.3)

96

Page 103: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

L =

−(k1 + k−5) k−1 0 0 0

k1 −(k−1 + k2) k−2 0 0

0 k2 −(k−2 + k3) k−3 0

0 0 k3 −(k−3 + k4) k−5

0 0 0 k4 −(k−4 + k5)

,

(A.4)

L+ =

0 0 0 0 k5

0 0 0 0 0

0 0 0 0 0

0 0 0 0 0

0 0 0 0 0

,

(A.5)

L− =

0 0 0 0 0

0 0 0 0 0

0 0 0 0 0

0 0 0 0 0

k−5 0 0 0

. (A.6)

Using the generating function defined as

P (z, t) =j=∞∑

j=−∞

zjF (j, t), (A.7)

the time evolution equations can be written as

dP

dt= LP + zL+P +

1

zL−P = A(z)P. (A.8)

97

Page 104: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

where

A(z) =

−(k1 + k−5) k−1 0 0 zk5

k1 −(k−1 + k2) k−2 0 0

0 k2 −(k−2 + k3) k−3 0

0 0 k3 −(k−3 + k4) k−5

1zk−5 0 0 k4 −(k−4 + k5)

. (A.9)

As shown by Elston, the velocity and effective diffusion constant can be obtained as,

v = dλ′

0(1), (A.10)

Deff =d2

2(λ′′

0(1) + λ′

0(1)). (A.11)

In the above expression, λ0 is the largest eigenvalue of matrix A(z), and the primes denotes

derivatives with respect to z evaluated at z = 1. An easy way to calculate these derivatives

is as follows. Write down the characteristic equation of A(z). Using the fact that the largest

eigenvalue is zero, we can write,

λ0(1 + α) = αλ′

0(1) +α2

2λ′′

0(1) + ...., (A.12)

Also, substitute z = 1 + α in the characteristic equation. Equating the coefficients of α to

be zero, the derivatives of λ can be easily determined.

For the five state models that we use, using the above technique, the expressions for

velocity and randomness can be obtained.

Let us define

98

Page 105: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

p1 ≡ k3k4(f)k5[ADP] + k3k4(f)k−1 + k3k5k−1 + k4(f)k5k−1 + k4(f)k5k−2[P]

+k4(f)k−1k−2[P] + k5k−1k−2[P] + k5k−1k−3(f) + k5k−2[P]k−3(f)

+k−1k−2[P]k−3(f) + k3k−1k−4 + k−1k−2[P]k−4 + k−1k−3(f)k−4

+k−2[P]k−3(f)k−4 + k1[ATP](k4(f)k5 + k4k−2[P ] + k5k−2[P]

+k5k−3(f) + k−2[P]k−3(f) + (k−2[P] + k−3(f)k−4 + k3(k4(f) + k5 + k−4)

+k2(k3 + k4(f) + k5 + k−3(f) + k−4)) + (k−1k−2[P] + k4(k−1 + k−2[P])

+k−1k−3(f) + k−2[P]k−3(f) + (k−1 + k−2[P] + k−3(f))k−4

+k3(k4(f) + k−1 + k−4))k−5[ADP] + k2(k−3(k5 + k−4) + (k−3(f) + k−4)k−5[ADP]

+k4(f)(k5 + k−5[ADP]) + k3(k4(f) + k5 + k−4 + k−5[ADP])), (A.13)

p2 ≡ k−1(k3k4(f)k5 + k4(f)k5k−2[P] + k−2[P]k−3(f)(k5 + k4)) + k1[ATP](k3k4(f)k5

+k5k−2[P](k4(f) + k−3(f)) + k−2[P]k−3(f)k−4 + k2(k5(k4(f) + k−3(f))

+k−3(f)k−4 + k3(k4(f) + k5 + k−4))) + (k−1k−2[P](k4(f) + k−3(f))

+(k−2[P]k−3(f) + k−1(k−2[P] + k−3(f)))k−4 + k3k−1(k4 + k−4))k−5[ADP]

+k2(k3k4(f)k5 + k−3(f)k−4k−5[ADP] + k3(k4(f) + k−4)k−5[ADP]), (A.14)

p3 ≡ k1[ATP]k2k3k4(f)k5, (A.15)

p4 ≡ k−1k−2[P]k−3(f)k−4k−5[ADP]. (A.16)

In terms of these quantities

v0 = d(p3 − p4)/p2, (A.17)

r0 = 1 +2

(p3 − p4)

(

p4 −p1(p3 − p4)

2

p22

)

. (A.18)

99

Page 106: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

If the probability of back steps is taken into account the formulae are modified as,

v = v0.(Pf − Pr), (A.19)

r =1

1 − 2Pr

+ (r0 − 1)(1 − 2.Pr), (A.20)

where Pr and Pf are the probability of forward and backward steps, respectively. The steady

probabilities PX for the intermediate state X read:

P(DP, E) = (k2k3k4(f)k5 + k−1(k3k4(f)k5 + k5k−2[P](k4(f) + k−3)

+k−2k−3(f)k−4))/p2; (A.21)

P(DP, T) = (k1[ATP](k3k4(f)k5 + k5k−2[P](k4(f) + k−3(f)) + k−2[P]k−3(f)k−4)

+k−2[P]k−3(f)k−4k−5[ADP])/p2; (A.22)

P(D,T) = (k1[ATP]k2(k4(f)k5 + k−3(f)(k5 + k−4))

+(k2 + k−1)k−3(f)k4k−5[ADP])/p2; (A.23)

P(T/DP, D)1 = (k1[ATP]k2k3(k5 + k−4)

+(k2k3 + k−1(k3 + k−2))k−4k−5[ADP])/p2; (A.24)

P(T/DP, D)2 = (k1[ATP]k2k3k4(f) + (k2k3k4(f) + k4(f)k−1(k3 + k−2)

+k−1k−2k−3(f))k−5[ADP])/p2. (A.25)

If the detachment rate from state X is denoted by kX , the net detachment rate during a

mechanochemical cycle is given by

kdet = k(DP, E)P(DP, E) + k(DP, T)P(DP, T) + k(D, T)P(D, T)

+k(T/DP, D)1P(T/DP, D)1 + k(T/DP, D)2P(T/DP, D)2, (A.26)

100

Page 107: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

and the run length is given by

L = v/kdet. (A.27)

101

Page 108: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Appendix B

Stochastic Energetics

Stochastic energetics due to Sekimoto [72] aims at developing an energetic picture of Langevin

dynamics. It can be used to determine the energetic quantities of motor-cargo system defined

as follows. In a particular state (say, 1) the motor is described by the Langevin equation

−∂v1/∂x + (−γmdx/dt + ξ(t)) = 0, (B.1)

where ξ(t) is the Gaussian noise with zero mean and 〈ξ(t)ξ(t′)〉 = (2/γm)δ(t− t′). The work

done by the heat bath on the motor is [72] given by

D = −∫ tf

ti(−γm

dx

dt+ ξ(t)) ◦ dx(t) =

∫ tf

ti−∂V1(x, y)/∂x ◦ dx(t), (B.2)

where ◦ denotes Ito’s circle product. Taking the average of D with the probability of state

obtained by solving the Fokker-Planck equation and adding the contributions from states 1

and 2, we get the motor heat exchange rate with the environment as,

〈Dm〉 =∫ 1

0dx∫

−∞

dy

(

−∂V1(x, y)

∂xJ1m(x, y) − ∂V2(x, y)

∂xJ2m(x, y)

)

. (B.3)

Similarly, the heat exchange rate of the cargo with the environment is

〈Dc〉 =∫ 1

0dx∫

−∞

dy

(

−∂V1(x, y)

∂yJ1c(x, y) − ∂V2(x, y)

∂yJ2c(x, y)

)

. (B.4)

102

Page 109: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

The cargo output rate is the same as the work done by the cargo against the load,

〈W 〉 = −∫ 1

0dx∫

−∞

dyF (J1c(x, y) + J2c(x, y)) . (B.5)

The efficiency of the system is defined as the ratio of the cargo output rate and the motor

input rate,

ηeff =〈W 〉〈Rm〉

. (B.6)

At steady state it can be shown that

〈Rm〉 = 〈Dm〉 + 〈Dc〉 + 〈W 〉, (B.7)

Using the numerical solution of Fokker-Planck equations and the above definitions, we get

0

2

4

6

8

0.75 1.25 1.75 2.25

Inp

ut/

Dis

sip

atio

n R

ate

s

log10(ks)

<Rm><D.

m><D.

c>

Figure B.1: Variation of the energetic parameters-the motor energy input rate (〈Rm〉), themotor dissipation rate (〈Dm〉) and the cargo dissipation rate (〈Dc〉) – with the stiffness ofthe link. The potential parameters (for notations see Fig. 3.1) are, u = 20, a = 0.3, Dc =0.1, γ = 1.6 and external load f = 0.

the energetics. Fig. B.1 shows the variation of energetic parameters with the link stiffness

103

Page 110: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

in the absence of any external force. The behavior of energetic quantities can be understood

intuitively if we consider the limit u ≫ 1 as in the velocity explanation. Transition from

state 2 to 1 hardly supplies any energy, because motor spends most of its time close to the

potential minima in state 2. Hence, the net energy input rate can be approximated as

〈Rm〉 = (γ/2)∫

−∞

dx∫

−∞

dyφ2(x)P (x, y, 1/γ), (B.8)

where P (x, y, 1/γ) is the probability distribution of being in state 1. Since∫

−∞dyP (x, y, 1/γ)

0

0.001

0.002

0.003

0.004

0.75 1.25 1.75 2.25

Eff

icie

ncy

log10(ks)

γ=1.6γ=160

Figure B.2: Variation of the efficiency of the motor with the stiffness of the link for twodifferent transition rates. The potential parameters (for notations see Fig. 3.1)are, u = 20,a = 0.3, Dc = 0.1, and external load f = -0.4.

is a Gaussian distribution peaked near x = 0, the greater the spatial spread of this Gaus-

sian, the greater the input energy rate, since this allows motor to make transitions to high

energy regions of state 2 when the motor state changes from 1 to 2. This is the case for

very low stiffness of the link. As the stiffness of the link is increased, the width of the motor

probability distribution narrows and the motor input rate decreases, since more and more of

the motor transitions will be to the low energy regions of state 2. The same reasoning also

104

Page 111: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

explains the decrease of motor dissipation with increased stiffness. For very low stiffness,

only a small fraction of motor input is transmitted to the cargo before being dissipated at

the motor. As the stiffness is increased, more and more of the motor input is transmitted to

the cargo and dissipated there.

Figure B.2 shows the efficiency of the motor (B.6) as a function of the link stiffness. For

moderate transition rate (γ), there is an optimal stiffness of the link which maximizes the

efficiency. The velocity variation for this transition rate also shows a peak at this stiffness

and the same explanation for velocity peak can explain the efficiency peak. At this stiffness,

due to increased velocity, motor is able to transmit more power to the cargo.

105

Page 112: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

References

[1] J. Howard, Mechanics of Motor proteins and Cycloskeleton, Sinauer AssociatesInc., Sunderland MA. (2001).

[2] M. Schliwa and G. Woehlke, Nature 422, 759 (2003)

[3] R. D. Vale, Cell 112, 467 (2003).

[4] K. M. Ruppel, J. Spudich, Annu. Rev. Cell. Biol. 12, 543 (1996).

[5] M. P. Koonce and M. Samso, Trends in Cell Biol. 14, 612 (2004).

[6] R. A. Mooney and R. Landick, Cell 98, 687 (1999).

[7] P. D. Boyer, Annu. Rev. Biochem. 66, 717 (1997).

[8] D. E. Smith, S. J. Tans., S. B. Smith, S. Grimes, D. L. Anderson, and C. Busta-mante, Nature 413, 748 (2001).

[9] S. P. Gilbert and A. T. Mackey, Methods 22, 337 (2000).

[10] M. A. Geeves and K. C. Holmes, Annu. Rev. Biochem. 68, 687 (1999).

[11] E. Mandelkow and A. Hoenger, Curr. Opin. Cell Biol. 11, 34 (1999).

[12] F. S. Kozliensky, S. Sack, A. Marx, M. Thormahlen, E. Schonbrunn, V. Biou, A.Thompson, E.-M. Mandelkow, and E. Mandelkow, Cell 91, 985 (1997).

[13] F. J. Kull and S. A. Endow, Trends in Biochem. Sci. 29, 103 (2004).

[14] R. L. Cross, Nature 427, 407 (2004).

[15] M. P. Sheetz and J. A. Spudich, Nature 303, 31 (1983).

[16] K. Visscher, M. J. Schnitzer and S. M. Block, Nature 400, 184 (1999).

[17] A. E.-M. Clemen, M. Vilfan, J. Jaud, J. Zhang, M. Barmann, and M. Rief,Biophys. J.88, 4402 (2005).

[18] S. Weiss, Science 283, 1689 (1999).

106

Page 113: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

[19] A. Yildiz, J. N. Forkey, S. A. McKinney, T. Ha, Y. E. Goldman, and P. R. Selvin,Science 300, 2061 (2003).

[20] A. F. Huxeley, Prog. Biophys. 7, 255 (1957).

[21] R. A. Bockmann and H. Grubmuller, Nat. Struct. Biol. 9, 198 (2002).

[22] M. Dittrich, S. Hayashi and K. Schulten, Biophys. J. 87, 2954 (2004).

[23] S. A. Endow, BioEssays 25, 1212 (2003).

[24] H. Wang and G. Oster, Nature 396, 279 (1998).

[25] R. D. Vale and R. A. Milligan, Science 288, 88 (2000).

[26] M. E. Fisher and A. B. Kolomeisky, Proc. Natl. Acad. Sci. (USA) 98, 7748(2001).

[27] C. V.Sindelar, M. J. Budny, S. Rice , N. Naber , R. Fletterick , R. Cooke, NatStruct Biol. 9, 844 (2002).

[28] P. Reimann, Phys. Rep. 361, 57 (2002).

[29] D. Keller and C. Bustamante, Biophys. J. 78, 541 (2000).

[30] G. Lattanzi and A. Maritan, J. Chem. Phys. 117, 10339 (2002).

[31] Molinger, A., A. J. Fisher, and R. J. Baskin, J. theor. Biol. 211, 143 (2001).

[32] R. F. Fox and M. H. Choi, Phys. Rev. E. 63, 051901 (2001).

[33] T. Duke and S. Leibler, Biophys. J. 71, 1235 (1996).

[34] C. M. Coppin, D. W. Pierce, L. Hsu, and R. D. Vale, Proc. Natl. Acad. Sci.(USA) 94, 8539 (1997).

[35] M. J. Schnitzer, K. Visscher, and S. M. Block, Nature Cell Biol. 2, 718 (2000).

[36] M. Nishiyama, E. Muto, Y. Inoue, T. Yanagida, and H. Higuchi. Nature CellBiol. 2, 425 (2002).

[37] N. J. Carter and R. C. Cross, Nature 435, 308 (2005).

[38] W. Hua, E. C. Young, M. L. Fleming, and J. Gelles, Nature 388, 390 (1997).

[39] M. J. Schnitzer and S. M. Block, Nature 388, 386 (1997).

[40] D. D. Hackney, Proc. Natl. Acad. Sci. (USA) 91, 6865 (1994).

[41] W. O. Hancock and J. Howard, Proc. Natl. Acad. Sci. (USA) 96, 13147 (1999).

[42] A. Yilditz, M. Tomishige, R. D. Vale, and P. R. Selvin, Science 303, 676 (2004).

107

Page 114: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

[43] C. L. Asbury, A. N. Fehr, and S. M. Block, Science 302, 2130 (2003).

[44] H. E. Higuchi, Muto, Y. Inoue, and T. Yanagida, Proc. Natl. Acad. Sci. (USA)94, 4395 (1997).

[45] S. Uemura, K. Kawaguchi, J. Yajima, M. Edamatsu, Y. Y. Toyoshima, and S.Ishiwata, Proc. Natl. Acad. Sci. (USA) 99, 5977 (2002).

[46] I. M.-T. C. Crevel, I. M.-T. C., A. Lockhart, and R. A. Cross, J. Mol. Biol. 257,66 (1996).

[47] S. P. Gilbert, M. L. Moyer, and K. A. Johnson, Biochemistry 37, 792 (1998).

[48] M. L. Moyer, S. P. Gilbert, and K. A. Johnson. Biochemistry 37, 800 (1998).

[49] M. M. Farrel, A. T. Mackey, L. M. Klumpp, and S. P. Gilbert, J. Biol. Chem.277, 17079 (2002).

[50] S. S. Rosenfeld, G. M. Jefferson, and P. H. King. J. Biol. Chem. 276, 40167(2001).

[51] S. Rice, A. W. Lin, D. Safer, C. L. Har, N. Naber, B. O. Carragher, S. M. Cain,E. Pechatnikova, M. Wilson-Kubalek, M. Whittaker, E. Pate, R. Cooke, E. W.Taylor, R. Milligan, and R. D. Vale, Nature 402, 778 (1999).

[52] M. Kikkawa, E. P. Sablin, Y. Okada, R. J. Fetterick, and N. Hirokawa, Nature411, 439 (2001).

[53] R. D. Vale, R. Case, E. Sablin, C. Hart, and R. Fletterick. Phil. Trans. R. Soc.Lond. B Biol. Sci. 355, 449 (2000).

[54] M. Tomishige, and R. D. Vale, J. Cell Biol. 151, 1081 (2000).

[55] M. Tomishige, U. M. Wiedemann, and R. D. Vale, Biophys. J. (Annual MeetingAbstracts) 2002:509a (2002).

[56] Y.-Z. Ma, and E. W. Taylor, J. Biological Chem. 272, 724 (1997).

[57] K. Kawaguchi, and S. Ishiwata, Science 291, 667 (2001).

[58] A. M. Hoenger, Thormahlen, R. Diaz-Avalos, M. Doerhoefer, K. N. Goldie, J.Muller, and E. Mandelkow, J. Mol. Biol. 297, 1087 (2000).

[59] T. Elston, J. Math. Biol. 41, 189 (2000).

[60] Y. Inoue, A. H. Iwane, T. Miyai, E. Muto, and T. Yanagida. Biophys. J. 81, 2838(2001).

[61] H. Sosa, E. J. G. Peterman, W. E. Moerner, and L. S. B. Goldstein, NatureStructural Biol. 8, 540 (2001).

108

Page 115: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

[62] J. Yajima, A. C. Alonso, R. A. Cross, and Y. Y. Toyoshima, Current Biol. 12,301 (2002).

[63] Y.-H. Song, A. Marx, J. Muller, G. Woehlke, M. Schliwa, A. Krebs, A. Hoenger,and E. Mandelkow, EMBO J. 20, 6213 (2001).

[64] C. D. Bustamante, Keller, and G. Oster. Acc. Chem. Res. 34, 412 (2001).

[65] C. S. Peskin and G. Oster. Biophys. J. 68, 202s (1995).

[66] G. Cappello, M. Badoul, A. Ott, J. Prost, and L. Busoni, Phys. Rev. E 68,021907 (2003).

[67] H. Noji, R. Yasuda, M. Yoshida and K. Kinosita, Nature 386, 299 (1997).

[68] T. Elston and C. Peskin, SIAM J. Appl. Math. 60, 842 (2000).

[69] T. Elston, D. You, and C. Peskin, SIAM J. Appl. Math. 61, 776 (2000).

[70] F. Julicher, A. Adjari, and J. Prost, Rev. Mod. Phys. 69, 1269 (1997).

[71] R. D. Astumian and M. Bier, Phys. Rev. Lett. 72, 1766 (1994).

[72] K. Sekimoto, J. Phys. Soc. Jpn. 66, 1234 (1997).

[73] R. Yasuda, H. Noji, K. Kinosita Jr., F. Motojima, and M. Yoshida, Cell 93, 1117(1998).

[74] C. Veigel, J. E. Molloy, S. Schmitz, and J. Kendrick-Jones, Nat. Cell. Biol. 5,980 (2003).

[75] A. J. Hunt, F. Gittes, and J. Howard, Biophys. J. 67, 766 (1994).

[76] K. Kinosita Jr., R. Yasuda, H. Noji, and K. Adachi, Phil. Trans. R. Soc. London355, 473 (2000).

[77] J. Xing, J.-C. Liao, and G. Oster, Proc. Natl. Acad. Sci. (USA) 102, 16539(2005).

[78] R. Yasuda, H. Noji, M. Yoshida, K. Kinosita Jr., and H. R. Itoh H. R., Nature410, 898 (2001).

[79] O. Panke and B. Rumberg, Biochim. Biophys. Acta. 1412, 118 (1999).

[80] R. A. Cross R. A. Trends. Biochem. 29, 301 (2004).

[81] S. M. Bloch, C. L. Asbury, J. W. Shaevitz, and M. L. Lang, Proc. Natl. Acad.Sci. (USA) 100, 2351 (2003).

[82] A. E.-M. Clemen, M. Vilfan, J. Jaud, J. Zhang, M. Barmann, and M. Rief,Biophys. J. 88, 4402 (2005).

109

Page 116: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

[83] M. Rief, R. S. Rock, A. D. Mehta, M. S. Mooseker, R. E. Cheney, and J. A.Spudich, Proc. Natl. Acad. Sci. (USA) 97, 9482 (2000).

[84] A. Vilfan, Biophys. J. 88, 3792 (2005).

[85] S. S. Rosenfeld and H. L. Sweeney, J. Biol. Chem. 279, 40100 (2004).

[86] E. M. De La Cruz, A. L. Wellsdagger, S. S. Rosenfeld, M. Ostapdagger, and H.L. Sweeney, Proc. Natl. Acad. Sci. (USA) 96, 13726 (1999).

[87] Y. R. Chemla, K. Aathavan, J. Michaelis, S. Grimes, P. J. Jardine, D. L. Ander-son, and C. Bustamante, Cell 122, 683 (2005).

[88] K. Kinosita Jr., K. Adachi, and H. Itoh, Annu. Rev. Biophys. Biomol. Struct.33, 245 (2004).

[89] J. P. Abrahams, A. G. W. Leslie, R. Lutter, and J. E. Walker, Nature 370, 621(1994).

[90] R. D. Astumian, Phys. Rev. Lett. 91, 118102 (2003).

[91] S. Saffarian, H. Qian, I. Collier, E. Elson, and G. Goldberg, Phys. Rev. E 73041909 (2006).

[92] Y. Q. Gao, W. Yang, and M. Karplus, Cell 123, 195 (2005).

[93] Y. Q. Gao, W. Yang, R. A. Marcus, and M. Karplus, Proc. Natal. Acad. Sci.(USA) 100, 11339 (2003)

[94] M. S. Liu, B. D. Todd, and R. J. Sadus, J. Chem. Phys. 118, 9890 (2003).

[95] R. Aggeler, M. A. Haughton, and R. A. Capaldi, J. Biol. Chem. 268, 20831(1993).

[96] R. Aggeler, M. A. Haughton, and R. A. Capaldi, J. Biol. Chem. 270, 9185 (1995).

[97] R. A. Brockmann and H. Grubmuller, Biophys. J. 85, 1482 (2003).

[98] T. K. Nishizaka, H. Oiwa, S. Noji, E. Kimura, M. Muneyuki, M. Yoshida, andK.Kinosita Jr., Nat. Struct. Biol. 11, 142 (2004).

[99] R. L. Cross, Biochim. Biophys. Acta. 1458, 270 (2000).

[100] W. Junge, O. Panke, D. A. Cherepanov, K. Gumbiowski, M. Muller, S. Engel-brecht, FEBS Lett. 504, 152 (2001).

[101] D. A. Cherepanov, A. Y. Mulkidjanian, and W. Junge, FEBS Lett. 449, 1 (1999).

[102] Y. Hirono-Hara, K. Ishizuka, K. Kinosita Jr., M. Yoshida, and H. Noji, Proc.Natl. Acad. Sci. (USA) 102, 4288 (2005).

110

Page 117: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

[103] O. Panke and B. Rumberg, Biochim. Biophys. Acta. 1412, 118 (1999).

[104] K. Shimabukuro, R. Yasuda, E. Muneyuki, K. Y. Hara, K. Kinosita, Jr., and M.Yoshida, Proc. Natl. Acad. Sci. (USA) 100, 14731 (2003).

111

Page 118: c 2006 by Prasanth Sankar. All rights reserved.research.physics.illinois.edu/Publications/theses/copies/Sankar_Prasanth.pdfPRASANTH SANKAR M. S., University of Illinois at Urbana-Champaign,

Author’s Biography

Prasanth Sankar was born on Nov. 1, 1976 in Kerala, India. He received MSc. degree in

Physics in 1998 from the Indian Institute of Technology, Kharagpur. He received his M.

S. in Physics, from the University of Illinois at Urbana-Champaign in 2000. While at the

University of Illinois he was supported by teaching and research assistantships.

112