Top Banner
Chemotaxis of Escherichia coli to controlled gradients of attractants: Experiments and Mathematical modeling Submitted in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY in CHEMICAL ENGINEERING by VUPPULA RAJITHA (Roll No. 04402602) Supervisor: Prof. Mahesh S. Tirumkudulu Co-Supervisor: Prof. K. V. Venkatesh Department of Chemical Engineering Indian Institute of Technology Bombay 2009
113

Chemotaxis of Escherichia Coli to Controlled Gradients of Attractants(1)

Nov 10, 2015

Download

Documents

alexa2108

quimiotaxis en E. coli, control de quimiotrayente
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
  • Chemotaxis of Escherichia coli to controlled

    gradients of attractants: Experiments and

    Mathematical modeling

    Submitted in partial fulfillment of the requirements

    for the degree of

    DOCTOR OF PHILOSOPHY

    in

    CHEMICAL ENGINEERING

    by

    VUPPULA RAJITHA

    (Roll No. 04402602)

    Supervisor: Prof. Mahesh S. Tirumkudulu

    Co-Supervisor: Prof. K. V. Venkatesh

    Department of Chemical Engineering

    Indian Institute of Technology Bombay

    2009

  • Declaration

    I declare that this written submission represents my ideas in my own words

    and where others ideas or words have been included, I have adequately cited and

    referenced the original sources. I also declare that I have adhered to all principles

    of academic honesty and integrity and have not misrepresented or fabricated or

    falsified any idea/data/fact/source in my submission. I understand that any vio-

    lation of the above will be cause for disciplinary action by the Institute and can

    also evoke penal action from the sources which have thus not been properly cited

    or from whom proper permission has not been taken when needed.

    Vuppula Rajitha

    (Roll No. 04402602)

    i

  • Abstract

    Chemotaxis is a phenomenon in which a microorganism or multi-cellular organisms

    are able to direct their movements in the presence of certain chemicals in their envi-

    ronment. This could be either towards favorable chemicals called chemoattractants

    or away from unfavorable chemicals called chemorepellents. The understanding

    of this phenomena is important in many biological processes including immune

    response, embryo-genesis, wound healing, bio-film formation and bio-remediation

    of subsurface contaminants. The objective of this thesis is to study the response

    of microorganisms in the presence of controlled gradients of chemoattractants us-

    ing a micro-capillary. This would involve developing mathematical models that

    describe the chemotaxis pathway in a chemotactic cell and to relate the kinetics

    to the motion of cell in the presence of chemoattractants. The predictions of the

    model will be compared with experiments where the motion of cell in the presence

    of chemoattractants will be tracked. The previous studies have reported average

    drift velocities for a given gradient and do not measure drift velocities as a func-

    tion of time and space. To address this issue, a novel experimental technique was

    developed to quantify the motion of E. coli cells to varying concentrations and

    gradients of chemoattractant so as to capture the spatial and temporal variation of

    the drift velocity. The statistics of single cell motions such as cell velocity, tumbling

    frequency, rotational diffusion, drift velocity and translational diffusivity are ob-

    tained independently. An existing two state receptor model of Barkai and Leibler

    (1997) was used for the intracellular pathway and into this the extra-cellular in-

    fluence such as ligand concentration and Brownian motion was incorporated to

    predict the response for the experimental conditions. The model predicted the ex-

    ii

  • perimentally observed increase in drift velocity with gradient and the subsequent

    adaptation. Our study revealed that the rotational diffusivity induced by the ex-

    tracellular environment is crucial in determining the drift velocity of E. coli. The

    model predictions matched with experimental observations only when the response

    of the intracellular pathway was highly ultra-sensitive to overcome the extracel-

    lular randomness. The parametric sensitivity of the pathway indicated that the

    dissociation constant for the binding of the ligand and the rate constants of the

    methylation/demethylation of the receptor are key to predict the performance of

    the chemotactic behavior. The study demonstrates the key role of oxygen in the

    chemotaxis response and that the response to a ligand cannot be analyzed in isola-

    tion to effects of oxygen. Further, the study noted that the chemotactic response

    also depends on the nature of the chemical used as a chemoattractant which in

    turn decides the energy status of the cell.

    iii

  • Contents

    Abstract ii

    List of Figures xii

    List of Tables xiii

    1 Introduction 1

    1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

    1.2 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

    1.3 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

    1.4 Organization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

    2 Materials and Methods 19

    2.1 Experimental Protocols . . . . . . . . . . . . . . . . . . . . . . . . . 19

    2.2 Modeling Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

    3 Mathematical modeling and Experimental validation of chemo-

    taxis under controlled gradients of methyl-aspartate in Escherichia

    coli 36

    3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

    3.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

    3.3 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

    3.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

    4 Chemotaxis of Escherichia coli to L-serine 56

    4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

    iv

  • 4.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

    4.3 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

    4.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

    5 Chemotaxis of Escherichia coli in the presence of glucose via phos-

    photransferase (PTS) pathway 76

    5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

    5.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

    5.3 Discussions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

    5.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

    6 Conclusions and Future study 88

    References 96

    v

  • List of Figures

    1.1 Schematic diagram of E. coli with multiple flagella. It is typically

    rod-shaped and is about 2 m long and 1 m in diameter. (Eisen-

    bach, 1994) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

    1.2 Swimming behavior of E. coli or Salmonella in the absence of

    chemotactant. It executes a random walk composed of runs and

    tumbles with zero drift. . . . . . . . . . . . . . . . . . . . . . . . . 4

    1.3 Swimming behavior of E. coli or Salmonella in the presence of

    chemotactant. It executes a biased random walk composed of runs

    and tumbles with significant drift. . . . . . . . . . . . . . . . . . . . 5

    1.4 Experimental data using fluorescence correlation spectroscopy: Re-

    sponse of individual motors as a function of CheY-P concentration.

    (Cluzel et al., 2000) Each data point describes a simultaneous mea-

    surement of the motor bias and the CheY-P concentration in an

    individual bacterium. CW bias was computed by analyzing video

    recordings for at least 1 min. . . . . . . . . . . . . . . . . . . . . . . 6

    1.5 Chemotaxis signaling pathways in E. coli (A) in the absence of

    chemoattractant, and (B) in the presence of chemoattractant. The

    dotted line represents the reactions with reduced rate in response to

    ligand binding to the receptor. It can be noted that on adaptation

    the pathway returns to steady state as shown in (A). . . . . . . . . 7

    2.1 Linear calibration curve: Fluorescence intensity with varying con-

    centration of 2-NBDG. . . . . . . . . . . . . . . . . . . . . . . . . . 22

    vi

  • 2.2 Micro-capillary experimental setup for establishing glucose gradi-

    ents. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

    2.3 Measured fluorescence intensity in the capillary for 10 min (solid

    line) and 540 min (dashed line) after the start of the experiment.

    It can be noted that the gradient was stable for more than 9 hours. 23

    2.4 (A) Agar plate containing 0.3 % agar and motility buffer (absence

    of glucose) with 107 cells introduced at the center of the plate.The image was taken after 6 hours of incubation. It can be noted

    that no ring was formed. (B) Agar plate containing 0.3 % agar,

    motility buffer and glucose (5 mM), with 107 cells introducedat the center of the plate. The image was taken after 6 hours of

    incubation. Clearly, the cells have migrated forming a distinct ring. 23

    2.5 Micro-capillary experimental setup for establishing MeAsp/serine

    gradients. x = 0 is located at the end of the pellet with a ligand

    concentration, L0. The experimental values of L0 and the estab-

    lished gradients (G) are given in Table 2.1 and these conditions

    were used in the model simulations. . . . . . . . . . . . . . . . . . . 26

    2.6 (A) A magnified view of the micro-capillary set-up (B) Stable linear

    gradients; Measured concentration profiles for two gradients: t =

    1.5 min (black line) and t = 5 min (blue line) forG = 0.016 M/m,

    t = 1.5 min (red line), t = 5 min (green line) and t =30 min (pink

    line) for G = 0.16 M/m respectively. . . . . . . . . . . . . . . . . 27

    3.1 Mean square displacements (MSD standard error) about themean value (from Gaussian fit) as a function of time. MSD was

    obtained using data from all experiments including those with and

    without gradients. The solid line represents the linearity obtained

    for a diffusive regime. The diffusive regime occurs beyond 3.8 sand is shown by a dotted arrow. . . . . . . . . . . . . . . . . . . . 39

    vii

  • 3.2 Measured drift velocity (u0 standard error) as a function of dis-tance in the absence of aspartate using normal buffer (), 100 Maspartate in normal buffer (N), and 10000 M aspartate in normal

    buffer (). The solid line represents the average trend for the three

    experiments. The empty circle () represents the drift velocity whenthe buffer is saturated with oxygen in the absence of aspartate. . . . 40

    3.3 Average run speeds ( standard error) as a function of aspartategradients at: 500 m (+), 1000 m (), and 1500 m (). Thesolid line represents the average value for the entire data set. The

    arrow indicates the average run speed of 18 m/s obtained in the

    absence of MeAsp. . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

    3.4 Drift velocity (uasp standard error) as a function of distance forvarious gradients of aspartate (G) and initial ligand concentration

    (L0) (A) G = 0.016 M/m and L0 = 16 M; (B) G = 0.08 M/m

    and L0 = 80 M; (C) G = 0.16 M/m and L0 = 160 M; (D)

    G = 1.6 M/m and L0 = 1600 M. represents data from ex-periments and solid line represents model prediction. . . . . . . . . 42

    3.5 (A) Drift velocity obtained through simulation as a function of as-

    partate gradients at four different spatial locations: dotted line in

    gray at 100 m, dotted line in black at 500 m, dashed line at

    1500 m and solid line at 3000 m, respectively (B) Normalized

    CW bias with time for four different gradients: solid line for G =

    0.016 M/m and L0 = 16 M, dashed line for G = 0.08 M/m

    and L0 = 80 M, dotted line in black for G = 0.16 M/m and L0

    = 160 M and dotted line in gray for G = 1.6 M/m and L0 =

    1600 M (C) Normalized active receptor concentration with time

    for four different gradients: solid line for G = 0.016 M/m and

    L0 = 16 M, dashed line for G = 0.08 M/m and L0 = 80 M,

    dotted line in black for G = 0.16 M/m and L0 = 160 M and

    dotted line in gray for G = 1.6 M/m and L0 = 1600 M. . . . . . 44

    viii

  • 3.6 (A) Drift velocity obtained through simulation as a function of lig-

    and dissociation constant for a gradient of 0.06 M/m at four

    different spatial locations: dotted line in gray at 100 m, dotted

    line in black at 500 m, dashed line at 1500 m and solid line

    at 3000 m, respectively (B) Drift velocity as a function of ratio

    of methylation and demethylation rate constants for a gradient of

    0.06 M/m at four different spatial locations: dotted line in gray

    at 100 m, dotted line in black at 500 m, dashed line at 1500 m

    and solid line at 3000 m, respectively (C) Drift velocity as a func-

    tion of gradient with varying Hill coefficients (n) at x = 500 m

    (D) Variation of Hill coefficient (n) with rotational diffusivity (Dr)

    to obtain a drift velocity of 1 m/s at x = 500 m, for a gradient

    of 0.06 M/m. The arrow indicates the experimentally observed

    rotational diffusivity, Dr = 0.32 rad2/s, corresponding to n = 50. . . 46

    3.7 (A) Measured drift velocity ( standard error) as a function of thespatial gradient of the logarithmic aspartate concentration: G =

    0.016 M/m (), G = 0.08 M/m (), G = 0.16 M/m (N)and G = 1.6 M/m () (B) Average value of the drift velocity

    ( standard error) at each d(lnL)/dx with location. The solid linerepresents a linear curve fit while the dashed line represents the

    trend observed by (Kalinin et al., 2009) . . . . . . . . . . . . . . . . 49

    ix

  • 4.1 (A) Mean square displacements about the mean value as a function

    of time for varying serine concentrations: 0 M (), 250 M (),1000 M (N), 5000 M (), 24000 M () and 240,000 M() (B) The translational diffusivity for various serine concentra-tions. The values of the diffusivities were obtained from rotational

    diffusivity and average run speed () and compared with those ob-tained using MSD (). (C) Average run speeds ( standard error)as a function of distance in the absence of serine (), and in thepresence of serine, 5000 M () when the buffer is saturated withair. Average run speeds in the absence of serine (), and in thepresence of serine, 5000 M (N) when the buffer is saturated with

    oxygen. (D) Average run speeds ( standard error) as a functionof serine concentration at 500 m. N indicates the average run

    speed of 18.65 0.62 m/s obtained in the absence of serine whensaturated with air. . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

    4.2 (A) Rotational diffusivity as a function of serine concentration at

    500 m. N indicates the rotational diffusivity 0.380.05 rad2/sobtained in the absence of serine when saturated with air. (B)

    Measured drift velocity ( standard error) as a function of distancein the absence of serine (), and in the presence of serine, 5000 M() when the buffer is saturated with air. Measured drift velocityin the absence of serine (), and in the presence of serine, 5000 M(N) when the buffer is saturated with oxygen (C) Measured drift

    velocity ( standard error) as a function of distance for varyingserine concentration when saturated with air: 0 M (), 50 M(), 250 M () and 5000 M (). . . . . . . . . . . . . . . . . 64

    x

  • 4.3 The measured drift velocity ( standard error) as a function ofdistance for various gradients of serine (G) and initial ligand con-

    centration (L0); (A) G = 0.0016 M/m and L0 = 1.6 M () (B)G = 0.016 M/m and L0 = 16 M () (C) G = 0.16 M/mand L0 = 160 M () (D) G = 1.6 M/m and L0 = 1600

    M () (E) G = 16 M/m and L0 = 16000 M ( N) and (F)

    G = 160 M/m and L0 = 160000 M (). . . . . . . . . . . . . . 67

    4.4 (A) Drift velocity obtained through experiments and simulation as a

    function of serine gradients at three different spatial locations: solid

    line and at 500 m, dashed line and at 1000 m, dotted lineand at 1500 m respectively. (B) Drift velocity obtained throughexperiments and modified predicted values by addition of an average

    drift velocity due to oxygen gradient and serine concentration as a

    function of serine gradients at three different spatial locations: solid

    line and at 500 m, dashed line and at 1000 m, dottedline and at 1500 m respectively. . . . . . . . . . . . . . . . . . . 68

    4.5 (A) Drift velocity obtained through experiments as a function of

    serine gradients at 500 m: solid line with represent our experi-mental data while the dotted line with represents that observedby Berg and Turner (1990). (B) Measured drift velocity ( stan-dard error) as a function of the spatial gradient of the logarithmic

    serine concentration: G = 0.0016 M/m (), G = 0.016 M/m(), G = 0.16 M/m (), G = 1.6 M/m (), G = 16M/m () and G = 160 M/m (N) (C) Average value of thedrift velocity ( standard error) at each d(lnL)/dx with location.The solid line represents a linear curve fit. . . . . . . . . . . . . . . 70

    xi

  • 5.1 (A) The number of E. coli cells as a function of displacement mea-

    sured over a time period of 3.3 s for uniform glucose concentra-

    tion of 28 M in the capillary: Experimental data:, Gaussianfit:solid line (B) Gaussian fits for zero gradient: solid line, and

    0.05 M/m, ...... . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

    5.2 (A) Mean square displacement (variance from Gaussian fit) for all

    experiments (B) Apparent diffusivity as a function of time for var-

    ious glucose concentrations and gradients. . . . . . . . . . . . . . . 79

    5.3 (A) Measured drift velocity for varying glucose gradients (B) Mea-

    sured drift velocity in the absence of glucose. The data excludes the

    contribution of non-chemotactic drift (C) Measured drift velocity

    for varying glucose concentrations (D) Drift velocity as a function

    of effective glucose gradients: for varying glucose concentrationexperiments in the absence of initial gradients (see Figure 5.3 (C)),

    for initially established glucose gradients (see Figure 5.3 (A)),and .... for Hill equation obtained by fit. . . . . . . . . . . . . . . . 81

    5.4 (A) Measured drift velocity as a function of the spatial gradient of

    the logarithmic glucose concentration. The dotted line represents

    the linear variation (see equation 5.2) with slope 1000 m2/s. (B)

    Average run speeds as a function of effective glucose gradients:for varying glucose concentration experiments in the absence of ini-

    tial gradients (Figure 5.3 (C)), for initially established glucosegradients (see Figure 5.3 (A)), for absence of glucose and ....for average value in the presence of glucose. . . . . . . . . . . . . . 83

    xii

  • List of Tables

    1.1 Mathematical models overview . . . . . . . . . . . . . . . . . . . . . 14

    2.1 Gradients and ligand concentration at x = 0 obtained through ex-

    periments. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

    3.1 Comparison of chemotactic parameters in E. coli with those re-

    ported in literature. (Berg and Brown, 1972; Liu and Papadopou-

    los, 1996; Berg and Turner, 1990; Ahmed and Stocker, 2008; Kalinin

    et al., 2009) These values were obtained by averaging ( standarddeviation) over all experiments. . . . . . . . . . . . . . . . . . . . . 38

    3.2 Parameters used to simulate the model. . . . . . . . . . . . . . . . . 48

    xiii

  • Chapter 1

    Introduction

    1.1 Background

    Chemotaxis is a phenomenon in which a microorganism or multi-cellular organ-

    isms are able to direct their movements in the presence of certain chemicals in

    their environment. Cells capable of chemotaxis are bacteria, protozoa, amoeba,

    cellular slime moulds, sperm and fibroblasts phagocytes. Chemotaxis is used by

    organisms to find food by swimming towards the highest concentration of food

    molecules, or to flee from poisons. Chemicals which are capable of eliciting such a

    response from chemotactic cells are called Chemotactants. Chemotactants may act

    as attractants, in which case the chemotactic cell will move towards them, thereby

    exhibiting a positive chemotactic response. On the other hand, some chemicals act

    as repellents, in which case the chemotactic cell will move away from them, thereby

    exhibiting a negative chemotactic response. Examples of chemoattractants include

    nutrients such as sugars, amino acids and small peptides, while chemorepellents

    include antibiotics and noxious chemicals such as phenol.

    Chemotaxis plays an indispensable role in our life from birth to death. During

    embryonic process cells migrate to their target destinations to become components

    of arms, legs, liver, heart, brain and other organs. In the brain, immature neurons

    move from their birthplace to final targets where they make the right connections

    and allow functions such as learning and memory. In the process of wound healing,

    1

  • white blood cells migrate towards the site of infection to destroy the foreign organ-

    ism. (Condliffe and Hawkins, 2000) In metastasis, cancer cells secrete their own

    chemotactic stimulus to direct their migration towards lymphatic vessels using the

    lymphatic system as a major route to spread throughout the body. (Carlos, 2001;

    Wood et al., 2006; Shields et al., 2007) Cell migration is also important in bio-

    film formation and bio-remediation of surface contaminants. (Duffy et al., 1997;

    Pandey and Jain, 2002) In all these processes, how the cells sense the environment

    and migrate to specific locations is not clear. Bacterial chemotaxis provides a

    basic platform for learning how these processes work which in turn will help in

    unraveling similar processes in more complex systems.

    Escherichia coli is a model organism for the study of bacterial chemotaxis,

    because its genetics is comparatively simple, is generally harmless, can be grown

    quickly on a wide range of media and much of the biochemical pathway is known.

    Thus, it has been used for experimental and theoretical studies of chemotaxis over

    the last 40 years. Since the discovery of chemotaxis by Pfeffer and Englemann in

    1880, many studies on bacterial chemotaxis have been reported in the literature.

    However, the much needed impetus to the study of bacterial chemotaxis began in

    the 1960s by Julius Adler who noted that bacteria used specific receptors to recog-

    nize the chemicals. Since then, numerous studies have identified these receptors,

    unraveled the workings of the receptors and the signaling pathway, recorded the

    motion in detail, in the presence of various chemoattractants and, more recently,

    tried to integrate this knowledge into detailed mathematical models to predict

    the motion of E. coli in the presence of chemicals. Many of these studies will be

    reviewed in the following pages of this chapter.

    E. coli are extremely tiny creatures. They are rod-shaped, about 2 m long

    and 1 m in diameter and are mostly motile. They can achieve motility via the

    utilization of specialized structures known as flagella. These flagella allow the

    bacteria to exhibit both positive and negative chemotactic responses. A single

    bacterial cell may possess multiple flagella as shown in Figure 1.1. (Eisenbach,

    1994)

    2

  • Figure 1.1: Schematic diagram of E. coli with multiple flagella. It is typically

    rod-shaped and is about 2 m long and 1 m in diameter. (Eisenbach, 1994)

    Swimming in the absence of chemotactant: In the absence of chemo-

    tactant, E. coli executes a random walk composed of runs and tumbles. Smooth

    swimming in a straight line is called a run and an abrupt turning motion is called

    a tumble as shown in Figure 1.2. A cell is propelled by a set of several helical

    flagellar filaments that arise at random points on its sides and extend several body

    lengths out into the external medium. Each filament is driven at its base by a

    rotary motor embedded in the cell envelope. The energy for the motor comes from

    proton motive force, i.e., by the flow of protons (hydrogen ions) from the outside

    to the inside of the cell. During the runs, the filaments coalesce into a bundle

    that pushes the cell forward. When viewed from behind the cell, the bundle ro-

    tates counterclockwise (CCW), and, to balance the torque, the cell body rotates

    clockwise (CW). Tumbles are initiated by CW motor rotation and during this the

    reversed filaments come out of the bundle and go through a series of polymorphic

    transformations from normal to semi-coiled and then to curly. (Berg, 2004) This

    change in course defines the tumble interval. When the motor switches back to

    CCW rotation, the filaments regain its normal conformation and rejoin the bun-

    3

  • dle. Turner et al. (2000) noted that tumbling occurs only when 25% or more ofthe flagellar filament on a given cell reverse to clockwise direction.

    Figure 1.2: Swimming behavior of E. coli or Salmonella in the absence of chemo-

    tactant. It executes a random walk composed of runs and tumbles with zero drift.

    Swimming in the presence of chemotactant: An increasing chemoat-

    tractant concentration or a decreasing chemorepellent concentration decreases the

    probability of CW rotation and, therefore, the probability of tumbles. Thus, the

    final outcome is a random walk of the bacterial cell, biased towards the chemoat-

    tractant or away from the chemorepellent as shown in Figure 1.3. Segall et al.

    (1986) studied the impulsive chemotactant response of the cell and observed that

    the cell compares the concentration over the past 1 s with that observed over

    the previous 3 s and responds to the difference. This implies that the concen-

    tration/gradient sensed by the bacteria is temporal in that the bacteria possess

    a memory, which compares the past information with the present information to

    make a decision. So the cell decides whether life is getting better or worse. If its

    getting better, they continue in the same direction and if its getting worse, they

    wont worry about it.

    The regulation of this chemotaxis phenomena in the bacteria is achieved by a

    network of interacting proteins. (Eisenbach, 1994) The basic mechanism in flagel-

    lated bacteria involves a receptor mediated phosphorylation of a cytoplasmic pro-

    tein CheY that binds to the flagellar motor and changes the frequency of tumbles.

    There are essentially six cytoplasmic proteins viz., CheA, CheB, CheR, CheW,

    4

  • Figure 1.3: Swimming behavior of E. coli or Salmonella in the presence of chemo-

    tactant. It executes a biased random walk composed of runs and tumbles with

    significant drift.

    CheY and CheZ that are needed to process the sensory signals and transmit con-

    trol signals to the flagellar motor. E. coli has five methyl-accepting chemotaxis

    proteins (MCPs) (referred to as chemotaxis receptors) that mediate responses to

    serine (Tsr), aspartate and maltose (Tar), ribose and galactose (Trg), and dipep-

    tides (Tap). These MCPs are transmembrane proteins that span from the internal

    to the external surface of membrane and are made of 550 amino acids. Of these

    receptors, Aer is the most specific one that senses oxygen and energy levels of the

    cell. The signaling domain of the MCP modulates the activity of the CheA to

    elicit chemotactic response. The CheW protein couples CheA to MCPs to form a

    complex. The CheA phosphorylates which in turn phosphorylates two other regu-

    lator proteins, CheY and CheB. The phosphorylated CheY activates the flagellar

    motor switch protein FliM (M) resulting in increased tumbling frequency. In vivo

    experimental studies (Cluzel et al., 2000) have reported CW bias (the fraction of

    time spent in CW rotation) as a function of CheY-P concentration (Yp) and is

    shown in Figure 1.4. It is known that for a fixed value of the ligand concentration

    there exists a fixed concentration of Yp which yields the corresponding CW bias.

    This in turn decides the probability of a run or a tumble event. If CheY-P concen-

    tration is very low, which means negligible CW value, cell will run and vise versa.

    Also, the CheZ promotes the CheY dephosphorylation. The phosphorylated CheB

    5

  • removes the methyl groups from MCPs, where as CheR continuously adds methyl

    groups to MCPs. The level of methylation of MCPs decides the tumbling fre-

    quency of the organism. (Falke et al., 1997) A schematic diagram of the signaling

    pathways in E.coli is shown in Figure 1.5.

    Figure 1.4: Experimental data using fluorescence correlation spectroscopy: Re-

    sponse of individual motors as a function of CheY-P concentration. (Cluzel et al.,

    2000) Each data point describes a simultaneous measurement of the motor bias

    and the CheY-P concentration in an individual bacterium. CW bias was computed

    by analyzing video recordings for at least 1 min.

    When an attractant is added, it binds to the receptor (specific to that attrac-

    tant) and the rate of autophosphorylation of CheA decreases. Consequently, the

    rate of phosphorylation of CheY and CheB decrease causing a reduction in the ac-

    tivity of FliM and decrease in the tumbling frequency. However, the CheR contin-

    ues to methylate MCPs, thereby increasing the level of methylation of MCPs. This

    progressively reduces the attractant binding affinity to the MCPs and increases

    the rate of autophosphorylation of CheA. The phosphorylated CheY and CheB

    return to their initial state. Thus, after the introduction of a chemoattractant, the

    6

  • Figure 1.5: Chemotaxis signaling pathways in E. coli (A) in the absence of

    chemoattractant, and (B) in the presence of chemoattractant. The dotted line

    represents the reactions with reduced rate in response to ligand binding to the

    receptor. It can be noted that on adaptation the pathway returns to steady state

    as shown in (A).

    7

  • tumbling frequency reduces for a short time, after which the tumbling frequency

    increases to the preattractant level. The initial reduction in tumbling frequency

    increases the run length and biases the motion of cell towards the attractant. On

    the other hand, in the presence of a repellent, the rate of autophosphorylation

    of CheA increases resulting in the increase in frequency of tumbles. (Eisenbach,

    1994) Quantification of this chemotaxis signaling pathway requires construction

    of mathematical models that describe the operation of the whole network.

    Experimentally, the chemotactic behavior of E. coli can be quantified using

    various parameters such as average run speed, clock-wise bias, drift velocity, cell

    diffusivity, rotational diffusivity, etc. Of the four commonly reported parameters,

    the first is the run speed that gives the average speed of the cells in between

    consequent tumbles. The second is the clock-wise bias which gives the fraction

    of time spent by the cells in tumble mode. Of the next two parameters, the

    drift velocity measures the mean speed of cells while the translational diffusivity

    characterizes its random motion. Note that the random motion is induced by the

    collisions of cells with the solvent molecules as well as due to their own tumbling

    and is calculated as the mean square distance traversed by the cells upon time.

    A related parameter is the rotational diffusivity calculated as the mean square

    angular deviation upon time. A wide range of experimental methods have been

    developed so far to analyze these chemotactic parameters. (Eisenbach, 1994) They

    include capillary (Adler, 1966a), swarming plate or ring forming (Wolfe and Berg,

    1989), stopped flow diffusion chamber (Ford et al., 1991), micro-capillary (Liu

    and Papadopoulos, 1996), diffusion gradient chamber (Widman et al., 1992) and

    microfluidic assays (Berg and Turner, 1990), (Hanbin et al., 2003), (Ahmed and

    Stocker, 2008), (Kalinin et al., 2009), all of which are reviewed in the next section.

    Experimental methods for analyzing chemotaxis

    Capillary assay: This was the first technique developed by Adler (Adler, 1966a) for

    measuring bacterial chemotaxis. A capillary tube containing a chemoattractant-

    free buffer is placed in a suspension of bacteria. The bacteria enter the cap-

    8

  • illary and accumulate in it. The extent of this accumulation is a measure of

    the motility of the bacteria. When the capillary contains a chemoattractant, a

    gradient is developed by diffusion. The bacteria follow this gradient and accumu-

    late in the capillary. For measuring negative chemotaxis, a capillary containing

    chemorepellent-free buffer is immersed in a chemorepellent containing suspension

    of bacteria. The bacteria escape from the chemorepellent into the capillary and

    accumulate in it. The bacteria in the capillary can be counted by plating them

    or by observing them under a microscope. The chemotaxis receptors become sat-

    urated when the capillary contains very high concentrations of chemoattractant.

    In this case the bacteria were unable to sense the gradient and a sharp drop was

    observed in the number of bacteria accumulated in the capillary. This is the most

    commonly used method for chemotaxis but the response depends on chemotactant

    transport, metabolism, and growth of E. coli.

    Swarming plate or Ring forming assay: Bacteria are placed at a certain spot

    in a plate containing semisolid agar and a low concentration of a metabolizable

    chemoattractant. The bacteria metabolize the chemoattractant and thereby pro-

    duce a chemoattractant gradient. The bacteria use up the local supply of chemoat-

    tractant and follow the induced gradient. Due to this, a continuous expanding ring

    of dense bacteria is formed. This ring marks the boundary between the region that

    has been depleted of chemoattractant and the region still containing the chemoat-

    tractant. When the plate contains a number of chemoattractants, a number of

    expanding rings are formed. Note that this method of quantifying chemotaxis is

    restricted to metabolizable chemoattractants only (Wolfe and Berg, 1989). Also,

    there is no control over the induced gradient and the final response depends on

    both metabolism and growth.

    Stopped-flow diffusion chamber assay: In the study by Ford et al. (1991), the

    bacterial migration behavior to the gradients was measured using a stopped flow

    diffusion chamber. In this assay, two bacterial suspensions differing in stimulant

    concentrations were contacted by impinging the two flows. As long as there is

    a flow through the chamber, there is no mixing between the two suspensions.

    9

  • However, once the flow is stopped, a transient attractant concentration gradient is

    created by diffusion and the variations in bacteria density can then be measured

    by light scattering. This method can be used to quantify the bulk properties of

    populations such as drift velocity and translational diffusivity.

    Micro-capillary assay: The micro-capillary assay used by Liu and Papadopou-

    los (1996) consisted of two reservoirs communicating through a long capillary of

    50 m inner diameter. The linear concentration profile for the chemoattractant

    was achieved by filling one of the reservoirs with motility buffer while the other

    one was filled with the chemoattractant solution. It was assumed that after 24

    hours, the concentration profile would attain the steady-state linear concentration

    profile. Here, the single cell parameters such as run speed, run length, turn angles

    etc were measured in the capillary with the help of a microscope.

    Diffusion gradient chamber assay: Widman et al. (1992) used a diffusion gradi-

    ent chamber assay that consists of a square arena bounded by a reservoir on each

    side. Each reservoir was separated from the arena by a semi permeable mem-

    brane. From the source and sink reservoirs, mediums containing attractants and

    containing no attractant respectively were pumped using a syringe pump. The

    system was allowed to establish the gradients partially after which the cells were

    inoculated at the center point in the arena of the chamber. Glucose and oxygen

    concentrations in the chamber were measured using micro-biosensor and the cell

    growth and migration patterns were observed with time. In this study, however,

    the concentration of the aspartate which was used as a chemoattractant was not

    measured.

    Microfluidic assays: Berg and Turner (1990) studied the chemotaxis in glass

    capillary arrays. In this assay, two stirred chambers were separated by a micro-

    channel plate comprising a fused array of capillary tubes. Here, the cells added to

    first chamber migrate to the second chamber via multichannel plate. The density

    of these cells were then calculated by measuring the scattering of light beam

    incident from a laser diode. The study found that the flux of bacteria increased

    on addition of an attractant into the second chamber compared to when there was

    10

  • no attractant. A linear concentration profile of the chemoattractant was assumed

    between the chambers. The drift velocity was calculated from the knowledge

    of total flux for varying chemoattractant gradients. In this study, the motility

    medium consisted of sodium lactate which is a carbon source for E. coli and

    whose effect on chemotaxis was neglected. Later, Ahmed and Stocker (2008) used

    a micro-channel connected at right angles to a side branch. Initially, a solution

    of chemoattractant and fluorescein was injected into the main channel. After the

    main channel was completely filled with the solution, the motility buffer containing

    cells was injected into the side channel at a constant flow rate so that the cells

    migrate to a constant gradient into the main channel established by diffusion. The

    trajectories of the bacteria along with the concentration gradient were recorded.

    The chemotactic parameter diffusivity was computed from the motion of a single

    cell as well as population.

    Recently, Kalinin et al. (2009) have used microfluidic assays to quantify the

    chemotaxis. In this study, three parallel channels were patterned in an agarose

    gel. A fluorescent solution initially flowed through the upper source channel which

    was later replaced with the chemoattractant while a blank buffer flowed through

    the lower sink channel. This established a linear chemical gradient in the central

    channel because of the diffusion that takes place through the agarose gel from the

    source channel to the sink channel. The cells were then introduced into the central

    channel and their trajectories were recorded. Using the tracked positions, the

    cell diffusivity and the chemotactic migration coefficient (CMC), i.e., the average

    vertical position of all the cells tracked with respect to the central position of the

    channel were calculated. Further, the study suggested that the cells respond to

    the spatial gradient of the logarithmic attractant concentration. However, it is

    important to note that the motility buffer used in this study contained lactic acid

    which is a carbon source and whose influence on chemotaxis was ignored.

    11

  • Mathematical studies on chemotaxis

    A number of mathematical models have been developed to describe different as-

    pects of chemotaxis at the level of a single cell (Levin et al., 1998; Novere and

    Shimizu, 2001; Lipkow et al., 2005) in addition to descriptions of entire bacterial

    populations (Keller and Segel, 1970, 1971; Alt, 1980; Rivero et al., 1989; Newman

    and Grima, 2004). Most of the population studies used Keller and Segel model

    (Keller and Segel, 1970) for analyzing chemotactic population behaviors, which

    was basically developed for modeling the movement of slime molds. According to

    this model, the flux of cells ~Jb will be given by,

    ~Jb = ~b+ b~c (1.1)

    where b is the cell density, c is the chemical stimulus concentration, is the ran-

    dom motility coefficient that measures the translational diffusivity of a population

    of bacteria resulting from the random walk behavior, while is the chemotactic

    sensitivity coefficient which represents the strength of the attraction of a popula-

    tion of bacteria to a given chemical. In physical terms, the bacterial flux comprises

    of two parts, namely, diffusion (or random motility) and chemotactic motion char-

    acterized by drift velocity.

    Lovely and Dahlquist (1975) were the first to relate the individual-cell observa-

    tions to the random motility coefficient (), a macroscopic cell transport parameter

    describing population-scale motility,

    =v2

    3(1 cos) (1.2)

    where v is the bacterial swimming speed, is the run time and is the turn angle.

    Recall that, E. coli chemotaxis relies on temporal sensing mechanism to detect

    spatial gradients as they swim through them and this mechanism is governed by

    the ability of an individual cell to compare the fraction of bound chemoreceptors

    on its cell surface at various points in time. A change in the number of bound

    12

  • chemoreceptors allows a cell to change its CW bias. Experimentally, Berg and

    Brown (1972) found that, when cell is traveling in the direction of an increasing

    spatial chemical gradient, it extends their run lengths to continue traveling in

    that direction. The study suggested that, in the presence of a chemical attractant

    the number of bound receptors on a cells surface increases with time resulting in

    reduced tumbling frequency. To describe this behavior Rivero et al. (1989) used

    the Keller and Segel model. At the population level, the chemotactic migration

    of bacteria was described using the drift velocity (vc),

    vc =2

    3v tanh

    (o2v

    Kd(Kd + L)2

    L

    x

    )(1.3)

    where v is the run speed, o is the chemotactic sensitivity coefficient and L is the

    attractant concentration. Kd, the attractant-receptor dissociation constant, is a

    measure of the ability of a membrane-bound chemoreceptor on the cell surface

    to detect a specific attractant. However, these population studies do not account

    for changes in the intracellular pathway to variations in extracellular attractant

    concentration.

    Chemotaxis intracellular signaling pathway of E. coli provides a unique op-

    portunity to identify and develop computational methods required to obtain a

    quantitative understanding of the intracellular signaling pathways. Recent models

    for chemotaxis incorporate the signaling pathway at molecular level and integrate

    it with motor response to predict bacterial motion. Examples include AgentCell

    (Emonet et al., 2005), E solo (Bray et al., 2007) and RapidCell (Vladimirov et al.,

    2008). AgentCell simulates the complete signaling pathway stochastically in a

    single cell and predicts the motion in a three-dimensional environment. Esolo

    is a deterministic model that solves ordinary differential equations represent-

    ing the signaling reactions in the pathway and predicts bacterial movement in

    two-dimensional environment. RapidCell utilizes the Monod-Wyman-Changeux

    (MWC) (Mello and Tu, 2005; Keymer et al., 2006) two-state receptor model for

    mixed receptor clusters, incorporates the adaptation dynamics and connects the

    CheY-P values to the cells tumble/run to predict the motion of E. coli in a

    13

  • Table 1.1: Mathematical models overview

    Model Description

    Spiro et al. (1997) Used a three methylation state model and simulated the

    response to temporal variation of aspartate.

    Barkai and Leibler (1997) Model ensured that adaptation is robust and showed

    that adaptation time is inversely proportional to

    receptor-complex activity.

    Levin et al. (1998) Receptor modification reactions catalyzed by CheR and

    CheB incorporated in Bacterial ChemoTaxis (BCT)

    model. Model explored the consequence of variation

    in protein expression CheY-P and the phenomenon of

    non-genetic individuality.

    Morton-Firth et al. (1999) Stochastic simulation of bacterial chemotaxis

    (StochSim): Here, the activity of receptor com-

    plexes was determined by free-energy changes due to

    both ligand binding and changes in methylation state.

    Shimizu et al. (2001) This extended the model of Morton-Firth et al. (1999)

    and allowed interactions between neighboring receptor

    complexes arranged in a regular lattice according to the

    free energy of a receptor complex determined by the

    activity states of its immediate neighbors. Model simu-

    lated with different lattice sizes and geometries to find

    the effect of the coupling energy between neighboring

    receptors on the signal to-noise ratio and gain.

    Mello and Tu. (2003) Deterministic model used to determine the full set of

    conditions under which the system achieves perfect

    adaptation.

    14

  • Rao et al. (2004) Used both Barkai and Leibler (1997) and Sourjik and

    Berg (2002) models to compare the intracellular path-

    ways of E. coli and B. subtilis. Both pathways shown

    to be adaptive and robust.

    Lipkow et al. (2005) Using the Smoldyn program, model simulated the diffu-

    sive movement of individual CheY molecules and their

    binding with receptor complexes, CheZ and FliM.

    Emonet et al. (2005) AgentCell, an agent-based program that relates stochas-

    tic intracellular processes to the behavior of individ-

    ual cells and bacterial populations. Cells represented

    as agents made up of chemotaxis proteins, motors and

    flagella that can move through a three dimensional en-

    vironment.

    Bray et al. (2007) E solo model which uses ordinary differential equa-

    tions of the signaling reactions in the pathway and gives

    the graphical display of bacterial movement in two-

    dimensional environment.

    Vladimirov et al. (2008) RapidCell model uses the Monod-Wyman-Changeux

    (MWC) model for mixed receptor clusters, adaptation

    dynamics and a model of cell tumbling to give the mo-

    tion of E solo in two-dimensional environment.

    15

  • two-dimensional environment. It is only recently that Kalinin et al. (2009) using

    MWC model, predicted the bacterial motion to varying chemoatttactant gradi-

    ents and compared them with experimental results. The computational single cell

    mathematical models used for bacterial chemotaxis are briefly listed in Table 1.1.

    In addition to the chemical stimuli, microorganisms also sense other stimuli

    such as light, temperature, electric field, etc. In all cases, the name of the response

    includes a prefix that describes the stimulus. The suffix taxis means moving to-

    wards or away from the stimulus. They include phototaxis movement directed by

    light, thermotaxis by temperature changes, electrotaxis by electrical field, mag-

    netotaxis by magnetic field, geotaxis by gravity, elasticotaxis by elastic force etc.

    (Berg, 2004)

    1.2 Motivation

    The traditional techniques to characterize chemotaxis were the agarose gel as-

    say (Wolfe and Berg, 1989) and the capillary assay (Adler, 1966a; Liu and Pa-

    padopoulos, 1996) due to their simplicity. In both these methods, E. coli moves

    up a gradient set by the consumption of the attractant and there is no control

    over the induced gradients. More recent experiments have been able to establish

    controlled gradients using microfluidic techniques (Kalinin et al., 2009) that al-

    low the measurements of drift velocities over a wide range of gradients. However,

    these techniques fail to capture temporal and spatial variation of the chemotactic

    response of E. coli to spatial variations of chemoattractants. In parallel with the

    experimental work, a number of extensive mathematical models have been pro-

    posed for E. coli chemotaxis, but there exists no study which predicts the bacterial

    motion along a gradient and compares them with experiments. Recall that the

    motion of E. coli not only depends on the local ligand concentrations but also on

    past history of concentration experienced by it. Further, the extent of influence

    of oxygen and energy source on the motion is not clear. How the intracellular sig-

    naling response reflects the extracellular response characterized by motion is still

    16

  • an active research area. The understanding of the available intracellular signaling

    pathway and absence of experimental and theoretical study which quantifies the

    chemotaxis in both time and space in controlled gradients motivated us to carry

    out the work presented in this thesis.

    1.3 Objectives

    The main objective of the current study is to quantify the response of microorgan-

    isms to varying chemoattractant concentrations/gradients using a micro-capillary.

    This would involve developing mathematical models that describe the chemotaxis

    pathway in a chemotactic cell and to relate the kinetics to the motion of cell in

    the presence of chemoattractants. The predictions of the model will be compared

    with experiments where the motion of cell in the presence of chemoattractants will

    be tracked.

    The overall objectives are summarized below :

    Develop a novel technique to quantify the chemotaxis in E.coli in the pres-ence of chemoattractants in controlled environment at the phenotypic level

    in a suitable device.

    Develop a mathematical model to quantify the motion of E.coli using avail-able signaling pathway and connect it to the motion.

    Validate the model results with experimental data to get better insight ofthe chemotaxis mechanism.

    1.4 Organization

    The work presented in this thesis is organized into six chapters describing the

    research methodology and results, followed by conclusions and recommendation

    for future work. In chapter 1, an overview of bacterial chemotaxis phenomena is

    presented. Chapter 2 describes the various experimental and theoretical methods

    17

  • used in the present work. Here, the methods developed for modeling the motion

    of E. coli and experimental techniques used for establishing various controlled

    gradients have been described. In chapter 3, we describe the mathematical model

    along with experimental validation of chemotaxis under controlled gradients of

    methyl-aspartate in E. coli while chapter 4 deals with a similar study in the

    presence of a metabolizable attractant, namely, L-serine. Chapter 5 discusses the

    phenomena of chemotaxis in the presence of glucose via the phosphotransferase

    (PTS) pathway. Finally, the overall conclusions and directions for future work are

    presented in chapter 6.

    18

  • Chapter 2

    Materials and Methods

    2.1 Experimental Protocols

    Microorganism

    Recall that E.coli has five methyl-accepting chemotaxis proteins (MCPs) that

    mediates metabolism-independent chemotactic responses in E.coli. The high-

    abundance chemoreceptors of this family (Tsr and Tar) mediate responses to serine

    and aspartate, respectively. The low-abundance chemoreceptors (Tap and Trg)

    are present at only 10 percent of the concentration of Tsr and while Tap mediates

    response to dipeptides, Trg senses galactose and ribose by means of their respective

    periplasmic binding proteins. (Li and Hazelbauer, 2004) In this study, we used

    Escherichia coli K-12 (MTCC 1302, IMTECH Chandigarh, India) strain and

    chemoreceptors characterization was done by Chromous Biotech Pvt Ltd. The

    detailed report can be found at the end of this chapter and shows that the strain

    used throughout our study has four MCPs (Tsr, Tar, Tap and Aer) excluding Trg

    receptor. The culture was revived over monthly intervals.

    Rectangular micro-capillaries

    Micro-capillaries were obtained from Arte Glass Associates Co., Ltd, Japan. The

    dimensions of the capillaries are 5 cm (L) 1000 m (W) 100 m (H).

    19

  • Chemicals

    Luria Bertani (LB) broth was obtained from Hi-media company. KH2PO4, K2HPO4,

    (NH4)2SO4, MgS04. 7H20, Ehylenediaminetetraacetic acid (EDTA), Polyvinyl

    pyrrolidine (PVP), D-glucose (glucose), -methyl-DL-aspartate (MeAsp) and L-

    serine (serine) were obtained from Sigma-Aldrich company. 2-(N-(7-nitro-benz-2-

    oxa-1,3-diazol-4-yl)amino)-2-deoxy-glucose (2-NBDG) was obtained from Invitro-

    gen Corporation. Tryptone (Difco) and Bacto agar (Difco) were obtained from

    BD Biosciences company.

    Media

    The motility buffer (MB) contained (/l of distilled water) K2HPO4, 11.2 g; KH2PO4,

    4.8 g; (NH4)2SO4, 2 g; MgS04. 7H20, 0.25 g; PVP, 1 g; and EDTA, 0.029 g. (Adler,

    1973) Luria-Bertani (LB) broth contained LB (/l of distilled water) 25 g. Tryptone

    medium contained (/l of distilled water) tryptone, 10 g and Nacl, 5 g. Glucose

    medium contained (/l of distilled water) glucose, 4 g and the salts (/l distilled wa-

    ter): K2HPO4, 11.2 g; KH2PO4, 4.8 g; (NH4)2SO4, 2 g; and MgS04. 7H20, 0.25g.

    Different concentrations of chemotaxis medium was prepared by adding different

    amounts of chemoattractant to the motility buffer under sterilized conditions.

    Growth conditions

    The growth conditions of the bacteria are critical to the success of the chemotaxis

    experiments. In glucose experiments, the bacteria were grown as per the following

    procedure: One loop full of culture from the slant was inoculated into LB media

    and allowed to grow for 9 hours in the exponential growth phase. The incubation

    was always carried out at 37oC and 240 rpm. After the culture was grown for

    9 hours, 1 ml of the culture sample was transferred to the sterilized LB medium

    and left to further grow for 6 hours (early exponential phase). Next, 1 ml of the

    culture broth from LB medium was transferred to the glucose medium and allowed

    to grow for 6 hours (exponential phase). To ensure that the cells are adapted to

    20

  • glucose, 10 ml of sample was transferred to glucose medium and grown again for

    4 hours (early exponential phase). The biomass was separated by taking 50 ml of

    the culture into sterilized tubes and centrifuged at 4000 rpm for 10 minutes. The

    supernatant was decanted and the settled pellet was gently re-suspended in 10 ml

    motility buffer. In order to provide energy to the bacteria, 28 M of D-glucose was

    added to the motility buffer in all experiments. The above procedure was repeated

    three times before introducing the cells into the capillary. High levels of bacterial

    motility were observed on viewing the cells under an optical microscope. Finally

    for the chemotaxis measurements, cells were introduced gently by touching the

    pellet with the mouth of the capillary. Then, concentration of cells in the capillary

    were calculated by diluting them in the buffer followed by plating. Approximately

    106-107 bacteria/ml were taken into the capillary in all experiments.

    We used similar procedure for MeAsp experiments and serine experiments with

    some minor protocol modifications. After the culture was grown for 9 hours, 1

    ml of the culture broth from LB medium was transferred to tryptone medium

    and allowed to grow for 6 hours (exponential phase). To ensure that the cells

    are adapted to tryptone medium, 10 ml of sample was transferred to tryptone

    medium and grown again for 4 hours (early exponential phase). After this, the

    washing procedure is similar except that the motility buffer does not contain any

    chemoattractant in the gradient experiments. However, we added chemoattractant

    to the motility buffer for observing the motion of E. coli for obtaining uniform

    concentrations (zero gradient) of the attractant.

    Calibration of intensity for 2-NBDG measurements

    The micro-capillaries were first sterilized and were marked with graduations spaced

    at 0.5 cm along its length. The concentration gradients were first established using

    different concentrations of 2-NBDG (fluorescent glucose) solutions using motility

    buffer. In order to calibrate the intensity of 2-NBDG solutions, a 5 cm plug of

    2-NBDG solution was drawn into the capillary and the ends were sealed with

    wax. Using a microscope with a 4X (NA 0.13) objective lens, the fluorescence

    21

  • intensity of 2-NBDG was measured at the center of the capillary. The fluorescence

    intensity was found to vary linearly with 2-NBDG concentration up to 1000 M.

    The calibration chart is shown in Figure 2.1.

    Figure 2.1: Linear calibration curve: Fluorescence intensity with varying concen-

    tration of 2-NBDG.

    Establishment of 2-NBDG gradients for glucose experiments

    Figure 2.2: Micro-capillary experimental setup for establishing glucose gradients.

    Experiments were performed to establish different gradients of 2-NBDG. A 2.5

    cm plug containing a fixed concentration of 2-NBDG was drawn into the capillary

    followed by a 2.5 cm plug of a lower concentration of 2-NBDG. Approximately 107

    22

  • Figure 2.3: Measured fluorescence intensity in the capillary for 10 min (solid line)

    and 540 min (dashed line) after the start of the experiment. It can be noted that

    the gradient was stable for more than 9 hours.

    Figure 2.4: (A) Agar plate containing 0.3 % agar and motility buffer (absence of

    glucose) with 107 cells introduced at the center of the plate. The image wastaken after 6 hours of incubation. It can be noted that no ring was formed. (B)

    Agar plate containing 0.3 % agar, motility buffer and glucose (5 mM), with 107

    cells introduced at the center of the plate. The image was taken after 6 hours of

    incubation. Clearly, the cells have migrated forming a distinct ring.

    23

  • bacteria /ml were taken into the capillary by contacting the pellet (as described

    earlier) into the mouth of the capillary. As before, the ends of the capillary

    were sealed with wax. A schematic diagram of the micro-capillary experimental

    setup is shown in Figure 2.2. Using a microscope with a 4X (NA 0.13) objective

    lens, the fluorescence intensity of 2-NBDG was measured along the length of the

    capillary as a function of time. The results are shown in Figure 2.3. It is noticed

    that the intensity profiles attains a steady state within 10 minutes and negligible

    change in the concentration profile was observed up to 540 minutes. These profiles

    were robust and easily reproducible and were not affected by the cell movements.

    Growth experiments with 2-NBDG demonstrated that there is no change in OD

    for 4 hours (results are not shown) indicating that 2-NBDG was not metabolized.

    Also, experiments were performed in agar plate assays to check the chemotactic

    response of the current strain to glucose. Clear rings were formed within 6 hours

    indicating that the E. coli strain lacking Trg shows normal chemotactic behavior

    towards glucose (Figure 2.4). Further, for the chemotaxis experiments, normal

    D-glucose was used instead of 2-NBDG as the cells did not respond to 2-NBDG.

    Establishment of 2-NBDG gradients for MeAsp and Serine

    experiments

    Initially, a 4.5 cm liquid plug containing a fixed concentration of 2-NBDG was

    drawn in the capillary followed by about 0.5 cm plug of a motility buffer without

    2-NBDG. Approximately 107 bacteria/ml were taken into the capillary by con-

    tacting the pellet (as described earlier) with the mouth of the capillary. Then, the

    capillary ends were sealed with wax. A schematic diagram of the micro-capillary

    experimental setup is shown in Figure 2.5. Using a microscope with an 4X (nu-

    merical aperture = 0.13 ) objective lens, the fluorescence intensity of 2-NBDG was

    measured over 0 < x < 1500 m as a function of time shown in Figure 2.6(A)

    and the results are shown in Figure 2.6 (B). It is noticed that within the first

    1.5 min, the intensity profiles attain the steady state and there is a negligible

    change in the concentration profile up to 30 min. These profiles were robust and

    24

  • easily reproducible and were effected neither by the cell movements nor by the

    consumption of 2-NBDG by this strain. Further, for the E. coli chemotaxis to

    MeAsp/serine, MeAsp/serine was used instead of 2-NBDG. It is assumed that the

    gradients obtained with MeAsp/serine is identical to 2-NBDG, since the molecular

    weight of both are approximately similar and so the diffusivity in water would also

    be approximately the same. The experimental values of L0 and the established

    gradients (G) are given in Table 2.1. These conditions were also used in the model

    simulation.

    Table 2.1: Gradients and ligand concentration at x = 0 obtained through experi-

    ments.

    S.No Concentration (L0, M) at x = 0 cm Gradient (G, M/m)

    1 1.6 0.0016

    2 16 0.016

    3 160 0.16

    4 1600 1.6

    5 16000 16

    4 160000 160

    Quantification of chemotaxis

    Image analysis was used to quantify the movements of E.coli. This involves the

    analysis of moving objects in image sequences. In this section, specifications of

    the optical microscope, image processing and procedure for calculating the cell

    movements are discussed.

    Specifications of optical microscope: IX71 Inverted Microscope (Olympus, Japan)

    was adopted as the optical microscope. Images were taken using Evolution VF

    cooled monochrome camera (Media Cybernetics, Japan) in Bright-field (BF) il-

    lumination mode with magnification of 40X objective lens (numerical aperture =

    0.75).

    Image processing procedure: Image-Pro Plus 6.0 image analysis program was

    used to locate E.coli in each frame and follow its motion in subsequent frames.

    25

  • Figure 2.5: Micro-capillary experimental setup for establishing MeAsp/serine gra-

    dients. x = 0 is located at the end of the pellet with a ligand concentration, L0.

    The experimental values of L0 and the established gradients (G) are given in Table

    2.1 and these conditions were used in the model simulations.

    26

  • Figure 2.6: (A) A magnified view of the micro-capillary set-up (B) Stable linear

    gradients; Measured concentration profiles for two gradients: t = 1.5 min (black

    line) and t = 5 min (blue line) for G = 0.016 M/m, t = 1.5 min (red line), t = 5

    min (green line) and t =30 min (pink line) for G = 0.16 M/m respectively.

    This program uses a single stack TIFF image which contains 500 sequential TIFF

    images taken at an interval of 0.11 s. In every experiment two stack images were

    recorded at 500, 1000 and 1500 m. Note that the image captured using 40X

    magnifications encompassed a physical area measuring 160 120 m2.Procedure for calculating the cell movement: To minimize errors in finding the

    cell movements, the following two assumptions were made. First, the cells moving

    close to each other or overlapping cell movements were neglected. Secondly, the

    cells not in the field of view for the entire time series (minimum 1.1 s) were

    neglected.

    Using Matlab c, each stack image was thresholded so as to clearly distinguishcells from the image background. Then, with the help of the auto tracking option,

    the movement of cells along the gradient direction (x) was tracked in a sequence of

    images. The raw data containing spatial location of each cell as a function of time

    was exported to an Excel c file. The procedure for calculating the movement ofcell over a period of time was straight forward. Starting from an initial position

    in the first frame, the displacements between two consequent frames 0.11 s apart

    27

  • were determined. The tracking data with time was collected from six identical

    experiments repeated on different days to capture the variability, if any.

    2.2 Modeling Methods

    Model for the intracellular pathway

    A two state receptor model proposed by Barkai and Leibler (1997) was used to

    simulate the intracellular signaling pathway. The model considers the methyl

    accepting chemotaxis proteins (MCPs), CheA and CheW, as a single entity (re-

    ceptor complex) and assumes that these receptor complexes, whose concentration

    is denoted by T , exist in either an active (TA) or an inactive (T I) state. Let

    Ti represent the concentration of receptor complexes with i residues methylated

    and i(L) denote the probability that the receptor complex Ti is active when the

    concentration of chemoattractant is L. The receptor complex can be in one of five

    methylation states with i = 0, 1, 2, 3 or 4 methyl groups. The total concentration

    of active receptors is given by,

    TA =40

    i(L)Ti, (2.1)

    while the total concentration of inactive receptors is given by,

    T I =

    40

    (1 i(L))Ti. (2.2)

    The binding kinetic equation for active receptor complex is given by,

    TAF + L [TAL] (2.3)

    The total active receptor complex concentration TAT is given by,

    TAT TAF + [T

    AL] (2.4)

    where TAF is free (non-ligand bound) active receptor complex concentration and

    [TAL] is the ligand bound active receptor complex concentration respectively. The

    28

  • fraction of free active receptor complex concentration from the above equation is

    given by,

    TAFTAT

    =KL

    KL + L(2.5)

    where KL is the ligand dissociation constant. Similarly, the fraction of ligand

    bound receptor complex concentration is given by,

    [TAL]

    TAT=

    L

    KL + L(2.6)

    The total probability of the receptor complex being in active state is the sum of

    the probabilities of the ligand bound and non-ligand bound receptors being in

    active state and is given by,

    i(L) =Li L

    KL + L+

    0iKLKL + L

    (2.7)

    where the parameters are assigned the following numerical values, L0 = 0, L1 =

    0, L2 = 0.1, L3 = 0.5,

    L4 = 1,

    00 = 0,

    01 = 0.1,

    02 = 0.5,

    03 = 0.75 and

    04

    = 1. These values are taken from Morton-Firth et al. (1999) and are estimated

    from the free energy states of methylation and the ligand occupancy of receptor

    complex for MeAsp. The corresponding phosphorylation rate equations with the

    corresponding rate constants (Emonent and Cluzel, 2008) are given by,

    dApdt

    = 23.5(TA)A 100(Ap)Y 10(Ap)B (2.8)

    dYpdt

    = 100(AP )Y 30(Yp) (2.9)

    dBpdt

    = 10(Ap)B (Bp) (2.10)

    Here, A, AP , Y , Yp, B and Bp represent, respectively, the concentrations of CheA,

    phosphorylated CheA, CheY, phosphorylated CheY, CheB and phosphorylated

    CheB. Li and Hazelbauer (2004) have measured these chemotaxis protein concen-

    trations for wild type and are given by, A + Ap = 5.3 M, B + Bp = 0.28 M,

    Y + Yp = 9.7 M and CheR (R) = 0.16 M. The total receptor concentration

    29

  • (Tar+Tsr), T0 + T1 + T2 + T3 + T4 = 17 M and R = 0.16 M reflect the

    reported (Sourjik and Berg, 2004) findings that both Tsr and Tar participate in

    sensing aspartate.

    Barkai and Leibler (1997) model assumes that CheR (R) binds to the inac-

    tive receptors (T I) and the phosphorylated CheB (Bp) binds to the active recep-

    tors (TA). Assuming that, the methylation and demethylation reactions follows

    Michaelis - Menten kinetics, the rate of demethylation and methylation is given

    by, respectively,

    rB =kbBp

    KB + TA(2.11)

    rR =krR

    KR + T I(2.12)

    where, kb = 0.6 s1 and kR = 0.75 s

    1 are the rate constants and KB = 0.54 M

    and KR = 0.39 M are the Michaelis constants (Emonent and Cluzel, 2008) for

    receptor demethylation and methylation, respectively.

    The rate of methylation is proportional to the concentration of inactive recep-

    tors (1-i(L))Ti, and the rate of demethylation is proportional to the concentra-

    tion of active receptors i(L) Ti. For the receptor Ti, the rate of demethylation

    is rB i(L) Ti and the rate of methylation is rB (1 i(L)) Ti, the mass balanceequations for the corresponding receptor can be given by,

    dT0dt

    = rR(1 0(L))T0 + rB1(L)T1 (2.13)

    dT1dt

    = rR(1 1(L))T1 + rB2(L)T2 + rR(1 0(L))T0 rB1(L)T1 (2.14)

    dT2dt

    = rR(1 2(L))T2 + rB3(L)T3 + rR(1 1(L))T1 rB2(L)T2 (2.15)

    dT3dt

    = rR(1 3(L))T3 + rB4(L)T4 + rR(1 2(L))T2 rB3(L)T3 (2.16)

    dT4dt

    = rR(1 3(L))T3 rB4(L)T4 (2.17)

    Using Matlab c, we solved simultaneously the steady state phosphorylationreaction equations and the mass balance equations for the receptors.

    30

  • E. coli motion model

    The ligand concentration decides the protein CheY-P concentration which is the

    output of the signaling pathway. In vivo experimental studies (Cluzel et al.,

    2000) using fluorescence correlation spectroscopy have reported CW and switching

    frequency (F ) as a function of CheY-P concentration (Yp). Recall that the CW is

    the fraction of time spent in clock-wise rotation of the flagella while the switching

    frequency is the number of times the motor switched its direction of rotation per

    unit time. Further, Cluzel et al. (2000) used the Hill equation to describe the CW

    bias as a function of Yp,

    CW =Y np

    (Y np ) + (Kn)

    (2.18)

    where n is the Hill coefficient and K is the half saturation constant. Cluzel et al.

    (2000) reported n = 10 and K = 3.1 M for their single cell experiments in-

    dicating a highly ultra-sensitive response (Figure 1.4). In our experiments, the

    experimentally measured CW bias was used to set the value of K so as to yield

    a steady state value of Yp. It was also observed that the switching frequency F

    qualitatively behaves as F dCW/d Yp. (Cluzel et al., 2000) So, the inverse ofthe switching frequency gives the time period containing one change of direction

    of rotation. The time spent in a single tumble mode is given by,

    ttum =CW

    F(2.19)

    while the time spent in a single run mode, also referred to as the run time is,

    trun =1 CW

    F(2.20)

    Assuming a Poisson process, (Vladimirov et al., 2008) we can now obtain the

    probability that E. coli will switch from tumble to run mode and vice versa. If

    the E. coli is in run mode, then the probability that it will switch to tumble mode

    in time dt is Pruntum = dt/trun while 1 (dt/trun) is the probability that it willcontinue in the run mode after each time step dt. Similarly, Ptumrun = dt/ttum.

    31

  • In the simulations, the E. coli starts from its initial position, x = 0, y = 0

    with the ligand concentration varying linearly with x, L(x) = L0 + Gx, where G

    is the gradient. At t = 0, the E. coli is made to run for time duration dt along

    x, after which Pruntum is determined. A number between 0 and 1 is randomly

    generated using a Matlab function that has a uniform probability in that range.

    If the number obtained is less than Pruntum, then the E. coli is made to tumble

    else it continues to run. In case of tumble, the position of the cell is held constant

    and a new direction of the motion is chosen from a gamma distribution of turn

    angles that is obtained independently from our experiments. The distribution

    yielded a mean turn angle of 71o 1.1 that was close to that observed by Bergand Brown (1972). In case of run, the cell is made to move at a constant run speed

    but at an angle chosen from a normal distribution with mean zero angle (about

    its previous angle) and a variance2Dr dt, where Dr is the rotational diffusivity

    and was determined independently from our experiments. Once the E. coli has

    tumbled or run, the above procedure is repeated at the new location for t = t+dt

    using the new local ligand concentration. The simulation was obtained for 1000

    cells and the mean properties were calculated for comparison with experiments.

    This model was used to simulate the motion of E. coli for varying -methyl-DL-

    aspartate (MeAsp) and L-serine (serine) concentrations/gradients. It is important

    to note that, MeAsp is a non-metabolized chemoattractant whereas serine is a

    metabolized (a less source of carbon) chemoattractant. Further, the model results

    are validated with the experimental results and are discussed in the next chapters.

    The following three pages presents a concise report on the presence of different

    chemotaxis receptors in the E. coli strain used in our experiments.

    32

  • Chapter 3

    Mathematical modeling and

    Experimental validation of

    chemotaxis under controlled

    gradients of methyl-aspartate in

    Escherichia coli

    3.1 Introduction

    In this chapter, we describe a novel technique to measure the chemoattractant

    gradients with respect to space and time in the absence of fluid flow. Unlike the

    previous studies, we track the motion of each cell and obtain the local drift velocity

    as a function of time and space for a broad range of concentrations and gradients

    of -methyl-DL-aspartate (MeAsp, a non-metabolized chemoattractant). We used

    an existing two state model (Barkai and Leibler, 1997) for the intracellular path-

    way and incorporate the extra-cellular influence such as ligand concentration and

    Brownian motion to predict the response for the experimental conditions. The

    statistics of single cell motions such as cell velocity, run angle, turn angle, tum-

    bling frequency, rotational diffusion, drift velocity and diffusivity are obtained

    36

  • independently. The predicted motion matched with the observation only when

    the response of the intracellular pathway was highly ultra-sensitive. The detailed

    comparison of the predictions with the observed behavior revealed bounds on the

    parameters describing the intracellular pathways. Further, our studies also show

    that oxygen plays a key role in the chemotaxis response and the response to a

    ligand cannot be analyzed in isolation to oxygen.

    Chemotaxis signaling model

    A two state receptor model proposed by Barkai and Leibler (1997) was used to

    simulate the intracellular signaling pathway. The model considers the methyl ac-

    cepting chemotaxis proteins (MCPs), CheA and CheW, as a single entity (receptor

    complex) and assumes that these receptor complexes, whose concentration is de-

    noted by T , exist in either an active (TA) or an inactive (T I) state. Further, a

    receptor is assumed to exist in one of the five methylation state. For a fixed value of

    the ligand concentration, the model solution yields the active and inactive receptor

    concentrations for each of the methylation states along with the concentrations of

    chemotaxis proteins in the phosphorylated and dephosphorylated states. Finally,

    the concentration of CheY-P decides the motion of E. coli. The model details of

    the intracellular signaling pathway and motion details are given in Materials and

    Methods (Chapter 2).

    3.2 Results

    Gradients of fluorescent glucose (2-(N-(7-nitro-benz-2-oxa-1,3-diazol-4-yl)amino)-

    2-deoxy-glucose, 2-NBDG) were established in the capillary and their variation

    with time were recorded. It was found that the gradient was established in the

    first 1.5 minutes after which the variation was negligible for almost 30 min (see Ma-

    terials and Methods). At x = 0, the concentration reaches a constant finite value

    within this short time beyond which (x > 0) a stable linear gradient was achieved.

    A wide range of gradients could be established by varying the concentration of the

    37

  • ligand in the liquid plug. It can be noted that the gradient experiments with E.

    coli were conducted for about 15 min after establishing the gradient. This implies

    that the stable gradients could be maintained during the period of the experiment

    (see Figure 2.6).

    The experiments were conducted with various gradients of MeAsp and with

    uniform concentrations in the absence of gradients, and the mean square displace-

    ment (MSD) of over 1000 cells from these experiments are presented as a function

    of time in Figure 3.1. Interestingly, all the measured points from various exper-

    iments collapse to give a unique trend, wherein the MSD increases quadratically

    up to t 3.8 s after which the increase is linear. The MSD was further usedto determine the translational diffusivity, ((x )2/2t) at large times (t > 3.8s) yielding a value of 110 26.22 m2/s, irrespective of the external conditions.Here, is the mean displacement in time t obtained using Gaussian distribution.

    The measured angular displacement data was fit to a Gaussian distribution and

    was used to calculate the rotational diffusivity, given by Dr = 2/2t where is the angular displacement and the angular brackets represent ensemble average.

    Table 3.1 presents the various parameters related to E. coli motion obtained from

    our experiments and the measured values are close to those reported in previous

    studies.

    Table 3.1: Comparison of chemotactic parameters in E. coli with those reported in

    literature. (Berg and Brown, 1972; Liu and Papadopoulos, 1996; Berg and Turner,

    1990; Ahmed and Stocker, 2008; Kalinin et al., 2009) These values were obtained

    by averaging ( standard deviation) over all experiments.

    Description Present study Reported

    Average run speed (v, m/s) 18.6 2.2 14.2 3.4Mean tumble angle (degrees) 71.80 1.1 62 26

    Rotational diffusivity (Dr, rad2/s) 0.32 0.07 0.06 - 0.28

    Steady state CW bias (CWss) 0.16 0.05 0.1Cell diffusivity (D, m2/s) 110 26.22 125 - 360

    38

  • Figure 3.1: Mean square displacements (MSD standard error) about the meanvalue (from Gaussian fit) as a function of time. MSD was obtained using data

    from all experiments including those with and without gradients. The solid line

    represents the linearity obtained for a diffusive regime. The diffusive regime occurs

    beyond 3.8 s and is shown by a dotted arrow.

    Next, the measured drift velocity, in the presence and the absence of MeAsp,

    was recorded at three different locations along the capillary and is presented in Fig-

    ure 3.2. In the absence of MeAsp, independent experiments were conducted with

    only motility buffer that was saturated with air and with pure oxygen. Figure 3.2

    shows that in the absence of MeAsp the cells show significant drift velocities (u0)

    initially and the velocity drops steadily beyond 1000 m from the entrance. The

    high concentration of the cells in the pellet would have rendered the interstitial

    fluid devoid of oxygen. Consequently, the cells respond to an oxygen gradient

    established when the cells are brought in contact with the motility buffer in the

    capillary. However, a significant decrease in drift velocity with increasing distance

    was observed beyond 5000 m and was probably due to a lack of endogenous en-

    ergy source since the cells become non-motile. Further, the run speeds did not

    vary up to 2500 m, suggesting that the endogenous energy source was not limit-

    ing for x < 2500 m. Figure 3.2 also includes the drift velocity when the buffer is

    saturated with oxygen. The velocities were higher by about 1.5 times everywhere

    than that for the normal case although the cells again became non-motile for dis-

    tances beyond 5000 m. The higher velocities could be attributed to the higher

    39

  • gradients of oxygen sensed by the cells.

    All experiments in the presence of MeAsp were performed at normal levels of

    oxygen (saturated with air). Experiments were conducted to first quantify the

    chemotactic response in the absence of gradient to uniform MeAsp concentrations

    of 100 (N) and 10000 M () and the results are also presented in Figure 3.2. The

    measured drift velocities for both the concentrations were similar to that observed

    in the absence of MeAsp. This clearly indicates that cells respond solely to oxygen

    in the absence of MeAsp gradients. To study the chemotactic behavior of E. coli,

    various gradients of MeAsp were established in the capillary. Figure 3.3 presents

    the measured average run speed as a function of the gradient of MeAsp at three

    different locations in the capillary. The plot also includes measurements in the

    absence of MeAsp (indicated by the arrow). It can be noted that the average run

    speeds are in the range of 16-21 m/s irrespective of the gradient or the location.

    An average value of 18 m/s was used in the model simulations.

    Figure 3.2: Measured drift velocity (u0 standard error) as a function of distancein the absence of aspartate using normal buffer (), 100 M aspartate in normalbuffer (N), and 10000 M aspartate in normal buffer (). The solid line represents

    the average trend for the three experiments. The empty circle () represents thedrift velocity when the buffer is saturated with oxygen in the absence of aspartate.

    The measured drift velocities contain the influence of both oxygen and MeAsp

    gradients. In order to separate the influence of oxygen gradient on the total drift

    velocity, we subtracted (uo) from measured drift velocity values assuming alge-

    40

  • Figure 3.3: Average run speeds ( standard error) as a function of aspartategradients at: 500 m (+), 1000 m (), and 1500 m (). The solid line representsthe average value for the entire data set. The arrow indicates the average run speed

    of 18 m/s obtained in the absence of MeAsp.

    braic addition of the responses. (Strauss et al., 1995) Thus, the drift velocity due

    to MeAsp alone is given by, uasp = u uo. Note that this procedure not onlyeliminates the effect of oxygen but also the non-chemotactic drift induced by the

    no flux condition on the plug side. The measured drift velocities at fixed locations

    in the capillary for four different gradients with their corresponding initial con-

    centrations are shown in Figure 3.4. For the lowest gradient (G = 0.016 M/m),

    the response was minimal indicating that the E. coli had completely adapted

    even at x = 500 m. However, on increasing the gradient to G = 0.08 M/m, a

    drift velocity of 1.4 m/s was measured at 500 m which monotonically decreased

    to 0.8 m/s at 1500 m demonstrating adaptation. On increasing the gradient

    further, a similar profile but with higher initial drift velocities was observed. How-

    ever, at the highest gradient tested (G = 1.6 M/m), the measured velocities at

    500 and 1000 m matched with that for G = 0.08 M/m, but was significantly

    lower at 1500 m indicating early adaptation. The model was simulated for the

    above mentioned gradients and the predictions are compared with experiments in

    Figure 3.4. Note that the simulations do not incorporate the effects of oxygen.

    Further, simulations were run in the absence of attractants to determine the non-

    chemotactic drift velocity as a function of distance. The non-chemotactic drift

    41

  • velocity decreases gradually from 0.35 m/s at x = 0 to a negligible value at x =

    1500 m (results not shown). These values were subtracted from the values ob-

    tained for gradients to compare with experimental data, where the effects of both

    oxygen and non-chemotactic drift have been eliminated. The spatial variation of

    the predicted drift velocities matched reasonably well with the measurements for

    all four gradients. Specifically, the model captures the increase in drift velocities

    with gradient close to the plug and the subsequent monotonic decrease with dis-

    tance. Further, the measured drift velocity indicates that the cells adapt faster at

    higher gradients, again in agreement with experiments.

    Figure 3.4: Drift velocity (uasp standard error) as a function of distance forvarious gradients of aspartate (G) and initial ligand concentration (L0) (A) G =

    0.016 M/m and L0 = 16 M; (B) G = 0.08 M/m and L0 = 80 M; (C) G =

    0.16 M/m and L0 = 160 M; (D) G = 1.6 M/m and L0 = 1600 M. represents data from experiments and solid line represents model prediction.

    42

  • The model was used to obtain the drift velocities at different spatial locations

    along the capillary as a function of varying gradients (see Figure 3.5 (A)). Below

    a gradient of 0.01 M/m, the drift velocity is negligible and does not vary with

    distance indicating no response. With increasing gradients, the drift velocity ini-

    tially increases at all four locations reaching a maximum after which the velocity

    drops. A maximum drift velocity of 1.8 m/s at 100 m was observed close to a

    gradient of 0.1 M/m. Note that the peak value of drift velocity also decreases

    with spatial distance. A negligible drift velocity is obtained for gradients greater

    than 10 M/m but for locations greater than 3000 m due to adaptation.

    The model also predicted the normalized CW bias with time for measured

    gradients and the corresponding normalized value (with respect to the steady

    state value) of the fraction of active receptors, TA/TAss (Figure 3.5 (B) and (C),

    respectively). Note that TA decides the concentration of CheY-P and therefore the

    CW. Perfect adaptation to a new environment requires that the TA and therefore

    the CW return to their steady state values after the initial transients due to change

    in the environment. At the lowest gradient tested in our experiments, TA drops

    by a small value after which it regains the steady state value suggesting quick

    adaptation. Consequently, this small change in TA leads to a step drop of CW

    bias drops at x = 0 after which it attains the steady state value within a short time

    (200 s). This indicates that the run length is higher than the steady state value

    only for a short distance beyond which there is negligible response. At the higher

    gradient of 0.08 M/m, the fraction of active receptors decrease further leading

    to a slower recovery of the normalized CW bias to its steady state value. The plot

    indicates that the cells do not completely adapt even after 2000 s. On increasing

    the gradient further, a similar initial response is observed although the normalized

    CW bias recovers faster to the steady state value. For very high gradients (G =

    1.6 M/m), the bias takes a negligible value up to 500 s indicating very large run

    lengths. However, the recovery to the steady state value is achieved within 1500

    s, indicating faster adaptation at higher gradients. Note that we were unable to

    measure t