3D Computational Modeling and Simulation o

Embed Size (px)

Citation preview

  • 7/31/2019 3D Computational Modeling and Simulation o

    1/16

    Computers in Biology and Medicine 38 (2008) 738753www.intl.elsevierhealth.com/journals/cobm

    3D computational modeling and simulation ofleukocyte rolling adhesion and deformation

    Vijay Pappu, Prosenjit Bagchi

    Department of Mechanical and Aerospace Engineering, Rutgers University, The State University of New Jersey, 98 Brett Road, Piscataway, NJ 08854, USA

    Received 2 November 2007; accepted 3 April 2008

    AbstractA 3D computational fluid dynamic (CFD) model is presented to simulate transient rolling adhesion and deformation of leukocytes over a

    P-selectin coated surface in shear flow. The computational model is based on immersed boundary method for cell deformation, and stochastic

    Monte Carlo simulation for receptor/ligand interaction. The model is shown to predict the characteristic stop-and-go motion of rolling

    leukocytes. Here we examine the effect of cell deformation, shear rate, and microvilli distribution on the rolling characteristics. Comparison

    with experimental measurements is presented throughout the article. We observe that compliant cells roll more stably, and have longer pause

    times due to reduced bond force and increased bond lifetime. Microvilli presentation is shown to affect rolling characteristics by altering the

    step size, but not pause times. Our simulations predict a significant sideway motion of the cell arising purely due to receptor/ligand interaction,

    and discrete nature of microvilli distribution. Adhesion is seen to occur via multiple tethers, each of which forms multiple selectin bonds, but

    often one tether is sufficient to support rolling. The adhesion force is concentrated in only 13 tethered microvilli in the rear-most part of a

    cell. We also observe that the number of selectin bonds that hold the cell effectively against hydrodynamic shear is significantly less than the

    total adhesion bonds formed between a cell and the substrate. The force loading on individual microvillus and selectin bond is not continuous,

    rather occurs in steps. Further, we find that the peak force on a tethered microvillus is much higher than that measured to cause tether extrusion.

    2008 Elsevier Ltd. All rights reserved.

    Keywords: Computational fluid dynamics; Immersed boundary method; Fluid structure interaction; Microcirculation; Leukocyte; Selectin; Receptors; Microvilli

    1. Introduction

    Adhesion of circulating leukocytes to vascular endothelium

    is a key event in inflammatory response [1]. The process,

    often called adhesion cascade, involves multiple steps that begin

    with initial arrest or tethering of leukocytes to the endothelium,

    followed by slow rolling of the cells. Subsequently, leukocytesfirmly adhere and spread over the endothelium, and then trans-

    migrate to the sites of inflammation. Extensive studies in the

    past have shown that the tethering and rolling are mediated

    by three types of adhesion molecules, P-, E-, and L-selectins,

    which bind to their respective ligands with high affinity [24].

    P-selectin-glycoprotein-ligand-1 (PSGL-1) is a common lig-

    and that is known to bind to all three selectins. Flow chamber

    Corresponding author. Tel.: +1 7324453656.

    E-mail address: [email protected] (P. Bagchi).

    0010-4825/$- see front matter 2008 Elsevier Ltd. All rights reserved.

    doi:10.1016/j.compbiomed.2008.04.002

    studies have also shown tethering and rolling of leukocytes over

    selectin-coated surfaces [58].

    Analysis of leukocyte trajectory revealed that a rolling cell

    does not flow continuously, but rather it moves in a stop-and-

    go manner [4,5,9,10] due to random formation and breakage of

    receptor/ligand bonds. Selectin bonds are known to have high

    association and dissociation rates. A threshold shear is requiredto mediate rolling [11,12]. However, above the threshold shear,

    cells roll stably with relatively less variation in rolling veloc-

    ity as shear rate increases 20 folds [8,10]. Chen and Springer

    [10] hypothesized an automatic braking system in which

    receptor/ligand bonds increase with increasing shear to stabi-

    lize rolling.

    Microvilli, which are protrusions from the cell surface, also

    plays a critical role in rolling [1315]. L-selectins and PSGL-1

    are concentrated on the tips of microvilli. When a leukocyte

    comes in proximity to the endothelium, bond formation is

    http://www.intl.elsevierhealth.com/journals/cobmmailto:[email protected]:[email protected]://www.intl.elsevierhealth.com/journals/cobm
  • 7/31/2019 3D Computational Modeling and Simulation o

    2/16

    V. Pappu, P. Bagchi / Computers in Biology and Medicine 38 (2008) 738753 739

    facilitated due to high accessibility of microvilli tips. Concen-

    tration of L-selectins and PSGL-1 on microvilli tips also sug-

    gests that selectin bonds are likely to form in clusters, rather

    than distribute uniformly over the entire cell/substrate contact

    zone.

    While circulating leukocytes maintain a spherical shape,

    rolling leukocytes are known to deform to tear-drop shapesat higher shear [8,16,17]. Firrell and Lipowsky [16] ob-

    served nearly 140% increase of WBC length as the shear rate

    increased from 50 to 800 s1. Cell deformability may affect

    rolling in several ways [9,13,18]. Upon initial tethering, the

    cell/substrate contact area becomes flat allowing formation of

    newer microvilli tethers to further stabilize the rolling. Adhe-

    sion via multiple microvilli would reduce the force on individ-

    ual selectin bonds and prolong bond lifetime. Cell deformation

    may also reduce the hydrodynamic drag, and hence the bond

    force. Comparison of rolling characteristics of ligand-coated

    microspheres, and fixed and normal leukocytes suggested that

    normal cells roll more smoothly with longer pauses compared

    to microspheres or fixed leukocytes [7,9]. The bond force was

    also estimated to be significantly lower in case of an untreated

    leukocyte than that on a microsphere.

    Among several computational models, the adhesive dynam-

    ics simulation (ADS) pioneered by Hammer and co-workers

    made a significant contribution to theoretical understanding of

    leukocyte rolling [1922]. In ADS, leukocytes are modeled

    as rigid spheres, and the receptor/ligand interaction is sim-

    ulated by stochastic Monte Carlo simulation. Recent works

    by the same group have incorporated microvilli deformation

    within the framework of ADS [23]. Deformation of an adher-

    ent leukocyte was considered in two-dimensions by Dong and

    co-investigators by modeling a leukocyte as a viscous liquiddrop surrounded by an elastic ring [8,24,25]. NDri et al. [26]

    modeled leukocytes as 2D compound liquid drops to study the

    effect of cell nucleus on deformation. The role of viscoelas-

    ticity, and microvilli extension during leukocyte adhesion and

    rolling were considered in a recent 3D model developed by

    Khismatullin and Truskey [27,28]. However, the characteris-

    tic stop-and-go motion of a rolling leukocyte, as observed in

    vitro and in vivo, was not reported by Dong and co-workers,

    NDri et al., and Khismatullin and Truskey.

    Recently, Jadhav et al. [29] developed a 3D model for rolling

    leukocytes by coupling cell deformation with stochastic simu-

    lation of receptor/ligand interaction. Their model was able toreplicate the stop-and-go motion of leukocytes. Such com-

    putational tools can be used to gain deeper insights into the

    biomechanics of cell rolling and adhesion. In this article, we

    present a similar 3D computational fluid dynamic (CFD) model

    to simulate rolling adhesion of deformable leukocytes over a

    P-selectin coated surface in a shear flow. Our computational

    model is based on the immersed boundary method (IBM) for

    cell deformation, and stochastic Monte Carlo simulation of

    receptor/ligand interaction. Our model has been able to predict

    the characteristic stop-and-go motion of rolling leukocytes.

    Using the model, we address three specific questions in this

    article: (i) How does deformation affect the rolling character-

    istics of cells? (ii) How does distribution of microvilli affect

    cell rolling? (iii) How is the adhesion force distributed in the

    cellsubstrate contact zone?

    2. Computational methodology

    The flow configuration is described in Fig. 1. The adhesive

    rolling motion of deformable leukocytes in shear flow over aP-selectin coated planar surface is considered. Shear rate is var-

    ied as 100, 300 and 500 s1. The initial shape of the leuko-

    cyte is spherical with diameter 8 m. Microvilli, each of length

    350 nm, are distributed randomly over the cell surface. Bonds

    are allowed to form in the microvilli tips. We consider two dif-

    ferent microvilli distributions, Nmv =21 and 155, where Nmv is

    the number of total microvilli. We consider deformation of the

    cell, but not of microvilli. Further parameters of the problem

    are listed in Table 1.

    The computational modeling is based on the IBM [3032]

    which is particularly suitable for the present study, as the leuko-

    cyte is modeled as a compound liquid drop surrounded by hy-

    perelastic membrane (discussed later). In IBM, a single set of

    equations is used to solve the fluid motion interior and exterior

    of the cell. The fluid motion is governed by the continuity and

    NavierStokes equations as

    u = 0, (1)

    ju

    jt+ u u

    = p + (u + (u)T), (2)

    where u is the fluid velocity, is the density, p is the pressure,

    and is the viscosity. The cell surface is then recognized by a

    source-like term F added to the r.h.s. of Eq. (2). For a rolling

    cell in shear flow, the force on the cell surface can arise fromtwo contributions: fe the elastic force due to cell deformation,

    and fa the adhesive force due to bond formation between the

    cell and the substrate. The source term F is related to fe and

    fa as

    F(x, t ) =

    jS

    [fe(x, t) + fa(x

    , t )](x x) dx. (3)

    Here x is a point in the flow domain, x is a point on the

    cell surface jS, and is the delta function which vanishes

    everywhere except at the membrane. Models for computation

    of fe and fa are described later.

    The NavierStokes equations are discretized on a fixed

    Eulerian grid, and the cellplasma interface is tracked in a

    Lagrangian manner by a set of marker points distributed on the

    cell surface (Fig. 1). The equations are first solved to obtain

    the fluid velocity and pressure in the entire flow domain. The

    velocity of the cell membrane is then obtained by interpolating

    the fluid velocity as

    u(x) =

    S

    u(x)(x x) dx, (4)

    where S denotes the entire flow domain. Cells are then

    advected as

    dx

    dt = u(x

    ). (5)

  • 7/31/2019 3D Computational Modeling and Simulation o

    3/16

    740 V. Pappu, P. Bagchi / Computers in Biology and Medicine 38 (2008) 738753

    leukocyte

    Nucleus

    shear flow

    Eulerian grid

    Y

    X

    X

    Y

    Z

    cell surface

    Microvillus

    Selectin bondsPlate

    Fig. 1. (A) Schematic of leukocyte rolling over a selectin-coated surface in a shear flow. A 2D slice is considered to show the Eulerian grid used to discretize

    the flow domain. (B) Actual 3D cell showing the Lagrangian mesh on the surface. The discrete locations of microvilli are shown by . (C) Schematic of one

    microvillus and selectin bonds.

    Table 1

    Parameter values used in simulations

    Parameter Value Source

    Shear rate 100, 300, 500 s1

    Leukocyte diameter 8m

    Membrane stiffness (Eh) 2.6, 0.9, 0.3 dyn/cm [29]

    Microvillus length 0.35m [14,43]

    Number of microvillus (Nmv) 21, 155 [14]

    Number of ligands 50/microvilli [29]

    Receptor site density 144/m2 [6]

    Selectin bond length (l0) 0.1m [48]

    Spring constant (kb) 1 pN/nm [48]

    Transition spring constant (kts) 0.99 pN/nm [6]

    Unstressed forward rate (k0f ) 3.30 s1 [49]

    Unstressed reverse rate (k0r ) 3.7 s1 [6]

    Distribution of the forces from cell surface to the surround-

    ing fluid grid (Eq. (3)), and interpolation of the fluid veloc-

    ity onto the cell surface (Eq. (4)) involve a 3D delta function

    which is constructed by multiplying three 1D delta functions as

    (x x) = (x x)(y y)(z z). For numerical imple-

    mentation, a discretized representation of the -function is used

    as [31]

    D(x x) =1

    643

    3i=1

    1 + cos

    2(xi x

    i )

    for |xi x

    i |2, i = 1, 2, 3,

    D(x x) = 0 otherwise, (6)

    where is the Eulerian grid size. The above representation

    approaches the analytical delta function as approaches zero.

    The discrete delta function is so constructed that the distribu-

    tion of the surface forces, or the interpolation for the surfacevelocity, is performed over a sphere of diameter equal to four

    Eulerian grid points surrounding each Lagrangian node. In dis-

    crete form, the integrals in Eqs. (3) and (4) are written as

    F(xj) = i D(xj xi )f(x

    i ), (7)

    u(xi ) = jD(xj xi )u(xj), (8)

    where i and j represent the Lagrangian and Eulerian grid points,

    respectively.

    We model leukocytes as compound liquid drops surrounded

    by thin hyperelastic membranes. The membrane is assumed to

    follow the neo-Hookean law. The strain energy function for themembrane is then given by

    W =Eh

    6(21 +

    22 +

    21

    22 3), (9)

    where E is the modulus of elasticity, h is the membrane thick-

    ness, and 1 and 2 are the principal stretch ratios. Note that

    the neo-Hookean model does not strictly represent a leukocyte

    membrane. Models which closely resemble large deformation

    of leukocytes during micropipette aspiration exist in the litera-

    ture [33,34], and can be implemented in IBM. We choose the

    neo-Hookean model as it is easy to implement, yet can predict

    cell deformation under rolling conditions [29]. Three values of

  • 7/31/2019 3D Computational Modeling and Simulation o

    4/16

    V. Pappu, P. Bagchi / Computers in Biology and Medicine 38 (2008) 738753 741

    Eh, 0.3, 0.9, and 2.6 dyn/cm, are considered in the simulations

    following Jadhav et al. [29].

    The cell surface is discretized using triangular elements

    (Fig. 1). We use a finite element model to compute the elastic

    force generated on the cell surface due to deformation [35].

    In this model, the elastic force fe is obtained at three nodes of

    each element by differentiating the strain energy function Wwith respect to the nodal displacement v as

    fe =jW

    j1

    j1

    jv+

    jW

    j2

    j2

    jv. (10)

    The main idea is that a general 3D deformation of the membrane

    can be reduced to a 2D problem by assuming that individual

    triangular element on the membrane remains flat even after

    deformation, and that the membrane force remains invariant

    under a rigid body rotation. This assumption still allows large

    deformation of a leukocyte in the model. The resultant force feat a membrane node is the vector sum of the forces exerted by

    all elements surrounding that node.Formation of receptor/ligand bonds between a leukocyte and

    the substrate is simulated using stochastic Monte Carlo method

    [19,29]. Bonds are assumed to behave as stretched springs under

    force loading following a Hookean model [36]. The probability

    of formation of a new bond, and that of breakage of an existing

    bond, in a time interval t, are given by

    Pf = 1 exp(kft ) (11)

    and

    Pr = 1 exp(krt ), (12)

    respectively, where kf and kr are the forward and reverse reac-tion rates which are computed as

    kf = k0f exp

    kts(l l0)2

    2KBT

    (13)

    and

    kr = k0r exp

    (kb kts)(l l0)

    2

    2KBT

    , (14)

    where k0f and k0r are the unstressed reaction rates, kb is the

    spring constant, kts is the transition state spring constant, l and

    l0 are the stretched and unstretched lengths of a bond, KB isthe Boltzmann constant, and T is the absolute temperature.

    Values of the parameters are given in Table 1. At a given time

    instance, two random numbers N1 and N2, between 0 and 1,

    are generated. A new bond is allowed to form if Pf > N1, and

    an existing bond is allowed to break if Pr > N2 [19,29]. Force

    in each bond fb is then obtained as

    fb = kb(l l0). (15)

    The adhesion force fa is the vector sum of the forces arising

    from all bonds formed in a microvillus tip.

    The NavierStokes equations are discretized spatially using

    a second-order finite difference scheme, and temporally using

    a two-step time-split scheme. In this method the momentum

    equation is split into an advectiondiffusion equation and a

    Poisson equation for the pressure. The body-force term is re-

    tained in the advectiondiffusion equation. The nonlinear terms

    are treated explicitly using a second-order AdamsBashforth

    scheme, and the viscous terms are treated semi-implicitly us-

    ing the second-order CrankNicholson scheme. The resultinglinear equations are inverted using an ADI (alternating direc-

    tion implicit) scheme to yield a predicted velocity field. The

    Poisson equation is then solved to obtain pressure at the next

    time level. Using the new pressure, the velocity field is cor-

    rected so that it satisfies the divergence-free condition. The cell

    surface is also advected using a second-order AdamsBashforth

    scheme. Details of the time-step scheme are given in Bagchi

    and Balachandar [37]. Computational domain is a rectangular

    box with the longest axis aligned parallel to the flow direction

    (X). The domain is assumed periodic in the X and Z directions.

    Typical Eulerian resolution used in the flow solver is 320 points

    in X, and 120 points each in the Y and Z directions. Typical

    Lagrangian mesh used to discretize the cell surface consists of

    1280 triangular elements.

    3. Validation of numerical method

    We first validate our IBM code against published results on

    cell deformation. For this purpose, we consider deformation

    of a spherical capsule as a model cell placed in a linear shear

    flow as shown in Fig. 2. A capsule is a liquid drop surrounded

    by an elastic membrane. In a shear flow a capsule deforms into

    an ellipsoidal shape. The deformation can be expressed using a

    dimensionless parameter D =(LB)/(L +B), where L and B

    are the major and minor axis of the ellipsoid in the plane of theshear (Fig. 2A). Time-history of D for various dimensionless

    shear rate is shown in Fig. 2B and compared with that obtained

    by boundary integral simulation by Ramanujan and Pozrikidis

    [38]. Excellent agreement between the two simulations is

    observed.

    Sensitivity of our results to the Eulerian and Lagrangian res-

    olutions is shown in Fig. 2C (and, in inset 2D) by considering

    three test simulations at different resolutions: (i) 803 Eulerian

    grids and 1280 Lagrangian elements, (ii) 1203 Eulerian grids

    and 1280 Lagrangian elements, and (iii) 1203 Eulerian grids

    and 5120 Lagrangian elements. No significant difference is

    observed between the three test cases.We also keep track of the cell volume during the simulations.

    The change in the cell volume is less than 0.1% from its

    initial volume. The projection method used here for flow solver

    satisfies the mass (or, volume) conservation up to 1014 at

    every grid point in the computational domain.

    4. Results

    Snapshots of a rolling leukocyte obtained from a simulation

    at 500 s1 and Eh = 2.6dyn/cm are presented in Fig. 3. Shown

    here is a sequence of initial tethering, cell deformation, and

    tether breakage. In Fig. 3A, initial bonds are just formed on

    microvillus 1. The cell shape is nearly spherical at this time.

  • 7/31/2019 3D Computational Modeling and Simulation o

    5/16

    742 V. Pappu, P. Bagchi / Computers in Biology and Medicine 38 (2008) 738753

    B

    L

    Undeformed Capsule

    Deformed Capsule

    00 2 4 6 8

    0.2

    0.4

    0.6

    0.8

    0.2

    0.025

    0.05

    0.1

    0 2 4 6 80

    0.1

    0.2

    0.3

    0.4

    0.5

    3.5 4 4.5

    0.38

    0.39

    0.4

    Fig. 2. Validation of IBM code. (A) Schematic of a spherical capsule deforming in shear flow. (B) Deformation index D versus time. : present results;

    Ramanujan and Pozrikidis [38]. Results are shown for various values of dimensionless parameter a/2Eh where a is cell diameter and is shear rate.

    (C) Grid resolution test. , 803 Eulerian points and 1280 Lagrangian elements; 1203 Eulerian points and 1280 Lagrangian elements;

    1203 Eulerian points and 5120 Lagrangian elements. (D) Inset showing the closeup.

    Upon initial tethering, the cell rotates clockwise due to the fluid

    torque (Fig. 3B). The rolling motion temporarily stops, and

    the cell deforms to make a flat contact area with the substrate

    (Fig. 3C). As a result, more microvilli become available for

    bond formation, such as microvillus 2 in Fig. 3C. Sub-

    sequently, microvillus 1 breaks, cell rolling commences

    (Fig. 3D), and the contact area decreases. The cell is eventually

    tethered via microvillus 2 (Fig. 3E). The contact area in-

    creases again, and microvilli 3 and 4 also become available

    for bond formation. Fig. 3F shows the breakage of microvillus

    2 followed by cell rolling. In Fig. 3G, the cell is shown to

    tether again by microvillus 4.

    The rolling sequence of a more compliant cell (Eh =

    0.3dyn/cm) is presented in Fig. 4. Fig. 4A shows the initial

    arrest of the cell by tethering of microvillus 1, followed by

    deformation of the cell into a tear-drop shape in Figs. 4BC.

    The contact area increases significantly, and four microvilli

    become available for bond formation. However, the cell is

    anchored only by microvillus 1 located in the sharp corner

    formed at the trailing edge. In Fig. 4D, microvillus 1 breaks,

    and the rear end of the cell retracts. Figs. 4E and F show

    formation of a new tether via microvillus 2, followed by cell

    spreading.

    Figs.3 and 4 show that during rolling adhesion, the cell shape

    deviates significantly from its spherical shape. Average contact

    area is shown in Fig. 5A. It increases with increasing shear

    rate and decreasing membrane stiffness. The present results

    show excellent agreement with the in vivo measurements by

    Firrell and Lipowsky [16]. We also compute a dimensionless

    parameter called deformation index, L/H, where L is the end-

    to-end length along the flow direction, and H is the height of the

    cell (Fig. 5B). The ratio increases with increasing shear rate and

    decreasing membrane stiffness. The qualitative trend and the

    range of values are in agreement with the in vivo measurements

    of Damiano et al. [17] and Firrell and Lipowsky [16] (shown in

    Fig. 5B), and 3D computational modeling of Jadhav et al. [29].

    Next we consider the trajectory of a rolling leukocyte.

    Fig. 6 shows the axial (x) displacement of the cell w.r.t. time.

    Cell motion is characterized by a series of steps during which

    the cell rolls, and pauses during which the cell is adherent. Our

    results show the role of shear-rate and membrane stiffness in

    modulating the steps and pauses. Effect of varying shear rate at

    a constant Eh = 0.3dyn/cm is illustrated by plots AC which

    show shorter pause times and more frequent steps with increas-

    ing shear. Effect of varying membrane stiffness at a constant

    shear is illustrated by plots CE, which show decreasing step

    sizes and longer pause times with decreasing stiffness. The

    instantaneous rolling velocity corresponding to the trajectories

    is also shown in the same figure (right panel). The stochastic

    nature of the cell motion is evident here by fluctuations in cell

  • 7/31/2019 3D Computational Modeling and Simulation o

    6/16

    V. Pappu, P. Bagchi / Computers in Biology and Medicine 38 (2008) 738753 743

    1

    A

    1

    B

    X

    Y

    Z

    YX

    Z

    1

    2

    C

    2 D

    2

    3

    4E

    4F

    4 G

    Fig. 3. Sequence of a rolling leukocyte at 500 s1 shear rate, Eh = 2.6dyn/cm, and Nmv = 21. The shear flow and the cell movement are from left to right.

    The left panel shows sideview, and the right panel shows bottomview. Lagrangian mesh on the cell surface is also shown. In the bottomview, microvilli

    forming bonds are marked by numbers 1, 2 etc.

    velocity. The fluctuations increase with increasing shear rate

    and membrane stiffness implying that a leukocyte rolls less

    stably at higher shear rate and membrane stiffness.

    The average pause time of a rolling leukocyte obtained from

    our simulations is shown in Fig. 7A. The average pause time

    varies between 0.04 and 0.6 s. It strongly depends on shear

    rate, and decreases with increasing shear rate. The pause time

    depends strongly on membrane compliance at low shear rate,

    but weakly at higher shear rate. It increases with increasing

    membrane compliance. In Fig. 7A, we compare our computed

    pause times with the experimental data of Smith et al. [6],

    and find reasonable agreement. Our simulations also predict

    that the average pause time does not depend on the microvilli

    distribution given by Nmv (not shown).

    The average step size of a rolling leukocyte obtained from

    our simulations is shown in Fig. 7B. The step size strongly

  • 7/31/2019 3D Computational Modeling and Simulation o

    7/16

    744 V. Pappu, P. Bagchi / Computers in Biology and Medicine 38 (2008) 738753

    1

    X

    Y

    Z

    1

    1

    2

    3

    4 C

    1

    2

    B

    XY

    Z

    1

    2

    A

    1

    2

    3

    4D

    2

    3

    4E

    2

    3

    4F

    Fig. 4. Same as in Fig. 3 except Eh = 0.3dyn/cm.

    depends on Nmv. For Nmv = 155, the average step size is in the

    range 0.2.0.6m, and it does not strongly depend on shear rate

    and membrane stiffness. For Nmv = 21, the step size increases

    significantly due to more sparse distribution of microvilli. The

    step size also decreases with decreasing membrane stiffness

    and increasing shear rate. For Nmv = 21, the step size depends

    strongly on the membrane stiffness at low shear, but weakly on

    shear rate at low stiffness.

  • 7/31/2019 3D Computational Modeling and Simulation o

    8/16

    V. Pappu, P. Bagchi / Computers in Biology and Medicine 38 (2008) 738753 745

    shear rate (1/s)

    100 200 300 400 50010

    20

    30

    40

    contactaream

    2

    shear rate (1/s)

    deformationindex(L/H)

    100 200 300 400 5001

    1.2

    1.4

    1.6

    1.8

    2

    Firrell &Lipowsky

    Fig. 5. Average contact area (A) and deformation index (B). Dash lines are best fit curve obtained by Firrell and Lipowsky [16] based on in vivo measurements;

    solid lines are present results. Eh = 2.6dyn/cm, Eh = 0.9dyn/cm, Eh = 0.3dyn/cm.

    Average rolling velocity is shown in Fig. 8. It increases with

    increasing shear rate, increasing membrane stiffness, and de-

    creasing Nmv. At Nmv = 21, the average rolling velocity ranges

    from about 4 to 112 m/s, whereas at Nmv = 155, it ranges

    from 3 to 11 m/s. At higher Nmv, and lower shear, the rolling

    velocity does not change significantly with varying membrane

    stiffness; but it depends strongly on membrane stiffness at lower

    Nmv and higher shear. In the figure, we compare our results

    with three experimental measurements, Kim and Sarelius [4],

    Yago et al. [9], and Ramachandran et al. [39], which show rea-

    sonable agreement. Our results at Nmv = 21 are in agreement

    also with in vivo data (30.

    50m/s) of Firrell and Lipowsky[16].

    As a quantification of the stochastic nature of cell rolling,

    we compute RMS (root-mean-square) of instantaneous rolling

    velocity in Fig. 10A. The RMS of the axial velocity shows

    a strong dependence on shear rate, membrane stiffness, and

    microvillus distribution. The RMS increases with increasing

    shear rate, and membrane stiffness. On the contrary, it decreases

    with increasing number of microvilli. At Nmv = 155, the RMS

    does not change appreciably w.r.t. shear rate and membrane

    stiffness, though it shows an increasing trend w.r.t. these pa-

    rameters. At Nmv = 21, several-folds increase in the RMS is

    observed at higher membrane stiffness, and higher shear rate.

    Hence, the cell rolls more stably with the denser population of

    microvilli, and at lower shear rate and membrane stiffness.

    Interestingly, our simulations predict that during the rolling

    motion, a leukocyte can undergo a significant sideway move-

    ment. Fig. 9 shows the instantaneous lateral velocity of the cell

    for a representative case. The lateral velocity, similar to the

    axial velocity, also shows fluctuations which increase with in-

    creasing shear rate and membrane stiffness. The RMS of the

    lateral velocity fluctuations is shown in Fig. 10B for Nmv = 21

    which shows similar trend as that of the axial velocity RMS.

    The lateral velocity RMS increases with increasing shear rate,

    and membrane stiffness, and decreasing number of microvillus.

    For Nmv = 21, the RMS of the lateral velocity is comparable

    to, though lower than, that of the axial velocity. At higher

    Nmv = 155, the RMS of the lateral velocity is found to be sig-

    nificantly low (not shown).

    Total adhesion force between the cell and the substrate is

    shown in Fig. 11A as a function of shear rate, membrane stiff-

    ness, and microvilli distribution. The adhesive force obtained

    in our simulations ranges from about 100 to 750 pN for shear

    rates 100.500s1. In comparison, the adhesive force estimated

    by House and Lipowsky [40], based on in vivo measurements,

    ranges from 110 to 7610 pN for shear rates 200.2500 s1. The

    adhesive force increases with increasing membrane stiffness,

    partly due to increased hydrodynamic drag on less compliantcells. The adhesive force also increases with increasing shear

    rate and increasing microvilli population.

    It is also of interest to examine how the adhesive force is dis-

    tributed within the cell-surface contact area. This is shown in

    Fig. 11B for a representative case at 500 s1, Eh = 0.3dyn/cm,

    and Nmv = 155. In our simulations, bonds are assumed to form

    at the microvilli tips only. Hence the distribution of the adhe-

    sive force is not continuous, rather discrete. In Fig. 11B, we

    show the 3D distribution of the microvilli over the cell surface

    in the contact area. Together Figs. 11B and d show how the

    adhesive force is distributed among all 20 microvilli that are

    present in the contact area. Nearly 90% of the total adhesion

    force is concentrated in three microvilli (marked by arrows in

    Fig. 11D) which form tethers in the rear end of the cell. The

    force on each of these tethered microvilli ranges from 200 to

    400 pN, while that on each of the remaining 17 microvilli is in

    the range 020 pN. The maximum force on a tethered microvil-

    lus is shown in Fig. 11C as a function of shear rate, membrane

    stiffness, and microvilli distribution. This force is obtained

    before a tether breaks away from the substrate. It ranges from

    about 80 to 420 pN.

    The average number of microvilli that form tethers, and the

    total number of microvilli available within the contact area are

    shown in Fig. 12. For Nmv = 155, the number of microvilli

    within the contact area increases with increasing shear rate

  • 7/31/2019 3D Computational Modeling and Simulation o

    9/16

    746 V. Pappu, P. Bagchi / Computers in Biology and Medicine 38 (2008) 738753

    Time (s)

    Displacement(micron)

    1 2 3

    10

    15

    20

    25

    C

    B

    A

    Time (s)

    Displacement(micr

    on)

    0.1 0.2 0.3 0.4 0.5

    20

    30

    40

    50

    60

    E

    D

    C

    Velocity(micron/

    s)

    0 0.5 1 1.5 2 2.5

    0

    100

    200

    0 0.5 1 1.5 2 2.5

    0

    200

    400

    600

    800

    0 0.2 0.4 0.6

    0

    200

    400

    600

    800

    0 0.2 0.4 0.6

    0

    200

    400

    600

    800

    Time (s)

    0 0.2 0.4 0.6

    0

    500

    1000

    Fig. 6. Axial displacement (left figures) and axial velocity (right figures) of a rolling leukocyte. A, B, and C show the effect of shear rate which varies as

    100, 300, and 500 s1, respectively, at a constant Eh = 0.3dyn/cm. C, D, and E show the effect of membrane stiffness which varies as Eh = 0.3, 0.9, and

    2.6dyn/cm, respectively, at a constant shear rate of 500s1.

    and membrane compliance due to increased deformation of the

    cell. For the most compliant cell considered here at 500 s1

    shear rate, nearly 35 microvilli are available in the contact area.

    However, only five of them form tethers. The number of teth-

    ers increases slightly with increasing shear rate and membrane

    compliance. For Nmv = 21, the average number of tethers is

    even lower, and in the range of 12.

    We next examine the average number of selectin bonds

    formed between a cell and the substrate in Fig. 13. Note that we

    assume 50 PSGL-1 ligands on each microvillus tip (Table 1).

    In contrast, we observe that only about half of them, at most,

    forms bonds. The force on these bonds ranges from 0 to

    100 pN (discussed later). However, even less than half of these

    bonds carry a force > 10 pN. Thus we consider a bond to be

  • 7/31/2019 3D Computational Modeling and Simulation o

    10/16

    V. Pappu, P. Bagchi / Computers in Biology and Medicine 38 (2008) 738753 747

    shear rate (1/s)

    pausetim

    e(s)

    0 100 200 300 400 500

    0

    0.2

    0.4

    0.6

    0.8

    Smith et al

    shear rate (1/s)

    stepsize(m

    icron)

    100 200 300 400 5000

    1

    2

    3

    4

    5

    Fig. 7. (A) Average pause time, and (B) step distance as functions of shear rate, microvilli distribution, and membrane stiffness. Eh = 2.6, Eh = 0.9, and

    Eh = 0.3dyn/cm. Solid lines represent Nmv = 21 and dash line Nmv = 155. represent in vitro data of Smith et al. [6]. Pause time does not depend on

    Nmv, hence only data for Nmv = 155 are shown.

    X

    X** *

    shear rate (1/s)

    100 200 300 400 500 6000

    5

    10

    15

    shear rate (1/s)

    a

    veragevelocity(micron/s)

    100 200 300 400 5000

    20

    40

    60

    80

    100

    120

    Kim &Sarelius

    Fig. 8. Average rolling velocity for (A) Nmv = 21 and (B) Nmv = 155. Present results: Eh = 0.3, Eh = 0.9, Eh = 2.6dyn/cm. Experimental

    results: - - - - Kim and Sarelius [4], Yago et al. [9], X Ramachandran et al. [39].

    stretched when the force on it is > 5 pN. The average number

    of total bonds ranges from 200 to 500, and agrees well with

    that estimated by Jadhav et al. [29]. Remarkably, however, the

    number of stretched bonds lies in the range 2080.Next we examine the history of the force on individual

    microvilli. A representative case at 500 s1, Eh = 0.3dyn/cm,

    and Nmv =21 is considered in Fig. 14A. The microvilli are iden-

    tified by numbers 1, 2 etc. The location of these microvilli

    on the cell surface was shown earlier in Fig. 4. Strikingly,

    our simulations show that the force on a microvillus devel-

    ops in steps. As an illustration, consider microvillus 4 which

    comes in contact with the substrate and forms adhesion bonds

    at around t= 0.03 s. However, the bonds on microvillus 4 are

    not stretched until t= 0.13 s as the cell is tethered by microvilli

    2 and 3. Hence the force acting on microvillus 4 remains

    less than 10 pN. At t = 0.13 s, microvillus 2 breaks. As the

    cell begins to roll, bonds in microvilli 3 and 4 are stretched,

    and the forces on them sharply increase. Subsequently, the cell

    is tethered by microvillus 3, and the force on it increases to

    about 350 pN. The force on microvillus 4 also increases to

    80 pN due to bond stretching. At aroundt =

    0.175 s, microvil-lus 3 breaks. The cell rolls again, and the bonds in microvillus

    4 are stretched further, and the force jumps to 280 pN.

    We next examine the force history in individual

    P-selectin/PSGL-1 bonds in Fig. 14B. Formation and breakage

    of individual bonds occur throughout the lifetime of a tether.

    When one bond breaks, forces on the remaining bonds in the

    cluster increase in a step-like manner. It also implies that mul-

    tiple bond breakage, rather than a single event, is necessary

    for a microvillus tether to break. We also note that for a few

    bonds, the peak force on individual bond is as high as 50 pN,

    while for others the force remains below 10 pN. It again sug-

    gests that, though multiple bonds are present, not all of them

    are effective in tethering the cell.

  • 7/31/2019 3D Computational Modeling and Simulation o

    11/16

    748 V. Pappu, P. Bagchi / Computers in Biology and Medicine 38 (2008) 738753

    The average bond rupture force, and the bond lifetime are

    shown in Figs. 15A and B. The rupture force varies in the

    range of 2080 pN, and it increases with increasing shear

    rate and membrane stiffness. The bond lifetime, which corre-

    lates well with the cell pause times shown earlier, decreases

    with increasing shear and membrane stiffness. Dependence

    of bond lifetime on membrane stiffness is stronger at lowshear.

    5. Discussion

    3D computational modeling and simulation are presented

    to study adhesive rolling of deformable leukocytes in a shear

    flow. We study the effect of cell deformation, shear rate, and

    microvilli distribution on the rolling characteristics. The simu-

    lations show the transient deformation of a tethered leukocyte

    in to a tear-drop shape as observed in previous experiments

    (Fig. 4). Computed cell deformation agrees well with earlier in

    vivo and in vitro measurements (Fig. 5). The average rolling ve-

    locity increases with increasing shear and membrane stiffness,

    and also agrees well with previous experimental measurements

    (Fig. 8).

    The sequence of cell shape presented in Figs. 3 and 4 sug-

    gested that the cellsubstrate contact area of a rolling leukocyte

    varies with timeit increases when the cell is adherent, and

    Time (s)

    0.1 0.2 0.3 0.4 0.5

    -200

    0

    200

    400

    E

    Fig. 9. Sideway velocity (m/s) of a rolling leukocyte. The case shown is

    the same as in Fig. 6E.

    shear rate (1/s)

    RMSvelocity(micron

    /s)

    100 200 300 400 5000

    50

    100

    150

    200

    250

    shear rate (1/s)

    100 200 300 400 5000

    50

    100

    150

    200

    250

    Fig. 10. RMS velocity fluctuation of (A) axial and (B) sideway motion. Nmv = 21 and - - - - - Nmv = 155. Eh = 0.3, Eh = 0.9, and Eh = 2.6dyn/cm.

    For (B) Nmv = 155 does not give any significant sideway motion, and hence data not shown.

    decreases when the cell is rolling, in agreement with in vitro

    observation of Dong et al. [8]. Previous in vivo and in vitro

    studies suggested that rolling leukocytes become flattened

    against the substrate upon initial tethering. Our results in Figs. 3

    and 4 suggest that upon initial tethering, the cell rotates about

    the tether followed by flattening of the cell surface which

    occurs very rapidly, and often within a fraction of a second.One distinct characteristics of leukocyte rolling observed in

    vivo and in vitro is that the cells do not roll continually, but

    rather in a stop-and-go manner. Previous computational models

    that considered deformation of adherent leukocytes did not re-

    port such intermittent motion [2428]. The stop-and-go motion

    is due to formation and breaking of receptor/ligand bonds, and

    can only be predicted by stochastic simulation. Coupling cell

    deformation with stochastic bond kinetics is relatively new in

    the context of modeling of leukocyte rolling. Our IBM simula-

    tion has been able to predict the stop-and-go motion of a leuko-

    cyte. Our simulations predict that the fluctuations in rolling ve-

    locity are reduced, and hence the rolling motion stabilizes, with

    increasing cell deformability, in agreement with earlier exper-

    imental observations [7,9].

    Our simulations predicted that a leukocyte can undergo a

    significant sideway motion during rolling. The sideway motion

    is purely due to the stochastic nature of bond formation, and

    discrete nature of microvilli presentation. A microvillus can

    form an initial tether that may be oriented at a non-parallel angle

    with the shear flow, causing the cell to move sideway once an

    earlier tether breaks. However, once the cell is tethered by the

    new microvillus, it quickly orients itself parallel to the flow.

    Not only the stochastic formation and breakage of selectin

    bonds, also the presentation of the microvilli affects the fluc-

    tuating motion of a leukocyte. By considering two differentmicrovilli distributions, we show that the average rolling veloc-

    ity and fluctuations are higher for the sparse distribution. The

    higher rolling velocity is caused by increasing step size which

    seems to correlate with the inter-microvilli distance. Presence

    of microvilli in our model allows formation of discrete bond

    clusters rather than uniform distribution of bonds over the

  • 7/31/2019 3D Computational Modeling and Simulation o

    12/16

    V. Pappu, P. Bagchi / Computers in Biology and Medicine 38 (2008) 738753 749

    shear rate (1/s)

    force(pN)

    100 200 300 400 5000

    200

    400

    600

    800

    shear rate (1/s)

    force(pN)

    100 200 300 400 5000

    100

    200

    300

    400

    500

    Flow

    Y X

    Fig. 11. (A) Average total adhesion force, (B) distribution of adhesion force among all bound microvilli in the cell/substrate contact area, (C) average peakforce on individual tethered microvilli, and (D) distribution of bound microvilli over the cell surface. The arrows indicate three microvilli forming tethers.

    Nmv = 21 and - - - - - Nmv = 155. Eh = 2.6 and Eh = 0.3dyn/cm.

    entire cell/substrate contact zone. In vivo measurement by

    Zhao et al. [42] obtained step size 2 m, while in vitro mea-

    surements by Alon et al. [46] for L-selectin mediated rolling

    estimated 3.4m. These data, together with our prediction

    of 0.5.4m step size for P-selectin mediated rolling sug-

    gested that the step-distance is independent of the nature of

    receptorligand complex, and is dependent on the microvilli

    distribution. The pause time between successive rolling steps

    is observed to depend not only on shear rate, but also on cell

    deformability. On the contrary, it did not depend on microvilli

    distribution.

    We found that leukocyte adhesion is via multiple tethers,

    varying from 2 to 5 in numbers. These numbers are, however,

    significantly less than the number of microvilli, often up to 35,

    which are available within the cell/substrate contact area, and

    are observed to form bonds. Increasing the shear rate by a factor

    of five only doubled the number of tethered microvilli from an

    average value of 2.5 to 5. Though multiple tethers are observed

    in most of our simulations, we also observe that a single tether

    is often sufficient to support rolling. We examined how the

    adhesion force is distributed among all microvilli that are bound

    to the substrate. Though nearly all microvilli in the contact

    area contain bonds, significant adhesive force is concentrated

    in 13 tethered microvilli located in the rear-most part of the

    cell.

    The peak force on a tethered microvillus predicted by

    our simulations is significantly above the force required for

    tether extrusion (61 pN) measured by Shao et al. [43]. Our

    results, therefore, predict that the microvilli in the rear end

    of the cell would develop tether extrusion, while those in the

    remaining part of the contact area would at most undergo

    elastic tension. Although we have not considered deformation

    of a microvillus in our model, such a consideration would

    reduce the tether force [23,43,44]. In that case, tether ex-

    trusion may occur only at higher shear rate and membrane

    stiffness.

    We observe that multiple, rather than single, selectin bonds

    form per microvillus. When a tether microvillus breaks and

    pulls from the substrate, several bonds, varying from 9 to 14,

    break simultaneously. We also observe that during a pause in the

    rolling motion, individual bonds within the tethered microvillus

    can break. But breakage of individual bond does not lead to the

  • 7/31/2019 3D Computational Modeling and Simulation o

    13/16

    750 V. Pappu, P. Bagchi / Computers in Biology and Medicine 38 (2008) 738753

    shear rate (1/s)

    100

    300

    500

    0

    10

    20

    30

    40

    shear rate (1/s)10

    0300

    500

    0

    10

    20

    30

    40

    shear rate (1/s)

    100

    300

    500

    0

    10

    20

    30

    40

    shear rate (1/s)

    100

    300

    500

    0

    10

    20

    30

    40155 microvilli

    Eh=0.3 dyn/cm

    155 microvilli

    Eh=2.6 dyn/cm

    21 microvilli

    Eh=2.6 dyn/cm

    21 microvilli

    Eh=0.3 dyn/cm

    Fig. 12. Average number of total bound microvilli (open bars) and tethered microvilli (solid bars) as functions of shear rate, cell compliance, and microvilli

    distribution.

    breakage of the tether, as other bonds quickly stretch to share

    the load. Simultaneous breakage of multiple bonds can only

    cause a tether to retract.

    The facts that multiple bonds form per microvillus, and

    the pause time of rolling motion was found to depend on cell

    compliance, are noteworthy in the context of experimental es-

    timation of bond dissociation rates. Bond dissociation rates are

    often measured based on the pause times obtained from leuko-

    cyte rolling in flow chamber. However, cell deformation cannot

    be easily controlled in such experiments. Assuming single bond

    dissociation, the Bell model [45] is fit to the measurements to

    obtain the bond dissociation rate. Direct measurements using

    atomic force microscopy or laser trap yielded different values

    than that obtained from flow chamber experiments [3,11,47].

    Our results suggest that two possible reasons for discrepancy

    are cell deformation and formation of multiple bonds per

    tether.

    Our simulations predicted the average number of total

    bonds in the range 100500, and it increases with increas-

    ing shear rate and membrane compliance due to increased

    cell-surface contact area resulting in more bound microvilli.

    However, we observe that only a few bonds residing in the

    tethered microvilli are significantly stretched. The average

    number of stretched bonds is in the range of 2080 which

    increases with increasing shear rate, due to increasing num-

    ber of tethers, but does not depend on membrane compliance

    due to the fact that the number of tethers did not increase ap-

    preciably with increasing cell deformability. Per microvillus,

    only 914 bonds are predicted to be significantly stretched,

    which agrees well with the measurements of Chen and

    Springer [10].

    The maximum and average rupture force of a P-selectin/

    PSGL-1 bond was estimated to be 100 and 60 pN, respectively,

    which is in the same range as measured recently by Marshal

    et al. [48] using atomic force microscopy. It is also in the

    same range as that obtained by Schmidtke and Diamond ([41],

    86172 pN at 100250 1/s). We observe that bond force in-

    creases with increasing shear and membrane stiffness. Depen-

    dence of bond force on cell deformability is also in agreement

    with previous in vitro studies using untreated neutrophil, and

    ligand-coated microbeads [7]. The bond force was measured to

    be 500 pN for microbeads and 124 pN for neutrophil at a shear

  • 7/31/2019 3D Computational Modeling and Simulation o

    14/16

    V. Pappu, P. Bagchi / Computers in Biology and Medicine 38 (2008) 738753 751

    shear rate (1/s)

    Bonds

    100 200 300 400 5000

    100

    200

    300

    400

    500

    Fig. 13. Average total bonds (solid lines) and stretched bonds (dash lines).

    Eh = 2.6 and Eh = 0.3dyn/cm.

    Time (s)

    force

    (pN)

    0.05 0.1 0.15 0.2 0.250

    100

    200

    300

    400

    500

    1

    2

    3

    3

    5

    4

    Time (s)

    force

    (pN)

    0.7 0.8 0.9 1 1.1 1.2 1.3 1.40

    10

    2030

    40

    50

    60

    70

    Fig. 14. (A) Force history of a few microvilli and (B) force history of a few bonds within a particular microvillus. In (A), microvilli are indicated by number

    1, 2, etc. which are same as shown earlier in Fig. 4. In (B) arrows indicate step increase in bond force in response to breaking of another bond.

    shear rate (1/s)

    Bondpeakforce(pN)

    100 200 300 400 5000

    20

    40

    60

    80

    100

    shear rate (1/s)

    Bondlifetime(s)

    100 200 400300 500

    0

    0.2

    0.4

    0.6

    0.8

    1

    Fig. 15. (A) Peak bond force and (B) bond life time. Eh = 2.6 and Eh = 0.3dyn/cm.

    rate 100 s1 (see also Smith et al. [6]). Significantly reduced

    bond force for neutrophil compared to that for microspheres

    was attributed to microvilli elongation. In contrast, microvilli

    extension was not considered in our model. Our result, there-

    fore, suggests that cell deformation helps reducing the bond

    force and prolong bond lifetime. Hydrodynamic shear and cell

    deformation have competing effects on bond dissociation. Forceon individual bond increases, and hence bond lifetime de-

    creases, with increasing shear. Despite rapid bond dissociation

    at high shear, leukocyte rolling is stabilized by cell deformation

    via two pathways. First, increased cell/substrate contact area

    causes more microvilli to be accessible for bond formation. In-

    creased number of tethers with increasing shear alleviates the

    force on individual bond. The second mechanism is the reduc-

    tion of the hydrodynamic drag on a deformed cell which also

    reduces bond force and hence the dissociation rate.

    In conclusion, we presented a 3D model to simulate a rolling

    leukocyte in shear flow. Our model predicted the stop-and-go

    motion of leukocytes. The major results are: (i) compliant cells

    roll more stably, and have longer pauses due to reduced bond

    force and increased bond lifetime, (ii) microvilli presentation

    affects rolling characteristics by altering the step size, (iii)

    adhesion force is concentrated in only 13 tethered microvilli

    in the rear-most part of a cell, (iv) number of effective bonds

    is much less than total adhesion bonds, (v) adhesion is via

  • 7/31/2019 3D Computational Modeling and Simulation o

    15/16

    752 V. Pappu, P. Bagchi / Computers in Biology and Medicine 38 (2008) 738753

    multiple tethers, each of which forms multiple selectin bonds,

    but often one tether is sufficient to support rolling, (vi) force

    loading on individual microvillus and selectin bond is not con-

    tinuous, rather occurs is steps, (vii) peak force on a tethered

    microvillus is much higher than that measured to cause tether

    extrusion. Thus, both cell deformation and microvillus defor-

    mation occur simultaneously during cell rolling, and need tobe considered in future computationalmodels.

    Conflict of interest statement

    None declared.

    Acknowledgments

    This research is partially supported by a Busch Biomedical

    Grant from Rutgers University, and NSF Grant BES-0603035.

    Computational support from National Center for Supercomput-

    ing Applications at Urbana, IL is acknowledged. Authors thank

    R.M. Kalluri and Sai Doddi for their help.

    References

    [1] T.A. Springer, Traffic signals on endothelium for lymphocyte

    recirculation and leukocyte emigration, Annu. Rev. Physiol. 57 (1995)

    827872.

    [2] M.B. Lawrence, T.A. Springer, Leukocytes roll on a selectin at

    physiologic flow rates: distinction from and prerequisite for adhesion

    through integrins, Cell 65 (1991) 859873.

    [3] L.J. Rinko, M.B. Lawrence, W.H. Guilford, The molecular mechanics

    of P- and L-selectin lectin domains binding to PSGL-1, Biophys. J. 86

    (2004) 544554.

    [4] M.B. Kim, I.H. Sarelius, Role of shear forces and adhesion molecule

    distribution on P-selectin-mediated leukocyte rolling in postcapillary

    venules, Am. J. Physiol. Heart Circ. Physiol. 287 (2004) H2705H2711.[5] R. Alon, D.A. Hammer, T.A. Springer, Lifetime of the P-selectin-

    carbohydrate bond and its response to tensile force in hydrodynamic

    flow, Nature 374 (1995) 539542 (Erratum, Nature 376 (1995) 86).[6] M.J. Smith, E.L. Berg, M.B. Lawrence, A direct comparison of selectin-

    mediated transient, adhesive events using high temporal resolution,

    Biophys. J. 77 (1999) 33713383.

    [7] E.Y. Park, M.J. Smith, E.S. Stropp, K.R. Snapp, J.A. DiVietro, W.F.

    Walker, D.W. Schmidtke, S.L. Diamond, M.B. Lawrence, Comparison

    of PSGL-1 microbead and neutrophil rolling: microvillus elongation

    stabilizes P-selectin bond clusters, Biophys. J. 82 (2002) 18351847.[8] C. Dong, J. Cao, E.J. Struble, H.H. Lipowsky, Mechanics of leukocyte

    deformation and adhesion to endothelium in shear flow, Ann. Biomed.

    Eng. 27 (1999) 298312.

    [9] T. Yago, A. Leppanen, H. Qiu, W.D. Marcus, M.U. Nollert, C. Zhu, R.D.

    Cummings, R.P. McEver, Distinct molecular and cellular contributions to

    stabilizing selectin-mediated rolling under flow, J. Cell Biol. 158 (2002)

    787799.

    [10] S. Chen, T.A. Springer, An automatic braking system that stabilizes

    leukocyte rolling by an increase in selectin bond number with shear, J.

    Cell Biol. 144 (1999) 185200.[11] R. Alon, S. Chen, R. Fuhlbrigge, K.D. Puri, T.A. Springer, The kinetics

    and shear threshold of transient and rolling interactions of L-selectin

    with its ligand on leukocytes, Proc. Natl. Acad. Sci. USA 95 (1998)

    1163111636.

    [12] M.B. Lawrence, G.S. Kansas, E.J. Kunkel, K. Ley, Threshold levels of

    fluid shear promote leukocyte adhesion through selectins (CD62L,P,E),

    J. Cell Biol. 136 (1997) 717727.[13] E.B. Finger, R.E. Bruehl, D.F. Bainton, T.A. Springer, A differential role

    for cell shape in neutrophil tethering and rolling on endothelial selectins

    under flow, J. Immunol. 157 (1996) 50855096.

    [14] R.E. Bruehl, T.A. Springer, D.F. Bainton, Quantization of L-selectin

    distribution on human leukocyte microvilli by immunogold labeling and

    electron microscopy, J. Histochem. Cytochem. 44 (1996) 835844.

    [15] U.H. von Andrian, S.R. Hasslen, R.D. Nelson, S.L. Erlandsen, E.C.

    Butcher, A central role for microvillus receptor presentation in leukocyte

    adhesion under flow, Cell 82 (1995) 989999.

    [16] J.C. Firrell, H.H. Lipowsky, Leukocyte margination and deformation in

    mesenteric venules of rat, Am. J. Physiol. 256 (1989) H1667H1674.[17] E.R. Damiano, J. Westheider, A. Tozeren, K. Ley, Variation in the

    velocity, deformation, and adhesion energy density of leukocytes rolling

    within venules, Circ. Res. 79 (1996) 11221130.

    [18] E.B. Lomakina, C.M. Spillmann, M.R. King, R.E. Waugh, Rheological

    analysis and measurement of neutrophil indentation, Biophys. J. 87

    (2004) 42464258.

    [19] D.A. Hammer, S.M. Apte, Simulation of cell rolling and adhesion on

    surfaces in shear flow: general results and analysis of selectin-mediated

    neutrophil adhesion, Biophys. J. 63 (1992) 3557.

    [20] K.C. Chang, D.A. Hammer, Adhesive dynamics simulations of sialyl-

    Lewisx/E-selectin-mediated rolling in a cell-free system, Biophys. J. 79

    (2000) 18911902.

    [21] K.C. Chang, D.F.J. Tees, D.A. Hammer, The state diagram for cell

    adhesion under flow: leukocyte rolling and firm adhesion, Proc. Natl.

    Acad. Sci. 97 (2000) 1126211267.[22] M.R. King, D.A. Hammer, Multiparticle adhesive dynamics. Interactions

    between stably rolling cells, Biophys. J. 81 (2001) 799813.

    [23] K.E. Caputo, D.A. Hammer, Effect of microvillus deformability on

    leukocyte adhesion explored using adhesive dynamics simulations,

    Biophys. J. 89 (2005) 187200.

    [24] C. Dong, X. Lei, Biomechanics of cell rolling: shear flow, cell-surface

    adhesion, and cell deformability, J. Biomech. 33 (2000) 3543.

    [25] X. Lei, M.B. Lawrence, C. Dong, Influence of cell deformation on

    leukocyte rolling adhesion in shear flow, J. Biomech. Eng. 121 (1999)

    636664.

    [26] N.A. NDri, W. Shyy, R. Tran-Son-Tay, Computational modeling of cell

    adhesion and movement using a continuum-kinetics approach, Biophys.

    J. 85 (2003) 22732286.

    [27] D.B. Khismatullin, G.A. Truskey, Three-dimensional numerical

    simulation of receptor-mediated leukocyte adhesion to surfaces: effectsof cell deformability and viscoelasticity, Phys. Fluids 17 (2005) 031505.

    [28] D.B. Khismatullin, G.A. Truskey, A 3D numerical study of the effect of

    channel height on leukocyte deformation and adhesion in parallel-plate

    flow chambers, Microvasc. Res. 68 (2004) 188202.

    [29] S. Jadhav, C.D. Eggleton, K. Konstantopoulos, A 3-D computational

    model predicts that cell deformation affects selectin-mediated leukocyte

    rolling, Biophys. J. 88 (2005) 96104.

    [30] C.S. Peskin, D.M. McQueen, A 3-dimensional computational method

    for blood-flow in the heart. 1. Immersed elastic fibers in a viscous

    incompressible fluid, J. Comput. Phys. 81 (1989) 372405.

    [31] G. Tryggvason, B. Bunner, A. Esmaeeli, N. Al-Rawahi, W. Tauber, J.

    Han, S. Nas, Y. Jan, A front tracking method for the computations of

    multiphase flow, J. Comput. Phys. 169 (2001) 708759.

    [32] R. Mittal, G. Iaccarino, Immersed boundary methods, Annu. Rev. Fluid

    Mech. 37 (2005) 239261.[33] C. Dong, R. Skalak, Leukocyte deformability: finite element modeling

    of large viscoelastic deformation, J. Theor. Biol. 158 (1992) 173193.

    [34] C. Dong, R. Skalak, K.L. Sung, G.W. Schmid-Schonbein, S. Chien,

    Passive deformation analysis of human leukocytes, J. Biomech. Eng.

    110 (1988) 2736.

    [35] J.M. Charrier, S. Shrivastava, R. Wu, Free and constrained inflation

    of elastic membranes in relation to thermoforming-non-axisymmetric

    problems, J. Strain Anal. 24 (1989) 5574.

    [36] M. Dembo, On Peeling an Adherent Cell from a Surface, Lectures

    on Mathematics in the Life Sciences, Some Mathematical Problems in

    Biology, vol. 26, American Mathematical Society, Providence, RI, 1994

    pp. 5177.

    [37] P. Bagchi, S. Balachandar, Steady planar straining flow past a rigid

    sphere at moderate Reynolds numbers, J. Fluid Mech. 466 (2002)

    365407.

  • 7/31/2019 3D Computational Modeling and Simulation o

    16/16

    V. Pappu, P. Bagchi / Computers in Biology and Medicine 38 (2008) 738753 753

    [38] S. Ramanujan, C. Pozrikidis, Deformation of liquid capsules enclosed

    by elastic membranes in simple shear flow: large deformations and the

    effect of fluid viscosities, J. Fluid Mech. 361 (1998) 117143.

    [39] V. Ramachandran, M. Williams, T. Yago, D.W. Schmidtke, R.P. McEver,

    Dynamic alterations of membrane tethers stabilize leukocyte rolling on

    P-selectin, Proc. Natl. Acad. Sci. USA 101 (2004) 1351913524.

    [40] S.D. House, H.H. Lipowsky, Leukocyte-endothelium adhesion:

    microhemodynamics in mesentery of the cat, Microvasc. Res. 34 (1987)361379.

    [41] D.W. Schmidtke, S.L. Diamond, Direct observation of membrane tethers

    formed during neutrophil attachment to platelets or P-selectin under

    physiological flow, J. Cell Biol. 149 (2000) 719729.

    [42] Y. Zhao, S. Chien, R. Skalak, H.H. Lipowsky, Leukocyte rolling in rat

    mesentery venules: distribution of adhesion bonds and the effects of

    cytoactive agents, Ann. Biomed. Eng. 29 (2001) 360372.

    [43] J.Y. Shao, H.P. Ting-Beall, R.M. Hochmuth, Static and dynamic

    lengths of neutrophil microvilli, Proc. Natl. Acad. Sci. USA 95 (1998)

    67976802.

    [44] Y. Yu, J.Y. Shao, Simultaneous tether extraction contributes to neutrophil

    rolling stabilization: a model study, Biophys. J. 92 (2007) 418429.

    [45] G.I. Bell, Models for the specific adhesion of cells to cells, Science 200

    (1978) 618627.

    [46] R. Alon, S. Chen, K. Puri, E.B. Finger, T.A. Springer, The kinetics of

    L-selectin tethers and the mechanics of selectin-mediated rolling, J. Cell

    Biol. 138 (1997) 11691180.

    [47] U.S. Scwarz, R. Alon, L-selectin-mediated leukocyte tethering in shear

    flow is controlled by multiple contacts and cytoskeletal anchorage

    facilitating fast rebinding events, Proc. Natl. Acad. Sci. 101 (2004)69406945.

    [48] B.T. Marshall, K.K. Sarangapani, J. Wu, M.B. Lawrence, R.P. McEver,

    C. Zhu, Measuring molecular elasticity by atomic force microscope

    cantilever fluctuations, Biophys. J. 90 (2006) 681692.

    [49] E.F. Krasik, D.A. Hammer, A semianalytic model of leukocyte rolling,

    Biophys. J. 87 (2004) 29192930.

    Prosenjit Bagchi received his PhD from the University of Illinois at Urbana-Champaign in 2002. Currently he is an assistant professor in Mechanical andAerospace Engineering at Rutgers University.

    Vijay Pappu is a graduate student in Mechanical and Aerospace Engineeringat Rutgers University.