flow modelling

Upload: deepika-singh

Post on 09-Apr-2018

227 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/7/2019 Flow Modelling

    1/16

    Journal of Biotechnology 119 (2005) 181196

    Flow modelling within a scaffold under the influence ofuni-axial and bi-axial bioreactor rotation

    H. Singh a,1, S.H. Teoh b,2, H.T. Low b,3, D.W. Hutmacher b,

    a Tissue Engineering Laboratory E3-05-04, National University of Singapore, 10 Kent Ridge Crescent, Singapore 119260, Singaporeb Division of Bioengineering, National University of Singapore, 9 Engineering Drive 1, EA 03-12, Singapore 117576, Singapore

    Received 20 September 2004; received in revised form 17 March 2005; accepted 29 March 2005

    Abstract

    The problem of donor scarcity has led to the recent development of tissue engineering technologies, which aim to create

    implantable tissue equivalents for clinical transplantation. These replacement tissues are being realised through the use of

    biodegradable polymer scaffolds; temporary/permanent substrates, which facilitate cell attachment, proliferation, retention and

    differentiated tissue function. To optimise gas transfer and nutrient delivery, as well as to mimic the fluid dynamic environment

    present within the body, a dynamic system might be chosen. Experiments have shown that dynamic systems enhance tissue

    growth, with the aid of scaffolds, as compared to static culture systems. Very often, tissue growth within scaffolds is only seen

    to occur at the periphery. The present study utilises the Computational Fluid Dynamics package FLUENT, to provide a betterunderstanding of the flow phenomena in scaffolds, within our novel bioreactor system. The uni-axial and bi-axial rotational

    schemes are studied and compared, based on a vessel rotating speed of 35 rpm. The wall shear stresses within and without the

    constructs are also studied. Findings show that bi-axial rotation of the vessel results in manifold increases of fluid velocity within

    the constructs, relative to uni-axial rotation about the X- and Z-axes, respectively.

    2005 Elsevier B.V. All rights reserved.

    Keywords: Computational fluid dynamics; Tissue engineering; Bioreactor

    Corresponding author. Tel.: +65 6874 1036; fax: +65 6872 3069.

    E-mail addresses: [email protected] (H. Singh),

    [email protected] (S.H. Teoh), [email protected]

    (H.T. Low), [email protected] (D.W. Hutmacher).1 Tel.: +65 6874 8870; fax: +65 6777 3537.2 Tel.: +65 6874 4605; fax: +65 6872 3069.3 Tel.: +65 6874 2225; fax: +65 6872 3069.

    1. Introduction

    In a number of long-term tissue engineering stud-

    ies under static conditions (e.g. Petri dish), it has

    proven extremely difficult to promote the high-density

    three-dimensional in vitro growth of cells that have

    been removed from the body and deprived of their

    normal in vivo vascular sources of nutrients and gas

    exchange. To optimise gas transfer and nutrient deliv-

    0168-1656/$ see front matter 2005 Elsevier B.V. All rights reserved.

    doi:10.1016/j.jbiotec.2005.03.021

  • 8/7/2019 Flow Modelling

    2/16

    182 H. Singh et al. / Journal of Biotechnology 119 (2005) 181196

    ery, as well as to mimic the fluid dynamic environment

    present within the body, a dynamic system has to be

    chosen. The bioreactor is one example that attempts

    to fulfil these requirements, particularly in the studyof tissue engineering, by simulating the physiological

    and fluidic conditions that occur in vivo (Freed and

    Vunjak-Novakovic, 2002; Papoutsakis, 1991; Cherry

    and Papoutsakis, 1986). Cells have often been seen to

    grow well along the periphery of 3D scaffold, while

    proliferation is often significantly affected at the centre

    of the scaffold, where necrotic neo-tissue can some-

    times be seen. This is partly dueto poor fluidictransport

    of media and scaffold design, among other reasons.

    Therefore, simulations that provide such flow visual-

    izations could greatly assist in the design of scaffolds

    as well as bioreactor.

    A number of experimental studies were previously

    reported by using different bioreactor systems (Freed

    et al., 1994; Martin et al., 2004; Vunjak-Novakovic et

    al., 1996). However, limited information has been pub-

    lished on studies that attempt to simulate the dynamic

    fluid environment prior to commencing experimen-

    tal studies. Advantages of computational simulations

    include the ability to modify and study the effects of

    bioreactor design with respect to the flow analysis,

    without having to develop and construct actual phys-

    ical models, or to run a large number of experiments.This is coupled with the significant savings in time and

    costs. Furthermore, visualisation of flow as enabled by

    the simulation package is a key factor in determining

    the efficacy of the system, and allows for design opti-

    mization prior to bioreactor design and modifications.

    NASA has taken significant strides relating to their

    rotating bioreactor studies, under the influence of unit

    and micro-gravity. While it is known that in the pres-

    ence of body forces, density differences between the

    cells attached to micro-carriers and the fluid medium,

    cause relative motion resulting in mechanical shearand increased mass transport. However, in the micro-

    gravity environment, buoyancy effects are greatly

    reduced. The gravity of Earth is replaced by centripetal

    acceleration as the dominant body force. For a typ-

    ical rotation rate of 2 rpm, within a 0.05 m diameter

    vessel, the magnitude of the body force is reduced

    (Kleis and Pellis, 1995) to approximately 0.001 m/s2.

    Kleis et al. (1996) proceeded further by carrying

    out studies on mass transport, with regards to their

    micro-gravity bioreactor model. Consequently Boyd

    and Gonda (1990) mathematically modelledthe motion

    of particles within the NASA bioreactor model at both

    unit and micro-gravity. However, the effects of gravity

    can be felt by almost everything on Earth, and is notignored here for practical reasons. Rotating (Bursac

    et al., 2003; Martin et al., 2004) and perfusion (Martin

    et al., 2004; Bancroft et al., 2003; Darling and

    Athanasiou, 2003; Sodian et al., 2002) bioreactors,

    each with their own flow characteristics, are currently

    being utilized for studies involving the culturing of

    boneand cartilage.Thesebioreactor models induce dif-

    ferent normal and shear forces that act on cells, with

    varying consequences.

    One aspect that this paper will focus on is that of

    shear stresses acting on cells and tissues. A response

    of an endothelial monolayer of cells subject to steady,

    laminar shear stresses is that of an alteration in mor-

    phology. A typical change would occur from an ini-

    tial polygonal pattern, to one, which depicts an elon-

    gated profile, being aligned to the direction of fluid

    flow (Levesque and Nerem, 1985; Levesque et al.,

    1989; Stathopoulos and Hellums, 1985). This change

    is basically accompanied by the restructuring of the

    cytoskeleton, or more specifically, the alignment of

    microtubules, followed by the formation of actin stress

    fibres. Similar experiments were conducted on bovine

    endothelial cells. It was found that the stiffness ofendothelial cells exposed to shear stresses of 2 Pa,

    increased with the duration time of exposure. However,

    after 24 h of exposure, the stiffness of the endothe-

    lial cells was similar all around the cell, indicating

    the ability of the cells to adapt to the changes in the

    environment. This concept includes that of stress fibre

    orientation, as mentioned in the paper by Yamada et al.

    (2000).

    Olivier and Truskey (1993) showed that the flow

    regime itself, and not just the magnitude of shear

    stress, plays a critical part. Their experiments involvedturbulent flows as generated in a cone and plate vis-

    cometer, which actually resulted in the detachment of

    anchorage-dependent cells at shear stresses of only

    1.5 dyn/cm2.

    At relatively higher levels of shear stress, rang-

    ing from 26 to 54 dyn/cm2, human endothelial cells

    were found to detach from its surface and correspond-

    ingly exhibited reduced viability (Stathopoulos and

    Hellums, 1985). In contrast to this, no cell loss was

    reported for bovine aorta endothelial cells, which were

  • 8/7/2019 Flow Modelling

    3/16

    H. Singh et al. / Journal of Biotechnology 119 (2005) 181196 183

    subject to a stress level (Levesque and Nerem, 1985)

    of approximately 85 dyn/cm2. Pritchard et al. (1995)

    found that adhesion generally increases with decreas-

    ing wall shear stress. It is also now well documentedthat cell metabolism is affected by fluid stresses. Shear

    stresses, within limits, tend to stimulate the release of

    specific enzymes and growth factors, thereby enhanc-

    ing cellular attachment and proliferation.

    Stathopoulos and Hellums (1985) found that HEK

    cells release urokinase, due to variations in shear stress

    levels. Furthermore, human umbilical vein endothelial

    cells were found to release a five-fold amount of prosta-

    cyclin at 10 dyn/cm2, as compared to near-zero stress

    conditions. Similarly, Vunjak-Novakovic et al. (1996,

    1998) studied the affects of dynamic seeding within

    polymer scaffolds with respect to chondrocytes, and

    found that these cells can be conditioned to grow

    under the correct shearstress levels. Chondrocyteshave

    been known to proliferate better under the influence of

    shear stresses and are able to withstand higher loads as

    compared to chondrocytes cultured statically (Sucosky

    et al., 2003). Both Porter et al. (2005) and Raimondi

    et al. (2004) utilised CFD techniques to model the

    effects of perfusion, to better comprehend the influence

    of perfusion and hence shear stresses, on 3D cultures.

    Lappa (2005) on the other hand, developed a rigorous

    set of equations to model fluid flow and algorithms toestimate soft tissuegrowth within a bioreactor. Redaelli

    et al. (1997) on the other hand, attempted to model

    pulsatile flow within arteries. This intriguing article

    describes how a FORTRAN algorithm was interfaced

    to FIDAP, a commercially available CFD package, and

    then simulated.

    2. Software and methodology

    2.1. GAMBIT and FLUENT

    GAMBIT (Fluent Inc.), being a powerful graph-

    ical modeling tool, was utilised as a pre-processor

    for the Computational Fluid Dynamics package, FLU-

    ENT. This general-purpose CFD tool can solve many

    fluid flow problems such as steady, unsteady, lami-

    nar and turbulent flows, heat transfer, Newtonian and

    non-Newtonian flows, among others. FLUENT also

    provides excellent mesh flexibility as well, allowing

    the grid to be refined or coarsened based on the require-

    Fig. 2.1. Actual bioreactor prototype model.

    ments of the flow solution. Flow solutions are based on

    the conservation of mass, momentum and energy equa-

    tions. GAMBIT and FLUENT were therefore used in

    tandem, to model the flow of fluid within our prototype

    bioreactor model.

    2.2. The bioreactor model

    A novel bi-axial bioreactor system was developed

    by a team from the National University of Singaporeand the Singapore Polytechnic. The unique design of

    the entire system adds some flexibility that is some-

    times found to be lacking in commercially available

    bioreactors. This novel design involves the bioreac-

    tor being vertically upright, instead of the conven-

    tional horizontal-type vessel that is so often seen (see

    Fig. 2.1). However, the fact that allows this design to

    particularly stand out is that it allows for rotation either

    about the vertical axis (spinning), the horizontal axis

    (tumbling), or even rotation simultaneously about both

    axes(gyroscopic action). Ideally, the utilisationof these

    features would lead to enhanced cell culture media flow

    throughout the entire vessel (see Fig. 2.2and Table 2.1).

    2.3. The tissue engineering scaffold

    The scaffolds utilised for the simulations are rou-

    tinely produced by our group (Zein et al., 2002;

    Hutmacher et al., 2004). The poly()caprolactone

    (PCL) scaffolds used have been studied extensively for

    bone engineering applications. The scaffold fibres are

  • 8/7/2019 Flow Modelling

    4/16

    184 H. Singh et al. / Journal of Biotechnology 119 (2005) 181196

    Fig. 2.2. Profile of novel bioreactor.

    300m diameter and form a 90 lay-down pattern, to

    form a scaffold of regular architecture with dimensions

    of approximately 5 mm 5 mm 5 mm (Fig. 2.3). As

    with the bioreactor models, the point of origin of the

    scaffold lies at its very centre.

    2.4. Scenarios, criteria and guidelines

    The following set of simulations were performed,

    and are as follows:

    Bioreactor: rotation about the horizontal X-axis

    (with scaffold);

    Bioreactor: rotation about the vertical Z-axis (with

    scaffold);

    Bioreactor: bi-axial rotation about the horizontal and

    vertical axes (with scaffold).

    Table 2.1

    Bioreactor dimensions

    Component Dimensions (mm)Vessel diameter 94

    Vessel height 130

    Inlet length 6

    Inlet diameter 6

    Inlet distance from centre (radially) 23.5

    Outlet tube length 100

    Outlet tube inner diameter 6

    Outlet tube outer diameter 9

    Outlet distance from centre (radially) 23.5

    Outlet tube through-hole diameter (6) 3

    Vertical distance between holes 16

    Fig. 2.3. MicroCT of a scaffold produced by rapid-prototyping

    (Hutmacher et al., 2004).

    Firstly, it must be noted that simulations (without

    scaffolds) were used to determine appropriate locations

    for placing the scaffolds. Moreover, flow phenomena

    were studied, and the locations of turbulent artifacts

    such as eddies and vortices were studied. Consistency

    and uniformity of flow throughout the chamber were

    also criteria that were particularly sought for. Where

    possible, regions of very low or stagnant fluid velocity

    were also avoided, as placing scaffolds within these

    regions would not possibly allow for adequate flow of

    culture media within and through the scaffolds, therebypotentially reducing the flow of nutrients to cells, as

    well as the ability of waste to be removed.

    Secondly, only the Z-axis (vertical centre-line pass-

    ing through the vessel) was utilised as a line of refer-

    ence, whereby specific positions along this line were

    identified for placement of scaffolds. The main reason

    for this is that shear stresses tend to be minimised at

    distances furthest from the walls. In this way, exces-

    sive shear stresses could potentially be avoided. Addi-

    tionally, points along this reference line having higher

    velocities were selected for scaffold fixation. This was

    done, bearing in mind that there were no eddies or tur-

    bulent artifacts within close proximity to the selected

    positions.

    2.5. Conditions, settings and meshing

    Table 2.2 indicates values and parameters that were

    used for the simulations. The contour and velocity plots

    were captured at equal intervals of 0.0235 m, dividing

    the entire vessel into sections of four. On the whole,

  • 8/7/2019 Flow Modelling

    5/16

    H. Singh et al. / Journal of Biotechnology 119 (2005) 181196 185

    Table 2.2

    Boundary conditions and settings

    Culture media

    Density (kg/m3) 1030

    Dynamic viscosity (Pa s) 0.0025

    Boundary conditions

    Inlet velocity (m/s) 0.5895

    Vessel rotational velocity (rpm) 35

    Gravity (+Zdimension) (m/s2) 9.81

    Schemes

    Solver: segregated, implicit

    Flow: laminar

    Pressure: body-force weighted

    Pressurevelocity coupling: PISO

    Momentum: second order upwind

    three sections were at equal intervals along the Y-axis

    (Planes 13), enabling viewing from the front. Another

    three equal sections were taken at the side, enabling

    analysis from the transverse perspective point of view.

    These transverse sections (Planes 46) were conse-

    quently captured along the X-axis.

    All the axial conventions must be appropriately

    noted and strictly adhered to. By plotting for the X-,Y- and Z-directions, the location of the maximum

    wall shear stress was determined and pinpointed. The

    shear stress at this specific location would thereforebe expected to have adverse effects on cells, and very

    possibly resulting in their death. Counter-measures

    could therefore be devised, so as to prevent damage to

    cells at such locations.

    The mesh applied (Fig. 2.4) here is that of a tetrahe-

    dral mesh. GAMBIT is generally capable of meshing

    Fig. 2.4. Sectioned mesh of scaffold via central plane.

    an entire volume at one go, by applying the same

    meshing constraints throughout, such as a constant

    interval count. The outlet tube, however, was meshed

    individually by meshing its faces due to complexcurved surfaces. Meshing was carried out face by face,

    with an interval count of 20. Each individual scaffold

    fibre length was meshed with a 10-interval count. The

    remaining volume was the meshed with an interval

    count of 50 throughout Mesh convergence analyses

    were carried out to prior to the actual simulations

    to select a suitable mesh density for this study. Four

    sets of 3D meshes of the bioreactor vessel alone were

    generated at varying degrees of mesh refinement at

    (i) 50606, (ii) 227191, (iii) 282440 and (iv) 321674

    tet. elements. All meshes converged successfully with

    a convergence error criterion of 0.001. Velocity pro-

    files were compared in FLUENT at specific planes and

    points. However, only a slight increase in accuracy was

    noted (approx. 3%) despite a significant increase in the

    number of elements and computational time from (iii)

    to (iv). Mesh (iii) was therefore selected based on its

    accuracy and its economy.

    Fig. 2.5indicates the sections taken of the bioreactor

    for our study. Fig. 2.6 presents a section of the scaffold

    as viewed from the front (sectioned vertically), with

    Points 15 labelled for our analysis. Fig. 2.7 presents

    the locations of Points 610 within the scaffold, asviewed from the underside when the scaffold is sec-

    tioned horizontally. Fig. 2.8depicts a single pore within

    the scaffold, visually presenting us with flow parame-

    ters that will be discussed throughout this article.

    Fig. 2.5. Plan view of vessel and corresponding sections along XY

    planes with planes spaced at equal intervals (quarter-distance) of

    0.0235m.

  • 8/7/2019 Flow Modelling

    6/16

    186 H. Singh et al. / Journal of Biotechnology 119 (2005) 181196

    Fig. 2.6. Scaffold sectioned along centralXZplane (front view) with

    specific points and axis labelled.

    Fig. 2.7. Scaffold sectioned along central XYplane (as viewed from

    underside of scaffold) with specific points and axis labelled.

    Fig. 2.8. Depiction of a scaffold pore with velocityvectors and shear

    stresses acting along all surfaces.

    Particular attention waspaidto the flowenvironment

    within the scaffolds. Prior to this (not shown here), the

    most suitable locations were determined for placing the

    scaffolds, for each and every rotational scheme. Thiswas done to maximise the performance of each con-

    figuration. The configurations were then finally com-

    pared.

    3. Results and discussion

    3.1. Rotation about X-axis

    Fig. 3.1.1 reveals the contour plots captured at

    Planes 13, while Fig. 3.1.2 depicts contour plots at

    side Planes 46. The scaffold centre was positioned at

    the middle of the bioreactor at coordinate (0,0,0).

    Fig. 3.1.3 displays the scaffold as sectioned verti-

    cally, through its centre. As with the previous scenario,

    Points 15 were selected for analysis, while Fig. 3.1.4

    reveals the presence of an eddy. This eddy can be

    observed to exist behind the scaffold. Additionally, it

    can also be seen from both figures that the vectors seem

    to be of higher intensity near the front-right corner

    of the scaffold. It is very likely that escalated levels of

    wall shear stresses would be expected to occur along

    the external surface of the scaffold at these regions.Table 3.1 displays the velocity magnitudes and u,

    v and w components for locations within the scaffold,

    as designated by Points 15. Points 5 and 4 experience

    the highest fluid velocity magnitudes at approximately

    3.52 103 and 3.01 103 m/s, respectively, while

    Points 1 and 2 experience the lowest velocities, at

    1.25 103 and 1.49 103 m/s. The primary flow

    direction is noted to be in the Y-direction. Point 7 can be

    seen to experience a relatively high velocity magnitude

    with respect to the other points, at 2.81 103 m/s.

    The dominant v velocity component is approximately2.6 103 m/s.

    Fig. 3.1.5 depictsthe wall shear stresses actingalong

    the surfaces of the scaffold fibres. The average wall

    shear stresses were found to generally lie between 0.8

    and 1.2 Pa, as data for a variety of locations within the

    scaffoldwere extracted and counter-checked.However,

    specific regions were seen to indicate high peaks in wall

    shear stresses.

    It can also be noticed that at the very front of the

    diagram, a lighter shade of blue can be seen represent-

  • 8/7/2019 Flow Modelling

    7/16

    H. Singh et al. / Journal of Biotechnology 119 (2005) 181196 187

    Fig. 3.1.1. Velocity contour plots captured at frontal Planes 13 for bioreactor with scaffold rotating about X-axis.

    Fig. 3.1.2. Velocity contour plots captured at side Planes 46 for bioreactor with scaffold rotating about X-axis.

    Table 3.1

    Velocities at Points 110 within scaffoldbioreactor rotation about X-axis

    Points

    1 2 3 4 5 6 7 8 9 10

    Magnitude (103 m/s) 1.25 1.49 2.12 3.01 3.52 1.77 2.81 1.78 2.04 1.68

    x (103 m/s) 0.35 0.50 0.29 0.25 0.31 0.13 0.40 0.30 0.33 0.50

    y (103 m/s) 1.20 1.40 2.10 3.00 3.50 1.25 2.60 1.75 1.75 1.00

    z (103 m/s) 0.04 0.12 0.08 0.04 0.14 1.25 1.00 0.05 1.00 1.25

  • 8/7/2019 Flow Modelling

    8/16

    188 H. Singh et al. / Journal of Biotechnology 119 (2005) 181196

    Fig. 3.1.3. Sectioned scaffold and velocity vector plot at y = 0 m plane within bioreactor rotating about X-axis.

    ing approximately 1.2 Pa in shear stresses. This occurs

    near the zones of contact between criss-crossing scaf-

    fold fibres.It is also noticed that forthe externalsurface,

    the colour-coded dark blue colour softens to cyan

    and light green, from left to right. This implies that

    the higher wall shear stresses are located towards the

    right side (positive X-direction) of the scaffold. One

    very likely reason for this phenomenon relates to thedesign of the bioreactor. It is suspected that the pres-

    ence of the outlet tube tends to deviate and diverge the

    swirling fluid, reducing fluid velocity andshear stresses

    especially in areas within close proximity to the outlet

    tube. The fluid then accelerates, increasing in veloc-

    ity due to the rotating action of the bioreactor, thereby

    resulting in higher shear stresses to the right (positive

    X-direction) of the scaffold. The location of the peak

    wall shear stress can be seen within Fig. 3.1.5, as theminiscule red spot (indicated by the arrow).

    Fig. 3.1.4. Sectioned scaffold and velocity vector plot at z = 0 m plane within bioreactor rotating about X-axis.

  • 8/7/2019 Flow Modelling

    9/16

    H. Singh et al. / Journal of Biotechnology 119 (2005) 181196 189

    Fig. 3.1.5. Wall shear stressrange of 04Pa forscaffold withinbiore-actor rotating about X-axis.

    3.2. Rotation of bioreactor with scaffold about

    Z-axis

    For this scenario, the scaffold was positioned at the

    coordinate (0,0,0.028), 28 mm above the origin rela-

    tive to our directional convention. Figs. 3.2.1 and 3.2.2

    depict similarities that can be seen to occur with respect

    to the Planes 2 and 5, where a central core of slow-

    moving fluid is noted. This is pronounced in Fig. 3.2.2.As previously mentioned, the formation of the core

    is possibly due to fluid velocity being a function of

    the angular velocity and radius of the chamber. As

    the angular velocity is held constant, the fluid velocity

    increases proportionally as the distance from the rotat-

    ing center increases. Furthermore, centrifugal forcestend to displace fluid outwards. In Fig. 3.2.2, aneddy

    can be seen to have formed, as shown in Plane 5. The

    location of this entity is approximately one third of

    the bioreactor height from the base, directly below the

    scaffold. This recirculation zone can be related to the

    corresponding contour plot in Fig. 3.2.2, within the

    dark-blue zone that is in the bottom half of the ves-

    sel. Plane 6 also reveals another eddy, corresponding

    in location to thecontour plot. This phenomenon occurs

    near the base of the vessel.

    Fig. 3.2.3 displays the sectioned scaffold (as viewed

    from the front) where velocity vectors seem to be

    deflected by the scaffold. This occurs on the right and

    bottom sides of the scaffold. Additionally, a small eddy

    can be seen to have formed just slightly to the left of

    the scaffold (Fig. 3.2.3). The next frame depicts the

    scaffold sectioned along the horizontal plane, wherez = 0 m. Again, the rear (bottom of figure) and right

    faces of the scaffold experience heightened levels of

    flow (Figs. 3.2.4 and 3.2.5).

    Table 3.2 indicates that Points 2 and 5 experi-

    ence similarly high fluid velocity magnitudes, being

    approximately 3.8 10

    3 and 3.76 10

    3 m/s. Thisrelates well to Fig. 3.2.3, which shows that the vec-

    tors approaching from the right are deflected upwards.

    Fig. 3.2.1. Velocity contour plots captured at frontal Planes 13 for bioreactor with scaffold rotating about Z-axis.

  • 8/7/2019 Flow Modelling

    10/16

    190 H. Singh et al. / Journal of Biotechnology 119 (2005) 181196

    Fig. 3.2.2. Velocity contour plots captured at side Planes 46 for bioreactor with scaffold rotating about Z-axis.

    Fig. 3.2.3. Sectioned scaffold and velocity vector plot at y = 0 m plane within bioreactor rotating about Z-axis.

    Table 3.2

    Velocities at Points 110 within scaffoldbioreactor rotation about Z-axis

    Points

    1 2 3 4 5 6 7 8 9 10

    Magnitude (103 m/s) 0.71 3.80 2.02 1.92 3.76 0.94 5.06 2.02 1.04 4.08

    x (103 m/s) 0.08 0.16 0.22 0.23 0.08 0.75 0.75 0.10 0.90 0.75

    y (103 m/s) 0.65 3.80 2.00 1.80 3.75 0.50 5.00 2.00 0.40 4.00

    z (103 m/s) 0.28 0.05 0.21 0.63 0.20 0.25 0.13 0.23 0.33 0.35

  • 8/7/2019 Flow Modelling

    11/16

    H. Singh et al. / Journal of Biotechnology 119 (2005) 181196 191

    Fig. 3.2.4. Sectioned scaffold and velocity vector plot at z =0.028 m plane within bioreactor rotating about Z-axis.

    This is further confirmed by the v velocity compo-

    nents, which prove that their major velocity contri-

    butions are particularly attributed to the fluid flow in

    the Y-direction. Point 1, though, experiences a lower

    magnitude of velocity, at about 0.71 103 m/s. This

    indicates that cells at this location may be derived of

    essential nutrient transport.The wall shear stress plot for the scaffold again indi-

    cates that higherstresses tend to occur at theouteredges

    and faces, which are in direct contact with the mov-

    ing fluid. It is within the scaffold that these stresses

    Fig. 3.2.5. Wall shear stressrange of 02Pa forscaffold withinbiore-

    actor rotating about Z-axis.

    decrease significantly, to a value of 1 Pa and less. Upon

    careful examination, it is revealed that higher stresses

    here not only occur along the outer surfaces of the scaf-

    fold, but especially at the areas where scaffold fibres

    intersect along the outer faces of the scaffold. The aver-

    age shear stresses along the scaffold fibres within the

    scaffoldrangefrom 0.2to 0.6 Pa approximately. In con-trast, shear stresses at the fibre intersections range from

    0.8 to 1.6 Pa. It is clear that the shear stresses acting on

    and within the scaffold are not uniform throughout,

    but are dependent on the direction of flow, too. This is

    despite the relatively small dimensions of the scaffold

    of approximately 5 mm length per side.

    3.3. Bi-axial rotation of bioreactor with scaffold

    This scenario involves positioning the scaffold at

    z =

    0.02 m, along the vertical centre-line of the ves-sel. From Fig. 3.3.1, pockets of slow-moving fluid, as

    indicated by the dark-blue region captured at Planes 1

    and 3. However, for Plane 2, a significant portion of this

    dark-blue contour seems to have filled up a large por-

    tion of the chamber, primarilyat the bottom-half region.

    Closer examination of the corresponding velocity vec-

    tor plot reveals the presence of an eddy at the bottom

    of the vessel, as indicated by the recirculating vectors.

    This small recirculation body corresponds in location

    to the small extension, as can be seen in Plane 2 of

  • 8/7/2019 Flow Modelling

    12/16

    192 H. Singh et al. / Journal of Biotechnology 119 (2005) 181196

    Fig. 3.3.1. Velocity contour plots captured at frontal Planes 13 for bioreactor with scaffold rotating bi-axially.

    Fig. 3.3.1, to be extending downwards to the bottom-

    right of the chamber (Fig. 3.3.2).

    Fig. 3.3.3 depicts velocity vectors being deflected

    upwards and to the left by the scaffold. This vector

    plot also suggests that the top-right corner would espe-

    cially experience relatively higher wall stresses, due to

    the intensity of the arrows at that location. A small

    eddy can also be seen slightly to the left of the scaf-fold. The next vector plot reveals an eddy to the left

    of the scaffold. Here, the faster-moving fluid flowing

    along the rear of the scaffold can be seen entering the

    scaffold and exiting from the front face of the scaffold,

    along the positive Y direction. It is of interest to note

    that from Table 3.3, all 10 locations experience sim-

    ilar magnitudes of velocity, ranging from 23 103

    to 29 103 m/s. This indicates that improved flow

    and mixing can potentially be achieved by adopting

    this scheme. We also see that the magnitudes haveincreased by 1 order, indicating that this rotational

    scheme improves fluid flow significantly. The data fur-

    Fig. 3.3.2. Velocity contour plots captured at side Planes 46 for bioreactor with scaffold rotating bi-axially.

  • 8/7/2019 Flow Modelling

    13/16

    H. Singh et al. / Journal of Biotechnology 119 (2005) 181196 193

    Fig. 3.3.3. Sectioned scaffold and velocity vector plot at y = 0 m plane within bioreactor rotating bi-axially.

    Table 3.3

    Velocities at Points 110 within scaffoldbioreactor rotation about bi-axial XZ-axes

    Points

    1 2 3 4 5 6 7 8 9 10

    Magnitude (103 m/s) 28.51 28.01 28.31 26.00 24.01 27.72 28.01 27.53 27.55 23.14

    x (103 m/s) 0.20 0.05 0.60 0.40 0.05 0.75 0.50 1.25 2.50 2.50

    y (

    10

    3

    m/s) 28.50 28.00 28.30 26.00 24.00 27.70 28.00 27.50 27.40 23.00z (103 m/s) 0.50 0.75 0.38 0.05 0.50 0.55 0.40 0.10 1.35 0.10

    Fig. 3.3.4. Sectioned scaffold and velocity vector plot at z =0.02 m plane within bioreactor rotating bi-axially.

  • 8/7/2019 Flow Modelling

    14/16

    194 H. Singh et al. / Journal of Biotechnology 119 (2005) 181196

    Fig. 3.3.5. Wall shearstress rangeof 08Pa forscaffold withinbiore-

    actor rotating bi-axially.

    ther indicates that the X-velocities generally do not

    seem to be major constituents of the respective magni-

    tudes, unlike the v or Y-velocities (Fig. 3.3.4).

    Fig. 3.3.5 displays the shear stresses along surfaces

    as well as within the scaffold. It is noted that shear

    stresses seem to increase in the positive X direction,

    indicating that the maximum wall shear stresses occurs

    at the right side of the scaffold. This is verified by

    Fig. 3.3.5, which isolates wall shear stresses within

    the range of 08 Pa. It can be seen that the intersect-ing fibres tend to experience higher stresses at points

    of intersection, as highlighted by the green-coloured

    streaks. The nominal shear stresses within the scaffold,

    however, seem to be generally lower, at approximately

    1.62 Pa, as indicated by the lighter-blue colour. This

    is in contrast to the stresses experienced at the fibre

    intersections, which fall within the range of 2.44.8 Pa

    approximately.

    In summary, the most outstanding differences relate

    to the bi-axial or gyroscopically rotating bioreactor.

    The velocity magnitudes at each of the locations fromPoints 1 to 10 have increased significantly by manifold,

    and up to one order of calculation in some cases. The

    data clearly prove that bi-axial rotationof thebioreactor

    results in the increase of fluid flow within the scaffold

    under the studied conditions. At point 1, a 22.79 ratio

    of fluid velocities is noted; by comparing the velocities

    generated by the bi-axial andXrotational schemes (see

    Table 3.1 and Fig. 3.4). Consequently, a ratio of 40.02

    for Point 1 was also noted, by comparing the veloci-

    ties due to the bi-axial and Z rotational schemes (see

    Fig. 3.4. Fluid velocity magnitudes at Points 15 within scaffold

    subjected to rotation about respective axes.

    Table 3.2 and Fig. 3.5). The above results indicate a

    significant enhancement of fluid velocity and mixing,

    due to the bi-axial mode of bioreactor rotation. Bi-axialrotation clearly enhances mixing of the fluid, as fluid

    particles rotate under the influence of two axes. This

    increases the penetrability of particles into the scaffold,

    as they can enter the pores of the scaffold from more

    than one direction. It is also noted that fluid mixing has

    improved due to the bi-axial rotation, as compared to

    uni-axial rotation. This is indicated by breaking up of

    contours, contrary to the uniform contours as noted

    for the uni-axial rotation cases. It must be mentioned

    that bi-axial rotation, however, may result in a slight

    increase of turbulent artifacts such as eddies.The shear stresses within the scaffolds do not vary

    greatly for both uni-axially rotating cases. These gen-

    erally values range from 0.4 to 0.6 Pa for the proto-

    type model. However, bi-axial rotation of the prototype

    model reports an average shear stress value of 1.8 Pa

    withinthe scaffold. This valueof shear stressrepresents

    those that occur within the scaffold. Findings also show

    that the wall shear stresses acting along the external

    Fig. 3.5. Fluid velocity magnitudes at Points 610 within scaffold

    subjected to rotation about respective axes.

  • 8/7/2019 Flow Modelling

    15/16

    H. Singh et al. / Journal of Biotechnology 119 (2005) 181196 195

    edges of the scaffolds tend to be approximately three

    to four times more than the internal shear stresses as

    indicated. This is notedto particularlyoccurat theinter-

    sections of scaffold struts/bars and at the circular edgesof thefibre end-faces. Onereason for this is that stresses

    tend to be concentrated at the edges, which are directly

    exposed to the dynamic fluid.

    The Reynolds number is a commonly applied non-

    dimensional parameter that is applied to assess the state

    of a system, where

    Re =Vd

    or Re =

    r2

    (3.1)

    whereby V is the fluid velocity, d the diameter, r the

    radius and is the rotational velocity. The Reynoldsnumber represents a ratio of inertial to viscous effects

    and indicates laminar, transient or turbulent flow. For

    example, a flow scheme would tend towards the lam-

    inar scheme due to a higher fluid viscosity, reducing

    the Reynolds number of a fluid. The Reynolds num-

    ber is of interest in many bioreactor systems as laminar

    flow regimes are often employed. However, our sys-

    tem is more complex because bi-axial rotation within

    our asymmetric vessel is coupled with inlet and out-

    let flows. Hence, we did not attempt to calculate Re.

    The gravity and buoyancy effects are of less signifi-cance with respect to our bioreactor system, and were

    therefore neglected in our simulations.

    4. Conclusions

    1. Three different flow configurations were discussed

    with respect to our prototype bioreactor design, pri-

    marily rotation about the X (tumble), Z (spin) and

    XZ(bi-axial) axes.

    2. Fluid velocities and shear stresses within the scaf-folds were studied and compared, and proved that

    bi-axial rotation results in significant improvements

    in terms of fluid transport through the scaffolds.

    This is due to the combined effects of the rota-

    tional velocity vectors about both axes, which

    when combined, would almost certainly result in a

    higher rotational velocity component, as compared

    to rotation about a single axis. However, a rise in

    shear forces was also reported. Greater transport

    within scaffolds for example, would therefore result

    in most cases (and not just ours) as a result of bi-

    axial rotation.

    3. These simulations assist in identifying critical

    issues and problems, for example, in approximat-ing the locations of recirculation zones. These

    zones may potentially damage cells and inhibit

    growth. Furthermore, these recirculating bodies

    may impede the flow of fluid into and out of the

    scaffold. The choice of flow regime is therefore of

    great importance.

    References

    Bancroft, G.,Sikavitsas, V., Mikos,A., 2003. Design of a flow perfu-

    sion bioreactor system for bone tissue-engineering applications.Tissue Eng. 9 (3), 549554.

    Boyd, E.J., Gonda, S., 1990. Mathematical modelling of the flow

    field and particle motion in a rotating bioreactor at unit gravity

    and microgravity. In: NASA/ASEE Summer Faculty Fellowship

    Program, Johnson Space Centre, August 23.

    Bursac, B., Papadaki, M., White, J., Eisenberg, S., Vunjak-

    Novakovic, G., Freed, L.E., 2003. Cultivation in rotating bioreac-

    torspromotesmaintenanceof cardiacmyocyteelectrophysiology

    and molecular properties. Tissue Eng. 9 (6), 12431253.

    Cherry, R.S.,Papoutsakis, E.T., 1986. Hydrodynamic effects on cells

    in agitated tissue culture reactors. Bioprocess Eng. 1, 2941.

    Darling, E., Athanasiou, K., 2003. Articular cartilagebioreactors and

    bioprocesses. Tissue Eng. 9 (1), 926.

    Freed, L.E., Vunjak-Novakovic, G., 2002. Spaceflight bioreactorstudies of cells and tissues. Adv. Space Biol. Med. 8, 177195.

    Freed, L.E., Vunjak-Novakovic, G., Biron, R.J., Eagles, D.B.,

    Lesnoy, D.C., Barlow, S.K., Langer, R., 1994. Biodegradable

    polymer scaffolds for tissue engineering. Biotechnology 12,

    689693.

    Hutmacher, D.W., Risbud, M., Sittinger, M., 2004. Evolution of

    computer aided design and advanced manufacturing in scaffold

    research. Trends Biotechnol. 22 (7), 354362.

    Kleis, S.J., Pellis, N.R., 1995. Fluid dynamic verification experi-

    ments on STS-70. In: NASA/ASEE Summer Faculty Fellowship

    Program, Johnson Space Centre, August 9.

    Kleis, S.J., Begley, C., Pellis, N.R., 1996. Bioreactor mass transport

    studies. In: NASA/ASEE Summer Faculty Fellowship Program,Johnson Space Centre, August 2.

    Lappa, M., 2005. A CFD level-set method for soft tissue growth:

    theory and fundamental equations. J. Biomech. 38, 185190.

    Levesque, M.J., Nerem, R.M., 1985. The elongation and orienta-

    tion of cultured endothelial cells in response to shear stress. J.

    Biomech. Eng. 107, 341347.

    Levesque, M.J., Sprague, E.A., Schwartz, C.J., Nerem, R.M., 1989.

    The influence of shear stress on cultured vascular endothelial

    cells: the stress response of an anchorage-dependent mammalian

    cell. Biotechnol. Progr. 5, 18.

    Martin, I., Wendt, D., Heberer, M., 2004. The role of bioreactors in

    tissue engineering. Trends Biotechnol. 22 (2), 8086.

  • 8/7/2019 Flow Modelling

    16/16

    196 H. Singh et al. / Journal of Biotechnology 119 (2005) 181196

    Olivier, A.L., Truskey, G.A., 1993. A numerical analysis of

    forces exerted by laminar flow on spreading cells in a par-

    allel plate flow chamber assay. Biotechnol. Bioeng. 42, 963

    973.

    Papoutsakis, E.T., 1991. Fluid-mechanical damage of animal cells inbioreactors. Trends Biotechnol. 9, 427437.

    Porter, B., Zauel, R., Stockman, H., Guldberg, R., Fyhrie, D., 2005.

    3-D computational modeling of media flow through scaffolds in

    a perfusion bioreactor. J. Biomech. 38, 543549.

    Pritchard, W.F., Davies, P.F., Derafshi, Z., Polacek, D.C., Tsao, R.,

    Dull, R.O.,Jones, S.A., Giddens, D.P., 1995. Effectsof wall shear

    stress and fluid recirculation on the localization of circulating

    monocytes in a three-dimensional flow model. J. Biomech. 28,

    14591469.

    Raimondi,M.T., Boschetti,F.,Falcone, L.,Migliavacca, F., Remuzzi,

    A., Dubini, G., 2004. The effect of media perfusion on three-

    dimensional cultures of human chondrocytes: integration of

    experimental and computational approaches. Biorheology 41,

    401410.

    Redaelli,A., Boschetti, f.,Inzoli,F.,1997. Theassignment of velocity

    profiles in finite element solutions of pulsatile flow in arteries.

    Comput. Biol. Med. 27 (3), 233247.

    Sodian, R.,Lemke, T., Fritsche,C., Hoerstrup, S.,Fu, P., Potapov, E.,

    Hausmann, H., Hetzer, R., 2002. Tissue-engineering bioreactors:

    a new combined cell-seeding and perfusion system for vascular

    tissue engineering. Tissue Eng. 8 (5), 863870.

    Stathopoulos, N.A., Hellums, J.D., 1985. Shear stress effects onhuman embryonic kidney cells in vitro. Biotechnol. Bioeng. 27,

    10211028.

    Sucosky, P., Osorio, D.F., Brown, J.B., Neitzel, G.P., 2003. Fluid

    mechanics of a spinner-flask bioreactor. Biotechnol. Bioeng. 85

    (1), 3446.

    Vunjak-Novakovic, G., Freed, L.E., Biron, R.J., Langer, R., 1996.

    Effects on mixing on tissue engineered cartilage. AIChE J. 42,

    850860.

    Vunjak-Novakovic, G., Obradovic, B., Bursac, P., Martin, I., Langer,

    R., Freed, L.E., 1998. Dynamic seeding of polymer scaffolds for

    cartilage tissue engineering. Biotechnol. Progr. 14, 193202.

    Yamada, H., Takemasa, T., Yamaguchi, T., 2000. Theoretical study

    of intracellular stress fiber orientation under cyclic deformation.

    J. Biomech. 33, 15011505.

    Zein, I., Hutmacher, D.W., Teoh, S.H., Tan, K.C., 2002. Poly(e-

    caprolactone) scaffolds designed and fabricated by fused depo-

    sition modeling. Biomaterials 23 (4), 11691185.