Anda di halaman 1dari 16

Journal of Biotechnology 119 (2005) 181–196

Flow modelling within a scaffold under the influence of


uni-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, Singapore
b 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 better
understanding 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

1. Introduction

In a number of long-term tissue engineering stud-


∗ Corresponding author. Tel.: +65 6874 1036; fax: +65 6872 3069. ies under static conditions (e.g. Petri dish), it has
E-mail addresses: g0301488@nus.edu.sg (H. Singh), proven extremely difficult to promote the high-density
mpetsh@nus.edu.sg (S.H. Teoh), mpelowht@nus.edu.sg three-dimensional in vitro growth of cells that have
(H.T. Low), biedwh@nus.edu.sg (D.W. Hutmacher).
1 Tel.: +65 6874 8870; fax: +65 6777 3537. been removed from the body and deprived of their
2 Tel.: +65 6874 4605; fax: +65 6872 3069. normal in vivo vascular sources of nutrients and gas
3 Tel.: +65 6874 2225; fax: +65 6872 3069. 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
182 H. Singh et al. / Journal of Biotechnology 119 (2005) 181–196

ery, as well as to mimic the fluid dynamic environment and Gonda (1990) mathematically modelled the motion
present within the body, a dynamic system has to be of particles within the NASA bioreactor model at both
chosen. The bioreactor is one example that attempts unit and micro-gravity. However, the effects of gravity
to fulfil these requirements, particularly in the study can be felt by almost everything on Earth, and is not
of tissue engineering, by simulating the physiological ignored here for practical reasons. Rotating (Bursac
and fluidic conditions that occur in vivo (Freed and et al., 2003; Martin et al., 2004) and perfusion (Martin
Vunjak-Novakovic, 2002; Papoutsakis, 1991; Cherry et al., 2004; Bancroft et al., 2003; Darling and
and Papoutsakis, 1986). Cells have often been seen to Athanasiou, 2003; Sodian et al., 2002) bioreactors,
grow well along the periphery of 3D scaffold, while each with their own flow characteristics, are currently
proliferation is often significantly affected at the centre being utilized for studies involving the culturing of
of the scaffold, where necrotic neo-tissue can some- bone and cartilage. These bioreactor models induce dif-
times be seen. This is partly due to poor fluidic transport ferent normal and shear forces that act on cells, with
of media and scaffold design, among other reasons. varying consequences.
Therefore, simulations that provide such flow visual- One aspect that this paper will focus on is that of
izations could greatly assist in the design of scaffolds shear stresses acting on cells and tissues. A response
as well as bioreactor. of an endothelial monolayer of cells subject to steady,
A number of experimental studies were previously laminar shear stresses is that of an alteration in mor-
reported by using different bioreactor systems (Freed phology. A typical change would occur from an ini-
et al., 1994; Martin et al., 2004; Vunjak-Novakovic et tial polygonal pattern, to one, which depicts an elon-
al., 1996). However, limited information has been pub- gated profile, being aligned to the direction of fluid
lished on studies that attempt to simulate the dynamic flow (Levesque and Nerem, 1985; Levesque et al.,
fluid environment prior to commencing experimen- 1989; Stathopoulos and Hellums, 1985). This change
tal studies. Advantages of computational simulations is basically accompanied by the restructuring of the
include the ability to modify and study the effects of cytoskeleton, or more specifically, the alignment of
bioreactor design with respect to the flow analysis, microtubules, followed by the formation of actin stress
without having to develop and construct actual phys- fibres. Similar experiments were conducted on bovine
ical models, or to run a large number of experiments. endothelial cells. It was found that the stiffness of
This is coupled with the significant savings in time and endothelial cells exposed to shear stresses of 2 Pa,
costs. Furthermore, visualisation of flow as enabled by increased with the duration time of exposure. However,
the simulation package is a key factor in determining after 24 h of exposure, the stiffness of the endothe-
the efficacy of the system, and allows for design opti- lial cells was similar all around the cell, indicating
mization prior to bioreactor design and modifications. the ability of the cells to adapt to the changes in the
NASA has taken significant strides relating to their environment. This concept includes that of stress fibre
rotating bioreactor studies, under the influence of unit orientation, as mentioned in the paper by Yamada et al.
and micro-gravity. While it is known that in the pres- (2000).
ence of body forces, density differences between the Olivier and Truskey (1993) showed that the flow
cells attached to micro-carriers and the fluid medium, regime itself, and not just the magnitude of shear
cause relative motion resulting in mechanical shear stress, plays a critical part. Their experiments involved
and increased mass transport. However, in the micro- turbulent flows as generated in a cone and plate vis-
gravity environment, buoyancy effects are greatly cometer, which actually resulted in the detachment of
reduced. The gravity of Earth is replaced by centripetal anchorage-dependent cells at shear stresses of only
acceleration as the dominant body force. For a typ- 1.5 dyn/cm2 .
ical rotation rate of 2 rpm, within a 0.05 m diameter At relatively higher levels of shear stress, rang-
vessel, the magnitude of the body force is reduced ing from 26 to 54 dyn/cm2 , human endothelial cells
(Kleis and Pellis, 1995) to approximately 0.001 m/s2 . were found to detach from its surface and correspond-
Kleis et al. (1996) proceeded further by carrying ingly exhibited reduced viability (Stathopoulos and
out studies on mass transport, with regards to their Hellums, 1985). In contrast to this, no cell loss was
micro-gravity bioreactor model. Consequently Boyd reported for bovine aorta endothelial cells, which were
H. Singh et al. / Journal of Biotechnology 119 (2005) 181–196 183

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 documented
that 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 Fig. 2.1. Actual bioreactor prototype model.
found that these cells can be “conditioned” to grow
under the correct shear stress levels. Chondrocytes have ments of the flow solution. Flow solutions are based on
been known to proliferate better under the influence of the conservation of mass, momentum and energy equa-
shear stresses and are able to withstand higher loads as tions. GAMBIT and FLUENT were therefore used in
compared to chondrocytes cultured statically (Sucosky tandem, to model the flow of fluid within our prototype
et al., 2003). Both Porter et al. (2005) and Raimondi bioreactor model.
et al. (2004) utilised CFD techniques to model the
effects of perfusion, to better comprehend the influence 2.2. The bioreactor model
of perfusion and hence shear stresses, on 3D cultures.
Lappa (2005) on the other hand, developed a rigorous A novel bi-axial bioreactor system was developed
set of equations to model fluid flow and algorithms to by a team from the National University of Singapore
estimate soft tissue growth within a bioreactor. Redaelli and the Singapore Polytechnic. The unique design of
et al. (1997) on the other hand, attempted to model the entire system adds some flexibility that is some-
pulsatile flow within arteries. This intriguing article times found to be lacking in commercially available
describes how a FORTRAN algorithm was interfaced bioreactors. This novel design involves the bioreac-
to FIDAP, a commercially available CFD package, and tor being vertically upright, instead of the conven-
then simulated. 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
2. Software and methodology about the vertical axis (spinning), the horizontal axis
(tumbling), or even rotation simultaneously about both
2.1. GAMBIT and FLUENT axes (gyroscopic action). Ideally, the utilisation of these
features would lead to enhanced cell culture media flow
GAMBIT (Fluent Inc.), being a powerful graph- throughout the entire vessel (see Fig. 2.2 and Table 2.1).
ical modeling tool, was utilised as a pre-processor
for the Computational Fluid Dynamics package, FLU- 2.3. The tissue engineering scaffold
ENT. This general-purpose CFD tool can solve many
fluid flow problems such as steady, unsteady, lami- The scaffolds utilised for the simulations are rou-
nar and turbulent flows, heat transfer, Newtonian and tinely produced by our group (Zein et al., 2002;
non-Newtonian flows, among others. FLUENT also Hutmacher et al., 2004). The poly(␧)caprolactone
provides excellent mesh flexibility as well, allowing (PCL) scaffolds used have been studied extensively for
the grid to be refined or coarsened based on the require- bone engineering applications. The scaffold fibres are
184 H. Singh et al. / Journal of Biotechnology 119 (2005) 181–196

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


(Hutmacher et al., 2004).
Fig. 2.2. Profile of novel bioreactor.

Firstly, it must be noted that simulations (without


300 ␮m diameter and form a 90◦ lay-down pattern, to scaffolds) were used to determine appropriate locations
form a scaffold of regular architecture with dimensions for placing the scaffolds. Moreover, flow phenomena
of approximately 5 mm × 5 mm × 5 mm (Fig. 2.3). As were studied, and the locations of turbulent artifacts
with the bioreactor models, the point of origin of the such as eddies and vortices were studied. Consistency
scaffold lies at its very centre. and uniformity of flow throughout the chamber were
also criteria that were particularly sought for. Where
2.4. Scenarios, criteria and guidelines possible, regions of very low or stagnant fluid velocity
were also avoided, as placing scaffolds within these
The following set of simulations were performed, regions would not possibly allow for adequate flow of
and are as follows: culture media within and through the scaffolds, thereby
potentially reducing the flow of nutrients to cells, as
• Bioreactor: rotation about the horizontal X-axis well as the ability of waste to be removed.
(with scaffold); Secondly, only the Z-axis (vertical centre-line pass-
• Bioreactor: rotation about the vertical Z-axis (with ing through the vessel) was utilised as a line of refer-
scaffold); ence, whereby specific positions along this line were
• Bioreactor: bi-axial rotation about the horizontal and identified for placement of scaffolds. The main reason
vertical axes (with scaffold). for this is that shear stresses tend to be minimised at
distances furthest from the walls. In this way, exces-
Table 2.1 sive shear stresses could potentially be avoided. Addi-
Bioreactor dimensions tionally, points along this reference line having higher
Component Dimensions (mm) velocities were selected for scaffold fixation. This was
Vessel diameter 94
done, bearing in mind that there were no eddies or tur-
Vessel height 130 bulent artifacts within close proximity to the selected
Inlet length 6 positions.
Inlet diameter 6
Inlet distance from centre (radially) 23.5
2.5. Conditions, settings and meshing
Outlet tube length 100
Outlet tube inner diameter 6
Outlet tube outer diameter 9 Table 2.2 indicates values and parameters that were
Outlet distance from centre (radially) 23.5 used for the simulations. The contour and velocity plots
Outlet tube through-hole diameter (×6) 3 were captured at equal intervals of 0.0235 m, dividing
Vertical distance between holes 16
the entire vessel into sections of four. On the whole,
H. Singh et al. / Journal of Biotechnology 119 (2005) 181–196 185

Table 2.2 an entire volume at one go, by applying the same


Boundary conditions and settings meshing constraints throughout, such as a constant
Culture media interval count. The outlet tube, however, was meshed
Density (kg/m3 ) 1030 individually by meshing its faces due to complex
Dynamic viscosity (Pa s) 0.0025
curved surfaces. Meshing was carried out face by face,
Boundary conditions with an interval count of 20. Each individual scaffold
Inlet velocity (m/s) 0.5895
fibre length was meshed with a 10-interval count. The
Vessel rotational velocity (rpm) 35
Gravity (+Z dimension) (m/s2 ) 9.81 remaining volume was the meshed with an interval
count of 50 throughout Mesh convergence analyses
Schemes were carried out to prior to the actual simulations
Solver: segregated, implicit to select a suitable mesh density for this study. Four
Flow: laminar
sets of 3D meshes of the bioreactor vessel alone were
Pressure: body-force weighted
Pressure–velocity coupling: PISO generated at varying degrees of mesh refinement at
Momentum: second order upwind (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-
three sections were at equal intervals along the Y-axis files were compared in FLUENT at specific planes and
(Planes 1–3), enabling viewing from the front. Another points. However, only a slight increase in accuracy was
three equal sections were taken at the side, enabling noted (approx. 3%) despite a significant increase in the
analysis from the transverse perspective point of view. number of elements and computational time from (iii)
These transverse sections (Planes 4–6) were conse- to (iv). Mesh (iii) was therefore selected based on its
quently “captured” along the X-axis. accuracy and its economy.
All the axial conventions must be appropriately Fig. 2.5 indicates the sections taken of the bioreactor
noted and strictly adhered to. By plotting for the X-, for our study. Fig. 2.6 presents a section of the scaffold
Y- and Z-directions, the location of the maximum as viewed from the front (sectioned vertically), with
wall shear stress was determined and pinpointed. The Points 1–5 labelled for our analysis. Fig. 2.7 presents
shear stress at this specific location would therefore the locations of Points 6–10 within the scaffold, as
be expected to have adverse effects on cells, and very viewed from the underside when the scaffold is sec-
possibly resulting in their death. “Counter-measures” tioned horizontally. Fig. 2.8 depicts a single pore within
could therefore be devised, so as to prevent damage to the scaffold, visually presenting us with flow parame-
cells at such locations. ters that will be discussed throughout this article.
The mesh applied (Fig. 2.4) here is that of a tetrahe-
dral mesh. GAMBIT is generally capable of meshing

Fig. 2.5. Plan view of vessel and corresponding sections along X–Y
planes with planes spaced at equal intervals (quarter-distance) of
Fig. 2.4. Sectioned mesh of scaffold via central plane. 0.0235 m.
186 H. Singh et al. / Journal of Biotechnology 119 (2005) 181–196

Particular attention was paid to the flow environment


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. This
was 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


Fig. 2.6. Scaffold sectioned along central XZ plane (front view) with Planes 1–3, while Fig. 3.1.2 depicts contour plots at
specific points and axis labelled. side Planes 4–6. 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 1–5 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,
Fig. 2.7. Scaffold sectioned along central XY plane (as viewed from as designated by Points 1–5. Points 5 and 4 experience
underside of scaffold) with specific points and axis labelled. the highest fluid velocity magnitudes at approximately
3.52 × 10−3 and 3.01 × 10−3 m/s, respectively, while
Points 1 and 2 experience the lowest velocities, at
1.25 × 10−3 and 1.49 × 10−3 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 × 10−3 m/s.
The dominant v velocity component is approximately
−2.6 × 10−3 m/s.
Fig. 3.1.5 depicts the wall shear stresses acting along
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
scaffold were extracted and counter-checked. However,
specific regions were seen to indicate high peaks in wall
shear stresses.
Fig. 2.8. Depiction of a scaffold pore with velocity vectors and shear It can also be noticed that at the very front of the
stresses acting along all surfaces. diagram, a lighter shade of blue can be seen represent-
H. Singh et al. / Journal of Biotechnology 119 (2005) 181–196 187

Fig. 3.1.1. Velocity contour plots captured at frontal Planes 1–3 for bioreactor with scaffold rotating about X-axis.

Fig. 3.1.2. Velocity contour plots captured at side Planes 4–6 for bioreactor with scaffold rotating about X-axis.

Table 3.1
Velocities at Points 1–10 within scaffold—bioreactor rotation about X-axis
Points

1 2 3 4 5 6 7 8 9 10
Magnitude (×10−3 m/s) 1.25 1.49 2.12 3.01 3.52 1.77 2.81 1.78 2.04 1.68
x (×10−3 m/s) 0.35 0.50 0.29 0.25 0.31 0.13 0.40 0.30 0.33 0.50
y (×10−3 m/s) −1.20 −1.40 −2.10 −3.00 −3.50 −1.25 −2.60 −1.75 −1.75 −1.00
z (×10−3 m/s) −0.04 −0.12 −0.08 −0.04 −0.14 1.25 1.00 0.05 −1.00 −1.25
188 H. Singh et al. / Journal of Biotechnology 119 (2005) 181–196

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 ence of the outlet tube tends to deviate and diverge the
near the zones of contact between criss-crossing scaf- swirling fluid, reducing fluid velocity and shear stresses
fold fibres. It is also noticed that for the external surface, especially in areas within close proximity to the outlet
the colour-coded dark blue colour “softens” to cyan tube. The fluid then accelerates, increasing in veloc-
and light green, from left to right. This implies that ity due to the rotating action of the bioreactor, thereby
the higher wall shear stresses are located towards the resulting in higher shear stresses to the right (positive
right side (positive X-direction) of the scaffold. One X-direction) of the scaffold. The location of the peak
very likely reason for this phenomenon relates to the wall shear stress can be seen within Fig. 3.1.5, as the
design of the bioreactor. It is suspected that the pres- miniscule 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.
H. Singh et al. / Journal of Biotechnology 119 (2005) 181–196 189

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 forces
tend to “displace” fluid outwards. In Fig. 3.2.2, an eddy
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 the contour plot. This phenomenon occurs
near the base of the vessel.
Fig. 3.2.3 displays the sectioned scaffold (as viewed
Fig. 3.1.5. Wall shear stress range of 0–4 Pa for scaffold within biore-
actor rotating about X-axis.
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
3.2. Rotation of bioreactor with scaffold about can be seen to have formed just slightly to the left of
Z-axis the scaffold (Fig. 3.2.3). The next frame depicts the
scaffold sectioned along the horizontal plane, where
For this scenario, the scaffold was positioned at the z = 0 m. Again, the rear (bottom of figure) and right
coordinate (0,0,−0.028), 28 mm above the origin rela- faces of the scaffold experience heightened levels of
tive to our directional convention. Figs. 3.2.1 and 3.2.2 flow (Figs. 3.2.4 and 3.2.5).
depict similarities that can be seen to occur with respect Table 3.2 indicates that Points 2 and 5 experi-
to the Planes 2 and 5, where a central “core” of slow- ence similarly high fluid velocity magnitudes, being
moving fluid is noted. This is pronounced in Fig. 3.2.2. approximately 3.8 × 10−3 and 3.76 × 10−3 m/s. This
As previously mentioned, the formation of the core relates well to Fig. 3.2.3, which shows that the vec-
is possibly due to fluid velocity being a function of tors approaching from the right are deflected upwards.

Fig. 3.2.1. Velocity contour plots captured at frontal Planes 1–3 for bioreactor with scaffold rotating about Z-axis.
190 H. Singh et al. / Journal of Biotechnology 119 (2005) 181–196

Fig. 3.2.2. Velocity contour plots captured at side Planes 4–6 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 1–10 within scaffold–bioreactor rotation about Z-axis
Points

1 2 3 4 5 6 7 8 9 10
Magnitude (×10−3 m/s) 0.71 3.80 2.02 1.92 3.76 0.94 5.06 2.02 1.04 4.08
x (×10−3 m/s) −0.08 −0.16 −0.22 −0.23 −0.08 −0.75 −0.75 −0.10 0.90 0.75
y (×10−3 m/s) −0.65 3.80 2.00 −1.80 3.75 −0.50 5.00 2.00 0.40 4.00
z (×10−3 m/s) −0.28 −0.05 −0.21 −0.63 −0.20 −0.25 −0.13 −0.23 −0.33 −0.35
H. Singh et al. / Journal of Biotechnology 119 (2005) 181–196 191

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- decrease significantly, to a value of 1 Pa and less. Upon
nents, which prove that their major velocity contri- careful examination, it is revealed that higher stresses
butions are particularly attributed to the fluid flow in here not only occur along the outer surfaces of the scaf-
the Y-direction. Point 1, though, experiences a lower fold, but especially at the areas where scaffold fibres
magnitude of velocity, at about 0.71 × 10−3 m/s. This intersect along the outer faces of the scaffold. The aver-
indicates that cells at this location may be derived of age shear stresses along the scaffold fibres within the
essential nutrient transport. scaffold range from 0.2 to 0.6 Pa approximately. In con-
The wall shear stress plot for the scaffold again indi- trast, shear stresses at the fibre intersections range from
cates that higher stresses tend to occur at the outer edges 0.8 to 1.6 Pa. It is clear that the shear stresses acting on
and faces, which are in direct contact with the mov- and within the scaffold are not uniform throughout,
ing fluid. It is within the scaffold that these stresses 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, primarily at 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.
Fig. 3.2.5. Wall shear stress range of 0–2 Pa for scaffold within biore- This small recirculation body corresponds in location
actor rotating about Z-axis. to the small “extension, as can be seen in Plane 2 of
192 H. Singh et al. / Journal of Biotechnology 119 (2005) 181–196

Fig. 3.3.1. Velocity contour plots captured at frontal Planes 1–3 for bioreactor with scaffold rotating bi-axially.

Fig. 3.3.1, to be extending downwards to the bottom- along the rear of the scaffold can be seen entering the
right of the chamber (Fig. 3.3.2). scaffold and exiting from the front face of the scaffold,
Fig. 3.3.3 depicts velocity vectors being deflected along the positive Y direction. It is of interest to note
upwards and to the left by the scaffold. This vector that from Table 3.3, all 10 locations experience sim-
plot also suggests that the top-right corner would espe- ilar magnitudes of velocity, ranging from 23 × 10−3
cially experience relatively higher wall stresses, due to to 29 × 10−3 m/s. This indicates that improved flow
the “intensity” of the arrows at that location. A small and mixing can potentially be achieved by adopting
eddy can also be seen slightly to the left of the scaf- this scheme. We also see that the magnitudes have
fold. The next vector plot reveals an eddy to the left increased by 1 order, indicating that this rotational
of the scaffold. Here, the faster-moving fluid flowing scheme improves fluid flow significantly. The data fur-

Fig. 3.3.2. Velocity contour plots captured at side Planes 4–6 for bioreactor with scaffold rotating bi-axially.
H. Singh et al. / Journal of Biotechnology 119 (2005) 181–196 193

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 1–10 within scaffold—bioreactor rotation about bi-axial XZ-axes
Points

1 2 3 4 5 6 7 8 9 10
Magnitude (×10−3 m/s) 28.51 28.01 28.31 26.00 24.01 27.72 28.01 27.53 27.55 23.14
x (×10−3 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 23.00
z (×10−3 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.
194 H. Singh et al. / Journal of Biotechnology 119 (2005) 181–196

Fig. 3.4. Fluid velocity magnitudes at Points 1–5 within scaffold


subjected to rotation about respective axes.

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


Fig. 3.3.5. Wall shear stress range of 0–8 Pa for scaffold within biore- significant enhancement of fluid velocity and mixing,
actor rotating bi-axially. due to the bi-axial mode of bioreactor rotation. Bi-axial
rotation clearly enhances mixing of the fluid, as fluid
ther indicates that the X-velocities generally do not particles rotate under the influence of two axes. This
seem to be major constituents of the respective magni- increases the penetrability of particles into the scaffold,
tudes, unlike the v or Y-velocities (Fig. 3.3.4). as they can enter the pores of the scaffold from more
Fig. 3.3.5 displays the shear stresses along surfaces than one direction. It is also noted that fluid mixing has
as well as within the scaffold. It is noted that shear improved due to the bi-axial rotation, as compared to
stresses seem to increase in the positive X direction, uni-axial rotation. This is indicated by “breaking up” of
indicating that the maximum wall shear stresses occurs contours, contrary to the “uniform” contours as noted
at the right side of the scaffold. This is verified by for the uni-axial rotation cases. It must be mentioned
Fig. 3.3.5, which isolates wall shear stresses within that bi-axial rotation, however, may result in a slight
the range of 0–8 Pa. It can be seen that the intersect- increase of turbulent artifacts such as eddies.
ing fibres tend to experience higher stresses at points The shear stresses within the scaffolds do not vary
of intersection, as highlighted by the green-coloured greatly for both uni-axially rotating cases. These gen-
streaks. The nominal shear stresses within the scaffold, erally values range from 0.4 to 0.6 Pa for the proto-
however, seem to be generally lower, at approximately type model. However, bi-axial rotation of the prototype
1.6–2 Pa, as indicated by the lighter-blue colour. This model reports an average shear stress value of 1.8 Pa
is in contrast to the stresses experienced at the fibre within the scaffold. This value of shear stress represents
intersections, which fall within the range of 2.4–4.8 Pa those that occur within the scaffold. Findings also show
approximately. that the wall shear stresses acting along the external
In summary, the most outstanding differences relate
to the bi-axial or gyroscopically rotating bioreactor.
The velocity magnitudes at each of the locations from
Points 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 rotation of the bioreactor
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 and X rotational 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- Fig. 3.5. Fluid velocity magnitudes at Points 6–10 within scaffold
ties due to the bi-axial and Z rotational schemes (see subjected to rotation about respective axes.
H. Singh et al. / Journal of Biotechnology 119 (2005) 181–196 195

edges of the scaffolds tend to be approximately three in most cases (and not just ours) as a result of bi-
to four times more than the internal shear stresses as axial rotation.
indicated. This is noted to particularly occur at the inter- 3. These simulations assist in identifying critical
sections of scaffold struts/bars and at the circular edges issues and problems, for example, in approximat-
of the fibre end-faces. One reason for this is that stresses ing the locations of recirculation zones. These
tend to be concentrated at the edges, which are directly zones may potentially damage cells and inhibit
exposed to the dynamic fluid. growth. Furthermore, these recirculating bodies
The Reynolds number is a commonly applied non- may impede the flow of fluid into and out of the
dimensional parameter that is applied to assess the state scaffold. The choice of flow regime is therefore of
of a system, where great importance.

ρVd Ωr 2
Re = or Re = (3.1) References
µ υ

whereby V is the fluid velocity, d the diameter, r the Bancroft, G., Sikavitsas, V., Mikos, A., 2003. Design of a flow perfu-
radius and Ω is the rotational velocity. The Reynolds sion bioreactor system for bone tissue-engineering applications.
Tissue Eng. 9 (3), 549–554.
number represents a ratio of inertial to viscous effects Boyd, E.J., Gonda, S., 1990. Mathematical modelling of the flow
and indicates laminar, transient or turbulent flow. For field and particle motion in a rotating bioreactor at unit gravity
example, a flow scheme would tend towards the lam- and microgravity. In: NASA/ASEE Summer Faculty Fellowship
inar scheme due to a higher fluid viscosity, reducing Program, Johnson Space Centre, August 23.
the Reynolds number of a fluid. The Reynolds num- Bursac, B., Papadaki, M., White, J., Eisenberg, S., Vunjak-
Novakovic, G., Freed, L.E., 2003. Cultivation in rotating bioreac-
ber is of interest in many bioreactor systems as laminar tors promotes maintenance of cardiac myocyte electrophysiology
flow regimes are often employed. However, our sys- and molecular properties. Tissue Eng. 9 (6), 1243–1253.
tem is more complex because bi-axial rotation within Cherry, R.S., Papoutsakis, E.T., 1986. Hydrodynamic effects on cells
our asymmetric vessel is coupled with inlet and out- in agitated tissue culture reactors. Bioprocess Eng. 1, 29–41.
let flows. Hence, we did not attempt to calculate Re. Darling, E., Athanasiou, K., 2003. Articular cartilage bioreactors and
bioprocesses. Tissue Eng. 9 (1), 9–26.
The gravity and buoyancy effects are of less signifi- Freed, L.E., Vunjak-Novakovic, G., 2002. Spaceflight bioreactor
cance with respect to our bioreactor system, and were studies of cells and tissues. Adv. Space Biol. Med. 8, 177–195.
therefore neglected in our simulations. 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,
689–693.
4. Conclusions Hutmacher, D.W., Risbud, M., Sittinger, M., 2004. Evolution of
computer aided design and advanced manufacturing in scaffold
1. Three different flow configurations were discussed research. Trends Biotechnol. 22 (7), 354–362.
with respect to our prototype bioreactor design, pri- Kleis, S.J., Pellis, N.R., 1995. Fluid dynamic verification experi-
ments on STS-70. In: NASA/ASEE Summer Faculty Fellowship
marily rotation about the X (tumble), Z (spin) and Program, Johnson Space Centre, August 9.
XZ (bi-axial) axes. Kleis, S.J., Begley, C., Pellis, N.R., 1996. Bioreactor mass transport
2. Fluid velocities and shear stresses within the scaf- studies. In: NASA/ASEE Summer Faculty Fellowship Program,
folds were studied and compared, and proved that Johnson Space Centre, August 2.
bi-axial rotation results in significant improvements Lappa, M., 2005. A CFD level-set method for soft tissue growth:
theory and fundamental equations. J. Biomech. 38, 185–190.
in terms of fluid transport through the scaffolds. Levesque, M.J., Nerem, R.M., 1985. The elongation and orienta-
This is due to the combined effects of the rota- tion of cultured endothelial cells in response to shear stress. J.
tional velocity “vectors” about both axes, which Biomech. Eng. 107, 341–347.
when combined, would almost certainly result in a Levesque, M.J., Sprague, E.A., Schwartz, C.J., Nerem, R.M., 1989.
higher rotational velocity component, as compared The influence of shear stress on cultured vascular endothelial
cells: the stress response of an anchorage-dependent mammalian
to rotation about a single axis. However, a rise in cell. Biotechnol. Progr. 5, 1–8.
shear forces was also reported. Greater transport Martin, I., Wendt, D., Heberer, M., 2004. The role of bioreactors in
within scaffolds for example, would therefore result tissue engineering. Trends Biotechnol. 22 (2), 80–86.
196 H. Singh et al. / Journal of Biotechnology 119 (2005) 181–196

Olivier, A.L., Truskey, G.A., 1993. A numerical analysis of Sodian, R., Lemke, T., Fritsche, C., Hoerstrup, S., Fu, P., Potapov, E.,
forces exerted by laminar flow on spreading cells in a par- Hausmann, H., Hetzer, R., 2002. Tissue-engineering bioreactors:
allel plate flow chamber assay. Biotechnol. Bioeng. 42, 963– a new combined cell-seeding and perfusion system for vascular
973. tissue engineering. Tissue Eng. 8 (5), 863–870.
Papoutsakis, E.T., 1991. Fluid-mechanical damage of animal cells in Stathopoulos, N.A., Hellums, J.D., 1985. Shear stress effects on
bioreactors. Trends Biotechnol. 9, 427–437. human embryonic kidney cells in vitro. Biotechnol. Bioeng. 27,
Porter, B., Zauel, R., Stockman, H., Guldberg, R., Fyhrie, D., 2005. 1021–1028.
3-D computational modeling of media flow through scaffolds in Sucosky, P., Osorio, D.F., Brown, J.B., Neitzel, G.P., 2003. Fluid
a perfusion bioreactor. J. Biomech. 38, 543–549. mechanics of a spinner-flask bioreactor. Biotechnol. Bioeng. 85
Pritchard, W.F., Davies, P.F., Derafshi, Z., Polacek, D.C., Tsao, R., (1), 34–46.
Dull, R.O., Jones, S.A., Giddens, D.P., 1995. Effects of wall shear Vunjak-Novakovic, G., Freed, L.E., Biron, R.J., Langer, R., 1996.
stress and fluid recirculation on the localization of circulating Effects on mixing on tissue engineered cartilage. AIChE J. 42,
monocytes in a three-dimensional flow model. J. Biomech. 28, 850–860.
1459–1469. Vunjak-Novakovic, G., Obradovic, B., Bursac, P., Martin, I., Langer,
Raimondi, M.T., Boschetti, F., Falcone, L., Migliavacca, F., Remuzzi, R., Freed, L.E., 1998. Dynamic seeding of polymer scaffolds for
A., Dubini, G., 2004. The effect of media perfusion on three- cartilage tissue engineering. Biotechnol. Progr. 14, 193–202.
dimensional cultures of human chondrocytes: integration of Yamada, H., Takemasa, T., Yamaguchi, T., 2000. Theoretical study
experimental and computational approaches. Biorheology 41, of intracellular stress fiber orientation under cyclic deformation.
401–410. J. Biomech. 33, 1501–1505.
Redaelli, A., Boschetti, f., Inzoli, F., 1997. The assignment of velocity Zein, I., Hutmacher, D.W., Teoh, S.H., Tan, K.C., 2002. Poly(e-
profiles in finite element solutions of pulsatile flow in arteries. caprolactone) scaffolds designed and fabricated by fused depo-
Comput. Biol. Med. 27 (3), 233–247. sition modeling. Biomaterials 23 (4), 1169–1185.

Anda mungkin juga menyukai