You are on page 1of 14

Journal of Membrane Science 303 (2007) 86–99

Momentum transfer inside a vertically orientated


capillary membrane bioreactor
B. Godongwana 1 , M.S. Sheldon ∗ , D.M. Solomons 2
Department of Chemical Engineering, Cape Peninsula University of Technology, P.O. Box 652,
Cape Town 8000, South Africa
Received 20 April 2007; received in revised form 22 June 2007; accepted 27 June 2007
Available online 10 July 2007

Abstract
Innovation in biotechnology research has resulted in a number of fungi being identified for diverse industrial applications. Much research
has been done in developing optimised membrane bioreactor (MBR) systems for the cultivation of these fungi as a consequence of their potent
industrial applications. This research has been hampered by the lack of a thorough understanding of the fluid mechanics through these devices.
In this article, analytical and numerical solutions to the Navier–Stokes equations were developed to describe the hydrostatic pressure and velocity
profiles in a single fiber membrane gradostat reactor (SFMGR). A generic equation for low wall Reynolds number (Rew = ρvw rH /μ) flows was
developed and solved for the case of negligible angular variations of the flow profiles. The mathematical expressions were proposed as solutions
to transient state, laminar, incompressible, viscous and isothermal flow through a membrane with a variable hydraulic permeability. These profiles
were developed for the lumen and shell sides, taking into account the osmotic pressure and gel formation that occurs when solute particles are
rejected on the membrane. The models developed are applicable to different orientations and configurations. A numerical scheme, with a complete
stability analysis, was developed to complement the analytical models. The models were tested on a vertically orientated MGR, operated in the
dead-end mode. The model solutions gave profiles that are in agreement with the experimental results.
© 2007 Elsevier B.V. All rights reserved.

Keywords: Capillary membrane bioreactor; Momentum transfer; Finite difference; Numerical technique

1. Introduction are inherent of hollow fiber (HF) and capillary MBR’s, for
the continuous production of enzymes from Phanerochaete
Innovation in biotechnology research has resulted in a num- chrysosporium. The MGR showed higher enzyme activities
ber of fungi being identified for diverse industrial applications. than previous and subsequent conventional bioreactor systems
Much research has been done in developing optimised mem- [1,2,4,5]. Following from the work done by Leukes [3], a number
brane bioreactor (MBR) systems for the cultivation of these of other investigations [6–9] have demonstrated the suitability
fungi as a consequence of their potent industrial applications and viability of a capillary polysulphone (PSu) MGR for lignin
[1–5]. Leukes [3] developed a membrane gradostat reactor peroxidase (LiP) and manganese peroxidase (MnP) production.
(MGR), which uses to its advantage the nutrient gradients that The performance of HF and capillary membranes is determined
in large by the transport rate of the key nutrients and/or wastes
[10,11]. It is therefore crucial to have a complete description
Abbreviations: BC, boundary condition; HF, hollow fiber; IC, initial con- and understanding of the momentum as well as the mass transfer
dition; LiP, lignin preoxidase; MnP, manganese preoxidase; P. chrysosporium,
Phanerochaete chrysosporium; SFMGR, single fiber membrane gradostat reac-
through these devices for an optimum MBR design.
tor; SSC, steady-state condition; WRF, white rot fungi A number of experimental and theoretical investigations have
∗ Corresponding author. Tel.: +27 21 460 3160; fax: +27 21 460 3282. been conducted with the aim of modelling the momentum trans-
E-mail addresses: sheldonm@cput.ac.za (M.S. Sheldon), fer inside HF and capillary membranes [10–18]. Most of these
Deon.Solomons@uct.ac.za (D.M. Solomons). investigations however, were unsuccessful in providing adequate
1 Tel.: +27 21 460 3160; fax: +27 21 460 3282.
2 Address: Department of Mathematics and Applied Mathematics, University mathematical models for pressure and velocity profiles, which
of Cape Town, Private Bag, Rondebosch 7700, South Africa. also account for the different modes of operations and/or orien-
Tel.: +27 21 460 3160; fax: +27 21 460 3282. tations of MBR’s. Out of the theoretical investigations [10–18],

0376-7388/$ – see front matter © 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.memsci.2007.06.070
B. Godongwana et al. / Journal of Membrane Science 303 (2007) 86–99 87

Kelsey et al. [10] gave more detailed mathematical models for


velocity and pressure profiles of nutrients flowing through a
HF MBR, operated at steady state. However, the models devel-
oped by Kelsey et al. [10] ignored osmotic pressure effects
that result from the rejection of solutes as well as gel layer
formation on the surface of the membrane lumen. These mod-
els were also restricted to a horizontally orientated bioreactor.
The Navier–Stokes equations are regarded as the fundamental
equations governing fluid motion. Although there are no analyt-
ical solutions for the complete Navier–Stokes equations, these
equations can still be solved numerically to attain a more descrip-
tive analysis than a simplified analytical solution. A number of
numerical algorithms have been proposed as solutions to the
problem of flow through HF and capillary membranes. Damak
et al. [16] used an implicit finite difference method to numeri-
cally develop pressure and velocity fields of crossflow filtration.
This model however does not take into account the different
modes of operation of membrane bioreactors and also ignores
Fig. 1. A schematic diagram of the vertical single fiber membrane gradostat
osmotic pressure effects. Currently there are no mathematical reactor system.
models to describe the momentum transfer through a vertically
Table 1
orientated capillary MGR, which also take into account osmotic
Model parameter values used to test the developed model
pressure effects, gel layer formation, as well as the mode of
operation and/or orientation of the bioreactor. This study was Model parameter Symbol Unit Basic value
aimed at developing these mathematical models for the purpose Initial hydraulic permeability km m2 2.18 × 10−17
of scaling-up the MGR system for continuous LiP and MnP pro- Fraction retentate f 0
duction. The transient-state models developed are applicable to Membrane inner radius RL m 6.63 × 10−04
Membrane outer radius Rx 1.37
horizontal and vertical orientations as well as dead-end, contin-
Distance from the centre of R2 8.92
uous open shell, closed shell, and shell side cross flow modes of the membrane to the centre
operation. The current study was restricted to membranes with of the extra capillary space
no growth on the external surface and future work will focus on of the glass manifold
modifying these models to account for the growth kinetics of Glass manifold inner radius R3 16.47
Glass bioreactor length Lb m 0.23
the P. chrysosporium.
Effective membrane length L m 0.16
Inlet hydrostatic pressure p0 Pa 117.01 × 103
2. Materials and methods Outlet hydrostatic pressure p1 Pa 113.57 × 103
Nutrient flowrate Q m3 /s 1.72 × 10−09
The vertical single fiber membrane gradostat reactor Nutrient medium viscosity μ Pa s 1.18 × 10−3
Nutrient medium density ρ kg/m3 998.00
(SFMGR) system that was used to test the models is the same
as that previously used and explained by Ntwampe and Shel-
don [19]. It consisted of a single capillary membrane, made of
surface modified polysulphone, encased in a glass bioreactor.
A schematic diagram of the SFMGR is shown in Fig. 1 and
the dimensions are given in Table 1. The polysulphone capillary
membranes were produced and supplied by the Institute of Poly-
mer Science at the University of Stellenbosch (South Africa).
These membranes were developed by Jacobs and Leukes [20]
and offer many advantages when used in a MBR, such as low
resistance to liquid transport and a large external surface area.
They are characterized by an internally skinned and externally
unskinned region of microvoids, approximately 0.15 mm long
and 0.015 mm thick, as shown in Fig. 2. The separation charac-
teristics of these membranes are summarized in Table 2.
A nutrient medium, described by Tien and Kirk [21], was sup-
plied to the membrane lumen at a flowrate of 1.72 × 10−9 m3 /s
using a Watson Marlow 505S peristaltic pump (Dune Engineer-
ing, RSA). The SFMGR was fitted with two pressure transducers
(Rhomberg Winter, South Africa) at the inlet and outlet of the Fig. 2. A longitudinal SEM of an internally skinned and externally unskinned
MGR as shown in Fig. 1. LabView, a data acquisition programme capillary membrane [7].
88 B. Godongwana et al. / Journal of Membrane Science 303 (2007) 86–99

Table 2
Characteristics of the internally skinned and externally unskinned polysulphone
capillary membrane
Outer diameter (m) 0.0018
Inner diameter (m) 0.0013
Wall thickness (m) ±200 × 10−6
Average pore size (␮m) 11
Burst pressure (kPa) 1400
Operating pH 4–11
Maximum temperature (◦ C) 50

developed by National Instruments, was utilised for acquiring,


transforming and displaying the online data from the pressure
transducers.
The models presented in this study will be applied to the
SFMGR operated in the dead-end mode with a constant shell
side pressure. The models will be used to predict the pressure
along the length of the membrane to the outlet. The predicted
outlet pressures will be compared to measured outlet pressures
as well as literature models predictions. The accuracy of the
developed models will be tested under varying conditions of
operation, including variations in membrane length and nutrient
flow rates.

3. Analysis

3.1. Model assumptions

The model solutions were based on the following imposed


operating conditions of the SFMGR: (1) the system is isother-
mal; (2) the flow regime within the fiber lumen is fully
developed, laminar, and homogenous; (3) physical and transport
parameters (e.g. density, viscosity and diffusivity) are constant;
(4) the nutrient flows in axial and in radial directions in the
lumen, whereas in the dense and in the spongy matrix of the
membrane the flux is only in the radial direction; (5) the momen- Fig. 3. Cross-section of a membrane showing the hypothesised profiles and
notation of the membrane regions.
tum transfer can be solved independently of the mass transfer.
The dimensionless parameter for indicating whether or not the  
mass transfer can be decoupled from the momentum transfer ∂vr ∂vr vθ ∂vr v2θ ∂vr
analysis is the Schmidt number (Sc = μ/ρDAB ). Sc larger than 1 ρ + vr + − + vz
∂t ∂r r ∂θ r ∂z
indicate that the concentration boundary layer is much smaller
than the momentum boundary layer [15]. The Sc for the nutrient    
1 ∂ ∂vr vr 1 ∂ 2 vr 2 ∂vθ ∂ 2 vr
solution used in this study was much greater than 1 (Sc = 744.5). =μ r − 2+ 2 2 − 2 + 2
r ∂r ∂r r r ∂θ r ∂θ ∂z
∂p
3.2. The generic equation − + ρgr (2)
∂r
The notation used and the classification of the membrane Eqs. (1) and (2) may be reduced by considering the fact that for
regions is shown in Fig. 3. The starting point for this analysis a small wall Reynolds number (Rew = ρvw rH /μ) the inertial
is the z and r-components of the non-conservative form of the terms will be negligible [10,14,18]. In most viscous flows normal
Navier–Stokes equations in cylindrical coordinates: stress effects, ∂2 vz /∂z2 , are not as important as shear stresses
  [22] and are negligible when the aspect ratio, RL /L, is less than
∂vz ∂vz vθ ∂vz ∂vz 10−2 , a condition that is satisfied in almost all HF membrane
ρ + vr + + vz
∂t ∂r r ∂θ ∂z devices [10]. Eq. (1) then simplifies to:
        
1 ∂ ∂vz 1 ∂ 2 vz ∂ 2 vz ∂p ∂vz 1 ∂ ∂vz ∂p
=μ r + 2 2 + 2 − +ρgz (1) ρ =μ r − + ρgz (3)
r ∂r ∂r r ∂θ ∂z ∂z ∂t r ∂r ∂r ∂z
B. Godongwana et al. / Journal of Membrane Science 303 (2007) 86–99 89

Eq. (3) may be written in dimensionless form by introducing the


following dimensionless groups:
vzz vzz
UL = = (4)
vz0 (p0 − p1 )R2L /4μL
where UL represents the dimensionless axial velocity in the
lumen of the membrane, p0 and p1 are the inlet and outlet hydro-
static pressures, respectively, vzz and vz0 are the axial velocities
in the lumen at a point z and at the entrance, respectively, RL the
membrane inner radius, L the length of the membrane and μ is
the fluid dynamic viscosity.
vrz vrz
VL = = (5)
vz0 (p0 − p1 )R2L /4μL
where VL is the dimensionless radial velocity in the membrane
lumen and vrz is the radial velocity in the lumen at a point z.
The velocity components, UL and VL , are expressed as ratios Fig. 4. An SEM of gel layer deposit on the surface of the capillary membrane
of the maximum axial velocity at the membrane entrance. The after ±120 h of operation of the dead-ended SFMGR.
maximum axial velocity at the membrane entrance is expressed
by Poiseuille’s equation: is visible on an SEM after ±120 h of operation for the specific
  conditions of operation given in Table 1. Since the permeability
R2 p0 − p1
vz0 = L (6) of any gel layer, kg , formed on the membrane surface is a strong
4μ L function of the diameter dg of the particles, the Carman–Kozeny
The dimensionless radial and axial velocities in the shell, VS equation is one of the most popular equations for particulate
and US , respectively, are also expressed as ratios of the maxi- filtration [24]:
mum axial velocity at the membrane entrance. In Eqs. (4) and
(5), therefore, the subscripts zz and rz should be replaced by zs dg2 ε3
kg = (13)
and rs, respectively, when referring to shell flows. The dimen- 180(1 − ε)2
sionless hydrostatic luminal pressure, PL , and the dimensionless
where ε is the porosity of the gel layer. Estimation of the actual
osmotic pressure in the lumen, Π, are defined as ratios of the
value of the gel layer permeability kg needs accounting for mass
total pressure drop along the membrane lumen:
transport towards the membrane. The definition of the dimen-
4pL sionless luminal hydrostatic pressure, PL , can also be extended
PL = (7)
p0 − p 1 to the dimensionless shell hydrostatic pressure, PS , as a ratio of
the total pressure drop along the membrane lumen. In terms of

Π= (8) the dimensionless variables, Eq. (3) becomes:
p0 − p 1
   
z ∂UL 1 ∂ ∂UL ∂PL
Z= (9) = R − −b (14)
L ∂τ R ∂R ∂R ∂Z
r
R= (10) where
RL
4ρgz L
μt b= (15)
τ= (11) p0 − p 1
ρR2L
To solve for Eq. (14) we make use of separable differential equa-
km kg L
κ= 2 (12) tions. This technique works by guessing the form of the solution
RL (dw + dg ) of the partial differential equation which turns the equation into
where Z is a dimensionless axial coordinate, R a dimensionless ordinary differential equations, that are easier to treat. If the
radial coordinate, τ a dimensionless time, κ a dimensionless solution of Eq. (14) is assumed to be of the form:
hydraulic permeability, km the membrane hydraulic permeabil-
UL (R, τ) = Ξ(R)K(τ) (16)
ity, kg the permeability of the gel layer, dw the membrane wall
thickness and ρ is the fluid density. The gel layer deposit on as a product of a radial function Ξ(R) and a function of time
the surface of the membrane is a function of both flux, which K(τ). The following assumption is made regarding the function
increases the mass rate of material retained at the membrane, K:
and cross-flow velocity, which reduces polarization by enhanc-
ing feed-side mass transfer [23]. An SEM of the gel layer deposit ∂K
= −α2 K (17)
in a SFMGR is shown in Fig. 4. The gel layer shown on this figure ∂τ
90 B. Godongwana et al. / Journal of Membrane Science 303 (2007) 86–99

in terms of a non-zero separation constant α. The basis of this Table 4


assumption is apparent when considering its solution, which is The boundary, steady state, and initial conditions of the SFMGR
IC/BC/SSC τ, R, Z UL , US , VL , VS , PL , PS Range
K = B0 e−α

(18)
IC τ =0 U = Uτ = U∞ = 0 (22a) 0 ≤ R ≤ 1, 0 ≤ Z ≤ 1
∂U
for a constant of integration B0 . This form of the function K(τ) SSC τ =∞ =0 (22b) 0 ≤ R ≤ 1, 0 ≤ Z ≤ 1
∂τ
is appropriate for the model under consideration, since it is to BC1 R=1 U=0 (22c) 0≤Z≤1
be expected that the velocity profile decays with time. Eq. (14) ∂U
BC2 R=0 =0 (22d) 0≤Z≤1
may then be written as ∂R
BC3 R=0 V =0 (22e) 0≤Z≤1
BC4 R=1 V = VM (22f) ␶=∞
d2 Ξ dΞ dPL
R2 +R + (α2 R2 )Ξ = f (Z, τ) (19) BC5 Z=0 PL =PL (0), =a (22g) 0≤R≤1
dR2 dR dZ
BC6 R=0 U = finite (22h) 0≤Z≤1
with the inclusion of an axial-temporal function: BC7 Z=0 US = 0, PS = PS(0) (22i) 0≤R≤1
∂US
2 BC8 R = R2 =0 (22j) 0≤Z≤1
eα τ R2 G ∂R
US = 0 0≤Z≤1
f (Z, τ) = , B0 = 0 (20) BC9 R = R3 (22k)
B0 BC10 R = R2 VS = 0 (22l) 0≤Z≤1
BC11 R = Rx VS = 0 (22m) 0≤Z≤1
Here-in, the factor G is given by
 
∂PL
G= −b (21) or
∂Z  
1 dPL
Eq. (19) may be thought of as a generic equation for small values U∞ = − (1 − R2 ) −b (26)
4 dZ
of the wall Reynolds number (Rew  1) flows through cylindri-
cal surfaces, because under the conditions specified in Table 3 The ‘no slip’ assumption of BC1 is based on the fact that the
it reduces to the familiar Bessel’s equation, Laplace’s equation parameter χL is 0.02 for the SFMGR system used. According to
or the more general Poisson’s equation, all of which have been Catapano et al. [11], the distortion of the velocity profile relative
extensively studied, particularly in literature on flow behaviour. to that established in non-porous wall systems is measured by
χL, where
3.3. Vertical orientation (constant shell side pressure)  
μkm 1/2
χ=4 (27)
R3L
The solutions presented in this section are only applicable
to a system with a constant shell side hydrostatic pressure. Eq. In Eq. (27) χ is in m−1 and km is in m/Pa s, and if the value of χL
(19) is solved by making use of the boundary (BC), steady-state is less than unity, the distortion of the velocity profile induced
(SSC) and initial conditions (IC) specified in Eqs. (22a)–(22m) by the porous wall is negligible. The steady-state function U∞
as listed in Table 4, and has a solution of the form: is zero before the time τ ∞ and non-zero at time τ ∞ :
Ξ = Ξ∞ − Ξτ or U = U∞ − Uτ (23) α2 Ξ∞ = 0 for τ0 ≤ τ ≤ τ∞ (28)
where U∞ and Uτ are the steady-state and transient axial velocity therefore, substituting Eqs. (25) and (28) back into Eq. (24), and
distribution, respectively, Ξ ∞ and Ξ τ are functions of U∞ and recalling the definition of K given in Eq. (18) results in
Uτ , respectively. Eq. (19) may therefore be written as  
  1 d dΞτ G α2 τ∞ 2
1 d d(Ξ∞ − Ξτ ) G R + α 2 Ξτ = (e − eα τ ) (29)
R + α2 (Ξ∞ − Ξτ ) = (24) R dR dR B0
R dR dR K
The transient velocity function Ξ τ is finite for τ < τ ∞ , but as
The steady-state axial velocity profile, U∞ , is obtained by apply- τ ∞ approaches infinity the function Ξ τ is undefined in Eq. (29),
ing the SSC, BC1 and BC2 in Table 4 to Eq. (19) to obtain: therefore G/B0 has to be zero in this equation. This condition is
only met if B0 is very large. Eq. (29) then simplifies to
1 G  
Ξ∞ = − (1 − R2 ) (25) 1 d dΞτ
4 K∞ R + α 2 Ξτ = 0 (30)
R dR dR
Table 3 Eq. (30) is a special case of Bessel’s equation and therefore
Generic equation of low wall Reynolds number flows through cylindrical
surfaces
should have a solution of the form:

α f(Z, τ) Resulting equation Ξτ = B1 J0 (αR) + B2 Y0 (αR) (31)


>0 =0 Bessel’s equation where B1 and B2 are constants and J0 (αR) and Y0 (αR) are Bessel
=0 >0 Poisson’s equation
=0 =0 Laplace’s equation
functions of zero order of the first kind and second kind respec-
tively. Because the function Y0 (αR) tends to minus infinity as R
B. Godongwana et al. / Journal of Membrane Science 303 (2007) 86–99 91

goes to zero, B2 in Eq. (31) has to be zero. Imposing the ‘no slip The dimensionless radial velocity is obtained by integrating Eq.
condition’ in Eq. (22c), the axial lumenal velocity, UL , should (39) and making use of BC3 to obtain:
be equal to zero at R = 1, Ξ must therefore also vanish at R = 1.   ∞

Since B1 cannot be set equal to zero without obtaining the trivial 1 R R2  J1 (αn R) −α2n τ
VL (Z, R, τ) = 1− −8 e
solution τ = 0, the Bessel function J0 (αR) must be set equal to 4 2 2 α4n J1 (αn )
n=1
0 in Eq. (31):
d2 PL
∴ J0 (αR) = 0 (32) × (40)
dZ2
But Eq. (32) has an infinite number of roots α (often called The velocity through the membrane matrix, VM , is derived from
eigenvalues) that will satisfy the boundary conditions of Eq. Darcy’s law, taking into account osmotic pressure, gel formation
(31). These roots are infinite and follow the sequence α1 = 2.405, and gravity, and is given by
α2 = 5.520, α3 = 8.654, . . ., αn . Hence, there are many solutions,
Ξ ␶ = B1n J0 (αn R) with n = 1, 2, 3, . . ., ∞, which will satisfy Eq. VM = −κ[PSbπ − PL ] (41)
(30) and the corresponding boundary conditions. Substituting
Eqs. (18) and (31) into Eq. (16), and considering the above where
 
arguments results in the following expression for Uτ : dw
PSbπ = PS + Π − b (42)

 L
−α2n τ
Uτ (R, τ) = Bn e J0 (αn R) (33)
where Π is the dimensionless osmotic pressure on the lumen
n=1
side of the membrane, and the term PSbπ is a function of PS , Π,
where and b. For most work, the van’t Hoff approximation for osmotic
pressure gives an adequate estimate [23]:
Bn = B0 B1n (34)
Φ = cw R∗ T (43)
The initial condition, Eq. (22a) stipulates that at τ = 0, Uτ = U∞ ,
therefore where cw is the total concentration of ions at the wall–membrane
   ∞ interface on the feed side in kmol/m3 , T the absolute tempera-
1 dPL
− (1 − R2 ) −b = Bn J0 (αn R) (35) ture of the solution in K, and R* is the universal gas constant
4 dZ 8.31451 J/(g mol) K. The determination of the wall concentra-
n=1
tion requires an estimation of the local mass transfer coefficient
Eq. (35), the Fourier–Bessel series, may be solved by making use
between the bulk feed stream and the membrane surface, which
of some standard properties of Bessel functions (e.g. Lommel
in turn requires an account of the mass transfer characteristics
integrals) to obtain the function Bn :
(e.g. the solute diffusion coefficient) of the membrane bioreac-
 
2 dPL tor. The dimensionless osmotic pressure, Π, inside the lumen
Bn = − 3 −b (36) is therefore obtained by substituting Eq. (43) into Eq. (9). The
αn J1 (αn ) dZ
luminal pressure profile, PL , can be obtained by combining Eqs.
Substitution of the new expression for Bn in Eq. (33) results in (40) and (41) and making use of BC4, and this substitution results

  in
 J0 (αn R) dPL
−α2n τ
Uτ = −2 e − b (37) d2 PL
α3n J1 (αn ) dZ − 16κPL = −16κPSbπ (44)
n=1
dZ2
where an is the eigenvalues of Eq. (30) as previously explained. Eq. (44) is a simple ordinary differential equation (ODE) that
The transient axial velocity distribution in the membrane lumen can be resolved using Laplace transforms with BC5, and has a
is therefore given by substituting Eqs. (26) and (37) into Eq. solution of the form:
(23):



PL (Z) = (P0 − PSbπ ) cosh(4 κ)Z
1  J0 (αn R) −α2n τ
UL (Z, R, τ) = − (1 − R ) − 82
e a √
4 α3 J (α ) + √ sinh(4 κ)Z + PSbπ (45)
n=1 n 1 n 4 κ
 
dPL where
× −b (38)
dZ √ √
4 κ(P0 − PSbπ ) sinh(4 κ) − b(1 + f )
The corresponding luminal radial velocity profile is obtained by a= √ (46)
f − cosh(4 κ)
making use of the continuity equation, which in its dimension-
less form for an incompressible fluid is given by The parameter f was first defined by Bruining [14] as

1 ∂(RVL ) ∂UL UL (1)


=− (39) f = (47)
R ∂R ∂Z UL (0)
92 B. Godongwana et al. / Journal of Membrane Science 303 (2007) 86–99

For a constant shell pressure, as was the case in the MGR, there Table 5
is no radial shell flow (i.e., VS = 0), and the axial shell velocity Discretization of the dimensionless Navier–Stokes equations
profile, US , is given by Differential Discretization Type
  ∞
n+1
Ui,j −Ui,j
n

b R 4 −α2n τ
∂U
∂τ n τ−U n
Forward difference
US = 2
2R2 ln + R3 − R −
2 2
Bn e J0 (αn R) ∂U
Ui,j+1 i,j−1
4 R3 b Central difference
n=1
∂R n 2R
Ui+1,j −Ui−1,j
n
∂U
(48) ∂Z n 2Z
Central difference
Ui+1,j −2Ui,j
n +U n
∂2 U i−1,j
∂Z2 ΔZn2 n
Central difference
where ∂2 V
n
Vi,j+1 −2Vi,j +Vi,j−1
Central difference
 
∂R2 n
Pi+1,j
R
−Pi,j
n
2

2 1bR22 1 ∂P
∂Z n Z
Forward difference
Bn = − ln R3 + J0 (αn ) ∂P
Pi,j+1 −Pi,j
n

J12 (αn ) 2 2 4 ∂R R Forward difference


     
bR22 1 1
+ 2 αn − J1 (αn ) + ln R3 + J2 (αn )
4αn 2Γ (2) 2 The solutions for the generic model for a vertically orientated
membrane bioreactor with a variable shell side pressure is pro-
b
− 3 J1 (αn ) (49) vided in Appendix A.
αn

In Eqs. (48) and (49), R1 , R2 and R3 are the membrane exit 3.4. Numerical scheme development
radius, the extra capillary space radius and the glass manifold
inner radius, respectively. The volumetric flow rate on the lumen The numerical solutions of the two-dimensional
of the membrane is obtained from the equation: Navier–Stokes equations were obtained by finite differ-
 1 encing the dimensionless partial differential equations as shown
ΩL = 4 UL R dR (50) in Table 5. These solutions are restricted to luminal flows only.
0 The assumptions used in the development of the analytical
Substituting Eq. (38) in Eq. (50) and integrating between the solutions are still maintained, however in the numerical scheme
limits 0 and 1 results in the following expression for the luminal the assumptions of negligible inertial and normal stress effects
volumetric flowrate: have been lifted. The extension of the numerical scheme to
 ∞ −α2 τ
 
account for these effects makes it a more accurate approach
1  e n dPL vis-à-vis the analytical, however, the convergence of this
ΩL = − −8 − b (51) scheme is also influenced by a number of factor as will be
4 α4n dZ
n=1
discussed. The dimensionless form of Eq. (1) is
A stream function is defined such that  
∂U ∂U ∂U
1 ∂ψ + Reb V + βU
U=− (52) ∂τ ∂R ∂Z
R ∂R  
1 ∂ ∂U ∂2 U ∂P
and, = R + β2 2 − β +b (56)
R ∂R ∂R ∂Z ∂Z
1 ∂ψ
V = (53) and the dimensionless form of Eq. (2) is
R ∂Z
 
∂V ∂V ∂V
From Eqs. (52) and (53) the stream functions in the lumen and + Reb V + βU
shell of the membrane are respectively given by ∂τ ∂R ∂Z
    
1 ∂ ∂V V ∂2 V ∂P
ψL =
1 1 dPL
−b = R − 2 + β2 2 − (57)
4 2 dZ R ∂R ∂R R ∂Z ∂R
 2   
R R2 dPL(1) dPL(0) where Reb is the bulk fluid Reynolds number, which is different
+ 1− − (54) from the wall Reynolds number (Rew ) that is calculated using
2 2 dZ dZ
the wall velocity. Reb is defined as
and
ρvz0 RL
Reb = (58)
R22 μ
ψS = −b [ln(R3 − R2 − ln R3 )](R23 − R22 )
4
and β is the aspect ratio defined as

R22 R23 R43 − R42
− (R23 − R22 ) + (R23 − R22 ) − (55) RL
8 8 16 β= (59)
L
B. Godongwana et al. / Journal of Membrane Science 303 (2007) 86–99 93

The corresponding axial and radial finite difference equations


are, respectively:
 
n+1
Ui,j −Ui,j
n U n −Ui,j−1
n i,j+1
n n
Ui+1,j −Ui−1,j
n
+Reb Vi,j +βUi,j
n
τ 2R 2Z
1 Ui,j+1 − Ui,j−1 − 2Ui,j
n n n
Ui,j+1 n + Un
i,j−1
= + 2
R 2R R
U n
i+1,j −2U n + Un
i,j i−1,j
n
Pi+1,j −Pi,j
n
+β2 −β +b (60)
Z2 Z
and,
 
n+1
Vi,j −Vi,j
n n
Vi,j+1 −Vi,j−1
n n
Vi+1,j −Vi−1,j
n
+Reb n
Vi,j +βUi,j
n Fig. 5. Dimensionless transient axial velocity profiles in the membrane lumen
τ 2R 2Z for the vertically orientated SFMGR, operated in the dead-end mode with con-
stant shell side pressure.
1 Vi,j+1 − Vi,j−1 − 2Vi,j
n n n
Vi,j+1 n + Vn n
i,j−1 Vi,j
= + −
R 2R R2 R2 velocity profiles from the developed analytical models, Eqs. (38)
Vi+1,j − 2Vi,j + Vi−1,j
n n n Pi,j+1 − Pi,j
n n
and (40), respectively, are plotted in Figs. 5 and 6. As can be seen
+β2 − (61)
Z 2 R from both Figs. 5 and 6, in the limiting case τ → ∞, the velocity
Eqs. (60) and (61) were solved using a revised form of the semi- profiles approach steady state. At steady state there is no change
implicit method for pressure-linked equations (SIMPLE). The in the velocity profiles with a change in time. For the model
procedure for the SIMPLE algorithm is as follows [22]: parameter values listed in Table 1 this corresponds to a τ = 1000
(7.5 min). The dimensionless radial velocity VL increases to a
maximum value of 1.0 × 10−6 , whereas UL decreases to zero
1. Start the iterative process by guessing the pressure field.
with an increasing R. This result is consistent with the ‘no slip’
Denote the guessed pressures as (P*)n .
assumption of BC1. A visualisation of this flow field is provided
2. Use the values of (P*)n to solve for Un and Vn from the
in Fig. 7 by means of streamlines. The method of producing a
momentum equations. Since these values are those associated
streamline from Eq. (54) is to set Ψ L equal to an arbitrarily cho-
with the values of (P*)n , denote them by (U*)n and (V*)n .
sen constant and plotting the R versus Z curve. Other streamlines
3. Since the values of (U*)n and (V*)n were obtained from
are obtained by setting Ψ L equal to various other constants. To
guessed values of (P*)n , they will not necessarily satisfy the
obtain R for a given Z value in Eq. (54) is an iterative process.
continuity equation. Hence, using the continuity equation,
Unfortunately, the solver algorithm that was used (Microsoft
construct a pressure correction (P )n which when added to
solver) does not converge for the range Z = 0–1. However, the
(P*)n will bring the velocity field more into agreement with
shape of the streamlines can still be extracted by recognising that
the continuity equation. That is, the corrected pressure Pn is
the function of the streamline is a polynomial of the 4th degree
given by
in R. These lines show the pattern of flow of ‘fluid elements’ as
P n = (P ∗ )n + (P )
n
(62) they move along the membrane at different radial positions in
the dead-end mode.
4. Designate the new value of Pn in Eq. (62) as the new value of When the model parameter values in Table 1 are fitted to Eq.
(P*)n . Return to step 2 and repeat the process until a velocity (45), for different hydraulic permeabilities, it can be deduced that
field which satisfies the continuity equation is obtained.

In this study, the SIMPLE algorithm was simplified by using


the expressions obtained from the analytical solution to develop
the pressure field in step 1. The pressure profile was assumed
to exhibit an exponential decay with time to some steady-state
value. The velocity profiles obtained from this pressure field
very closely satisfies the continuity equation.

4. Results

4.1. Analytical model solutions

As indicated in Section 2, these models were applied to a


vertically orientated MGR operated in the dead-end mode with Fig. 6. Dimensionless transient radial velocity profiles in the membrane lumen
a constant shell side pressure. Using the parameters specified in for the vertically orientated SFMGR, operated in the dead-end mode with con-
Table 1, the resulting transient-state axial, UL , and radial, VL , stant shell side pressure.
94 B. Godongwana et al. / Journal of Membrane Science 303 (2007) 86–99

small hydraulic permeabilities of the membrane. The models


presented in Section 3.3 of this paper give identical results to the
models proposed by Kelsey et al. [10] when ignoring gravitation
effects, and osmotic effects. It may also be noted from Eq. (41)
that the velocity through the matrix of the membrane is a function
of the osmotic pressure. As the osmotic pressure increases in
a SFMGR, with increasing concentration on the surface of the
membrane, the permeation velocity VM decreases. Recalling that
the performance of a capillary membrane is determined in large
by the transport rate of the key nutrients to the microbe on the
surface of the membrane [10], this information is crucial for
optimising the efficiency of the SFMGR. This result highlights
Fig. 7. Streamlines in the upper half of the capillary membrane operated in the the error of ignoring osmotic effects in previous publications
dead-end mode with constant shell side pressure (ψ represents stream functions
as defined by Eq. (54)).
aimed at modelling flow through these devices.
A comparison of Eq. (45) with two of the most widely used
models for predicting pressure profiles in HF and capillary mem-
the pressure drop along the membrane decreases with decreasing
brane devices is shown in Fig. 9. The model parameter values
membrane hydraulic permeability. By way of illustration, when
used for this comparison are listed in Table 1. Fig. 9 clearly shows
the dimensionless permeability κ decreases from 5.82 × 10−1 to
the divergence of the Bruining [14] and Kelsey et al. [10] mod-
5.82 × 10−3 , as would be the case with increasing growth of the
els from the experimental data. The developed model gives an
fungus, the dimensionless pressure drop, PL , also decreases
average percentage error of 0.5%, whereas the Bruining [14] and
from 17.5 to 4. In Fig. 8, PL is the difference between PL at
Kelsey et al. [10] models give percentage errors of 1.5% and 3%,
Z = 0 and at Z = 1. In the limiting case κ → 0, the pressure profile
respectively for a MBR with an effective membrane length of
will resemble that of a straight circular tube of constant cross
16 cm. The percentage error was calculated as: (the experimental
section (Poiseuille flow). This result is also expected, since the
value − theoretical value)/(experimental value). This discrep-
pressure drop in a vertically orientated SFMGR is merely due to
ancy is amplified with increasing membrane length, as shown
permeation of the luminal fluid and gravity. Another important
on Fig. 9. The difference between the model developed in this
feature of Fig. 8 is that the pressure drop is only linear for very
study, Eq. (45), and the Bruining [14] and Kelsey et al. [10]
models is that this model accounts for osmotic pressure and

Fig. 8. Dimensionless luminal hydrostatic pressure profile as a function of the Fig. 9. A comparison of the developed model with literature models and exper-
dimensionless membrane hydraulic permeability (κ). imental data.

Table 6
A comparison of the developed analytical pressure model against experimental pressure data for different nutrient flowrates and membrane lengths
Membrane length (m) Flowrate (ml/min) Experimental inlet pressure (Pa) Experimental outlet pressure (Pa) Developed model outlet pressure (Pa)

0.233 79.2 124151.6 120688.0 121312.4


0.233 158.0 129110.5 124388.0 126271.3
0.233 238.0 147601.7 142910.4 144762.5
0.290 79.2 121511.0 116111.4 118671.8
0.290 158.0 139899.0 134351.1 137059.8
0.290 238.0 156660.7 150806.1 153821.5
0.550 79.2 111782.9 104395.7 108943.7
0.550 158.0 120483.9 112724.6 117644.7
0.550 238.0 130665.5 122660.1 127826.3
B. Godongwana et al. / Journal of Membrane Science 303 (2007) 86–99 95

gravitational force, whereas the others do not. Eq. (45) indicates points then the round off error ε may be expressed as the series
a linear decline in the hydrostatic pressure along the length of of the form:
the membrane. This result is in agreement with Catapano’s [11]
contention that the axial pressure profile is linear when there 
N/2

is ‘no-slip’ velocity. A summary of other experimental condi- ε(Z, R, τ) = exp(am τ + ikm R + ilm Z) (65)
tions of operation including variations in membrane length and m=1

nutrient flow rates, aimed at validating the developed model, are


in terms of a constant am that may possibly depend upon m
given in Table 6.
and km = lm = 2πm, where m = 1, 2, 3, . . ., N/2. If μ is the exact
solution of the finite difference form of the differential Eq. (63),
4.2. Finite difference scheme formulation then setting:

The difference equations developed and solved in this study, U(Z, R, τ) = μni,j + εni,j (66)
Eqs. (60) and (61), each contain only one unknown and can be
solved explicitly for this unknown in a straightforward man- results in a difference equation for εni,j
n+1 n+1
ner. The unknowns are Ui,j and Vi,j , since we assume that
U and V are known at all grid points at time level n. By defi-
i,j − εi,j
εn+1 n εni,j+1 − 2εni,j + εni,j−1 εni,j+1 − εni,j−1
nition therefore, the solution of Eqs. (60) and (61) follows an = +
explicit approach. Besides its simplicity compared to the implicit τ (R)2 2RR
approach, a major disadvantage of an explicit approach is that εn
2 i+1,j
− 2εni,j + εni−1,j
there are restrictions on the values of the independent variables. +β 2
(67)
(Z)
An exact stability analysis of the difference representation of the
nonlinear Navier–Stokes equations does not exist [22]. Since a time step has size τ, an axial space step Z and a
radial space step size R, there is an amplification factor for the
4.2.1. Possible sources of error rounding off error ε, viz.
Sources of error in the numerical scheme could possibly arise
 n+1   
ε 
 i,j  am τ 
in the accuracy of the experimental determination of the input 2[1−cos(km R)] sin(km R)
variables (e.g. κ, p0 , p1 , π, ρ), and the fit of the criteria for  n  =e = 1 − τ −
 εi,j  (R)2 RR
convergence of the scheme (i.e. the continuity equation). The

most sensitive of the input variables are the inlet and outlet pres- 2β2 [1 − cos(lm Z)] 
sures, in that, an increase of 1 kPa in the inlet pressure will + ≤1 (68)
(Z)2
result in a 17% error in the prediction of the axial velocity
n+1
for the next time step Ui,j . The convergence of the scheme This inequality has to be satisfied in order for the linearized form
is also very sensitive to changes in the hydraulic permeability, of the difference Eq. (56) to have a stable solution, and implies
κ. Slight changes/errors in the other input variables will also that
affect the accuracy of the scheme but to a very small degree in 
comparison to the input and output pressures and the hydraulic 2[1 − cos(km R)] sin(km R)
τ −
permeability. (R)2 RR

2β [1 − cos(lm Z)]
2
4.2.2. Von Neumann stability analysis + ≤2 (69)
If one looks for a stability criteria for the Navier–Stokes dif- (Z)2
ferential Eqs. (56) and (57), a fairly accurate treatment entails
besides the fact that the curly brackets of (68) is non-negative, i.e.
ignoring the inertial and normal stresses for negligible bulk
2[1 − cos(km R)]/(R)2 + 2β2 [1 − cos(lm Z)]/(Z)2 ≥ sin(km
Reynolds number Reb  1, so that Eqs. (56) and (57) become
R)/RR. [This latter stability condition simplifies to
linear in the functions U and V. Despite an exact treatment not
(1 + β2 )Rkm = 2πR(1 + β2 )m ≥ 1 in the limit that R and Z
being presented here, one can still get an idea of the stability
tend to 0.] The stability requirement for the solution of the
requirement upon the difference Eqs. (60) and (61) by studying
difference Eq. (67) [having set sin(km R) = 1, cos(km R) = 0
the linearized Navier–Stoke equations:
and cos(km Z) = 0, at the extremity of the curly brackets of
 
∂U 1 ∂ ∂U ∂2 U ∂P (69)] is the constraint:
= R + β2 2 − β +b (63)
∂τ R ∂R ∂R ∂Z ∂Z
(ΔR)2
and τ ≤ (70)
1 − 1/2(R/R) + β2 (R/Z)2
  2
∂V 1 ∂ ∂V V 2∂ V ∂P
= R − 2
+ β 2
− (64) Here (R/R) is the fractional radial increment. Also of inter-
∂τ R ∂R ∂R R ∂Z ∂R
est is the differential equation for V, Eq. (64). The round
for negligible values of the bulk Reynolds number or by ignoring off error δ(Z, R, τ) of its linear difference equation satisfies
some of the inertial and normal stress effects. If there are N grid another difference equation similar to that of the previous case
96 B. Godongwana et al. / Journal of Membrane Science 303 (2007) 86–99

for ε: Table 7
Grid independence of the flow field
i,j − δi,j
δn+1 n δni,j+1 − 2δni,j + δni,j−1 δni,j+1 − δni,j−1
= + τ U V (Z; R)
τ (R)2 2RR
0.1 2.33 × 10−6 9.74 × 10−8 (0.38; 0.09)
δn
2 i+1,j
− 2δni,j + δni−1,j 0.05 2.29 × 10−6 1.38 × 10−7 (0.38; 0.09)
+β (71) 0.01 2.26 × 10−6 1.70 × 10−7 (0.38; 0.09)
(Z)2 0.005 2.26 × 10−6 1.75 × 10−7 (0.38; 0.09)
0.001 2.25 × 10−6 1.78 × 10−7 (0.38; 0.09)
the main difference being the presence of an additional curvature
Analytical solution 2.29 × 10−6 1.79 × 10−7 (0.38; 0.09)
term −δnij /R2 . Generalise the round off error to

δm = exp(bm τ + ikm R + ilm Z) (72)

The constants bm may vary with m, while the constants km and


lm also depend upon m: km = lm = 2πm, for m = 1, 2, 3, . . ., N/2 as
was seen earlier in the stability analysis for the difference equa-
tion version of Eq. (63). Then the amplification factor |δn+1
i,j /δi,j |
n

is given by
 

1 − τ 2[1 − cos(km R)] − sin(km R) + 1
 (R)2 RR R2

2β2 [1 − cos(lm Z)] 
+ ≤1 (73)
(Z)2
In the limit that R and Z tends to zero, one of the two
possibilities is that
2πm 1
4(1 + β2 )π2 m2 − + 2 ≥0 (74) Fig. 10. Convergence of the numerical solution as a function of grid spacing.
R R
and the other [having put sin(km R) = 1, cos(km R) = 0, and The rate of convergence of the numerical solution is influ-
cos(lm Z) = 0] is that enced by a number of factors including the aspect ratio, the
Reynolds number and the grid spacing. In Fig. 9 the convergence
(ΔR)2
τ ≤ is defined as Un + 1 /Un , and the solution is said to be converged
1 − (1/2)R/R) + (1/2)(R/R)2 + β2 (R/Z)2 when the value of the convergence is unity. The influence of
(75) the grid spacing is that the rate of convergence increases with
This is the stability condition for the linearized difference Eq. increasing grid spacing, and this is demonstrated by a compar-
(61), and is in fact a stronger form of stability condition (70). ison of a grid spacing of 0.02 and 0.0007 in Fig. 10. On the
Unfortunately a treatment of non-linear Eqs. (60) and (61) will other hand, the accuracy of the solution decreases with increas-
not be done here, but constraint (75) at least gives us a good idea ing grid spacing. The grid spacing required therefore and the
of the stability condition for linearized Navier–Stokes equations. corresponding accuracy will be dictated by the application of
The fractional radial increment is always less than unity, except the numerical solver, and a trade off will have to be reached
possibly at the start j = 0. The denominator of (75) is positive between a comprehensive solution that takes up a lot of comput-
definite and greater than one, which guarantees that the numer- ing time and a little less precise solution that is fairly quick to
ical solution is stable in almost all cases where the increments solve [22].
R and Z are chosen to be much less than one.
A tenable solution of a differential equation, when using a 5. Conclusion
finite difference scheme (or finite volume), should be indepen-
dent of the number of grid points (or volume cells) used. If this A generic equation for small Rew flows through SFMGR’s
condition is not met then the solution is not stable. This means was developed. The model was solved for the specific case of
that the steady-state values of U and V should be independent of negligible angular variations in the flow profiles (i.e., m = 0).
the time increment chosen. Generally to remedy this situation From this generic equation, detailed models of momentum trans-
the number of grid points used must be increased, that is, the fer accounting for gel formation, osmotic pressure, the mode
time increment needs to be decreased. In Table 7, decreasing of operation, and orientation of a HF and/or capillary MGR
τ from 0.01 to 0.005 improves the numerical solution only were developed. The accuracy of these models was tested on a
marginally; therefore the solution that uses τ equal to 0.01 is vertically orientated SFMGR operated in the dead-end mode.
essentially grid-independent. The model parameter values cor- These models were based on a clean membrane without any
responding to the velocity values in Table 7 are listed in Table 1, bio-organism; however, since the application of interest of the
and the spatial increments R and Z were chosen to be 0.02. developed models is the optimisation of enzyme production,
B. Godongwana et al. / Journal of Membrane Science 303 (2007) 86–99 97

future work will incorporate the kinetics of the P. chrysosporium 


fungus with the momentum transfer. f (R) = R42 − R4 − 2R23 (R22 − R2 )
 
R 2 R R
2 3
6. Future work − 8R22 2
ln − R2
2 R3 4 2
In future work, these models will be modified for systems  2 
R R R3 2
with growth on the shell side of the bioreactor. − ln − R (A7)
2 R3 4
Acknowledgements
ω2 f (Rx )(PL(0) − PS(0) )
The authors feel indebted to the Mandela-Rhodes Foundation A= (A8)
f (Rx ) − Rx
and the National Research Foundation for their financial support
towards the completion of this study. ω3 f (Rx )2 (PL(0) − PS(0) ) sinh ω
The Institute of Polymer Science at the University of Stellen- C= (A9)
[f (Rx ) − Rx ][f (Rx )(f − cosh ω)Rx (f − 1)]
bosch for providing the capillary membranes.
To Dr. van der Merwe for her contribution in validating some f (Rx )(PL(0) − PS(0) )
D = PL(0) − (A10)
of the mathematical principles in this study. f (Rx ) − Rx

Appendix A. Vertical orientation (variable shell side ωRx f (Rx )(PL(0) − PS(0) ) sinh ω
pressure) H =b+
[f (Rx ) − Rx ][f (Rx )(f − cosh ω) + Rx (f − 1)]
(A11)
The solutions developed in this section are applicable to a
system with a shell side hydrostatic pressure that is a function the dimensionless shell volumetric flowrate, ΩS , is given by
of the axial position. The treatment is the same as for the con-  

stant shell side pressure bioreactor system. The axial and radial 1 R23 1
ΩS = R2 ln R3 +
2
− +
velocity profiles, as well as the luminal volumetric flowrate UL , 2 2 4
VL , ΩL are still governed by the same expressions; i.e., Eqs. ∞
 
(38), (40) and (51), respectively. The new expressions for the 4  dPS
Bn e−αn τ J1 (αn )
2
+ −b (A12)
shell profiles are obtained, as before, from Eq. (1) neglecting αn b dZ
n=1
the inertial terms:
     the stream function in the shell side of the membrane bioreactor
∂vzS 1 ∂ ∂vzS ∂pS
ρ =μ r − + ρgz (A1) is given by
∂t r ∂r ∂r ∂z  
dPS 1  4
From Eq. (A1) making use BC7–11, the new expressions for the ψS = −b + R2 − R4 − 2R23 (R22 − R2 )
shell side profiles are given by dZ 16
   

  ∞
R22 R2 R22 R2 R R2
1 R 4 −α2n τ − 8R22
ln − − ln −
US = − 2
2R2 ln +R3 −R −
2 2
Bn e J0 (αn R) 2 R3 4 2 R3 4
4 R3 b
n=1  
  dPS(1) dPS(0)
dPS × − (A13)
× −b (A2) dZ dZ
dZ
where
1 d 2 PS R22 R2
VS = f (R) (A3) = [ln(R3 − R2 ) − ln R3 ](R23 − R22 ) − 2 (R23 − R22 )
16R dZ2 4 8
ARx cosh ωZ CRx sinh ωZ R23 2 R4 − R42
PS = + + HZ + D (A4) + (R3 − R22 ) − 3 (A14)
ω2 f (Rx ) ω3 f (Rx ) 8 16
where A, C, D and H are constants. The corresponding luminal
pressure profile is Nomenclature
A cosh ωZ C sinh ωZ
PL = + + HZ + D (A5) a dimensionless entrance pressure drop
ω2 ω3 A constant in pressure equation (bioreactor of vari-
where able shell side pressure)
  b dimensionless function of gravitational accelera-
f (Rx ) − Rx
ω = 16κ
2
(A6) tion
f (Rx )
98 B. Godongwana et al. / Journal of Membrane Science 303 (2007) 86–99

Bn constants of integration, n = 0, 1, 2, 3, . . . ω arbitrary constant


C constant in pressure equation (bioreactor of vari- Ω dimensionless flowrate
able shell side pressure)
dg gel layer thickness in the membrane lumen (m) Subscripts
dw membrane wall thickness (m) 0 membrane entrance
D constant in pressure equation (bioreactor of vari- 1 membrane exit
able shell side pressure) 2 extra capillary space radius (distance from the
f fraction retentate centre of the membrane to the centre of the extra
g gravitational acceleration (m/s2 ) capillary space of the glass manifold)
H constant in pressure equation (bioreactor of vari- 3 glass manifold inner radius
able shell side pressure) ∞ steady state
i horizontal grid coordinate L membrane lumen
j vertical grid coordinate M membrane matrix
Jn (α) Bessel function of order n of the first kind S shell side of membrane
kg gel layer permeability (m2 ) Sbπ function of PS , Π and b
km membrane hydraulic permeability (m2 ) τ transient state
K function of dimensionless time w wall
L effective membrane length (m) z, r, θ cylindrical spatial co-ordinates
Lb glass bioreactor length (m)
p pressure (Pa)
P dimensionless pressure References
Q volumetric flowrate (m3 /s)
r radial coordinate (m) [1] H. Willershausen, A. Jager, H. Graf, Ligninase production of Phane-
R dimensionless radial coordinate rochaete chrysosporium by immobilization in bioreactors, J. Biotechnol. 6
(1987) 239–243.
RL membrane inner radius (m)
[2] R. Venkatadri, R.L. Irvine, Cultivation of Phanerochaete chrysosporium
Reb bulk fluid Reynolds number (ρvz0 rH /μ) and production of Lignin peroxidase in novel biofilm reactor systems: hol-
Rew wall Reynolds number (ρvw rH /μ) low fiber reactor and silicone membrane reactor, Water Res. 27 (1993)
Sc Schmidt number (μ/ρDAB ) 591–596.
t time (s) [3] W.D. Leukes, Development and characterisation of a membrane gradostat
bioreactor for the bioremediation of aromatic pollutants using white rot
T absolute temperature (K)
fungi, Ph.D. Thesis, Rhodes University, Grahamstown, 1999.
U dimensionless axial velocity [4] M.T. Moreira, G. Feijoo, C. Palma, J.M. Lema, Continuous production of
v velocity (m/s) Manganese Peroxidase by Phanerochaete chrysosporium immobilized on
V dimensionless radial velocity polyurethane foam in a pulsed packed-bed bioreactor, Biotechnol. Bioeng.
Yn (α) Bessel function of order n of the second kind 56 (1997) 130–137.
[5] A. Dominguez, I. Rivela, R.S. Couto, M.A. Sanroman, Design of a new
z axial coordinate (m)
rotating drum bioreactor for ligninolytic enzyme production by Phane-
Z dimensionless axial coordinate rochaete chrysosporium grown on an inert support, Process Biochem. 37
(2001) 549–554.
Greek letters [6] S. Govender, Optimisation studies on a membrane gradostat bioreactor
α independent variable in Bessel function for ligninase production using white rot fungi, MTech Thesis, ML Sultan
β aspect ratio of the membrane Technikon, Durban, South Africa, 2000.
[7] M.S. Solomon, Membrane bioreactor production of lignin and manganese
Γ gamma function
peroxidase, MTech Thesis, Cape Technikon, Cape Town, 2001.
ε is the porosity of the gel layer [8] C.J. Garcin, Design and manufacturing of a membrane bioreactor for the
θ angle in cylindrical coordinates (tan−1 (y/x)) cultivation of fungi, MSc Thesis, Rhodes University, Grahamstown, 2002.
Θ dimensionless function of the angle θ [9] M.S. Sheldon, H.J. Small, Immobilisation and biofilm development of
κ dimensionless membrane permeability Phanerochaete chrysosporium on polysulphone and ceramic membranes,
J. Membr. Sci. 263 (2005) 30–37.
μ fluid dynamic viscosity (Pa s)
[10] L.J. Kelsey, M.R. Pillarella, A.L. Zydney, Theoretical analysis of convec-
Ξ function of the radial coordinate tive flow profiles in a hollow fiber membrane bioreactor, Chem. Eng. Sci.
Π dimensionless osmotic pressure 45 (1990) 3211–3220.
ρ fluid density (kg/m3 ) [11] G. Catapano, G. Iorio, E. Drioli, C.P. Lombardi, F. Crucitti, G.B. Doglietto,
τ dimensionless time M. Bellantone, Theoretical and experimental analysis of a hybrid bioartifi-
cial membrane pancreas: a distributed parameter model taking into account
 dimensionless function of bioreactor dimensions
Starling fluxes, J. Membr. Sci. 52 (1990) 351–378.
Φ osmotic pressure (Pa) [12] J.P. Thakaran, P.C. Chau, Operation and pressure distribution of immo-
χ distortion of the velocity profile (m−1 ) bilised cell hollow fiber bioreactors, Biotechnol. Bioeng. 28 (1986)
ψ dimensionless stream function 1064–1071.
[13] G. Catapano, G. Iorio, E. Drioli, M. Filosa, Experimental analysis of a
cross-flow membrane bioreactor with entrapped whole cell: influence of
B. Godongwana et al. / Journal of Membrane Science 303 (2007) 86–99 99

transmembrane pressure and substrate feed concentration on reactor per- [19] S.K.O. Ntwampe, M.S. Sheldon, Quantifying growth kinetics of
formance, J. Membr. Sci. 35 (1988) 325–338. Phanerochaete chrysosporium immobilised on a vertically orientated
[14] W.J. Bruining, A general description of flows and pressures in hollow fiber polysulphone capillary membrane: biofilm development and substrate con-
membrane modules, Chem. Eng. Sci. 44 (1989) 1441–1447. sumption, Biochem. Eng. J. 30 (2006) 147–151.
[15] Y. Moussy, Bioartificial kidney. I. Theoretical analysis of convective flow in [20] E.P. Jacobs, W.D. Leukes, Formation of an externally unskinned polysufone
hollow fiber modules: application to a bioartificial hemofilter, Biotechnol. capillary membrane, J. Membr. Sci. 121 (1996) 149–157.
Bioeng. 68 (1999) 142–152. [21] M. Tien, T.K. Kirk, Lignin peroxidase of Phanerochaete chrysosporium,
[16] K. Damak, A. Ayadi, B. Zeghmati, P. Schmitz, A new Navier–Stokes and Methods Enzymol. 16 (1988) 238–249.
Darcy’s law combined model of fluid flow in crossflow filtration tubular [22] J.D. Anderson, Computational Fluid Dynamics: the Basics with Applica-
membranes, Desalination 161 (2004) 67–77. tions, McGraw Hill, New York, 1995.
[17] M. Elshahed, Blood flow in capillary under starling hypothesis, Appl. Math. [23] R.H. Perry, D.W. Green, J.O. Maloney (Eds.), Perry’s Chemical Engineer’s
Comput. 149 (2004) 431–439. Handbook, 7th ed., McGraw Hill, New York, 1998.
[18] A.S. Berman, Laminar flow in channels with porous walls, J. Appl. Phys. [24] X.-Y. Li, B.E. Logan, Permeability of fractal aggregates, Water Res. 35
24 (1953) 1232–1235. (2001) 3373–3380.

You might also like