Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

ARTICLE

pubs.acs.org/IECR

Mathematical Analysis of the Meso-Scale Flow Field in Spiral-Wound


Membrane Modules
Margaritis Kostoglou*,† and Anastasios J. Karabelas‡

Department of Chemical Technology, School of Chemistry, Aristotle University of Thessaloniki, Thessaloniki, Greece

Chemical Process Engineering Research Institute, Centre for Research and TechnologyHellas, sixth km Charilaou-Thermi Road,
GR 570 01, Thermi-Thessaloniki, Greece

ABSTRACT: The use of flat-sheet, spiral-wound, membrane modules for reverse osmosis and nanofiltration applications is very
extensive. Design and performance optimization of these modules requires sound mathematical modeling. This study focuses on the
mathematical analysis of the mesoscale hydrodynamic equations for the narrow channels with spacers, of the entire membrane sheet,
previously derived from the microscale momentum conservation laws14. The mathematical problem is enhanced by considering a
spatial dependence of the retentate channel and membrane permeabilities to account for fouling/scaling, aiming at future use of the
proposed techniques to simulate long time fouling dynamic behavior of the process. The formal mathematical treatment of the
original problem leads to several levels of approximation (depending on the problem parameter values) which admit either analytical
or numerical solutions with reduced dimensionality, or numerical solutions with reduced convergence difficulties. To confirm the
validity of conclusions obtained by following these procedures, several results from simplified cases are compared with numerical
solutions of the original problem. Furthermore, all possible simplifications and analytical solutions of the particular problem have
been obtained, as well as the conditions under which they hold, thus forming the basis for more comprehensive modeling, including
mass transfer and scaling/fouling phenomena. Specific criteria are also provided for selecting appropriate simplified solutions to
specific cases, helpful in the development of flow and membrane fouling simulators.

1. INTRODUCTION the modules. The problem becomes much more difficult if mass
The spiral wound membrane modules (SWM) dominate the transfer is taken into account, where a very high spatial resolution
field of high pressure membrane processes; that is, reverse is needed close to the wall because of the rather high Schmidt
osmosis and nanofiltration.1 The commercial success of this type number of the solute.4,5 Moreover, the effect of dynamic
of modules intensifies the need for optimization, regarding both phenomena such as fouling and scaling, which manifest them-
their geometrical characteristics and operating conditions, in selves at a rather large time scale, pose additional challenges to
order to improve their long-term operation and thus reduce the the modeling task.
operating cost.2 An important aspect of an integrated optimiza- Direct comprehensive modeling of the SWM module opera-
tion process is the development of advanced computational tools tion, given the above complexities, is out of the question; thus,
for the simulation of module operation. Through the reliable several simplified approaches have been employed, including
computation of key parameters, such as the spatial distribution of reduction of dimensionality,6 spatial scale separation using the
trans-membrane flux, as well as the pressures and velocities at the unit cell concept,7 ignoring the intrinsically transient character of
retentate and permeate sides, throughout membrane sheets/ the flow.8 The unit cell approach is promising,9,10 in that
envelopes, one can obtain significant insight into the conditions “constitutive relations” are derived from detailed, direct numer-
favoring fouling and scaling, in addition to average quantities ical simulations11 of the elementary unit and, assuming repetition
necessary for system optimization (e.g., axial pressure drop at of the flow pattern, they are incorporated in the conservation
retentate side, lateral pressure variation at the permeate side). equations for the entire membrane sheet.12,13 In a previous
Unfortunately, modeling transport phenomena inside the com- publication,14 a detailed derivation of the mesoscopic hydro-
plicated geometry of SWM modules is not a straightforward task. dynamic equations from the detailed filament scale momentum
SWM modules are comprised of several large size membrane transfer equation can be found. The resulting mesoscale liquid
envelops separated by a net-type spacer at the retentate flow mass balance is very important because it defines the distribution
channel; a porous cloth/filler is placed in-between the membrane of wall velocity on the membrane, which in turn determines the
sheets of an envelop, at the low pressure permeate side. Thus, the fouling and scaling rates and correspondingly their distribution
two flow fields interact through the membrane. The very on the membrane. Colloidal fouling and scaling due to sparingly
significant modeling difficulties are due to the complicated
retentate-side geometry (characterized by several highly dispa-
rate size scales, from the small net-filament diameter to the large Received: October 13, 2010
sheet dimensions), as well as to the unsteady, essentially Accepted: February 24, 2011
chaotic,3,4,11 flow field generated by the spacers, at usual liquid Revised: February 7, 2011
velocities. These features characterize only the hydrodynamics of Published: March 15, 2011

r 2011 American Chemical Society 4653 dx.doi.org/10.1021/ie102083j | Ind. Eng. Chem. Res. 2011, 50, 4653–4666
Industrial & Engineering Chemistry Research ARTICLE

soluble salts are major problems,1 causing membrane perfor- former case the correct approach is not to use directly the
mance degradation and reduction of overall plant efficiency. constitutive relation but to incorporate it in a flow field model
Over the years several approximate analytical solutions have like the one presented in the present work; in the latter case, the
been proposed for particular problem formulations, either for the complexity of the geometry prevents the direct modeling of flow
pure hydrodynamics problem or as part of modeling the filtration in the retentate channel suggesting the “periodic unit cell”
operation.1518 In the general case, these equations must be approach. Although the highly inertial and rather chaotic char-
solved numerically, but according to the suggestion of Aris19 an acter of the flow in the feed-side channel creates some reserva-
equation must be first analyzed mathematically in order to derive tions regarding the accuracy of this particular approach,14 it is the
all possible simplifications as well as analytical and asymptotic best alternative up to now, considering that direct numerical
results, and, thus, understand the structure of the problem. This treatment of the membrane sheet scale is computationally
knowledge is usually important even for the appropriate numer- prohibitive. To proceed, advantage is taken of extensive unit cell
ical handling of the problem. Fortunately, for the particular CFD simulations,11 showing that the pressure drop is related to
problem at hand, the values of parameters typical of practical the superficial, cross sectional average, flow velocity U as follows
applications fall in a region of the parameter space where
Δp U2F
considerable simplification is possible. The present work deals ¼  f 1 Re f 2 ð1Þ
with a formal mathematical analysis of the hydrodynamic equa- ΔL D
tions (derived in a previous work14) that are enhanced by where the Reynolds number is defined as Re = UDμ/F and F and
introducing permeability distributions to account for membrane μ are the fluid density and viscosity, respectively. The retentate
fouling/scaling phenomena; these equations hold at the mem- channel pressure is p and the spacer filament diameter is D. The
brane-sheet spatial scale in SWM, and lead to unification, common (net-type) spacers dealt with are comprised of two
correction/confirmation and rationalization/explanation of all layers of parallel unwoven filaments; thus, the channel gap is
previous approximate analytical solutions derived on the basis of twice the filament diameter. Detailed geometric description of
heuristic arguments. The spatial distributions of retentate chan- spacers is provided elsewhere.11 The coefficients f1 and f2 depend
nel and membrane permeabilities considered in this work, allow on the spacer geometry (pitch to diameter ratio H/D and
future incorporation of the proposed methods to simulation filament crossing angle β) and on the flow incidence angle θ.
codes for tackling the complete membrane operation problem. For cases of nearly unidirectional flow, the angle dependence can
The structure of the present work is as follows: First, the be ignored and the normal incidence (θ = π/2) values of
governing equations are presented and properly nondimensiona- parameters can be used, corresponding to isotropic perme-
lized to derive the dimensionless parameters of the problem. The ability.14 As outlined above, to generalize the problem handled
numerical approach for the solution of equations to be used for the in the present work, it is considered that fouling or scaling may
assessment of the approximate techniques is described next. Then affect the permeability at the retentate channel. The distribution
using a formal and systematic approach, a hierarchy of simplifica- of such deposits may be represented by multiplying the left-hand
tions is presented depending on the values of the parameters. This side of constitutive relation 1 by a spatially dependent relative
hierarchy includes analytical solutions, dimensionality reductions permeability K(x,y) where the maximum value K = 1 corre-
or simplification of the original problem, still requiring however a sponds to the clean case. This is equivalent to assuming that only
(much easier) numerical solution. Finally, several numerical f1 and not f2 depends on the local degree of fouling. The only
results are presented to examine and confirm the validity of the restriction associated with the above fouling consideration is the
simplified approaches derived in the present work. implicit assumption that the nonlinear pressure drop-Re relation
holds for any degree of fouling. This assumption is necessary
2. PROBLEM FORMULATION since direct modeling of flow field for cases with fouling is
extremely difficult; indeed, the pressure drop depends on the
2.1. Derivation of Equations. 2.1.1. Retentate Channel. Be- exact distribution of fouling in the unit cell, the determination of
fore proceeding with problem formulation, it should be recognized which appears to be far more difficult than the computation of
that the flow field at the retentate side is complicated not only the flow field. On the other hand, it is a reasonable leading order
because of the presence of the net-type spacers but also because of approximation based on a mathematical expansion of the con-
commonly occurring phenomena of fouling and scaling. These stitutive relation, around the “clean” membrane case.
phenomena have the following effects in regard to the flow field Generalizing the relation 1 by considering a differential pressure
modeling: First, the effective permeability in the narrow retentate drop in an arbitrary direction and substituting for Re leads to
channels tends to be reduced because of scale20 or foulants21
depositing on the spacers and partly on the membrane, not KD1 þ f 2
B¼ 
U rp ð2Þ
necessarily uniformly, with a concomitant effect on the flow field. f 1 F1  f 2 μf 2 U 1  f 2
Second, thin fouling layers (e.g., comprising organic foulants, or
colloids) on the membrane22 tend to reduce its effective perme- By taking the magnitude of both sides, of the above vectorial
ability, without significantly affecting the channel permeability, identity, solving for U and replacing again in eq 2 one obtains
thus directly affecting permeate flux. These effects will be ac- !1=ð2  f 2 Þ
counted for in the following problem formulation. KD1 þ f 2
B¼ 
U jrpjðf 2  1Þ=ð2  f 2 Þ rp ð3Þ
A key feature in modeling the flow field in the spacer-filled f 1 F1  f 2 μf 2
SWM module is the “constitutive” expression relating local
pressure drop in the channel to the local cross sectional average where the velocity vector is averaged along a direction normal to
flow velocity. Such a constitutive relation can be obtained either membrane surface.
by fitting appropriate expressions to experimental pressure drop 2.1.2. Permeate Channel. For the permeate channel, a linear
vs flow rate data, or from subscale CFD simulations. In the resistance law based on the prevailing smaller flow rates and smaller
4654 dx.doi.org/10.1021/ie102083j |Ind. Eng. Chem. Res. 2011, 50, 4653–4666
Industrial & Engineering Chemistry Research ARTICLE

pore characteristic sizes can be typically assumed. The absence of


fouling in this side implies a uniform permeability leading to the
following relation for the cross sectional average velocity u:
k
u ¼  rP
B ð4Þ
μ
where P is the permeate side pressure and k the corresponding
permeability. The permeate velocity vector is averaged along the
direction normal to the membrane. The membrane flux is related to
the membrane permeability kw, membrane thickness δ and the
local trans-membrane pressure through the following linear relation
(Darcy’s law):
kw
uw ¼ ðp  PÞ ð5Þ
δμ
where as positive direction for uw is assumed by convention the
direction from retentate to permeate (the actual one). As
outlined above, to generalize the model for membranes that
have suffered a reduction of their effective permeability because
Figure 1. Schematic of the membrane element geometry considered:
of a thin fouling layer, the right-hand side of the above equation
(a) flow field cross section and (b) top view.
is multiplied by a spatially dependent relative permeability
M(x,y) with a maximum value M = 1 corresponding to the
nonfouled surface. The differential mass balances, in x and y the least resistance. Therefore, the only correct choice for the
directions, for the retentate and permeate channels lead to the retentate channel pressure is to define it at x = 0 and x = Lx, that
following equations: is, to consider the pressure drop as an input parameter and the
flow rate as the outcome of the problem solution. The inlet
Retentate side Lrz r 3 U
B ¼  uw ð6Þ manifold of the membrane module may create a nonuniform
inlet pressure, but in any case, it is expected to be small (since the
Permeate side Lpz r 3 B
u ¼ uw ð7Þ pressure drop in the membrane module is much larger than in the
manifold), and the only way to take it into account is to simulate
where Lrz and Lpz are the retentate and permeate channel gaps,
the flow into the entire manifold, which is out of the scope of the
respectively.
present work.
Combining eqs 37 with the previously mentioned general-
Considering that the inlet flow is toward the module
izations, a system of two coupled partial differential equations for
and the exiting retentate in the same direction, leads to
the pressure fields p and P is obtained:
the following expressions for the inlet and outlet velocity
r 3 K 1=ð2  f 2 Þ ðx, yÞjrpjðf 2  1Þ=ð2  f 2 Þ rp profiles:
!1=ð2  f 2 Þ   !1=ð2  f 2 Þ
KD1 þ f 2 Dp
D1 þ f 2 Mðx, yÞkw U in ðyÞ ¼  1  f f ð10Þ
¼ ðp  PÞ ð8Þ f 1F 2μ 2 Dx x ¼ 0
f 1 F1  f 2 μ f 2 δμLrz

  !1=ð2  f 2 Þ
Mðx, yÞkw KD1 þ f 2 Dp
rP¼2
ðP  pÞ ð9Þ U out ðyÞ ¼  1f f ð11Þ
kδLpz f 1F 2μ 2 Dx x ¼ Lx
This system needs appropriate boundary conditions. In Fig-
The average inlet and outlet velocities Uin,av and Uout,av
R can be
ure 1, the no penetration (zero velocity boundary) conditions at
obtained by applying the averaging operator 1/Ly L0 y...dy to
the three boundaries of the permeate side (at x = 0, x = Lx, and y =
the above functions. The flow recovery fraction R (permeate
Ly) are evident. The boundary condition at the outflow edge of
outlet flow/retentate inlet flow) can be computed as R = 1 
the permeate side (y = 0) can be set to be zero; it is noted that this
[Uout,av/Uin,av]. An alternative way is to use the flux through
is not restrictive at all since only pressure differences appear in
the membrane, that is,
the equations. The no penetration condition also holds for the
Z Lx Z Ly
two boundaries of the retentate side (at y = 0 and y = Ly). The
remaining boundary conditions are not straightforward; in R ¼ uw dxdy=ðLy Lrz U in, av Þ ð12Þ
0 0
particular, for a one-dimensional problem, it is reasonable to
choose as input parameters the flow rate and outflow pressure
and to obtain the inlet pressure (i.e., the pressure drop) from 2.2. Nondimensionalization. The mathematical problem is
problem solution. This approach cannot be extended to the now complete and contains many physical parameters. To
2-dimensional case, considered in the present work; indeed, the analyze it, a grouping is needed of the physical parameters to a
inlet velocity is not uniform but its profile is part of the problem few dimensionless ones. Let the retentate inlet pressure be pin
solution. The fundamental parameters are the inlet and the outlet and the outlet pressure po. All pressures are normalized by pin and
pressures; consequently, the flow adjusts itself to a route of all lengths and spatial coordinates by Ly. The dimensionless form
4655 dx.doi.org/10.1021/ie102083j |Ind. Eng. Chem. Res. 2011, 50, 4653–4666
Industrial & Engineering Chemistry Research ARTICLE

of eqs 8 and 9 is R ¼ 1  U out, av =U in, av ð26cÞ


1þR R
r3K ðx, yÞjrpj rp ¼ AMðx, yÞðp  PÞ ð13Þ
uw ¼ Mðx, yÞðp  PÞ ð26dÞ
r2 P ¼ Mðx, yÞACðP  pÞ ð14Þ
The term before the pressure gradient in eq 13 is used to
where obtain a reference value to the total permeability. The existence
of reference values is important for the following mathematical
f2  1
R¼ ð15Þ treatment of the problem. The exponent R takes values between
2  f2 0 (linear pressure drop relation) to 0.5 (fully inertia dominated
!1=ð2  f 2 Þ !ðf 2  1Þ=ð2  f 2 Þ flow). Unit cell CFD simulation revealed11 that typical values for
f 1 F1  f 2 μ f 2 Ly kw L2y practical flow conditions and commercial spacers are between
A¼ ð16Þ 0.35 and 0.5. The parameter A is related to the ratio of the
D1 þ f 2 pin δLrz membrane permeability to the retentate channel permeability,
whereas the parameter C is related to the ratio of the retentate
kw L2y 1 channel to the permeate channel permeability. The use of
C¼ ð17Þ another parameter B = AC instead of C is more straightforward
kδLpz A
but the particular choice of parameters in the present work
with boundary conditions facilitates the mathematical analysis.
The dimensionless problem is now complete. Inputs include the
Dp parameters A, C, R, L
¼ 0 at y ¼ 0, y ¼ 1 ð18a, bÞ hx, hp0 and the two functions K(x,y) and M(x,y);
Dy the main outputs are the flow recovery fraction R, the average inlet
velocity Uhin,av, and the membrane flux distribution huw(x,y).
DP
¼ 0 at y ¼ 1 ð19Þ
Dy 3. NUMERICAL SOLUTION
The mathematical problem described above is solved using a
DP commercial finite element code (COMSOL Multiphysics). De-
¼ 0 at x ¼ 0, x ¼ Lx ð20a, bÞ
Dx spite the smoothness of the solution, it seems that the non-
linearity of the velocity-pressure drop relation introduces
P ¼ 0 at y ¼ 0 ð21aÞ stiffness to the problem. Convergence of the steady state
problem using nonlinear stationary solvers is not possible.
p ¼ 1 at x ¼ 0 ð21bÞ Instead, the handling of the problem as a (pseudo-)transient
one and the use of implicit ordinary differential solvers leads to
_ convergence; but this convergence requires a very large number
p ¼ po at x ¼ L x ð21cÞ of time steps. In the present work, a relatively coarse unstructured
The inlet and outlet velocities are normalized by the velocity Uo triangular grid consisting of 1000 quadratic elements is em-
defined as ployed. Comparison of the results with those obtained by a
!1=ð2  f 2 Þ denser grid of 4000 elements shows differences of the order of 3
D1 þ f 2 pin  po per thousand for the main variables. The computational time for
Uo ¼ ð22Þ the denser grid is significant (between 5 and 10 min) pointing to
f 1 F1  f 2 μf 2 Lx
the usefulness of the simplified solutions derived in the present
and the membrane flux is normalized by the reference value uwo work. The results of the numerical code will be used in this work
defined as for the assessment of the asymptotic approximate solutions.

kw 4. ANALYTICAL TREATMENT
uwo ¼ p ð23Þ
δμ in
4.1. On the Values of A. Henceforth only dimensionless
Also a reference value for the flow recovery fraction R can be variables are treated; thus, the overbar is omitted for clarity of
defined as R0 = uwoLx/(LrzUin,av). Using the above normal- presentation. To proceed with the analysis, it is recognized that
ization, results in the following expressions for the output the membrane module operation, under conditions encountered in
variables: practice, sets some limits in the values of the main problem
_ !1 þ R parameters. On the basis of the geometric dimensions of the
KL x Dp commercial modules and the fact that the pressure drop in the
U in ðyÞ ¼  ð24Þ
1  po Dx retentate and permeate sides is much smaller than the feed
x¼0
pressure, leads to the following estimates: Lx is approximately
_ !1 þ R unity, C is of order one, po is between 0.8 and 1, and R is between
KL x Dp 0 and 0.5. Regarding the parameter A, an estimate can be
U out ðyÞ ¼  ð25Þ
1  po Dx _ derived through its relation to R. Assuming the simplest case with
x ¼ Lx
no permeate pressure drop and linear pressure drop law (C = 0,
Z 1 Z 1
R=0), the retentate pressure equation can be solved to give
U out, av ¼ U out dy, U in, ave ¼ U in dy ð26a, bÞ pffiffi pffiffi
0 0 p ¼ c 1 e  Ax þ c 2 e A x ð27aÞ

4656 dx.doi.org/10.1021/ie102083j |Ind. Eng. Chem. Res. 2011, 50, 4653–4666


Industrial & Engineering Chemistry Research ARTICLE

P ¼ Pð0Þ þ APð1Þ þ A2 Pð2Þ þ ::: ð32Þ


The above expansions must be substituted in eqs 13 and 14. A
problem arises from the nonlinear term containing a power of the
pressure gradient magnitude. Its expansion to a power series of A
is cumbersome; thus, it will be described in detail.
jrpjR ¼ ðp2x þ p2y ÞR=2
ð1Þ 2 R=2
¼ ððpð0Þ ð1Þ 2 ð0Þ
x þ Apx Þ þ ðpy þ Apy Þ Þ
0 0 12 1R=2
!2
B pð0Þ pð1Þ C
¼ @pð0Þ2 1 þ A xð0Þ þ pð0Þ2 @1 þ A y A A
x y ð0Þ
px py

Figure 2. Value of parameter A (for C = 0, R=0) versus flow recovery ð33Þ


fraction R, for several values of dimensionless pressure po. Only the first order expansion with respect to A is of interest;
so, using (1 þ x)2 ≈ 1 þ 2x results in
where
pffiffi jrpjR  ðpð0Þ2
x þ pð0Þ2
y þ 2Aðpð0Þ ð1Þ ð0Þ ð1Þ R=2
x px þ py py ÞÞ
e AL x
 po
c1 ¼ pffiffi pffiffi ð27bÞ R=2
e ALx  e AL x ¼ ðpð0Þ2
x þ pð0Þ2
y Þ ½1 þ 2Aðpð0Þ ð1Þ ð0Þ ð1Þ
x px þ py py Þ
R=2
pffiffi =ðpð0Þ2
x þ pð0Þ2
y Þ ð34Þ
po  e ALx
c2 ¼ pffiffi pffiffi ð27cÞ An additional term of the Taylor expansion (1 þ x)R/2 ≈ 1 þ
e ALx  e ALx
Rx/2 leads to the following first order expansion of the nonlinear
The flow recovery fraction R is given as term:
pffiffi pffiffi
c 2 e AL x  c 1 e  AL x jrpjR  jrpð0Þ jR ½1 þ ARðpð0Þ ð1Þ ð0Þ ð1Þ ð0Þ2
x px þ py py Þ=ðpx þ pð0Þ2
y Þ
R ¼ 1 ð28Þ
c2  c1 ð35Þ
Using the above relations, after some simple algebra, leads to the Substituting 31, 32, and 35 into 13, performing some algebra and
following equation relating A to Lx, po, and R. equating terms of zero order and first order in A results in the
pffiffi pffiffi 2po ð1  RÞ þ 2 following two equations:
e AL x
þ e ALx
¼ ð29Þ
po þ 1  R r 3 K 1 þ R ðx, yÞjrpð0Þ jR rpð0Þ ¼ 0 ð36Þ
The values of A are presented as a function of R, for Lx = 1 and 2 3
several values of po in Figure 2. In practice, R for a reverse osmosis pð1Þ ð0Þ ð1Þ ð0Þ
x px þ py py
membrane element does not exceed 0.20.3 because the con- r 3 K 1 þ R ðx, yÞjrpð0Þ jR 4rpð1Þ þ R ð0Þ2 ð0Þ2
rpð0Þ 5
px þ py
comitant increase of solute concentration, concentration polar-
ization and osmotic pressure set an upper limit to flow
recovery.1,2 An effective operation also requires po > 0.8. Under ¼ Mðx, yÞðpð0Þ  Pð0Þ Þ
these conditions, A resulting from Figure 2 is a small number. ð37Þ
Given the small value of A, a Taylor series expansion of the
exponential terms in eq 29 leads to the following simplified The first term in the brackets, at the left-hand side, corresponds
relation to the firstorder correction ignoring permeability contribution,
whereas the second term is the correction for the pressure drop
1 2Rð1  po Þ dependent permeability. The computations showed that these
A¼ ð30Þ
Lx 2 po þ 1  R terms are of comparable value, so simplification is not possible.
In the same way, retaining terms up to second order for A,
which describes the curves of Figure 2 with a maximum error less eq 14 is transformed to
than 0.5% (observed at large values of A). The complication
arising from non-zero C and R values has a small quantitative r2 Pð0Þ ¼ 0 ð38Þ
influence on A values, but it does not change the general picture,
that is, that A is much smaller than 1 under conditions prevailing r2 Pð1Þ ¼  Mðx, yÞCpð0Þ ð39Þ
in practice.
4.2. Small A Expansion. The smallness of parameter A
r2 Pð2Þ ¼ Mðx, yÞCðPð1Þ  pð1Þ Þ ð40Þ
suggests a regular perturbation expansion of the unknown
pressure fields with respect to A, that is, The boundary conditions for P(i) fields are the same as those of P
field. The boundary conditions for p(0) are exactly the same as
p ¼ pð0Þ þ Apð1Þ þ A2 pð2Þ þ ::: ð31Þ those for p. Finally, the boundary conditions for p(i) with i > 1 are
4657 dx.doi.org/10.1021/ie102083j |Ind. Eng. Chem. Res. 2011, 50, 4653–4666
Industrial & Engineering Chemistry Research ARTICLE

zero normal derivative at y = 0 and y = 1, and zero value at x = 0 The solution for p(0) is rather simple, and it is given as
and x = Lx. R x 1
A very important aspect of the problem is that P(0) is governed ð0Þ K ðxÞdx
p ¼ 1  ð1  po ÞR L0x ð45Þ
1
0 K ðxÞdx
by the Laplace equation augmented by four homogeneous
boundary conditions leading to the trivial solution P(0) = 0. This
explains why three terms are retained for P. The actual expansion Taking into account that p(1) is also a function of x (as it can be
has the form P = AP(1) þ A2P(2), that is, the permeate pressure formally proven) and replacing the expression for p(0) in eq 36,
field itself is of order A, so an accurate permeate pressure field results after some algebra in the following simple equation for
calculation requires the P(2) field. On the other hand, p(0) is of p(1)(x):
order 1 and p(1) is the solution of a generalized Poisson equation !R
(as it will be clearly shown in the following) with a source term (i. Lx D 1 þ R Dpð1Þ
ð1 þ RÞ R Lx K ðxÞ
e., p(0) since P(0) = 0) of order 1 and two Neumann and two 1 Dx Dx
0 K ðxÞdx
Dirichlet homogeneous boundary conditions. The solution of !
the above Poisson equation leads clearly to a function being R x 1
0 K ðxÞdx
everywhere much smaller than 1 and accordingly than p(0). The ¼ MðxÞ 1  ð1  po ÞR Lx ð46Þ
1
same analysis can be extended to the next expansion term A2p(2), 0 K ðxÞdx
which can be shown to be completely negligible with respect to
p(0). This is why the very complicated equation for p(2) is not The above differential equation can be solved using the boundary
written down above. Consequently, p(0), p(1), P(1), and P(2) are conditions for p(1) to obtain
needed for a good approximation of practical flow fields. The   Z Z x0
ð1Þ 1 kðLx Þ R x 1
membrane flux distribution is computed as p ¼ 1 þ R ðx0 Þð  Z þ Mðx00 Þ½1
1 þ R Lx 0 K 0
uw ¼ Mðx, yÞðpð0Þ þ Aðpð1Þ  Pð1Þ Þ  A2 Pð2Þ Þ ð41Þ  ð1  po Þkðx00 Þ=kðLx Þdx00 Þdx0 ð47aÞ
The inlet velocity can be computed from where
Z Z x0
" #! 1 þ R 1 Lx
1
KLx Dpð0Þ Dpð1Þ Z¼ Mðx00 Þ½1  ð1  po Þkðx00 Þ
U in ðyÞ ¼  þA ð42Þ kðLx Þ 0 K 1 þ R ðx0 Þ 0
1  po Dx Dx
x¼0 =kðLx Þdx00 Þdx0 ð47bÞ
A similar computation for the outlet velocity is not equally Rx
accurate since the smallness of the value of the second order and the function k(x) = 0K1(x)dx has been used for clarity of
expansion term of p does not hold for its derivative at x = Lx. This presentation.
is why, using only first-order expansion of p, R turns out to be 4.3.2. Permeate Side Functions. An analytical solution is
independent of C (permeate side pressure drop). To recover the possible only for the P(1) field and it is shown in detail in
correct order relation for R, avoiding the need for the use of p(2), Appendix A. This analytical solution is rather complicated; thus,
the alternative way for estimating R must be considered, that is, it is better to resort to approximate techniques for the solution of
the governing eqs 39 and 40. For a relatively small variation of the
Z 1 Z Lx
function M(x), it has been observed that the y-profile of the
R=R o ¼ uw dxdy=U in, ave ð43Þ permeate pressure field is self-similar with respect to the x-
0 0
position, that is, the profiles P(x,y)/P(x,0) coincide for all x
Up to this point, the only achievement is that the original set of values. This implies that a solution of the form F(x)S(y) exists for
two coupled nonlinear PDEs was replaced by four linear PDEs, the pressure fields P(1) and P(2). In general, the variation of these
which can be solved successively. However, the advantage is that fields with x are expected to be small since the only x variation is
numerical stiffness has been removed and the new equations can due to the right-hand side source term. This typically small
be solved directly using stationary linear solvers. For the cases variation is smoothed out by the action of the x-diffusion term
considered here, the reduction of computation time achieved by combined with the two no-flux boundary conditions. On the
the linearization procedure is 1015 times. This is very im- other hand, the Dirichlet condition at y = 0 imposes a consider-
portant in cases where the quasi-steady flow field must be solved able variation of the flow field in the y direction. This large
many times, as is the case in foreseen model applications of difference in the variation of the pressure field, between the two
fouling deposit evolution. directions, leads to the particular product type of solutions
4.3. Closed Form Solutions for K and M Not Depending on observed. Seeking a solution of the product type, one can
y. 4.3.1. Retentate Side Functions. For this particular case, some substitute it in eq 39, which, after separating the x- and y-terms,
exact solutions can be derived for the perturbation flow fields. results in an equality of the x and y terms. But as is well-known
The absence of explicit appearance of y in eqs 36 and 37 from the separation of variables technique, each side of the
combined with the two zero-derivative boundary conditions at equation can be assumed to take a constant value. Thus, for the y
y = 0 and y = 1 reveals that p(0) and p(1) are functions only of x. terms
The corresponding equations take the form (taking into
account the negative sign of dp(0)/dx): D2 S
¼c ð48Þ
Dy2
!R þ 1
D Rþ1 Dpð0Þ
K ðxÞ  ¼0 ð44Þ where c is an arbitrary constant. The solution of the above
Dx Dx
equation employing the boundary condition at y = 0 and y = 1
4658 dx.doi.org/10.1021/ie102083j |Ind. Eng. Chem. Res. 2011, 50, 4653–4666
Industrial & Engineering Chemistry Research ARTICLE

Figure 3. Hierarchical decision tree, showing several levels of simplification and suggesting the appropriate solution approach for each level, based on
the findings of the present work.

and choosing c, to have an average of S equal to unity, leads to diffusive terms of the right-hand side and the no flux boundary
! conditions tend to smooth out this y dependence. Consequently,
ð1Þ y2 ð1Þ p(1) is weakly dependent on y, and, since its contribution to p is
P ¼ 3 y F ðxÞ ð49Þ
2 small, one can ignore this dependence, assuming that p(1) is
(1)
R 1 (1) in y, and seek its y-average form (i.e., p (x) =
uniform
This form can be assumed to be a one term expansion of the 0 p (x,y)dy). It is noted that here the y-independence is just
unknown function in bases functions in the y-direction. The equation an assumption, unlike in the case M(x,y) = M(x) examined above,
for F(1)(x) can be derived by using a weighted residual approach (in where the y-independence is a formal mathematical result.
particular the method of moments23). Substituting the trial form 49 Assuming p(1) is y-independent, taking the integral of eq 37 for
in equation and taking the integral of both sides with respect to y, y from 0 to 1 and employing the zero derivative boundary
from y = 0 to y = 1, the following equation is derived: conditions at these
R points results in the one-dimensional eq 46
but with M(x) = 01M(x,y)dy. Therefore, the analytical solution
D2 F ð1Þ ðxÞ
 3F ð1Þ ¼  CMðxÞpð0Þ ð50Þ derived above, for M being a function only of x, holds by simply
Dx2 averaging M(x,y) in the y direction.
The same procedure is followed for P(2). The function S turns out to The situation with P(1) and P(2) is quite different. The y-
be the same as for P(1), that is, ! dependence in the right-hand side of eqs 39, 40, “interacts” with
ð2Þ y2 ð2Þ the Dirichlet boundary conditions and cannot be ignored in any
P ¼ 3 y F ðxÞ ð51Þ case (since the y-dependence is not induced by M but it is
2
intrinsic to the problem). On the other hand, the weighted
Substituting in eq 40 and integrating with respect to y leads to the residual approach used for M(x) can be easily extended to the
following equation for F(2): case of M(x,y), but it is found that it has only partial success since
the y-dependence of M modifies the trial function y  y2/2. The
D2 F ð2Þ ðxÞ
 3F ð2Þ ¼ CMðxÞðF ð1Þ  pð1Þ Þ ð52Þ only way to obtain an accurate solution for any form of M(x,y) is
Dx2 by solving numerically the Poisson eqs 39 and 40. However, as
Procedures for the approximate solution of the eqs 50 and 52 can be analytical or approximate solution of these equations is no longer
found in Appendix B. The particular case of clean membrane (K = M possible, there is no reason to keep them separately and not go
= 1) is treated in Appendix C. back and combine them to an equation, which gives directly the
4.4. Approximate Solution for Functional Dependencies permeate pressure field, that is
K(x) and M(x,y). This is a very important case, where a significant
simplification of the problem is possible (compared to the case of D2 P D2 P
K being a function of both x and y). The fact is that p(0) does not þ ¼ Mðx, yÞðP  pð0Þ  Apð1Þ Þ ð53Þ
Dx2 Dy2
depend on M; thus, it remains a function of x. It will be recalled that
the first correction term Ap(1) is only a few percent of the dominant The reason that renders this back step advantageous is that the
term p(0). In addition, in the present case, the y dependence of p(1) new equation is also linear. Thus, it was shown that in the case of
is induced by the right-hand side term of eq 37, whereas the K(x) and M(x,y) functions the approximately one-dimensional
4659 dx.doi.org/10.1021/ie102083j |Ind. Eng. Chem. Res. 2011, 50, 4653–4666
Industrial & Engineering Chemistry Research ARTICLE

Figure 5. Inlet velocity profiles corresponding to the three flow fields of


Figure 4.

typical results are shown here. All the results shown below refer
to a typical domain of square shape (i.e., Lx = 1) representative of
membrane elements used in practice. It will be noted, however,
that the problem can be also formulated for other membrane
sheet (Lx 6¼ 1) geometries.
First, the influence of the function K(x,y) on the p(0) field will
be presented; it will be also shown why its y dependence
precludes any simplification. As it is obvious from eq 37, the
only parameter for the p(0) field is the exponent R. In Figure 4 the
flow streamlines are shown for the p(0) field, for R = 0.4 and
K(x,y) = 1 everywhere except at the drawn disk where K(x,y) =
Kc, for three values of Kc. These streamlines are moderately
deformed at the boundaries of the disk for Kc = 0.5 (Figure 4a),
very deformed for Kc = 0.2 (Figure 4b) and extremely deformed
for Kc = 0.01 (Figure 4c) for which actually no fluid is passing
through the low permeability region. This streamline deforma-
tion not only precludes simplification of eq 37 for p(0) but
Figure 4. Flow streamlines corresponding to the pressure field p(0) for renders questionable the validity of the original eqs 8 and 9 (at
R = 0.4. The function K = 1 everywhere except in the drawn circle least locally), which were based on the assumption of small
where K = Kc. (a) Kc = 0.5, (b) Kc = 0.2, and (c) Kc = 0.01. deviations from unidirectional flow.14 To support this assump-
tion, it is stressed that the strong y-dependence of K(x,y) is the
retentate-side pressure field can be determined analytically and only possible way leading to its violation. The inlet flow profiles
the permeate side pressure field can be obtained by solving corresponding to the three values of Kc are shown in Figure 5.
numerically a simple linear partial differential equation. This is a Obviously, the existence of the low permeability region influ-
large reduction of the computational effort, compared to the ences the inlet flow profile. This confirms the assertion that fixing
system of two coupled partial differential equations, one of them the inlet pressure is the physically correct boundary condition, and
being nonlinear, prone to create numerically stiff behavior in the not fixing the inlet velocity, which is associated with an unrealistic
problem. uniform inlet velocity profile. It should be noted, parenthetically,
that the reduction of the retentate velocity (i.e., usually referred
to as cross-flow velocity), upstream of fouled channel region, may
5. RESULTS AND DISCUSSION further aggravate fouling, because of the reduced flow shearing at
All findings of the present work are summarized in Figure 3, the membrane surface.
where the starting point is the complete set of equations; all the To show the accuracy of the analytical solution, the numer-
approaches proposed in the present work, where problem ical and the analytical results for several quantities of interest
reduction is possible are included. In this section no parametric and for seven cases with realistic values of the parameters are
analysis of the model will be presented since it is not within the presented in Table 1, for K(x,y) = M(x,y) = 1. It will be recalled
scope of the present work. Only typical numerical results that these values (K = M = 1) represent the case of membranes
confirming the theoretical statements of the previous analysis operating with no foulants that can affect both retentate
will be shown (i.e., comparison between exact and approximate channel and membrane permeability. Two values of A
results). It is stressed that the theoretical development is (accounting for membrane permeability), two values of po
supported by the full spectrum of the results (actually the (accounting for retentate side pressure drop), two values of R
derivation was based on combined formal mathematical argu- (resistance law exponent), and three values of C (accounting for
ments and numerical computations) but by necessity only some permeate side pressure drop) are employed in seven

4660 dx.doi.org/10.1021/ie102083j |Ind. Eng. Chem. Res. 2011, 50, 4653–4666


Industrial & Engineering Chemistry Research ARTICLE

Table 1. Parameters for the Cases 1 to 7 and Corresponding


Numerical Results
case A C R po R uwave Pave(y = 1) σvw (100)

1 0.1 1 0.4 0.9 0.45 0.908 0.0448 3.316


2 0.05 1 0.4 0.9 0.246 0.928 0.023 3.068
3 0.05 5 0.4 0.9 0.234 0.872 0.1068 4.554
4 0.1 1 0.3 0.9 0.379 0.909 0.0449 3.303
5 0.05 1 0.4 0.8 0.125 0.88 0.0218 6.131
6 0.05 5 0.4 0.8 0.101 0.827 0.0101 7.039
7 0.1 2 0.4 0.8 0.227 0.834 0.0818 6.764

Figure 7. Function q(x) = (1 þ R)p(1)(x) for several values of po ; K =


1, M = 1.

presented in Table 1 will be used to confirm several analytical


results for spatially distributed quantities (in addition to the
integral values shown in Table 1).
The analytical distribution of uw is so close to the numerical
one that a graphical comparison is without scope. The contours
of uw for the cases 1 and 6 are shown in Figure 6a and b,
respectively. Obviously, the uw distribution depends on the
problem parameters exhibiting, for the uniform membrane
permeability case, a maximum at the corner between inlet reten-
tate and outlet permeate sides and a minimum at the opposite
corner. The exact knowledge of this distribution is important for
the prediction of fouling phenomena.
The first-order correction field for the retentate pressure
and no y dependence of K and M can be written as p(1) = q/(1
þ R). This can be decomposed into two terms, that is, the first
is equal to q, representing the constant permeability first-order
correction, and the second term, equal to [Rq/(1 þ R)], is
the correction arising from the nonlinearity of the perme-
ability constitutive law. The function q(x) for K(x) = M(x) = 1
depends only on po, and it is shown in Figure 7 for several po
values. It is obvious that q takes small values (maximum up to
0.12, average about 0.07, which confirms that the first order
correction to retentate pressure field is already very small and
higher-order correction is not needed. The nonlinear resis-
tance expression can increase p(1) up to twice (for R = 0.5),
but even in the worst case, the average contribution of p(1) to p
Figure 6. Permeate flux uw contours: (a) Case 1 and (b) case 6. is less than 2%. Therefore, the mathematical statements of the
Parameters values listed in Table 1. previous sections, regarding the retentate pressure field, are
confirmed numerically. The average deviation of 2% may seem
representative combinations. The results presented are the quite small, thus raising the question of the practical signifi-
permeate recovery fraction R, the average wall velocity uwave, cance of the term p(1) for determining the pressure field. It will
the average permeate pressure at the y = 1 side (which can be be noted, however, that this term cannot be ignored because
considered as the permeate pressure drop) and the standard its contribution to the local pressure slope (related to the flow
deviation of the wall velocity σvw. It should be stressed that, for through eq 24) is significant.
these cases, the small deviation of analytical results from the Next, some numerical results are presented in Figure 8 to
presented results is smaller than the accuracy of the numerical assess (and confirm) the validity of the self-similar profile of
methods; thus, they are practically exact and there is no need to permeate pressure in the y direction (i.e., shape 2y  y2). The
be presented. The analytical techniques provide accurate pre- normalized (with respect to their values at y = 1) y profiles of P
dictions even in the rather extreme cases where the retentate for the case 6, at three values of x, are shown in Figure 8. The
pressure drop is 20% and the permeate pressure drop is 10% of profiles are indistinguishable from each other and from the basis
the feed pressure, the flow recovery fraction is 45% and the function [2y - y2] in the scale of the figure, even though the
membrane flux standard deviation is 9% of its average value. selected case corresponds to a significant permeate side pressure
The numerical results from the cases with parameter values drop. This confirms the validity of our technique to approximate
4661 dx.doi.org/10.1021/ie102083j |Ind. Eng. Chem. Res. 2011, 50, 4653–4666
Industrial & Engineering Chemistry Research ARTICLE

Figure 8. Normalized permeate pressure y-profiles at several x-posi-


tions (x = 0, 0.5, 1) compared to the basis function 2y  y2 (case 6,
Table 1). The four curves are indistinguishable.

Figure 11. Retentate side (a) pressure contour and (b) flow streamlines
for A = 0.1, C = 0, R = 0.4, po = 0.8, and M = 1 everywhere except in the
drawn circular region where M = 0.

shown in Figure 9 to examine the significance of the R exponent.


According to the above analysis, the difference for p is [1/(1 
0.3)  1/(1  0.4)]q(x) = 0.238q(x). This is confirmed from
the numerical results shown in Figure 9. The wall velocity
profiles are shown to confirm that the y pressure profiles (and
the corresponding uw profiles) have the same self-similar shape,
which is independent of the exponent R. As one would have
Figure 9. Retentate side pressure p versus x for y = 0 and membrane flux
expected, the local permeate fluxes uw at y = 0 are significantly
uw versus 1  y at x = 0 and x = 1 for case 4 (continuous lines) and case 1 larger compared to those at y = 1 because of the greater local
(crosses). Parameter values listed in Table 1. differences of the trans-membrane pressure at y = 0.
The numerical results for the permeate pressure P(x,1) and
the first-order analytical approximation based on eqs 49 and
C4 are depicted in Figure 10, for several cases. In each case,
there is a constant difference between the two curves, which is
exactly accounted for by adding the uniform value approxima-
tion for the F(2) function (combination of eqs 51 and C5). It is
also noted that the third order polynomial approximation for
F(1) leads to results indistinguishable form those shown in
Figure 10 (based on the analytical solution for F(1)). This is
important to be noted since the analytical solution exists only
for M = 1, and the polynomial approximation can be used for
arbitrary M(x).
Finally, several results will be shown to confirm the assertion
that the y-dependence of the retentate pressure field p for
arbitrary membrane permeability can be ignored.
The problem is solved numerically for A = 0.1, C = 0, R = 0.4,
po = 0.8, and two shapes of the function M(x,y). In the first case
Figure 10. Numerical solution (dashed lines) and first-order analytical (Figure 11) the membrane permeability exhibits a sharp variation
approximation P(1) (solid lines) of the permeate pressure at the edge y = (a disk with zero permeability), and in the second (Figure 12), a
1 along the x-direction. smooth variation M(x,y) = y. It is noted that the selected
parameters correspond to the most extreme limit of conditions
the fields P(1) and P(2) (eqs 49 and 51). The retentate pressure examined in this work, that is, they correspond to F = 0.35. The
field p and the uw profiles at x = 0 and x = 1, for cases 1 and 4, are pressure contours and the flow streamlines for the case of a zero
4662 dx.doi.org/10.1021/ie102083j |Ind. Eng. Chem. Res. 2011, 50, 4653–4666
Industrial & Engineering Chemistry Research ARTICLE

how would the findings of the present work be affected if the


assumption about constancy of f2 were relaxed? The answer is
that the linearization procedure still holds but this is not the case
for the derived analytical solutions.

6. CONCLUSIONS AND COMMENTS


In general, by employing a formal mathematical analysis, the
present work unifies, confirms, and corrects, wherever needed,
previous analytical solutions to the problem which are based
on heuristic arguments, offering a generalized guide, for selecting
models at the appropriate level, for the simulation of the
hydrodynamics of flat-sheet spiral wound modules. In particular,
it is formally proven here that the pressure at the retentate side
essentially varies only in the x direction, whereas the pressure at
the permeate side is only y-dependent, as is usually assumed in
the literature for clean membranes. In addition, it is shown that,
for the case of a fouled membrane, the dominant retentate
pressure terms do not necessarily depend on the permeate
pressure field, as previously suggested.14
It is shown that to compute the total flow resistance and the
wall velocity distribution for flat-sheet membrane modules with
spacers, characterized by specific constitutive relations for the
Figure 12. Retentate side (a) pressure contour and (b) flow streamlines local pressure drop-velocity, the direct solution of the nonlinear
for A = 0.1, C = 0, R = 0.4, po = 0.8, and M = y. Linear membrane system of two 2-dimensional mass balances is not necessary. A
permeability variation. formal small-parameter mathematical expansion is always pos-
sible, for operating conditions of practical interest, leading to a
simplification of the problem. The present analysis covers the
cases of arbitrary spatial distributions of retentate channel and
membrane permeability. Several exact or approximate solutions
of the mathematical problem are derived, which are related to
the spatial dependence of these permeabilities. The validity of
proposed solution approaches was confirmed by comparing
their results with those from the numerical solution of the
original problem. These approaches are summarized in the
chart of Figure 3 of this paper and include all possible
approximate and asymptotic solutions obtained up to now for
the problem at hand. Moreover, this chart can be considered as
a guide for the selection of the appropriate level of approxima-
tion for solution of the mathematical problem, depending on a
particular set of parameters. By following this general approach,
one can avoid in most cases an unnecessary complexity, thus
facilitating large scale simulations.
Figure 13. Retentate side pressure y-profiles at several x positions for From a more pragmatic point of view, long time-scale
the conditions of Figure 11 (dashed lines) and Figure 12 (solid lines). simulations of the operation of flat-sheet membrane modules,
considering also the development of scaling/fouling deposits,
requires the solution of the system of eqs 13 and 14 resulting
permeability circular patch are shown in Figure 11; it is obvious from mass and momentum balances. These equations must be
that even in this worst case “scenario” the flow field can be safely solved at each time step since the functions M and K evolve as
considered unidirectional and the pressure field essentially uni- fouling/scaling proceeds; obviously, coupling with a realistic
form in the y direction. The corresponding graphs for the linear model for these processes is needed. The stiffness of the
permeability profiles are shown in Figure 12; there is a slight system, owing to its nonlinearity, renders the long time
tendency of the flow to move toward the high permeability side simulation task computationally arduous and any simplifica-
but it is small to invalidate the unidirectional flow approximation. tion of the system of the core eqs 13 and 14 is crucial.
To provide a better view, for the deviations of the pressure profile Fortunately, the mathematical structure of the problem allows
from the uniformity in the y direction, the numerical results for simplifications for any set of practical conditions. First, the
p(x,y) are shown in Figure 13, for three values of x and the two original stiff system can be replaced by the linear set of
aforementioned types of M(x,y). It is noted that the maximum eqs 3640, which can be solved sequentially. In case of
local deviation, between the approximate scheme proposed in channel blocking being a function only of the axial coordinate
the previous section (see guidelines in Figure 3) for p(x) with (along the flow direction), the retentate pressure field can be
arbitrary M(x,y) and the numerical results, is 0.2% for the linear determined by the integral relations 45 and 47a and the
permeability profile. A question that may be raised here is, permeate pressure field from the solution of the linear partial
4663 dx.doi.org/10.1021/ie102083j |Ind. Eng. Chem. Res. 2011, 50, 4653–4666
Industrial & Engineering Chemistry Research ARTICLE

differential eq 53. For example, in the case of thin fouling layers where λi = (i þ 1/2)π are the eigenvalues of the problem (i =
(K = 1) the system 45, 47a, and 53 can be used instead of 13 and 0, 1, 2, ...). The expansion coefficients must be determined by
14 greatly facilitating the computation of the evolution of the the non-homogeneous boundary condition at x = Lx. Sub-
spatial distribution of the membrane fouling. Work is in stituting the infinite series in the boundary condition results
progress to implement, and expand, the above approach to in
account for fouling, which is aided by experiments. ¥
Θ¼ ∑ c1i λi ðeλ L i x
 eλi Lx Þsinðλi yÞ ðA7Þ
’ APPENDIX A: ANALYTICAL SOLUTION FOR THE i¼0
FIRST-ORDER PERMEATE PRESSURE FIELD P(1) FOR K Multiplying the above relation by the jth eigenfunction (j = 0,
AND M NOT DEPENDING ON Y 1, 2, ...) and integrating with respect to y from y = 0 to y = 1,
Unlike the case of p(1), in the permeate side the y-dependence the orthogonality of the eigenfunctions leads to the result
of the pressure field is introduced by the boundary condition (in 2Θ
particular the Dirichlet boundary condition at y = 0); thus, it c1i ¼ ðA8Þ
cannot be disregarded. The equation that must be solved for P(1) λ2i ðeλi Lx  eλi Lx Þ
is Regarding Q2, the eigenfunction expansion (in x direction
DPð1Þ DPð1Þ
þ 2 ¼  CMðxÞpð0Þ ðA1Þ where the non-homogeneity appears), which satisfies the
Dx2 Dy boundary conditions at x = 0, x = Lx, and y = 1, is
This equation with the appropriate boundary conditions can be ¥
solved analytically as follows: At first P(1) is split into an harmonic Q2 ¼ ∑ c2i ðeμ y þ eμ ðy  2Þ Þcosðμi xÞ
i¼0
i i ðA9Þ
part (satisfying Laplace equation) and a particular solution to
account for the non-homogeneous terms in A1: where μi = iπ/Lx are the eigenvalues of the problem.
Pð1Þ ¼ Q L ðx, yÞ  Q in ðxÞ ðA2Þ Substitution of the above series in the non-homogeneous
boundary condition at y = 0, results in
Substituting this in A1 and requiring that QL fulfills the Laplace
¥
∑ c2i ð1 þ e2μ Þcosðμi xÞ
equation results in
Q in ðxÞ ¼ i ðA10Þ
D2 Q in i¼0
¼  CMðxÞpð0Þ ðA3Þ
Dx2 Exploiting the orthogonality property of the eigenfunctions
which admits the solution in a way similar to that of the Q1 problem leads to the
Z xZ x0
following expressions for the expansion coefficients:
Q in ¼ C Mðx00 Þpð0Þ dx00 dx0 ðA4Þ Z
1 Lx
0 0 c20 ¼ Q in ðxÞdx ðA11Þ
2Lx 0
The boundary conditions for QL can be found by substituting
eq A2 in boundary conditions for P(1): Z Lx
2
Q L ¼ Q in ðxÞ at y ¼ 0 ðA5aÞ c2i ¼ Q in ðxÞcosðμi xÞdx ðA12Þ
Lx ð1 þ e2μi Þ 0

DQ L Summation of Qin, Q 1, and Q2 leads to the exact solution for


¼ 0 at y ¼ 1 ðA5bÞ the P(1) field. No exact solution is possible for P(2) since P(1)
Dy
introduces y to the right-hand side of the governing eq 40.
DQ L
¼ 0 at x ¼ 0 ðA5cÞ ’ APPENDIX B: APPROXIMATE SOLUTION OF EQS 50
Dx
AND 52 FOR THE FUNCTIONS F(1) AND F(2) REPRE-
Z Lx SENTING THE X DEPENDENCE OF THE PERMEATE
DQ L DQ L
¼ ¼C MðxÞpð0Þ dx ¼ Θ at x ¼ Lx ðA5dÞ PRESSURE FIELD
Dx Dx 0
As has already been discussed, the x variation of the permeate
The Laplace equation must be solved with two non-homo- pressure field, and correspondingly that of F functions, is small. A
geneous boundary conditions. Thus, the unknown harmonic zero order approximation is equivalent to assuming that F(1) and
function is split in two harmonic functions, each one having F(2) have no x-dependency. In such a case, the basic representa-
one non-homogeneous boundary condition, that is, QL = Q 1 tion of these functions is their average values. To determine these
þ Q2, where Q 1 should satisfy the boundary conditions of QL values, the integral of eqs 50, 52 with respect to x are taken (from
(eqs A5b, A5c, and A5d) and Q1 = 0 at y = 0, and Q2 should x = 0 to x = Lx), and then they are divided by Lx to obtain
satisfy A5a, A5b, and A5c with (∂Q2)/(∂x) = 0 at x = Lx. Each Z Lx
one of the sub-problems for Q1 and Q2 can be solved by the 1
F ð1Þ ¼ C MðxÞpð0Þ dx ðB1Þ
separation of variables technique. Regarding Q1 the eigen- 3Lx 0
function expansion (in y direction where the non-homoge-
Z Lx  
neity appears), which satisfies identically the boundary 1 1
conditions at y = 0, y = 1, and x = 0 is F ð2Þ ¼ C MðxÞ pð1Þ  CMðxÞpð0Þ dx ðB2Þ
3Lx 0 3
¥
Q1 ¼ ∑ c1i ðeλ x þ eλ x Þsinðλi yÞ
i¼0
i i
ðA6Þ Unfortunately, the x dependence of F functions cannot be
recovered by a systematic correction of the above results. The
4664 dx.doi.org/10.1021/ie102083j |Ind. Eng. Chem. Res. 2011, 50, 4653–4666
Industrial & Engineering Chemistry Research ARTICLE

structure of the governing equation is typical of a singular correction term is represented by R in the denominator of the
boundary value problem with the outer expansion obtained expression 1/(1 þ R), which shows that for the values of R
directly by the elimination of the derivative term.24 Then the considered here this contribution is comparable to the main
inner expansion is needed in order to satisfy the boundary permeability correction term.
conditions. What makes inevitable this approach (although the The exact P(1) field can be obtained as
shape of the profile of F(1) indicates such an approach, as it can be !
x 2
ð1  p Þ x3
seen later) is the boundary layer thickness for the particular Pð1Þ ¼ C  o
þ CLx ð1 þ po Þ
problem, which is 30.5, that is, the boundary layer covers the 2 Lx 6
whole domain of the definition of the problem for a typical Lx ¥ λi x
þ eλi x Þ 3 þ po
close to one. The exact functions F can be found by solving
analytically (possible only for some simple forms of K(x) and
 ∑ 2ðe
i ¼ 0 λ ðeλ L i x  eλi Lx Þ
sinðλi yÞ þ CLx 2
24
i
M(x)) or solving numerically the two-point linear boundary ¥
value problems (50) and (52). The boundary conditions on F are þ ∑ c2i ðeμ y þ eμ ðy  2ÞÞcosðμi xÞ
i¼1
i i ðC3Þ
zero derivative at x = 0 and x = Lx.
A better approximation of F(1) than the uniform value can be where
derived by applying the weighted residuals method to eq 50. In
particular, a third-order polynomial is assumed for the unknown 2C ð  1Þi Lx 1  po
function, that is, F(1) = ∑4i=1cixi1. This is the lowest-order c2i ¼ 2μ
þ
Lx ð1 þ e i Þ μi 2 6Lx
polynomial capable of predicting a nonuniform solution. Two
of the unknown coefficients can be found from the two boundary " ! #!
2
conditions, and the third one from the known average value of i 3Lx 6 6
 ð  1Þ  4 þ 4
F(1) (determined by taking the zeroth moment with respect to x μi 2 μi μi
of the governing equation). The resulting expression is
Z ! The above exact solution can be approximated, extremely
ð1Þ C Lx ð0Þ 2 3 L2x accurately, by the proposed shape in eq 49. In this particular
F ¼ MðxÞp dx þ c x  2
x  ðB3Þ case the governing equation for F(1) can be solved analytically to
3Lx 0 3Lx 6
give
The last unknown coefficient c can be determined by taking the pffiffi pffiffi pffiffi pffiffi
1p
first moment with respect to x of eq 50 and substituting the F ð1Þ ¼ C pffiffiffi o ½ð1  e 3Lx Þe 3x þ ð1  e 3Lx Þe 3x 
expression B3. After some algebra, the following result is derived: 3 3Lx
 
Z Lx C 1  po
Lx þ 1 x ðC4Þ
c¼  C MðxÞpð0Þ dx 3 Lx
2 0
Z Lx  ! The x-dependence of F(2) is extremely small, so the computation
2 4
L L of the average value of F(2) is sufficient. This value depends only
xMðxÞpð0Þ dx =
x x
C þ ðB4Þ
0 3 10 on the average value of F(1), which is correctly given by B1, that is,
it is the same as the average value of the analytical solution C4.
The weighted residuals method of moments, with polynomial After some algebra, one obtains
basis function employed here, is the same with the well known
integral method used in boundary layer theory.25 The x depen- C po Lx 2 C2
dence of F(2) is extremely small, so there is no reason to seek an F ð2Þ ¼   ð1  ð1  po Þ=2Þ ðC5Þ
3ð1 þ RÞ 12 9
approximation better than its average value, given by eq B2. It is
noted that this result is derived using only the average value of
F(1); thus, an update for the use of polynomial expansion for F(1)
is not needed since the average value remains the same. ’ AUTHOR INFORMATION

’ APPENDIX C: EXPLICIT SOLUTION FOR PRESSURE Corresponding Author


FIELDS FOR THE CLEAN MEMBRANE CASE (K = M = 1) *E-mail: kostoglu@chem.auth.gr.
This is the simplest possible case and the solutions will be ’ REFERENCES
given in detail. The retentate side pressure field is given as
follows: (1) Water Treatment: Principles and Design. Membrane Filtration in
Water Treatment Principles and Design; Wiley: New York, 2005; Chapter
1  po 12.
pð0Þ ¼ 1  x ðC1Þ (2) Schwinge, J.; Neal, P. R.; Wiley, D. E.; Fletcher, D. F.; Fane, A. G.
Lx
Spiral wound modules and spacers. Review and analysis. J. Membr. Sci.
  ! 2004, 242, 129–153.
ð1Þ 1 x2 1  po x3 Lx Lx (3) Koutsou, C. P.; Yiantsios, S. G.; Karabelas, A. J. Numerical
p ¼  þ ð1  po Þ  x simulation of the flow in a plane channel containing a periodic array of
1þR 2 Lx 6 6 2
cylindrical turbulence promoters. J. Membr. Sci. 2004, 231, 81–90.
ðC2Þ (4) Koutsou, C. P.; Yiantsios, S. G.; Karabelas, A. J. A numerical and
experimental study of mass transfer in spacer-filled channels: Effects of
The above terms can be identified in eq 37. The contribution of spacer geometrical characteristics and Schmidt number. J. Membr. Sci.
the nonlinear pressure drop-velocity relation to the permeability 2009, 326, 234–251.

4665 dx.doi.org/10.1021/ie102083j |Ind. Eng. Chem. Res. 2011, 50, 4653–4666


Industrial & Engineering Chemistry Research ARTICLE

(5) Fimbres-Weihs, G. A.; Wiley, D. E. Review of 3D CFD modeling


of flow and mass transfer in narrow spacer field channels in membrane
modules. Chem. Eng. Process. 2010, 49, 759–781.
(6) Wardeh, S.; Morvan, H. P. CFD simulations of flow and
concentration polarization in spacer filler channels for application to
water desalination. J. Chem. Eng. Res. Des. 2008, 86, 1107–1116.
(7) Ranade, V. V.; Kumar, A. Comparison of flow structures in
spacer filled flat and annular channels. Desalination 2006, 191, 236–244.
(8) Picioreanu, C.; Vrouwenvelder, J. S.; van Loosdrecht, M. C. M.
Three dimensional modeling of biofouling and fluid dynamics in feed
spacer channels of membrane devices. J. Membr. Sci. 2009, 345, 340–
354.
(9) Geraldes, V.; Semiao, V.; de Pinho, M. N. The effect of the
ladder-type spacers configuration in NF spiral-wound modules on the
concentration boundary layers disruption. Desalination 2002, 146, 187–
194.
(10) Geraldes, V.; Semiao, V.; de Pinho, M. N. Flow management in
nanofiltration spiral wound modules with ladder type spacers. J. Membr.
Sci. 2002, 203, 87–102.
(11) Koutsou, C. P.; Yiantsios, S. G.; Karabelas, A. J. Direct
numerical simulation of flow in spacer-filled channels: Effect of spacer
geometrical characteristics. J. Membr. Sci. 2007, 291, 53–69.
(12) Van der Meer, W. G. J.; Van Dijk, J. C. Theoretical optimization
of spiral wound and capillary nanofiltration modules. Desalination 1997,
113, 129–146.
(13) Boudinar, M. B.; Hanbury, W. T.; Avlonitis, S. Numerical
simulation and optimization of spiral wound modules. Desalination
1992, 86, 273–290.
(14) Kostoglou, M.; Karabelas, A. J. On the fluid mechanics of spiral-
wound membrane modules. Ind. Eng. Chem. Res. 2009, 48, 10025–
10036.
(15) Evangelista, F. Optimal design and performance of spiral
wound modules. II. Analytical method. Chem. Eng. Commun. 1988,
72, 81–94.
(16) Evangelista, F. An improved analytical method for the design of
spiral wounds modules. Chem. Eng. J. 1988, 38, 33–40.
(17) Kim, A. S. Permeate flux inflection due to concentration
polarization in cross-flow membrane filtration: A novel analytic ap-
proach. Eur. Phys. J. E. 2007, 24, 331–341.
(18) Karode, S. K. Laminar flow in channels with porous walls,
revisited. J. Membr. Sci. 2001, 191, 237–241.
(19) Aris, R. Mathematical Modeling Techniques; Dover: New York,
1994.
(20) Tzotzi, Ch.; Pahiadaki, T.; Yiantsios, S. G.; Karabelas, A. J.;
Andritsos, N. A study of CaCO3 scale formation and inhibition in RO
and NF membranes. J. Membr. Sci. 2007, 296, 171–184.
(21) Vrouwenvelder, J. S.; Graf von der Schulenburg, D. A.;
Kruithof, J. C.; Johns, M. L.; van Loosdrecht, M. C. M. Biofouling of
spiral-wound nanofiltration and reverse osmosis membranes: A feed
spacer problem. Water Res. 2009, 43, 583–594.
(22) Sioutopoulos, D.; Yiantsios, S. G.; Karabelas, A. J. Relation
between fouling characteristics of RO and UF membranes in experi-
ments with colloidal organic and inorganic species. J. Membr. Sci. 2010,
350, 62–82.
(23) Villadsen, J.; Michelsen, M. L. Solution of Differential Equation
Models by Polynomial Approximation; Prentice-Hall: New York, 1978.
(24) Kevorkian, J.; Cole, J. D. Multiple Scale and Singular Perturba-
tion Methods; Springer: New York, 1996.
(25) Deen, W. M. Analysis of Transport Phenomena; Oxford Uni-
versity Press: New York, 1998.

4666 dx.doi.org/10.1021/ie102083j |Ind. Eng. Chem. Res. 2011, 50, 4653–4666

You might also like